[John Villadsen, M. L. Michelsen] Solution of Diff(BookFi.org)

231

Click here to load reader

Transcript of [John Villadsen, M. L. Michelsen] Solution of Diff(BookFi.org)

  • Solution of Differential Equation

    Models by Polynomial Approximation

    JOHN VILLADSEN MICHAEL L. MICHELSEN

    Instituttet for Kemiteknik Denmark

    PRENTICE-HALL, INC. Englewood Cliffs, New Jersey 07632

  • Library of Congress Cataloging in Publication Data VILLADSEN, JOHN.

    Solution of differential equatiOn models by polynomial approximation.

    (Prentice-Hall international senes in the physical and chemical engineering sc1ences)

    Includes bibliographies and index. 1. Differential equations-Numerical solutions.

    2. ApproximatiOn theory. 3. Polynomials. 4. Chem1cal engmeering-Mathematical models. I. Michelsen, M. L., joint author. II. Title. QA.371.V52 515'.35 77-4331 ISBN 0-13-822205-3

    197 8 by Prentice-Hall, Inc. Englewood Cliffs, N.J. 07632

    All rights reserved. No part of this book may be reproduced in any form or by any means without permission in writing from the publisher.

    10 9 8 7 6 5 4 3 2 1

    Printed in the United States of America

    Prentice-Hall International, Inc., London Prentice-Hall of Australia Pty. Limited, Sydney

    Prentice-Hall of Canada, Ltd., Toronto Prentice-Hall of India Private Limited, New Delhi

    Prentice-Hall of Japan, Inc., Tokyo Prentice-Hall of Southeast Asia Pte. Ltd., Singapore Whitehall Books Limited, Wellington, New Zealand

    1.

    2.

    Contents

    Preface

    A Preliminary Study of Some Important Mathematical Models from Chemical Engineering

    Introduction, 1. 1.1 The General Mathematical Model, 3. 1.2 Steady State Homogeneous Flow Model for a Reacting

    System, 10. 1.3 Steady State and Transient Models for Solids, 23. 1.4 Heterogenous Model for a Reacting System, 34. 1.5 A Model for Hollow-Fiber Reverse-Osmosis Systems, 40. 1.6 Flow of Polymer Melts in Extruders, 45. 1.7 Solution of Linear Differential Equations, 50. Exercises, 59. References, 61.

    Polynomial Approximation-A First View of Construction Principles Introduction, 67. 2.1 A Taylor Series Approximation, 68.

    vii

    xi

    1

    67

  • viii

    3.

    4.

    5.

    Contents

    2.2 The Lowest-Order MWR Approximations Y1(x), 73. 2.3 Higher-Order MWR Approximations y~x), 77. 2.4 Nonlinear Problems, 92. 2.5 Reformulations of the Nth-Order Approximation YN(x ), 98. Exercises, 108. References, 109.

    Some Important Properties of Orthogonal Polynomials-Formulation of a Standard Collocation Procedure 111

    Introduction, 111. 3.1 The Power Series Representation of PK:13 )(x ), 112. 3.2 Zeros of Orthogonal Polynomials, 115. 3.3 Differentiation and Integration of Lagrange Interpolation

    Polynomials, 118. 3.4 Program Description, 131. 3.5 Discretization of Differential Equations in Terms of

    Ordinates, 135. Exercises, 138. References, 141.

    Solution of Linear Differential Equations By Collocation 143

    Introduction, 143. 4.1 Solution of a Linear Boundary Value Problem, 144. 4.2 Integration of Initial Value Problems, 148. 4.3 Linear Parabolic Differential Equations, 166. 4.4 Collocation Solution of a Linear PDE Compared to Exact

    Solution, 175 4.5 Construction of Eigenfunctions by Forward Integration, 183. 4.6 An Ordinary Differential Equation at the Boundary of a

    PDE, 188. Exercises, 191. References, 196.

    Noklinear Ordinary Differential Equations

    Introduction, 198. 5.1 Multiple Solutions of a Trivial Boundary Value Problem,

    200.

    198

    Contents

    6.

    7.

    8.

    5.2 Existence and Uniqueness Theorems, 204. 5.3 Application of Comparison Differential Equations, 207. 5.4 Global Collocation Solution of a Nonlinear Differential

    Equation, 212. 5.5 Concentration Profiles for Nonisothermal Reactions, 215. Exercises, 228. References, 230.

    One-Point Collocation

    Introduction, 232. 6.1 Application of One-Point Collocation

    to Ordinary Differential Equations, 233. 6.2 Application of One-Point Collocation

    to Partial Differential Equations, 253 6.3 One-Point Collocation for Initial Value Problems, 266. Exercises, 269. References, 271.

    Global Spline Collocation

    Introduction, 273. 7.1 Global Spline Collocation for Two-Point Boundary Value

    Problems, 274. 7.2 Eigenvalues and Entry Length Problems by Spline

    Collocation, 285. Exercises, 293. References, 295.

    Coupled Differential Equations

    Introduction, 297. 8.1 On the Numerical Structuring of Coupled Differential

    Equations, 299. 8.2 The General Initial Value Problem-

    Accuracy, Convergence, and Stability Considerations, 305. 8.3 Sensitivity Functions, 328. 8.4 Partial Differential Equations, 332. Exercises, 335. References, 344.

    IX

    232

    273

    297

  • X

    9.

    Contents

    Selected Research Problems 347

    Introduction, 347 9.1 The Graetz Problem with Axial Conduction, 348. p , .f 9.2 Asymptotic Stability of a Catalyst Particle, 365. fej ace 9.3 Fixed Bed Reactor Dynamics-Transfer Functions and State

    Space Formulation by Collocation, 394. Exercises, 405. References, 413.

    Appendix A: Computer Programs with Test Examples 417

    Subject Index 441

    Author Index 445

    The principal aim of this book is to support the engineer-specifically the chemical engineer-who is interested in quantitative treatment of physical models. Engineering models often appear in the guise of differential equations and we present some tools for solving these models by polyno-mial approximation.

    Many years of teaching experience have made it abundantly clear that .chemical engineering students and graduates are by no means less mathematically gifted than their colleagues working in other engineering sciences. Once they understand that even very complex models can be attacked with the aid of only a few, basic methods and that these methods are available as computer codes, many of the students develop a vora-cious appetite for mathematical studies and may have to be pushed gently away from their newly acquired hobby of mathematical modeling before they become estranged from their true vocation of chemical engineering: inventiveness in the field of chemical processes, the actual large- or small-scale synthesis of chemical compounds, and the interplay between the chemical industry and society as a whole.

    Scandinavian technical universities have a long tradition of giving their students a solid background in classical mathematics-at least with respect to the number of required courses. Still it is an undeniable fact that their lack of preparedness for numerical manipulations with models is often lamentable. Since the same observation is also made by col-leagues in American universities, it may express some rather fundamental misconception in our way of teaching mathematics to young people even before the university level.

    xi

  • xii Preface

    The graduate student or engineer working on a mathematical project is interested in the solution of the model and in the physical insight that the solution can give him. Unless he feels that the mathematics course helps him understand the physical background of his own problem, he will approach it as just another required course that stands between him and his true interests. Not only may a successful course in mathematics improve the student's understanding of physical sciences, but it may also prompt him at some later stage of his development to study specific subjects of mathematics in depth.

    Countless books exist on applied mathematics and quite a few on numerical methods. Hildebrand's series is an example of excellent publi-cations from the first category, while Lapidus' Digital Computation for Chemical Engineers has taught many generations of students how to talk sense with a computer.

    Our text tries to incorporate model understanding and numerical model analysis into a single book. It does not deal with numerical methods per se (classical methods such as finite differences are hardly touched); neither is it a book about applied (or engineering) mathematics. Classical methods for solving linear differential equations are reviewed in chapter 1 and well-known techniques such as approximation by perturba-tion series occur frequently. Yet there are much better and more com-plete treatises on these subjects.

    On reflection, the book seems to owe much to Lanczos' beautifully written Applied Analysis. He tried to enamor the student with the philosophy of numerical work rather than presenting the subject as either a string of theorems or a stack of digested subroutines on punched cards. Each physical problem must be treated on its own merits-at least if it is worth any research effort-and a careful analysis of the problem is more than half its solution. Once the ephemeral "feeling" for numerical work has been instilled into the student, he can produce remarkable results with only a few tools.

    The first chapter presents what we regard as the basic units of mathematical models in chemical engineering: fluid flow, diffusion, and chemical reaction. In developing the model for a fixed bed reactor, the intimate coupling between transport processes that leads to an almost untractable mathematical complex is clearly seen. Much emphasis is put on the simplifications that are possible by judicious model approximation. The student who understands how the model can be simplified without sacrificing t~ desired essential features will save himself many hours of fruitless computer work, and he is likely to emerge from his analysis with a far better understanding of the physical process.

    The personal interests of the authors in chemical reactor simulation as an academic research subject and in industrial practice is apparent not only in the first chapter but in the entire book. We make no amende

    Preface XIII

    honorable for this preoccupancy with chemical reactions and transport phenomena, but hopefully the last two examples of chapter 1 will give some inspiration to the reader who is interested in fluid mechanics and rheology. The numerical methods of the following chapters can certainly also be applied to these subjects.

    Chapters 2 to 4 give a systematic treatment of weighted residual methods applied to linear problems. These chapters may be regarded as the core of the book, and the student who becomes familiar with the application of the algorithms in chapters 3 and 4 will be able to solve many models of practical interest. The orthogonal collocation method is our preferred choice of weighted residual method, for reasons that are stated clearly in chapters 2 and 4. We are not convinced, however, that this very popular method is a vademecum for the solution of differential equations. We believe that collocation is a convenient, mechanical method that may give even more desirable results in unison with quite different techniques such as perturbation methods. There are numerous examples of this combination of methods in chapters 2, 4, 6, 8, and 9.

    Nonlinear problems are treated in chapters 5 and 8. Here simplifica-tions of the numerical procedure-similar to the model simplifications of chapter 1-are pursued whenever possible. To reduce the numerical work, many different tricks have to be played. In chapters 5 and 8, many of these tricks are advocated, most notably the sensitivity analysis that makes tracing of the solution as a function of model parameters much easier.

    Specific difficulties of the problem may require special techniques. Chapter 7, which describes global spline collocation as a method of solving entry length problems, is specifically noteworthy. Part of chapter 8 treats an efficient procedure for integration of sets of stiff equations, a numerical problem of tremendous interest in current research. The procedure is based on semi-implicit Runge-Kutta methods that are related to the collocation methods of chapter 4 but are easier to apply.

    The techniques of chapter 6-one-point collocation methods-are eminently suited for qualitative study of the problem. The one-point collocation method is mentioned in numerous papers from the last few years. Different variants of the method are reported, and usually the object is to study the influence of model parameters on the solution.

    In chapter 9 we report results on three larger research projects. The problems are all from very active areas of research, and their solution might be interesting in itself, but the intention of the chapter is to show how closely all the methods of the previous chapters work together. Perhaps this will give the reader determination to analyze his own problems in a similar fashion.

    Finally, in the appendix we have collected a number of subroutines that should help the reader to solve a majority of computational problems

  • xiv Preface

    in differential equations. The routines have been used for over 6 years at this university and in some large Danish engineering companies. The routines are available in the libraries of computer centers at the Univer-sity of Houston and at the University of California at Berkeley and are used at a number of other American universities; in Salford, Great Britain; at the Technion in Haifa; and in various Scandinavian univer-sities. We believe our long experience with the programs have made them almost failproof.

    The readers of this text are usually from various fields of engineering or science. The problems may be stated in chemical engineering terms but they are usually of a quite general nature. In our university (with a 5-year curriculum for the degree of MSc in engineering) it is used in the fourth year as a textbook in a one-semester course {4 credit hours). In American universities it is probably best suited for a first-year graduate course, although some of the material can certainly be taught at an undergraduate level. Sections 1.1 to 1.4, 1. 7, and 8.2, and chapters 2, 3, 4, and 6 contain material for approximately one semester. Very brief continuing-education courses, supplemented by one week of intense problem solving, have been given from the appendix alone with good results.

    Exercises are given in each chapter. Many of these have been selected to support the understanding of a specific point in the text. Another group of exercises points to applications of the methods that are not immediately obvious from the text. Finally, a number of exercises are formulated as small research problems-either as an analysis of a recent paper from the literature or as an extension of previously reported results. This last group is the most important to us because we wish to equip the student for individual research in numerical treatment of mathematical models. Naturally some basic computer experience should be required as a prerequisite to the course, but in many of the computer exercises, no more than a straightforward application of the algorithms of the appendix is needed.

    The introductions to each chapter should help focus the attention on the major ideas of the chapter and after each chapter there is a brief list of references (with comments about the most important ones). No attempt has b~en made to present a complete set of references for each subject.

    In finishing a project of this size the authors have incurred debts to many persof.ts. J. Villadsen wishes to thank Professor Warren E. Stewart of the University of Wisconsin for an invigorating experience more than 10 years ago. It was in cooperation with this dedicated scientist that the first version of orthogonal collocation was conceived.

    Thanks are also due to the many chemical engineering departments in different parts of the world who have provided funds for short courses on

    Preface XV

    part of the material; most recently a full-semester course was offered at the University of Houston, where the graduate students helped to draw attention to weak passages.

    Drs. Bruce A. Finlayson and Bruno van den Bosch read the manu-script and suggested many valuable corrections. The authors graciously acknowledge their contributions.

    Finally, the secretaries at Instituttet for Kemiteknik have labored over several years to produce the editions of notes that have now finally reached a measure of completeness. Mrs. Bente Hansen and Mrs. Johanne Nielsen bore the brunt of this work. For their patience we owe them our sincere thanks.

    Lyngby, Denmark JOHN VILLADSEN MICHAEL L. MICHELSEN

  • Solution of Differential Equation

    Models by Polynomial Approximation

  • A Preliminary Study of Some Important

    Mathematical Models from Chemical Engineering

    Introduction

    In this chapter, we present the main items of the book:

    1

    1. Define the models that are to be solved by approximate methods and, in a purely qualitative way, discuss the expected behavior of the mathematical solution.

    2. Present a few important mathematical tools that have been used to obtain "exact" or "closed-form" solutions of various differential equation models.

    The physical systems to be considered are composed of one or several phases, each of which is supposed to be continuous on the microscopic scale. Differential mass balances and heat balances for each phase and equilibrium relations between the phases are our main modeling tools. If the system is a flowing fluid, both mass and heat balance are influenced by the velocity field of the fluid, and this is described by a momentum balance. Our starting point is equations (1) to (3) of section 1.1, and these equations are used as far as possible in an attempt to give a unified treatment of the model formulation. The equations are certainly not an exact description of the physical systems, and the reader should note the restrictions that the application of a simple set of basic equations imposes on the treatment. Model formulation must necessarily be based on actual physical systems, and sections 1.2 to 1.6 are a catalog of models that we have chosen to use as examples. The choice of examples is directed by

  • 2 Mathematical Models from Chemical Engineering Chap. 1

    the interests of the authors, and the reader who finds that too much emphasis has been put on models for chemical reactors can hopefully be convinced that models for these systems exhibit most of the important features of the vaguely defined model building concept. The examples are organized as shown in table 1.1.

    Section

    1.1

    1.2

    1.3

    1.4

    TABLE 1.1 ORGANIZATION OF MODELS DESCRIBED IN CHAPTER 1

    Description of model

    General model: momentum-energy-and mass balances. Explicit models for velocity distribution in unidirectional flow.

    Models for a homogeneous fluid phase with or without reaction. No interac-tion fluid-bed packing.

    Steady state and transient models for a solid phase on which a chemical reaction occurs.

    The interplay between the flowing fluid phase and the stationary solid phase. Steady state and transient fixed bed reactor models.

    1.5 Purification of a fluid by reverse osmosis in hollow fibers. The example shows how complicated models can sometimes be treated gradually by suitable assumptions.

    1.6 Two-dimensional flow and heat transfer to a non-Newtonian fluid. Material beyond the level of section 1.1 is used in the model formulation. The final model has an unusual mathematical appearance.

    The models of sections 1.2 to 1.4 are all characterized by rather simple flow patterns: The convective fluid flow is one-dimensional in the axial direction of a cylinder. Diffusion gradients inside solids are con-sidered only across a thin plate, in the axial and radial directions of a cylinder and radially in a sphere. With these restrictions added to the assumptions of the basic model (1) to (3), an apparently complete treatment of these frequently occurring systems, all of which are treated in subsequent chapters by numerical methods, can be given.

    The layout of the chapter would give the reader too bright a picture of model formulation as an easily systematized subject if no more examples were included. Some feeling for the complexity of the subject may be obtained tJy"ough a study of the last two examples in sections 1.5 and 1.6. These examples are not very complicated, although they include a few features such as a two-dimensional velocity field that was not present in the previous examples.

    Results from classical mathematical analysis occur throughout the book interspersed with the approximative methods that will be our main tools for solving models.

    Sec. 1.1 The General Mathematical Model 3

    Two subjects, however, are given separate treatments. These are the solution of coupled linear differential equations with constant coefficients and the Fourier series solution of linear partial differential equations.

    As fur.ther ar?ued in the introduction to section 1. 7, these specific mathematical topics are used in almost any type of numerical treatment of much more complicated nonlinear differential equations.

    1.1 The General Mathematical Model

    A differential mass balance for component A in a dilute binary mixture takes the form

    dCA iit + ( v . v c A) = (V . D v c A) - RA (1) A corresponding thermal energy balance is almost identical to (1):

    pc.[~~ + (v VT)] = (V kVT) + Q (2) The velocity field v, which must be known in order to solve these two equations, is given by a momentum balance for the fluid.

    For a constant-density, constant-viscosity Newtonian fluid one obtains the f~llowing vector differential equation, the so-called N~vier-Stokes equatiOn:

    p[: + (v Vv)J = J.LV2v + pG ~ Vp (3) Equation (3) int~oduces a new variable, the pressure gradient Vp. In the examples of sect10~s 1.2 ~o 1.4, the pressure is known at every point of the system. EquatiOn (3) IS supplemented by the equation of continuity

    dp at = -[V . (pv)]

    or for a flow at constant density, simply

    v. v = 0 (4) . The reader _who is familiar with the subject of transport phenomena

    will have n.o difficulty in recognizing this set of four equations as the source of mnumerable mathematical models of interest to chemical engineers. He will also know that these leave some physical phenomena and .syst~ms ~ntouched. If he is in doubt about their applicability in a specific SituatiOn, he should consult a standard text such as Bird (1960, chapters 3, 10, and 18) for an almost complete guide to the model building process.

  • 4 Mathematical Models from Chemical Engineering Chap. 1

    We shall never need a more general basis for the models solved in the present text. If any phenomenon that is not cov.ered by the ~ssumptions of (1) to (3) appears, the system will be very simple othe~wi.se and the appropriate differential balances can be set up from firs~ pnnciple~.

    The two terms of the left-hand side of the equations descnbe the accumulation of mass, thermal energy, and momentum per unit time in a unit volume of the fluid as seen by an observer who travels through the system, passively drifting along the streamlines. Thi~ total a~cumulation term-the substantial derivative or D/ Dt in the notatiOn of Bud (1960, P 73)-may also be conceived as the sum of two perh~ps more ~ell-known terms. The first term is the conventional accumulation term m a volume that is fixed in space, and the second term is the rate of change of the property by the convective flow through the element.

    The right-hand sides of the equations also consist of two terms. The first term is the change of molecular diffusion flow across the element. Fick's law, Fourier's law, and Newton's viscosity law are used to correlate the flux q of an extensive property (mass, heat, ?r mom~ntum) t? the gradients of concentration, temperature, . and flmd. vel?~Ity by simple semi-empirical proportionality relations with mass diffusivity D, thermal heat conductivity k, and viscosity JL as proportionality constant~.

    As written in (1) and (2), D and k are molecular properties of the fluid, and these quantities are well known either fr~m experiments or from theoretical calculations. We shall often use equations (1) and (2) for systems that are much too complex to allow a complete solution of (1) to (3). An empirical velocity distribution is inserted in (1) and (2), but now the transport parameters are empirical quantities without relation to k and D as they stand in (1) and (2): A turbulent axial mixing coefficient, a mass transport coefficient inside the pore system of a catalyst, and a radial heat conductivity in a fixed bed reactor will all be called D and k with only a brief reference to their physical significance. The r~ason. that this violation of the assumptions of the general model does not mvahdate the results is that the mechanisms of the more complicated transport phenomena are empirically found to be similar to those leading to the diffusion terms of (1) and (2), and consequently the mathematical models look alike although their foundation is different.

    The last terms-RA, Q, and (pG- Vp)-are volumetric production terms. RA is the rate of disappearance of A per unit volume by a chemical re.action. Q is a corresponding heat of reaction by the chemical reaction or ttnother continuous heat source in the fluid. Internal friction, absorption of radiant heat, or dielectric heating all contribute to 0, while external heating from a steam jacket surrounding the fluid has to be treated differently in an accurate model, although we shall see that an approximation of a boundary heat source by an average volumetric heat source is quite often used.

    Sec. 1.1 The General Mathematical Model 5

    Finally, pG and -Vp in equation (3) describe the production of momentum by body forces proportional to the fluid density (the action of gravity is an example) or by a pressure gradient V p in the flowing system.

    Equations (1) and (2) are scalar equations, and their solutions are the scalar quantities cA and T as functions of the space coordinates and time t. Equation (3) is a vector equation with three components, one equation for each of the three unknown scalar components of the fluid velocity v.

    We denote the coordinates of the normal rectangular coordinate system as either (x, y, z) or (x1, x2, x3), whichever is more convenient. The axial coordinate in cylinder geometry is called z, and the radial coordinate in cylinder and spherical geometry is called r. We shall never treat problems in the angular coordinates of these last two geometries-the dependent variables will be constant in these directions. This restric-tion makes a complete description of the relevant vectorial quantities of equations (1) to (3) for all three geometries quite easy.

    The three components of v are vxb vx2, vx3 or, in a more convenient notation, vb v2, v3. The gradient of the scalar quantities cA and T is a vector with the partial derivatives in the 1-, 2-, and 3-directions as components. In rectangular coordinates, the scalar products v V c A and v VT of equations (1) and (2) are

    3 acA 3 aT I v;- and I v;-1 ax; 1 ax;

    In equation (3), the quantity Vv-the gradient of the vector v-is a second-order tensor, which is a (3 x 3) matrix M with element m;i = av/ ax; in row i and column j.

    The scalar product v Vv is the product of the (row) vector v = (v11 v2, v3) and M. It is thus a (1 X 3) vector, where each component is the scalar product of v and a column of M.

    av1 av2 av3 ax1 ax1 ax1

    v Vv = ( v11 v2, v3) av1 av2 av3 ax2 ax2 ax2 (5) av1 av2 av3 ax3 ax3 ax3

    = ( f Vav1 1 ' i=1 axi

    f av2 Vj ' j=l axj

    f vav3) i=1 'axi

    Assume for the moment that D and k are constant; i.e., they are the same in all directions and they are. independent of cA and T. In this case, the first term on the right-hand sides of (1) and (2) can be written DV2cA

  • 6 Mathematical Models from Chemical Engineering Chap. 1

    and k V2 T. V2 is called the Laplacian operator and, in rectangular coordinates, the diffusion terms contribute with the following scalars in equations (1) and (2) and in the ith equation of (3):

    ~ a2cA k ~ a2T ~ a2vi D ~..J 2, ~..J 2' JL ~..J 2

    J=l axj j=laxj j=lax, The scalars V2 vi (i = 1, 2, 3) that appear, one in each of the three scalar equations (3), are, of course, only the correct contributions for the viscous forces when the fluid is Newtonian (and has a constant density: V v = 0). For more general fluids the correct viscous term -V T must be used in the momentum balance in place of -JL V2v. Tis the shear stress tensor, a (3 x 3) matrix with components rji, and the vector-matrix product

    . . .

    3 arji . -V Tis similar in form to (5) and gives a contnbut10n- I- to the zth

    ,=lax, equation of (3).

    The curvilinear expressions for the components of V and - V T and for the scalar V2 are summarized in table 1.2 for the r- and z-directions of a cylinder ( v = 0 and v, Vz independent of cp) and for the r-direction of a sphere. (v8 = v = 0 and v, is independent of () and of .)

    Cylinders

    r-directlon

    z -direction

    Spheres

    r-direction

    ~

    Exam.Jies:

    TABLE 1.2 EXPLICIT EXPRESSIONS FOR SOME TERMS THAT

    OFTEN APPEAR IN EQUATIONS (1) TO (3) IN CYLINDRICAL AND SPHERICAL GEOMETRY

    Components of V The scalar \;P Components of -V T/ J.L

    a [1 a J a2 v, a - - -(rv,) + - 2

    ar ar r ar az 1 a ( a) a2 -; ar "a; + az 2

    a ! ~(~) + a2 v; az r ar ar az

    _!_ ~( , 2 av,) _ ; v, a 1 a(za) ar ;a;.'a;. r2 ar ar r

    1. The energy balance in spherical geometry with no variation in the 8- or -directions:

    c (aT+ vaT\ = _!_ ~(r2kaT\ + 0 (6) p p at r a-;} r2 ar a-;}

    Sec. 1.1 The General Mathematical Model 7

    2. The momentum balance in cylinder geometry for a constant-property Newtonian fluid. Only the z-component and the ,_ component are .shown. It .is assumed that v = 0 and that Vz and v, do not change m the cp -direction.

    ( avz + avz avz) P at v,-;;; + Vz--;;;

    (7)

    = JL{~[! ~(v,r)] + a2v,} + pG - ap ar r ar az 2 r ar

    In the flow ?roblems of the following sections we shall assume that (3) can be solved mdependently of (1) and (2). This leads to an enormous s~mpli~cation of the computational work since the three coupled equa-tions_ m (3) can be solved separately for the components of v and the solutiOn v(x, t) can afterward be inserted into the two coupled equations ~1) ~nd. (2),. which are finally solved for cA and T. The assumptions tmphed m t~Is computational scheme are that v and the parameters p and JL of (3) are mdependent of cA and T and that the chemical reaction must occur withou~ volume change. These assumptions-except perhaps the temperature mdependence of p and JL-are generally quite reasonable for diluted reacting mixtures.

    It _is convenient to present a few classical solutions of Navier-Stokes e~uat10ns since these explicit solutions will be used throughout the text without reference to equation (3).

    For laminar flow in a cylinder with inpenetrable walls v and v are d h ' r

    zero an t e ~xial .ve!ocity ~z is determined by the axial pressure gradient and by the flmd fnct10n. Smce the fluid is considered incompressible and any volume change due to chemical reaction is neglected, Vz is indepen-dent of z. It_ does change in the r-direction since a dissipation of z-momentum IS necessary to maintain the fluid flow in the z-direction and the r-component of the z-momentum vector is a function of avz/ar. C:onse_quently for _steady state laminar flow of a constant density and viscosity, Newtoman fluid vz(r) is obtained by solution of the z-component of (3) with

    avz avz V = v = - = - = G = 0

    r az acp z

    1 d( dvz) dp JL--; dr '--;I; = dz = constant

  • 8 Mathematical Models from Chemical Engineering Chap. 1

    With a no-slip condition vz = 0 at the tube wall r = R, the solution, which is finite at r = 0, is

    ( dp)R2

    ( r2

    ) v (r) = -- -- 1 - ~ z dz 4~-t R

    = Vmox( 1 - ~22) = 2Vav( 1 - ~22) (8)

    where either the maximum velocity Vmax at r = 0 or the average velocity Vav defined by

    1 - ___!_ - _f!_ - 1 - !._ dr2 JR( d )R2( 2)

    Vav = A LV, dA - R2 o dz 4/L R2 1( dp)R 2 _ 1

    = 2 - dz 4~-t - 2 Vmax

    may be used to normalize the velocity profile. If the fluid is non-Newtonian, the elements Tji of the shear stress

    tensor must be used in the viscous terms of (3) rather than the velocity gradients. . .

    For cylinder symmetric flow in the z-d1rect10n (vr = v

  • 10 Mathematical Models from Chemical Engineering

    1.2 Steady State Homogeneous Flow Model for a Reacting System

    1.2.1. Derivation of a one-dimensional model

    Chap. 1

    Dimensionless independent variables { = z/ L and x = r/ R are intro-duced in (12) and (13):

    (14)

    (15)

    These equations are extremely simple compared to (1) to (3): The momentum balance has been solved independently of the mass and energy balances, the time dependence of c A and T has been dropped, and the transport properties are assumed constant. Even so, the resulting model consists of two coupled nonlinear partial differential equations (RA and Q depend nonlinearly on cA and T); its complete solution is a major numerical enterprise that is beyond the scope of this text.

    Simplifications of the two equations are indeed desirable and often possible in practice. Here we shall discuss how these simplifications can be brought about by discarding small terms and by suitable averaging methods that transform (14) and (15) into two coupled nonlinear first-order ordinary differential equations (25) and (26). In subsection 1.2.3, the axial dispersion terms on the right-hand sides are reintroduced-but the model still consists of ordinary differential equations. Finally, in subsection 1.2.4, the total set of side conditions of (14) and (15) are discussed and a particularly simple variant of the equations is treated a little further.

    In equations (14) and (15) the coefficients of the second-order deriva-tives with respect to { are frequently several orders of magnitude smaller than the coefficients of the first-order derivatives. Consequently these terms can frequently be neglected.

    On the other hand, the coefficients of the radial diffusion terms are quite large si\ce L/ R is usually 1. T?us, e~cept for unusually r.apid reactions, one may assume that the radial vanatwn of cA and T IS of minor importance; this indicates that the mean values of cA and T over the cross section of the reactor may be used rather than the x- and (-dependent variables in (14) and (15) if the purpose of the calculations is to compute the conversion of a given feed in a reactor of given length L.

    Sec. 1.2 Steady State Homogeneous Flow Model 11

    . In~egrating (14) ~nd (15) over the cross section, neglecting the axial dtffus10n terms and Imposing a condition of zero flux across the cylinder axis, yields

    I1 Vz acAd 2 LDr (acA) f1 L -- X = 2-2- - - 2 -R X dx 0 Vav a( R Vav ax x=1 0 Vav A (16) i 1 .!!:__ aT dx 2 = 2 2Lk r (aT) + 2 i 1 ~ Qx dx 0 Vav a( R VavPCp ax x=1 0 VavPCp (17)

    For an impermeable wall the concentration gradient at x = 1 is zero while by the film theory the temperature gradient is proportional to th~ difference between the reactor wall temperature T w and the fluid temper-ature Tx= 1 just inside the wall.

    kr(aT\ R a-;J x=

    1 = U(Tw- Tx=I) (18)

    Inserting in (16) and (17) yields

    I1 II Vz acA L 2 --;xdx = -2 -RAxdx 0 Vav a!:> o Vav (19) I

    1 Vz aT LU I 1 L 2 - -;:xdx = 2 . (Tw- Tx=I) + 2 --Qxdx

    0 Vav a!:, RvavPCp 0 VavPCp (20)

    The average concentration and temperature in the cross section of the reactor are defined by

    CAVav = I 1 2vzCA ((, x)x dx 0

    Tvav = I 1 2vzT({, x)x dx 0

    (21)

    (22) and fo~ sn:all radial concentration and temperature gradients the right-hand Side mtegrals of (19) and (20) are well approximated by

    I1 L L _ 2 -RAx dx = -RA (cA, T) (23) 0 Vav Vav f

    1 L L _ 2 --Qxdx = --Q(cA, T) (24)

    o VavPCp VavPCp

    Finally we wou~d.like to s~bstitute T w - Tx=I in (20) by a corresponding temperature '!nvmg force m terms of the average temperature f of (22). Now ITw- Tl is necessarily greater than ITw + Tx=II and in order to compensate for this a modified heat transfer coefficient [J (which is smaller than U) is introduced at the same time.

  • 12 Mathematical Models from Chemical Engineenng Chap. 1

    Our model (14) and (15) n_?W reduces to a one-dimensional form in terms of the variables cA and T defined by (21) and (22):

    (25)

    (26)

    A suitable choice of 0 is obtained by

    (27)

    where the best choice of a depends on the velocity profile as discussed in chapter 6. For a flat velocity profile, a = i is best. . . .

    Let us consider (25) and (26) for an nth-order Irreversible reaction where the rate constant k is given by an Arrhenius form:

    RA = kcA = a exp (- ~T)c':.. (28) and the heat production term Q arises due to the heat of reaction

    (29)

    In (28), a is a temperature-independent constant, E is the activation energy, and R 0 the universal gas constant. .

    Let the reactor inlet temperature and concentration ?e f = To and cA = co, respectively. RA can be rewritten into a dimensiOnless form:

    (30)

    where()= T/To, y = cA/co, k(To) = ko, and 1' = E/RaTo. Equations (25) and (26) now become

    ~ = - Da y" exp [ ~ 1 - ~) J (31)

    Sec. 1.2 Steady State Homogeneous Flow Model

    ~; = P Day~ exp [ ~ 1 - ~) J - Hw(IJ - Ow) D Lko n-1 a= -c0

    Vav

    {3 = c0(-dH) pcPT0

    Hw = 20 __!:__ R CpPVav

    13

    (32)

    The Damkohler number, Da, is a measure of rate of reaction at inlet conditions, y is a dimensionless activation energy, {3 is the adiabatic temperature rise relative to inlet temperature T0 if all the reactant is consumed, and Hw is a dimensionless heat transfer coefficient (the number of heat transfer units in the reactor).

    Thus the resulting reactor model is a set of two nonlinear first-order ordinary differential equations that are much easier to solve than (14) and (15). The major model simplification lies in the averaging process (16) and (17), which can also be used if the axial diffusion (or dispersion) terms of (14) and (15) are retained.

    In certain cases (31) and (32) can be solved in closed form. Thus for an adiabatic reactor ( 0 = 0), multiplication of (31) by {3 and addition to (32) yields

    d d((O + {3y) = 0

    or () + {3y is equal to a constant, which is determined from the inlet conditions 0 = y = 1 to be 1 + {3.

    or

    Now() can be eliminated from (31):

    dy n [ y{3(1 - y) J d( = -Day exp 1 + {3(1 _ y) = -F(y)

    C- Jl du Y F(u) (33)

    For each value of y the integral (33) must be found by a numerical method (except for yf3 = O) but tabulation of ( as a function of y is nevertheless a trivial problem. If y is desired at a fixed grid of C-values, (33) is interpreted as an algebraic equation in y. Inverse interpolation in a table of ( versus y may give sufficiently accurate values for y at the grid points.

  • 14 Mathematical Models from Chemical Engineering Chap. 1

    The final step in the solution of (31) and (32) for an adiabatic reactor is to determine the temperature profile from the algebraic equation () = 1 + {3[1 - y(()], which, of course, does not present any problem once y (() is known.

    It is of some interest to note that an analytical solution of (31) and (32) is possible also for a nonadiabatic reactor if E = 0, i.e., if the rate constant is temperature independent as is the case in the high tempera-ture S02-converter beds. After solving the mass balance and inserting y(() in the energy balance, this becomes a first-order inhomogeneous equation with constant coefficients. A "hot spot," i.e., a maximum in 0((), may well occur also when E = 0. After the hot spot, all reactant has been burnt off and the reactor acts as a simple heat exchanger. Exercise 1.1 treats a reactor where this simple solution is applicable.

    1.2.2. Limiting solutions to the full fluid-phase model

    The maximum effect of the simplifications leading from (14) and (15) to (31) and (32) can be estimated by a study of two limiting cases-either infinite or zero radial diffusivity. This will be done for an isothermal first-order reaction where (14) and (15) as well as (31) and (32) are particularly simple to solve.

    The full model for laminar Newtonian flow with first-order isothermal reaction in an empty tube is

    2 ay D'L 1 a ( ay) 2(1 - x )- = ---- x- - Day a( R 2 Vav X ax ax

    when the axial diffusion is neglected. The side conditions of (34) are CA y =- = 1 at ( = 0 Co

    ay = 0 at x = 0 and x ax

    Two limiting cases are considered.

    D'L 1. R 2~ ~ oo or y((, x) = y(() for all (

    1 for ( 2:: 0

    (34)

    With no radial gradients the solution of (34) and (31) are identical and the average outlet concentration

    Y(i: = 1) = r 4(1 - x2)y(1, x)x dx

    Sec. 1.2 Steady State Homogeneous Flow Model 15

    is found by solution of (31) with () = 1 and n 1: y(( = 1) = exp (-Da)

    2. In the absence of radial diffusion (D' = 0), a fluid element retains its identity throughout its passage of the reactor and

    y(( = 1) = 4 i 1 x(l - x 2 ) exp ( Da 2 ) dx o 2(1 -X )

    The ave:age ~o~centration at the rea_ctor outlet obtained from (34) for a fimte D hes between the solution of these limiting cases that are compared in table 1.3 for various values of the Damkohler number.

    TABLE 1.3 AVERAGE OUTLET CONCENTRATION FOR INFINITE AND ZERO

    RADIAL DIFFUSIVITY

    Da D' ~ oo D' = 0

    0.1 0.905 0.910 0.2 0.819 0.832 0.5 0.607 0.649 1.0 0.368 0.443 2.0 0.135 0.219

    It is seen that the model simplification is without practical conse-quences when the reaction is slow since both limiting models give the same result for small values of Da. It is shown at the end of the next subsection that a finite radial dispersion can be treated by means of an equivalent axial dispersion term; even though this introduces a second-or_der derivat~ve i? (31),. the resulting model is still considerably simpler than the partial differential equation (34).

    1.2.3 Axial dispersion of mass and energy

    The averaging process of subsection 1.2.1 can be performed even though the axial diffusion terms are retained. One obtains the following set of second-order nonlinear equations:

    p~M ~;; - ~~ - Day" exp H 1 %) ] = 0 (35) 1 d 2 e de [ ( 1)] PeH d(z- d( + {3 Dayn exp 'Y 1-0 - Hw(O- Ow)= 0

    (36)

  • 16 Mathematical Models from Chemical Engineering Chap. 1

    The two dimensionless groups PeM and PeH are axial Peclet numbers defined by

    P _ VavL eM--- and D

    The transport coefficients D and k which enter into the Peclet numbers are empirical quantities which bear no relation to the molecular properties. Experiments for a packed bed [Gunn {1971)] show that PeM is between 0.8 and 2 times the bed length per particle diameter ratio L/ dP for various packings. PeH is smaller, but of the same order of magnitude. Since L/ dP is very large in most industrial reactors, the numerical coefficient of the second-order terms are small. Since the second deriva-tives are of the same order of magnitude as the first derivatives, it is certainly reasonable to neglect the axial dispersion terms in a steady state calculation as we did in subsection 1.2.1.

    Models (35) and (36) are interesting, however, from a mathematical point of view; as we shall see at the end of the subsection, inclusion of an axial diffusion term may lead to a useful interpretation of the often far more important radial diffusion terms.

    Solution of the set of two second-order differential equations (35) and (36) requires two side conditions for each dependent variable.

    The correct specification of these side conditions is by no means an easy task and it might be helpful to consider the slightly simplified model where Hw = 0. Here y(~) is a monotonically decreasing function of ~ and () (~) is an increasing function when the reaction is exothermic. Mass or energy is carried toward ( = 1 by convective transport. The diffu-sion mechanism acts in the direction of decreasing y and () and conse-quently by this mechanism energy is carried toward ~ = 0 while mass is carried toward ( = 1.

    Both effects tend to decrease y and increase () near the reactor entrance and it would certainly be incorrect to take the side condition y = () = 1 at ~ = 0 from section 1.2 since obviously y < 1 and () > 1 at ~ = 0.

    Bischoff {1961) gives a very clear presentation of the side condition problem, and his procedure will be briefly reviewed for the mass balance since the resulting side conditions apply to the present example as well as to a more cpmplicated example in the next subsection.

    Assume *p;;tt the reaction section 0 ::::; ~ ::::; 1 is preceded by an en-trance section -oo 1.

    Sec. 1.2 Steady State Homogeneous Flow Model 17

    Mass balances for the two outer sections are

    1 d 2y dy Pe_ d(2 - d~ = 0 for ~ < 0 with y = 1 at ~ ~ -oo

    1 d 2 y dy Pe+ d~2 - d~ = 0 for ~ > 1 with y finite at ~ ~ oo

    The solutions of these equations are

    y = 1 + A 1 exp (Pe_ ~) for ' < 0 y = A 2 + A 3 exp (Pe+ ~) for ~ > 1

    (37)

    where A 3 = 0 to satisfy the condition at~~ oo. The solutions for the outer sections tell us that y decreases from 1 to

    1 + A 1 when~ increases from -oo to 0 and that y is constant (=A 2 ) for ~ > 1.

    Mass balances for differential volumes around ~ = 0 and ~ = 1 are

    1 dy_ 1 dy+ Y- - Pe_ d~ = Y+ - PeM d~

    1 dY_ 1 dY+ y_ - PeM d~ = y+ - Pe+ d~

    {38)

    where Y+ and Y_ are values of y inside the reactor, Y- = 1 + A 11 Y+ = A 2 , and the derivatives in the outer sections are found by differen-tiation of {37):

    dy _ _ dyj _ dY+ _ dyj _ dr - drl - - A1 Pe_ and d - d I - - 0 ~ ~ (-0- ( ~ (-1+

    Inserting these results into {38) one obtains 1- - _1_ dy+

    - Y+ PeM d( 1 dY_

    y_ - PeM d~ = Az

    {39)

    (40)

    We finally apply that the concentration profile is continuous across the boundaries at ~ = 0 and ~ = 1. Thus Y+ = Y- = 1 + A 1 and Y_ = Y+ = A 2 whereby {40) is reduced to

    1 dY_ dY_ PeM d( = 0 or d~ = 0 for any finite PeM (41)

    Equations (39) and {41) are the required set of boundary values for {35).

  • 18 Mathematical Models from Chemical Engineering Chap. 1

    The boundary conditions are independent of the rate expression in (35) and also independent of the specific values of Pe+ and Pe_, i.e., on the conditions in the outer sections.

    An identical derivation for the heat balance shows that the side conditions for (36) are completely analogous to (39) to (41) with (0, PeH) instead of (y, PeM)

    If Pe_ ~ oo, y very rapidly attains the value 1 for ( < 0 - which is the reason why (39) is often described as implying a discontinuity of y (or 0) at this point. The dependent variables are both continuous of course but the first derivatives are discontinuous except when Pe_ = PeM, i.e., when the diffusivity is the same in the entrance sector and in the reactor.

    It is also noted that the gradient at ( = 1 is zero only when PeM is finite and in fact nothing may be concluded in advance about the gradient at ( = 1 when PeM ~ oo. This is in full agreement with our expectations: (39) degenerates for PeM~oo to y = 1 at ( = 0 and (35) degenerates to the first-order equation (31) for which only one side condition is neces-sary.

    It is finally noted that direct integration of (35) and (36) for Hw = 0 and insertion of the side conditions derived above yields

    or

    1 - y(( ~ 1) ~ Da r exp [ ~ 1 - ~) k d( IJ(( = 1)- 1 = fJ Dar exp [ ~ 1- ~) k d(

    0 = 1 + {3(1 - y) at ( = 1

    Furthermore if PeM and PeH both approach zero, the model degenerates to the adiabatic stirred tank model for which 0 and y are constant and equal to their exit values 0(( = 1) and y(( = 1).

    Consequently the model with side conditions (39) to (41) and Hw = 0 correctly reduces to the adiabatic stirred tank model in the limiting case (PeM, PeH) = 0:

    0 = 1 + {3(1 - y) (42)

    Axial mass diffusion in an isothermal reactor implies that y(( = O) < 1 and that the concentration gradient is less steep than in an ideal plug flow reactor. The plug flow reactor gives a higher exit conversion than

    Sec. 1.2 Steady State Homogeneous Flow Model 19

    any reactor with axial dispersion andy(() must cross the plug flow profile exactly one time in 0 < ( < 1.

    The nonisothermal reactor might behave quite differently since the transport of energy toward ( = 0 greatly enhances the chemical reaction in the inlet section of the reactor and the conversion may in fact become higher than in a plug flow reactor.

    It is well known that the limiting case (42) of an adiabatic stirred tank reactor may exhibit three steady states even for first-order reactions. The other limiting case (31) and (32) is an initial value problem with no singularities in the derivatives and it has one and only one solution whatever the values of the parameters.

    Consequently one suspects that for a small value of the Peclet number and large enough values of the heat sensitivity parameters {3 and y, several solutions of the adiabatic reactor model with axial mixing may exist in a certain range of Da-values.

    Over the last 10 years numerous articles from chemical engineering literature have treated the two equations (35) and (36) and many interest-ing phenomena of multiple solutions have been found also for Hw = 0. A short list of references to some interesting papers on this subject is given at the end of the chapter.

    In table 1.3 we considered two limiting solutions of the partial differential equation (34), wJ:lich describes first-order isothermal reaction for laminar flow of the reactant in an empty tube. The case D' ~ oo corresponds to a plug flow reactor while the caseD'= 0 corresponds to a strong apparent axial dispersion of reactant: The fluid that reaches ( = 1 along an x -coordinate close to 1 has had an extremely long residence time in the reactor while the fluid in the center of the tube reaches the outlet after L/2vav time units. The result of the uneven distribution of residence times is clearly a much smaller conversion for large values of Da.

    The interpretation of a less than perfect radial mixing in terms of an axial dispersion coefficient was originally given by Taylor (1953). For flow in empty tubes a one-dimensional model approximation for a system with laminar flow requires the use of an "effective" axial diffusivity given by

    Dz ~ Dz + _...!:_ R 2Va/ eff- 48 D' or

    1 1 1 VavR 2 --=--+---PeM,eff PeM 48 D'L (43)

    If we consider the effect of the last term alone, we may compare the average concentration at the reactor outlet as calculated by (34) and by (35) for various values of the dimensionless group DrL/vavR 2 Table 1.4 shows the result of a calculation by the complete model (34) and by the "improved" averaged model (35) for Da = 1.

  • 20 Mathematical Models from Chemical Engineering Chap. 1

    TABLE 1.4 IMPROVED ONE-DIMENSIONAL REACTOR MODEL

    D'L/VavR 2 Yc=I (full model) Yc=I [by (35) and (43)]

    0.1 0.4038 0.4179 0.2 0.3931 0.3982 0.5 0.3809 0.3818 1 0.3750 0.3752

    It is seen that the Taylor dispersion concept permits a much better model approximation than the crude averaging process of subsection 1.2.1 that leads to either Yc= 1 = 0.3679 or to Yc= 1 = 0.443 for all entries of table 1.4 depending on whether complete mixing or no mixing at all is assumed.

    Wicke (197 5) reviews the application of Taylor dispersion for packed beds where numerical coefficients are selected that are different from those of (43).

    1.2.4 Boundary conditions for two-dimensional model-the extended Graetz problem

    After having discussed some important and (from a practical point of view) very reasonable simplifications of the complete model for the fluid phase, we shall briefly return to (14) and (15) and review the boundary conditions of these equations.

    For laminar flow in an empty tube, D and k can be taken to be equal in the radial and axial directions and they are molecular transport properties. The dimensionless model is

    ay 1 [a2y L 2 1 a ( ay)] n [~ 1)] Vz(x)-=----+---x- -Day exp 1-- =0 ac PeM a(2 R 2 X ax ax ()

    (44) ao 1 [a20 L 2 1 a ( ao)] n [~ 1)] vz(x)- = - - + --- x- + {3 Day exp 1 -- = 0 ac PeH a(2 R 2 X ax ax ()

    The assumption of Dz = D' and kz = k' is not correct for flow in a packed bed and the resulting model will become considerably more complex than ( 44).

    Sec. 1.2 Steady State Homogeneous Flow Model 21

    Boundary conditions at ( = 0 and at ( = 1 are taken from the discussion in subsection 1.2.3.

    c = 0: 1 ay 1 ao y---=1 and ()---=1 PeM ac PeH ac

    l = 1: ay = ao = 0 ac ac

    Symmetry across the cylinder axis gives

    ay = ao = o at x = o ax ax

    The tube wall is assumed to be impenetrable for the reactant

    ay = 0 at x = 1 ax

    (45)

    (46)

    (47a)

    The side conditions at x = 1 may be much more complicated for the energy balance. Assume that the tube wall is insulated from ( = 0 to C = ( 1 and that heat transfer to an exterior medium of temperature T w(C) is admitted for C 1 ~ C ~ 1:

    ae . - = 0 at X = 1 for 0 :::; ( < (1 ax (47b)

    Ow may be a constant or it may vary with (. For a given mass flow of coolant and a given inlet temperature T w(C = 0) or T w(C = 1), a total heat balance over part of the reactor plus heat exchanger gives an ordinary differential equation to determine T w(C) simultaneously with the solution of the partial differential equation model for the reactor. This type of side condition is discussed in chapter 4 for a considerably simpler model.

    It should be noted that ( 44) with its side conditions ( 45) to ( 4 7) is a boundary value problem in ( as well as in x. A considerable simplifica-tion is obtained if the axial diffusion terms, which are often quite insignificant, are neglected; the equations then become of first order in ( and can be solved by a marching technique from ( = 0. The somewhat complicated boundary condition ( 4 7b) at x = 1 presents no difficulty, the adiabatic and the nonadiabatic reactor differ only in the side condition for the energy balance and not in the structure of the differential equations.

  • 22 Mathematical Models from Chemical Engineering Chap. 1

    It may finally be observed that the two equations (44) as well as the side conditions can be decoupled if PeM = PeH. This leads to another significant simplification; however, this is not based upon a true physical description of the system since PeM is usually considerably greater than PeH.

    We shall repeatedly have occasion to discuss the solution of the linear partial differential equation that appears when the heat generation term is dropped from the heat balance (15).

    2vav0 - x2)aT = +[_!_ ~(x aJ\ + R2a2~] az R pep x ax a-; J az (48)

    A detailed analysis of this problem may give some insight into the difficulties of solving the complete model (14) and (15) without neglecting the axial diffusion terms.

    As a reference for the numerical solution of ( 48), which will be started ./ in chapter 4 and continued in chapter 9, we shall here give an outline of the dimensionless quantities that can be derived from the equation.

    The boundary conditions are imposed at z ~ -oo and at z ~oo and we shall assume that the tube is insulated from z = -oo to z = 0 and that the wall temperature is constant, T w for z 2::: 0. The Newtonian fluid enters the tube at z ~ -oo with T= T0 , flows in fully developed laminar flow through the tube, and attains the temperature T = Tw when z --? oo. The absence of a natural scale for the z -direction permits a small simplification of the model in comparison with ( 44) if the dimensionless variables are chosen as follows:

    () = 1 for (--? -oo a8 = o at x = 0 ax

    !~=0 at x= 1 ax

    z

    PeR

    and () is finite ( ~o) for ( ~ oo

    for all (

    ,

  • 24 Mathematical Models from Chemical Engineering Chap. 1

    the same position. This is certainly not true: The concentration of adsorbent is lower in the fluid phase inside the pore system of the adsorbent than in surrounding bulk fluid phase-otherwise the fluid phase would not be purified during its passage of the adsorption column. In steady state operation the temperature on the surface of a catalyst particle is higher than in the fluid at the same tubular position when the reaction is exothermic, and the external surface is again cooler than the interior of the particle since the major part of the conversion takes place on the large interior surface of the catalyst pellet.

    In this section we shall review the most common mass and heat transfer models for steady state and transient behavior of porous parti-cles. This is a necessary preliminary step if the reactor model of the previous section is to be refined-and the arguments for this refinement seem pretty strong. It is, however, important to strike a balance between degree of refinement and computational work. Model ( 44) for the fluid phase is quite complicated even without axial dispersion, and the occur-rence of a complicated coupling mechanism between fluid phase and solid phase may push the model beyond the capability of even a large computer-especially for transient studies or in steady state optimization of a reactor operation.

    We shall visualize the catalyst particle as a porous solid with twisting pores of different diameter and length, with crevices, dead-end pores, and perhaps a micropore system imbedded in a system of larger pores. These large pores serve as main passages for the fluid reactant that penetrates the solid and reacts on the pore walls.

    The heat of reaction Q = ( -dH)RA -which may be positive or negative-is conducted through the pore system and the solid matrix to the pellet surface and from there into the surrounding bulk fluid phase.

    Formally the mass and energy balances for pure diffusional flow of one reactant in a large surplus of inert are derived from (1) and (2) when the convective terms are dropped.

    aT_ ke ~zT 0 Pc- -v + pat - R 2 (56)

    The two transport coefficients De and ke, which are considered to be independent of spatial position in the pellet (either directly or implicitly through cA and T), are composite properties that reflect the different transport mechanisms inside the complex pore structure. Experimental values for De and ke are found in Satterfield (1970) pp. 65 and 171.

    Sec. 1.3 Steady State and Transient Models for Solids 25

    V2

    is the Laplacian operator for the solid phase. It may contain one, two, or at most three spatial coordinates in the pellet. For one-dimensional diffusion,

    (57)

    where ~ = 0, 1, an? 2 for planar, cylindrical, and spherical geometry, respectively, and x Is the ratio of the distance from the center measured relative to the pell~t radius or (for slabs) the half thickness of the pellet.

    The accumulatiOn term on the left-hand side of the mass balance contains the pellet porosity e since it is assumed that the accumulation occurs in the ga~ phase onl~. Correspondingly, p and cP are pellet density and heat capacity, respectively, and not properties of the solid matrix alone.

    A number of other physical processes of considerable interest to ch~mical engineering, e.g., leaching of solids, adsorption of vapors on sohds: an~ thermal treatment of solids, are described by one or both equatiOns m (56). The detailed treatment that we give these equations in sev~ral ~hapter~ of this text is well justified by their importance to engmeenng design and also by the importance of linear variants of the equations in theoretical numerical analysis.

    The pr_?duction. terms - RA and Q are, of course, not always caused ?Y a chemical reaction. In a physical adsorption process, RA is the rate of mcrea~e of adsorbed material on the solid and Q is the heat of adsorption per_ umt adsorbed material. The physical adsorption may be described as an mstantaneous reaction-the immobilized material on the solid surface is ~l~ays in equilibrium with material in the fluid phase at the same positiOn.

    For one spatial variable x the side conditions at x = 0 and x = 1 are

    acA aT - =- = 0 atx = 0 ax ax (58)

    acA RhM l i); = De [cb - cA(X = 1)] aT Rh at x = 1 ax = -,z-[Tb - T(x = 1)]

    (59)

    Tb a?d cb are p:op~rties of the bulk fluid phase. They are generally functiOns of positiOn m the reactor and of time but in the present section we shall regard them as known quantities. hM and h are mass )l.Dd-

    ,.""",#. ~

    // ~ .... -' . ~:f

  • 26 Mathematical Models from Chemical Engineering Chap. 1

    transfer coefficients for the film that surrounds the pellets. The two dimensionless groups

    (60)

    and

    (61)

    are called the Biot numbers for mass and heat transfer, respectively. The pellet model is seen to be a set of coupled nonlinear partial

    differential equations, which are again coupled to the fluid phase mass and energy balances through the boundary conditions (59). The fluid phase balances are themselves coupled partial differential equations with time and one or two spatial coordinates as independent variables.

    The solution of the complete set of balances for the coupled phases is a considerable task even in the steady state case, and development of a simplified pellet model is certainly warranted.

    Our objective is (1) to obtain a solution of the pellet model for the steady state problem, and (2) to discuss model simplifications for the transient calculations.

    One dimensionless form of (56) is

    ay 2 2 [ ~ / 1)] n aT = V y - exp 1'\ 1 - 0 y (62) (63)

    for an nth-order irreversible reaction.

    CA T y=- and T=-Co To

    where c0 is a reference concentration and T0 a reference temperature. T is a dimensionless time given by

    ffi

    (64)

    The dimensionless groups are the Thiele modulus , the Lewis number Le, {3, and y; {3 and y are the dimensionless heat generation and

    Sec. 1.3 Steady State and Transient Models for Solids

    activation energy, respectively.

    2 R2

    [ ( E )] n-1 =- aexp --- Co De RaTo

    Le = DePCp ke

    E y=--RaTo

    Finally, the dimensionless spatial boundary conditions are

    ay __ ao __ 0 at x = 0

    ax ax

    (ay) . ax x=l = BIM (yb - Yx=l)

    1.3.2 The steady state solution of the pellet model

    27

    (65)

    (66)

    (67)

    (68)

    We shall first consider the steady state solution of (62), (63), (66), (67), and (68). For convenience (c0 , T0) are chosen as (cb, Tb) and the steady state model is

    (69)

    (70)

    X= 0: dy = d(} = 0 dx dx

    :: = BiM (1 - y) and :: = Bi (1 - 0) (71)

    X= 1:

  • 28 Mathematical Models from Chemical Engineering Chap. 1

    In particular we are interested in the so-called effectiveness factor ,, the volume averaged reaction rate relative to the rate at bulk phase tempera-ture and concentration.

    J~ 2 rate (y, 8) dxs+l e s+l 'Y/ = J~ 2 rate (1, 1) dxs+I = J0 rate (y, 8) dx (72)

    In section 1.4 we use 'Y/ wherever possible to represent the solution of the pellet problem when the combined fluid-phase pellet model is solved.

    For industrially important problems, Bi is quite often small and the intraparticle temperature gradient may be negligible. This leads to a significant simplification of (69) and (70):

    Integrate (70) over the pellet volume to obtain

    (s + 1) Bi(1 - ii)- -{3

  • 30 Mathematical Models from Chemical Engineering Chap. 1

    energy balances for the fluid phase, an almost unsurmountable numerical task is at hand. Thus simplifications of the dynamic pellet model are certainly desired.

    1.3.3 Linearized dynamic model for the pellet

    For small deviations from a steady state solution, the rate expression in (62) and (63) may be linearized around this steady state.

    Let (Yss' Bss) be the solution to (69) to (71) and define deviation variables

    y(x, T) = y(x, T)- Yss(X) B(x,T) = B(x,T)- Bss(x)

    2y" exp [ ~ 1 - !J)] = 2y;,exp [ ~ 1- ~)] + YRy +OR, where

    The linearized transient equations are

    ay - t'72 " R " - R e" aT - v Y - yY o ao 2 A A A

    Le- = V B + {3Ryy + {3R 68 aT

    with boundary conditions

    ay = ao = 0 ax ax

    ay s "' o - + lMY = ax

    ae A -+BiB= 0 ax

    at x = 0

    atx = 1

    (79)

    (80)

    (81)

    (82)

    (83)

    Substitution of an exponential time dependence for y and 0 yields an eigenvalue problem. If all eigenvalues have negative real parts, the solution (y, 0) of (82) will decrease to zero for all x when T ~ oo. Under

    Sec. 1.3 Steady State and Transient Models for Solids 31

    these circumstances the steady state (Yss' Bss) is said to be asymptotically stable.

    This linear stability analysis is treated in detail in chapters 5 and 9.

    1.3.4 Simplifications of the nonlinear transient model

    If the effective heat conductivity ke of the solid is sufficiently large in comparison with the film transfer coefficient h (i.e., if Bi is small), the whole temperature increase occurs over the film, and the energy balance in (56) can be integrated over the system volume to give an ordinary differential equation in the average pellet temperature f:

    - - Il dT h - s+l pcP-d = (s + 1)-(Tb - T) + (-Il.H)RA dx t R o (84)

    where the replacement of Tb - Tx = 1 by Tb - f may as in the steady state model be compensated for by using a smaller value h of the heat transfer coefficient

    1 1 R - h ==-+a- or h =----h h P ke 1 + ap Bi

    The mass balance cannot be si~ilarly simplified since BiM 1:

    BiM = hMR = .!(2hMR)D0 = _! Sh D 0 De 2 Da De 2 De

    The Sherwood number (based on the reactant diffusion coefficient D 0 through the film) is never less than 2. D 0 may be several orders of magnitude larger than De, the "effective" diffusion in the porous pellet, and consequently BiM 1.

    Simplification of the transient mass balance is, however, possible but for quite different reasons.

    The parameter Le is generally very large, of the order of 10 to 1000, for most industrial reactions. This means that the concentration changes much more rapidly than the temperature of the pellet.

    Under these circumstances (pcP/ke)R 2 is a more natural time scale than (e/ De)R 2 that was used in (64), and (62) and (63) will appear as

    1 ay 2 2 n [~ 1)] - -- = V y - y exp 1 - -Le aT1 B

    :~ = V2 1J + {32y" exp [ ~ 1 - !J) J I ke

    Tt = -z-R peP

    (62a)

    (63a)

  • 32 Mathematical Models from Chemical Engineering Chap. 1

    Our assumption that the concentration profile is instantaneously estab-lished-the so-called quasi-stationary assumption for the mass balance-means that the left-hand side of (62a) can be equated to zero.

    The energy balance ( 63a) is handled by the previous simplification 1 (84) and we obtain the following set of equations:

    (85)

    Solution of (85) analytically (for n = 1) or numerically gives the concen-tration profile as a function of if just as in (74) and (75). The integral in (86) is evaluated and the dynamics of the pellet is described by a single ordinary differential equation in ii.

    This extremely simplified-and yet realistic-model retains several important characteristics of the complete model (62) and (63). For example, it allows a "quenched" state with a small reaction rate to slide into an "ignited" state with a large reaction rate and a significant temperature drop across a boundary film.

    1.3.5 Pellet parameters for a sulfuric acid catalyst

    A numerical example will indicate typical values of the pellet dimen-sionless groups for a gaseous reactant.

    Villadsen (1970) and Livbjerg and Villadsen (1972) present the fol-lowing data for a silica-based so2 oxidation catalyst:

    Pellet: = 0.4, De = 5.2 X 10-2 cm2 /s, ke = 7 X 10-4 cal/s em K, (pep) = 0.4 cal/ cm3 K, R = 0.3 em Reaction conditions: Inlet of an industrial converter. Pressure 1 atm, Cbo = Csoz = 1.6 X 10-6 mol/cm3 for 10 vol 0/o so2 in air, T0 = 77\K, superficial gas velocity v0 = 100 cm/s, and bed porosity Eb = 0.4, heat of reaction, 23,100 cal/mol Gas-phase properties at 500 C: Ph = 4.57 x 10-4 g/cm3, (cph = 0.24 cal/g K, Db = 0.655 cm2/s, J.Lb = 3.58 x 10-4 g/cms. kb = 9.19 x 10-5 caljcm s oK

    Sec. 1.3 Steady State and Transient Models for Solids 33

    With these data one obtains

    R 2 v0 Rpb e = = 77 J.Lb ' Pr = (J.L;p) b = 0.94, Sc = (..1!:_) pD b 1.26

    . 0.357 Jv [Satterfield (1970), p. 82] = eb Re0 .359 = 0.19

    jH jD = 1.37 for most gaseous reactants [Satterfield (1970), p. 83]*

    h = jH(pcPhvo Pr-213 = 2.97 X 10-3 caljcm2 s K . Rh

    BI = k = 1.27 e

    BiM _ hM ~ _ (Pr\ 213 jD ke ke Bi - h D - &J -:- (p ) n = 5450D = 73

    e }H CP /Y-Je e c (-dH)D {3 = bO k T. e = 3.6 X 10-3

    e 0

    L DePCp e = -- = 74 kee

    The value of {3 is so small that. in the absence of heat and mass film resistance the pellet will be at substantially the fluid-phase temperature To (d T = 2.~ oK if all reactant is co~verted in the pellet). For spherical e_ellets and B1 = 1.27 we predict an h value of 0.80h, while the value of h/ h that is obtained from the full model is 0. 78 (see section 6.2). Thus our model sim~li.fication (84) is fully justified. BiM is large and the simpler boun~ary cond1~10n. y = 1 at x = 1 could be used rather than ( 67)-the resultmg reduction m complexity of the model is, however, small.

    If the catalyst is extremely active, the reaction occurs on the pellet surface and ~ --- 0 at any point in the pellet, including x = 1. jj is given by (77) and It can be as large as 1 + {3(BiM/Bi) = 1 + 73{3 = 1.26-or d T = 200 K across the film.

    This maximum temperature increase is seen to be

    {3 BiM = jv(Pr)213(cb0(-dH)) ~ B. . S (p ) 0.6f3bo 1 ]H C CP bOTbo

    where f3bo is the dimensionless adiabatic temperature rise for a pure gas-phase conversion of the reactant.

    In practice this large temperature gradient across the film is of course never obtained since the reaction proceeds at a finite rate. ' '

    *Bird (1960, p. 647) uses iH = fv

  • 34 Mathematical Models from Chemical Engineering Chap. 1

    The parameter Le for the pellet dynamics is very large and our assumption of quasi-stationarity for the mass balance (62a) is certainly justified.

    Many theoretical studies of the dynamic behavior of catalyst pellets have been made with Le < 1 where-as introduced in chapter 9-most of the mathematically interesting phenomena occur. If for a moment we assume that Le 1, the energy balance (63) instantaneously follows even rapid variations in the concentration profile. Now, consider a pellet whose pore volume is filled with reactant at concentration cbo The pellet is suddenly ignited and an instantaneous temperature rise

    is registered if the rate of reaction is very fast. The transient temperature rise may be much larger than {3 when Le is

    close to zero; but as we have seen from the numerical example, the Lewis number is in reality much larger than unity and the energy balance reacts sluggishly to changes in concentration.

    1.4 Heterogeneous Model for a Reacting System

    In section 1.2 we derived homogeneous models for a reactor, that is, models where pellet reactant concentration and temperature are the same as in the fluid outside the catalyst pellet. In section 1.3, steady state and dynamic models for the pellet were discussed. A combination ~f res~lts from the two sections leads to a heterogeneous reactor model m wh1ch pellet phase and fluid phase may have different reactant concentration and temperature at the same position in the reactor. The two phases are coupled through a set of boundary conditions.

    1.4.1 Parameters and variables in heterogeneous model

    The dyq.amics of the reactor bed are characterized by two time constants tu~and tH. The first time constant tM is the ratio of the free reactor volume to the volumetric inlet flow rate, i.e., a measure of fluid residence time. The second time constant tH is the ratio of the reactor heat capacity to the heat capacity of the inlet stream. tH is consequently a measure of response time to thermal disturbances and is called the

    Sec. 1.4 Heterogeneous Model for a Reacting System

    thermal residence time.

    Leb (M=-Vo

    35

    Subscript b refers to the fluid phase and subscript 0 to inlet fluid conditions. When the same quantities are used for both fluid and solid phases, they are shown without subscript for the solid phase.

    The bed porosity eb is between 0.3 and 0. 7 and with a reactor pressure of the order of 1, the last term in tH is several orders of magnitude larger than 1. Hence the thermal residence time is much larger than the fluid residence time tM and in practice all capacitance terms except the thermal capacitance of the solid phase may be neglected. We shall see that this leads to a tremendous model simplification for transient calculations.

    It will be assumed that axial dispersion can be neglected and that radial gradients can be accounted for by using an effective wall heat transfer coefficient Ub defined by (27) with a = !.

    The fluid phase is taken to be a nearly ideal gas and the pressure drop in the reactor is neglected. For '!n equimolar chemical reaction the linear fluid velocity at reactor temperature Tb is

    Pbo Tb v = v0- = Vo-

    Pb Tbo

    1.4.2 Models for the two coupled phases

    For spherical pellets the fluid-phase mass and energy balances are

    2 a 2acb 2 3 -7TRb-(vcb) = eb7TRb- + (1 - eb)7rRb-hM(cb - C8 )

    az at R

    2a[ (p ] 2a(pcpTh 2 3 -7TRb- v cPTh = eb7TRb + (1 - eb)7TRb-h(Tb - Y's)

    az at R

    +27TRbUb(Tb - Tw) (I's, cs) are temperature and concentration, respectively, of the reactant on the surface of the pellets. (Tb, cb), the fluid-phase variables, are assumed to be constant in the cross section of the bed.

    Dimensionless variables Yb = cb/cb0 , ()b = Tb/Tbo, and ( = z/L are introduced. v = v0(Tb/Tb 0); (cph is independent of temperature and

  • 36 Mathematical Models from Chemical Engineering Chap. 1

    reactant concentration.

    The solid-phase balances are (56):

    ae D 2 ) e- = ---zV e - RA(e, T

    at R aT k 2 peP---;;( = R 2 V T + (-liH)RA (e, T)

    The subscript e that was used to denote effective transport coefficients ke and De has been dropped here to avoid confusion of nomenclature. Boundary conditions for the solid-phase balances are

    where x is the dimensionless pellet coordinate. The model simplification (85) and (86) is introduced into the solid

    balances:

    D 2 -R 2 V e = RA (e, T)

    at 3h - I 1 - 3 pep- = -R (Tb - n + (-liH) RA (e, T) dx at o

    The mass balance can be solved first, and the concentration profile e (x) is obtained as a function of eb and f.

    An effectiveness factor 17 is defined by the following relation:

    11RA (ebo, Tbo) = 11RAo = I 1 RA (e, t) dx 3 0

    Now the mass balance can be formally integrated- over the pellet volume

    D hMR 3 R2 v(eb - es) = 11RAo

    Sec. 1.4 Heterogeneous Model for a Reacting System 37

    and the energy balan_ce becomes

    or

    1 - eb peP aii 3h 1 - eb 1 - 1 - eb (-liH)e R ----- = -----(()b _ ()) + __ bO AO

    eb (pepho at R eb (pepho eb Tb 0(pepho rr;;;:;; The f~ctor ( -liH)ebo/ !'bo(peP ho is the dimensionless adiabatic temp-

    erature ~1se f3bo ~efined m subsection 1.3.5. The factor of aiij at was defined m subsectiOn 1.4.1 as (tHftM) - 1. Note that ii is a function of t as well a~ of ~he bed a~ial variable C, which explains why the pellet energy balance IS still a partial differential equation even after the averaging process over the pellet variable x.

    W_e shall finally introduce Hb and Hw, two dimensionless groups that contam the number of heat transfer units, fluid to pellets and fluid to reactor wall, for the reactor of length L:

    Hb = tM 1 - eb _1_ 3h = (1 - eb)L 3h eb (pepho R v0 (pepho R

    Hw = tM 1 2 Ub = 2LUb eb(pepho Rb Vo(pcphoRb

    The combined fluid-phase and pellet model is now

    -() ayb - aob - ayb 1 - eb RAo ayb baY Yb aY - tM-at + --tMrr-- = tM- + Da 11 ~ ~ eb ebo at

    aob -- ac = Hb(Ob - 0) + Hw(Ob - Ow) (87)

    ( )ao -tH - tM at = Hb(()b - 0) + f3bo Da 11 where Da = [(1 - eb)/eb]tM(RA 0 /eb 0 ) is the Damkohler number as defined in section 1.2 but based on the catalyst volume of the bed and the inlet superficial gas velocity v0

    The steady state model that is obtained by dropping the time deriva-tives in (87) is

    dyb aob -()b dC - Yb--;;f = Da 11

    40b -- dC = Hb(Ob - 0) + Hw(Ob - Ow) (88)

    0 = Hb(()b - ii) + /3bo Da 11

  • 38 Mathematical Models from Chemical Engineering Chap. 1

    The model consists of two coupled first-order ordinary differential equations coupled to an algebraic equation. The pellet steady state model is hidden in YJ = YJ(O, yb), which is obtained by (73) and (74) or for large values of Bi from the solution of (69) to (71).

    If the reactor is adiabatic, further simplification of (88) is possibl~ similar to (33) in section 1.2. Since Hw = 0 for an adiabatic reactor, (} can be eliminated between the second and third equation to give

    d(}b -- = -f3bo Da YJ

    d{

    The first equation is multiplied by f3bo and the two equations are added:

    d (f3bo0bYb + (}b) = 0 d{

    and (}b can be eliminated to give a final model that consists of one ordinary differential equation for Yb and two algebraic equations for (}b and if.

    The coupling of a set of ordinary differential equations with algebraic equations or the coupling of partial differential equations with ordinary differential equations in the boundary conditions as seen in subsection 1.2.4 is a very common feature of chemical engineering models.

    The transient model (87) is frequently-and with full justification according to the arguments of subsection 1.4.1-simplified by dropping the accumulation terms with factor tM. The resulting model is

    aYb aob -(}ba( - Yba( = Da YJ

    - aob = Hb(Ob - if) + Hw(Ob - Ow) a{

    ao --- = Hb((}b - 0) + f3bo Da YJ at/tH

    (89)

    The model is still a set of three coupled partial differential equations with yb, ob; ,&and if as dependent variab.les and with the pellet dynamics hidden in YJ, which is given by the solutiOn of (85) and (86).

    The time constant t for transport of heat from the pellet to the fluid p -phase is obtained if (86) is divided by 3 Bi for spherical pellets:

    R C d 0 - {3 cf.>2 [ ~ { 1 ) J J 1 n 3 P _P - = ( ob - (}) + -= exp 'Y\ 1 - -= y dx 3h dt 3 Bi 0 o

    Sec. 1.4 Heterogeneous Model for a Reacting System

    or

    t = Rp~p = tH P 3h Hb

    as also seen from the last equation of (89).

    39

    tP may well be much smaller than tH if Hb is large but it is certainly much larger than the internal pellet time constant, which has been neglected in our previous averaging process (84).

    If tP is much smaller than tH, that is, if we assume if to follow (}b instantaneously-the so-called "chromatographic assumption" of Aris and Amundson (1973, p. 35), because it corresponds to an instantaneous local equilibrium between the solution and adsorbed species in a chromatographic column-then additional simplification of the model (89) is possible. Subtraction of the middle equation of (89) from the last equation yields

    (90)

    The consequence of this last simplification is best seen by consideration of a nonreactive system (Da = O) where (90) is a first-order linear partial differential equation in (}b This is solved by the method of characteristics as described in Aris and Amundson (1973) to give a temperature front that marches with constant velocity litH through the reactor without changing shape.

    Thus by neglecting the heat transfer resistance between pellets and fluid phase, our dynamic model becomes unable to predict the broadening of sharp temperature disturbances that is observed in experimental reac-tor dynamic studies. Slowly changing temperature disturbances may still be accurately simulated but the model is unable to cope with high-frequency disturbances.

    A dispersive element may be included in the transient reactor model even 'when the pellet dynamics is neglected. Most industrial reactors are built of high heat capacity material and the dynamics of the reactor wall may well be much more important (large time constant) than the pellet dynamics [see, e.g., Hoiberg and Foss (1971)]. This coupling element has not been considered in our models but it may easily confound a labora-tory study of reactor dynamics.

    It should finally be mentioned that axial dispersion can be included in the model without increasing its complexity. Some investigators, e.g., Wrede Hansen (1974) and S0rensen (1976), collect all dispersive effects (pellet fluid, fluid wall, etc.) into an effective axial heat dispersion term in the same way as real gradients are handled in subsection 1.2.3. If a suitable Pe-number can be derived from experiments to describe this

  • 40 Mathematical Models from Chemical Engineering Chap. 1

    lumping process, the method leads to the same dynamic results as in the present treatment; this does, however, have the benefit of operating with real physical properties.

    1.5 A Model for Hollow-Fiber Reverse-Osmosis Systems

    The flow systems of the previous sections were all characterized by a unidirectional convective flow. This greatly simplified the solution of Navier-Stokes equations and the complications arose due to diffusional effects in the fluid or in a solid phase that was coupled to the fluid phase.

    In this and the next section we deviate somewhat from the main stream of our text, which by and large deals with solution of models taken from the previous sections. The excursion into reverse osmosis here or into extrusion of polymers in the next section is not motivated by a desire to treat these subjects on their own merits. Rather we wish to stress that model building is certainly more complicated than we have shown so far.

    1. Description of the process in a hollow-fiber membrane: Heavy-walled hollow cylindrical membranes with an internal diameter of 50-20 microns are spun from a semipermeable material and imbedded in an epoxy tube sheet that is housed in an exterior shell. Pressurized feed solution flows over the fibers in the shell. The product water permeates the fibers and flows inside them toward the open ends, which are at atmospheric pressure. A very large membrane area per unit volume is provided and concentration polarization-a rna jar source of efficiency loss in reverse osmosis (or ultrafiltration)-is negligible due to the permissible small pro-duct flux. These features of the hollow-fiber membrane make it popular in a variety of situations: desalination and waste water treatment in general-as well as in biological separations where the treatment of uremia by hemodialysis is only one example.

    The setting up and analysis of a model for the highly complicated flow pattern in a hollow-fiber agglomerate is important to establish rational design criteria for wall thickness, outside fiber radius, and fiber length and to determine the capacity at given feed flow rate, operating pressure, and inlet feed ~ncentration. Optimal values of outside fiber radius r0 and fiber length L can be determined for given operating conditions and this has been the object of the study by Gill (1973). His model looks considerably more complicated than the one that is derived below, but when the various assumptions that he makes through his derivation are introduced immediately, the results become identical.

    Sec. 1.5 Hollow-Fiber Reverse-Osmosis Systems 41

    2. Assumptions in the mathematical model: The arrangement of fibers is shown in figure 1-1(a). Variations in concentration and

    (a)

    -----------------------------

    Phase 1

    --------------------(b)

    Figure 1-1. (a) Hollow fiber reverse osmosis. (b) Arrangement of fibers and fluid flow directions through phases 1-3.

    pressure are small around the periphery of the imaginary hex-agonal boundary th~t surrou~ds each fiber, and this boundary is replaced by an eqmvalent cucle having the same annular area between it and the outside fiber wall.

    ~e radius re of the equivalent annulus is determined by the porosity of the bed s and the outside fiber radius r0 :

    - ( 1 ) 1/2 r = -- r e 1- c 0 (91)

  • 42 Mathematical Models from Chemical Engineering Chap. 1

    The velocity vector v for the fluid flow through the system shown in figure 1-1 (b) has two components Vr and Vz Since the system is composed of three phases, (1) outside the fibers (i = _1), (2) in the fiber wall (i = 2), (3) in the passage on the product si~e of the fiber (i = 3), and the two velocity components are used m all three phases, it is convenient to use u; for Vzi and vi for vri to avoid double indexing.

    It is assumed that the fluid is Newtonian and that steady state laminar flow exists in the system. Both pressure and concentration decrease significantly in the axial z -direction due to friction loss and product removal.

    The simultaneous solution of nine coupled nonlinear partial differen-tial equations (two equations for v and one mass balance f~r e~c~ pha~e) is an almost impossible task. Luckily a number of Simphfymg cir-cumstances exist that drastically reduce the computational work.

    Assumption 1: The equation of motion can be solved s_eparat~ly. The present system admits to the simplification discussed _m sect10n 1.1 since v is independent of concentration and the process Is assumed to be carried out isothermally.

    Assumption 2: Axial diffusion terms are negle_cted in compariso_n with axial convective terms, and the radial velocity component V1 Is thus independent of axial distance z. The resulting velocity field (92) and (93) for phase 1 contains considerably fewer terms than the parent equations (7):

    u au1 + v au1 = _ .!_ ap1 + v_! ~(r au1) 1

    az 1 ar p az r ar ar (92)

    v av1 = _.!_ ap1 + v .!(.! ~(rv1)] 1 ar p ar ar r ar

    (93)

    Assumption 3: The small flow v1w out of phase 1 probably rules o_ut the possibility of concentration polarization that would set up radml concentration gradients. Thus c1 is assumed to be independent of r and a mass balance over the annular volume V1 = 1r(r; - r~) dz will give a relation between the loss of solute from the bulk phase 1 and the am~nt that has penetrated the membrane at radius ro:

    2 2 d(u1avC1w) (1 K)2 1r(r - Yo) = - - 1TYoV1wC1w e dz (94)

    K is called the rejection coefficient of the membrane. If K is close to unity, almost no solute enters the fiber wall; a value of K close to

    Sec. 1.5 Hollow-Fiber Reverse-Osmos1s Systems 43

    zero means that the solute is strongly retained (absorbed) by the fiber material. K is clearly an empirical constant for the fiber membrane. It is found to vary only a few percent with the main design objective, which is the productivity , defined by the area-averaged linear velocity in phase 1,

    = U1av(Z) - U1av(Z = 0) U1avCz = 0) (95)

    and the representation of c 1 w by (94) with a constant K is reasonable for less than approximately 0.5-0.6.

    ~e area-averaged axial velocity u1av is defined similarly to (94) by a ~md_ b_alance from z = 0 to z using the wall radial velocity v

    1w,

    which IS mdependent of z by assumption 2.

    ( )=_!_J ( _ 2r0 U1av z A u1 r, z) dA1 - U1av(Z = 0)- -2--2v1wZ 1 At Ye - Yo (96)

    ~quation (96) shows that du 1av/ dz = -[2r0 /(r; - r~)]v 1 w, which is mserted into (94) to give

    (97)

    Integration of (97) from z = 0 to z with the initial condition c = c0

    and U1av = U1aiz = 0) at z = 0 immediately gives the concentration of solute at the wall as a function of z:

    C1w(z) = [U1av(Z = O)]K Co U1av(z) (98)