You are on page 1of 203
TRANSPORT IN POROUS CATALYSTS R. JACKSON Department of Chemical Engineering University of Houston, Houston, Texes, U.S.A. ELSEVIER SCIENTIFIC PUBLISHING COMPANY AMSTERDAM ~ OXFORD— NEW YORK = 1977 CHEMICAL ENGINEERING MONOGRAPHS Edited by Professor S.W. CHURCHILL, Department of Chemical Engineering, University of Pennsylvania, Philadelphia, Pa. 19104, U.S.A. Vol.1 Polymer Engineering Vol. 2 Filtration Post-Treatment Processes Vol. 3 Multicomponent Diffusion Vol. 4 Transport in Porous Catalysts ELSEVIER SCIENTIFIC PUBLISHING COMPANY 335 Jan van Galenstraat P.O. Box 211, Amsterdam, The Netherlands Distributors far the United States and Canada: ELSEVIER NORTH-HOLLAND INC, 52, Vanderbilt Avenue New York, N.Y. 10017 Library of Congress Cataloging In Publication Data Jackson, Roy, 1932- Transport in porous catalysts, (Caemical engineering onogeaphs ; 7. 4) Bibliography: Pe Includes indexes. 1, Catalysts. 2. Transport theory, I. Title, BES SHL 395 TI B292 TW O-W44 35 93-9 ISBN: 0-444-41295-6 (series) ISBN: 0-444-41593-9 (vol. 4) © Elsevier Scientific Publishing Company, 1977. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechan- ical, photocopying, recording or otherwise, without the prior written permission of the publisher, Elsevier Scientific Publishing Company, P.O. Box 330, Amsterdam, ‘The Netherlands Printed in The Netherlands PREFACE This book is concerned with diffusive transport in porous media, with particular emphasis on the modeling of chemical reaction in porous catalyst pellets. Much of the existing literature in this field restricts attention to binary mixtures, Knudsen diffusion, or some other simple special case. However, it is rare indeed that only two chemical species participate in reactions of practical importance, and in gas phase reactions it is comaon for the catalyst pore size distribution to span the “awkward range in which the pore diameter and molecular mean free path lengths are comparable in size. Particular attention has therefore been given here to developing the physical theory in a form general enough to apply to practical situations with multicomponent mixtures and pores in the intermediate size range. By collecting the rather scattered experimental and theoretical results currently available in this fleld it is hoped to foster a better under standing of the influence of physical processes in catalytic reactions of realistic complexity and stimulate further work in the numercus areas where our knowledge is incomplete, The greater part of the material for this monograph was assembled while the author was on sabbatica) leave at Imperial College, London, duting the spring of 1975, Thanks are due to Prof. R, W. H. Sargent for arranging this visit and to the Science Research Council for some financial support during this peried. In organizing my thoughts on this subject I have also benefited greatly from discussions with a number of my colleagues; in particular, T should mention R, Aris, W. E. Stewart and D. Luss. A number of diagrams are based on published results from vartous sources, as indicated in the captions. I am grateful to the respective publishers for permission to make use of this material. Houston, R. JACKSON Texas TABLE OF CONTENTS PREFACE, v CHAPTER 1, INTRODUCTION, 1 CHAPTER 2. FLUX RELATIONS FROM SIMPLE MOMENTUM TRANSFER ARGUMENTS, 6 CHAPTER 3. THE DUSTY GAS MODEL, 18 CHAPTER 4, GAS MOTION IN A LONG TUBE AT THE LIMIT OF BULK DIFFUSION AND VISCOUS FLOW, 25 CHAPTER 5. ALGEBRAIC MANIPULATIONS AND LIMITING FORMS OF THE DUSTY GAS MODEL EQUATIONS, 34 CHAPTER 6, SOME IMPORTANT EXPERIMENTAL RESULTS ON GAS MOTION IN POROUS MEDIA AND CAPILLARIES, 50 CHAPTER 7. SURFACE DIFFUSION, 59 CHAPTER &. MODELS OF FLOW AND DIFFUSION IN POROUS MEDIA, 63 CHAPTER 9. FLUX RELATIONS UNDER REACTIVE CONDITIONS, 77 CHAPTER 10, EXPERIMENTAL CHARACTERIZATION AND TESTING OF FLUX MODELS, 38 CHAPTER 11, STEADY STATE MATERIAL AND ENTHALPY BALANCES IN POROUS CATALYST PELLETS, 110 1.1. Introduction, 110 11.2, Stoichtometric relations, 112 11.3. Material balance relations, 114 11.4, Dimensionless form of the material balances, 122 11.5. Pressure variations within a catalyst pellet, 128 11.6. Approximate neglect of pressure variations in the intermediate diffusion range, 132 11.7. The range of validity of the stoichiometric relations, 139 11.8. Problems with mora than one chemical reaction, 150 41.9. Enthalpy balance for non-isothermal pellets, 156 CHAPTER 12, MATERIAL AND ENTHALPY BALANCES IN UNSTEADY STATES, 159 12.1. General formulation of the balance equations, 159 12.2. Explicit balance equations for simple cases, 162 12.3. Dimensionless form of the equations, 168 12.4, The stability of steady states, 171 viii APPENDIX I, THERMAL TRANSPIRATION AND THERMAL DIFFUSION, 177 APPENDIX IL, THE EXPERIMENTS OF THOMAS GRAHAM, 186 REFERENCES, 190 AUTHOR INDEX, 193 SUBJECT INDEX, 194 Chapter 1, INTRODUCTION Reaction rates and selectivity in porous catalyst pellets are strongly influenced dy rate limitations associated with diffusion of reactants be- tween the surface of the pellet and active sites deep within the porous structure, and there is now a very extensive literature concerned with the behavior of this type of system. In industrial practice the reaction mix- tures almost invariably contain more than two substances, yet a large pro~ portion of the theoretical results te be found in the literature rest on expressions for the material fluxes within the pellet which are valid only in binary mixtures, Not infrequently additional assumptions are introduced, either explicitly or implicitly, which further limit the validity of the re- sults to a simple isomerization reaction, the limiting case of Knudsea diffusion, or some other unrealistic situation of quite restricted practical interest. In real catalyst pellets, on the other hand, multicomponent mixtures diffuse through a porous matrix composed of irregular channels of various sizes, ranging from micropores of dimensions much smaller than mean free path lengths in che reaction mixture, to macropores of dimensions substan~ tially greater than the mean free path lengths. To model diffusion in such systems it is necessary to use flux relations for multicomponent mixtures, valid throughout the entire range between the limits of Knudsen diffusion control ia very small pores and bulk diffusion control in large pores. hough the basis for this was established in the 19th century, it is only comparatively recently that detailed forms for the flux relations have been proposed and serious thought has been given to their use in describing the behavior of porous catalysts. The fruits of this thought are scattered widely but thinly in the literature of physical chemistry and chemical engi- neering, and the principal purpose of this monograph is to draw together material from the technical literature of the last hundred years or so and give a connected account of those aspects of transport in porous media which are important in modeling catalyst pellets. So little systematic information is available about transport in liquids, or strongly non-ideal gaseous mixtures, that attention will be jimitea throughout to the behavior of ideal gas mixtures. It is not intende thereby, to minimize the importance of non-ideal behavior in practice. we However, in the study of thermodynamics and transport phenomena, the behavior of ideal gases and gas mixtures has historically provided a norm against which their more unruly brethren could be measured, and a signpost to che systematic treatment of departures from ideality, In view of the complexity of transport phenomena in multicomponent mixtures a thorough understandiag of the behavior of ideal mixtures is certainly a prerequisite for any pro gress in understanding non-ideal systems. The starting point for any mathematical description of porous catalysts is a set of flux relations expressing the flux vectors of the various species in terms of gradients in pressure, temperature and composition. Provided the gradients are not too large in magnitude, these are invariably assumed to be linear and therefore have the following general form ac} ue a Gig QP. Terad x, + GS, @oP,Terad p + Go(x.p,T)grad T (1.1) where is the molar flux vector of species r, x, the mole fraction of this species in the mixture, and p and T denote pressure and temper- ature, respectively, The set of mole fractions is denoted by x , and each of the coefficients G., Sp, and Gy may depend on x, p, and T, as in- dicated, Thus the transport properties of a mixture of o species are Specified by n(a+1) coefficients, each of which depends on a+1 vari- ables. To evaluate these coefficients experimentally over the entire rele- vant ranges of pressure, temperature and composition would clearly require an experimental program of quite impracticable scope, even in mixtures with Ro more than three or four componeats. it is therefore necessary to invoke Some form of flux model, vhose purpose is to replace the general relations (1.1) by equations of more specific algebraic form, based on physical Teasoning and containing only a small number of adjustable parameters, ‘The values of these parameters can then be determined from a comparably small set of transport measurements in the medium or, preferably, from independent experimental or theoretical considerations. In this way the task of charac- terizing the transport laws experimentally is reduced to manageable dimen- Sions, Unfortunately, however, physically based flux models invariably lead to equations which are implicit in the fluxes, and of such a form that ex- plicit solutions are cumbersome to the point of impracticability for mixtures of more than two or three substances. ‘Thus the difficulty of experimental characterization associated with explicit, but empirical flux relations of the form (1.1) is traded for a difficulty of interpretation and application. Indeed, this is the main obstacle to theoretical progress in modeling the behavior of catalyst pellets with reaction mixtures of realistic complexicy. At the present time there exist no flux relations with a completely sound theoretical basis, capable of describing transport in porous media over the whole range of pressures or pore sizes. All involve empiricism to a greater or less degree, or are based on a physically unrealistic representa~ tion of the structure of the porous medium. Existing models fall into two main classes: in the first the medium is modeled as a network of inter- connected capillaries, while in the second it is represented by an assembly of stationary obstacles dispersed in che gas on a molecular scale. The first type of model is closely related to the physical structure of the medium, but its development is hampered by the lack of a solution to the problem of transport in a capillary whose diameter is comparable to mean free path lengths in the gas mixture. The second type of model is more tenuously re- lated to the real medium but more tractable theoretically, Flux models based on capillary networks are, of course, founded on transport laws for a single capillary. These are simple in form and can be derived by sound physical arguments when the capillary diamater is small compared with mean free path lengths in the gas mixture. This is the well kaown problem of Knudsen streaming, Though apparently much less widely re- alised, it is also possible to give a detailed derivation of the transport laws at the opposite limit, where the capillary diameter is very large com- pared with mean free path lengths and continuum theory is appropriate. This will be discussed in Chapter 4. For a capillary whose diameter is comparable with free path lengths the transport problem has so far defied solution, and it is necessary to resort to empirical or semi-empirical methods of interpo- lation between the known behavior at the limits of lerge and small diameters, The most convincing arguments of this type, which are based on momeatum- transfer considerations, are described in Chapter 2. The second type of approach to flux modeling, the so-called "dusty gas model," is developed in Chapter 3. In view of its completely different physical basis it is remarkable that its predictions are in complete agree- ment with those of the capillary model. it is worth remarking that the development of both types of model, like so many other aspects of the kinetic theory of gases, relies heavily on ideas of Clerk Maxwell. Some of these were rediscovered by later workers, but there is remarkably little that was not anticipated, at least in outline, by Maxwell. As noted earlier, the equations obtained from the models are implicit in the fluxes--a feature which they owe to the implicit form of the well known Stefan-Maxwell diffusion relations. I¢ is therefore not entirely a trivial matter to derive their physically important limiting forms or to obtain explicit solutions for the fluxes, and these matters are examined in detail in Chapter 5. The success of transport models must be measured by their ability to describe the results of flow and diffusion measurements in porous media. While a large number of measurements of this type can be made, there are certain experiments whose results are of such general importance that all flux models would be expected to conform with them. These experiments, which form the empirical basis for an understanding of transport in porous media, are described in Chapter 6. Just as the theory of transport in por- ous media owes much to ideas of Clerk Maxwell, the experimental basis is dominated by the work of another 19th ceatury scientist, Thomas Graham. Graham's work has been extensively misquoted, with the consequence that his results were largely ignored until their eventual rediscovery about a cen- tury later. However, E. A. Mason, who has been instrumental in the modern development of the dusty gas model, has actively drawn attention to the true nature and importance of Graham's contributions, which are here summarized briefly in Appendix Il. Within the class of capillary models it is possible to generate a great variety of flux relations, differing ia algebraic form and reflecting the nature of particular assumptions regarding the orientations, pattern of connections and size distribution of the pore system. Each set of assump- tions introduces certain parameters into the equations, and for practical reasons the structural assumptions must be such that the number of parameters is not too large. The development and classification of these models has unfortunately been less than systematic, acd some of its ramifications are explored in Chapter 8. At present we have only fragmentary information on which to base a judgment of the relative merits of different models. The evaluation of the available parameters of a particular model and subsequent testing of its predictions requires careful experimental planning, and few systematic investigations have so far been reported. Chapter 10 discusses the strategy of experimental testing, illustrates some of the pitfalls of superficial interpretation of the results, and describes one or two of the more noteworthy investigations in some detail. The final chapters, 11 and 12, are concerned with the particular appli- cation of transport theory to which this monograph is principally directed, namely the modeling of porous catalyst pellets. The behavior of a porous catalyst is described by differential equations odtained from material and enthalpy balances. Tn addition to pressure, temperature and composition, these contain vectors representing material and enthaipy fluxes, To re- duce them to closed form it is then necessary to relate the flux vectors to the local gradients in pressure, temperature and composition, and this is the role of the flux modeis. In principle, then, given the flux relations there is ao difficulty in constructing differential equations to describe the behavior of a catalyst pellet in steady or unsteady states. Ia practice, however, this simple procedure is obstructed by the implicit nature of the flux relations, since an explicit solution of usefully compact form is obtainable only for binary mixtures. In steady states this impasse is avoided by using certain re- lations between the flux vectors which are associated with the stoichiometry of the chemical reaction or reactions taking place in the pellet, and the major part of Chapter Ll is concerned with the derivation, application and limitetions of these stoichiometric relations. Fortunately they permit practicable solution procedures to be constructed regardless of the number of substances in the reaction mixture, provided there are only one or two stoichiometrically independent chemical reactions. In unsteady states the situation is less satisfactory, since stol- chiometric constraints need no longer be satisfied by the flux vectors. Consequently differential equations representing material balances can be constructed only for binary mixtures, where the flux relations can be solved explicitly for the flux vectors. This severely limits the scope of work on the dynamical equations and their principal field of application-- the theory of stability of steady states, The formulation of unsteady material and enthalpy balances is discussed in Chapter 12, which also ia- cludes a brief digressioa on stability problems. ” The equations developed in Chapters 11 and 12 are quite complicated and in many practical applications it may be desirable to replace them by simpler approximate equations. ‘The construction of such approximations, of adequate accuracy, is a field still largely unexplored. Nevertheless, it is important to understand the structure of the complete equations if approximations are to be made deliberately, rather than by inadvertent omission. Chapter 2. FLUX RELATIONS FROM SIMPLE MOMENTUM TRANSFER ARGUMENTS A form for the flux relations in porous media which accounts correctly for many phenomena can be obtained rather simply from momentum transfer con- siderations, These originate in treatments of the motion of gas mixtures given by Maxwell [L]*and Stefan (2} and in Kaudsen's (3] work on the flow of a rarified gas through a capillary. However, they also introduce as- sumptions which cannot entirely be justified within the framework of the argument, $o they are perhaps best regarded as a means of developing some physical feeling for the underlying phenomena. The type of treatment described here was originally introduced by Scott and Dullien {4], who confined attention to isothermal isobaric dif- fusion in binary mixtures, Similar equations were independently published shortly after by Rothfeld (5], and the method was later extended to multi-~ component mixtures by Silveston [6]. Perhaps the most complete exposition is given by Mason and Evans [7]. When a pure gas flows through a channel the accompanying fall in pressure is accounted for partly by acceleration of the flowing stream and partly by momentum transfer to the stationary walls. Since a porous medium may be regarded as an assembly of channels, similar considerations apply to flow through porous media, but in the diffusional situations of principal interest here accelerational pressure loss can usuaily be neglected. If more than one molecular species is present, we are also interested in the relative motions of the different species, so momentum transfers by colli- sions between different types of molecules are also important. To be more specific, consider a gas mixture which is in some form of steady axial motion in a tube, as indicated in Figure 2.1, By introducing Figure 2,1, Axial motion of a gas mixture in a tube, *AI] references ate to the bibliography on pages an ensemble of systems which are macroscopically identical, a velocity dis- tribution can be defined for each molecular species at any point P. Since the flow pattern is steady and the tube will be regarded as infinitely long, the mean values of the transverse components of velocity vanish, but each species r may have a non-vanishing mean value v, for its axial velocity component. In general v_ will depend on the location of P in the eross~ seetion and, at any given location, will take different values for different species, A molecule from ? will move unimpeded in a straght Line until it encounters either the wall of the tube, at a poiat such as W, or another molecule, at a point such as Q. In the Latter case, the molecule encoun- tered at Q may belong either to the same species rr, or to a different species $. Now encounters between molecules, or between a molecule and the wall are accompanied by momentum transfer, Thus if the wall acts as a diffuse reflector, molecules colliding with it lose all their axisl momentum on average, so such encounters directly change the axial momentum of each species. In an intermolecular collision there is a lateral transfer of mo- mentum to a different location in the cross-section, but there is also a net change in total momentum for species r if the molecule encountered belongs to a different species. Furthermore, though the total momentum of a particular species is conserved in collisions between pairs of molecules of this same species, the successive lateral transfers of momentum associ~ ated with a sequence of collisions may terminate in momentum transfer to the wall. ‘Thus there are three mechanisms by which a given species may lose momentum in the axial direction: (4) by direct transfer to the wall as a result of molecule-wall colli- sions, (44) by transfer to another species as a consequence of collisions be- tween pairs of unlike molecules, and (iit) by indirect transfer to the wall via a sequence of molecule-mole- cule collisions terminating in a molecule-wall collision. Clearly che general situation is very complicated, since all three mechanisms operate simultaneously and might be expected to interact in a complex manner. Indeed, this problem has never been solved rigorously, and the momentum transfer arguments we shall describe circumvent the difficulty by €irst considering three simple situations in which each of the three separate mechanisms in turn opetates alone. Im these circumstances the re- lations between Eluxes and composition and/or pressure gradients cam be found without too much difficulty, Rules of combination, which are essen— tially empirical guesses, are then proposed to construct the corresponding relations for situations in which all three mechanisms are simultaneously operative. When the diameter of the tube is small compared with molecular mean free path lengths in the gas mixture at the pressure and temperature of interest, molecule-wall collisions are mich more frequent than molecule- moleeule collisions, and the partial pressure gradient of each species is entirely determined by momentum transfer to the wall by mechanism (i). As shown by Kaudsen (3) it is not difficult to estimate the rate of momentum transfer in this case, and hence deduce the flux relations. For this purpose the shape of the tube's cross-section is not important; only its area A and perimeter P are needed. If a, denotes the mean sumber of molecules of species x par unit volume and ©, their mean speed, it is well known from elementary kinetic theory that rhe rate of incidence of molecules on unit area of the inner surface of the tube is given by 9,0, . Thea, if v, is the mean value of the axial component of velocity and m, the molecular mass, the momentum flux to the wall is Foon? per unit length. This also gives the aet rate of momentum transfer if we assume that the wall acts as a diffuse reflector, for then the mean axial momentum of the reflected molecules is zero. Neglecting acceleration we can equate the rate of transfer to the wall and the gradient in the axial momentum flux for species r, which is given by —Adp./dz where p, denotes the partial pressure and 2 is an axial co-ordinate. Thus we have «acmy SOR ery (2.1) But the molar flux of species 1 per unit cross-sectional area is given by Heo mM May 2.2) where 1%, is the Avogadra number, and the mean speed is well known to be given by . = (BL\e o "9 where M_ is the molecular weight, R denotes the molar gas constant and T is the absolute temperature, Using (2.2) and (2.3), equation (2.1) then *Strictly speaking, this expression is correct for a semi-infinite region bounded by a plane wall and containing a gas at rest. Here it is applied to a bounded region surrounded by a curved wall, and the molecules have a drift velocity parallel to the wall. Knudsen was concerned that this drift velocity might invalidate the treatment, but Pollard and Present {8} showed that this is not the case, reduces to vole ~dp = P ‘SRT 2.4 dz 7 4a Ga) Ce Ia isothermal conditions this can also be written in terms of the molar con- centration c,, since p,=cRT. Thus we obtain N= Di de /dz (2.5) where dD. is the Knudsen diffusion coefficient for species r in the tube, defined by : ma (8D DL” OF (fe) @.8 In particular, for a tube of circular cross-sectioa, this reduces to + a {8RT oe Cir) @.7) where a is the radius, This may be compared with a corresponding ex~ pressioa obtained from @ much more detailed and rigorous application of kinetic theory to a long capillary of circular cross-section {9], which gives _ 2a (RT e t Expressions (2.7) and (2.8) differ by a factor 3/8, or about 1,18, and this is typical of the order of error associated with simple momentum trans~ fer arguments.” The flux relation (2.5) can also be written in the alter~ native form (2.9) when the right hand side clearly represents the rate of momentum transfer to the wells per unit internal volume of the tube, for species r . A porous medium may be regarded as a complex assembly of interconnected channels and, if each of these is regarded as a capillary tube, one is led to speculate that a flux relation of the algebraic form (2.5) might be appropriate when all the channels have diameters much smaller than the mean TE, instead of assuming diffuse reflection at the wall, it is postulated that a fraction f£ of the incident molecules is scattered diffusely and the rest suffer specular reflection, the right hand side of equation (2.8) must be multiplied by a factor (2-£)/£ . 10 free path leagths. However, to give this speculation an unambiguous mean- ing, it is necessary to consider carefully just what is meent by @ flux vector of a concentration in this context. Clearly the flux vector defined at a point within the gas filled channels may vary widely in both magnitude and direction between adjacent chaaneis, and even from poiat co point within a given channel. But for practical purposes these Fluctuations on the scale of individual channels are not of primary interest, Rather we wish to con- sider a local average flux vector Ny , which determines the flux rate of species r across surface elements which are large compared with che dimensions of individual channels. Specifically, if 48 is an arbitrary element of area, with unit normal vy, the vector X is defined in such a way that the molar flow of species r through the element in the direction of y is given by y+ WAS, provided 48 is large compared with the channels but still small compared with the region of interest in the porous medium, Notice that X_, so defined, is a superficial average rather than an interstitial average, in the sense that it determines the flow per unit total cross-section, rather than the flow per unit void cross section. Local average values of concentrations and partial pressures, rather chan point values, are also used in the flux relations. However, in this case it is convenient to use interstitial local averages, based on the void volume rather than the total volume, since these are ote closely related to the properties of bulk gases. With this interpretation of the fluxes and concentrations, it is reasonable in view of (2.5) to expect that che flux relations in a porous medium will take the form _ pe ao DE grad c, (2.10) provided all channels are small compared with mean free path lengths. Here DE denotes an effective Knudsen diffusion coefficient for the medium, and is a sealar factor when the medium is isotropic. In anistropic media it is replaced dy a second order tensor, but we shall concentrate attention on isotropic media throughout. in view of (2.8) we might expect De to have the form L- at bo . «() , (ay where K, is a constant, characteristic of the scale and geonetry of the porous medium, which should be proportional to the mean pore diameter and to the void fraction. It should also contain a factor which takes account of the fact that the channels are not uniform cylinders, nor are they all aligned with the direction of mean flow, Thus K, = 2ea/31 (2,12) where ¢ is the void fraction and a@ the mean pore diameter, while 7 ts usually called a tortuosity factor. in a given porous medium , is prob- ably best determined by direct measurement of pressure driven flows (see Chapter 10). Consider next momentum transfer by mechanism (ii) above, with molecule of species r gaining or losing momentum ia collisions with molecules of other species. If we consider a gas mixture in the absence of any solid material belonging to a porous matrix, momentum transfer by mechanism (i) is certainly eliminated, since there are no solid walls present. Let us furthe assume that the mixture is at constant pressure and is uniformin the x,y co-ordinate directions of a set of rectangular axes, though there are compo” sition gradients in the z-direction. ‘then clearly there is no net transfer of momentum by intermolecular collisions in any direction normal to the g-axis, and mechanism (iii) above is eliminated, We have thus constructed a situation in which momentum traasfer occurs only as a result of mechanism (ii), and we attempt to relate the fluxes describing diffusion ia the a-direction to composition variations. The idea of treating diffusion ia such a system in terms of momentum transfer between unlike species appears to have originated with Maxwell in nis classic memoir on the dynamical theory of gases [1], and 2 similar treatment was subsequently given by Stefan {2}, who was apparently unaware of Maxwell's work. A rigorous kinetic theory treatmeat of diffusion is quite complicated and had to await methods developed by Enskog {10] aad Chapman (see Chapman and Cowling (11}) for solving the Boltzmann equation. In view of the difficulty of formal kinetic theory there have been numerous attempts to treat the phenomena of diffusion, viscosity and thermal conduc: tivity by simple "mean free path" arguments. The fruits of these can be found in many elementary textbooks, and they are quite successful in de~ scribing viscosity and thermal conduction in dilute gases. On the other hand, in the case of diffusion, they predict that the diffusion coefficien will be strongly composition dependent in mixtures of substances with wide differing molecular weights; a result quite at variance with the experimen facts. The much earlier arguments of Maxwell and Stefan, using momentum transfer, ate free of this defect. Since they are just as elementary as the "mean free path" theories, 2 it is therefore difficult to understand their neglect for so long. A de- tailed comparisoa of the two types of elementary theory has been given by Furry [12] and clearly demonstrates the superiority of the momentum transfer approach. A good exposition of the momentum transfer arguments is given by Present [9], but for cur purpose it is necessary oaly to quote the result, For a binary mixture this takes the form Tap, Px %_O¥ 7 YD (2.13) dz ay vhere p is the total pressure (assumad constant), p, is the partial pressure of species 1, x, and x, are the mole fractions of the two 1 2 1 and v, are the mean values of the 2-components of their molecular velocities, The symbol species, and v Aj. denotes the mutual diffusion co~ efficient of the two species and the theory gives an explicit expression for this.* Io line with our interpretation of equation (2.9) above, the right hand side of (2.13) must be regarded as the rate of momentum transfer, per unit volume, from molecules of species 1 to molecules of species 2 by inter- molecular collision. With this interpretation it is easy to guess the generalization to more than two species; there must be terms like the right hand side of equatioa (2.13) to represent momentum transfer from species 1 to each of the remaining species 2,3,....n separately, giving apy _. S px x, (y= ¥,) 4 go? is Since any species may be numbered 1, this may also be written in the more symmetric form (2,14) without any real gain in generality. The molar fluxes of the various Species are related to the velocities v_ by where c¢ is the total molar concentration, while the partial pressures and mole frections are related ty p, = x,p . Using these, and the fact that ‘Note the use of a script & for the binary pair mtual diffusion coefficient, as distinct from the Roman D already used to represent Knudsen diffusion coefficients. This convention will be adhered to throughout. 13 p {8 constant, equation (2.14) can then be rewritten in the following forms: (2.15) or (2,16) Equations (2.15) or (2.16) are the so-called Stefan-Maxwell relations for multicomponent diffusion, and we have seen that they are an almost obvious generalization of the corresponding result (2.13) for two components, once the right hand side of this has been identified physically as an inter- molecular momentum traasfer rate. In the case of two components equation (2.16) degenerates to Ni, - x,N = vebiy ax, fda (2,47) 1 Y where N=N,+N8,, aad this is just the familiar Fick equation. There are a Stefan-Maxwell relations in an n-component mixture, Dut they are not independent since each side of (2.16) yields zero on summing over r from 1 to a. Physically this is not surprising, since they describe only momentum exchange between pairs of species, and say nothing about the total momentum of the mixture. in order to complete the determin~ ation of the fluxes WN. Na» the Stefan-Maxwell relations must be supple~ L mented by an overall momentum balance. in the presence of solid boundaries this then leads to a result which depends on the rate of momentum transfer to the boundaries (see Chapter 4}, ‘The simple result quoted above gives no indication how thé equations should be modified to take iato account the in- fluence of gradients in temperature and total pressure, though Furry {123 shows that a similar approach can be used to explain the effect of tempera~ ture gradients, A complete kinetic theory treatment, based on the first Chapman-Enskog approximation to the solution of the Boltzmann equation, correctly represents the effects of pressure, composition and temperature gradients, and this will be used in Chapter 3 when discussing the “dusty gas" model. The Stefan-Maxwell equations have been presented for the case of a gas in the absence of a porous medium. However, in a porous medium whose pores are all wide compared with mean free path lengths it is reasonable to guess that the fluxes will still satisfy relations of the Stefan-Maxwell form Since intermolecular collisions st dominate molecule-wall collisions. wa However, we would expect the biaary paic diffusion coefficients to be re- placed by effective diffusion coefficients in the medium, so that -d9, aN - * Ne SE = Rr x es (2.18) str rs We would also expect that (2,19) where K, is a factor determined by the geometry of the pore structure only. K, should be independent of the pore size, provided this is large compared with mean free path lengths, but it should be proportional to the void fraction, and should reflect the fact that the available directions of molecular drift are constrained by the orientations of the pores. Thus it is common to find K, expressed in the form = eft (2.20) where 1 is a tortuosity factor determined by the statistics of pore orientations. Diffusion coefficients such as 8 and 4°, will be referred to as bulk diffusioa coefficients, as distinct from the Knudsen diffusion co- efficients introduced earlier. Finally we require a case in which mechanism (ifi) above dominates momentum transfer, In flow along a cylindrical tube, mechanism (i) is certainly insignificant compared with mechanism (iii) when the tube diameter is Large compared with mean free path lengths, and mechanism (ii) ean be eliminated completely by limiting attention to the flow of a pure substance. We then have the classical Poiseuille [13] problem, and for a tube of cir- cular cross-section aolution of the viscous flow equations gives aarp dp N= ByRT dz (2.21) where y is the viscosity of the gas. Ia a porous medium we might antici- pate a relation of similar form “Bp di Ng (2,22) where B, is 2 factor characteristic of the scale and geometry of the pore 15 structure, and known as the permeability of the medium. An equation of this form is usually attributed to d!Arcy [14] and known as “d‘Arcy’s law" com- paring equations (2.21) and (2.22), we see that Bo 2/8 (2.23) in the case of a single tube of circular cross-section, and in @ porous me- dium we would also expect B, £0 be proportional to the square of the pore diameter. When mixture of gases replaces the pure gas, but there is no relative diffusion of the separate species, the total flux is given by equation (2.22). The flux of each species separately is then given by BoP ay RT de (2.24) N Equations (2.10), (2.18) and (2.24) provide the flux relations in situations where each of the three separate mechanisms of momentum transfer dominates, However, there remains the problem of finding the flux relations in intermediate” situations where all three mechanisms may be of comparable importance, This has been discussed by Mason and Evans [7], who assumed first that the rates of momentum transfer due to mechanisms (i) and (ii) should be combined additively, If we write equation (2.10) in the form -dp, RTL dz > r The right hend side represents the rate of momentum transfer from species by mechanism (i) and, combining this with the rate of transfer by mech- anism (if) as given by equation (2.18), we obtaia x StS (2,25) which is a set of equations in the so-called “diffusive fluxes" NU. Jt is important to realize that these equations are independent, and there- fore serve to determine the unknowns N) uaiquely, despite the fact that only {na-1) of the Stefan-Maxwell relations (2.18) are independent. The physical reason for this is clear. As mentioned earlier, equations (2.18) do not contain aay mechanism for momentum transfer to the stationary walls, “DtArcy's work was published in a book with the unlikely title "Les fontanes publiques de Ia ville de Dijon." The porous media in question were filter beds through which the water for the fountains circulated. 16 but this shortcoming is removed in equatioas (2.25) by introduction of the terms RIN-/a® . Pe It remains only to take inte account viscous effects described by equations (2,24), and Mason and Evans assume that this can be done simply by adding a viscous flux x to each diffusive flux, giving nL amen (2.26) with the viscous fluxes determined by (2.27) as in equation (2.24). The procedure of Mason and Evans has the electrical analog shown in Figure 2.2, where voltages correspond to pressure gradients and currents to fluxes, As the argument stands there is no real justification for this pro- cedure; indeed, it seems improbable that the two mechanisms for diffusive momentum transfer will combine additively, without any interactive modifica- tion of their separate values. It is equally difficult to see why the effect of viscous velocity gradients can be accounted for simply by adding N rc oP, x dz (Kaudsen) Figure 2.2, Electrical analog of the procedure of Mason and Evans. ll viscous fluxes of the d'Arcy form to the diffusive fluxes, It is, there- fore, all the more remarkable that in isothermal systems the final flux relations, represented by equations (2.25), (2.26) and (2.27), are in com- plete agreement with the results of a more sophisticated analysis based on the dusty gas model, to be described in Chapter 3. 18 CHAPTER 3. THE DUSTY GAS MODEL Even in a cylindrical tube or a porous medium of equally regular geometry Lt is difficult to refine the simple momentum transfer arguments of Chapter 2 into a complete and internally consistent theory. As we have seen, a complete theory can be constructed when the tube diameter is small compared with mean free path lengths, and at the opposite limit of large tube diameter, where continuum theory can be used, it will be shown in Chapter 4 that a reasonably satisfactory description is also possible. How- ever, in the "intermediate" range, where the tube diameter is comparable with mean free path lengths, the molecular velocity distribution varies in a complicated way over the cross section, No solution is currently known for this general intermediate case, though some progress has been made by Pollard and Present [8] with the additional constraint of constant total pressure and, more recently, Hesse [15] has attempted a treatment using Oasager's principle of minimum energy dissipation. In the case of a porous medium the characteristic difficulties of the intermediate case can be circumvented by a device originally introduced by Maxwell [16] and Later exploited ia some detail by workers from Oak Ridge National Laboratory and the University of Maryland {17]+[21], The idea was stated very succinctly by Maxwell as follows: "We may suppose the action of the porous material to be similar to that of a aumber of particles, fixed in space and obstructing the motion of the particles of the moving systems”. In effect, the a gaseous species present in the diffusing mixture are supplemeated by an (n+1)°° “dummy species of very massive molecules, constrained by unspecified external forces to have zero drift velocity at all times. ‘The Chapman-Enskog kinetic theory is then applied to the pseudo- gas composed of the gaseous species and the (at+1)" dummy species, known as the "dust". Interaction of the gas molecules with the dust mole- cules simulates their interaction with the immobile solid matrix of a porous medium, but one is no longer faced with the difficult problem posed by flux and composition variations across the section of a pore, since the porous medium, in the guise of its representative dust, is dispersed throughout the gas oa a molecular scale. As a consequence of this simplification, certain physical features 19 of the veal porous medium are lost. For example, at the level of kinetic theory approximation used, the volume occupied by the dust particles is a very small fraction of the total volume, so the void fraction or porosity of the real porous medium has no physical analog in the model. Also, since there are no channels of finite side occupied only by gas, corresponding to the pores in the real medium, the formal development of the equations intro- duces no terms corresponding to viscous fluxes. As in the case of the simple theory presented ia Chapter 2, these must be added empirically, though Mason et al. [21] present arguments to justify the additive combina~ tion of diffusive and viscous contributions to the fluxes. Despite the fact that there are no analogs of void fraction or pore size in the model, by varying the proportion of dust particles dispersed among the gas molecules it is possible to move from 3 situation where most momentum transfer occurs in collisions between pairs of gas molecules, to one where the principal momentum transfer is between gas molecules and the dust. Thus one might hope to obtain at least a physically yeasonablé form for the flux relations, over the whole range from bulk diffusion to Knudsea streaming, The starting point for developing the model is the set of diffusion equations for a gas mixture in the presence of temperature, pressure and composition gradients, and under the influence of external forces.“ These take the following form grad p B.D with the molar fluxes Ne for the separate species related to the vectors 1 x by 1_yt N. 7 M7 BE grad(in 7). (3.2) Mw r Here w, denotes the mass fractioa of species rt in the mixture, x, is #The derivation of the flux relations given here is a simplified treatment which yields the correct algebraic form but does aot provide estimates of the various diffusion parameters. A more complete discussion is given by Mason et al. [21]. 20 its mole fraction and c, its molar concentration while E, is the external rT is the molecular weight of species r and the coefficients aL are thermal force (e.g., gravity) acting per mole on the molecules of species r. M diffusion coefficients for the respective species, The three terms on the right hand side of equation (3.1) represent ordinary diffusion (driven by composition gradients), pressure diffusion and forced diffusion, respectively, and the relative magnitudes of these terms are determined, since there are no adjustable parameters on the right hand side of this equation. In con- trast, the importance of thermal diffusion depends on the magnitudes of the thermal diffusion coefficients vr in equation (3.2). To assess the sig- nificance of thermal diffusion effects consider a binary mixture, and combine equations (3.1) and (3,2) to give T > XN) x DIIM, = 505M aed ” -at ee a grad(in T) (3.3) wR 2 where 1 = grad xy + 50° eE, “oS os (3.4) 3 and c is the total molar concentration, related to p and T by c= p/RI, Since and grad(1n T) each have dimensions of reciprocal length, the quantity in braces on the right hand side of equation (3.3) is a@ dimensionless measure of the relative importance of thermal diffusion and the other diffusion mechanisms. Values for this quantity, called the ther- mal diffusion ratio, are tabulated by Hirschfelder et al [22] and range between 107! and 107? . This suggests that thermal diffusion effects might be neglected without great error, s0 for the present we shall omit them, replacing the vectors xt by N. in equations (3.1), However, this approximation is by no means always justified, and Appendix I gives an account of the phenomena of thermal transpiration and thermal diffusion associated with the neglected terms. With slight differences in notation the equations presented above can be found in Chapter & of Hirschfelder et al [22]. To obtain equations describing the dusty gas model, equations (3.1) must be applied to a pseudo mixture of (nt1) species, in which the extra species, numbered n+l, represents the dust. We must also require (a) that the dust concentration ¢ be spatially uniform, ml (b) that the dust remain at rest, on average, so that Nii) () that Mi.) ~7®, cotresponding to massive dust particles. and In order to satisfy requirement (b) the-dust molecules must be constrained by external forces Physically these represent the forces exerted Ear 2 on the porous medium by whatever external agency prevents it from moving in response to gas pressure gradients, There are no external forces on the gaseous species, so FL #0 for r= lt Tt is important to bear in mind that the pressure, total concentration, mole fractions and mass fractions appearing in equations (3.1) will now refer to the pseudo mixture which includes the dust molecules, and not to the gas itself, This distinction will be maintained by writing p'. ef, xt ané ul for the quantities referred to the pseudo mixture, and p, Sy ¥,s%, for the same quantities referred to the gas. The latter, of course, are the quantities of physical interest. in the case of species concentrations c,, partial pressures p, and flux vectors N,, there is ao Gistinetion between the two, but we must write ay for the diffusion coefficients in the pseudo mixture. Writing out the first a terms on the left hand side of equations (3.1 separately from the (a+1)"" term, these equations take the following form for our pseudo mizture: oO xR + Rit xt OND Ss'r oss | thy eee s#r mm wl at = ad x + ( =)grad p+ cgeikaet * (c= ye. .0) G3.5) Now the foree per unit volume exerted on the porous medium by the pressure gradient in the gas is -gradp, where p, a8 distinet from Py is the physical pressure of the gaseous mixture. This is the force which must be balanced in our model by the external forces acting on the dust particles, so Coareaey = Brad P (3.6) It then remains only to translate from the primed variables associated with the pseudo mixture to the unprimed variables associated with the gas. For this purpose let us first combine the first two terms on the right hand side of (3.5), writing (3.7) We may also write pls ctRT = (ote, RT = pte, RT 3.8 and hence te grad p' + grad pte iB grad T (3.5) since c.,, is constant, Using this to evaluate grad p, equation (3.6) becomes CmetEey 7 Brad PT - EGR grad T (3.10) Again, from equation (3.8) and the fact that p_ = ¢ RT, we have ¢ = grad (c RT) 4 gead o +2% grad (in t) 3.1) L L =, grad p. = —>s pr Pp v {oto )R Using equations (3.7) - (3.11), equations (3.5) may now be rewritten in the form cy 8S he . 2 or wm SAL, sotl sft L ‘ + Rygrad c+ oF grad(la 7) Cay yR Brad T (3.12) where the mole fractions have been expressed as x! = ci/e' . Multiplying throughout by c!, and recalling that p'*c!RT and x =e /e, these be~ come a xN 7M, xt st Fs Sat " ». Se A, = grad c+ (©, Me, ))erad(in T) (3.43) rs rath s=1 str Furthermore 80 a3 Mou 2, wi—0, and the last term on the right hand side of (3.13) vanishes, We may then define = eet ; p= 4 zcthl de 5 0, - Bm 8 el! ne (3.14) and equations (3.13) simplify to 23 = + grad ¢, = ¢, grad (ia T) (3.15) The right hand sides of these equations may be expressed in a sumber of alter- native forms, Perhaps the most compact is obtained by writing ¢) = p_/RT. Then e+ d(ia 7) = a(2) 42s, ated grad grad c, +e, grad(ia 1) = grad{ae) +—b grad T= ge ered py RT so equations (3,15) become A «N+ 38 ose “Ss 28 1 rs ae (3.16) These are the flux relations associated with the dusty gas model. As explained above, they would be expected to predict only the diffusive con- tributions to the flux vectors, so they should be compared with equations (2,25) obtained from simple momentum transfer arguments. Equations (3,16) are then seen to be just the obvious vector generalization of the scalar equations (2.25), so the dusty gas model provides justification for the simple procedure of adding momentum transfer rates. As in Chapter 2 the complete fluxes must be constructed by adding viscous contributioas to the diffusive contributions predicted by equations (3.16), so we obtain finally Dey ye Sade 7s L WEES Eo ap seal e, (3.17) str rs r and x Bop we - oe grad p (3.18) with the complete flux vectors N given by D Vv Neth (3.199 In these equations, the constants les clearly represent the effective e c xepresent the effective Knudsen diffusion coefficients. Equations (3.14) do binary pair bulk diffusion coefficients in the porous medium, while D not provide any physical basis for estimating these quantities, bur it is interesting to note that they do indicate correctly the expected effect of 24 varying the dust concentration. As the dust concentration increases both the ratio c'/e and the mole fraction x increase, $0 the # increase " and the oe decrease, correspoading to a nove towards Kaudsen diffusion control. Conversely, as the dust concentration decreases, we move to the limit of bulk diffusion control. For the reasons expiained in Chapter 2 we might expect the diffusion coefficients to be given by equations of the form (2.11) and (2.19) a(R): 82, so we see that the dusty gas model contains three parameters characteristic of the structure of the porous medium, namely Kk, K, and B,. These must be determined by suitably designed experiments, as described in Chapter 10, The relation between the dusty gas model and the physical structure of a real porous medium is rather obscure. Since the dusty gas model does not even contain any explicit representation of the void fraction, it certainly cannot be adjusted to reflect features of the pore size distributions of different porous media. For example, porous catalysts often show a strongly bimodal pore size distribution, and their flux relations might be expected to reflect this, but the dusty gas model can respond only to changes in the three parameters K,, K), and B,. It can be argued that the distribution of free path lengths for collisions between gas molecules and dust particles simulates a distribution of pore sizes in the actual medium, but the form of the distribution of path lengths is determined by the principles of dilute gas kinetic theory, and is not available to model the real pore size dis~ tribution. Controversy on the physical significance of the dusty gas model is unlikely to be resolved but, regardless of this, it provides the simplest effective method presently available for representing flow and diffusion throughout the intermediate region between the extremes of Knudsen streaming and bulk diffusion with viscous flow. Chapter 5 will be devoted to exploring its algebraic structure and examining limiting cases of physical interest. Thermal transpiration and thermal diffusion effects have been neglected in developing the dusty gas model, and will be neglected throughout the rest of the text. The physics of these phenomena and the just{fication for neglecting them are discussed in some detail in Appendix I. 25 Chapter 4. GAS MOTION IN A LONG TUBE AT THE LIMIT OF BULK DIFFUSION AND VISCOUS FLOW tn Chapter 3 the dusty gas model was introduced to circumvent the difficwlty of solving the complete problem of flow and diffusion in a long cylindrical tube whose diameter is comparable with the gaseous wean free path lengths. We have earlier seen that a direct solution of the tube prob- lem is possible when the diameter is very small compared with the mean free paths, and it is reasonable to speculate that a solution may also be possible at the opposite limit of large diameter, where a continuum description of the gas motion is appropriate. ‘The influence of the tube walls is then taken imto account through boundary conditions on the continuum variables, A familiar example of this approach is the Poiseuille solution for laminar flow through a tube under the influence of @ pressure gradient, and the non~ slip condition at the well then provides the required boundary condition. The complete problem with composition gradients as well as a pressure gradient, may be regarded as a "generalized Poiseuille problem", and its solution would be valuable for comparison with the limiting form of the dusty gas model for small] dust concentrations. Indeed, it is the “large diameter” counterpart of the Knudsen solution in tubes of small diameter. It also has interesting historical roots and, as we shall see, it raises fundamental questions regarding the boundary conditions at the wall, which apparently are not widely understood. We shall therefore investigate what is involved in formulating and solving this problem. The continuum description of diffusion ia a multicomponent gas mixture is provided by equations (3.1) and (3.2). In the absence of a temperature gradient these may be condensed to the following form (1 is the molecular mean velocity for species rc, related to the flux M by (4.2) 26 Only o-1 of the o equations (4.1) are independent, since both sides vanish on suming over x, so a further relation between the velocity vectors vw, is required. It is provided by the overall momentum dalance for the mixture, and a weil known result of dilute gas kinetic theory shows that this takes the form of the Navier-Stokes equation L 2 ie tye | = ptt wary) + uy, (4.3) et written in terms of the mass mean velocity Sy defined by ao ore : (4,4) r Here p denotes the density of the mixture and 9, the mass of species r per unit volume, while A and wp are the bulk and shear viscosities, respectively. Consider now the problem of steady motion in an infinitely long cylindrical tube of circular cross-section and radius a , and let (1,2) denote cylindrical coordinates about the tube axis, Since yu satisfies the Navier-Stokes equation, if we also impose the non-slip boundary condi - tion y,+0 at the tube wall, we simply obtain the Poiseuille solution. Thus p is independent of r and v, is axially directed, vith magnitude given by vate) = > ag BG ED 4.8) Thea, knowing ¥., one relation between the individual species velocities is provided by equation (4.4), and this can be used in conjunction with the“ n-1 independent equations (4.1) to determine all the ye and hence complete the solution, Hovever, before embarking on this procedure let us examine the particular case in which there is so pressure gradient, The Poiseuille solution (4,5) then reduces to Yn FO and it follows, of course, that the mass flux through the tube vanishes. But, as we shall see in Chapter 6, this is completely at variance with experimental experience. In the case of a binary mixture, for example, it is very well established that the isobaric molar fluxes satisfy Graham's famous relation or =0 4.8) 27 rather than the condition of zero mass flux NM) + NM, = 0 Clearly then, the continuum approach as outlined above is faulty. Further- more, since our erroneous result depends only on the non-slip boundary con dition for vy) together with the Navier-Stokes equation (4.3), one or the other of these must be in error, There is no reason to suspect vw, of not satisfying the Navier-Stokes equation, since this is firmly grounded in dilute gas kinetic theory. Instead we need to re-examine our demand that the mass mean velocity should vanish at the wall, as this is not a uaique generalization of the non-slip condition for a pure fluid. The basis for the familiar non-slip boundary condition is a kinetic theory argument originally presented by Maxwell [23]. For a pure gas Maxwell showed that the tangential velocity v and its derivative normal to a plane solid surface should be related by 4D where x is a coordinate normal to the surface and 2 {2 G <3 (2-1 (4,8) Here £ denotes the fraction of molecules diffusely scattered at the surface and £ is the mean free path. If distance is measured on a scale whose unit is comparable with the dimensions of the flow channel and vw is sone suitable characteristic fluid velocity, such as the center-line velocity, then dv/dx qv? and £<<1, Provided a significant proportion of incident molecules are scattered diffusely at the wall, so that f is hot too small, it then follows from (4.8) that G< oY 0 (4.17) © Equations (4.16) and (4.17) can therefore be solved for the diffusion velocities and, siace each of the variables x,, % and p depends oaly on 2, the u, are clearly independent of radial position, Knowing the u,, boundary condition (4.14) provides the value of v,(a), which may then be used in equation (4.15) to generate the complete radial profile of the mass mean velocity. The individual species velocities then follow from equations (4,12) and the solution is formally complete. However, it is not quite what we want, For practical purposes we need to translate the result jato relations betweea the cross-sectional average values x, of the species fluxes. These relations can thea be compared with the buik diffusion limit BL of the dusty gas model as a check on the consistency of the two different approaches to the flux relations. The cross-sectional average molar fluxes are defined by a ge iit) #4 IEE or er ty, (r)det cpa, since ¢, and u, are independent of radial position. Using equation (4.15) for v(x), this becomes x a do 7 OX v.68 - Bn ds * a (4.18) Now consider the cross-sectional average N of the total moler flux and the cross-sectional average diffusion fluxes J,, defined by (4.19) (4.20) Using equation (4.20) it is then easy to show that xo ¥ J, _ xt, (@, 745) 3 s om Fs S rs ox, making use of equations (4.16) 5 8S, dx x 8 sir7*es op Se re} dp way a at dz Rt | dz . 5 rs gut from the definition of the diffusion fluxes given above it follows immediately that 32 (4.22) and equations (4.21) and (4.22) thea provide a independent relations which may be solved for the diffusion fluxes, Expressing the mass fractions w, in terms of mole fractions, equations (4.21) can aiso be written in the equivalent form xJp 7 ed, dx, x For "es. 2-5. > Re RT dz et (4.23) s Le remains to determine the total flux N by introducing into equation (4.19) the value of v,(a) given by the boundary condition (4.14), Thus we obtain 2 2 2 a (uo 8) i pa” dp. =: + (4.24) uRT dz Se is But equation (4.20) shows chet xpJp Xp, = OX Ry CU, 7 Uy) sor rs 3 and using this on the right hand side of equation (4.24) gives wy BDA eFe-*T an The summation on the right hand side of this can be simplified by recalling that x = 0, with the final result chet r (4.25) Once the a, and § have been found from equations (4.21), (4.22) and (4.25), ail the fluxes x, are, of course, determined, Thus we now have 33 the complete flux relations for the isothermal sotion of a gaseous mixture in a long tube, at the Limit where all mean free paths are very short com- pared with the diameter of the tube, The treatment of the boundary conditions given here is @ generaliza- tion to multicomponent mixtures of a result originally obtained for @ bi- nary mixture by Xraners and Kistemaker [25].” These authors also obtained results equivaient to the binary special case of our equations (4.21) and (4.25), and integrated their equations to calculate the pressure drop which accompanies equimolar counterdiffusion in @ capillary. their resules, and the important accompanying experimental measurements , will be discussed in Chapter 6+ Eganeve and Kistemaker evaluated the momentum transfer to che wall by nolecular impacts in a Frame of reference moving with the mole mean velocity Teiertae some algebraic advantage over working in the rest frame of the tube M4 Chapter 5. AL RALCG MANEPULATLONS AND LIMITING FORMS OF THE DUSTY GAS MODEL EQUATIONS The dusty gaa model flux equations (2.17) - (3,19) describe flow and diffusion over the complete range of relative values for the pore diameters and the mean free path lengtas. Their algebraic form is rather complicated, so it is important to identify physical situations in which they can be solved explicitly for the fluxes, or at least approximated by simpler equa- tions. OE particular interest are the limiting forms associated with mean free path lengths which are short and long, vespectively, compared with the pore diameters, We shali, therefore, first derive several equivaient forms of the complete equations which are merely algebraic rearrangements of each other, Next we shall examine with some cate those limiting situations in which the flux relations can be approximated by simpler equations, and final- ly we shall solve the equations for the few cases in which an explicit solu- tion is practicable. Thermal diffusion effects will be neglected throughout, so the flux relations are given by equations (3.17) - (3.19), which are repeated here for convenience; Gu) (3.2) Da Koay 6.3) These can easily be condensed into a single set of a equations for the flux vectors x From (5.2) and (5.3) Bip 2) . db a etade and using this in equation (5.1), together with the substitution p= px,, we obtain 35 N x N- sN. SP ragx & = ppared x, (5.4) This is the most compact form for the complete flux relations. In Chapter 4 we introduced the total Zlux N and the diffusion fluxes » related to the N_ by © tr (5.3) The J, and N together forn a set of a independant vectors, since ve =o, ¥ and there is a one-to-one relation between this set and the flux vectors x Thus any diffusional situation can be described equally well ia terns of the Nj» oF in terms of Nand the J., and it is consequently useful to rewrite the flux equations in terns of this alternative set of vectors. From (5.2), (5.3) and (5.5) x, 3p + — SL tS + gp op ered? and, using this on the left hand side, equation (5.1) can be written L yp BoP x sk z +N +a Brad? +> = => pp ered PL (5.6) r stt ts On summing over r from i to n and solving for N, we then obtain Se pe s/s grad p (3.7) L Sx Joe, s/s 5 This expression for N may aw be substituted back into the left hand side of equation (5.6) and, after some rearrangement, we obtain . joe x J x uses =. Be Avs RT grad p (S.8) x rt where A, is defined by 34 .9) Equations (5,8) are not all independent, of course, since both sides vanish = 9 they can on summing over r, but when supplemented by the condition >. r be solved for the J.. Having found these, N then follows from equation (5.7) and all the flux vectors are determined, Thus equations (5.7) and (5.8) pro- vide the required alternative formulation of the dusty gas model. Yet another form for the equations, this time expressed in terms of the Ry» san be obtained by sudsticuting for Nand the J, from equations (5.5) inco (5,7) and (5.8), After a little rearrangement the result can be written Bp wx, i+ + = )eradp (5.10) Foe and # N+ x LN. x ahr” “ees *r 2. te =~ gp radx_ - ap grad p (5.11) Equations (5.11), of which only n-i are independent, may be regarded as a generalization of the Stefan-Maxwell relations, Let us now tura attention to situations in which the flux equations can be replaced by simpler limiting forms, Consider first che limiting case of "dilute solutions" where one species, present in considerable excess, is regarded as a solvent and the remaining species as solutes, This is the sim- plest Limiting case, since it does not involve any examination of the relative behavior of the permeability and the bulk and Knudsen diffusion coefficients. veri, Then for rén If species n is the solvent, x,~1 and x,<<1 for r= equations (5.1) reduce approximately to Pf, t\oi4 Sips te ls - ge ered, (5.12) BD a. ao The solute species therefore diffuse independently, rather as in Knudsen dif- fusion, but with effective diffusion coefficients Ds where dete t (5.43) r a) 3 aa a relation attributed to then obtained as the sum equations (5.2) and (5.3) o evo geared yy . ~ OE grad x at Brae x Here utes are present only in The limiting cases which the mean free path with the pore dianerers, 37 Bosanquet {26], The total fiux of species r is of the viscous and diffusive contribucions, Using grad p x Bop ~ gpl Olt sm jerade (5.14) RTP ru, the viscosity of the gas mixture has been taken to be p, since the sol- low concentration. of greatest interest correspond to conditions in lengths are large and small, respectively, compared Recall from the discussion in Chapter 3 that che effective Knudsen diffusion coefficients Do are proportional to pore diameter and independent of pressure, while che effective bulk diffusion coefficients 82, are independent of pore diameter and inversely proportional to pressure. Thus, when the pore diameters are sufficiently smail or the pressure suffi- ciently low, 0° << 32), for all x, s and t, This is the Limit of Knudsen streaming and the first term on the left hand side of equation (5.1) is clea: ly dominant in these circumstances, giving aS DE Er 3t2d P, & Hovever, without further examination of the relative magnitudes of oe as and B,, it is not immediately obvious that xu can be neglected, At che opposite limit of large pore diameter or high pressure, See << of for all r, s and t, and consequentiy the coefficients of the terms in the summation over s on the lef: hand side of equation (5.1) are all large compared with 1/Dr. Ie is therefore tempting simply to drop the term N0/O? from the left hand side at this limic, Nevertheless, such a procedure is quite incorrect, since the remaining terns on the left hand sides of equations (5,1) vanish on sum ming over r, while in general the right hand sides do not. Thus an incon~ sistent set of equations has been generated by destroying the independence of the left hand sides. From what has been said, it is clear that both physical and mathemati- cal aspects of the limiting processes require more careful examination, and we will start this by examining the relative values of the various diffusion coefficients and the permeability, paying particular attention to their deper dence on pore diameter and pressure. 38 The Knudsen diffusion coefficients are given by equations (2.11) ia which K, is independent of pressure and proportional to pore diameter 2, so that we can write be = ad (5.15) where >. is independent of pore diameter and pressure. Similarly, since the factor K, depends only on pore geometry and void fraction, it follows from 1 equation (2.19) that we can write 3, e . rs a6 (5.16) with By independent of pore diameter and pressure, Finally, since the per- meability B| is proportional to a” and independent of pressure, we can write B= ai (5.17) with B, once again, independent of pore diameter and pressure. The limiting form of the flux equations for large pore diameters or high pressure is best approached starting from equations (5.7) and (5.8). Let us first examine the factor 4. given by equation (5.9), The ratio of the two terms on the right hand side of this equation is reSat a,x, : 2 = des as Since the factor in brackets does not depend on a” or p it is clear that os i e rs vit > x ne seat whenever a and/or p is sufficiently large, then 6, may be replaced by Bes in equation (5.8) with negligible error, giving S tebe Mle 3 = (5.18) 8. ser rs This determines the diffusion fluxes at the limit of bulk diffusion control. a0 Now consider the ratio of the two terms within the brackets on the ripht hand side of equation (5.7) In gases at moderate pressures the viscosity is almost independent of pressure so the factor in brackets on the right hand side does not depend on a or p, and it follows that Bop 2 >> 4 oe L Sx, /0 v whenever a and/or p is sufficiently large. Then the second term within the brackets may be neglected on the right hand side of equation (5.7), which reduces to ye ~ See stad (3.19) 28% This determines the total flux at the limit of viscous flow. Equations (5.18 and (5.19) theregore describe the limiting form of the dusty gas model for high pressure or large pore diameters -- the limit of bulk diffusion control and viscous Flow. At this point it is important to emphasize that, by changing a and p, it is not possible to pass to the limit of viscous flow without simulta~ neously passing to the limit of bulk diffusion control, and vice versa, since physical estimates of the relative magnitudes of the factors D Bee and B show chac — 8. rs e “Ss e BD, xD, ; whenever 40 Day s (See section 11.4). A different limiting form of the dusty gas model for bulk diffusion control and viscous flow may also be obtained quite easily starting from equations (5.10) and (5.11), As we have seen gt Bee. Ses 7 Meg Bnd 2s DP “ be ror act this Limit, so equations (5.10) and (3.11) reduce to (5.20) and (5.21) respectively, It is easy to check that equations (5.20) and (5.21) are entirely equivalent to (5,18) and (5.19), from which they can be obtained by the substitutions: There are n-1 independent equations of type (5.21), since both sides vanish on summing over x, Equation (5,20) then provides one furcher relation between the fluxes to complete the set. There is a further simplification which is often justifiable, but not by consideration of the flux equations above. The nature of many problems is such that, when the permeability becomes large, pressure gradients become very small ieuier cndu liuses oecomng very large. in catalyst peliets, ror example, reaction rates limit the attainable values of the fluxes, and it then follows from equation (5,19) that gradp ~O as Bl~=, But then the second term on the right hand side of equation (5.18) becomes negligible, and this equation may be replaced by Al x Jo - «J Astx “s 2. 2 praax (5.22) & = RT < r3 This simplification must be used with caution, of course, making sure that the specification of the problem does not determine the magnitude of the pressure gradient, but ic is very useful in the important case of a coarsely porous catalyst pellet. It may seem curious that Knudsen diffusion coefficients still appear in equations (5.18) and (5.19), which supposedly give the flux relations at the limit of bulk diffusion control, However, inspection reveals that oaly ratios of these coefficients are effectively present, and from equation (2,11) it follows that ee o = p/ay = Ot /M) Using this, (5.18) an¢ (3,19) may be written ia the form grad p 5.23) aad y= - — - 2 gradp (5,24) eliminating the apparent dependence on the Knudsen coefficients. The opposite limiting form of the general flux relations, for small values of a and/or p, is most simply obtained starting from equation (5.4), At this limit it follows from equations (5.15), (5.16) and (5.17) that BoP +4 >»> + and << 1 a x *ts v, Thus thé first term dominates the left hand side of equation (5.4) and oa the right hand side the factor in brackets may be replaced by unity, so that the equation reduces to e x =~ 2 grad B. = 7 gr 88d 2, (5.25) 42 as anticipated, At the Limit of Knudsen streaming the flux relations (5.25) derexmine che fluxes explicitly in terms of partial pressure gradients, but the general flux relations (5.4) are implicit in the fluxes and their solution does not have an algebraically simple explicit form for an arbitrary number of com- ponents, It is therefore important to identify che few cases in which rea- sonabdly compact explicit solutions can be obtained. For a binary mixture, simultaneous solution of the two flux equations (5.4) is straightforward, and the result is important because most experimental work on flow and dif- fusion in porous media has been confined to pure substances or binary mix- tures, The flux vectors are found to be given by ups N 2 grad x Ph Me PEP x Uv) + sty Bp 3 r yt, to grad p (5,26) ee "elt Pep FahPL Pe and Ny (5.27) which are, of course, interchangeable by interchange of the suffixes 1 and 2. However, the reader should be aware that the dusty gas flux relations for a binary mixture appear in the literature in a variety of algebraic disguises, which can be very confusing. For example, in the isobaric case some authors use the algebraic form originally given by Scott and Dullien [4], namely (5.28) 43 N where g = 1+". Nowever, a brief inspection reveals thar this is nothing more than wt NxM, 7 4,8 Ay Meh he Le grads oe 2 RE 1 1 12 which is just equation (5.4) applied to an isobaric binary mixture. Consideration of the behavior of equations (5.26) and (5,27) when the pressure or the pore diameter becomes large illustrates the care which must be exercised in passing to this limit, Multiplying numerator and denominator in equation (5.26) by Deel gives e Sfp arad x, 2 RT fe @ Sip Nabe ss e+ %2 yy 1 e x D, Bo -a ak 4 2 | grad p ie » x, D, v2 @ 2 4 But also ac this limit, Bjp/y >> Dj, and the first term in square brackets may be neglected compared with the second, giving AOE course, ehis form only has meaning in one-dimensional diffusion, where Xj, N, and gradx, are collinear, and the ratio N)/K, can be interpreted onambiguous ly. ab e a x Be i ~ 22 - oe XN # grad, > gp. —$° stade (5.29) Palye * Je 1 72 with a corresponding expression for N, obtained by interchanging the suffixes Lan 2, However, if (5.29) and its companion equation for N are used to calculate the diffusion flux J,, we find =~ da}, eradx, (5,30) ‘which contains ao terms represeating pressure diffusion. Such terms never- theless should be present, as can be seen by writing out the general limit- ing equation (5.18) above for the special case of a binary mixture, with the vesulc Lod ee xX. DSO. 2 se _ a2 .e {Po Ph T By. Brad x, ar Fix 7% parade (5.32) 1 2 waar ad to be compared with (5.30), The coefficient of grad p on the right hand side of equation (5.31) is not small compared with the coefficient of gradx, at the bulk diffusion limit, The correct expression for N) at this limit caa be found from equations (5.18) and (5,19), specialized to the case of 2 binary mixture, and it is 2-2 an 2] tL e sty 2 Bt xy Bo 2 an erat pay Sead (5.32) of, % 2 eae 1 22 There is no doubt that the coefficient of the chird term oa the right hand side of this equation is much larger than the coefficient of the second term at the bulk diffusion limit, and this justifies our original form (5.29). However, on constructing J, from N, and Ny, using equation (5,32) and its AS companion for Ny, the individually dominant terms containing B, cancel out, and it is the other terms in grad p which generate the pressure diffusion tera in (5.31). In problems where 3, is very large, and aa a consequence grad p+, the use of (5.29) is, of course, fully justified. Explicit expressions for the fluxes can also be found in the case of a ternary mixture, though they are appreciably more complicated than those for a binary mixture, The best starting point is equatioas (5.7) and (5.8). when there are three components in the mixture it is easy to check that equa- tions (5.8) and che condition yee = 0 are satisfied by F a(S 3) S93 N23 831 Aaa By oh5 8. (5.33) 12°23°3Y Xyhog + Xobay + Soho and the two equations obtained from this by cyclic interchange of the suf- fixes 1, 2, 3. Here the factors 4. are given by equations (5.5), and x. = RE A, & gp grad a+ (5.34) As "RT Le with corresponding equations for N, and Ny. Using (5,7) for N, this can be written, (5.35) which expresses N, in terms of the diffusion fluxes, Using equation (5.33) and its two companions to evaluate the vectors i, on the right hand side of equation (5.35), we finally obtain, after some reatrangement 46 Syohoatsy Oryhy3 7 Xb a1 * X3812? (5.36) This equation, together with the corresponding results obtained from it by cyclic permutation of the suffixes 1, 2, and 3, provides the required explicit form of the flux relations. For more than three components extremely heavy algebra is generated in attempting to solve the implicit flux relations, and in general no usefully compact explicit solution is obtained, ‘However, there are two interesting special cases in which explicit flux relations caa be obsained with an arbi- trary number of components in the mixture, Neither would be expected to correspond accurately with physical situations of practical interest, but they may provide useful qualitative, or semi-quantitative pointers to the dehavior of more accurate nodels. the first case corresponds to a situation im which all Knudsen diffu- sion coefficients are equal, and all binary pair bulk diffusion coefficients are equal: f=, gt = p= 0; ar .= 9 G@lieys) (6.37) These are conditions which could be satisfied, even approximately, only in a mixture of isomers of rather similar structure. Then the general dusty gas flux equations (5.4) reduce to 1,3) %8 oe a “ety BoP x +f) - S++ Baran, - Gli +h lace (5.38) Summing over all r, we obtain BP a oa H+ +3] eer (5.39) and substitutiag for N from this into equation (5,38) gives 4) p BM, Ry Bop - ot gn [0 + — | aad (5.40) = (E43) RT py rade Dy 3 which is the required solution. the second case follows froa a suggestion of Bird [27] to the effect that binary pair bulk diffusion coefficients might be approximated by expres- sions of the forn a a 6.4) where © is a function of temperature and pressure characteristic of the particular substances forming the mixture but, at least over Limited ranges, independent of their proportions ia the mixcure, while F, is a factor char acteristic of substance r and dependent only on temperature. As 2 conse~ quence of equasions (5.41) and (2.19) we can also write (say) (5.42) in a porous medium, With this form for the coefficients ory we will show that it is possible to obtain an explicit solution of the large pore (or high pressure) limiting form of the flux relations, represented by equations (3.19) and (5.22). Using equation (5.42) in equation (5,22): e ps = - oe grad x, *F ep Teds Fle tse (5.43) 3 = Dividing both sides of this by > suming over r, and usiag the fact chat oo 0, it follows that pe, Dias d grad x, and this may be substituted back into the right hand side of equation (5.43), which can then be solved for te giving (5.44) 43 This is an explicit solution of the Stefan-Maxwell equations for the diffu- sion fluxes, The species flux vectors are then given by Substituting for the J, from equation (5.44) and simplifying then gives ye x, < © grad x, ofr. x Bop 2-8 iT. 888 \_ ol 4, RT S. ~ F * s € Br J” etade G45) x_F r x /D ss ‘s/s which is the required explicit solution for the fluxes, Finally, before leaving our exploration of the dusty gas model, we must compare the large pore (or high pressure) Limiting form of its flux relations with the corresponding results derived in Chapter 4 by detailed solution of the continuum equations in a leng capillary. The relevant equa- tions are (4.23) and (4,25), to be compared with the corresponding scalar forms of equations (5.23) and (5.24). Equations (4.25) and (5.24), are seen to be identical, while (4.23) and (5.23) differ only in the pressure diffu- sion tern, which takes the form in (5.23), Frequently (as in most problems of catalyst modeling) the pressure 4g diffusion terms are acgligibie at the limit of bulk diffusion control, where pressure gradients are small, Thus the difference between the two sets of equations may aot be of great practical importance, though it is a signifi- cant point of principle, Indeed, it is gratifying that two so dissimilar aoproaches to the Elux relations should yleld such closely similar results. 53 Chapter 6, SOME IMPORTANT EXPERIMENTAL RESULTS ON GAS MOTTON IN POROUS MEDIA AND CAPILLARIES Proposed flux models for porous media invariably contain adjustable parameters whose values must be determined from suitably designed flow or diffusion measurements, and further measurements may be made to test the relative success of different models, This may involve extensive programs of experimentation, and the planning and interpretation of such work forms the topic of Chapter 10, However, there is in addition a relatively small number of experiments of historic importance which establish certain general features of flow and diffusion in porous media. These provide criteria which must be satisfied by any proposed flux model and are therefore of central importance in the subject. They may be grouped into three classes. In the first, motion is generated by composition gradients in the absence of any temperature or pressure gradient, in the second, motion is a conse- quence of pressure gradients alone, and in the third,both pressure and composition gradients are present but a constraint of zero aet molar flux is imposed. These will be discussed in turn. In the first class the earliest, and still some of the most extensive and careful experiments reported, were those of Thomas Graham in the middle of the 19th century, Graham devoted many years to the study of various aspects of the flow of gases and gas mixtures through porous plugs, capillary tubes and orifices, but of principal interest for our present purpose is his earliest work [28] on the interdiffusion of pairs of gases through porous plugs of stucco. A brief review of the remainder of his work is given in Appendix IT. His apparatus, which was extremely simple, is sketched in Figure 6.1, A glass cylinder containing a test gas B, closed at its upper end by a stucco plug, is inverted over @ liquid seal in which the gas B is insol- uble (Graham used mercury). The liquid surfaces inside and outside the tube are level initially, 0a exposing the apparatus to air (A), the test gas B diffuses out through che plug and air diffuses in to replace it until the tube contains only air, At the same time the level of Liquid within the tube changes in response to the difference between che volume of B diffusing out and the volume of air diffusing in. Thus these two volumes are both known when the final steady state is reached, Using a variety of SL gases B, Graham reached the simple but important conclusion that the ratio of the volumes diffusing in and out is the inverse of the ratio of the square roots of the densities of the two gases. Figure 6.1, Graham's diffusion apparatus The experiment can obviously be criticized on two grounds; first, it does not determine the instantaneous fluxes, but rather their integrals over the period of the experiment and, second, conditions are not strictly isobaric because of the difference in head of mercury which develops as diffusion proceeds. Graham was aware of the problem of interpretation posed by the developing pressure difference; indeed he remarks that his conclusion is "distorted" when the porous structure of the plug is too loose, and correctly attributes the "distortion" to the influence of bulk flow driven by the pressure gradient. ‘The difficulty is eliminated, of course, if the height of the tube is continually adjusted so that the mercury surfaces in- side and outside remain level. Because of its integral nature, Graham's experiment does not prove that the ratio of the instantaneous fluxes of the two gases is at all times the inverse ratio of the square roots of their densities, but at least it is not inconsistent with such a hypothesis. Since the pressures and temperatures are the same in all his experiments, and the pressures are sufficiently Low that the gases behave ideally, volume flow is proportional to molar flow and density is proportional to moleculer weight, and this hypothesis can therefore be expressed in the form 1 M+ 8, vi, .D It is aot surprising that such a relation should hold at the limit of Knudsen diffusion, since the Knudsen diffusion coefficients are themselves inversely proportional to the square roots of molecular weights, but the pore diameters in Graham's stucco plugs were certainly many times larger than the gaseous meaa free path lengths at the experimental conditions. Thus his experiments were the first to indicate the surprising result that yelation (6.1) remains valid even in conditions where bulk diffusion re- sistance is completely dominant. Accordingly (6.1), perhaps the most im- portant single experimental result on diffusion in porous media, will be veferred to as Graham's relation. Though well known to Graham's immediate successors, this striking re- tation later appears to have been lost sight of, and there are repeated assertions in the literature that equimolar counterdiffusion occurs under isobaric conditions in the bulk diffusion regime. ‘This assumption of equi- molar counterdiffusion wes challenged in 1953 by Hoogschagen [29], [30], who was apparently unaware of Graham's work. His apparatus, which is sketched in Figure 6.2, was designed for steady state operation under com- pletely isobaric conditions, thereby eliminating the objections which might be raised against Graham's work. Measurements were made on the inter- diffusion of oxygen and the "inert" gases He, Ng, and co,. Referring to Figure 6.2, a stream of oxygen at atmospheric pressure is passed across the lower face of the porous plug, and inert gas circulates past its upper face driven by natural convection in the closed Loop of tubing above the plug. Oxygen diffusing upwards is removed from the circulating gas by reaction with the heated copper, while inert gas diffusing downwards is replaced from the gas burette, whose moving reservoir is adjusted at all times to maintain atmospheric pressure within the graduated cylinder. The flux of inert gas can then be obtained directly from the change of Liquid level in the gas burette, and the flux of oxygen can be deduced from the gain in weight of the heated copper. Hoogschagen made his porous plug by compressing granules in the size range 1 = 84, which should give pores much larger than the gaseous mean free path lengths at atiospheric pressure. It was confirmed that this was, indeed, the case by noting that the flow of pure nitrogen through the plug uader a given total pressure difference greatly exceeded its diffusion flux under the same partial pressure difference when mixed with a second gas. Though the results obtained were relatively few in number, they con- formed fully to Graham's relation within the experimental accuracy. Subd Sequently further evidence has been provided by the accurate experiments of 53 inert gas FE] porous plug Figure 6.2, Apparatus of Hoogschagen (From Hoogschagen [29}) Wicke and Hugo [31] and by other measurements including those of Wakao and Smith [32], Rothfeld [5], and Remick and Geankoplis [34], all of which con- firm Graham's relation. Thus we are justified in requiring consistency with equation (6.1) as a primary constraint to be satisfied by any accept- able flux model. Indeed, it is easy to check that the dusty gas model, which is the only comprehensive flux model described so far, satisfies this condition, For a binary mixture the flux vectors of the two species are given by equations (3.26) and (5.27). Setting grad p= 0, noting that grad 7 grad Xp. and taking the ratio of these gives immediately ome a. Bom Dp, =~ My/My which is just Graham's relation. Lt is interestiag to note that the dusty gas model equations also 54 suggest the appropriate form for a generalization of Graham's relatioa to multicomponent mixtures. Setting grad p = 0 in equation (5.10) gives 2 1 this may alse be written (6.2) which is the requixed generalization, applicable to isobaric diffusion in multicomponent mixtures. Ic should be satisfied throughout the whole range between the limits of Knudsen streaming and bulk diffusion control. The second important class of experiments referred to earlier studies the flow of gases under the influence of presgure gradients, but without composition or temperature gradients. The classical investigations of Poiseuille [13] on flow through capillaries and d'Arcy [14] on flow through porous media used liquids rather than gases, and the most important early work on gas flow was that of Knudsen [3], [33], [34] on the flow of gases through long capillaries over a very wide range of pressures. To appreciate the questions raised by Knudsen's results, consider first the relation between nolar flow and pressure gradient for a pure gas flowing through a porous plug, rather than a capillary. The form predicted by the dusty gas model can be obtained by setting x, = 1, grad %, = 9 in equation (5.26), with the result Bop = - Liye 4 22 )ee N= cap (2) + 1 te (6.3) where £ denotes a co-ordinate measured in the direction of flow. If the thickness of the plug is L and p, and are the pressures at its two Pa faces, equation (6.3) can be integrated, giving NL a nea P, +P, PyoPp RT | * 2 (6.4) Thus, if measured values of the left hand side of equation (6,4) are plotted against the mean pressure (Pyt Py)/2, the theory predicts a straight line with slope B,/,RE and and intercept Dy/RT at (py+p,)/2 = 0, and experiments of Carman (35} confirm that this is the case.” 1a contrast, ootnote oa next page. be o Knudsen's very careful experiments on a long uniform capillary show that WL Py 7 Po) passes chrough a marked minimum when plotted as 4 function of (p, + 22)/2, at a value of the mean pressure such that the capillary diameter and the mean free path leagth are comparable. At higher values of the mean pressure, N,L/(p)-p,) rises linearly, as in the case of # porous medium, The reason for this difference in behavior was elucidated by Pollard and Present [8]. These authors were not able to solve the complete problem of pressure-driven flow in a capillary throughout the intermediate pressure yange; indeed, as mentioned in Chapter 3, this remains unsolved today, In- stead they considered the simpler case of isobaric self-diffusion, in which the tube contains molecules of a single species, some of which are considered to be “tagged”. It is then possible to extend the well known theory of Kaudsen streaming {9] to calculate the diffusioa rate of the tagged molecules ander the influence of a gradient in their concentration, and this can be Gone throughout the range of pressures over waich the mean free path and the tube diameter are comparable, As p70 the flux approaches the well known limit given by equations (2.5) and (2,8), but as p increases frou zero the flux decreases as a result of the added resistance to motion posed by inter- molecular collisions, In Pollard and Present's calculation this decrease is monotonic--there is 10 minimum--but their system differs from that in- vestigated experimentally by nudsen, since pressure gradients can generate wiscous streaming in the latter case when the pressure is sufficiently high. As the mean pressure increases, the increase in Flow as a consequence of Viscous streaming ia the experimental situation exceeds the decrease re- sulting from intermolecular collisions, as estimated by Pollard and Present, Thus the retio of flow to pressure drop passes through a minimum.and begins to rise again. Though by no means a complete theory, this is at least a reasonable explanation of the Knudsen minimum, and it then remains to explain why the Pollard and Present minimum is not observed for flow through porous medi: attributed this to the limited length/diameter ratio of the channels in @ typical porous medium and gave a plausible argument in favor of this view. In a later, but less detailed analysis of flow in the intermediate pressure range, Hiby and Pahl [37] suggested that the minimum should be absent when the leagth/diameter ratio of a capillary is less than about sixteen. Since it is likely that this is the case for the chaanels in most porous media of *Other measurements by Adzumi [36] are often quoted as further support for this statement. However, it is doubtful whether Adzumi‘s results extend to low enough pressures to be sure of this. 56 interest, we would not expect to observe a Knudsen minimum, and the pre- dictions of the dusty gas model should not be regarded as unrealistic oa this account. As Graham's relation shows, the molar fluxes in a binary mixture are not equal in magnitude under isobaric conditions, so it follows that equi- molar counterdiffusion can occur only in the presence of a pressure gradient. The prediction and neasurenent of the necessary pressure gradient is the objective of the third class of fundamental investigations mentioned at the beginning of this chapter. The classic experiment was carried out by Kramers and Kistemaker [25], who measured the pressure difference across a capillary in circumstances where the molar fluxes of the two species are constrained to be equal in magnitude, The diameter of their capillary was large compared with the mean free paths, so their results can be compared with predictions basad on the detailed theory of Chapter 4. Their apparatus is shown schematically in Figure 6.3. Figure 6,3. The Kramers-Kistemaker apparatus. (From Kramers and Xistemaker (25]) Chambers A and B are separated by an elastic rubber membrane whose motion in response to pressure differences is transmitted to the needle N. Variations in the pressure difference can then be followed by observing the displacement of the point of the needle. Initially chamber A and the capillary are filled to the desired pressure by opening cocks C, and C,, L 3 with C, and C, closed. Then C, {s opened and chamber 8 is filled 2 with the second gas until the position of the needle indicates no pressure 57 difference between A and B.* On opening cock C, the gases may then jaterdiffuse through the capillary. After an initial short transient period the pressure difference ig found to settle to a value which thereafter changes only very slowly, and this is the value hich is recorded, Since the vhole apparatus is isothermal and the volumes of chambers A and B are constant when the membrane no longer moves, it follows that the numbers of moles in A and 3 separately become constant once their pressures settle to constant values. The molar fluxes of the two species through the capillary mst then be equal in magnitude and opposite in direction, so attaching the ends of the capillary to large chambers of constant volume ha: the required effect of imposing equimolar counter diffusion.” The theory of the experiment can be based on equation (5.29) and its companion obtained by interchanging the suffixes 1 and 2. When 3,= ay these equations agree with the detailed solution of Chapter 4, as we have seen, provided a is large enough that pressure gradients are small and pressuce diffusion terms are negligible. They can be writtes V% S19 949 oxy ap dp --2 — Th 2. 2 =" RF Va +x, i de “2 BRT de fs Mt %) ,) N and these are just the equations derived directly by Kramers and Kistemake: using the arguments of Chapter 4 specialized to the case of two components (it was in this paper that the proper slip boundary condition for a mixtur was first presented.) 0, so it follows from th Under the experimental conditions N, +N, above equations that Fn view oF the sensitivity of the manometric system, which has a full seale range of about 2 mm of mercury, the final pressure must be reached in stages, repeating the operations described a number of times. i Ia Fact, the pressures will never settle completely to constant values until diffusion is complete and the composition is the same in both chambers, but dy making the chambers sufficiently large this rate of drift of the pressures can be reduced below any desired level. 58 Integrating this between the ends of the capillary at £ = (0,L), we obtain tn esol he Ba + Bi £8 ee OL a where x denotes x,, the mole fraction of species 1. In particular, whea (6.5) the chambers contain pure gases, x(L) = 1 and x({0) = 0, so this reduces to* dp © p{L) ~p(0) * “eu ia { (6.6) a % Since &,, 1/p, equation (6,6) predicts that the product pap should take a value which can be calculated from the diameter of the capillary and known properties of the gases, provided the pressure is high enough that mean free paths are short compared with the capillary diameter. Kramers and Kistemaker actually measured values of pdp equal to about 66% of the theoretically predicted value, so the agreement of their theory with experiment is, at first sight, not impressive, However, the observed values of dp were only of the order of 20, Hg, and it is significant that the theory predicts these very small pressure differences needed to establish equimolar counterdiffusion within a factor of two. We must, therefore regard the experiments of Kramers and Kistemaker as generally supporting the theoretical predictions of Chapter 4. ® For greater precision the variation of y with composition should be taken into account on integrating along the capillary. This leads to equatioas more complicated than (6.5) and (6.6), of course. oe Chapter 7. SURFACE DIFFUSION So far, in discussing diffusion at a molecular level, it has been assumed that molecules collide with each other or with the pore walls, taen separate instantly in directions determined by their incident configuration and the laws of mechanics. However, all gases are, to a greater or less extent, adsorbed at solid surfaces, and additional motion of the adsorbed species might be expected to occur by migration from site to site on the surface, Such a surface diffusion mechanism was proposed by Damkéhler 138] and was observed by Wicke and Kallenbach [39] whea mixtures of CO, and Ny diffused through a porous plug. If the wechanism of surface diffusion is regarded as @ "hopping" from site ta site when the adsorbed molecule has enough energy to surmount the barrier between adjacent sites, the surface flux might be expected to oe proportional to the gradient in surface concentration, (Metzner and Weaver [40}, Sandler and Roybal [41].) In other words, if ds is a line element and itsuntt aormal, both drawn in the surface, the flux of species r across this line element in the direction of g should be given by an ex- pression of the form s $ 3 Br a Mids = - de yids (7.2) s , : i where cy is the surface concentration of species r, in moles per unit area, 0S is a surface diffusion coefficient for this species, and x is a surface flux vector, It is reasonable to expect that De will depend on the number of vacant sites in the neighborhood of a given adsorption site, ve of adsorbed Yr species. In view of the activated nature of the surface migration process, $ and hence on the total surface concentration ¢ . 3 jl we might also expect D) to have an Arrhenius temperature dependence, so we can write d ES /RT o . 25 eSy0 + (7.2) If we assume that the rates of adsorption and desorption are both large compared with the surface migration rate, the surface and bulk concentra~ tions of each species will be almost in equlibrium, and hence will be ov related by the equilibrium adsorptioa isotherm, which has the general form 8 : om fey ep a) The form of the functions £. depends on the particular isotherm used; for example the Langmuir isotherm gives the familiar relation c= (7.4) where the Kare the adsorption equilibrium constants. Using an equation such as (7.3) the surface concentrations in equation (7.1) can be replaced by bulk concentrations at adjacent points. Thus s Bee See Bee on ac, }n t t and it follows from (7,1) chat (7.5) In general, therefore, the surface fiux of each substance is linearly re- lated to all the concentration gradients in the adjacent bulk phase. The coefficients in this linear relation depend on the bulk phase concentrations, through the concentration dependence of both 8 and the derivatives dE /de, . Each coefficient is also temperature dependent, A rather simpler situation arises when the bulk concentrations are sufficiently small that the adsorption isotherms approach linearity. Then (7.4), for example, shows that * be, " Ka) (ter) 8 o (otherwise) and equation (7.5) simplifies to $ ae, ar XN ds = D KT) ta ds (7a) ol showing that the surface flux of species xr is proportional to the bulk concentration gradient of the same species, For a capillary of circular cross-section and radius a, the surface flux in the axial direction, ex- pressed in moles per unit cross-sectional area of the capillary, is then given by 5 Be be, Noe RE 7-7) where 2 is an axial co-ordinate. The form of this flux relation is the same as equation (2.5) for Knudsen diffusion, but the dependence on physical and geometric conditions of the factor relating NE to de,/dz is quite different. First note that the Knudsen diffusion coefficient {s independent of composition and pressure, while the factor DS in equation (7.7) depends on the total adsorbed con- centration, and hence on the composition and pressure of the bulk phase. Secondly, recall that the Knudsen diffusion coefficient 1s proportional to the tube radius a, while the corresponding factor in equation (7.7) is inversely proportional to a. Thirdly, the temperature dependence of equation (7.7) is not at all Like that of the Knudsen diffusion coefficient, which is simply proportional to YT. In equation (7.7), on the other hand, the temperature dependence of the factor > is given by equation (7.2), and as a first approximation the adsorption equilibrium constant K_ might be expected to have an Arrhenius temperature dependence, and to decrease with inereasing temperature: B2/RT r (7.8) Combining equations (7.2), (7.7) and (7.8), we therefore obtain oS 8 aod 20% ce) 2 - BS) ART she ee TO 7.9% “dz Sote that the effective diffusion coefficient relating N= and de /dz ta equation (7.9) may cither increase or decrease with increasing temperature, depending on the relative values of E2 and zi, These are well-establish cases of negative temperature coefficients of surface diffusion [42], though measured surface diffusion coefficients are act usually found to have very marked temperature dependence. The form of equations (7.7) and (7.9) suggests a corresponding relation 62 for a porous medium, of the type e - 2s KC) grad ce, (7.10) where pes is the product of Bi and the surface area per unit volume, multiplied by the void fraction and divided by a suitable tortuosity factor to take account of the distribution of pore orientations. Measurements of surface diffusion can be made in a Wicke-Xallenbach diffusion cell [39], which is best operated at low pressure so that gaseous diffusion is of the Knudsen type, and its magnitude is not overwhelmingly large compared with surface diffusion. The contribution of the Knudsea diffusion flux to the total flux is then subtracted out by making separate measurements with a non-adsorbed gas* to determine its Knudsen diffusion coefficient. Since this is known to be inversely proportional to the square root of the molecular weight, the corresponding value for the original test gas follows immediately. The constants & and Ee associated with the adsorption isotherm can be obtained from equilibrium adsorption measure- ments quite independent of any measurements of diffusion fluxes. However, the form of the composition, pressure and temperature dependence 2£ D°° must be deduced from a sufficiently large number of direct measurements of surface diffusion. Accordingly, sone effort has gone into developing a more detailed theory of the mechanism of surface transport, with a view to expressing De in terms of parameters whose values can’ be found without diffusion measurements, In this way the number of parameters which must be fitted directly to diffusion measurements can be reduced to one or two. (See Metzner and Weaver (40].) : The incorporation of surface diffusion into a model of transport in a porous medium is quite straightforward, since the surface diffusion fluxes simply combine additively with the diffusion fluxes in the gaseous phase. It is therefore reasonable to go ahead with the construction of models for the gaseous phase diffusion without considering surface diffusion, though of course it must be incorporated before the models are used predictively, This may be particularly important when modeling the behavior of porous catalysts, where adsorption is strong and surface fluxes may be substantial. The literature of surface diffusion is now quite extensive. A review of the basic ideas, with reference to many of the earlier papers, is given by Dacey [44], and a good selection of references including more recent work can be found in Aris [45} * There is then the problem of identifying such a gas. See Hwang and Kammermeyer [43]. Chapter 8, MODELS OF FLOW AND DIFFUSION IN POROUS MEDIA one flux model for a porous medium--the dusty gas model--has already been described in Chapter 3, Although it is perhaps the most important and generally useful nodel currently available, it has certain shortcomings, and other modeis have been devised in attempts to rectify these. However, before describing these, we will review certain general principles to which all reasonable flux models must conform, A porous medium may be wodeiled in two basically dissimilar ways for the purpose of predicting flux relations. In the first, atteation is focussed on the pore network through which the gaseous Fluxes pasa, while in the second, attention is focussed on the obstacles to gaseous motion presented by the solid matrix. The earliest of ail models, due to Maxwell [16}, and its modern development as the dusty gas model, is the sole representative of the second type. Other treatments are all of the first type and include the early discussion of diffusional limitations in porous catalysts by Thiele [46], the macropore-micropore model of Wakao and Smith [32], the model of Johnson and Stewart {47}, Foster and Butt's [48} model based on convergent and divergent parallel pore arrays, and Feng and Stewart's [49] class of models developed from the earlier work of Johnson and Stewart. The Feng and Stewart models, in particular, are rather comprehensive in seope, and today it is probably reasonable to regard these and the dusty gas model as the principal contenders for serious attention.” When a model is based on a picture of an interconnected network of pores of finite size, the question arises whether it may be assumed that the compo~ sition of the gas in the pores can be represented adequately by a smooth function of position in the medium, This is always true in the dusty gas model, where the solid material is regarded as dispersed on a molecular scale in the gas, but is by no means necessarily so when the pores are pictured more realistically, and may be long compared with gaseous mean free paths. To see this, consider a reactive catalyst pellet with long non-branching pores. The composition at a point within a given pore is * Algebraically, the dusty gas flux relations are identical with one of the many particular cases of the Feng and Stewart model, as we shall see. How- ever, the two differ conceptually in their approach to deriving the flux relations. 64 then determined by the location of the point along the axis of the pore, rather than its position relative to axes fixed in the pellet, and the re~ lation between the two is influenced by the shape of the pore, This is illustrated by Figure 8.1, which depicts two unbranched pores in a catalyst pellet surrounded by a gaseous mixture of uniform composition, Points A and 38 indicated are close together in space, but the compositions at these points are clearly quite different, 8 lies near the mouth of its pore, where the composition is similar to that of the gas outside the pellet, while A lies almost at the mid-point of its pore, where the mixture is richer in reaction products and leaner in reactants. Figure 8.1, Porous pellet not describable by smooth composition fields. As noted, the dusty gas model avoids this situation by dispersing the solid medium, conceptually, on a wolecular scale, while the Feng and Stewart models, as we shall sea, require the pore structure to be "thoroughly cross- Iinke: demands of the pore geometry. In fact, it is easier to turn the question It is difficult to say, in precise terms, just what this condition tound and define a thoroughly cross-linked pore system as one in which smooth functions of position provide an adequate description of the composi- tion variations. To te more precise, suppose £ is a unit vector ia the direction of the axis for any pore segment in the medium. Then if there exist concentration fields e,& » Smooth functions of position x in the medium, such that the axial concentration gradients for the gaseous species in this pore segment are adequately represented by Z.gradc., ve shall say that the smooth field assumption is valid, Correspondingly, the pore geometry will be said to correspond to a smooth field model of the medium, It is only for smooth field models, in this sense, that partial differential equations relating species concentrations to position in space can de written down, However, a pore geometry which is consistent with the smooth field assumption in non-reactive conditions may ao longer be so when chemical reactions take place at the pore walls, This point will be taken up in detail in Chapter 9. Another important distiaction relating to pore geometry is that between “through” pores, with two open ends, and "dead-end" pores with only one. the Latter contribute to the fluxes ia time-varying conditions and provide source or sink terms in the presence of chemical reaction, but they have no influence on steady state diffusion or flow measurements in a non-reactive system. As a result of the discussion in Chapters 1 to 6, we are now in a position to formulate certain conditions which must be satisfied by any acceptable model for the gaseous phase fluxes in a porous medium.“ These are very useful, as it turns out that they are sufficiently restrictive to determine completely the formulation of certain problems, without the need to appeal to any particular flux model, All the following conditions refer to isothermal systems, Conditioa (i). If the pores are sufficiently small or the pressure sufficiently low that mean free paths are long compared with all pore diameters in the medium, then the fluxes and partial pressure gradients must be related by simple equations of the Knudsen form 2 Nose Pc grad p. (8.1) with the coefficients De taking the form 4 e BRT ve = de (8.2) where M, is the molecular weight of the species rand K, is a factor determined by the void fraction, the pore size distribution and the pore geometry. Condition (1i). If the pores are sufficiently large or the pressure sufficiently high that mean free paths are short compared with all pore diameters in the medium, then (a) when pressure variations are small compared with the absolute pressure, so that pressure diffusion is F Rg pointed out in Chapter 7, the surface fluxes may be neglected ia de- veloping models for gaseous phase transport and re-introduced at a later stage. 66 negligible, the fluxes must satisfy the Stefan-Maxwell (8.3) With coefficients 4°) of the forn e sf KL, (8.4) where K, is a factor determined by the pore geometry and the void fraction, and (b) in systems of uniform composition the total molar flux must be related to pressure gradient by an equation of the d'Arey form: (8.5) where B, is a factor depending on the void fraction, the pore size distribution and the pore geometry. Condition (iti), For isobaric diffusion in a binary mixture the flux vectors of the two species must satisfy Graham's relation Spire whatever the relative magnitudes of the mean free path (8.6) lengths and the pore diameters. Condition (iv). For equimolar counter diffusion in a binary mixture, the model must predict a pressure gradieat consistent with the result of the Kramers-Kistemaker experiment. When the mean free paths are long compared with all pore diameters, condition (i) above determines the form of the flux relations completely in isothermal systems, and no further modelling is required if K, is regarded as an empirical parameter which must be evaluated for each particular medium, At the opposite limit, where mean free paths are short compared with all pore diameters, conditions (ii) (a) and (b) do not determine the form of the flux relations completely, since fluxes in the absence of pressure gradients and fluxes in the absence of composition gradients cannot be combined in a simple manner when both types of gradient are present. However, as we shail see in Chapter 11, conditions (ii)(a) alone may, in certain circumstances, 67 provide all that is needed to solve steady state chemical reaction problems in porous catalyst pellets; once again no appeal to a detailed flux model is necessary. Unfortunately many important industrial catalysts do ace satisfy either of these simple limiting conditions, since their pore sizes span the intermediate range where mean free path lengths and pore diameters are comparable, so there is still a real aced for a complete flux model with a broad range of applicability. The simpiest and most commonly used flux model is provided by the dusty gas equations (3.27)-(3-19}. AL the conditions (i)-(iv) above are satis- fied by these equations, and the three parameters K,, X), and BL, intro~ duced in formulating the consistency conditions, are the only adjustable parameters in the model. Their experimental evaluation will be discussed in Chapter 10, The relation between the dusty gas model and the actual struc~ ture of a porous medium is obscure, since the "dust" particles invoked mathe- matically do not fill a substantial fraction of space, like the solid material of the actual porous medium, Feng and Stewart [49] take the view that the dusty gas equations are appropriate oaly for a medium with a single pore size, or a narrow distribution of pore sizes, and a similar position is adopted by Scott and Dullien [4] in discussing their own equations, which correspond to a special case of the dusty gas equations. However, the architects of the dusty gas model disagree, arguing that the distribution of free path lengths for collisions between gas molecules and dust particles corresponds to a distribution of pore sizes. Whichever view is taken, it is undeniable that the dusty gas model makes ao provision for varying this distribution to match a measured pore size distribution in a real porous medium, and it is diffieult to see, for example, how it should be formulated to represent the behavior of a medium with a strongly bimodal pore size distribution. It is mainly for this reason that models based on a more realistic picture of the pore structure are needed. When the pore structure is viewed as an assembly of interconnected channels, the flux vectors are formed by adding contributions from the separate channels, so before a model of this type can be constructed it is necessary to have equations which will predict the fluxes in a single chann: throughout the range from very small to very large digmaters. But we have already seea that no such equations are presently available. The classic Knudsen theory gives the fluxes for very small pores’ or very low pressures, and the continuum theory developed in Chapter 4 gives the fluxes for large pores or high pressures, but in intermediate cases there is no adequate vheory. However, by setting Kj = 29/3, K,-1 and BL - a’/@ in the dusty gas model flux equations, we know that the correct expressions for the fluxes in a single long capillary are obtained at the Knudsen diffusion Limit and, except for the pressure diffusion term (which is frequently negligible), the correct flux relations for a long tube are also obtained at the limit of bulk diffusion control and viscous flow. In the absence of more direct evidence it is therefore both tempting and reasonable to use flux relations of the dusty gas form, with the above choice of parameters, to represent the fluxes in a Single pore of radius a. This leads to one obvious discrepancy with experimental observation--there is no Knudsen minimum in the case of a pressure driven flux, However, it will be recalled from the discussion in Chapter 6 that the minimum is to be expected only in long tubes, In short tubes, which are probably more realistic representations of individual pores in a porous medium, no minimum is observed, so the failure of equations of the dusty gas model form to predict a minimum need not necessarily cause concern, Finally, the rough momentum transfer arguments of Chapter 2 also suggest flux relations of this form for a single pore, and in our present state of knowledge we can probably do no better than this. The simplest way of introducing the pore size distribution into the model is to permit just two possible sizes--micropores and macropores~-and this simple pore size distribution is not wholly unrealistic, since pelleted materials are prepared by compressing powder particles which ace themselves porous on a much smaller scale. The small pores within the powder grains are then the micropores, while the interstices between adjacent grains form the macropores, An eatly and well known model due to Wakao and Smith (32} Represents such a material by the idealized structure shown in Figure 8.2. Figure 8.2. The model of Wakao and Smith (From Wakao and Smith [32] ) Only one-dimensional diffusion is considered and the particles are repre- sented in section by squares, as shown in the diagram. -The authors then consider three routes by which a molecule may move between the planes ab and cd, indicated by broken lines. These are: (a) passage through a macropore of the upper layer followed by a macro- pore of the lower layer (path 1), (b) passage through the micropores of the upper Layer followed by the micropores of the lower layer (path 2), and (e) passage through a macropore of the upper layer followed by micro- pores of the lower Layer (path 3). qhe total flux is then obtained by adding the separate fluxes associated with each of these three paths. Physically, this is a difficult procedure to justify. On planes iater- mediate between ab and cd the compositions at corresponding points of each of the three paths will not be the same, so there should be lateral fluxes introduced by these Lateral composition gradients--it is not con- sistent to assume that the directions of motion are confined to vertical paths, as implied by the diagram. To evaluate the flux by each of the three patha, flux relations spanning the range between the micropore and macropore sizes are needed, and Wakao and smith confined their attention to binary mixtures, using the equations of Seott and Dullien (4}, which take the form dx, /ae -- L “7 Rr 1 ® 7) + go-jl-« lt ED 1 #) and ax, fat =~ 2 2. n= -2 (8.8) By ex, [1+ 2 | These are equivalent to the dusty gas model equations, but are valid only for isobaric conditions, and this fact severely limits the capability of the model to represent the behavior of systems with chemical reaction, To see this we need only remark that (8.7) and (8.8) together imply that N,/X, = “Dd Pos or Wty ve ubich {g Just Graham's relation, However, the condition of conservation of mas in a chemically reacting binary system requires that Num + NM) = 0 iv and these two requirenents are mutually contradictory unless M, = M), corresponding to the special case of limited practical iaterest in which Substances 1 and 2 are isomers. Though illustrated here by the Scott and Dullien flux relations, this is an example of a general principle which is often overlooked; namely, an isobaric set of flux relations canaot, in general, be used to represent diffusion in the presence of chemical reactions. The reaSon for this is the existence of a relation between the species fluxes in isobaric systems (the Graham relation in the case of a binary mixture, or its extension (6.2) for multicomponent mixtures) which is inconsistent with the demands of stoichiometry. If the fluxes are to meet the constraints of stoichiometry, the pressure gradient must be left free to adjust itself accordingly. We shall return to this point in more detail in Chapter 11. Of course, these shortcomings of the Wakao-Smith flux relations in- duced by the use of equations (8.7) and (8.8) can be removed by replacing these with the corresponding dusty gas model equations, whose validity is not restricted to isobaric systems. However, since the influence of a strongly bidisperse pore size distribution can now be accounted for more simply within the class of smooth field models proposed by Feng and Stewart [49], it is hardly worthwhile pursuing this.* Since pore size distributions caa be measured in some detail by mercury porosimetry, it is desirable to have a means of modelling media with dis- tributions rather more general than the simple micropore-macropore system considered by Wakao and Smith. A model proposed by Foster and Butt [48} has this facility, but its relation to the structure of the real porous medium is rather obscure. It is not a smooth field model, in the sense defined above, but follows in detail the composition variations through two sets of pores in parallel, each composed of segments in series with successively varying diameters. In one set of pores the sequence of segments increase in diameter on moving towards the center of the pellet, while in the other set they decrease. These two principal pore systems are also cross linked at intervals, at points where the compositions match within specified tolerances, The result is a structure of some complexity which requires computer sclution to determine the fluxes, However, like the Wakac-Smith model it uses the isobaric relations (8.7) and (8.8) for the fluxes in each pore segment, and consequently it suffers the same limita~ tions on its ability to describe chemically reacting systems, *¥ More recently, the Wakao-Smith type of model has been generalized by Cunningham and Geankoplis [50] to admit two different sizes of macropore. mW Tne class of models formulated by Feng and Stewart [49] is conceptually simpler than its predecessors and, at the same time, offers more Freedom in the description of the pore system. The structure of the medium is repre- sented by a statistically specified network of pores, closely cross-linked so that the smooth field assumption is valid, and such that the fractional void apace associated with pores in the radius interval da, and with orientations in the solid angle element dij about direction %, can be written f£(a,a)da du. Lf dead-end pores are present, it must be emphasized that this represents only that part of the void fraction associated with through pores. Let by denote a unit vector in the direction of the axis of a pore with orientation 2». me Jy ads f Sfp f(a, dada}. grad c, (8.15) 3 Dyadic notation is used in this equation, so As, denotes the outer product. a3 of the vector £5 Now suppose e(a) denotes the total void volume associated with pores with itself, a second order tensor. of radii [6 sor, ge | + grad po Ter + grad p 3 (8.29) 7 This is the most general flux relation of the Feng and Stewart class. Equations (8,20) are not sufficiently specific for practical purposes, so it is important to consider special cases leading to simpler relations. When the pore orientations are isotropically distributed, the second order tensors x, 8 and y are isotropic and are therefore scalar multiples of the unit tensor. Thus equation (8.20) simplifies to x8 5 8%, a(ayP, deca) [grad p, - oe grad p- Ys i, grad c. N NX (8.21) The disposable parameters in these equations are the permeability 8, the surface diffusion factor y , the tortuosity function 4(a) (which also accounts for the fraction of pore space which is dead-ended), and the sur- face diffusion coefEicients Da » to the extent that they cannot be esti- mated by independent means. Since the matrix elements F_, depend on pressure and composition, as well as pore radius, the pressure aad composi- tion depeadence of the coefficients relating to the terms grad Ps in equation (8.20) depend oa the form of the function ¥(a). Thus, by varying the form of x(a), flux relations of quite different algebraic structures can be obtained. Equations (8.21) stil! contain too many adjustable parameters to be of much value for predictive purposes, and Feng and Stewart propose three simpler special cases which may be of practical value. Gase (a). The through pores are of uniform radius a, and isotropic orientation, Also the surface coverage of adsorbed species is low, so that only one term survives in the last summation on the right hand side of equation (8.21), as discussed in Chapter 7. Then the flux relations reduce to x,p3 2 Nes em Die) grad p, - Thy atad p > 8, grad c, (8.22) = where s at, 6. = wo lim — Pr iP Be, The available parameters are %, 6, a, and the coefficients A. (P= l,...4a). Neglecting the surface diffusion fluxes equations (8,22) are then identical in form with the solution of the dusty gas model flux relations. Gase (b). x x a= Case (c). ” The 75 The tortuosity factor is independent of pore size, the pore orientation is isotropic and the surface coverage of adsorbed species is low, Then (8,21) reduces to x P38 _. > we - c = > J, Fesd@(@) | ered 0, - Sap ered p ~ 4, grad cy (8,23) ras Furthermore, if there are no dead end pores it is aot dif. alt to show from equation (6.16) that ne} under the assumption of isotropic pore orfentation, and it then follows from equation (8.18) that Le 2 wl? 42 (a) Since the void fraction distribution is independently measurable, the only remaining adjustable parameters are the 4, so when surface diffusion is negligible equations (8.23) provide a completely predictive flux model, Unfortunately the assumption that (a) is independent of a is unlikely to be realistic, since the proportion of dead end pores will usually increase rapidly with decreasing pore radius. The pore size distribution is strictiy bimodal, with macropores of radius a, and micropores of radius a, both of which are isotropically oriented, 1£ the surface coverage of adsorbed species is also assumed to be low, equation (8.21) then reduces to the following form: x Ps grad p - A, grad ¢. * (8.24) Here W,, W, 8 and the 4, are adjustable parameters to be fitted ta flow and diffusion measurements. This type of model might be expected to give a good representation of pelleted material formed from microporous particles. It therefore covers the same situation as that envisaged by Wakao and Smith, but it is not limited in application to binary mixtures, and is not based on purely isobaric flux relations which restrict its use in de- seribing chemically reacting systems. pore size distribution function ¢(a) appears parametrically in the flux relations of Feng and Stewart, so their models certainly cannot be completely predictive in nature unless this distribution is kaown. It is 76 measurable in principle by mercury porosimetry, but che measurements are extensive and tedious. Thus it is reasonable to ask whether any use can be made of limited information about this distribution function, less than a knowledge of {ts complete form, This question has been addressed by Petty (S11, who showed how limited statistical information about the pore size distribution could be translated into bounds on the transport parameters. 7 Chapter 9. FLUX RELATIONS UNDER REACTIVE CONDITIONS Though a porous medium may be described adequately under non-reactive conditions by a smooth field type of diffusion model, such as one of the Feng and Stewart models, it does not necessarily follow that this will scill be the case when a chemical reaction is catalysed at the solid surface. In these circumstances the smooth field assumption may not lead to appropriate expressions for concentration gradients, particularly ia the smaller pores. ‘Though the reason for this is quite simple, it appears to have been largely overlooked. To be specific let us have in mind a picture of a porous catalyst pellet as an assembly of powder particles compacted into a rigid structure which is seamed by a system of pores, comprising the spaces between ad jacen particles, Such a pore network would de expected to be thoroughly cross- linked on the scale of the powder particles. It is useful to have some quantitative idea of the sizes of various features of the catalyst structur so let us take the powder particles to be of the order of 50y in diameter. Then it is unlikely that the macropore effective diameters are much less than 10,000 A, while the mean free path at atmospheric pressure and ambient temperature, even for smali molecules such as nitrogen, does not exceed about 1,000 8. thus, at only moderately elevated pressures, diffusion in the macropore system will be controiled mainly by intermolecular collisioas and we will assume that the bulk diffusion-viscous flow limiting form of th dusty gas model is adequate to describe it. It will further be assumed tha each powder particle is seamed with micropores of uniform radius 15 4, some of which pass through the particle or are interconnected {n such a way that there is a through passage between two points on the surface of th particle. These will be referred to as through micropores. Others have only one opening located at the particle surface, and will be referred to as dead-end micropores. For simplicity it will be assumed that the through micropores have uniform length, of the same order of magnitude as the par- ticle diameter, while the dead-end micropores are assumed to have the same radius but only half the length, Ia view of the value quoted above for the mean free path length, it is clear that diffusion in the micropores is of the Knudsen type. The orientations of both macropores and micropores in ar element of the pellet containing many powder particles will be assumed to t 78 isotropically distributed. Figure 9,1 shows a simple idealized picture of a structure consistent with this description. Salil ay Figure 9.1, Idealized catalyse structure. let us denote the micropore and macropore radii by a, and ay respectively, and the length of the throughmicropores by 22, Let ty #, and €, de the contributions to the void fraction due to macropores, through micropores and dead-end micropores, respectively. In non-reactive conditions concentrations in the gaseous phase might be expected to vary significantly only over distances large compared with the powder particle size, and consequently it should be possible to define smooth concentration fields ¢_ such that the concentration gradients in any pore (micro or macro) are given by f+ gradc,, where 2. is a unit vector along the axis of the pore.” Smooth field models of the Feng and Stewart type (Chapter 8) should therefore be adequate to describe the fluxes and, since we have idealized to a strictly bimodal pore radius distribution, their general equations (8.21) reduce to 1 x P8 "- a Pie laa) * Fi (4,) grad Pe od grad p. (9.4) N ue 8 where the factor 4 results from the assumption of isotropic orientations of both micro and macro pores, and surface diffusion has been neglected because it is not relevant to the present discussion, The matrix elements Fig (3q) are obtained by solution of the dusty gas equations in their limiting form for bulk diffusion control, and are quite complex in general. Fig(a)s on the other hand, follow from the equations of Knudsen diffusion and have the simple form * Dead-end micropores are excluded here, of course, since they carry no concentration gradients in steady non-reactive conditions. 7a Dita y F(a.) =a Fe (9.2) 39) “RE rs uheve £,, is the Kronecker delta, and D.(a,) is the Kaudsen diffusion coefficient for species r in the micropores, Viscous flow is confined to che macropores, end for a randomly oriented assembly of straight cylindrical macropores, it follows from equation (8.18) that (9.3) Then, using (9.2) and (9.3), equation {9,1) can be written Dita) ake om ai x, 3 t RT grad Pe 73 Fyg(8q) grad Ps 5 2 ®nan and “ie RF etd P ow which is the final form for the steady state flux relations, Note thac they are completely predictive in form, containing no adjustable parameters, provided eps aq and a are known, When a chemical reaction takes place at the solid surface, we expect a smooth varietion in gas composition in the macropores on a scale compar- able with the whole pellet, provided the reaction rate is aot too high. The mixture will be richest in reactants near the pellet surface and richest in reaction products ear its center. ‘Thus we might expect the smooth field assumption to remain valid for the macropores. In the micropores, however, the situation is quite different. Here the ratio of surface area to volume is comparatively very large, and there may consequently be a marked varia~ tion in composition along the length of a single micropore. The smoothed concentration fields, which correctly estimate the concentration gradients ia macropores through the formula £. + grad c,, may therefore lead to totally misleading concentration gradients if applied to the micropores under chemically reacting conditions. Consequently, if grad p, denotes the gradient of the smoothed partial pressure field associated with species r in the marropores, an expression of the form 1, rf I) 2 75%, TRE Brad Py O-9) nay no longer correctly assess the contribution of the through micropores to

You might also like