7“ nuwycn rr.y 5‘ I v. o. 043. .‘hl'a 7 41;" .-' ‘ fi-v'u #53:; - 399135552”ij . ‘- (2‘15": .51.? 'b'lA4 )j': I"... 151;} \‘Irl‘lm ’33:; m. $29“ 1": I“: ‘< ‘3 A t,;2_: .. "um ”i 1‘ r523... 31%; ’50} “3’;”r...c 92w ;.5. '25: 5 .. ' r‘flj .5335 52"“355'52 "(5' '55‘3’ ' «'55. . '3; #52135?!“ -—n~ “$2.5,ou y“ ’ " t' ‘ :II '- l ~2». firm OLVI we??? ‘ l . I ‘3'“; ‘ '12.".‘u . . 8}" H" 51"? ”PM" ' . H'- . W? J 2 WHY! ‘JMX . 22' . h ’35. . 52-5: '._v.;,."§m{h “7;” n fir' 3’3“, ”i; ‘fiv‘ "12"", "W ”I“ 5;“. “. $741; “"2 .55. .1 $4 2.12.»... . W... L rid", .. '... I“$5.”; PM“: '51-: 55- ‘ firm“ .. .. mu 4. y’a- WK :4.‘ ,t‘ W" 0‘" t“: Luggagl‘lk» :3: ~91 We?“ ‘2 $.13, 35651.. 5% ‘7 vnvlul "'",“:\' ¥ 7.557 J Méé‘fl ,{L W .3555; {5552.555 ¢ , ~35”! 74325 ,_{:r&§,._ 1“ 3333143.“,«9 gifts- '1;' ,. "r. "'J i 5595;... , 2U" w‘f‘i‘ owl”! vESlURASRTYBRIE 1111 11111121111 111 111111111111111111111111111111 31293 0091 This is to certify that the dissertation entitled THE NUMERICAL SOLUTION OF CONFINED LAMINAR FLOW PAST A MOVING BOUNDARY presented by Bashar Sulaiman AbdulNour has been accepted towards fulfillment of the requirements for Doctor of Philosophy degree in Mechanical Ergneering Maggie Major professor Merle C. Potter MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 r LIBRARY Michigan State University \_ I T‘ PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE 1F 11 11 filT—T MSU I: An Affirmative Action/Equal Opportunity Institution emulation”: THE NUMERICAL SOLUTION OF CONFINED LAMINAR FLOW PAST A MOVING BOUNDARY By Bashar Suleiman AbdulNour A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 1990 ABSTRACT THE NUMERICAL SOLUTION OF CONFINED LAMINAR FLOW PAST A MOVING BOUNDARY By Bashar Suleiman AbdulNour The solution of the time—dependent, viscous, incompressible, two-dimensional, confined flow past a moving boundary is considered. The problem is modeled as a laminar flow in the entry region of a rectangular channel with uniform inlet velocity upstream and a fully-developed parabolic profile downstream. A numerical solution is obtained from the continuity and Navier-Stokes equations and the appropriate boundary conditions. It provides an exact solution over the entire domain, in the sense that it utilizes all terms in the describing equations. The vorticity-stream function formulation is selected for the treatment of the problem. After discretizing the non-dimensional governing equations in finite-difference form, the Altemating-Direction Implicit method is used to solve the vorticity equation while the iterative Successive Over-Relaxation method is used to solve the stream function equation. The vorticity wall condition is derived by expanding in a Taylor series. This second-order scheme maintains the diagonal dominance of the system of equations at high Reynolds number, thereby expanding the domain of convergence. The vorticity, stream function, and velocity components are calculated at each grid point for Reynolds numbers ranging between 10 and 5000. The plane channel flow is first investigated. This steady-flow solution is obtained as the asymptotic solution for large time of the unsteady flow. The results confirm two primary conclusions: the local maxima in the streamwise velocity profile near the wall is found to be minimal, and the inviscid-core region is approximately one-fifth of the entrance length with a profile-development region accounting for the rest. The use of accelerating parameters to control the rate of convergence has significantly improved numerical stability and reduced the time of computation. When a finite step is fixed to the lower boundary, however, distinct separated regions with pronounced eddies of recirculating fluid are developed on both sides. Motion of part of the smooth lower boundary is then included; both steady and oscillatory motion are studied. The size of a separated region, formed when the boundary is moving opposite to the flow, depends on the Reynolds number, Strouhal number, and velocity of the boundary. Cepyright by Bashar Sulaiman AbdulNour 1990 To my father Sulaiman and my mother Lydia for their infinite love and unequivocal support. ACKNOWLEDGMENTS Foremost, the author wishes to express his enormous gratitude and indebtedness to his major professor, Dr. Merle C. Potter. Mere words cannot convey this since Merle was not only an academic advisor, but also a true friend and a significant influence. Most importantly, the author greatly appreciates the remarkable interest and the invaluable educational experience gained from his association with Professor Potter. The author is also much obliged to his guidance committee members Professor James V. Beck and Professor David H. Y. Yen for their time and contributions. In addition, sincere thanks are extended to Dr. Craig W. Somerton for his constructive comments and to Dr. Reinier B. J. Bouwmeester for his early participation. Above all, without the strong encouragement and understanding that the author received from his family during his college career, this effort would have not been possible. Finally, the author’s faith in the Lord Jesus Christ was a primary source of inspiration. TABLE OF CONTENTS LIST OF TABLES ............................................................................................................ ix LIST OF FIGURES ........................................................................................................... x NOMENCLATURE ....................................................................................................... xvi 1. INTRODUCTION .................................................................................................... l 2. BACKGROUND ...................................................................................................... 8 2.1 Entry Flow in a Pipe ........................................................................................ 9 2.2 Entry Flow in a Channel ................................................................................ 14 2.3 Internally Separated Flows ............................................................................ 20 2.4 Flow in a Driven Cavity ................................................................................. 23 2.5 Externally Separated Flows ........................................................................... 27 2.6 Time-Dependent Flows .................................................................................. 28 2.7 Numerical Treatments of Flow Problems ...................................................... 3O 3. DESCRIPTION OF THE PROBLEMS ................................................................. 32 3.1 The Flow Problems ........................................................................................ 32 3.2 Mathematical Formulation ............................................................................. 35 3.3 The Problem for Solution: Vorticity-Stream Function Formulation ............. 39 3.4 The Steady Flow Problem .............................................................................. 42 3.5 Initial and Boundary Conditions .................................................................... 43 4. THE NUMERICAL METHODS ............................................................................ 51 4.1 The Computational Domain ........................................................................... 52 4.2 The Alternating-Direction Implicit Method for the Vorticity Equation ....... 52 4.3 The Successive Over-Relaxation Method for the Stream Function Equation ......................................................................................................... 61 4.4 Initial and Boundary Conditions .................................................................... 67 5. THE COMPUTATIONAL SCHEME .................................................................... 72 5.1 Accelerating Parameters ................................................................................ 72 5.2 Outline of the Computational Procedure ....................................................... 75 5.3 Computing Environment ................................................................................ 80 6. ANALYSIS ............................................................................................................. 81 6.1 Error Estimate for Vorticity Boundary Conditions ........................................ 81 6.2 Consistency and Errors of a Numerical Solution ........................................... 83 6.3 Numerical Stability of the Differenced Navier- Stokes Equations ................ 85 7. DISCUSSION OF RESULTS AND CONCLUSIONS .......................................... 89 7.1 Entry Flow in a Plane Channel ...................................................................... 89 7.2 Channel Flow with a Finite Step .................................................................... 95 7.3 Channel Flow with a Moving Boundary ........................................................ 98 7.4 Channel Flow with an Oscillating Boundary ............................................... 101 7.5 Accelerating Parameters .............................................................................. 104 7.6 Conclusions .................................................................................................. 108 APPENDIX A - COMPUTER PROGRAM .................................................................. 170 BIBLIOGRAPHY .......................................................................................................... 188 viii Table 1. Table 2. Table 3. Table 4. Table 5. LIST OF TABLES The inviscid-core length Li, the profile-development length L d, and the entrance length L e for various Reynolds numbers. ........................ 91 Comparison of the ratio L¢/(I-I-Re) as reported in different channel studies for Reynolds number = 200. .............................................. 92 Length of the aft separated region L2 for the channel flow with a finite step. .................................................................................................... 96 Optimum accelerating parameters, CPU time, and number of outer iterations for the channel flow with a uniform mesh of size 0.05 at various Reynolds numbers; the solution is based on one half the height of the channel. ................................................................................ 105 Optimum accelerating parameters, CPU time, and number of outer iterations for the channel flow with a uniform mesh of size 0.05 at various Reynolds numbers; the solution is based on the full height of the channel. ........................................................................................... 106 Figure 1. Figure 2. Figure 3. Figure 4. Figure 5. Figure 6. Figure 7. Figure 8. Figure 9. Figure 10. Figure 11. Figure 12. LIST OF FIGURES A schematic representation of the flow problem including the moving boundary. ....................................................................................... 33 Laminar entry flow in a two-dimensional channel. .................................... 34 Schematic representation of the sequence of entry flow problems. ............ 36 Dimensionless boundary conditions. .......................................................... 48 The flow problem with the finite step placed on the lower boundary. ....... 49 Fume-difference representation of the computational domain. .................. 53 Flow chart of the computational procedure. ............................................... 79 Development of streamwise velocity profiles through the entry region of a two-dimensional channel at various Reynolds numbers. ....... 111 Streamwise velocity profiles at locations near the inlet section for various Reynolds numbers. ....................................................................... 114 Development of streamwise velocity profiles through the entry region of the channel using the boundary-layer equations at Re = 200. ........................................................................................................... 115 Development of streamwise velocity profiles through the entry region of the channel using a cascade inlet condition at Re = 200. .......... 115 Development of streamwise velocity profiles through the entry region of the channel using an actual experimental inlet velocity profile at Re = 200. ................................................................................... 115 X Figure 13. Figure 14. Figure 15. Figure 16. Figure 17. Figure 18. Figure 19. Figure 20. Figure 21. Figure 22. Figure 23. Figure 24. Figure 25. Figure 26. Figure 27. Streamwise velocity development along the channel at different transverse locations across the channel for Re = 200 and 2000. .............. 116 Centerline velocity along the channel for various Reynolds numbers. 117 Development of normal velocity profiles through the entry region of a two-dimensional channel at various Reynolds numbers. ................... 118 Development of vorticity profiles through the any region of a two-dimensional channel at various Reynolds numbers. .......................... 119 Vorticity at the upper boundary along the channel at various Reynolds numbers. .................................................................................... 120 Vorticity at the upper boundary along the channel calculated from different formulas for Re = 200. ............................................................... 120 Vorticity contours of the flow in the entry region of a two- dimensional channel at Re = 200. ............................................................. 121 Streamlines of the flow in the entry region of a two-dimensional channel at Re = 200. ................................................................................. 121 Normalized pressure gradients at the wall along the channel for various Reynolds numbers. ....................................................................... 122 Normalized pressure gradients at the centerline along the channel for various Reynolds numbers. ................................................................. 122 Normalized pressure gradients at the wall and centerline along the channel for Re = 200. .......................................................................... 123 Streamwise velocity as a function of the mesh size at selected locations; Re = 200. .................................................................................. 124 Flow separation due to a finite step of various width; X1 = 1.0, B = 0.25, Re = 20. ......................................................................................... 125 Flow separation due to a finite step of various height; X1 = 1.0, D = 0.05, Re = 20. ......................................................................................... 127 Flow separation due to a finite step of various width; X1 = 2.0, B = 0.20, Re == 200. ....................................................................................... 129 Figure 28. Figure 29. Figme 30. Figure 31. Figure 32. Figure 33. Figure 34. Figure 35. Figure 36. Figure 37. Figure 38. Figure 39. Figure 40. Flow separation due to a finite step; Re = 200, X1 = 2.0, B = 0.25, D = 0.05. ................................................................................................... 130 Flow separation due to a finite step of various height; X1 = 1.0, D = 0.05, Re = 200. ....................................................................................... 130 Effect of step location on the flow separation due to a finite step; X1=8.0,B=0.20,D=0.05,Re=200. ................................................... 131 Velocity vectors of the flow in the presence of flow separation due toafinitestep;Xl=1.0,B=0.20,D=0.05,Re=20. ............................ 131 Development of streamwise velocity profiles along the channel in the presence ofafinite step;Xl = 1.0,Re=20. ....................................... 132 Development of streamwise velocity profiles along the channel in the presence of a finite step; X 1 = 2.0, B = 0.2, Re = 200. ....................... 133 Development of streamwise velocity profiles along the channel in the presence of a finite step; X1 = 1.0, B = 0.25, D = 0.05, Re 2 200. ........................................................................................................... 134 Vorticity contours in the presence of flow separation due to a finite step; X1 = 1.0, Re = 20. ................................................................... 134 Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB; X1 = 1.0, X2 = 2.0, Re = 20. ........... 135 Velocity vectors of the flow with the lower boundary section moving opposite to the inlet velocity at UB = -10.0; X1 = 1.0, X2 = 2.0, Re = 20. .............................................................................................. 136 Streamlines of the flow with the lower boundary section moving in the direction of the inlet velocity at UB; X1 = 1.0, X2 = 2.0, Re = 20. ............................................................................................................. 137 Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB; X1 = 1.0, X2 = 2.0, Re = 200. ......... 138 Streamlines of the flow with the lower boundary section moving in the direction of the inlet velocity at UB = 1.0; X 1 = 1.0, X2 = 2.0, Re = 200. ................................................................................................... 139 Figure 41. Figure 42. Figure 43. Figure 44. Figure 45. Figure 46. Figure 47. Figure 48. Figure 49. Figure 50. Figure 51. Figure 52. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB = -1.0; X1 = 1.0, X2 = 5.0, Re = 200. ........................................................................................................ 139 Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB; X1 = 4.0, X2 = 5.0, Re = 200. ......... 140 Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB; X1 = 4.0, X2 = 5.0, Re = 200. (A close up of Frg. 42.) ............................................................................. 141 Streamlines of the flow with the lower boundary section moving at U13;Xl =2.0,X2=4.0,Re=200. ........................................................... 142 Streamlines of the flow with the lower boundary section moving at UB; X1 = 2.0, X2 = 6.0, Re = 200. ........................................................... 143 Streamlines of the flow with the lower boundary section moving at UB; X1 = 8.0, X2 = 10.0, Re = 200. ......................................................... 144 Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB = -0.5; X1 = 10.0, X2 = 14.0, Re = 1000. ................................................................................................. 145 Streamlines of the flow with the entire lower boundary moving in the direction of the inlet velocity (Couette-Poiseuille flow) at UB = 1.0; Re = 500. ........................................................................................ 145 Development of streamwise velocity profiles along the channel with the lower boundary section moving at UB; X1 = 2.0, X2 = 4.0, Re a 200. ................................................................................................... 146 Development of streamwise velocity profiles along the channel with the lower boundary section moving at UB; X1 = 2.0, X2 = 6.0, Re a 200. ................................................................................................... 147 Development of streamwise velocity profiles along the channel with the lower boundary section moving opposite to the inlet velocity at UB = -0.5; X1 = 10.0, X2 = 14.0, Re = 1000. ......................... 148 Development of streamwise velocity profiles along the channel with the entire lower boundary moving in the direction of the inlet velocity (Couette-Poiseuille flow) at UB . ................................................. 148 xiii Figure 53. Figure 54. Figure 55. Figure 56. Figure 57. Figure 58. Figure 59. Figure 60. Figure 61. Figure 62. Figure 63. Figure 64. Vorticity contours with the lower boundary section moving at UB; Xl =1.0, X2 = 2.0, Re = 20. ..................................................................... 149 Vorticity contours with the lower boundary section moving at UB; X1=1.0,X2=2.0,Re=200. ................................................................... 150 Vorticity contours for the entire lower boundary moving in the direction of the inlet velocity (Couette-Poiseuille flow) at UB = 1.0; Re = 500. ............................................................................................ 150 Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 1.0, X2 = 2.0, Re = 20, St= l.0,A= l.0,n=8. ............................................................................. 151 Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 2.0, X2 = 4.0, Re = 200, St= 1.0,A=2.0,n=25. ........................................................................... 152 Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 2.0, X2 = 6.0, Re = 200, St = 0.5, A = 2.0, n = 14. ........................................................................... 153 Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 1.0, X2 = 2.0, Re = 200, St= 1.0,A= l.0,n=21. ........................................................................... 154 Size of the separated region as a function of the Strouhal number; X1 =1.0, X2 = 2.0, Re = 200, A = 1.0, t = (2n+1)n, UB = -1.0. ............. 155 Size of the separated region as a function of the Strouhal number; X1 = 2.0, X2 = 4.0, Re = 200, A = 1.0, t = (2n+1)rt, UB = -1.0. ............. 156 Size of the separated region as a function of the Strouhal number; X1: 2.0, X2 = 6.0, Re = 200, A = 0.5, t = (2n+ l)rt, UB = -0.5. ............. 157 Flow sueamlines showing the size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 1.0, x2 = 2.0, Re = 20, St = 2.0,t=(2n+1)1t. ................................................. 158 Size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 1.0, X2 = 2.0, Re = 200, St = 1.0, t= (2n+l)rt. .............................................................................................. 159 Figure 65. Figure 66. Figure 67. Figure 68. Figure 69. Figure 70. Figure 71. Figure 72. Figure 73. Figure 74. Figure 75. Size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 2.0, X2 = 4.0, Re = 200, St = 2.0, t = (2n+ l)x. .............................................................................................. 160 Size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 2.0, X2 = 6.0, Re = 200, St = 1.0, t= (2n+l)1t. .............................................................................................. 161 Streamlines of the flow with the lower boundary section oscillating at UB = Acos(t); Re = 1000, A = 1.0, t = (2n+l)1t, UB = -1.0. ........................................................................................................ 162 Effect of the location of the oscillating section of the boundary on the size of the separated region; X1 = 5.0, X2 = 6.0, Re = 200, St =1.0,A= l.0,t=(2n+l)1t,n=28,UB =-l.0. ........................................ 163 Effect of the length of the oscillating section of the boundary on the size of the separated region; X1 = 1.0, X2 = 3.0, Re = 200, St =1.0,A=1.0,t=-=(2n+1)rt,n=26,UB =-1.o. ........................................ 163 Effect of the location and length of the oscillating section of the boundary on the size of the separated region; X1 = 8.0, X2 = 10.0, Re=200,t=(2n+l)x. ............................................................................. 164 Development of streamwise velocity profiles along the channel at the instant when the boundary is moving at maximum velocity in the direction of the flow; Re = 200, t = 2n1t. ............................................ 165 Development of streamwise velocity profiles along the channel at the instant when the boundary is moving at maximum velocity in the direction opposite to the flow; Re = 200, t = (2n+1)1t. ...................... 166 Vorticity contours for the case of the lower boundary section oscillating at UB = Acos(t); X1 = 1.0, X2 = 2.0, Re = 20, St = 1.0, A= l.0,n=8. ........................................................................................... 167 Vorticity contours for the case of the lower boundary section oscillating at UB = Acos(t); X1 = 1.0, X2 = 2.0, Re = 200, St = 1.0,A=1.0,n=21. .................................................................................. 168 Vorticity contours for the case of the lower boundary section oscillating at UB = Acos(t); X1 = 8.0, X2 = 10.0, Re = 200, St = l.0,A=1.0,n=24. .................................................................................. 169 XV gvazz_rargrwrmmauw> NOMENCLATURE amplitude of oscillation of boundary velocity height of the finite step length of the finite step over-relaxation factor for the stream function equation gravitational acceleration channel height mesh size (equal in the x- and y-directions) time step profile-development length entrance length inviscid-core length number of grid points in the x-direction number of grid points in the y-direction number of cycles of oscillation for convergence pressure kinetic pressure Reynolds number Strouhal number time uniform velocity at the inlet velocity of the moving section of the boundary streamwise velocity component (in x-direction) normal velocity component (in y-direction) velocity component in the z-direction upstream location of the oscillating boundary or the step downstream location of the oscillating boundary or the step streamwise coordinate xvi y = normal coordinate z = coordinate perpendicular to the xy-plane Greek symbols or = f'mite-difference operator for the x-direction [3 - finite-difference operator for the y-direction y weighting factor for the stream function A incremental increase (in space or time) 8 tolerance error i; z-component vorticity v kinematic viscosity p density of the fluid 0' = weighting factor for the vorticity \l’ stream function or = frequency of oscillation of the boundary Vector symbols i unit vector in the x-direction i = unit vector in the y-direction k = unit vector in the z-direction V = velocity vector x position vector Subscripts B = nodes on the lower or upper boundaries CL = centerline nodes i = x-location index of an interior node j = y—location index of an interior node M = row of nodes on the upper boundary N = column of nodes on the downstream boundary Superscripts m = index of inner iterations for the stream function k = index of time step (") = relaxed value (') = weighted value ( )* = dimensional variable xvii INTRODUCTION Exact analytical solutions are available only for a limited number of fluid dynamics problems. In general, simplifying assumptions are applied to the describing equations of motion to facilitate a solution. Unfortunately, for most complicated flow problems such assumptions are not suitable. Consequently, these flows can be approached experimentally or by solving the equations of motion either numerically or by means of approximate analytical techniques. In complex flow problems it is necessary to determine the detailed velocity and pressure fields before an integral quantity of interest, such as an energy requirement or a wall shear force, can be calculated. From such calculations we can then predict how the integral quantity is influenced by the flow field. One fundamental way to achieve this task is to obtain a numerical solution to the mathematical model of the physical problem at hand. Advancements in high speed digital computing have resulted in a widespread interest in numerical solutions of the Navier-Stokes equations. Harlow and From (1965) remarked that computer calculations can give rapid and inexpensive, as well as accurate, solutions and that it is possible to perform sophisticated experiments on 1 2 the computer alone. Recent developments and advances in computational fluid dynamics were reviewed by Krause (1985). A comprehensive insight into the interdisciplinary nature of computational fluid dynamics and its application to carry out numerical experimentation was also explored by Aref (1986). The finite-difference approximation is one among a number of numerical methods that have been developed and studied extensively. In principle, furite-difference solutions can be obtained for flows having arbitrary Reynolds numbers. At high Reynolds numbers, however, convergence of the numerical solution depends on the numerical method, spatial grid resolution, and size of the time step. Difficulties in attaining convergence is usually associated with the loss of diagonal dominance of the system of difference equations approximating the flow field. The accuracy of the numerical solution is also dependent on the grid size and the error associated with the numerical approximation and the computation. For the majority of practical flows involving relatively high Reynolds numbers, the flow field must be represented adequately in order for a numerical solution to be sufficiently accurate. Consequently, the number of nodal points may become prohibitively large from the practical standpoint of computer time and storage. In addition, the lack of uniqueness in the numerical solution, due to the nonlinearities of the Navier-Stokes equations or inappropriate numerical representation of the boundary conditions, may prevent meaningful results. The present study is motivated by the potential importance of specifying solutions to flow problems comprising moving boundaries as well as investigating finite-difference methods effective in analyzing hi gh-Reynolds-number laminar flows. 3 Flow fields containing a moving boundary have many important applications in unsteady fluid mechanics. These applications include oscillating airfoils, helicopter rotors, impellers, compressor blades, baffle configurations, vessel geometries, washing machine agitators, agitated tanks, and industrial mixers. The flow patterns in such flows are usually complicated and affected by the fluid, dynamics of the flow, flow field geometry, and motion and shape of the moving boundary. Most importantly, moving boundaries can produce flow separation. The phenomenon of separation is one of the important features of fluid flow. It is an old subject in the literature of fluid mechanics. Using flow visualization techniques, Leonardo da Vinci observed and sketched recirculating eddies formed in the flow over objects with various configurations. Still, a thorough understanding of the subject of flow separation has not been obtained despite this long history. Laminar flow separation and reattachment of various kinds occur frequently in complex flows. One specific example is the laminar separated regions occurring in regimes of low-Reynolds-number flow around airfoils. Another example is the separated bubble generated by the flow near the leading edge of a thin airfoil. Unpredictable rapid growth of these separated regions can cause stall. The problem of flow separation is rather general, however, but it is the separation problem due to a moving boundary that is being addressed here. This will be illustrated by a steady and an unsteady motion of the boundary. The effect of boundary motion on a two-dimensional, viscous, incompressible, confined flow is considered in this study. The problem is modeled as a time- dependent, isothermal, laminar flow of a Newtonian fluid past a moving boundary 4 section placed within the entry region of a semi-infinite, straight, rectangular channel. The channel has a uniform velocity at the inlet and fully-developed Poiseuille flow far downstream. Both steady and oscillatory boundary motions are investigated. Our main objective is to numerically simulate the fluid flow when a moving boundary exists thereby developing a numerical technique suitable to treat problems that contain time-dependent boundary conditions. The solution can be obtained numerically by considering the continuity and the Navier-Stokes equations along with the appropriate boundary conditions. The full Navier—Stokes equations do not invoke the boundary-layer assumptions and, therefore, represent an exact solution in the sense that no terms in the x- and y- momentum equations are neglected. The vorticity-stream function formulation is selected for the treatment of the problem. The finite-difference equations are obtained from the partial differential equations of motion using second-order discretizations. After expressing the initial and boundary conditions in terms of the stream function and the vorticity in finite-difference form, a solution to the system of equations can be pursued to yield the vorticity and the stream function. From this information the velocity field, which is of primary interest, can be calculated. The Alternating-Direction Implicit (ADI) method is used to solve the vorticity equation while the iterative Successive Over-Relaxation (SOR) method is used to solve the stream function equation. Finite-difference approximations used when modeling physical problems often lead to the solution of large-order sparse linear systems. For large Reynolds numbers, however, it is essential to ensure the stability of the numerical solution by maintaining the diagonal dominance of the system of equations. An upwind differencing applied to the convective terms in the vorticity 5 equation helps avoid numerical instability. Second-order accuracy is then recovered at convergence via a correction technique. The power of this ADI/SOR computational scheme lies in the structure of the difference equations, which for all Reynolds numbers, yields a diagonally dominant system of linear algebraic equations. The present numerical solution is second-order accurate in its entirety since both the difference equations and boundary conditions have second-order accuracy. The use of accelerating parameters can further improve the convergence characteristics of the numerical procedure. Although somewhat artificially inserted into the difference equations, such parameters expand the domain of convergence and improve the computational speed as well. The general notion of accelerating the solution is well known. Nevertheless, a systematic determination of the optimum values of the accelerating parameters suitable for the nonlinear set of simultaneous equations resulting from the discretization of the Navier-Stokes equations has not been established. Therefore, a search must be made for the optimum values of the accelerating parameters in order to minimize the computing time and identify the domain of convergence. The optimum values of these parameters depend on the shape of the computational domain, mesh size, and the Reynolds number. They are determined primarily by numerical experimentation. The development of the steady flow in the entry region of a two-dimensional channel is first investigated using the full Navier-Stokes equations. Conduit flows are important to many engineering and biological applications, such as piping, heat exchanger design, short passages leading to nozzles and diffusers, and blood flow in arterial systems. Furthermore, the entry flow problem provides a prelude for developing laminar flow dreary and serves as an excellent example for testing 6 numerical schemes for solving the equations of motion. The streamwise velocity profile undergoes a change from an assumed inlet profile to that of a fully—developed, parabolic profile at a location downstream. The two regions making up the entrance flow are identified; the inviscid-core region, in which a viscous layer is assumed to exist near the wall, is followed by the profile-development region, in which viscous effects influence the entire channel flow. The results confirm two primary conclusions: the overshoot in the streamwise velocity profile near the wall is found to be minimal, and the inviscid-core region is approximately one-fifth of the entrance length with the profile-development region accounting for the rest. A second problem of entry flow in a channel constricted with a step is also discussed. When a finite step is fixed to the lower boundary, distinct separated regions with pronounced eddies of recirculating fluid are developed on both sides. A third class of problems involves an extension of the first; the channel flow is investigated when part of the lower boundary is allowed to move in its own plane, with either a constant or an oscillating velocity. Interest is focused on the problem of boundary motion that leads to the formation of a separated region during the upstream motion of a section of the lower boundary; the effect of the velocity, length, and location of the moving boundary on the size of the separated region is of particular interest. Prediction of such separated flows was previously done by patching the numerical solution in the vicinity of the separated region to the reminder of the flow field. Advancements in computing power make it possible to obtain a numerical solution for the entire flow field, thus avoiding difficulties associated with patching the 7 local numerical solution to the reminder of the flow. Moreover, a complete numerical solution is attractive for flows in which the boundary—layer theory is questionable for the range of Reynolds numbers considered, as is the case with separated flows, where the boundary-layer equations do not properly describe the separated portion of the flow. The present scheme provides a numerical solution over the entire domain utilizing all terms in the Navier-Stokes equations. It is the purpose of this study to also use different approximations in expressing the wall boundary conditions on the vorticity, examine several techniques for evaluating the coefficients of the convective terms in the vorticity equation, investigate the influence of several terms in the describing equations of motion, and study the effect of the mesh size and Reynolds number on the numerical solution. Numerical solutions are obtained for Reynolds numbers ranging fiom 10 to 5000. Furthermore, the over-relaxation and weighting factors for the present computational domain are determined. Their use in controlling the rate of convergence has significantly improved the numerical stability. The computer time is also reduced by up to a factor of three using the optimum values of these accelerating parameters. Reviews of work pertinent to this study are highlighted in Chapter 2. The present flow problems and their mathematical formulation are outlined in Chapter 3. Chapter 4 contains the details of the numerical methods used while Chapter 5 illustrates the mechanics of the computational scheme. Important considerations in the solution of the Navier-Stokes equations are discussed in Chapter 6. Finally, the results, along with a discussion and conclusions, are presented in Chapter 7. BACKGROUND In this chapter, it is our intention to highlight works relevant to the flow problem considered in this numerical study. First, a historical perspective into the laminar entry flows in pipes and channels is given. The flow in the entry region of pipes and channels is a classic problem in fluid mechanics and one that has important applications as well. In general, entry flows are studied by analytical methods, which include integral methods, series expansions, and linearization techniques; numerical methods; and experimental methods. Most of the analytical and numerical analyses involve some form of boundary-layer approximation in which a boundary layer exists near the walls together with an inviscid flow at the core near the inlet section. Bibliographies of laminar entry flows in pipes and channels, classified according to the method of solution, are given by Schmidt and Zeldin (1969), Lew and Fung (1970), Van Dyke (1970), and Fargie and Martin (1971). Secondly, a variety of numerical studies, in finite-difference form, pertinent to the flow problems of this study are presented. In particular, numerical considerations of internal and external separated flows, in addition to time-dependent flows of similar nature to the present investigation and some special numerical treatments to 9 flow problems, are discussed. Although numerical solutions to flow problems somewhat similar to parts of the present study, with some using correlated numerical techniques, are presented in the literature, no work could be identified with direct application to flows with moving boundaries or time-dependent boundary conditions. A brief survey of the specific numerical methods used in this study will be presented along with the numerical methods in Chapter 4. 2.1 Entry Flow in a Pipe Due to its resemblance in the physical setting and mathematical approach to the entry flow problem in a channel, the analogous problem in a circular pipe is included in this review. A variety of approximate analytical and numerical methods as well as experimental studies have been employed for the determination of the flow characteristics in the entry region of a pipe, as reported in the literature. Although not strictly considered a boundary-layer problem (Schliehting, 1979), this axisymmetric flow has been studied with the aid of methods similar to boundary-layer solutions. An assumed infinitesimally thin boundary layer on the wall of a pipe at the inlet section, grows in thickness as the flow passes downstream until it becomes equal to the radius of the pipe. The fluid at the core is practically uninfluenced by the viscosity. The fully-developed parabolic velocity distribution is approached asymptotically at a distance well beyond when the boundary layer thickness becomes equal to the pipe radius. At a large distance downstream from the inlet, the velocity distribution approaches the theoretical parabolic Poiseuille flow over the cross- section of the pipe. 10 Prandtl and Tietjens (1934) presented a summary of early analyses of the flow in pipes and channels. Independently, Hagen and Poiseuille (Prandtl and Tietjens, 1934) were the first to experimentally investigate the flow in straight tubes and include the influence of viscosity. Primarily, they were concerned with the application of the equations of hydrodynamics to estimate the pressure drop. Their findings, however, apply only to flow at a sufficient distance from the inlet section. On the other hand, using a perturbation of the parabolic distribution, Boussinesq (1891) was the first to make theoretical studies of the phenomenon, but his calculations for the velocity distribution did not agree with the experimental results. In the entry region, it is clear that no parabolic distribution exists. Prandtl (Prandtl and Tietjens, 1934) suggested a velocity profile approximated by two parabolic curves near the walls, with zero velocities exactly at the walls, joined by a flat middle part near the pipe axis. At the inlet section, the width of the parabolic curves is zero. It then increased as the flow advances downstream until the two curves are united into a single parabola. The flow in the uniform velocity core was assumed to satisfy Bemoulli’s equation. Schiller (1922) was the first to use an approximate method based on an equilibrium condition between momentum, pressure drop, and viscous drag to form an equation similar to the momentum integral equation. Schiller dropped the viscous terms and employed Bernoulli’s equation in his analysis to substitute for the pressure term in his integration of the Navier-Stokes equations. He designated the entrance length to be the streamwise location where the two parabolic curves are combined. Carrrpbell and Slattery (1963) corrected the original derivation of Schiller (1922) to account for the loss of energy due to viscous dissipation within the boundary layer. ll Employing both integral and differential momentum equations, Fargie and Martin (1971) simplified the treatment of the previous authors. They presented a thorough review of previous studies and compared their results as well. Lately, Mohanty and Astlrana (197 8) investigated the laminar flow in the entry region of a pipe using an integral method. They integrated the boundary-layer equations in the region closer to the inlet section according to the von KArmén- Pohlhausen scheme. Using fourth-degree polynomial approximations to the velocity profiles, they solved the Navier-Stokes equations with an "order-of-magnitude" analysis in the region further away from the inlet. They also showed that the laminar viscous layers meet at the pipe axis, when the inviscid core terminates, at a distance of approximately one-quarter of the entrance length. The previous literature did not indicate the location of such a meeting of the viscous layers in an entry flow, although the profiles of Schlichting (1979) appear to give such an impression. The unpublished work of Atkinson and Goldstein for the flow near the inlet of a pipe, presented by Goldstein (1965), used a variation of the method due to Schlichting (1934) for the corresponding problem in a two-dimensional channel. From the equations of motion in cylindrical coordinates, a solution is obtained for large Reynolds number by generalizing Blasius’ solution of Prandtl’s two-dimensional boundary-layer equation. A numerical integration of the equation resulting from applying a power series expansion yields the velocity distribution. Results agreed with the calculations of Schiller (1922) except near the wall. Away from the inlet, Atkinson and Goldstein continued the solution by calculations similar to those of Boussinesq (1891). The combination of these two solutions produced more accurate results than those of Boussinesq (1891) and Schiller (1922). 12 Langhaar (1942) analyzed the flow in the why region of a straight tube by means of a linearizing approximation to the Navier-Stokes equations. He defined the velocity profiles by Bessel functions and obtained the entrance length. He also determined the pressure drop in the entry region by applying the energy equation and the computed velocity field. Throughout his work, the inlet velocity was constant, the pressure was assumed to be a function of the streamwise coordinate only, and the second derivative with respect to the streamwise coordinate was neglected. The laminar flow development in the entry region of ducts of any cross section, including circular tubes and rectangular ducts, were also analyzed by Sparrow et al. (1964). Their analytical solution, which is modified from the method due to Langhaar (1942), involved a stretched axial coordinate and a linearization of the inertia terms of the equation of motion; hence, the boundary layer model does not need to be invoked. A correction term was used to compensate for the accumulated wall shear stresses exerted by the developing flow relative to the fully-developed flow. Experimental velocity measurements in the entry region of tubes with well- rounded inlets were first made by J. Nikuradse and published by Prandtl and Tietjens (1934). Results showed that the velocity distribution around the core close to the inlet section is independent of friction and that the parabolic reduction in velocity toward the walls is justified. After the boundary layers meet at the axis, the flat velocity at the core ceases to exist. Nikuradse’s tests checked well with the theoretical values obtained by Boussinesq (1891) in the region away from the inlet section and were in good agreement with the calculations of Schiller (1922) only in the first third of the entrance length. The complete transition to fully-developed flow according to Schiller (1922) seemed to be rather premature. This discrepancy is 13 apparently due to the acceleration of the fluid near the axis which causes the pressure drop in the inviscid core to increase compared to that of a fully-developed flow. At larger distances from the inlet section, Schiller’s calculations predicted a high rate of increase of centerline velocity. As a result, his entrance length was rather short. The numerical solution for laminar entry flow in a pipe by Hornbeck (1964) used essentially the same procedure as Bodoia and Osterle (1961), which is discussed in Section 2.2. Christiansen and Lemmon (1965) numerically analyzed the uniform entry flow in a pipe on the basis of linearized inertia terms without postulating a boundary- layer model. They also compared their results with previous theoretical and experimental results. The work of Vrentas et al. (1966) used the boundary-layer equations and the method of Wang and Longwell (1964), as outlined in Section 2.2. Laminar flow in the entry region of a pipe at low and moderate Reynolds numbers (in the range 0 to 500) was also studied numerically by Friedmann et al. (1968) without the boundary-layer approximations. They observed the overshoot phenomenon. To determine whether the bulges in the velocity profiles might not merely be a consequence of singularity at the leading edge, they treated the limiting case of vanishingly small Reynolds number analytically and modified the axial velocity profile near the wall to eliminate the singularity. Their results showed that the bulges persisted in the region near the wall. The entry flow into blood vessels was modeled by Lew and Fung (1970) as steady axisymmetric flow in circular tubes. A numerical solution, without the boundary layer approximation, was provided using a series of two sets of eigenfunctions for Reynolds numbers between 0 and 100. 14 2.2 Entry Flow in a Channel The laminar, incompressible flow in the entry region of a straight channel is an example of two-dimensional boundary-layer flow. Owing to viscous friction, boundary layers are formed on both walls, and their thickness increases as the flow propagates downstream. The development of the flow is primarily concerned with changes in the streamwise velocity component over the entrance length. Due to the influence of viscosity, the velocity profile gradually changes to assume the Poiseuille parabolic distribution at a sufficiently long downstream distance. Since the pioneering work of Prandtl, boundary-layer theory has been the centerpiece to all analytical solutions of the entry flow. Although considered as a powerful tool, it is well known that the boundary-layer assumptions are invalid in the vicinity of the leading edge of a flat plate as is the case near the inlet section to a channel. In this region the derivative azu/ax2 is not negligible relative to aZu/ayZ. In addition, the y-component velocity is not negligible and the pressure gradient in the y-direction is not necessarily small so that the y-component momentum equation does not vanish. In his early work, Schlichting (1934, 1979) analyzed the flow in the entry region of a two-dimensional channel using the boundary-layer model. The entry region was considered to be composed of two zones: first, a section of two boundary layers developing on the walls with an inviscid core near the centerline followed by a section where the velocity profile gradually develops into the parabolic distribution of Poiseuille flow. At small distances from the inlet section, he assumed that the two boundary layers on the lower and upper walls will grow in the same way as they 15 would along a flat plate at zero incidence. The resulting velocity consists of two boundary-layer profiles near the walls joined by a flat profile at the core. Since the flow rate at each and every cross section must be the same, the reduction in the flow rate near the walls due to friction must be compensated for by a corresponding increase near the centerline. Mathematically, Schlichting used the continuity equation and the definition of the displacement thickness to define the velocity outside the boundary layer as a function of the streamwise coordinate. Following a procedure similar to Blasius’ solution for laminar boundary-layer flow over a flat plate, he expanded the stream function in a series and integrated in the downstream direction to find the boundary-layer growth and the velocity distribution. The boundary-layer growth in this case is distinguished from Blasius’ solution by the effect of an accelerated external flow. The boundary-layer solution is then matched with the uniform flow at the core. In the region after the two boundary layers meet, where the flow is nearly fully- developed, Schlichting perturbed the Poiseuille velocity profile by performing an asymptotic power series expansion leading to progressive deviation from the parabolic distribution as the integration proceeds in the upstream direction. The two solutions, expressed in series form with a sufficient number of terms, are patched at some appropriate streamwise location where they are simultaneously valid. Thus, the flow in the entire entry region is determined. The entrance length was calculated by Schlichting as L,/H = 0.04 Re, where H is the height of the channel and Re is the Reynolds number. Roidt and Cess (1962) improved Schlichting’s (1934) solution by the inclusion of additional terms in the perturbation series. Van Dyke (1970) introduced an 16 asymptotic solution that is uniformly valid for large Reynolds numbers. He corrected Schlichting’s technique by adding second-order terms for the stream function to account for the displacement thickness in the boundary-layer region near the inlet; due to the displacement effect of the boundary layer, the inviscid core near the inlet does not resemble a uniform flow. The first approximation of Van Dyke’s expansion was the leading edge solution for a semi-infinite plate, which was calculated earlier by Davis (1967). Van Dyke also pointed out that the vorticity at the inlet is not zero because of its upstream diffusion. A similar procedure was used by Wilson (1971); Wilson (1969) considered the transition in the profile-development region to a fully- developed flow in more detail. In their works, Collins and Schowalter (1962, 1963) showed that the boundary- layer' solution approaches the finite-difference solution (Bodoia and Osterle, 1961) by considering more terms than had previously been retained by Schlichting (1934). Abarbanel et al. (1970) used a similarity solution to check whether the overshoot in the velocity profile of channel flow is part of the exact solution and not a numerical consequence of the truncation error. They investigated the Stokes’ flow over a half- plane with flat inlet velocity profile. Their velocity profiles exhibited the same sort of bulges as had been found in the numerical solutions. Moreover, the flow pattern in the vicinity of the leading edge agreed with that of channel flow. Recently, Li and Ludford (1980) examined the entry flow analytically to determine whether the overshoot in the velocity profile near the channel walls, found numerically, is real; both uniform and special irrotational inlet conditions were considered. They used a second-order boundary-layer analysis, up to a Reynolds number of 1000, and concluded that the axial velocity must have an overshoot at all 17 Reynolds numbers. This overshoot was found to lie inside the boundary layer for the uniform inlet condition and at the edge for the special irrotational inlet condition. Wiginton and Wendt (1969) simplified the transformation used in the linearization solution of Sparrow et al. (1964) and extended their techniques to solve the eigenvalue problem for two-dimensional flows. Also, Wiginton and Dalton (1970) modified the technique used by Sparrow et al. (1964) to analyze the laminar flow in the entry region of a rectangular duct with different aspect ratios; they noted that the pressure gradient reached a fully-developed state in about half the distance of the velocity profile. An analytical method for determining the pressure drop due to flow development in the entry region of ducts of arbitrary cross sections, without solving for the velocity field, were formulated by Lundgren et al. (1964). They used the momentum and energy equations with correction factors to account for the change in momentum between the inlet and downstream sections. An experimental investigation of the development of the velocity and pressure fields for a laminar flow in the entry region of a rectangular duct was carried out by Sparrow et al. (1967). Reynolds numbers were in the range 1000 to 5000. An interesting phenomenon was the higher velocity seen in the lower half of the duct as compared to the upper half. Comparison of their data with the results of other authors yielded a close resemblance. Feliss et al. (1977) also examined the incompressible channel flow experimentally and found that the entrance length, as a linear function of the Reynolds number, correlated well with the analytical predictions of Schlichting (1934). l8 Numerically, Bodoia (1959) and Bodoia and Osterle (1961) were the first to obtain a solution for the development of plane Poiseuille and Couette flows of an incompressible fluid in a straight channel. They used a first-order finite-difference approximation to Prandtl’s boundary-layer equation. Their solution was for Reynolds number of 75. Wang and Longwell (1964) studied the steady-state laminar flow in the entry region of a channel at 11 Reynolds number of 150 with both uniform and cascade inlet conditions. They employed Southwell’s relaxation method (Allen and Southwell, 1955) to solve the complete Navier-Stokes equations expressed in terms of the vorticity and the stream function, and used a first-order backward-difference approximation to stabilize the numerical solution. Their vorticity boundary condition, borrowed from Thom (1933), was first-order as well. The results were in better agreement with Schlichting (1934) in the region downstream of the inlet. The overshoots in the velocity profile near the inlet section were nonexistent for the cascade inlet condition. A solution for Reynolds number of 250 was obtained by Gillis and Brandt (1964) and echoed in their subsequent work (Brandt and Gillis, 1966) in which they extended the method to the equations of magrretohydrodynamics, specifically to the problem of the development of Hartrnann flow. A fourth-order equation in the stream function was used for the linearized Navier—Stokes equation. As in the case of Wang and Longwell (1964), the uniform inlet flow develops profiles with local minima on the channel axis and a pair of symmetrically placed maxima on either side of it. These kinked profiles persist for a distance downstream of the order of magnitude of the channel height. l9 McDonald et al. (1972) solved the Navier-Stokes equations in the entry region of both a tube and a channel for Reynolds numbers between 0.75 and 37 500. They developed a stable numerical scheme based on the Alternating-Direction Implicit method m solve the vorticity and stream function equations; their representation of the vorticity boundary condition was second-order accurate. An overshoot in the velocity profile existed for the uniform inlet case but was less pronounced for a zero- vorticity inlet condition. Using quasi-linearized Navier-Stokes equations in terms of the primitive variables, Morihara and Cheng (1973) solved the entry flow between parallel plates; their quasi-linearization involved a truncated Taylor series expansion of the first-order. A numerical solution for Reynolds numbers in the range 0 to 1000 was provided. The effect of the upstream condition on the laminar flow development in a channel for Reynolds numbers between 1 and 1000 was studied numerically by Sparrow and Anderson (1977). Their results showed that, due to the upstream interaction at low and intermediate Reynolds numbers, the sharply turning fluid near the walls tends to overshoot at first then readjust as it progresses downstream. In all the numerical solutions mentioned above, either some simplifying assumptions were made, such as the assumptions inherent in boundary-layer theory (that is, both the streamwise derivative azu/ax2 and the pressure gradient normal to the walls, ap/By, were neglected) and in the linearization of the Navier-Stokes equations, or a first-order vorticity boundary condition was used, or the numerical solution of the full Navier-Stokes equations was presented for relatively low Reynolds numbers. The recent numerical work by Abdulla (1987) used a successive-over relaxation procedure including accelerating parameters to solve the N avier—Stokes equations for 20 the may flow for Reynolds numbers up to 2000. Although he used a second-order vorticity boundary condition, his difference equations were only first-order accurate; he used forward] backward difl‘erencing to improve the convergence of the Navier- Stokes equations expressed in terms of the vorticity and the stream function. A study of transient magnetohydrodynamic duct flow was conducted by Lundgren et al. (1961). They considered the parallel flow of an electrically conducting viscous, incompressible fluid with transverse magnetic field and determined the exact solutions for the velocity and magnetic fields in the form of convolution integrals. Using a finite-difference solution, Hwang and Fan (1963) also investigated the development of velocity profiles in the entry region of a flat rectangular duct for an electrically conducting fluid and transversely applied magnetic field The subject matter of the next few sections is concerned with further literature review of work relevant to this study. A wealth of analytical, numerical, and experimental studies, for steady and time-dependent flows, are cited in the literature. A comprehensive survey is bound to be cumbersome and, therefore, will not be attempted. However, a number of investigations (predominantly numerical), with pertinence to the flow problem and numerical procedure of this study, are reviewed. The studies discussed hereafter are conveniently classified by flow type. 2.3 Internally Separated Flows Internal flows mainly containing separation and wakes are of interest to this study. These include flows with sudden contractions or expansions, flows in 21 constricted channels and pipes, flows in conduits with obstructions, flows in driven cavities, and flows past blunt or streamlined bodies placed in confined flow fields. Greenspan (1969a) studied the steady flow in a channel with both a forward- facing and a finite step for Reynolds numbers up to 10 000; the approaching flow was considered to be parabolic. He used the Successive Over-Relaxation method with accelerating parameters and first-order boundary vorticity to solve the vorticity and stream function equations. An approach consisting of a numerical procedure similar to that of Greenspan (1969a) for low Reynolds numbers and modified with upwind differencing for moderate and high Reynolds numbers, was used by Friedman (1972) to study the laminar flow in a channel with a finite step. Optimum accelerating parameters were also used to reduce the computer time needed for convergence. Taylor and Ndefo (1970) used the method of splitting, which was originally developed by Yanenko (1971), to compute the flow in a two—dimensional channel with straight walls and a backward-facing step for Reynolds numbers up to 100. Abdulla (1987) applied the same over-relaxation procedure that he used for entry flow (which is also similar to that of Greenspan (1969a)) to predict the laminar flow separation in a channel constricted with a forward-facing step, a backward-facing step, and a finite step. An experimental and numerical investigation of a two-dimensional channel flow over a single backward-facing step was performed by Armaly et al. (1983). They reported the downstream velocity distribution and reattachment length for Reynolds numbers between 70 and 8000. Laser-Doppler measurements revealed the presence of separation regions downstream of the step and on both sides of the channel, in 22 addition to the primary recirculating zone. The numerical results obtained showed good agreement with the experiment. Results for the time-dependent impulsively starting backward-facing step mounted in a two-dimensional duct were presented by Durst and Pereira (1988). They used a finite volume numerical approach for Reynolds numbers 10, 389, and 648. Experimentally, they showed that no recirculating flow region existed on the wall opposite to the step for Reynolds number S 420. A numerical solution of the N avier—Stokes equations for steady, laminar flow in a channel constricted with a semi-infinite forward-facing symmetric step on both walls, was provided by Dennis and Smith (1980). They studied separation for Reynolds number up to 2000 using different size mesh. Comparisons of the upstream separation position and the wall vorticity showed good agreement with the asymptotic solution of Smith (1979). The effect of an asymmetric downstream disttn'bance, in the form of a constriction, branching, bend, or flow injection, in a channel was also studied by Smith (1977). The boundary-layer equations were solved numerically. The flow becomes separated on the wall ahead of the disturbance and the finite disturbance prohibited a smooth joining with the original Poiseuille flow. Two different types of internally separated flows were studied by Kumar and Yajnik (1980). The flows in a channel with a symmetric sudden expansion and in a channel "with a base" at large Reynolds numbers were solved by means of a small perturbation in the eigenfunctions of the Poiseuille flow; this reduces the problem to solving nonlinear first-order ordinary differential equations. Results for the problem with the sudden expansion compared well with the finite-difference solution obtained from an iterative procedure. The separated flow due to an obstruction placed in a two- dimensional channel was explored by Nallasamy (1986). The characteristics of the 23 separated flow behind the obstruction were obtained from the finite-difference solution of the N avier-Stokes equations for Reynolds numbers up to 1500. A solution of the N avier-Stokes equations of a confined wake problem was obtained numerically by Paris and Whitaker (1965). The two—dimensional problem was formed by the merging of two Poiseuille flow streams. The Altemating-Direction Implicit method was used in the solution of the vorticity and stream function equations for Reynolds numbers as high as 647. At high Reynolds numbers, the boundary layers equations, without an order of magnitude analysis (i.e., the y- momentum equation was not neglected), were used. The results were in agreement with the solution of the complete equations. The flow in a circular pipe with recessed walls was numerically investigated by Friedman (1970). Separation and vortices were studied for Reynolds numbers up to 500. Also, the steady, laminar, axisymmetric flow through vascular stenoses was modeled in the study of Deshpande et al. (1976) as flow in a rigid tube with a contoured constriction. They used the full N avier-Stokes equations in cylindrical coordinates. For the Reynolds numbers at which the flow is known to be laminar, their numerical results converged without difficulty and compared well with experimental results. 2.4 Flow in a Driven Cavity Because of the nonlinearity of the Navier-Stokes equations governing the flow in a driven cavity, and because of the presence of regions in which the solution changes rapidly, a large number of authors used the driven cavity problem to analyze 24 their numerical techniques. The importance of the problem to fluid mechanists, however, stems from its relevance to separated flows. One of the earliest attempts at solving this problem was made by Kawaguti (1961) who studied the flow in a rectangular cavity with different length/depth ratios. He integrated the Navier-Stokes equations numerically and found out that the relaxation procedure failed to converge at high Reynolds numbers. The Navier- Stokes equations were solved iteratively by Mills (1965) for the flow generated in a driven rectangular cavity at a Reynolds number of 100. Mills used the numerical procedure advocated by Thom (Thom and Apelt, 1961). Burggraf (1966) assessed the effects of nonlinearities and of noncircular boundaries by considering numerical solutions of the full N avier-Stokes equations for a square cavity. First, his analysis, based on a linearized theory restricted to an eddy having a circular boundary, revealed the development of a recirculating eddy. He then confirmed the validity of the linear analysis as a description of separated eddies by the results of the numerical solution. These solutions were carried out by means of a relaxation procedure for Reynolds numbers in the range 0 to 400; the limiting cases, as the Reynolds number goes to zero or infinity, were solved analytically. At higher Reynolds numbers an inviscid core develops, but secondary eddies are present in the bottom corners of the square at all Reynolds numbers. Results obtained numerically as well as experimentally by Pan and Acrivos (1967) for a rectangular cavity, confumed the presence of a single inviscid core of uniform vorticity with viscous effects being confined to thin shear layers near the boundaries. Using a successive over-relaxation method with accelerating parameters, Greenspan (1967, 1968, 25 1969b) also studied the flow in a driven cavity. His numerical solutions were for Reynolds numbers between 50 and 100 000. Numerical solutions of the elliptic transport equations of a laminar flow with heat transfer for a two—dimensional driven cavity problem were presented by Runchal et al. (1969). Their solution featured unsymmetrical formulation of convection terms, Gauss-Seidel iteration procedure, and a grid with a nonuniform mesh. The influence of mesh size and the vorticity boundary condition on the accuracy and convergence of the solution were considered for a wide range of Reynolds numbers. Their computation results agreed with previous knowledge and showed that secondary eddies appear in the lower corners of the cavity, even for creeping flows. Bozeman and Dalton (1973) obtained solutions for the flow in rectangular cavities by solving various implicit approximations of the Navier-Stokes equations. They determined the effect of the Reynolds number and mesh size on the different schemes and also examined the effect of vorticity boundary conditions; the behavior of a vortex center and secondary vorticity were exhibited as well. Several numerical techniques for the treatment of the viscous, incompressible flow in a driven cavity were reported by NASA (Rubin and Harris, 1975). Their results showed that convergence is more rapid and accurate with the vorticity-stream function formulation compared to the solution of the equations of motion in terms of the primitive variables. The authors also attributed the higher-Reynolds-number solutions to the stability of the Alternating-Direction Implicit method. In particular, Morris (1975) used the Alternating-Direction Implicit method with a first—order vorticity boundary condition to solve the vorticity equation. 26 The two—dimensional driven cavity flow remains as a model problem in the study of different numerical schemes. Roache (1975) used the driven cavity problem to compare three different numerical methods and also to examine the effect of boundary conditions on his solution. This was also the case with Rubin and Khosla (1977) in their development of a higher-order approximation for viscous flow calculations. A review of many numerical methods, complete with error and convergence comparisons, for computing recirculating flows in the cavity problem, was presented by Tuann and Olson (1978). Benjamin and Denny (1979) discussed convergence properties of several finite-difference schemes for solving the two- dimensional flows in a cavity at Reynolds numbers as high as 10000. They used a relaxed Alternating-Direction Implicit procedure in their second-order accurate solution of the vorticity and stream function equations. Ghia et al. (1982) used the cavity flow problem to test the effectiveness of their numerical method; they used a fine mesh in hi gh-Reynolds-number flows. In addition, the method of false transient was used by Mei and Plotkin (1986) in a finite-difference scheme applied to the driven flow in a square cavity for Reynolds numbers up to 5000. Lastly, Duck (1982) studied the flow inside a square cavity generated by an oscillating motion of the upper boundary. He solved the unsteady Navier—Stokes equations for Reynolds numbers as large as 600. His results showed that the steady component of the flow comprises two pairs of equal counter-rotating eddies situated at the upper and lower halves of the cavity. 27 2.5 Externally Separated Flows External flows such as the flow over a flat plate and past cylindrical, spherical, and other different shape objects are associated with separation. Two of these studies for flows with boundary-layer separation are of particular interest. heal and Acrivos (1969) obtained a numerical solution for the Navier-Stokes equations for the steady flow over a flat plate in which a constant finite slip velocity in the direction opposite to the stream is imposed. Their study was concerned with the structure of the closed streamline flows within a boundary layer. These streamlines were found to be embedded in the viscous layer adjacent to the plate at all Reynolds numbers; Reynolds numbers were up to 80. The parabolic boundary-layer equations were found suitable for the solution of the flow including the recirculating zone near the plate. Briley (1970, 1971) analyzed a two—dimensional, viscous, incompressible flow containing separation bubbles by patching the numerical solution in the immediate vicinity of the bubble to the boundary-layer solution and the inviscid solution of the outer flow. He presented numerical solutions for flows with Reynolds numbers of the order of 50 (based on momentum thickness). Briley used a fourth-order Lagrangian polynomial for interpolating the stream function to derive a formula for the vorticity at the boundary based on a special relationship for the x-component velocity at the points adjacent to the boundary, thus providing higher order coupling between the vorticity and the stream function at the wall. The Altemating—Direction Implicit method, with an iterative procedure and variable time steps, was used to solve both 28 the vorticity and stream function equations; he preferred the marching approach in solving his steady-flow problem. 2.6 Time-Dependent Flows A general treatment of the time-dependent, two-dimensional flow of viscous, incompressible fluids can be found in many publications. Several time-dependent flow solutions are outlined below. A numerical method for time-dependent problems concerning the flow of viscous, incompressible fluids was applied by From and Harlow (1963) to study the development of a vortex street behind an impulsively accelerated plate placed in a channel. The Reynolds number range was between 15 and 6000. The numerical scheme employed was discussed in more detail by From (1963, 1964). It is an explicit scheme with an iterative procedure based on Liebmann’s method, as outlined by Thom and Apelt (1961); a special fust-order vorticity boundary condition was derived. Fromm also analyzed the stability of the vorticity equation. In his later work, From (1972) described his method for the numerical computation of the Navier-Stokes equations at high Reynolds numbers and its application to several time-dependent, viscous incompressible flow problems. He also presented alternative discretization schemes for the convective derivatives. A general finite—difference computational procedure for time-dependent viscous, incompressible flow problems was described by Pearson (1964, 1965a, 1965b). It involves solving a nonlinear, fourth-order equation in three independent variables and the use of a second-order approximation for the boundary vorticity. The method was 29 applied to examples such as the injection of fluid into a rectangular region, the time- dependent rotating disk problem, and vortex formation and break up in a two- dimensional region. He also discussed the choice of difference equations, relaxation procedures, approximation to the boundary conditions, and the resulting computational stability, speed, and accuracy. Pao and Daugherty (1969) studied the time-dependent, two-dimensional, viscous, incompressible flow past a finite flat plate. For the vorticity equation, they used both an explicit method and the Alternating-Direction Implicit method with forward] backward differencing, while for the stream function equation, they used the iterative Successive Over-Relaxation method. The vorticity boundary condition was approximated by the second-order expression due to Jenson (1959). Also, three types of coordinates were used to construct the computational grid. The details of the flow field around the plate were computed for Reynolds numbers equal to 4 and 993. It was found that the velocity at the outer edge of the boundary layer exceeds that of the free stream and that the Alternating-Direction Implicit method was two to three times faster than the explicit method. The accuracy of the different numerical methods and the effects of the far field boundary conditions and mesh size were also studied. A numerical procedure for a two-dimensional, compressible, unsteady boundary- layer flow was presented by Kwon et al. (1988). They studied the sudden acceleration of an incompressible fluid over a semi-infinite flat plate and the steady sinusoidal oscillation of a plate in a semi-infinite fluid, better known as Stokes’ second problem. Their results were in good agreement with the available theoretical solutions. 30 2.7 Numerical Treatments of Flow Problems A numerous number of numerical techniques have been applied to flow problems of various sorts. As a closure to this review, a number of these numerical treatments will be briefly highlighted; they are mostly studies related to the development of the present numerical scheme. In their integration of the Navier-Stokes equations for steady, two-dimensional flow, Dennis and Chang (1969) studied the convergence of the iterative procedure used to solve the steady flow past a finite plate and elliptic cylinders. The vorticity and stream function were used as dependent variables. They discussed the finite- difference approximation of the vorticity equation necessary to secure the diagonal dominance of the system while still retaining the higher-order accuracy of the central differences. Gosman et al. (1969) also analyzed the use of upwind differences to provide convergence to numerical solutions of flow problems. Several numerical methods for inviscid and viscous flows were compared by Taylor et al. ( 1972). For viscous flows, they concluded that implicit schemes are preferable to explicit or Dufort-Frankel differencing and that for the inertial terms, any scheme which is stable for inviscid flow and has controllable artificial viscosity at low Reynolds numbers, may be used. A more thorough review of methods used in the numerical simulation of viscous incompressible flows was given by Orszag and Israeli (1974). They discussed the choice and accuracy of available numerical approximations and cited an enormous number of references. They also studied the error associated by several formulas (i.e., those of Briley (1971), From (1963), Jenson (1959), Woods (1954), and Thom (1933)), used in evaluating the vorticity on 31 the no—slip boundaries and showed that results given by the different formulas can differ. Another extensive comparison between results given by the above various formulas in the case of a driven cavity flow has been made available by Gupta and Manohar (1979). To identify possible causes of instability, the effect of the finite- difference approximations to the no-slip conditions on the linear stability of the numerical solution of unsteady flows was analyzed by Taylor (1970). All conditions used were shown to be stable without the presence of suction and conditionally stable if there is suction; in this case, the choice of mesh size is restricted. The selection of the proper integration method was shown to be critical. Khosla and Rubin (1974) suggested a diagonally dominant second-order accurate implicit scheme to replace the first-order upwind differencing usually used to stabilize the solution of the nonlinear Navier-Stokes equations. Finally, the use of optimized over-relaxation factors was discussed by Thompson (1969). The number of iterations required for convergence for the time- dependent Navier-Stokes equations, with iterations performed at each time step, was reduced by more than one order of magnitude over Gauss-Seidel iteration. The restrictions placed on the maximum size of the time step were also removed. DESCRIPTION OF THE PROBLEMS The sequence of the basic flow problems and their mathematical modeling are discussed in this chapter. The describing equations of motion, the Navier-Stokes and the continuity equations, can be expressed in different forms. Here, the vorticity- stream function formulation is adapted for the solution. The initial and boundary conditions required in the solution are also presented. 3.1 The Flow Problems A number of flow problems are investigated in this study. The flow past a moving boundary section placed within the entry region of a rectangular conduit is of primary interest. The physical situation is illustrated by the two-dimensional, isothermal, incompressible, laminar flow of a Newtonian, isotropic, homogeneous fluid with constant density and kinematic viscosity in the entry region of a straight channel. The channel is made up of two semi-infinite, flat plates lying parallel to one another in the xz-plane; the x-axis of the rectangular Cartesian coordinate system coincides with the lower wall and the y-axis is normal to the flow. At the inlet where x = 0, a uniform flow with velocity U is assumed across the channel of height H. A 32 33 y t Uniform Profile Parabolic Profile Figure l. A schematic representation of the flow problem including the moving boundary. parallel Poiseuille flow is considered to exist far downstream. The defining geometry is shown in Fig. 1. The moving section of the boundary, which is assumed to be moving either steadily or oscillating periodically with a constant frequency and specified amplitude, has a velocity UB and is defined to exist between the streamwise locations x = XI and x = X2. The steady flow in the entry region of a plane two-dimensional channel is considered first. This entry flow gives rise to a growth of boundary layers on the two walls. The most detailed description to this problem has been that of Schlichting (1934, 1979). The fluid enters the channel with a uniform velocity across the inlet section and its velocity is assumed to be zero at the wall for all positive x. Due to the influence of friction, the fluid lying near the walls is retarded; the viscosity makes itself felt in a layer near the wall. The boundary layer grows in thickness as the flow 34 progresses downstream. The inner part of the flow is inviscid and its velocity is essentially uniform. Since the continuity of mass requires the flow rate at all cross- sections to be constant, the fluid near the centerline of the conduit is accelerated. At some streamwise distance from the inlet section, the two boundary layers merge together at the centerline (when the boundary-layer fills half the channel’s height). At this stage, the pressure drop per unit length is higher than that for the fully-developed flow. The flow will continue to accelerate until equilibrium between viscous forces and pressure forces is finally reached at a location downstream. As a result, the initially uniform velocity profile is transformed asymptotically into the parabolic velocity distribution of a Poiseuille flow. Since Bulax at 0 in the entry region, the flow is not one-dimensional, and the velocity components depend on the streamwise coordinate x as well as the normal coordinate y. The entrance length, , t Invlscid Core Parabolic Profile 4 L. :L: r.d __. l Inv'ncid-Core Length Profile-Development Length L _ Entrance Length Figure 2. Laminar entry flow in a two-dimensional channel. 35 which includes the inviscid-core length and the profile-development length, is considered to exist whenever the velocity at the centerline is less than 99% of the parabolic centerline velocity (Prandtl and Tietjens, 1934). The flow and its distinct regions are shown schematically in Fig. 2. The channel is then constricted by placing a finite step on its lower boundary. As a result, the flow becomes separated before and aft of the step. A moving section introduced on the lower boundary is also studied. When the boundary section moves uniformly in its own plane with a constant velocity in the direction opposite to the flow, a separated region exists in the near vicinity of the moving section. Motion of the boundary section in the direction of the flow, however, causes the neighboring fluid to accelerate without the presence of any separation; the problem of Couette- Poiseuille flow results when the entire lower boundary moves in the direction of the flow. The two cases of boundary motion with and against the direction of the flow also occur in the case of an oscillating boundary section where the flow becomes time- dependent. A schematic representation of the sequence of flow problems is shown in Fig. 3. 3.2 Mathematical Formulation Viscous, incompressible flows are described by the Navier-Stokes and the conservation of mass equations. To facilitate the solution, the differential equations of motion are expressed in non—dimensional form. The following characteristic quantities are selected for normalizing the various variables of this study: length = H, velocity = U, time = %. pressure = pUz (1) rig a) Entry flow in a two-dimensional channel. b) Entry flow with a step. c) Entry flow with a moving lower boundary section. <—> d) Entry flow with an oscillating lower boundary section. Figure 3. Schematic representation of the sequence of entry flow problems. 37 where H is the distance separating the lower and upper boundaries, U is the uniform inlet velocity, u) is the frequency of oscillation of the moving boundary section, and p is the density of the fluid. With the above characteristic quantities, the following dimensionless variables result: * it: t x = x , t = 0) t* ’ v = V , p = —L2— (2) H U pU where the boldface refer to vector quantities and the asterisk is used to denote a dimensional variable. In the above definitions, x is the position vector (xi + yj + zk ), t is the time, V is the velocity vector (ui + V} + wk), and p is the pressure; x, y, and z are the three spatial coordinates; u(x,y,t), v(x,y,t), and w(x,y,t) are the velocity components in the x-, y—, and z-directions respectively; and i, j, and k are unit vectors in the x—, y-, and z-directions, respectively. Hereafter, all variables and equations are expressed in dimensionless form unless otherwise stated. Applying the transformation of variables indicated above, the dimensionless Navier-Stokes equation in vector form is written as 3V V V V V 1 v2‘ at ( ) pk Re ( ) where V is the gradient operator and V2 is the Laplacian operator, both in non- dimensional form. The Strouhal number, St, is defined as St = —- <4) and the Reynolds number, Re, based on the channel’s height is given by 38 Re == —— (5) where v is the kinematic viscosity. In Eq. (3), since the body forces do not influence the velocity field of confined flows, the effect of body forces is ignored by including the body-force term in the pressure term to form the kinetic pressure, i.e., _ 2* (6) where p is the pressure, g is the gravitational acceleration, and z“ is the vertical coordinate; gz‘lU 2 is the dimensionless Froude number. Furthermore, the conservation of mass equation of an incompressible fluid is v.v = o . (7) It is evident that the flow is influenced by two dimensionless parameters, namely, the Reynolds number and the Strouhal number. The Reynolds number represents the ratio of inertial to viscous forces, while the Strouhal number reflects the effect of oscillatory boundary conditions. For a plane flow independent of the z-coordinate, we have w = 0 and em = 0. Then the three dimensionless momentum equations and the continuity equation, as given by Eqs. (3) and (7), respectively, in component form are reduced to .22. _a_u_ super. _1 __a2u _82u St at+u3x+vdy 3x+Re (8x2+3y2) (8) .31. fl. ELLE}; 1 32v _32v St dt+udx+vay 3y+Re (3x2+ay2) (9) 39 an 3v = 1 3x + ay 0 ( 0) where Eqs. (8) and (9) are the dimensionless x- and y-component Navier-Stokes equations, respectively, and Eq. (10) is the two-dimensional continuity equation. The above three equations are the describing equations of motion in terms of the three primitive variables u(x,y,t), v(x,y,t), and p(x,y,t). 3.3 The Problem for Solution: Vorticity-Stream Function Formulation In general, solving Eqs. (8)- (10) with the appropriate boundary conditions yields the velocity and pressure fields directly. The continuity and the Navier—Stokes equations, however, can be presented in several alternative forms for the solution. For two-dimensional flows, it is often desirable to introduce the vorticity and the stream function as dependent variables. This approach is attractive since it reduces the number of equations to be solved from three to two; and, second it ensures that the continuity equation is automatically satisfied. Failure to satisfy continuity using the primitive variables can lead to numerical instability (Mallinson and de Vahl Davis, 1973). On the other hand, the advantage gained by eliminating the pressure and its boundary conditions from the governing equations shows up as a difficulty in expressing the boundary conditions on the vorticity. By differentiating Eq. (8) with respect to y and Eq. (9) with respect to x and subtracting and invoking Eq. (10), we thus eliminate the pressure terms from the describing equations and reduce the number of momentum equations from two to one. There results 4o 3 an an 3 ii. fl. .1. .32.- it. 8' dt(ay dx)+u3x(3y 8x)+v8y(8y ax) = 1 82 Bu __§v_ + 32 On av (11) Re 8x2 3y 8x 3y2 3y 8x ° In plane flows, only the z-component C(x,y,t) of the vorticity vector exists; it is defined as 3v Bu = — - —. 12 C ax 3y ( ) Substituting the above definition into Eq. (11) yield the z-component vorticity equation a; a; at. 1 32: __§_32 , 8‘ at H ax H ay Re 8x2 + ay2 (13) The above equation is referred to as the vorticity transport equation. It states that the local and convective rates of change of vorticity are balanced by the rate of dissipation of vorticity. To fm'ther simplify the problem statement, the x- and y-components of velocity are expressed in terms of the stream function w(x,y,t) through u=—aN’—. v=-—a-‘JL. (14) By 8x The above definition satisfies the continuity equation identically. 41 It follows that the stream function is related to the vorticity by a Poisson-type equation derived from substituting the definition of Eqs. (14) into Eq. (12). The result is the stream function equation 32 82 Upon substituting for u and v in terms of \y from Eq. (14) into Eq. (13), there results a; 81 BQ _ 31$: 1 821; 32g St at + By 8x 8x 8y Re (8x2 + By2 . (16) The above equation is the z-component vorticity transport equation in terms of the vorticity and the stream function. Equations (15) and (16) are the describing equations that represent this two-dimensional flow. The result is a nonlinear system of two partial differential equations which involve the two unknowns C and w. Because of the nonlinearity of the problem, the solution is by no means straightforward. Hence, a simultaneous solution of Eqs. (15) and (16) with a prescribed set of initial and boundary conditions in the region bounded by the upstream condition, the far field downstream condition, and the lower and upper boundaries is required to give the solution in terms of C and v. Having found v, the velocities in the x- and y-direction can then be determined using the definition of the stream function given by Eq. (14). The above system of equations describing the motion does not lend itself to an analytical solution, hence, a numerical solution is attempted to yield the desired information. 42 For the sake of completeness, it is essential to mention that another way of expressing the describing equations is through a single fourth-order equation in terms of the stream function alone; this equation can be derived by further substitution of Eq. (15) into Eq. (16). The advantage of solving two second-order simultaneous differential equations in the vorticity and stream function instead of one fourth-order equation lies in the much greater rate of convergence of Poisson’s equation compared to the fourth-order equation (Woods, 1954). 3.4 The Steady Flow Problem The steady flow is, of course, a special case of the time-dependent problem. When the flow is steady, the problem has no independent characteristic time. In this case, the characteristic time is different fiom that given by Eq. (1) and is selected as H/U. Then, the dimensionless time becomes * t= if? . (17) The above definition allows the time-dependent equations to be reduced to steady-state equations. The two momentum equations given by Eqs. (8) and (9) are then rewritten as .22. an. _3“_=-_§L 1 32.. a2. at +0 8x +v 3y 3x + Re (3x2 + Byz) (18) _a_v_ .22. ALLA 1 82v 82v at +u 3x +v 3y 3y + Re (3x2 + Byz) (19) 43 where the Strouhal number has been eliminated. The continuity relation of Eq. (10) remains unaltered. It is noteworthy that even for the steady flow, the time derivatives are retained in the equations since the marching in time numerical methods require the presence of the time derivatives. In such a case, the time is fictitious. The asymptotic solution of the unsteady equations for large time is considered to be the solution to the steady-flow problem. This is referred to as a pseudo-unsteady method (Peyret and Taylor, 1983) or a false transient method (Mallinson and de Vahl Davis, 197 3). In terms of the vorticity and the stream function, the describing equation of motion for the steady flow are represented by Eq. (15) and the following modification1 of Eq. (16): _3£.+§1_§£_-§1_i§_=_1_ fire—2!... (20) at 3y 3x 8x 8y Re 3x2 8y2 3.5 Initial and Boundary Conditions To obtain a solution to the above describing equations, we must prescribe the appropriate initial and boundary conditions. In non-dimensional terms, the inlet is identified to exist at x = 0, the lower boundary at y = O, and the upper boundary at y = 1. The initial condition is assumed to be the quiescent state since the fluid is at rest fort S 0. In terms of velocities, 1 When the term 32C/3x2 is not included, Eq. (16) becomes equivalent to Prandtl’s boundary- layer equation. 44 u(x,y,O) = 0, v(x,y,O) = 0. (21) The velocity boundary conditions are the uniform velocity upstream, parabolic profile far downstream, and the no-slip boundary conditions on both the lower and upper boundaries. For t > 0, these are written as u(0.y.t) = 1. V(0.y.t) = 0 (22) u(xZLe,y,t) = 6y - 6y2, v(xZLc,y,t) = o (23) { 0 (fixed) u(x,0,t) = v(x,0,t) = 0 (24) U3 (moving, XleSXZ) u(x,1,t) 0, v(x,l,t) = 0 (25) where U]! is the velocity of the moving section of the lower boundary. The entrance length L° is defined as the distance from the inlet to the point where the centerline velocity reaches 99% of the fully-developed parabolic centerline value. At streamwise locations beyond Le the flow is independent of the x-coordinate. To ensure that the solution is not influenced by the downstream boundary condition, we select the location of the downstream boundary sufficiently larger than L.B such that the solution at the downstream end does not depend on x; the numerical solution is limited to the region influenced by x, i.e., x 5 Le. The choice of the x-location of the downstream boundary condition, given by Eq. (23), is validated by subsequent examination of the computed solutions. 45 The assumption of a constant inlet velocity was investigated in the experiments of J. Nikuradse (Prandtl and Tietjens, 1934). Nikuradse’s tests demonstrated that a uniform velocity profile is justified for tubes with well-rounded inlets. Furthermore, the influence of an actual experimental profile for the inlet velocity, obtained by fitting a polynomial near the wall to a uniform central section using measurements from a hot-wire anemometer, on the numerical solution was examined by Abdulla (1987). The difference in the streamwise velocity distributions along the channel due to the experimental profiles was found to be insignificant. The vorticity-stream function system of equations needs two initial conditions and eight boundary conditions. In terms of the stream function and the vorticity, the initial conditions are given by v(X.y.0) = y. C(X.y.0) = 0. (26) The boundary conditions on C and \y are derived from the known velocity boundary conditions. From Eqs. (15) and (16), it is apparent that the stream function equation requires four boundary conditions on \y while the vorticity equation demands four boundary conditions on C. On the upstream and downstream boundaries, the stream function for all y-locations is derived from the known velocity by integrating Eq. (14), i.e., w(x,y,t) = ] udy. (27) Upstream, where the inlet velocity is uniform, we obtain WOJJ) = Y- (28) 46 On the far field boundary downstream, where the streamlines are parallel to the walls, Eq. (23) and Eq. (27) lead to V(XZLe,y,t) = 3y2 - 2y3. (29) Obviously, even with the presence of a moving section, no fluid can flow across the lower boundary; hence, it forms a streamline which can arbitrarily be assigned a stream function with zero value, i.e., v(X.0.t) = 0. (30) On the upper boundary the velocity components are zero and the stream function v is given a value which is equal to the net flow rate; hence, we require that v(x,l,t) = 1 (31) where the dimensionless not flow rate is equal to 1. The vorticity boundary conditions both upstream and downstream are derived from the vorticity definition of Eq. (12). Since the y—velocity component is zero on either of these boundaries, Eq. (12) is then reduced to ___§_u_ C- ay' (32) It follows that the vorticity on the upstream and downstream boundaries are given by C(0.Y.t) = 0 (33) 47 C(x2L0,y,t) = 12y - 6. (34) On the lower and upper boundaries, the vorticity boundary conditions are not known explicitly. From Eq. ( 15), however, we can relate the boundary vorticity to the stream function in order to obtain a meaningful expression of the boundary vorticity. Since ‘1’ is constant on these boundaries, all its derivatives with respect to x, including azwaxz, are equal to zero. Equation (15) then gives 82 C(X,0,t) = ‘ -ainl_ -231, C(X.1.t) ay, (35) (36) The above two relations for boundary vorticity are rather general and need to be modified for subsequent use in the numerical solution. In the present form, however, no modification is necessary to accommodate the moving boundary; motion of the boundary shall be imposed when Eq. (35) is modified for the numerical solution. The boundary conditions, in terms of the stream function and the vorticity, used to solve Eqs. (15) and (16) are summarized below and displayed in Fig. 4. Upstream: v(0.y.t) = y. C(0.y.t) = 0 downstream (x 2 Le): w(x,y,t) = 3y2 - 2y3, C(x,y,t) = 12y - 6 lower boundary: v(x,0,t) = 0. 9090") = " $322 (37) (38) (39) 48 upper boundary: v(x,l,t) = 1, C(XJJ) = - %§-:g- (40) For the steady-flow problem, boundary conditions are dependent on x and y only despite the presence of fictitious time in the describing equations. Nevertheless, the boundary conditions are still given by Eqs. (22)-(24) in terms of the velocities and by Eqs. (37)-(40) in terms of the vorticity and the stream function; one exception is the plane entry flow problem where the boundary motion, given in terms of UB in Eq. (24), vanishes. The inclusion of steady motion of the entire lower boundary in the direction of the flow give rise to a Couette-Poiseuille flow problem. 112mm! (221mm I=0,V=0 n=6y~6y2,v=0 2 v=1oC=-—g-§- V=3yz-2y’.c=12!-6 y i=1 v=0 0 ’ I: v=0 V=Y9C= { Figure 4. Dimensionless boundary conditions. 49 In the presence of a finite step, as shown in Fig. 5, boundary conditions on the lower boundary must be modified. Specifically, boundary conditions in C and w on the step sides facing upstream and downstream must be accounted for. While ‘I’ on the entire lower boundary remains equal to zero, as given by Eq. (30) for a straight lower boundary, the vorticity boundary conditions on the upstream and downstream faces of the step becomes §(an9t) = T 933%", x = X19x2;05ySB (41) X since Nut/3y2 is equal to zero on these faces. Vorticity boundary conditions on the lower boundary segments both before and after the step as well as on the top surface of the step remain to be represented by Eq. (35). Figure 5. The flow problem with a finite step placed on the lower boundary. 50 To further examine the effect of the inlet velocity profile on the downstream development of the steady entry flow, an imaginary upstream extension to the channel walls, in the form of streamlines may be assumed. Taking the velocity distribution at the inlet section (x = O) to be unknown, the following boundary conditions are imposed in the region x < 0: u(-H,y,t) = 1, v(-H.y.t) = 0 (42) $6.0.” = 0, v(x,0,t) = 0 (43) 3y 113,14) = 0, v(x,l,t) = O (44) 3y where the uniform flow initiates upstream at x = -H instead of directly at the inlet. This is referred to as a cascade condition as it resembles an infinite cascade of equally-spaced plates in a uniform oncoming stream. In terms of the vorticity and the stream function, the above boundary conditions in the assumed region upstream of the inlet,become W(-H.y.t) = y. C(-H.y.t) = 0 (45) v(x.0.t) = 0. C(x.0.t) = 0 (46) v(x,l,t) = 1, C(x,l,t) = o. (47) Moreover, an inlet condition representing an experimental velocity distribution obtained from hot-wire measurements is used to compare the development of the streamwise velocity along the channel to that obtained using a uniform inlet velocity. THE NUMERICAL METHODS The main purpose of this investigation is to provide a detailed description of the velocity field by solving the nonlinear equations of motion. The finite-difference approximation, which is commonly used in solving partial differential equations, is selected to solve the describing equations numerically. A rectangular computational grid is constructed with a mesh size carefully selected since the total number of nodal points is limited by the available computer capabilities. The describing equations, expressed in terms of the vorticity C and the stream function \y, are first approximated by finiteedifference equations. In finite- difference form, the vorticity equation is approximated using the Alternating-Direction Implicit method while the stream function equation is approximated by the iterative Successive Over-Relaxation method. The present numerical solution is second—order accurate in its entirety since the numerical methods and boundary conditions have second-order accuracy. The objective is to compute the vorticity and the stream function at all the grid points for each time step. The two methods used in computing our time-dependent, viscous, incompressible, plane, flow problem are presented and discussed in details in the following sections. 51 52 4.1 The Computational Domain For this finite-difl‘erence scheme, a rectangular grid in the xy-domain, with grid lines parallel to the x- and y-axes, is superimposed to exactly fit the geometry of the flow problem including the moving boundary. As shown in Fig. 6a, the channel is divided into a finite number of increments such that the total number of nodes (including boundary nodes) is M in the x-direction and N in the y-direction; this results in (M-l) equally sized intervals in the x-direction and (N-l) equally sized intervals in the y—direction. Each node denoted by the subscripts ”i, j " represents a point (x,y) in the flow located at x = (i-1)Ax and y = (j-1)Ay where Ax and Ay are the mesh size in the x- and y-directions, respectively. The inlet is taken to exist at i = l, and the downstream condition is assumed to exist at i = M; the lower and upper boundaries are placed at j = 1 and j = N, respectively. Figure 6b illustrates the five nodes (x,y), (x+h,y), (x,y+h), (x-h,y), and (x,y-h) with h being the mesh size, equal in the x- and y-directions (h = Ax = Ay is to be used later in writing the difference equations); this representation for a typical interior node is used throughout our work to express the derivatives of the describing equations in discrete form. 4.2 The Alternating-Direction Implicit Method for the Vorticity Equation After a thorough search of the literature, it was decided that the Alternating- Direction Implicit (ADI) method would be used to solve the vorticity equation; the reasons are outlined below. The ADI method was first introduced simultaneously in two fundamental papers by Peaceman and Rachford (1955) and Douglas (1955); 53 ,1 _,| I, |,_ noun I 2 i-l 1 1+1 M-l M x a) finite—difl'erence grid structure and field boundaries. (Kym) . i9j+l (It-ha) (Xi!) (XHIJ) . . . i-l,’ Li “I” (xix-h) 0 hi1 b) finite-difference cell and notation for a typical interior grid point. Figure 6. Finite-difference representation of the computational domain. 54 further details were given by Douglas and Rachford (1956). The principles, variants, and improvements of the method were also discussed in numerous articles such as Birkhoff and Varga, ( 1959), Birkhoff et al. (1962), Mitchell and Fairweather (1964), Fairweather and Mitchell (1967), and Lynch and Rice (1968); a generalized form can be found in Douglas and Gunn (1964). A general description of the method is presented in many texts including Amos (1977), Mitchell (1969), and Varga (1962). Moreover, the method of Fractional Steps (or method of Splitting), which is somewhat similar, and in some cases identical, to the ADI method, can be found in Yanenko (1971). In addition, early applications of the ADI method in the numerical solutions to thermal-fluid problems were given in many publications, for example, Wilkes and Churchill (1966), Samuels and Churchill (1967), and Aziz and Hellums (1967) used the method in their calculations of natural convection. Kuskova (1968) used what he called the method of Variable Directions in his solution of the flow in a rectangular cavity. The specific implementation of the method to the Navier-Stokes equations along with detailed discussions can be found in Roache (1982) and Peyret and Taylor (1983). Indeed the ADI method continues to be used in all forms of flow calculations. The ADI method is currently the most popular approach to solve viscous flow problems (Roache, 1982). Most recently, Caughey and Iyer (1989) used the ADI method in their computation of the supersonic flow of two-dimensional inlet flow fields, using Euler’s equation. The ADI method is a multistep, multidimensional, second-order accurate implicit scheme used to solve partial differential equations. It is characterized by unconditional stability when used with the proper boundary conditions; unconditional 55 stability in the case when the Dirichlet-type boundary conditions are not time- dependent can be expected. It also provides a reduction in computing time by a factor of two or more (Roache, 1982) since it allows larger time steps than the explicit methods. In the ADI method, each time step is split into two halves resulting in two finite-difference equations each of duration At/ 2 where At is the time step. The first equation is implicit only in the x-direction while the second equation is implicit only in the y-direction. Equations are used in turn over successive time steps. Moreover, the ADI method permits the construction of an efficient and implicit scheme that requires only the inversion of a tri-diagonal matrix; full implicit schemes yield a tri- diagonal system of equations only for one-dimensional problems. In the x-equation, all implicit terms associated with x-derivatives are evaluated at t = (k+l/2)At, where k denotes the number of time steps, while all terms associated with the y-derivatives are evaluated at t = k At. In the y-equation, however, all terms associated with the x-derivatives are evaluated at t = (k+1/2)At, while those implicit terms associated with the y-derivatives are evaluated at t = (k+1)At. Using the notation Cit.) to represent C(x,y,t), the two ADI equations for an interiornode (i = 1, 2,..., M-1;j = 1, 2,..., N-l) in discrete form are (k+1/2) (1!) St ‘11 ' Cir At/2 (k) (k) (hr/2) (k+1/2) (k) (k) (k) (k) Vi.j+r ' V1.14 Ci+r.j ' Curd _ i+1,j ' ‘Vi-r.j Ci.j+r ' Cid-1 2Ay 2Ax 2Ax 2Ay (43) (RH/2) (k+l/2) (k+1/2) (k) (k) (k) l [ Cm; ' 2911.1 + Ci—Lj + Cider ' 29m + Cigar ] 2 Re (Ax)2 (Ay) (k+l) (k+-1’2) C - C id in‘ 3‘ At/2 (k+l/2) (k+1I2) (k+1I2) (k+1/2) (k+1/2) (k+ll2) (k+l) (k+l) + i,j+l ' i,j-l Ci+1,j ' Ci-IJ _ Vi+l¢ ' Wing Ci,j+1 ' Cid-1 2Ay 2Ax 2Ax 2Ay (ht/2) (k+1/2) (Int/2) (k+l) (k+1) (k+l) = 1 Cm.) ' 29m +Ci-l,j + Ci.j+r ' 29rd + Cid-l . (49) Re (Ax)2 (Ay)2 In Eqs. (48) and (49), the coefficients of the nonlinear terms, namely the velocity components u and v, are given in terms of the stream function using the definition of Eq. (14). In difference form, they are given by urn- = mi; 3"“ (50) Vid’ = Vi'l'jzaxmfl'j ' (51) All derivatives in the above equations are expressed using second-order central—difference discretizations. In computing C i, j at a new time increment, ‘Vi.j and its first-order space derivatives are considered taken from the previous time step and thus treated as known coefficients. The vorticity equation then becomes a linear, parabolic difference equation for C. This effectively decouples the vorticity and the stream function equations. The ADI method has a formal truncation error of o [ (1102, (Ax)2, (Ay)2] (Anderson et al., 1984). Ideally, in order to preserve the second—order accuracy in . . . 1 . time, the coefficrents of the nonlrnear terms must be evaluated as “if; I2) and via? 1n 57 it: 1/2) (k+1). and vi]- Eq. (48) andu 1n.Eq (49). Since u and v are given in terms of w, which is calculated from Poisson’s equation, this evaluation procedure requires the coupled solution of the vorticity and stream function equations at both time steps (k+1/2) and (k), which is impractical. Several alternatives and their associated error were discussed by Roache (1982). When the old values do?) and v."‘.’ are used In both 141-] equations, the formal accuracy deteriorates to O[At, hz] (Son and Hanratty, 1969). Briley (1971) calculated (if j)and vf" and linearly extrapolated forward tou 9‘1“”) and 1 - via: I2) using stored values of “(1:11) and v3?) ; the procedure was stable and second- order accurate but requires additional storage. The second—order accuracy is also retained (Roache, 1982) when the stream function derivatives are reevaluated after (HI/2) the first half of the time step by solving the stream function equation for Vi j . The “(ht/2) (k+1/2) . and v in nonlinear coefficients then become u:k j)and vfk j )in Eq. (48)an U M Eq. (49). Using a uniform mesh with spacing h = Ax = Ay, Eqs. (48) and (49) become St— 2( C(k+-1’2) - (k l/2)_ C(k 1/2) +(‘l’i,j+l Vi,j 1)(ci+l,j C131 ) (vi+)1,j Vi- (k1,j) (cm i,j+l Cid-1 (k+1/2)_ C(k+l/2)+ (k+l/ 2) (k) [ ( c1341 Ci- 1 J )+ ( C1j+l- Cid) +ci,j-l ] (52) 58 (k+1) C(kJ+l/‘2) St— 2.,(t (k+1l2) (RH 12) (k+1/2)_c(k+1/2) (k+ll2)_ WGHIIZ) (k+1) Cowl) + (VLj-t-l V111 )(Ci-l-IJ Ci- 1,j ) (Wi+l,jVi-1j )( Ci ,j+1 ci .j-l (k+lf2)_ C(kJH/Z) C(kH/Z) (k+1) C-(km (1m) —R4—c-[( Ci‘i'loj Ci- -,lj )+ +(Ci .-j+1 + Cid-1 ] e (53) (HI/‘2) The first equation is solved for the intermediate value Cij (assuming all variables at time step (k) are known) which is, in turn, used in the second equation; this intermediate value can be considered as an approximation of the exact solution2 (Peyret and Taylor, 1983). The second equation then gives the solution Cg?” at the end of the time interval. Each of Eqs. (52) and (53) represents a system of linear algebraic equations which leads to a tri-diagonal matrix. The two sets of (M-2)x(N-2) equations are solved simultaneously at each time step. A number of suitable direct and iterative methods are available for solving such systems. Tri-diagonal systems, however, can be readily solved using the Thomas algorithm, which is a simplified adaptation of the Gaussian elimination method in the form of a recursion formula. 2 lnusingguraalmuldstepmethodadnquesdonofhowwobmindrevaluesofthe intennediatequantitiesontheboundaryarises. FortheADlmethod,inwhichtheequationusedin eachnepiswnsiswntwimdwmighmlequadon,ahrgemcadmamrwuhmspeamdme appears fa points adjacent to the boundary (Fairweather and Mitchell, 1967, Mitchell, 1969) if the exacttime-dependentboundaryvalueisusedtodeterminetheprovisimalvalunggl’z)onthe boundariesparallelmthey-axis,requiredtosolveEq.(52). Onewaytoovercornethisdifficultyis todefineaboundaryvaluesuchthatthecombinationofbothtimestepsyieldssecond—orderaccuracy atpointsadjacenttotheboundariesaswellasirrnerpoints. Thiscanbeaccomplishedifthevalue C931”) atthcbomdariesisdeducedfromaveraging its valueattimcsteps(k)and(k+1). Other alternatives are presented in Peyret and Taylor (1983). 59 Due to the diagonal dominance of the coefficient matrix, the ADI method can significantly improve stability at high Re and, hence, extend the domain of convergence. Nevertheless, time—dependent boundary conditions can destroy the unconditional stability of the ADI method (Peyret and Taylor, 1983). In this case, a limitation on the time step is necessary to ensure the condition of diagonal dominance of the coefficient matrix for Cid" The diagonal dominance is preserved if we have luid-l S 2/(Ax Re) in Eq. (48) and lvid-l S 2/(Ay Re) in Eq. (49). If these conditions do not hold, however, the constraints on the time step are given by (Peyret and Taylor, 1983) 2112 2112 At = . At = _ luiJle " 2/RC lvi’ley - 2/RC (54) These satisfy the condition of diagonal dominance. The effect of the boundary conditions on the stability of the method and the coupling between the vorticity and the stream function on the boundary has been studied by Bontoux et al. (1980). Although the limitation on the time step is not very restrictive at low Reynolds number, for flows with moderate and high Reynolds numbers the stability criteria require such extremely small time steps and/or mesh size that the computation becomes costly; the maximum magnitude of the eigenvalues exceeds unity so that the von Neumann stability condition cannot be satisfied for any At/h. To remedy the situation, the first-order derivatives of C in the nonlinear convective terms in the vorticity equation are represented with forward/backward differences instead of the central differences used in Eqs. (48) and (49) (Forsythe and Wasow, 1960). The resulting scheme is stable. Consequently, the second-order accuracy of the central- difference approximation in the space variables is lost since the forward] backward 60 differences have only first-order accuracy. By correcting the signs of the nonlinear terms, this technique maintains the diagonal dominance of the system of equations at high Reynolds number and, hence, helps avoid the instability of the numerical solution. Schemes using such an approximation are found in the works of Barakat and Clark (1966), Torrance (1968), Greenspan (1969a); also, Lilly (1965) used this technique to stabilize his numerical solution of geophysical problems. The technique, first suggested by Lelevier (Richtrnyer and Morton, 1967), requires the use of backward differences for positive velocity components It and v (expressed in terms of derivatives of v) and the use of forward differences for negative velocity components. In so doing, the stability criterion becomes weakly dependent on Reynolds number (Pao and Daugherty, 1969). The stability criterion for this forward] backward difference approximatioh is 4 lvi.'l [UL-I :1 —T + —-'— + —-l— At 5 l. [ Reh h h (55) The second-order accuracy is recovered at convergence (Peyret and Taylor, 1983), however, by alternating the direction of the forward] backward differences, representing the first derivatives of C, at each time step (according to the signs of the coefficients of the nonlinear terms, i.e., the velocity components u and v). This is achieved by considering upwind differencing for the implicit part and downwind differencing for the explicit part. During the transient stage, the scheme is only first- order accurate in time and at steady-state the truncation error becomes of 0[ (At h), hz]; this technique gives a centeral-difference approximation at convergence. Implementation of the scheme associated with this upwind/downwind non-centered 61 approximation requires the replacement of the following differences in Eqs. (52) and (53): (CH-1,j ' Ci-lJ) = (1' a)(§i+1,j ' Cu) + (1 '1'“) (Cid ' CL”) (56) (Ci.j+l ' C1,j-l ) = (1" B)(Ci,j+1 ' Chj) + (1+B)(Ci,j ' Cidq) (57) where a = sign(ui.j) and B = - sign(vi.j) for the x-equation, and a = - sign(ui,j) and B = sign (Vid) for the y-equation. The unconditional stability is preserved (Peyret and Taylor, 1983) so that at each time step, the matrix is diagonally dominant. The above technique was proposed by Peyret (1971) for the solution of the incompressible Navier—Stokes equations in the primitive variables, and was applied by Daube and Ta Phouc Lee (1979) and by Ta Phouc Lee and Daube (1981) for the vorticity-stream function formulation. Finally, the solution to the steady-flow problem is obtained as the asymptotic solution for large time of the unsteady equations (thus, the term ac/ax is retained in the vorticity equation). Time steps are fictitious and each is merely considered equivalent to an iteration. The use of this pseudo-unsteady method has been introduced by Burggraf (1966) in association with an explicit-implicit discretization with respect to time. 4.3 The Successive Over-Relaxation Method for the Stream Function Equation A commonly used technique of solving a system of simultaneous equations is the iterative Successive Over-Relaxation method. It has been successfully utilized in 62 the solution of many important problems of mathematical physics including the subclass of elliptic, linear, two-dimensional partial differential equations, such as the Laplace and Poisson equations with Dirichlet or N eumann boundary conditions on a closed contour. The method, introduced independently by Frankel (1950) and Young (1954), is thoroughly described by Forsythe and Wasow (1960), Greenspan (1965), Wachspress (1966), and Varga (1962). General applications of the method to fluid flow problems can be found in Chow (1979) and Arpaci and Larsen (1984). Mostly, the SOR method is desirable since, due to its simplicity, it can be readily implemented. Furthermore, it does not require large computer storage, which is a critical problem when a large number of nodes are involved. In the SOR method, however, inner iterations are required at each time step. An initial guess at the solution is improved with subsequent successive approximations until the solution converges to within specified accuracy after a finite number of iterations. The stream function for iteration (k+-l) is combined with the kth iteration using a relaxation factor. The use of a relaxation factor reduces the number of iterations required to achieve a solution, and therefore, improves the convergence behavior of the SOR method over other iterative procedures. The stream function equation, given by Eq. (15), in finite-difference form is Yum ' 2‘Vi.i+ Vi-l,j + Vi,j+l ' 2‘1’i.j + Vij-l (AK)2 (Ay)2 = - Cr,- (58) where second-order central-difference approximations are employed to approximate the second-order derivatives at all interior mesh points with an error of O[(Ax)2, 63 (Ay)2]. With the vorticity Cid. considered as a known constant, the above represents a linear, elliptic partial difference equation for Win" In the explicit point-iteration method (i.e., the Jacobi method) we march through the grid from node to node solving Eq. (58) explicitly for the new value “(31m using old values VIE-l) at adjacent nodes known from the previous iteration; here (m) indicates the number of iterations. This is often referred to as Richardson’s iteration. To speed up convergence, as in the Gauss-Seidel method, old values of adjacent nodes are replaced by new ones as soon as they become available. Rearranging Eq. (58), and using a uniform mesh of size h, we obtain Liebmann’s iterative formula (n+1) = _1_ (m1) (m) (n+1) (m) (m) U 4 1: “J + Vi+l.j + Vin-1 + Vi.j+1 + h2§i.i :1 (59) As a refinement to the above methods to further improve the speed of convergence, the SOR method employs the correction formula in which the stream function for iteration (m+1) is combined with its value from iteration (m) using the over-relaxation factor FS (if? 1’ = FS mg?” + (1 - FS) (if? (60) where (riff 1’ is a relaxed value of the new value Vi?” obtained from Eq. (59); this relaxed value is an adjusted (or extrapolated) value to be used in lieu of V331“) explicitly in the next iteration; vi? is the value from the previous iteration as modified by a previous application of this formula if over-relaxation is being applied 64 successively at each iteration. In the above equation the over-relaxation factor F S controls the rate of convergence. It has a value 1 S FS S 2 and must be optimized for a given computation in order to maximize the rate of convergence (i.e., reduce the number of iterations) of the SOR method. For Poisson’s equation, the optimum value of FS depends on the mesh size, the shape of the domain, and the boundary conditions. For the present Dirichlet problem in a rectangular domain with uniform grid of size (M-1)h by (N-1)h, it has been shown (Chow, 1979) that the optimal value of the relaxation factor is calculated from the analytical expression _ 2 (l -/ 1- §2 ) F809, _ g2 (61) where g :3 COS[I/(M-l)] :- COSIE/(N‘ln (62) With the determination of the optimum over-relaxation parameter FS, it is possible to reduce the required number of overall iterations in the solution by more than one order of magnitude from that required by Gauss-Seidel iteration (Thompson, 1969); in addition, we may reduce the restriction placed on the maximum size of the spatial step imposed by the Gauss-Seidel technique. Utilizing Eq. (60), we can then rewrite Eq. (59) as 65 .. (m+1) _ _F§_ (m+1) (m) (m+1) (m) Vii ‘ 4 [ i-1.i + i+l.j + ‘Vioi-l + Vi.i+l + “2 Cid] + (1 - Fs) Vi? (63) The above is an iterative method which improves the approximation one point at atime. The iterative SOR by blocks (or lines) method is used to implicitly solve Eq. (15) for the stream function. The time of computation can be reduced by iterating on an entire block of points; a block represents a column or a row of node. For the present inner iteration (m+1), the stream functions for a block of nodes, considered to be columns, are determined from the values at adjacent nodes. At each column i = 2, 3,..., M-1, Eq. (40) is rewritten as W331) _ 4Worm) + (n+1) = - [Won-+1) + “[010 + h2 Cm." ] . (64) i,j i,j+l i-l,j i+l.j i,j . . l l 1 A column of nodes wrtlr stream functron values vi? ), v(im; ),. . ., Vim; ) for a given i constitutes a system of (N - 1) simultaneous linear algebraic equations with a tli—diagonal matrix. One iteration for each column i will implicitly solve for WES-1+1) values at all nodesj = 2, 3,..., N-l. This procedure is repeated for all i = 2, 3,..., M- 1 , one column at a time, successively updating known Vi? values from the previous +1) I) Relaxed values ‘17:“; iteration by new values Vi“? . are computed using Eq. (60) l as soon as tug-H) become available. These relaxed values are used in the computation at the next adjacent column and in the subsequent iteration as well. 66 Solution of the system of equations associated with columns i = 2, 3,..., M-1, yields the values of the stream function at all interior nodes. This approach is expected to speed convergence since the implicitness makes boundary conditions at j = 0 and j = N felt at all interior nodes along a given column (Arpaci and Larsen, 1984). The old incorrect values in adjacent columns and the added complexity of the solution of the tri-diagonal matrix, still slow the convergence. Ames (1977) indicates that, compared to the Gauss-Seidel iterations, the SOR by lines method requires only 1N2 as many iterations to reduce the initial errors by the same amount. In the SOR method, the solution is improved with successive iterations until it converges to within a specified accuracy after a finite number of inner iterations. 1 Except for the points where Vita)“ ) = 0, the iteration process is terminated as soon as the relative error criterion (n+1) (m) . . - \ll. . 1.1 1.1 (m+l) S E (65) i,j is satisfied at all interior nodes, where e is a preassigned tolerance error. Conversely, in order to save computing time, if the above is not satisfied, the process shall be terminated when a preselected maximum number of iterations is attained, even if the results have not yet reached the desired degree of accm'acy. In general, the maximum number of inner iterations required for convergence is directly proportional to the square of the number of equations to be solved (Roache, 1982), i.e., mm ~ [(M-2)x(N-2)]2. 67 4.4 Initial and Boundary Conditions In order to solve the difference equations for the vorticity and the stream function, initial and boundary conditions on C are needed for Eqs. (52) and (53) and on \y for Eq. (64) (or Eq. (59)). The initial conditions are given in Eq. (26). The boundary conditions are given by Eqs. (28)-(31) for \y and Eqs. (33)—(36) for C. The finite-difference form of these initial and boundary conditions are presented herein. First, the initial conditions (at time t = 0) on the stream function and vorticity canbeputintlreform ‘l’i,j = (I'Dh. chj = 0° (66) In the case of the stream function, stating the boundary conditions in finite- difference form is straightforward. At all time levels, these boundary conditions are asfollows: upstream: \le = (j-1)h (67) downstream: VMJ = “(1'1)“2 - 2[(.i-1)h]3 (68) lower boundary: (rm = o (69) upper boundary: ‘Vi.N = 1. (70) Likewise, the vorticity boundary conditions on the upstream and downstream boundaries are directly written as upstream: CLj = 0 (71) downstream: CM.j = 12[(j-1)h] - 6. (72) In addition, an explicit vorticity boundary condition is necessary to replace the second-order derivatives in Eqs. (35) and (36). Vorticity is generated if there is a no-slip condition at the boundary; thus it must be incorporated in the numerical solution. Expressing these conditions in finite-difference form is complicated since the no—slip boundary conditions are difficult to handle computationally (Pearson, 1964, 1965a). The motion of the lower boundary also enters these conditions. The vorticity at each boundary point is determined by expanding the stream function at the node adjacent to the lower boundary in a Taylor series, in terms of the stream function at the boundary, and keeping the third-order term, i.e., ..—.. h_ay_ 1‘2 32‘! i131- H...73 “'2 “’14- (3Y)i.r+ 2 (3Y2 1.1+ 6 3Y3 i.r+ OT ( ) where derivatives of w on the boundary in the x-direction are zero. Neglecting terms of higher order, and noting that for the moving section of the boundary (av/8y)“ = ui.l = UB and that, from Eq. (15), (32W/3y2)i.1 = - C“, we get 2 3 Via = ‘Vi,r + 11 U3 ' “1'15— “ - 316—635.)“. (74) Upon substituting the approximation (fill 3 Ci ‘Cir (75) By h into Eq. (74) and rearranging, we obtain the second-order (error is of O[h2]) "Woods’ condition" (Woods, 1954) (k) (k) (k) (k+” ‘9‘“) __ 3(Vi,l ‘ Wm) _ i,2 3UB (76) "1 ‘ h2 2 + h which gives the vorticity on the lower boundary in terms of the stream function and vorticity obtained from the previous iteration; the superscripts (k+1) and (k) denoting the present and previous time steps, respectively, are added to indicate the time of the available information; since the value of boundary vorticity depends on time, it is recalculated at every time step. It is also clear that motion of the boundary enters the problem through the vorticity boundary conditions. The velocity of the moving portion of the boundary can be a constant as is the case with steady motion or a known function of time expressing the oscillation of the lower boundary (U3 = 0 on the fixed portion of the boundary). Woods’ condition has been used in particular by Bozeman and Dalton (1973). An equivalent formula can be derived for the upper boundary, with does not include boundary motion; that is (k) (k) (k) C(1+1) 3 (th ' VLN-l ) C._._12 (77) i.N = h2 - 2 . Another frequently used second-order formula for determining the vorticity at the boundary was introduced by Jenson (1959). It is derived in a manner analogous 70 to Eq. (76). A linear combination between the expressions of the stream function at two points adjacent to the boundary allows us to evaluate az‘wayz. The result is (k) (k) (k) (k+” (k+!) = 7V}; - 8WB+1 + VB+2 + flL. (73) B 2112 h Here the subscript "B” represents a boundary point and "B+1" a neighboring internal point in a normal direction, regardless of the orientation of the boundary. The same expression can be obtained from approximating \y near the boundary by a third-order polynomial. The above formula was used by Pearson (1964, 1965a) and Wilkes and Churchill (1966). The stability of Woods’ condition was considered by Taylor (1970). Briley (1971) considered a formula of the same type which gives very little change in CB but converges faster since it permits larger time steps. However, extra programing effort is required since the resulting matrix is not tri-diagonal. Briley’s rational and derivation are reproduced in Roache (1982). An extensively used first-order condition for the boundary vorticity was proposed earlier by Thom (1933). After ignoring the second-order term in Eq. (73) and following a similar analysis to that leading to Eq. (77), the following expression is obtained: (k) (k) (k+1) (k+!) _ 2 (VB ' VB-t-l + ZUB (79) B ’ h2 h ' An alternative first-order formula was derived by From (1963, 1964) to give the boundary vorticity in terms of the stream function. Of course, expressing the 71 vorticity on the boundary in terms of the stream function can also be obtained from a first- or second-order finite-difference representation of C = - azwayz. A comparison of the errors associated with Eqs. (76) and (79) is presented in Section 6.1. Because C at the boundary is calculated using previous values, it can lag behind the values in the internal region by At (Roache, 1982). For large At, the calculation can become inaccurate and an iteration procedure is required. An alternative technique is the use of weighting factors to slow down the solution in the interior region; this is described in Chapter 5. Inthepresenceofafixedsteponthelowerboundary,avorticityboundary condition equivalent to Eq. (76) is used to replace Eq. (41) on the side surfaces of the step. Also, for the problem with a cascade inlet condition, the condition Cid = 0 is used at all boundary points having zero vorticity as required by Eqs. (46) and (47). A closing remark concerning the vorticity boundary condition is in order. Methods employing creation-type conditions were discussed by Anderson (1989). In the methods based on vorticity creation a constraint, in the form of an integral, is imposed on all boundary points. These constraints are equal in number to the number of unknown boundary values and, therefore, are used to close the system of equations. The value of the vorticity on the boundary is determined when the time derivative of the constraint vanishes as the vorticity is evolved from the solution of the vorticity equation. Anderson demonstrated that vorticity creation boundary conditions, when used in numerical solutions, are generally equivalent to those which exploit the relationship between the vorticity and the stream function. THE COMPUTATIONAL SCHEME The mechanics of the computational scheme adapted for this study is disclosed in the following sections. The computational scheme, used in conjunction with accelerating parameters, is comprised of the Alternating-Direction Implicit (ADI) method and the Successive Over-Relaxation (SOR) method. A brief description of the implementation of the scheme and the computing environment are also given. 5.1 Accelerating Parameters Fume-difference approximations used when modeling physical problems often lead to the solution of large-order sparse linear matrices. It has now been established that central-difference approximations provide convergent and accurate numerical solutions only for low Reynolds numbers. For large Reynolds numbers (greater than 200), however, upwind differencing is essential to ensure the stability of the numerical solution by maintaining the diagonal dominance of the system of equations. As a supplement to the numerical methods, the vorticity values obtained from solving Eqs. (52) and (53) and the stream function values obtained from solving Eq. 72 73 (64), are modified by introducing certain weighting factors into the numerical scheme. These factors are often referred to as accelerating parameters. Another type of accelerating parameters is the time step for the vorticity equation. The over- relaxation factor used in conjunction with the SOR method in the solution of the Poisson-type elliptic stream function equation is also considered as an accelerating parameter (Thompson, 1969). In general, accelerating parameters are used to minimize the time of computation. The use of accelerating parameters, however, have a significant effect on the convergence characteristics of the difference equations; they provide numerical stability by maintaining the diagonal dominance of the system of difference equations, hence, expand the domain of convergence. The idea behind the introduction of weighted averages is to avoid any possible divergence that results from any over-correction to the value of the vorticity or stream function; the numerical solution is slowed down whenever over-correction occurs. This procedure stabilizes the numerical solution by compensating for the coupling between the vorticity and stream function equations. More so in the case of the vorticity equation, the function of the weighting factor is to slow down the solution in the interior region in order to reduce the lag in vorticity boundary conditions behind the solution in the interior region, since the vorticity boundary condition is calculated from the stream function and vorticity values of the previous time step. Weighting factors are used to modify the values of C and v in the interior region. They allow for different fractions of the old and the new values obtained from subsequent time steps to be used during the matrix solution. Smoodrin g formulas are 74 introduced for this purpose (Greenspan, 1969a, 1974). The weighted averages of t; and v at the end of each time step are taken as linear interpolations of the new values calculated at the end of the present time step (k+1) obtained, respectively, from Eqs. (53) and (64) and the weighted values of the previous time step (k); that is, CE?” = (1 - a) Q?" + 0 cf]? 0 $051 (80) 4M) _ (k+» (k) i,j — (l - 'y) Vi.j + 7 Vid 0 57s 1 (81) where o and y are the weighting factors, and CE?” and fig“) are the weighted averages of the vorticity and the stream function, respectively. The values of the above weighting factors fall within the range of 0 to l and are not necessarily equal for the vorticity and the stream function. A similar formula can be used to give a weighted average for the value of the vorticity at the boundary (Peyret and Taylor, 198 3). However, the weighting factors corresponding to the interior points and the boundary points are not necessarily equal. The effect of the values of the accelerating parameters on the convergence of the solution and the computing time depend on the shape of the computational domain, mesh size, and the Reynolds number; in essence, it depends on the number of nodal points in the computational domain. Although the objective of using these parameters is well understood, the determination of their optimum values is not straightforward. Optimum values of the accelerating parameters (At, FS, 0, and y), with the exception of the over-relaxation factor F S for the stream function equation, are determined primarily by numerical experimentation. The optimum value of FS used in 75 the iterative SOR solution of Poisson’s equation can readily be calculated from Eq. (61). It should be noted that the stability and rate of convergence of the numerical solution is dependent on the size of the time step At . 5.2 Outline of the Computational Procedure The details of the computational scheme used to implement the numerical methods, discussed in Chapter 4, for computing our time-dependent, viscous, incompressible, two-dimensional flow problem are illustrated in this section. The objective is to compute the stream function \y and the vorticity C at all the grid points for each time step by solving the governing equations in their finite-difference form. The ADI method is used to solve the vorticity equation, while the SOR method is used to solve the stream function equation. The x- and y-component velocities, as well as the pressure gradients, are then calculated from the converged solution. After discretizing the equations and boundary conditions, the computational domain is defined; it depends on the Reynolds number. The mesh size and the time step are then chosen. The computation begins at t = 0 by assuming C8.) = 0 and WE? = (j- l) h as initial values at all nodal points (Eq. (66)). Boundary conditions given by Eqs. (67)-(72) are then calculated. The values of the vorticity {B on the lower and upper boundaries are also computed using Eqs. (76) and (77). When boundary motion exists, the velocity of the boundary is determined at the end of the time step and used in the calculation of CB. 76 The vorticity C115) at all interior nodes is computed at the end of the first time step by solving Eqs. (52) and (53) sequentially using the ADI method and available 2) values for the stream function VF}; new values for V3,! ate obtained from Eq. (64) before solving Eq. (53) for the second half of the time step. Then, utilizing the known values of the vorticity cf]? as constants, the stream function will.) is obtained by solving Eq. (64) using the iterative SOR by columns method and employing the relaxation formula of Eq. (60) (for the SOR by points method Eq. (63) is used). The inner iteration process of the SOR method is continued until the assumed error criterion, given by Eq. (65), is satisfied at which point the iteration is terminated; the tolerance error e = 10" is used. If convergence does not occur for some preselected maximum number of iterations, the computational procedure proceeds to the next time step to compute new values for Cid and WM. (1) The values for cf}; and Win am used in turn to determine the boundary conditions for the vorticity and the coefficients of the nonlinear terms in the vorticity equation for the next time step, k = 2. Subsequently, CS.) and WE? are determined. This above computational procedure can be generalized and repeated in the same fashion with general steps (k) and (k+1) as the old and present time steps, respectively, to yield a new estimate of the vorticity Cit-+1) and stream function ng’tl) at the end of each time step. (k+l) i,j (k+1) The computed values C M are corrected at the end of each time and \y step using the weighting factors. Incorporating the weighting factors to modify the 77 values of the stream function and vorticity in the interior region ensures convergence of the numerical solution. The computation continues as the computational scheme progresses in time. The computational procedure ends as soon as the relative error criterion, given by (RH) _ fit) iii hi _ -6 fowl) S e — 10 (82) i,j f selected as a condition for convergence, is satisfied at all interior nodes; nodes with Chi = 0 and Via‘ == 0 are excluded. Here a is the preassigned tolerance error and f represents either Cid. or Via" The numerical solution converges when Eq. (82) is satisfied at all nodal points for two successive time steps. For the time-dependent flow with an oscillating boundary, the above error criterion is enforced over an entire time period before convergence can be assumed. The computational procedure is terminated after a preselected number of time steps even if convergence is not attained. In summary, the numerical procedure is executed according to the following steps: 1. Assume initial values for Ci.j and ‘Vid' at all interior nodal points as given by Eq. (66). 2. Calculate the boundary conditions on C and W from Eqs. (67)-(72). 3. Calculate the values of vorticity CB on the lower and upper boundaries using Eqs. ('76) and (77). 78 4. Successively for every interior nodal point: a. Compute the vorticity Cid“ from Eqs. (52) and (53) using values for via. from the previous time step. For the second half of the time step, however, compute new values for V” by solving Eq. (64) using values of the vorticity Ciu’ obtained from the first half of the time step. b. Compute the values of the stream function ‘Vid’ from Eq. (64) using the already available values of CM as constants. Carry on the iterative process for some preselected maximum number of inner iterations or until a certain error criterion is satisfied. c. Correct the values of (i,j using Eq. (80). d. Correct the values of Wm using Eq. (81). 5. Continue to advance the solution in time until the relative error criterion of Eq. (82) is satisfied at all interior nodal points. 6. Return to step (3) and repeat the process if this error criterion is not satisfied. The above computational procedure is illustrated by the flow chart of Fig. 7. Finally, the velocity components u and v are calculated from the values of the stream function using a five-point, finite-difference approximation of the definition of the stream function, as given by Eqs. (50) and (51), respectively. The streamwise pressure gradients of interest, at the wall and centerline, can also be calculated in a similar manner; the pressure gradient can be obtained from Eq. (18) and then expressed in finite-difference form. 79 New time step (iteration) Figure 7. Flow chart of the computational procedure. 5.3 Computing Environment Computer programs were written to carry out the computational procedure as outlined in the previous section. The FORTRAN programing language is used with double precision for the real variables. The code for the oscillating boundary problem is reproduced in Appendix A. Since the algorithm is written in general terms, the other codes are similar and, hence, will not be presented. A cluster that includes a DEC VAX/VMS 11/750 mainframe computer with 8 megabytes of memory and three DEC MicroVAX-II minicomputers, each with 9 megabytes of memory have been used to perform the computations. When the number of nodal points becomes quite large such as in the case of high Reynolds numbers and small mesh size, both the number of operations at each time step and the number of time steps required to achieve convergence increase dramatically; the time of computation increases accordingly and the convergence process becomes tediously slow. A larger computer would handle the computation more quickly. Also, since vector processing can speed up the computation, a vectorized computer program executed on a supercomputer or a machine with parallel processors would significantly reduce the time of computation. ANALYSIS The topics discussed in this chapter are related to the characteristics of the numerical solution of the Navier-Stokes equations. In any numerical solution, it is often useful to look into the stability and consistency of the solution, as well as the error associated with it. The difference representation of the partial differential equations of motion must meet the conditions of consistency and stability in order to be considered acceptable; generally, this implies convergence. These issues will be dealt with in the following sections. Furthermore, the finite-difference approximation of the vorticity boundary condition is analyzed to determine its influence on the numerical error; this will be addressed first. 6.1 Error Estimate for Vorticity Boundary Conditions It is often postulated that the overshoot in the velocity profile near the boundary is due to the order of approximation of the vorticity boundary condition. In order to support the hypothesis that this overshoot is the outcome of an inadequate numerical approximation, a simple error analysis is presented herein. 81 82 Let the velocity profile be represented by the polynomial u(y) = a + by 4» cy2 + dy3 where a, b, c, and d are arbitrary constants. At the centerline, the velocity is uCL and its derivatives are (Bu/By)“ = O and (Eizu/Eiyz)CL as 0. Upon incorporating this information into the definition of the velocity and solving the resulting system of equations, we obtain the following values for the constants: a =-- 0, b = 2uCL, c = -3uCL, d = “or; (83) The error associated with the first-order accurate vorticity boundary condition (Thom, 1933) is considered first. In the derivation leading to Eq. (79), the third-order terms of the Taylor series expansion are neglected. The error’s leading term, therefore, becomes proportional to ~(h3/3)(8CB/8y). From Eq. (12), we have CB = - Bu/By since avlax = O at the no-slip boundary. It follows that at the boundary, where y = 0, we obtain 3:: = -2c = on“. (84) From the above analysis, it is obvious the that the error in the expression of CB is ~ - 2 h uCL < 0. The injected vorticity is being overestimated by 2 h “c1.- For the second-order accurate condition (Woods, 1954), however, the third- order term in the Taylor series expansion is retained and the leading term of the error is proportional to -(h2/8)(32CB/8y2). On the boundary, we require that 83 2 33528 _.._. '6d = '6“CL° (85) The above approximation obviously underestimates the injected vorticity CB by 3/4 hzuCL since the error is ~ 3/4 h uCL > 0. Indeed the overshoot can be thought of as being physically incorrect since its presence cannot be substantiated by experimental data. Morihara and Chang (1973) claimed that overshoofing is necessary to satisfy the continuity equation. Other authors explained the phenomenon as a consequence of the accelerating inviscid core or the large velocity gradients near the inlet. Others attributed these bulges to the upstream diffusion of the vorticity which renders the use of a uniform inlet velocity as incorrect; a cascade inlet condition was suggested instead. All previous analytical analyses justifying the overshooting were either conducted for vanishing Reynolds numbers or used in association with the boundary-layer approximation. 6.2 Consistency and Errors of a Numerical Solution Consistency of a numerical solution is a measure of the extent to which the finite-difference equations approximate the partial differential equations. A finite- difference scheme is consistent if the difference between the partial differential equation and its difference representation vanishes under grid refinement. This difference is defined as the truncation error. The truncation error associated with any numerical approximation is the sum of all terms not included in the Taylor series expansions used in the discretization of the derivatives. The truncation error is considered acceptable if the difference approximation is consistent and stable. 84 The difference between the exact and numerical solutions of any partial differential equation is attributed to the error associated with the solution of the finite- difference scheme. This error is the sum of the discretization error of the finite- difference approximation and round-oflr error of the calculations. The discretization error is caused by replacing the continuous derivatives by a discrete representation. This is defined as the difference between the exact solution (even though an exact solution may not be available) of the partial differential equations and the solution to the difference equations. This definition includes the truncation error of the difference representation of the partial differential equations plus the error introduced form the finite-difference approximations of the boundary conditions. The round-off error results fi'om rounding the number of digits in the arithmetic operations. The round-off error is proportional to the number of arithmetic operations and may depend on other factors such as the number of nodal points of the computational domain; in such a case, while the truncation error may be reduced from grid refinement, the round-off error will increase. The round-off error also depends on the machine accuracy and the precision of the computational algorithm; these are related to the number of significant digits carried during the calculations. One additional source of error is the tolerance error imposed by convergence criteria for the termination of the computation. In all numerical solutions, certain preassigned error criteria are used to decide whether additional computation (iterations for steady problems or time steps for marching methods) is necessary. In this study, these criteria are represented by Eqs. (65) and (82). 85 6.3 Numerical Stability of the Differenced Navier-Stokes Equations Numerical stability is a subtle issue closely related to the convergence of the numerical solution. It is related to the error of the finite-difference approximation. A numerical scheme is considered stable if the errors do not grow as the solution marches in time. It is not easy, however, to formally establish whether a certain numerical scheme is stable or not. Restrictions limiting the size of the spatial increments and/or time step are often required to avoid instability. If these restrictions are not met, the numerical results exhibit oscillations that grow very rapidly, and after a few time steps their amplitude becomes infinite so that an "overflow" is registered by the computer; this phenomenon is characteristic of instability. The convergence of a numerical solution is understood to mean that the finite- difference solution tends to the exact solution of the boundary-value problem as the increments in space and the time step are reduced in size. Although a certain finite- difference approximation may be consistent, the solution must be stable so that it has a chance of converging. A proof of this principle applied to a well-posed, linear, initial- value problem is given by the Equivalence Theorem due to Lax (Richtmyer and Morton, 1967). Lax’s theorem states that when the consistency condition is satisfied, stability is the only necessary and sufficient condition for convergence. On the other hand, the theorem has never been proven true for a nonlinear partial differential equation such as the Navier-Stokes equation. The present numerical scheme requires the simultaneous solution of the vorticity and stream function equations. It is common knowledge, however, that 86 Poisson’s equation possesses excellent convergence properties when solved alone. Therefore, to address the stability problem of the numerical solution, it is reasonable to assume that the more likely source of any instability is the nonlinear vorticity equation. Each of Eqs. (52), (53), or (64), represents a general system of N linear algebraic equations in N unknowns. In matrix form, they are written as [allitl = [bl - (86) where [a] is the square coefficient matrix of order N xN, [q] is the solution column vector (containing the unknowns C or v), and [b] is the column matrix representing the known quantities on the right-hand side. In the case of the elliptic stream function equation, let us assume that the system of equations, given by Eq. (64), is irreducible (cannot be arranged such that some of the unknowns can be determined by solving less than n equations). If we further assume that [a] is nonsingular then the solution [\v] exists and is unique. By ordering the system of equations such that the coefficient of the unknown at each row (or equation) is largest in magnitude, then the leading diagonal elements of [a] become largest in magnitude. It turns out that when the elements aij of the coefficient matrix satisfy the condition i=N laiil 2 2 Iaijl (87) i=1 jti 87 for all i, in general, except at points adjacent to the boundaries where it applies to a ij not multiplying boundary values of v, then the solution converges as the matrix is assured to be diagonally dominant (V arga, 1962). The strict inequality (the greater than sign) must be true for at least one value of i and will normally be satisfied at points adjacent to boundaries. The interpretation of this sufficient condition is clear in light of the required diagonal dominance of the coefficient matrix. For each equation, it requires that the magnitude of the diagonal elements be greater than the sum of magnitudes of all other non-diagonal coefficients with the inequality applied to at least one equation usually near the boundary. The diagonal dominance imposed by the above condition is forcing the error, which is defined as the difference between the calculated value and the exact solution, to become smaller and smaller as the iteration is repeated cyclically. Since the above is only a sufficient condition, convergence may sometimes be attained even though the above condition is not met. A necessary condition shall not be stated since it is impractical to evaluate (Anderson et al., 1984). The above stability condition also holds for all individual equations resulting from Eq. (63) for the SOR by points method. The stability requirements for the system of parabolic linear difference equations, resulting from Eqs. (52) and (53), can be established from the von Neumann (or Fomier') analysis (Anderson, et al., 1984) for marching methods applied to a linearized system of equations. For stable finite-difference calculations, the maximum eigenvalue of the amplification matrix associated with the system has a magnitude less or equal to one. This leads to the requirement that Ilsa-955- $1 (88) where A.“ is the largest eigenvalue of the Jacobian matrix of the system. The above condition ensures that the error continue to be bounded as the solution progresses in time. It should be noted that the von Neumann stability analysis presented above does not include the effect of boundary conditions even though a matrix notation for the system is used. The influence of the boundary conditions is not easily included in the analysis. However, the matrix method can be used for one-dimensional linear equations to demonstrate the effect of boundary conditions on the stability (Anderson et al., 1984). In reality, the vorticity boundary condition puts restrictions on the time step, in contrast to the unconditional stability inferred from von Neumann’s condition (Roach, 1982). Situations can occur, practically at large Reynolds numbers, when the convergence condition is violated at some points. This has led to an alternative formulation in which the diagonal dominance is maintained by approximating the convective terms with schemes of lower-order accuracy. This leads to coefficients that satisfy Eq. (88) under all circumstances (Dennis and Chang, 1969). The higher- order accuracy can be preserved by the correction technique of Eqs. (56) and (57). A comprehensive analysis of stability, including other methods of analysis with many theorems and proofs, is contained in Richtmyer and Morton (1967). DISCUSSION OF RESULTS AND CONCLUSIONS The subject matter of this chapter is divided into two major topics. The results obtained from the present numerical study are presented. The flow problems, in addition to the role of the accelerating parameters, are discussed separately. The related figures are displayed at the end of the chapter. Based on these results, conclusions are then drawn. Convergence is attained by advancing the solution in time until it no longer varies significantly. An error criterion is used to decide whether additional time steps are necessary; a tolerance error e = 10" is selected as a condition for convergence. The transient solution is of no interest (time steps are fictitious for the steady solution). The present numerical solution is second-order accurate since both the numerical scheme and boundary conditions have second-order accuracy. 7.1 Entry Flow in a Plane Channel The starting point for this analysis is the consideration of the full Navier-Stokes equations in terms of the vorticity and Poisson’s equation for the stream function, describing the viscous, incompressible flow in the entry region of a two-dimensional 89 90 channel. The vorticity C, stream function \v, streamwise velocity u, and normal velocity v are calculated at each nodal point for a uniform inlet velocity profile and Reynolds numbers ranging between 10 and 5000. A uniform grid with a spacing size of 0.05 is used. The number of meshes in the x-direction is equal to Re/ (20 H). Run CPU times on our available computer ranged from 2.5 minutes for Re = 20 to approximately 12 hours for Re = 2000 using the full channel height and optimum values of the accelerating parameters. Due to symmetry, the number of nodal points of the computational domain is reduced by considering one half the height of the channel. Consequently, the required CPU time is reduced to 0.5 minutes for Re = 20 and nearly 2 hours for Re = 2000. The two regions making up the entry flow (as shown in Fig. 2) are identified as the inviscid-core region, in which a viscous layer is assumed to exist near the wall, and the profile-development region, in which viscous effects influence the entire channel flow. The inviscid-core region ends when the thickness of the viscous layer becomes equal to half of the channel height. This is determined numerically to occur when the velocity at the centerline becomes larger than the velocity at the first node below the centerline. The entrance length L,= is defined as the distance from the inlet to the point where the centerline velocity reaches 99% of the parabolic centerline velocity. These regions and their ratios are obtained for various Reynolds numbers and are presented in Table l. The results show that the inviscid-core region is approximately one-fifth of the entrance length and increases with a decrease in the Reynolds number below 100. This is somewhat shorter than one-fourth the entrance length reported by Mohanty and Asthana (1978) for flow in a circular pipe. No significant difference was noted 91 Table l. The inviscid-core length Li, the profile-development length L d, and the entrance length L a for various Reynolds numbers. Re _L_i_ .113. Le Li Ld Le Li H H H H-Re H-Re H-Re Le 10 0.15 0.35 0.50 0.0150 0.0350 0.0500 0.3000 20 0.25 0.60 0.85 0.0125 0.0300 0.0425 0.2941 50 0.50 1.60 2.10 0.0100 0.0320 0.0420 0.2381 100 0.80 3.25 4.05 0.0080 0.0325 0.0405 0.1975 200 1.55 6.65 8.20 0.0078 0.0332 0.0410 0.1899 500 4.10 16.50 20.60 0.0082 0.0330 0.0413 0.2000 1000 7.85 33.35 41.20 0.0079 0.0334 0.0412 0.1897 2000 15.55 66.70 82.25 0.0077 0.0334 0.041 1 0.1892 5000 39.10 167.40 206.50 0.0078 0.0335 0.0413 0.1893 when the entrance length was compared with that of other researchers as well as the well known empirical relationship: L,/H = 0.04 Re. Moreover, when L e is calculated using different inlet conditions, the entrance length was found to be marginally influenced by the inlet velocity distribution. A summary of L .3 values for a number of channel studies is shown in Table 2. The streamwise velocity profile undergoes a change from an assumed inlet profile (U = 1.0) to that of a fully-developed, parabolic profile at a location 92 Table 2. Comparison of the ratio L,/(H-Re) as reported in different channel studies for Reynolds number = 200. Analytical Schlichting (1934) 0.0400 Experimental Feliss et al. (1977) 0.0400 Numerical Bodoia and Osterle (1961) 0.0440 Roidt and Cess (1962) 0.0454 Hwang and Fan (1963) 0.0422 Gillis and Brandt (1964) 0.0456 Morihara and Cheng (197 3) 0.0452 Abdulla (1987) 0.0442 Present study 0.0410 with boundary-layer equations 0.0405 with cascade inlet condition 0.0415 with experimental inlet profile 0.0417 downstream with centerline velocity uCL = 1.5. For a uniform inlet velocity, the streamwise velocity profiles for various Reynolds numbers are shown in Fig. 8. Several of these profiles are enlarged in Fig. 9. They demonstrate that for x-locations in the inviscid-core region, the velocity profiles include a minimum on the axis and symmetrically located maxima on either side of the centerline. The maximum velocity in these profiles is at most 4.9% higher than that at the centerline. This is in contradiction to reported results (Gillis and Brandt, 1964; Brandt and Gillis, 1966; Morihara and Cheng, 1973) in which these local maxima are much more pronounced over larger downstream distances and at low Reynolds numbers; the overshoot diminishes for Reynolds numbers > 200. The consistency in the velocity 93 profiles over the range of Reynolds numbers considered here suggests that the overshoot is a result of the numerical approximations. By solving the full Navier- Stokes equations, it is found that these local maxima are essentially nonexistent. This point is also evident in the work by Wang and Longwell (1964) who used a finite-difference formulation that is less accurate in the representation of the vorticity boundary condition than that used in this study. Their boundary condition (Thom and Apelt, 1961) is only first-order accurate thereby overestimating the boundary vorticity that leads to the overshoot; our method retains the second-order terms in the Taylor series expansion used to approximate the boundary vorticity. In addition, it can be shown that in the case of Li and Ludford (1980), the use of boundary-layer equations near the entrance leads to an incorrect representation of the initial vorticity. These results are supported by the numerical findings of Abdulla (1987). The effects of several factors on the develoth of the streamwise velocity profiles are investigated. The profiles obtained from the use of boundary-layer equations are shown in Fig. 10 for Re = 200. The profiles due to cascade inlet condition for Re = 200 are displayed in Fig. 11. An experimental profile for the inlet velocity, obtained by fitting a polynomial near the wall to a uniform central section using measurements from a hot-wire anemometer (Backus, 1989), was also used. The streamwise velocity distributions along the channel for this experimental profile at the inlet are shown in Fig. 12 for Re = 200. The difference due to the above three flow conditions, as compared to Fig. 8e, is clearly insignificant. The streamwise velocity as a function of the location across the channel is shown in Fig. 13 for Re = 200 and 2000. Also, the streamwise velocity at the centerline is compared in Fig. 14 for Re = 20, 50, 100, 200, and 2000. The shape of the 94 centerline velocity is identical to that of other studies where, with the increase of the Reynolds number, higher velocities are reached at shorter distances from the inlet. Profiles of normal velocity components along the channel are shown in Fig. 15 for Re = 200, 1000, and 2000. The changes in v are most apparent at distances near the inlet. Although not completely fully-developed, the flow in the profile-development region becomes parallel at a short distance after the inviscid—core region. Profiles of the vorticity along the channel for Re = 200, 1000, and 2000 are shown in Fig. 16. Figure 17 shows the values of the vorticity at the upper boundary forRe-20,200,and2000. Thelasttwofiguresindicatethatthevalueofboundary vorticity levels out to reach the fully-developed value (with magnitude of 6.0) as the flow progresses downstream; the vorticity on the lower boundary is equal but opposite in sign. The changes, as functions of x/Re, are more rapid for higher Reynolds numbers. A comparison of boundary vorticity using different formulas is presented in Fig. 18. It further supports, as indicated in Section 6.1, the hypothesis that the second-order Woods’ condition underestimates the boundary vorticity in contrast to the first-order Thom’s condition. Vorticity contours and streamlines of the flow are displayed in Figs. 19 and 20, respectively. It is clear that most changes take place in the immediate vicinity of the inlet section. The normalized pressure gradients along the channel at the wall and centerline are shown in Figs. 21, 22, and 23, for different Reynolds numbers. Both pressure gradients converge to the constant value (- Re/ 12) of a fully-developed flow. The details of the computational procedure for determining the proper time steps and mesh sizes must be given some special attention. Concerning the procedure for 95 determining the mesh sizes, there are two problems involved. One is to determine whether the size of the computational grid is large enough to impose the downstream boundary condition and match the required computations with the available computer capabilities. The other is to determine whether the chosen mesh sizes are small enough for accurate results. These two problems are obviously interrelated. Conversely, the size of the time step is important in determining the computation time and the numerical stability of the solution. To check whether the solution is independent of the mesh size, mesh sizes in the range of 0.02 to 0.1 are used at selected Reynolds numbers. The effect of the mesh size on the accuracy of the results is found to be minimal; for example, when compared to a grid size of 0.025, our standard grid (0.05) produces a maximum difference of 4.6% in the values of the streamwise velocity. The variation of the streamwise velocity with the mesh size at selected locations is shown in Fig. 24 for constant At and h/ At. The relationship between u and h2 is nearly linear, therefore, confirming the second-order accuracy of the solution in the space variables. 7.2 Channel Flow with a Finite Step Channel flow in the presence of a fixed finite step on the lower boundary is identified with separation. Two distinct separated regions with pronounced eddies of recirculating fluid are developed on both sides of the step. The size of these separated regions depends on the Reynolds number, the height (B), and length (D) of the step; the separated region upstream of the step is considerably smaller in size thanthataftofthestep. AchannelwithauniformgridandspacingofsizeOflS is used in the numerical solutions for Reynolds numbers 20 and 200. A finite step of variable height and length is placed at different locations Xl along the channel, as 96 shown in Fig. 5. A summary of the cases considered is presented in Table 3. The effect of the length of the step on the separation is demonstrated in Fig. 25 for Re = 20. Steps, located at X1 = 1.0, with length D = 0.0, 0.05, 0.1, 0.25, and 0.3 andheightB=0.25arestudied. ThestepoflengthD=0.0isaspecialcasewhere the step is represented by the grid line between two adjacent meshes. Table 3. Length of the aft separated region L2 for the channel flow with a finite step. Re Xl B D L2 20 1.00 0.20 0.00 0.368 1.00 0. 25 0. 00 0.441 1.00 0. 25 0. 05 0.468 1.00 0. 25 0.10 0.456 1.00 0. 25 0. 25 0.431 0. 85 0.25 0. 30 0.412 1.00 0.10 0. 05 0.190 1.00 0.20 0.05 0.386 1.00 0.30 0.05 0.552 1.00 0.40 0.05 0.663 1.00 0.50 0.05 0.821 200 1.00 0.20 0.05 1.794 1.00 0.25 0.05 2.175 2.00 0.20 0.05 1.733 2.00 0.20 0.25 1.506 2.00 0.20 1.00 1.116 2.00 0.25 0. 05 2. 309 8.00 0.20 0.05 1.702 For the cases 97 considered,thelengthLloftheseparatedregioninfrontofthestepappearstobe uninfluenced by an increase in D while the length L2 of the separated region aft of the step is slightly decreased with an increase in D. Conversely, the role of the height of the step in determining the size of the separated region is shown in Fig. 26 for Re = 20. The step’s height B varies between 0.1 and 0.5 with a constant length D = 0.05. Increasing the height of the step causes the size of the aft separated region to increase, as expected. In all of the above cases, the length of the front separated region is observed to be somewhat fixed with value L1 ~ 0.10. TheseparationduetoastepofheightB=0.2andRe=200isshowninFig. 27 for steps of length D = 0.05, 0.25, and 1.0 located at X1 = 2.0. The separated region becomes smaller with the increase in D. For Re = 200, Fig. 28 shows an increase in thesizeoftheseparatedregionduetoanincreaseintheheightofthesteme= 0.25; compared to Fig. 27a, the length of the separated region in the aft of the step is nearly doubled. Figure 29 shows the separation due to a step placed further upstream (at X1 = 1.0) for Re = 200, D = 0.05, and B = 0.2 and 0.25, respectively. The aft separated region is again larger for B = 0.25. Evidently, the size of the separated region is unaffected by the streamwise location of the step. This result is also confirmed by Fig. 30 where the step is placed at X1 = 8.0 where the flow is fully-developed; the size of the separated region remains the same regardless of the location of the step. Velocity vectors for the flow in the presence of a step with B = 0.2, D = 0.05, and Re a 20 are shown in Fig. 31. The velocity gradients are high in the vicinity of the step as exhibited by the direction of the vectors. Streamwise velocity profiles for 98 several combinations of height B and length D are shown in Fig. 32 for Re = 20 and X1 = 1.0, and in Figs. 33 and 34 for Re = 200 and X1 = 2.0 and 1.0, respectively. The above figures reveal that the presence of a step has an upstream effect on the shape of the velocity profile and that the velocity distribution becomes parabolic at a sufficient distance downstream. Inaddition,acomparisonofthecasesstudiedshowsthattheincreaseinthe Reynolds number is accompanied by an increase in the length of the separated region aft of the step. The size of the separated region upstream of the step, however, seems to be relatively independent of the Reynolds number. The vorticity contours, shown in Fig. 35, emphasize that the changes in the flow field are most pronounced in the vicinity around the step. 7.3 Channel Flow with a Moving Boundary For this flow, a section of the lower boundary is allowed to move with a constant velocity in a direction the same as or opposite to the entry flow. The channel studied here has a grid with a uniform spacing of size 0.05. The vorticity C, stream function \y, and streamwise velocity u, are calculated for flows with boundary velocity UB ranging from -10.0 to 5.0 and Reynolds numbers between 20 and 1000. The moving boundary is defined to exist between x = XI and x = X2 (as shown in Fig. 1). Results demonstrated by plots of streamlines, velocity profiles, and vorticity contours, show the formation of a separated region when the boundary section is moving in the upstream direction. In general, the size of the separated region depends on the Reynolds number and boundary velocity. Streamlines of the flow for 99 Re a 20 with the lower boundary section (between X1 = 1.0 and X2 = 2.0) moving opposite to the inlet velocity at UB = -10, -5, -2, and -1, are shown in Fig. 36. The separated region becomes larger with an increase in the magnitude of boundary velocity. Velocity vectors for the case Us = -10 are shown in Fig. 37 indicating the recirculating pattern of the flow inside the large separated region. Figure 38 shows the flow streamlines when the boundary section is moving in the direction of the inlet velocity at UB = 1, 2, and 5 for Re = 20. Observe that no separation occurs since the boundary is moving in the direction of the flow. The fluid in the near vicinity of the moving boundary section is accelerated due to the motion of the boundary. The above observations are revealed in Figs. 39 and 40 for Re = 200. Streamlines of the flow with boundary velocity UB = -2.0, -1.0, -0.5 and 1.0 are displayed. The influence of the Reynolds number on the size of the separated region is demonstrated by comparing Fig. 36 and Fig. 39 for U}, = -1.0 and -2.0. It is clear thatthesizeoftheseparatedregiondecreasesasReincreases. Thenumberof iterations needed to attain convergence to a steady-state solution increases with an increase in Re but decreases with a decrease in the magnitude of U3. The size of the separated region is also related to the length and location of the moving boundary section. As shown in Fig. 41, the separated region extends along the length of the moving boundary with no apparent increase in its height due to the increase in the length of the moving section of the boundary. Figures 42 and 43 show that placing the moving section of the boundary further downstream (between X1 = 4 and X2 = 5) produces no significant change in the size of the separated region; in this case, however, a larger number of iterations is required due to the extended 1w computational domain. Further examples regarding the effects of the length and position of the moving boundary section are shown in Figs. 44 and 45, for Re = 200 and Us 2 -1.0, -0.5, and 1.0. In addition, placing the moving boundary section in the fully-developed region of the flow showed no noticeable difference in the streamlines of the flow or the shape of the separated region, as shown in Fig. 46. For Re a 1000, streamlines are shown in fig. 47 when the lower boundary section, positioned between X1 = 10 and X2 = 14, is moving opposite to the inlet velocity at UB = -0.5; at Re = 1000 the separated region is reduced in size since it is apparently suppressed by the relatively high velocity incoming flow. Flow streamlines for the Couette-Poiseuille flow, when the entire lower boundary is moving in the direction ofthe flow at UB =1.0, for Re = 500 are shown in Fig. 48. Development of streamwise velocity profiles along the channel are also shown in Figs. 49 and 50 for UB = -1.0 , -0.5, and 1.0 and Re = 200, and in Fig. 51 for UB = -0.5 and Re = 1000. The velocity at the boundary between X1 and X2 is equal to the boundary velocity as required by the no-slip boundary condition. The parabolic velocity profile is attained at some distance beyond the influence of the moving boundary. Velocity profiles for the Couette-Poiseuille flow case are presented in Fig. 52 for Re = 200 and 500. Vorticity contours with and without the presence of a separated region are shown in Figs. 53 and 54 for Re = 20 and Re == 200, respectively. Evidently, changes in the vorticity are concentrated in the immediate vicinity of the moving boundary section. For the Couette-Poiseuille flow case, vorticity contours are shown in Fig. 55, as well. 101 7.4 Channel Flow with an Oscillating Boundary An extension of the previous problem is the introduction of an oscillatory motion to the moving section of the lower boundary, which it is permitted to oscillate sinusoidally in its own plane. The vorticity C, stream function w, and velocity component u are calculated at each nodal point for a flow with a uniform inlet velocity profile and boundary velocity U3 = A Cos (t) where A is the amplitude of oscillation of boundary velocity. A mesh size of 0.05 is used with Reynolds numbers of 20, 200, and 1000. Again, the oscillating boundary is defined to exist between x = XI and x = X2 (as shown in Fig. 1). As predicted, plots of streamlines, as well as velocity profiles and vorticity contours, demonstrate the existence of a separated region generated in the vicinity of the oscillating boundary section during its upstream motion. Generally, the size of the separated region depends on the Reynolds number, Strouhal number, the amplitude of boundary velocity, and time; its size is a maximum at t = (2n+1)rt where n is an integer representing the number of cycles of oscillation before the solution converges. Streamlines for the flow with Re = 20, St = 1.0, and A = 1.0 are shown in Fig. 56 at time t = 2m: and (2n+1)lt. For Re = 200, and different St and A, streamline patterns are shown in Figs 57 and 58 at t = 2n1c and (2n+1)x. Note that no separated region appears at the instant t = 2nrt since the boundary is moving with the flow, as in Fig. 56a, and that the formation of a separated region is clear at t = (2n+ 1):: when the boundary is moving at maximum velocity in the direction opposite to the flow, as in Fig. 56b. The influence of the oscillating boundary on the progression of the flow is further illustrated in Fig. 59 where a full cycle of oscillation we is shown; for this example, Re = 200, St = 1.0, and A = 1.0. The size of the separated region increases with time but starts to decrease for t > (2n+l)1t. At t = (2n+l)rt, the height of the separated region is equal to 0.098. As the velocity of the boundary becomes > 0, the separated region ceases to exist. The influence of the Strouhal number on the size of the separated region is demonstrated by Figs. 60-62 for St = 0.5, 1.0, and 2.0, Re = 200, and various A and length of the oscillating boundary. In all three cases, a separated region exists since themotionoftheoscillatingboundaryisoppositetotheflowatthatpointintime(t= (2n+l)rt). Clearly, the size of the separated region is inversely proportional with St which is directly related to the fiequency of oscillation. The effect of amplitude of oscillation A of boundary velocity is shown in Figs. 63—66 for A between 0.5 and 2.0 with different Re, St, and length of the oscillating boundary. Increasing the amplitude of oscillation causes the separated region to increase in size. Figure 67 displays stream function contours of the flow at t = (2n+-1): for Re = 1000 and different St and A. The separated region barely exists since it is apparently swept away by the relatively high velocity incoming flow. The effect of the Reynolds number on the size of the separated region can be established from the previous figures, in particular, Figs. 56b, 60b, and 67a, for Re = 20, 200, and 1000, respectively, where for all three flows St = 1.0, A = 1.0, X1 = 1.0 and X2 = 2.0. It is evident that, for fixed St and A, the size of the separated region decreases as Re increases. Comparing Figs. 63a and 60c for St = 2.0, A = 1.0, X1 = 1.0, X2 = 2.0, and Re = 20 and 200, respectively, also leads to the same conclusion. The number of iteration cycles it necessary to obtain convergence increases with an increase in Re and St. 103 In addition, the size of the separated region may depend on the location and length of the oscillating boundary section. Placing the oscillating section of the boundary further downstream, as shown in Fig. 68, has no significant effect on the size of the separated region (as compared to Fig. 60b); in this case, however, a larger number of iteration cycles is required due to the extended computational domain. On the other hand, Fig. 69 shows that the separated region extends along the length of the oscillating boundary with no apparent increase in its height due to the increase in the length of the oscillating section of the boundary. The streamlines of the flows shown in Fig. 70 combine the effects of both location and length. The oscillating boundary section is placed between X1 = 8.0 and X2 = 10.0; this is equivalent to a parabolic inlet profile. When compared to Figs. 61b and 69, for the flow with Re = 200, St = 1.0 and A = 1.0, Fig. 70b illustrates that placing the oscillating boundary section further downstream has no influence on the separated region. Moreover, from Figs. 65a and 70c it is clear that the above is true for the case Re = 200, St = 2.0, and A = 0.5. Developments of streamwise velocity profiles along the channel, for several flow conditions, are also shown in Fig. 71 at t = 2th and in Fig. 72 at t = (2n+l)rt. Variations in the velocity profiles due to the presence of the oscillating boundary appear to be local. The velocity at the boundary between X1 and X2 is equal to the boundary velocity as required by the no-slip boundary condition. A parabolic velocity profile prevails at the downstream end. Vorticity contours for the flow with St = 1.0, A = 1.0 XI: 1.0, and X2 = 2.0 are shown in Figs. 73 and 74 for Re = 20 and 200, respectively; these contours are at t = 104 (2n+1)rt with UB = -1.0 and t = 2n: with UB = 1.0. Vorticity contours are also shown in Fig. 75 for the flow with the oscillating boundary placed in the fully- developed region. The above figures demonstrate that changes in the vorticity are substantial only in the near vicinity of the oscillating section of the boundary. It is interesting to notice that the vorticity at the centerline is almost zero, which suggests that the oscillating boundary has a negligible effect on the flow in the upper half of the channel as long as the distance between the lower and upper boundaries is large relative to the height of the separated region. 7.5 Accelerating Parameters For the present numerical solution, the rate of convergence of the difference equations is significantly enhanced through the use of accelerating parameters. Optimum values of the time step (At) for the vorticity equation, over-relaxation factor (F8) for Poisson’s equation, and the optimum values of the weighting factors (a and y) for the vorticity and stream function, respectively, are presented below. In order to determine the optimum values of the accelerating parameters, numerical experimentation with a large number of combinations of these parameters is attempted for a channel with a uniform mesh of size 0.05. First, optimum values for At are obtained using the optimum FS given by Eq. (61). The optimum values of F8 are then refined by trying F8 in the range 1.0 to 1.9; FS = 1.0 corresponds to the Gauss-Seidel iterative scheme. Weighting factors for the vorticity and stream function are then applied, as prescribed by Eqs. (80) and (81), respectively. Values of a and y are assumed to vary between 0.0 and 0.9. H5 Optimum values of the accelerating parameters as functions of the Reynolds number, based on the numerical solution over one half the height of the channel, are given in Table 4 for Re = 20, 50, 100, 200, 500, 1000, and 2000. The number of outer iterations and computing time for the accelerated solution along with the number of iterations for the Gauss-Seidel solution, considered to be the unaccelerated case (F S = 1.0, o = 0.0, and y = 0.0), are also given in Table 4. The same information is Table 4. Optimum accelerating parameters, CPU time, and number of outer iterations for the channel flow with a uniform mesh of size 0.05 at various Reynolds numbers; the solution is based on one half the height of the channel. rmmMrMFPmmurd' Re At FS 0 7 CPU time am Wed 20 0.03 1.70 0.00 0.05 38.5 58 l 15 50 0.09 1.68 0.00 0.05 44.9 54 147 100 0.16 1.67 0.00 0.05 1:267 55 160 200 0.33 1.65 0.00 0.05 3:084 57 170 500 0.63 1.64 0.00 0.10 9:415 74 202 1000 0.67 1.64 0.00 0.15 31:59.9 122 228 2000 0.67 1.64 0.00 0.25 2:10:02.1 247 "' l.(TUfimekgwmmhmmee 1L DwanwmmnammmmkuwmwdhwqfimmnmakmmmpmmMms 3. TheunacceleratedcaserefasmflleGauss-SeidelitemdcnwidropfimumALF821.0.6 a 0.0, andyc 0.0. t ‘hflhmxmadnamfiaubwnmumwnpmemmnmmmmm 1% Table 5. Optimum accelerating parameters, CPU time, and number of outer iterations for the channel flow with a uniform mesh of size 0.05 at various Reynolds numbers; the solution is based on the full height of the channel. Number of Number of Re At FS 0 7 CPU time Iterations; Iterations; accelerated umccelerated 20 0.01 1.78 0.00 0.05 2:05.9 113 145 200 0.09 1.80 0.00 0.05 9:05.8 92 214 1000 0.24 1.80 0.00 0.10 2:18:29.0 266 406 2000 0.24 1.80 0.00 0.20 11:51:08.0 687 913 Notes: 1. CPUtirneisgiveninhrzminzsec. 2. Theaccelaatedeasehlcltxlefllemeofflleopfimumaccelerafingpamneters. 3. ThemacceleratedeasenfaswdreGamSeiddiwrafimwithopfimumALFS=lno - 0.0, and y :- 0.0. presented in Table 5 for Re 2 20, 200, 1000, and 2000 for the solution based on the full height of the channel. It is found that the optimum value of At increases with an increase of Re until it achieves a constant value at Re > 500. For the solution of one-half the height of the channel, this constant value is equal to 0.67, whereas for the full height of the channel it is equal to 0.24. For all Reynolds numbers, the number of iterations required for convergence initially decreases as At increases until the optimum value of At is attained. Increasing At above the optimum value causes the number of iterations, and hence the time of computation, to increase considerably. The solution, however, 107 becomes divergent as At continues to increase beyond a certain critical value. The numerical results also demonstrate that the optimum value of FS is related to the number of nodal points MxN of the computational domain, which is directly proportional to the Reynolds number and mesh size. The computing time decreases as FS increases from 1.0 until a minimum number of outer iterations is reached at the optimum value of FS. Continued increases in the value of FS result in an increase in the number of iterations. It is clear that the optimum value of FS decreases as the Reynolds number increases up to Re = 500. However, for higher Reynolds numbers the optimum value of FS approaches constant values equal to 1.64 and 1.8, respectively, for solutions based on one half and the full height of the channel. It is also noted that the results for Re < 500 converge most rapidly when the value of FS is slightly higher than that given by the analytical expression of Eq. (61); this value is approached asymptotically at higher Reynolds numbers. The weighting factors 0 and 7 also have a significant effect on the convergence of the numerical solution. Again, they depend on the total number of nodal points of the computational domain. The computing time decreases as 0‘ decreases to zero for all Reynolds numbers. This result can be explained as being the consequence of the refinement of the optimum value of At; an optimum At can produce an optimum convergence rate of the vorticity equation, thus implying that the rate of convergence of successive iterations needs not be slowed down. Unlike a, a minimum solution speed is obtained for values of 7 above zero. The optimum value of 7 has a constant value of 0.05 for the lowest Reynolds numbers and increases steadily for Re > 200 to reach a value as high as 0.25 for Re = 2000. Increasing the value of 7 above the 108 optimum value increased the number of iterations. With the use of the optimum values of the accelerating parameters, the required number of outer iterations is reduced approximately by a factor of 2 to 3 as compared to that required by the unaccelerated case. This is evident from the number of iterations displayed in Tables 4 and 5. Furthermore, for the range of Reynolds numbers considered, the time of computation for this ADI/SOR scheme is better than 2.5 times less than that required by the iterative solution employed by Abdulla (1987). Finally, the numerical results show that the optimum values of the accelerating parameters which are used to reduce the computing time for the plane channel flow, also represent the optimum values for the channel flow with a step or a moving boundary. 7.6 Conclusiom In closing, several conclusions are supported by the results of this numerical study. These are summarized below. 1. For the two-dimensional, viscous, incompressible entry flow, the entrance length was found to be compatible with known experimental results, i.e., L e = 0.04 Re. The length of the inviscid-core region is approximately one-fifth of the entrance length. The influence of a cascade inlet condition or an experimental inlet velocity profile (with velocity gradients near the walls) on the development of streamwise velocity profiles is insignificant. The local maxima resulting from the overshoot in the streamwise velocity profile near the walls, as predicted by other researchers, are essentially negligible. 109 For the flow in a channel constricted with a finite step, separation occurs on both sides of the step. The size of the upstream separated region is smaller and does not depends on the Reynolds number. The size of the aft separated region is directly influenced by the Reynolds number and the height of the step but is unaffected by the length of the step. Also, the size of these separated regions is independent of the streamwise location of the step. For the channel flow with a boundary section moving at a constant velocity, the upstream motion of the boundary clearly generates a separated region. The flow in the vicinity of the moving boundary is accelerated when the boundary is moving in the direction of the flow. The size of the separated region is a function of the Reynolds number and boundary velocity. It decreases as Re increases but increases with the magnitude of velocity Us of the moving boundary section. The increase in Re and the magnitude of UB causes the number of iterations needed to attain convergence to a steady-state solution to increase. Changes are concentrated in the immediate vicinity of the moving boundary. For the channel flow with an oscillating boundary section, the formation of a separated region is clear when the boundary is moving opposite to the flow; no separated region appears when the boundary is moving in the direction of the flow. The size of the separated region is inversely proportional to Re and St but directly proportional to the amplitude of boundary oscillation. The number of iteration cycles (i.e., time) needed to attain convergence increases with the increase in Re and St. The effect of the streamwise location of the oscillating boundary is insignificant. 110 The incorporation of accelerating parameters extends the domain of convergence and improves the stability of the numerical solution, as well as minimizes the time of computation by as much as a factor of 3 for certain Reynolds numbers. Their optimal values are functions of the number of nodal points in the computational domain; for each Reynolds number and mesh size there is an optimum combination of these parameters. The optimum values of the four accelerating parameters determined in this study for a plane channel flow, are also applicable for a channel flow with a step or a moving boundary section; these parameters should serve as a guide to reduce the computing time for other flow situations, which use this system of equations and have the same computational domain. The ADI/SOR scheme, utilizing optimum accelerating parameters, is numerically stable for high Reynolds numbers with the mesh size considered and appears to be a suitable technique for treating flow problems that contain time-dependent boundary conditions. It also provides the advantage of being more rapidly convergent than the usual iterative schemes. Results obtained using a finer mesh produce a maximum difference of 4.9% in the values of the streamwise velocity. 111 1.00 . 0.75: y 0.50; 0.25-1 0.00 l , . , . T . , . , . 0.0 0.1 0.2 0.3 0.4 0.5 0.6 x a) Re=10. 1.m m V I 0.75-4 y 0350'1 0.25- anol « 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 x b) Re=20. 1.00‘ N 0.75- y 0.50“ 0.25~ 0.00 / 0.0 1.0 2.0 ISIVO 4.0 50. 60. W70 8.0 10.0 11.0 12.0 x c) Re=200. Figure 8. Development of streamwise velocity profiles through the entry region of a two-dimensional channel at various Reynolds numbers. 112 1.00 m '1 0.75-: y0.50j 0.25-j .,.,.1r1r,.T.r1j., 0.00 0.0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 2.7 3.0 x d) Re=100. 1.00 '\ 1 I r 0.75- y 0.50-J 0.25« 0.00 / , efi . , . 0.0 0.5 .5 W25? 3? 3.5 4.04.5 5.0 5.5 6.0 x e) Re=200. r 1.00. N V l 075* y 0.50: 0.25; 0.00‘ . r i 0.0 2.5 #570 7.5 10.0 12.5 15.0 17.5 20.0 22.5 25.0 27.5 30.0 X f) Re=500. Figure 8. Development of streamwise velocity profiles through the entry region of a two-dimensional channel at various Reynolds numbers. 113 1.00 .\ 1.1 0.75; y 0.50; 0.25-1 'TfiT' 'rrrfiflrt'lrfi1 0.00 fi/ 1 I 0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0 40.0 45.0 50.0 55.0 50.0 x g) Re=1000. 1.00 .\ r1 e 0.75: )1 0.50“: 0.25- 0.00. / 0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0100.0110.0120.0 x h) Re=2000. 1.00 ~\ 0.75: y 0.50“: 0.25: 0.001 / 0.0 25.0 50.0 75.0100012501500175020002250250027503000 X i) Re = 5000. Figure 8. Development of streamwise velocity profiles though the entry region of a two-dimensional channel at various Reynolds numbers. 114 .G-OR.=20,x=O.1 .MRO=200,x=U.1 HR.-2W,x-U.5 HR.-200.x-1.0 G-EIROSZCKKJ,x=5.D T I V T 'H ,, .. .ffifit fir i 0.00 0.25 0.50 0.75 1.00 1 .25 1 .50 Figure 9. Streamwise velocity profiles at locations near the inlet section for various Reynolds numbers. F1 115 IoOOJ \ 0.75e y 0.50‘ 0.25.. . M T'l‘l'l'l‘l'l' 4.0 5.0 6.0 7.0 8.0 0.00 v v 0.0 1.0 2.0 3.0 >> 1 I Y I ' fi l r 9.0 10.0 11.0 12.0 X Figure 10. Development of streamwise velocity profiles through the entry region of the channel using the boundary-layer equations at Re = 200. 1 .00 v 0.75; y 0.50-1 0.25: CHOW—F": -1.0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 X Figure 11. Development of streamwise velocity profiles through the entry region of the channel using a cascade inlet condition at Re = 200. 171'11I' 4050 1.00 0.75j y 0.501 0.25; 0.001 . I . r I j’ I ' I t T ' I I ' 0.0 1.0 2.0 3.0 . . 6.0 7.0 8.0 9.0 10.0 11.0 12.0 X Figure 12. Development of streamwise velocity profiles through the entry region of the channel using an actual experimental inlet velocity profile at Re = 200. 116 1.5tttrIrtvII1v1_1_ITIII:Jxfi 1.2- 0.9— 3 d 3 1 0.6A 0.1% 00.. ,--..,.f-fi. ffifi--. 0 00 O 01 O 02 0.05 0.04 0 05 x/Re a) Re=200. 1’5 rTffil"'fffi'*'1 w 1.2“ d 0.9— 3 . 3 1 0.6- 0.3-4 0.0 ....,....T....,...-1.... 0.00 0.01 0.02 0.03 0.04 0.05 x/Re b) Re=2000. Figure 13. Streamwise velocity at different y-locations across the channel for Re = 200 and 2000. 117 x/Re Figure 14. Centerline velocity along the channel for various Reynolds numbers. 118 1.00 0.75-1 y 0.50-1 0.25: 0.00. v I T 0 0.1 0. T r ' l ' I I 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 .0 a) Re=200. 1.00 035‘ l ‘ y 0.5m 0.254 0.00 I ' l ' l 0.0 1 .0 2.0 3.0 5.0 6.0 7.0 8.0 9.0 1 0.0 *4 o b) Re=1000. 1.00 ’ 7’ 0.75; ‘ y 0.503 0.25: l L | 0.000 r16.o'26.o'3or.o' r fr ' 0. 40.0 50.0 60.0 70.0 80.0 90.0 100.0 X c) Re = 2000. Figure 15. Development of normal velocity profiles through the entry region of a two-dimensional channel at various Reynolds numbers. 119 1°00 ' I I l ' l ' l I r I I 7 I ' I I l .1 0.75- y 0.50“ I . 0.25-j D-fii ' I r I I F V I 1 I 1 I I I r I I I I 0.1 0.2 0.3 f 0.0 0.4 0.5 0.6 0.7 0.8 0.9 1 .0 X a) Re = 200. 1.m T I Y I I I T I 1 1 I I I I I I I I I 0.75- y 0.50~ 0.215q 0.“) T 1 1 I f r T I I I I I I I r I U I T I 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 1 0.0 x b) Re=1000. 1-00'T'I'r'rTIv1rII. r 0.75-j y 0.501 0.25: 0.001 I T . I I I I I ' f I I I I I r ' I 0.0 1 0.0 20. 0 30.0 40 .0 50.0 60.0 70.0 80.0 90.0 100.0 X c) Re = 2000. Figure 16. Development of vorticity profiles through the entry region of a two- dimensional channel at various Reynolds numbers. 120 35.0 0 Re - 20 A Re - 200 30.0 o D R. = 2000 25.0 20.0 15.0 10.0 5.0~ 0.0 T W T T—I v I I v I I I t r I I I T wiI T T T 1 0.00 0.01 0.02 0.03 0.04 0.05 x/Re Figure 17. Vorticity at the upper boundary along the channel at various Reynolds numbers. 35.0 are Woods' “ H Jenson! 300.1 H Thom'l 1 25.0— 20.0“ c, . 15.04 10.0- 5.0- 0.0 I j T rfi I r I I W T r f I I I I I T I I V I 0.00 0.00 0.00 0.01 0.01 0.01 x/Re Figure 18. Vorticity at the upper boundary along the channel calculated from different formulas for Re = 200. 121 I IjleIIIIIIIIIIFIITIIIIIfiITIIITiI - n—q - ‘ -q fl q 2. f... 3' fl c—I q q — q — 1 llllllllllllllllLllLlllllllLJJlll a)Re=200. a) Re = 2000. Figure 19. Vorticity contours of the flow in the entry region of a two-dimensional channel at Re = 200. Figure 20. Streamlines of the flow in the entry region of a two-dimensional channel at Re = 200. 122 35.0 4 o O R. - 20 A Re - 200 30°C? c: R: = 2000 =3 25.0: A é 20.o-« . \ J 3 15.0- 53 10.0- \ J . cg 5.0-r T 1. faint-IIIIIIII-I'I-.--qu.nu.luau-ul-ut 0.0 canon. o n 0 —5.0 —10-0 I ‘% I T I I I I I T I I T I T I I I T I T 0.00 0.01 0.02 0.03 0.04 x / Re Figure 21. Normalized pressure gradients at the wall along the channel for various Reynolds numbers. 12.5 3 0 R0 - 20 . A Re = 200 a U R. = 2000 8 10.0-‘ 3 3 5 1 3? 7.5: I v I1 \ 4 o. 4 . 3 5.0-j I"\ '1 g (1:. 1% A E 2.5: ”fig“: - .0. ' o o . - . . g : "lIIII-Iu-ollI:-I.Ial.-Po-lu-Ouloll .1 I 0.0: -2.5 1} I. T T r I r 1 I I r fi I I I I 0.00 0.01 0.02 0.03 0.04- x/Re Figure 22. Normalized pressure gradients at the centerline along the channel for various Reynolds numbers. 123 30.0 c I 0 Vol I A Centerline x 20.0— U . \ . a d U .l RT 10.0—g \ j ‘ Q) a 9:2 ‘ o “a I 4 1“...."“‘0llIt.no.0.0000000000000000 0.0— .o -10.0 .2 1 I I I I I I I I f I I I l' I I I I I 0.00 0.01 0.02 0.03 0.04 x/Re Figure 23. Normalized pressure gradients at the wall and centerline along the channel for Re = 200. 124 HOW) 8 ‘0 (7| J [1 I o x=D.2, y=0.2, h/M=I d = 1.080 El x-o.z, y-0.2, At 0.05 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 0.000 0.002 0.004 0.605 0.008 0.010 a) x = 0.2, y = 0.2. 1.445 1 11 —-L A 4.. N l U D lllLllllllllJJlJllllLll DO 1 .439 U(x.y) 2?: OD 1 .433 o x=D.5, y=0.5, h/M=l 1.430 E’fi’i'Pi' 2"9‘5’: 933.035. 0.000 0.002 0004' ' ' ' 0606 0.008 0.010 hZ I IIIII'IIIIII'TIIT‘FIII b) x = 0.5, y = 0.5. Figure 24. Streamwise velocity as a function of the mesh size at selected locations; Re = 200. 125 a) D = 0.0. b) D = 0.05. c) D = 0.1. Figure 25. Flow separation due to a finite step of various width; X1 = 1.0, B = 0.25, Re = 20. 176 d) D = 0.25. e) D = 0.30, (x1 = 0.85). Figure 25. Flow separation due to a finite step of various width; X1 = 1.0, B = 0.25, Re = 20. a) B =0.1. b) B = 0.2. c) B = 0.3. Figure 26. Flow separation due to a finite step of various height; X1 = 1.0, D = 0.05, Re = 20. a) D = 0.05. b) D = 0.25. c) D = 1.0. Figure 27. Flow separation due to a finite step of various width; X1 = 2.0, B = 0.20, Re = 200. 130 Figure28. Flow separation due to a finite step; Re = 200, X1: 2.0, B = 0.25, D = 0.05. a) B = 0.20. b) B = 0.25. Figure 29. Flow separation due to a finite step of various height; X1 = 1.0, D = 0.05, Re = 200. l3] Figure 30. Effect of step location on the flow separation due to a finite step; X 1 = 8.0, B = 0.20, D = 0.05, Re = 200. ---—---------_---_-----— ------------- —.....—.— .—.———._._. “.4“... _._-_._._.. —.—_———_—__—..._._.— .4-.- — ...................... Figure 31. Velocity vectors of the flow in the presence of flow separation due to a finite step; X1 = 1.0, B = 0.20, D = 0.05, Re = 20. 132 1.00 m T j I 0 y 0.50; 0.25: m:_...=J 0.0 0.1 0.2 0.3 0.4 0-5 0.6 0.7 0.8 0.9 1-0 1.1 1.2 a) B = 0.25, D = 0.0. 1.00 1 NW ' I r 0.75~ , 0.503 0.25; 0.00‘ 1“") 0.0 0.1 0.2 0.3 0.4 0-5 0.6 0.7 0.8 0.9 1-0 1.1 1.2 b) B = 0.25, D = 0.05. 1 .00 .w . \ D.75~ y 0.50-1 0.25- J 0’00 I F I W 0.0 0.1 0"2 0.3 W04'05106'107T0‘Br01‘9 1.0 1.1 1.2 1.3 X c) B = 0.5, D = 0.05. Figure 32. Development of streamwise velocity profiles along the channel in the presence ofa finite step; Xl =1.0, Re = 20. 133 1.00 .\ 0.75: y 0.501 0.251 ‘ / 0.00 I 0.0 1.0 20 31:0 40- 60. T70 80 10.011.012.0 x a) D=0.05. 1.00 N 0.75~ 1 y 0.50“ 0.25- 0.00 ./ 0.0 1.0 I20 .30 40. 60. 7'0. 8.0 10.011.012.0 x b) D=0.25. 1.00 I\ ' l 0.75% y 0.50“ 0.25~ 0.00 ./ . , . . . 0.0 1.02.0 .0 60. 7.0 «80 10.0 11.0 12.0 x c) D=1.0. Figure 33. Development of streamwise velocity profiles along the channel in the presence of a finite step; X1 = 2.0, B = 0.2, Re = 200. 134 1 .00 N 0.75 1 4 y 0.50 r . 0.25 ~ 0.00 / 0.0 1.0 2.0 ':30 4.0 60. r{7078'10 10.011.012.0 Figure 34. Development of streamwise velocity profiles along the channel in the presence of a finite step; X1 = 1.0, B = 0.25, D = 0.05, Re = 200. IIIIIIIIIITITIIIIIIIII‘IITIIIWT II llll E I L 1 l1 llJl ll LlJlllLlLlLL a) B = 0.20, D = 0.05. lJlJllllli/llLlllll 1 141 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 b) B = 0.25, D = 0.25. Figure 35. Vorticity contours in the presence of flow separation due to a f'mite step; Xl =1.0, Re = 20. 135 _lll|lllllllllllllllllrITlllllIllIIIITIIIIIIIIIIIIIIIlllllll a) UB=-10.0. b) UB =-5.0. c) UB=-2.0. Figure 36. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB;X1=1.0, X2 = 2.0, Re = 20. d) UB=-l.0. Figure 36. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB; X1 = 1.0, X2 = 2.0, Re = 20. ‘.----—_-—-—~—~‘--~~na Figure 37. Velocity vectors of the flow with the lower boundary section moving opposite to the inlet velocity at UB = -10.0; X1 = 1.0, X2 = 2.0, Re = 20. 137 a) Un=1-0- b) 0,, = 2.0. TI—TIIIIITllllllllllIlllllllllllllllllllllllIllllllllilllill xx. 960 c) UB = 5.0. Figure38. Streamlines of the flow with the lower boundary section moving in the direction ofthe inlet velocity at UB;X1=1.0, X2 = 2.0, Re = 20. a) UB=-2.0. c) UB=-0.5. Figure 39. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity atUB;X1=1.0, X2 = 2.0, Re = 200. 139 Figure 40. Streamlines of the flow with the lower boundary section moving in the direction of the inlet velocity at UB = 1.0; X1 = 1.0, X2 = 2.0, Re = 200. Figure 41. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB = -l.0; X1 =1.0, X2 = 5.0, Re = 200. 140 b) UB = -1.0. Figure 42. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB; X1 = 4.0, X2 = 5.0, Re = 200. 141 a) U3 3 -2.0. b) UB=-l.0. Figure 43. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB; X1 = 4.0, X2 = 5.0, Re = 200. (A close up of Fig. 42.) 142 a) U3 3 -l.0. c) UB=1.0. Figure 44. Streamlines of the flow with the lower boundary section moving at UB; X1 = 2.0, X2 = 4.0, Re = 200. 143 a) U3 = -l.o. b) UB 2-0.5. c) UB = 1.0. Figure 45. Streamlines of the flow with the lower boundary section moving at UB; X1: 2.0, X2 = 6.0, Re = 200. 144 a) U3=-1.0. b) UB=-0.5. c) UB = 1.0. Figure 46. Streamlines of the flow with the lower boundary section moving at UB; X1 = 8.0, X2 =10.0, Re = 200. 145 Figure 47. Streamlines of the flow with the lower boundary section moving opposite to the inlet velocity at UB = -0.5; X1 = 10.0, X2 = 14.0, Re = 1000. .960 .960 .960———: ~ :1 ha.“- . 729- .72! _.729—_-: H“ -‘ Rh; : H __ M J“ 480 Jae—g :- g. —. 240 .240 .240—-—-;l 1- Figure 48. Streamlines of the flow with the entire lower boundary moving in the direction of the inlet velocity (Couette-Poiseuille flow) at UB = 1.0; Re = 500. 146 1.00 \ . , . 0.75; y 0.50-: 0.25: 0.00‘ ./ 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0 X a) Una-1.0. 1.00 .\ . r 4 0.75% y 0.50: 0.25; 0.001 v r 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0 x b) Un=-0.5. 1.00 .\ . r . 0.75; y 0.50: 0.253 0.00. / 4 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0 X C) U3 3 1.0. Figure 49. Development of streamwise velocity profiles along the channel with the lower boundary section moving at UB; X1 = 2.0, X2 = 4.0, Re = 200. 147 1&1 .\ . , . 0.751 y 0.50-1 0.25: 0"" / . - r , . . r 1 . . 0 3.0 4.0 5.0 6.0 .w I l , 1 0.0 1. 2.0 7.0 8.0 9.0 10-0 11.0 12.0 X a) UB='1.0. .30 '\ II 7' I 4 0.75- y 0.50- 0.25m onn‘ / 0W ' ' 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10-0 11.0120 X b) UB=-0.5. 1.00 .\ . u . 0.75— y 0.50“ 0.25— 0.00 '/ ' I 1 I I I I I I I I I 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0 X c) UB=1.0. Figure 50. Development of streamwise velocity profiles along the channel with the lower boundary section moving at UB; X1 = 2.0, X2 = 6.0, Re = 200. 148 1.00 .\ . 1 0.75; y 0.50j 0.25: 0.001 .1” 0.0 5.0 10.0 15-0 20.0 25-0 30.0 35.0 40.0 45.0 50.0 55.0 60.0 X Figure 51. Development of streamwise velocity profiles along the channel with the lower boundary section moving opposite to the inlet velocity at UB = -05; x1 = 10.0, X2 = 14.0, Re = 1000. 1.m '\ r r I 0.754 J y 0.504 0.25-1 1 01134 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0 X a) UB = 0.5.11: = 200. 1.00 I\ ' I ‘ 075‘ \ y 0.50 4 0.25 - 1 0.00 0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0 22.5 25.0 27.5 30.0 X b) UB = 1.0.11: = 500. Figure 52. Development of streamwise velocity profiles along the channel with the entire lower boundary moving in the direction of the inlet velocity (Couette-Poiseuille flow) at UB . 149 J. a l 6 j.- in 11111111111111111 - - 1— —1 - t: -1 1— u- p- -1 1— —1 - H 1— —I p— - F- -. ' —1 p— .11 .I ’ q . .. - -1 m‘ 4 I. " ..---. . Figure 53. Vorticity contours with the lower boundary section moving at UB; X1 = 1.0, X2 = 2.0, Re = 20. {é ’1 111‘111[IIIII II l i a a ’1111 1 1 1111111; a) U3 = -0050 .r- _ F- -4 C 2 1- -I *- -1 _ d——-——. f ' -' '_.___ fi.‘ _‘ 1— .1 1- .— E a L. I - Figure 54. Vorticity contours with the lower boundary section moving at UB; X1 = 1.0, X2 = 2.0, Re = 200. —4-W l I 1111 1111111 JLJ Figure 55. Vorticity contours for the entire lower boundary moving in the direction of the inlet velocity (Couette-Poiseuille flow) at U B = 1.0; Re = 500. 151 a) t = 2n, Us = 1.0 (at the instant when the boundary is moving at maximum velocity in the direction of the flow). b) t = (2n+l)11:, UB = -1.0 (at the instant when the boundary is moving at maximum velocity in the direction opposite to the flow). Figure 56. Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 1.0, X2 = 2.0, Re = 20, St = 1.0, A = 1.0, n = 8. 152 a) t = 201:,UB = 2.0. b) t=(20+1)rt,UB==-2.0. Figure 57. Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 2.0, X2 = 4.0, Re 2 200, St = 1.0, A = 2.0, n = 25. 153 a) t= 2111:,0B = 2.0. b) t=(2n+1)1t, UB = -2.0. Figure 58. Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 2.0, X2 = 6.0, Re = 200, St = 0.5, A = 2.0, n =- 14. 154 a) t = 2n1t, UB =1.0. e) t=(2n+l)11:, UB = -1.0. b) t= (2n-1~l/4)1t,UB = 0.707. f) t= (2n~1~5/4)1t',UB = -0.707. c) t = (2n+1/2)1t, UB = 0.0. g) t = (2n+3/2)1t, UB = 0.0. d) t= (20+3/4)1c, UB = -0.707. h) t= (2n+7/4)1t, UB = 0.707. Figure 59. Streamlines of the flow at different time with the lower boundary section oscillating at UB = Acos(t); X1 = 1.0, X2 = 2.0, Re = 200, St = 1.0, A = 1.0, n = 21. 155 a) St = 0.5, n = 13. b) St = 1.0, n = 21. c) St = 2.0, n = 27. Figure 60. Size of the separated region as a function of the Strouhal number; X1 = 1.0, X2 = 2.0, Re = 200, A =1.0, t = (20+1)1t, UB = -1.0. 156 a) St = 0.5, 0 =15. b) St = 1.0, n = 23. c) St = 2.0, n = 42. Figure 61. Size of the separated region as a function of the Strouhal number; X1 = 2.0, X2 = 4.0, Re = 200, A = 1.0, t = (20+ l)1t, UB = -1.0. 157 a) St = 0.5, n =16. b) St = 1.0, n = 26. c) St = 2.0, n = 45. Figure 62. Size of the separated region as a function of the Strouhal number, X1 = 2.0, X2 = 6.0, Re = 200, A = 0.5, t = (2n+1)1t, UB = -0.5. a) A =1.0, n = 8, UB = -1.0. b) A = 2.0, n = 8, UB = -2.0. Figure 63. Flow streamlines showing the size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 1.0, X2 = 2.0, Re = 20, St = 2.0, t = (2n+1)1t. 159 a) A = 0.5, n = 25, UB = -0.5. b) A = 2.0, n = 20, 0,, = -2.0. Figure 64. Size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 1.0, X2 = 2.0, Re = 200, St = 1.0, t = (20+1)1c. 160 a) A = 0.5, n = 46, UB = -0.5. b) A = 2.0, n = 25, U13 = -2.0. Figure 65. Size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 2.0, X2 = 4.0, Re = 200, St = 2.0, t = (2n+1)1t. 161 a) A = 1.0.11 = 24, UB = -1.0. b) A = 2.0.11 = 25, 0B = -2.0. Figure 66. Size of the separated region as a function of the amplitude of oscillation of the boundary; X1 = 2.0, X2 = 6.0, Re = 200, St = 1.0, t = (2n+1)11:. 162 a) X1=1.0,X2=2.0,St=1.0,n = 68. b) x1 = 10.0, x2 =14.0,s:= 1.0.11 =101. 0) x1 = 10.0, x2 = 14.0.31: 0.5, n = 56. Figure 67. Streamlines of the flow with the lower boundary section oscillating at UB = Acos(t); Re =1000, A = 1.0, t = (20+1)rt, U3 = -1.0. 163 Figure 68. Effect of the location of the oscillating section of the boundary on the size of the separated region; X1 = 5.0, X2 = 6.0, Re = 200, St = 1.0, A = 1.0, t = (2n+l)1r, n = 28, UB = -1.0. Figure 69. Effect of the length of the oscillating section of the boundary on the size of the separated region; X1 = 1.0, X2 = 3.0, Re = 200, St = 1.0, A = 1.0, t = (20+1)x, n = 26, U], = -1.0. a) St = 0.5, A = 2.0, n = 17, U3 = -2.0. b) St = 1.0, A =1.0, n = 24, U3 = -l.0. c) St = 2.0, A = 0.5, n = 60, U3 = -0.5. Figure 70. Effect of the location and length of the oscillating section of the boundary on the size of the separated region; X1 = 8.0, X2 = 10.0, Re = 200, t = (2n+ Du. . ’1 We 7. 165 1.00 .\ . . , . 0.75: y 0.50: 0.sz 0.001 ./ - 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10-0 11.0 12.0 X a) XI: 2.0, x2 = 4.0, s: = 1.0, A =1.0,n = 23, UB =1.o. 1.00 .\ ' 1 0.753 y 0.50: 0.25; 0.00‘ ./ - -='—~- 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0 X b) x1 = 2.0, x2 = 4.0, s: = 0.5, A = 2.0, n = 13, UB = 2.0. 1.00 .\ . r . 0.75: y 0.50: 0.251 0.001 ./ . . - 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10-0 11.0 12.0 X 0) x1: 2.0,X2=6.0, St= 1.0, A = 1.0, n =24, UB =10. Figure 71. Development of streamwise velocity profiles along the channel at the instant when the boundary is moving at maximum velocity in the direction of the flow; Re = 200, t = 2n1r. 166 1.00 .\ 0.75; y 0.50: 0.25: ‘ Z 0.00 i 0.0 1.0 2.0 X a) xl =2.0,x2 =4.0, St= l.0,A = 1.0.11 =23, UB = -1.0. tT70 rBrio 10-0 11.0 12.0 1.00 .\ 1 0.75-1 i y 0.50 '- 0.25 - 4 0.00- 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 X 10.0 11.0 12.0 b) x1 = 2.0, x2 = 4.0, s: = 0.5, A = 2.0, n =13,UB = .20. 1.00 N 12 T 0.751 y 0.50; 0.25-j 0.004 ./ ' 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 X 10.0 11.0 12.0 0) XI: 2.0, x2 = 6.0, s: = 1.0, A =1.0,n = 24, UB = -1.0. Figure 72. Development of streamwise velocity profiles along the channel at the instant when the boundary is moving at maximum direction opposite to the flow; Re = 200, t = (2n+1)1c. velocity in the 11 «0.1 0.0 g_g_____-_._— 0.! : .0 Z I _. 1 N. .- .a. m a .m.....---____- a) t= 2nrt, UB =1.0. 0 0 0 D —0.0 7 1', 12 0 3 "M... ______ , m.- .. a. '4 b) t= (2n+1)1t, UB = -1.0. Figure 73. Vorticity contours for the case of the lower boundary section oscillating at UB = Acos(t);X1=1.0, X2 = 2.0, Re = 20, St = 1.0, A =1.0, n = 8. 168 ‘_ I 1— n- P —I — - h _ I— .1 o a.-—_ 4.0 g g In- — 1— a— E — r- _ .. F A ‘— a) t= 2nx’ UB = 100. fi 14 —1 I -1 w .4 -1 J x '— ' A — "I§.‘s u _- b) t=(2n+1)1t, UB = -1.0. Figure 74. Vorticity contours for the case of the lower boundary section oscillating at UB = Acos(t);Xl=1.0, X2 = 2.0, Re = 200, St = 1.0, A =1.0, n = 21. 169 P- I- _ - - - 1- — d 1— _. 1— .4 1- —1 _ I - -1 - d - i -1 E ' l ‘ 1a : I ’7?,73.35:"I."'"‘,'.:—.‘-‘Ei‘:‘?::.::r:::;:::42553;.“1- Tull-n- ........ - a) t= 2n1r, UB =10. P - 1- -1 - -1 F _ _ _ ‘ 0.0 0.0 90.0----- b - — ‘ E _ - 1- .1 P - w b I- b) t=(20-1-1)1c,UB =-1.0. Figure 75. Vorticity contours for the case of the lower boundary section oscillating at U13 = Acos(t); X1 = 8.0, X2 =10.0, Re = 200, St = 1.0, A = 1.0, n = 24. APPENDIX 00000000000000000000OOOOOOOOOOOOOOOOOOOOOOO A COMPUTER PROGRAM PROGRAM ENTRY File Name: ENTRY.FOR Written By: Bashar S. AbdulNour - December 1988 A computer program for the numerical solution of the vorticity transport equation and stream function equation describing the time- dependent, viscous, incompressible flow in the entry region of a two-dimensional channel with part of the lower boundary oscillating. A finite-difference scheme, using the Alternating-Direction Implicit (ADI) method to solve the vorticity transport equation for the vorticity and the Successive Over-Relaxation (SOR) method to solve Poisson's equation for the stream function, is employed. DESCRIPTION OF VARIABLES M . M1 - M2 - N a I, J = number of grid points in streamwise direction (X-dir.); along channel's length number of grid points in X-dir. to the left of the oscillating boundary number of grid points in X-dir. to the right of the oscillating boundary number of grid points in normal direction (Y—dir.); along channel’s height subscripts corresponding, respectively, to nodal points in the x- and Y-directions I=1 is the left boundary 12M is the right boundary J=1 is the lower boundary J=N is the upper boundary J-NC is the centerline [domain of interior reigon : I=2 to M-l, J-2 to N-l] mesh size; equal in x- and Y—directions Reynolds number based on channel height and inlet velocity Strouhal number based on channel height and inlet velocity [- WL/U, where W is frequency of oscillation] 170 000000OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO00000000000000000000000 INC ' KT - KK,KKA(T)' T . TA(T) TKK KTP KINC AA PHI PHIA(T) X(I) Y(J) XR(I) F(I,J) Q(I.J) FS(I,J) QS(I,J) FC(I,J,T) QC(IIJIT)- FT(I,J,T)- QT‘IIJIT)- RFS AKS AKV BKV ITMAX ITER ITERA(T) ITMXF - ITMXQ ITERF ITERQ EPSILON EPSILONF EPSILONQ ERFO, ERQO ERF, ERQ DERFO, DERQO KOUNT NCYCLE U(I,J) V(I,J) U0(I,T) ILE LE 171 number of time increments in one time period time step used in the ADI method counters for time steps time array to store time (when convergence begins) time from start of the 2*PI period time fraction of the 2*PI period lable for time steps included in the convergence period amplitude of oscillation of boundary velocity input forcing function (i.e., velocity of the oscillating boundary which is a given function of time) array to store values of PHI at successive time steps streamwise (i.e., axial) coordinate [X=0 to (M-1)*H] normal (i.e., transverse) coordinate [Y=0 to 1] X(I)/RE calculated values of stream function calculated values of vorticity stored values of stream function stored values of vorticity values of stream function at selected locations used to check the convergence values of vorticity at selected locations used to check the convergence values of stream function at specified time during the last period before convergence; values are used in output values of vorticity at specified time during the last period before convergence; values are used in output over—relaxation factor for Poisson's equation weighting factor for stream function weighting factor for vorticity in interior region weighting factor for vorticity boundary conditions preselected maximum number of outer iterations (or time steps) number of outer iterations for the stream function and vorticity array to store ITER (when convergence begins) maximum permitted number of inner iterations for the stream function maximum permitted number of inner iterations for the vorticity number of inner iterations for the stream function number of inner iterations for the vorticity tolerance error (infinitesimally small quantity) stream function tolerance error vorticity tolerance error maximum error in stream function and vorticity, respectively, for the outer iteration maximum error in stream function and vorticity, respectively, for inner iterations denominators of the relative error terms for the stream function and vorticity, respectively, used in the outer iteration to avoid division by zero counter to insure that the outer iteration is convergant for consecutive iterations over one full period number of cycles needed for the solution to converge streamwise velocity (in the X-direction); u-U(X,Y) normal velocity (in the Y-direction); V=V(X,Y) velocity of the boundary (a function of time) grid point along the centerline representing the entrance length entrance length defined when the centerline velocity reaches 99% of its non-dimensional value for a fully- 000 0000OOOOOOOOOOOOOOOOOOOOOOO GOO GOO GOO 1T2 developed parabolic velocity profile LERE - LE/RE ~ MSKIP - counter to skip X-location data points in order to reduce printing of output at high RE MMl, NMl, PI, ETA, H2, TH, C1, C2, GI, 62: are constants defined throughout the program to facilitate the computation. They are defined as they occur in the program. DESCRIPTION OF SUBROUTINES Subroutine Function Q_ADI Solving the vorticity equation using the ADI method F_COL Solving the stream function equation using the SOR by columns method TRI( ) Solving a system of equation with tri-diagonal matrix using Thomas algorithm (it is used within other routines) MORE( ) Writing the complete results at different time into the output file Declaration of Variables (double precision for all real variables): INCLUDE 'INCLUDE.FOR’ ! include file INCLUDE.FOR to define arrays dimensions IDIM, JDIM, and KDIM; declare variables; and define common blocks REAL*8 AKS, AKV, ETA REAL*8 FC(IDIM10,JDIM,KDIM), QC(IDIM10,JDIM,KDIM) REAL*8 FT(IDIM,JDIM,KDIM8), QT(IDIM,JDIM,KDIM8) REAL*8 T, AA REAL*8 UO(IDIM10,KDIM) REAL*8 ERFO, DERFO REAL*8 ERQO, DERQO REAL*8 LE, LERE INTEGER INC, KK, KINC INTEGER KOUNT, KSKIP, KNDEX, K INTEGER NCYCLE INTEGER ILE INTEGER LUNITR, LUNITS, LUNITW Define Logical Unit Numbers: LUNITR - 12 ! logical unit number of input data file LUNITS - 6 ! logical unit number of screen (default output) LUNITW - 14 ! logical unit number of output file Read Input Data from Datafile: OPEN (LUNITR,FILE-'ENTRY.DAT',STATUS='UNKNOWN') READ (LUNITR,*) M,N,M1,M2,RE,ST,INC,RFS,AKS,AKV,BKV, & EPSILON,EPSILONF,EPSILONQ,ITMAX,ITMXF,ITMXQ CLOSE (UNIT-LUNITR) Compute Grid Points and Coordinates: MMl - M-l NMl - N-l NC (N+1)/2 ! label centerline nodes H - l.0/FLOAT(N-1) X(1) 173 - 0. XR(1) - 0. DO 9110 I-2,M X(I) - X(I-1) + H 9110 XR(I) - X(I)/RE Y(1) - 0 no 9120 3-2,n 9120 Y(J) for the 0000 - Y(J-1) + H Calculate the Optimum Value of the Over—Relaxation Factor Stream Function: PI - ETA - RFS - 000 4.*DATAN(1.D0) 0.5*(COS(PI/MM1) + COS(PI/NM1)) 2.*(l.-SQRT(1.-ETA**2))/ETA**2 Introduce and Calculate Some Constants and Parameters: 0 H IIIIIII COMPUTE 2.*PI/INC H*H 2.*H RPS/4. 1. - RFS KT/(RE*H) 2.*Gl BOUNDARY CONDITIONS ***********i*************** Stream Function Boundary Conditions: l. Upstream - Uniform Profile at Inlet 0000000 DO 9215 J-2,NM1 F(1,J) ‘0 N H U‘ - Y(J) FS(1,J) - Y(J) C 2. Downstream DO 9220 J-2,NM1 F(M,J) - 3.*(Y(J)**2) - 2.*(Y(J)**3) 9220 FS(M,J) - F(M,J) C C 3. Lower and Upper Boundaries C DO 9230 I-1,M F‘I’l) - O. FS(I,1) - 0. F(I,N) - 1.0 9230 FS(I,N) - 1.0 C C Vorticity Boundary Conditions: C C l. Upstream - Uniform Profile at Inlet C DO 9255 J-1,N Q‘lrJ) - 0. 9255 QS(1,J) - 0. C C 2. Downstream C DO 9260 J-1,N Q(M,J) - 12.*Y(J) - 6. 9260 QS(M,J) 3 Q(M,J) 174 3. Lower and Upper Boundaries [to be evaluated later since they depend on Q(I,2) and Q(I,N-l), respectively] Initialization of Values on Grid of Interior Region: 0()0(30(3C) DO 9400 I-Z,MM1 DO 9400 J-2,NM1 F(I,J) - Y(J) 9400 Q(I,J) - 0. C C Initialization of Values of F and Q Selected to Test Convergence: C DO 9420 I-11,M-10,10 DO 9420 J-2,NM1 DO 9420 KK-1,INC II - (I - 1)/10 FC(II,J,KK) - 0. 9420 QC(II,J,KK) - 0. C C Initialization of Values of F and 0 Selected for Output: C DO 9422 I-2,MM1 DO 9422 J-2,NM1 D0 9422 K=1,KDIM8 FT(I,J,K) - o. 9422 QT(I,J,K) - o. c C BEGIN OUTER ITERATION FOR THE STREAM FUNCTION AND VORTICITY C ***t******************************************************* KR - O KOUNT 0 KSKIP INC/8 ! number of time steps skipped KNDEX ! time step index for FT and QT K ! same as KNDEX ITERF ITERQ (DOCDO DO 9590 ITER=1,ITMAX IF (KK.EQ.INC) KK - 0 RR - KK + 1 T - ITER*KT PHI - AA*DCOS(T) Storing All Old Values of the Stream Function and Vorticity (Values from the Previous Time Step): 0(3C)0 DO 9450 I=Z,MM1 DO 9450 J82,NM1 FS(I,J) - F(I,J) 9450 QS(I,J) - Q(I,J) C SOLVING THE VORTICITY TRANSPORT EQUATION FOR THE VORTICITY *******t************************************************** 3. (Continued) Calculate New Values of Lower and Upper Vorticity Boundary Condition at Begining of Vorticity Inner Iteration [at new time step] Using the Second-Order Woods’ Formula with Values of the Stream Function and Vorticity Known from Previous Iteration C)0(1()0(7()0 D0 9512 I'Z,M1-1 QS‘Ill) - Q‘Irl) 9512 9514 9516 9520 C C C C C C C C C 9530 9540 550 560 0000090 0000000 00000 175 Q‘Irl) - 3-*(F(I11)’F(Ilz))/Hz " Q(IIZ)/2o Q(I,1) I BKV*QS(I,1) + (1.-BKV)*Q(I,1) QS(M1I1)- Q‘er 1) Q(Ml,1) I 3.*F(M1,1)/HZ - 3.*(F(M1-l,2)+F(M1+1,2))/(2.*H2) - Q(Ml,2)/2. + 3.*PHI/TH Q(Ml,1) I BKV*QS(M1,1) + (1.-BKV)*Q(M1,1) DO 9514 IIM1+1,M2-1 QS(II1) ' Q‘Irl) Q(I,l) - 3.*(F(I,l)-F(I,2))/H2 - Q(I,2)/2. + 3.*PHI/H Q‘Ill’ ' BKV*QS(II1) + (1.-BKV)*Q(I,1) QS(M2.1)- 0(M2.1) Q(M2,l) - 3.*F(M2,l)/H2 - 3.*(F(M2-1,2)+F(M2+1,2))/(2.*H2) - Q(M2,2)/2. + 3.*PHI/TH Q(M2.1) - BKV*QS(M2.1) + (1.-BKV)*Q(M2.1) DO 9516 IIM2+1,MM1 QS(I.1) - 0(I.1) 0(I.1) - 3.*(F(I,1)-F(I.2))/H2 - 0(I,2)/2. Q(I,1) - BKV*QS(I,1) + (l.-BKV)*Q(I,1) DO 9520 II2,MM1 QS(IIN) ' Q(IIN) Q‘IIN) ' 3-*(F(I:N)'F(I:NM1))/HZ - Q(IINM1)/20 C(IrN) ' BKV*QS(I:N) + (1.-BKV)*Q(I,N) Calling Vorticity Equation Solver: CALL Q_ADI End of inner iteration for the vorticity. AAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAA Recalculate Q(I,J) Using Weighting Factor: ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ DO 9540 II2,MM1 DO 9540 JI2,NM1 Q(I,J) I AKV*QS(I,J) + (1.-AKV)*Q(I,J) SOLVING POISSON EQUATION FOR THE STREAM FUNCTION t***********t*********************************** Calling Poisson's Equation Solver: CALL F_COL End of inner iteration for stream function. AAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAA Recalculate F(I,J) Using Weighting Factor: ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ DO 9560 II2,MM1 DO 9560 JI2,NM1 F(I,J) I AKS*FS(I,J) + (1.-AKS)*F(I,J) Test the Values of the Stream Function and Vorticity Recalculated from the Outer Iteration for Convergence (by comparing selected values over one time period): ERFO I 0.7 ERQO I 0. DO 9580 II11,M-10,10 DO 9580 JI2,NM1 176 II I (I - 1)/10 DERFO I F(I,J) + EPSILON DERQO I Q(I,J) + EPSILON ERFO I DMAX1(ERFO,DABS((F(I,J)-FC(II,J,KK))/DERFO)) 9580 ERQO I DMAX1(ERQO,DABS((Q(I,J)-QC(II,J,KK))/DERQO)) C IF ((ERFO.LE.EPSILONF).AND.(ERQO.LE.EPSILONQ)) THEN IF ((KK-l).EQ.0) KK I INC + 1 IF ((KOUNT.GT.0).AND.(ITER-ITERA(KK-1)).NE.1) KOUNT I 0 IF (KK.GT.INC) KK I 1 9581 KOUNT I KOUNT + 1 ITERA(KK) I ITER KKA(KK) I KK TA(KK) I T PHIA(KK) I PHI DO 9584 IIM1,M2,10 II - (I - 1)/1o 9584 U0(II,KK) - pa: IF (KOUNT.GE.KK) THEN KNDEX I INT(KK/KSKIP) IF (KNDEX.EQ.(K+1)) THEN K I KNDEX DO 9586 I-1,M ' storing values of the stream DO 9586 JIl,N ! function and vorticity of FT(I,J,K) - F(I,J) ' the last full period before 9586 QT(I,J,K) I Q(I,J) ' exiting ENDIF ENDIF IF ((KOUNT.LT.INC).OR.(KK.LT.INC)) GO TO 9588 GO TO 9600 ENDIF Storing Selected Values of the Stream Function and Vorticity at the End of Each Time Step (to be Used to Check for Convergence): 900000 588 D0 9589 I=11,M-10,10 DO 9589 J=l,N II - (I - 1)/1o FC(II,J,KK) I F(I,J) 9589 QC(II,J,KK) I Q(I,J) C 9590 CONTINUE C C Error Message if Number of Outer Iterations Exceed ITMAX: C IF(ITER.GE.ITMAX) THEN 9980 WRITE (LUNITS, 9981) ITMAX 9981 FORMAT (/, 2X,'ERROR ,/, 2x, ----- ', 5 /, 2X,’JOB was ABORTED DUE To OUTER ITERATION PROBLEM. & /, 2X,'NUMBER OF OUTER ITERATIONS ', & 'HAS EXCEEDED ITMAX - ',I4,/) ENDIF Calculate the Number of Cycles for Convergence: 600 NCYCLE I ITER/INC End of outer iteration. AAAAAAAAAAAAAAAAAAAAAAA 0000\0000 177 Inlet and Downstream End: 000 U(1,1) I U(l,NC) U(l,N) I U(l,NC) Open Output File: WRITE (LUNITW,9801) 801 FORMAT (/.1x,'F11e Name: ENTRY.OUT',/) Write Data and Calculated Results into Output File: 000“) 0 000 WRITE (LUNITW,9810) 9810 FORMAT (/,1x,'Data:', Compute the Streamwise Velocity on the Lower Boundary at the a /,1x,'---- ',/) C WRITE (LUNITW,9811) RE,ST 9811 FORMAT (2X,’REYNOLDS NUMBER, RE - ’,F10.1,/, & 2X,’STROUHAL NUMBER, ST - ',F12.3,/) C WRITE (LUNITW,9815) M,N,H 9815 FORMAT (2X,’TOTAL x-DIR. GRID POINTS, M - ',I8 ,/, & 2x,'TOTAL Y-DIR. GRID POINTS, N - ’,I8 ,/, & 2X,'MESH SIZE IN x- AND Y-DIRECTION, H - ',F12.3 ) C WRITE (LUNITW,9817) KT 9817 FORMAT (2x,'TIME STEP, KT - ',F12.3,/) C WRITE (LUNITW,9822) RFS OPEN (LUNITW,FILEI’ENTRY.OUT’,STATUSI'NEW',FORM='FORMATTED') 9822 FORMAT (2X,'STREAM FN OVER-RELAXATION FACTOR, RFS - ',F7.3) WRITE (LUNITW,9824) AKS,AKV,BKV 9824 FORMAT (2X,'STREAM FN WEIGHTING FACTOR, AKS - ',F7.3,/, & 2X,'VORTICITY WEIGHTING FACTOR, AKV - ',F7.3,/, & 2X,'VORTICITY B.C. WEIGHTING FACTOR, BKV - ’,F7.3,/) C WRITE (LUNITW,9831) EPSILONF,EPSILONQ 9831 FORMAT (2x,'STREAM FN PERMITTED ERROR, EPSILONF - ',E14.6,/, & 2X,'VORTICITY PERMITTED ERROR, EPSILONQ - ',El4.6,/) C WRITE (LUNITW,9833) ERF,ERQ,ERFO,ERQO 9833 FORMAT (2X,’STREAM FN ERROR (INNER ITERATION), ERF - ',El4.6,/, & 2X,’VORTICITY ERROR (INNER ITERATION), ERQ - ',E14.6,/, & 2X,'STREAM FN ERROR (OUTER ITERATION), ERFO - ’,E14.6,/, & 2X,'VORTICITY ERROR (OUTER ITERATION), ERQO - ',El4.6,/) WRITE (LUNITW,9842) ITMAX,ITMXF,ITMXQ,ITERF,ITERQ,ITER 9842 FORMAT ( 2X,’PERMITTED MAXIMUM NO. OF ITERATIONS, ITMAX 2X,’PERMITTED MAX. NO. OF STREAM FN ITERATIONS, ITMXF 2X,’PERMITTED MAX. NO. OF VORTICITY ITERATIONS, ITMXQ 2X,'FINAL ACTUAL NO. OF STREAM FN ITERATIONS, ITERF 2X,’FINAL ACTUAL NO. OF VORTICITY ITERATIONS, ITERQ 2X,’ACTUAL NO. OF OUTER ITERATIONS, ITER hhhhhh WRITE (LUNITW,9846) NCYCLE 9846 FORMAT ( & 2X,'NUMBER OF CYCLES FOR CONVERGENCE, NCYCLE WRITE (LUNITW,9350) 9850 FORMAT (1X,'Results:’,/, 8 1x" _______ "/) \D 9! 986 966 9872 9877 9875 9879 (Itvnnnn lfll'l‘fllf" Q ~~ 9855 9859 9870 9867 9861 9872 9877 9875 9879 (UC)0<1()0 9886 C 9888 C C C C 178 WRITE (LUNITW,9855) X(M) FORMAT (2X,’TOTAT LENGTH OF CHANNEL - ’,F8.4 ,/) WRITE (LUNITW,9859) FORMAT (’1',/) WRITE (LUNITW,9870) FORMAT (2X,'VORTICITY AT LOWER AND UPPER BOUNDARIES FOR ', 'ONE FULL PERIOD BEFORE ',/, 2X,'EXITING ', '(ONE FULL PERIOD AFTER START OF CONVERGENCE) ',//) bl!” DO 9879 KINCI1,INC TKK - KKA(KINC)*KT KTP - TKK/(2.*PI) WRITE (LUNITW,9867) FORMAT (2x,' ', I I) WRITE (LUNITW,9861) KKA(KINC),ITERA(KINC),TA(KINC),TKK,KTP FORMAT (/,2x,'KK - ',16 ,/, & & 2X,'ITER - ',16 ,/, & 2x,'TIME, T - ',F9.4,/, & 2X,'TKK - ',F9.4,/, & 2X,'KTP - ',F9.4,/) WRITE (LUNITW,9872) FORMAT (2x,' x ' , 5 5X:' U(XIOIT) 'IZXI' Q(X:0:T) ',2X,' Q(X:H,T) ',/) WRITE (LUNITW,9875) X(1),U(l,1),Q(l,l),Q(l,N) DO 9877 II11,M-10,10 II - (I - 1)/1o WRITE (LUNITW,9875) X(I),U0(II,KINC),QC(II,1,KINC),QC(II,N,KINC) WRITE (LUNITW,9875) X(M),U(M,1),Q(M,1),Q(M,N) FORMAT (2x,F7.3,2x,3(2x,F10.6)) CONTINUE Call Subroutine MORE to Continue Writing Results (X, Y, U, V, F, and Q) at Selected Time Steps (Within One Full Period Before Exiting) into Output File; Values of Stream Function and Vorticity are Reassigned First: DO 9838 K=1,KDIM8 KINC ' K*KSKIP DO 9886 II1,M DO 9886 J=I,N F(I,J) I FT(I,J,K) Q(IIJ) ' QT(IIJIK) CALL MORE (ILE,KINC,LUNITW) Close Output File: CLOSE (UNIT-LUNITW) STOP END ** ******************************************************************** 00000000 0 0000000000000000000000000000000000000000000000000 0000 179 SUBROUTINE Q_ADI A subroutine for solving vorticity transport equation for the vorticity using the Alternating-Direction Implicit (ADI) method. The ADI method splits each time step into two halves. For the first half, the discrete form of the vorticity equation is implicit in X. For the second half the difference equation is implicit in Y. Each equation yields a tri-diagonal system of simultaneous linear algebraic equations. The resulting systems are solved using Thomas algorithm which is a special adaptation of the Gaussian elimination. This process is repeated over successive time steps. DESCRIPTION OF VARIABLES EXECLUSIVE TO THIS ROUTINE old values of the vorticity from the previous inner iteration intermediate values of the vorticity at the end of first half of time step U(I,J)*K/2 V(I,J)*K/2 finite-difference multiplier used to convert the central difference approximation of the first order derivatives of the vorticity with respect to x, contained in the convective terms, to forward/ backward difference approximation ALPHAl-l give backward difference approximation ALPHAlI—l give forward difference approximation ALPHA2( ) same as above but for derivatives with respect to Y Al(I), I one-dimensional arrays containing elements of the A2(I), upper, main, and lower diagonals, respectively, of A3(I) the tri-diagonal matrix of the vorticity equation; it applies to both the X-equation (first half of time step) and the Y-equation (second half of time step) BB(I) I one-dimensional array containing elements of the right-hand-side column matrix of the vorticity equation WW(I) I column matrix of the solution of the vortivity equation DERQ I denominator of the relative error term for the vorticity; used in inner iteration to avoid division by zero QOLD(I:J7 QINT(I,J) UDT(I,J) VDT(I,J) ALPHA1( ) INCLUDE 'INCLUDE.FOR’ ! include file REAL*8 QOLD(IDIM,JDIM), QINT(IDIM,JDIM) REAL*8 UDT(IDIM,JDIM), VDT(IDIM,JDIM) REAL*8 ALPHA1(IDIM,JDIM), ALPHA2(IDIM,JDIM) REAL*8 DERQ Calculate the Velocities and difference operators at All Interior Nodes: DO 8200 I=2,MM1 DO 8200 J-2,NM1 U(I,J) - (F(I,J+1)-F(I,J-1))/TH IF (U(I,J).GT.0.0) THEN 3 C CC GUCCCC C C (c 84C 55C Ab ECCC Ch» CC 8200 0 ()0(3 250 0(36)0(3()m 0 8300 C C 8400 C 8500 0(1()0 8550 0(0630 180 ALPHA1(I,J) I 1.0 ELSEIF (U(I,J).LT.0.0) THEN ALPHA1(I,J) I -1.0 ELSE ALPHA1(I,J) ' 0.0 ENDIF UDT(I,J) I U(I,J)*KT/2. V(I,J) - (F(I-1,J)—F(I+1,J))/TH IF (V(I,J).GT.0.0) THEN ALPHA2(I,J) - -1.o ELSEIF (V(I,J).LT.0.0) THEN ALPHA2(I,J) - 1.0 ELSE ALPHA2(I,J) - 0.0 ENDIF VDT(I,J) - V(I,J)*KT/2. Begin ADI Iteration to Solve the Vorticity Equation: DO 8900 ITERQI1,ITMXQ DO 8250 II2,MM1 DO 8250 J=2,NM1 QOLD(I,J) ' Q(IIJ) COMPUTE THE VORTICITY FOR INTERIOR REGION: l. X-Equation for the First Half of Time Step; Solve for the Intermediate Value of the Vorticity DO 8500 JI2,NM1 DO 8300 I=2,MM1 A1(I) I -(1+ALPHA1(I,J))*UDT(I,J) - Gl A2(I) I (2*ALPHA1(I,J))*UDT(I,J) + 62 + TH*ST A3(I) I (l-ALPHA1(I,J))*UDT(I,J) - 61 38(1) I (( (1+ALPHA2(I,J))*VDT(I,J) + 61) *Q(I,J-l) +(—(2*ALPHA2(I,J))*VDT(I,J) - GZ + TH*ST)*Q(I,J) +(-(1-ALPHA2(I,J))*VDT(I,J) + 61) *Q(I,J+l)) CONTINUE BB(2) I BB(2) - Al(2)*Q(1,J) A1(2) I 0. BB(MMI) I BB(MMl) - A3(MM1)*Q(M,J) A3(MM1) I 0. CALL TRI (2,MM1) ! tri-diagonal system solver; solves for Q at fixed J DO 8400 II2,MM1 QINT(I,J) I WW(I) CONTINUE 2. Rename the Intermediate Values of the Vorticity X-Boundary Conditions at End of First Half of Time Step DO 8550 J31,N QINT(1,J) ' Q(1aJ) QINT (“I J) - Q (“I J) 3. Compute the Stream Function and recalculate the Velocities and difference operators at All Interior Nodes: 8580 C C C C 00 8600 C C C 8700 C 8800 C 0 000 8850 8900 & & 181 CALL F_COL DO 8580 II2,MM1 DO 8580 JI2,NM1 U(I,J) I (F(I:J+1)-F(I,J-1))/TH IF (U(I,J).GT.0.0) THEN ALPHA1(I,J) - —1.0 ELSEIF (U(I,J).LT.0.0) THEN ALPHAl(I,J) - 1.0 ELSE ALPHA1(I,J) - 0.0 ENDIF UDT(I,J) - U(I,J)*KT/2. V(I,J) - (F(I-1,J)-F(I+1,J))/TH IF (V(I,J).GT.0.0) THEN ALPHA2(I,J) - 1.0 ELSEIF (V(I,J).LT.0.0) THEN ALPHA2(I,J) - -1.o ELSE ALPHA2(I,J) - 0.0 ENDIF VDT(I,J) - V(I,J)*xT/2. 4. Y-Equation for the Second Half of Time Step; Compute a New Value for the Vorticity at End of Time Step DO 8800 II2,MM1 DO 8600 JI2,NM1 A1(J) I -(1+ALPHA2(I,J))*VDT(I,J) - G1 A2(J) I (2*ALPHA2(I,J))*VDT(I,J) + GZ + TH*ST A3(J) - (l-ALPHA2(I,J))*VDT(I,J) - Gl BB(J) I (( (1+ALPHA1(I,J))*UDT(I,J) + G1) *QINT(I-1,J) +(-(2*ALPHA1(I,J))*UDT(I,J) - 62 + TH*ST)*QINT(I,J) +(-(1-ALPHA1(I,J))*UDT(I,J) + G1) *QINT(I+1,J)) CONTINUE BB(2) I BB(2) — A1(2)*Q(I,1) A1(2) I 0. BB(NM1) - BB(NM1) - A3(NM1)*Q(I,N) A3(NM1) - 0. CALL TRI (2,NM1) ! tri-diagonal system solver; solves for Q at fixed I DO 8700 JI2,NM1 Q(I,J) - WW(J) CONTINUE IF (ITMXQ.EQ.1) RETURN Test the Vorticity for Convergence: ERQIO. DO 8850 II2,MM1 DO 8850 JI2,NM1 DERQ I Q(I,J) + EPSILON BRO " DMAX1(ERQ,DABS( (Q(I:J)-QOLD(I,J) ) /DERQ) ) IF (ERQ.LE.EPSILONQ) GO TO 8950 CONTINUE 182 \O U C) RETURN END ********************************************************************** SUBROUTINE F_COL A subroutine for solving Poisson's equation for the stream function using the Successive Over-Relaxation (SOR) method. The difference form of the stream function equation yields a system of simultaneous linear algebraic equations. The SOR by blocks (or lines) method is used to solve the resulting system of equations. Blocks are chosen to be columns of the computational domain. DESCRIPTION OF VARIABLES EXECLUSIVE TO THIS ROUTINE FOLD(I,J) I stored values of the stream function (from previous iteration) AI(J), I one-dimensional arrays containing elements of the BI(J), upper, main, and lower diagonals, respectively, of CI(J) the tri-diagonal matrix of Poisson's equation (representing nodes J in column I) DI(J) I one-dimensional array containing elements of the right-hand-side column matrix of Poisson’s equation WI(J) I column matrix of the solution of Poisson's equation DERF I denominator of the relative error term for the stream function; used in inner iteration to avoid division by zero INCLUDE ’INCLUDE.FOR' ! include file REAL*8 DERF Begin SOR Iteration to Solve Poisson's Equation: DO 5700 ITERF-1,ITMXF 0 000 0 000000000000000000000000000000 00000 (:30 DO 5300 II2,MM1 DO 5300 JI2,NM1 5300 FOLD(I,J) I F(I,J) C C COMPUTE THE STREAM FUNCTION FOR INTERIOR REGION: C Start Solution of Block-Iteration by Columns and SOR: C DO 5500 II2,MM1 C AI(2) I 0. 81(2) I -4. CI(2) I l. DI(2) - -H2*Q(I,2) - F(I-1,2) - F(I+1,2) - F(I,1) DO 5400 JI3,NM1-1 AI(J) I 1. BI(J) I I4. CI(J) I l. 5400 5450 C 5500 0 0 000 5600 5700 5950 183 DI(J) -H2*Q(I,J) - F(I—1,J) - F(I+1,J) AI(NMl) - 1. BI(NM1) I -4. CI(NM1) - 0 DI(NM1) - IH2*Q(I,NM1) - F(I-1,NM1) — F(I+1,NM1) - F(I,N) CALL TRI (2,NM1) ! tri-diagonal system.solver; solves for F at fixed I DO 5450 JI2,NM1 F(I,J) I WI(J) F(I,J) I RFS*F(I,J) + C2*FOLD(I,J) CONTINUE IF (ITMXF.EQ.1) RETURN Test the Stream Function for Convergence: ERF I 0. DO 5600 II2,MM1 DO 5600 JI2,NM1 DERF I F(I,J) + EPSILON ERF - DMAXl(ERF,DABS((F(I,J)-FOLD(I,J))/DERF)) IF (ERF.LE.EPSILONF) GO TO 5950 CONTINUE RETURN . END *‘k'k****i************************************************************** 00000000000000000000000000 00000 SUBROUTINE TRI (K,L) A subroutine for solving a system of simultaneous linear algebraic equations with a tri-diagonal coefficient matrix. It uses the Thomas algorithim method which is a special adaptation of the Gaussian elemination method in the form of a The equations are numberd from K through L. diagonal, and superdiagonal coefficients are A, B, and C, respectively, each of dimension of length (L) containing the right—hand side The computed solution is the vector W(L). recursion formula. Their subdiagonal, stored in the arrays (L). D is a vector of the linear system. DESCRIPTION OF OTHER VARIABLES EXECLUSIVE TO THIS ROUTINE K, L I subscripts corresponding, respectively, to the first and last equations to be solved [usually KIl for a system of L equations] BETA(L), GAMA(L) I vectors of the intermediate coefficients used in the solution of the system as demanded by the recursion formula INCLUDE ’INCLUDE.FOR’ ! include file 0(30 2100 0 0 0(30 2200 0 184 REAL*8 BETA(IDIM), GAMA(IDIM) INTEGER K,L INTEGER KP1,LMK Computing intermediate arrays: BETA(K) I B(K) GAMA(X) I D(K)/BETA(K) KPl - x + 1 DO 2100 IIKP1,L BETA(I) - B(I) - A(I)*C(I-1)/BETA(I-1) GAMA(I) - (D(I) - A(I)*GAMA(I-1))/BETA(I) Computing the solution vector: W(L) I GAMA(L) LMK I L I R DO 2200 JI1,LMK I I L I J W(I) I GAMA(I) I C(I)*W(I+1)/BETA(I) RETURN END ********************************************************************** 0 0()0(00()0(30 0(30030 0()0 0 9620 9630 9635 SUBROUTINE MORE (ILE,KINC,LUNITW) Subroutine to compute the velocities and write data and calculated results at any time step during the one full period before the final time step (at convergence) into the output file. INCLUDE 'INCLUDE.FOR' ! include file INTEGER KINC, ILE, MSKIP INTEGER LUNITW PHI I PHIA(KINC) Compute the Streamwise Velocity: DO 9630 I-1,MM1 U(I,l) - o. U(I,2) - (I25.*F(I,2)+48.*F(I,3)-36.*F(I,4)+16.*F(1,5) I3.*F(I,6))/(12.*H) Do 9620 JI3,N-2 U(I,J) - (-F(I,J+2)+8.*F(I,J+1)-8.*F(I,J-1)+F(I,J-2))/(12.*H) JINMl U(I,J) - ( 25.*F(I,J)-48.*F(I,J-1)+36.*F(I,J-2)-16.*F(I,J-3) +3.*F(I,J-4))/(12.*H) U(I,N) - 0. Do 9635 I-M1,M2 U(I,1) - PHI O 96 9E 000w D ‘0 no 981 986 185 DO 9640 JI1,N 9640 U(M,J) I 6*Y(J) I 6*(Y(J)**2) U(1,1) I U(l,NC) U(l,N) I U(l,NC) C Compute the Normal Velocity: DO 9650 II1,M 9650 V(I,1) I 0. DO 9660 JI2,N V(1,J) I o. V(2,J) - I(I25.*F(2,J)+48.*F(3,J)I36.*F(4,J)+16.*F(5,J) & I3.*F(6,J))/(12.*H) DO 9655 II3,M-2 9655 V(I,J) I I(-F(I+2,J)+8.*F(I+1,J)I8.*F(I-l,J)+F(I-2,J))/(12.*H) V(MM1,J) I I(25.*F(MM1,J)I48.*F(MI2,J)+36.*F(MI3,J)I16.*F(MI4,J) & +3.*F(MIS,J))/(12.*H) 9660 V(M,J) I 0. C C Continue Writing Data and Calculated Results into Output File: WRITE (LUNITW,9859) 9859 FORMAT ('1',/) C WRITE (LUNITW,9860) 9860 FORMAT (2X,’OUTPUT FROM THE LAST PERIOD BEFORE EXITING',/) C TKK I KKA(KINC)*KT RTP - TKK/(2.*PI) WRITE (LUNITW,9861) KKA(KINC),ITERA(KINC),TA(KINC),TKK,KTP 9861 FORMAT (/,2X,'KK I ',16 ,/, & 2X,'ITER I ',16 ,/, & 2X,'TIME, T - ',F9.4,/, & 2X,'TKK I ',F9.4,/, & 2X,'KTP - ’,F9.4,/) WRITE (LUNITW,9862) 9862 FORMAT ( I. & 22x,' VELOCITY’,2X,’ VELOCITY', & 4x,' STREAM FN',2X,' VORTICITY’,/, & 2x,' x ' ,2x,' Y ' , a 4x,' U ',2x,' v ', 8 4X,’ F ',2X,' Q ',/) C I I 0 9863 IF (I.LT.M) THEN IF (I.LE.100) THEN I I I + 1 ELSE MSKIP I 2*(INT(I/100)+1) I I I + MSKIP IF (I.GT.(100*MSKIP/2 + 1)) I I 100*INT(I/100) + 1 ENDIF DO 9866 JI1,N WRITE (LUNITW,9864) X(I),Y(J),U(I,J),V(I,J),F(I,J),Q(I,J) 9864 FORMAT (2(2X,F7.3),2X,2(2X,F10.6),2X,2(2X,F10.6)) 9866 CONTINUE WRITE (LUNITW,9867) 9867 FORMAT (2X,' ', & l I) GO TO 9863 ENDIF 9881 MSKIP I INT(MM1/50) IF (MSKIP.LT.1) MSKIP I 1 0 0 00000000000000000000000000000000000000 186 RETURN END *******************t****‘k********************************************* ** File Name: INCLUDE.FOR A FORTRAN file to include the following FORTRAN statements which are shared between the main program and all subroutines: 1. PARAMETER statements defining arrays dimensions IDIM, JDIM and KDIM; these are the maximum allowable number of array elements in the x and Y directions and in time, respectively. 2. DECLERATION statments for all global variables shared between the main program and all subroutines. 3. Common statements representing common variables and arrays; this will reserve the same memory location for these arrays. To include this file in the FORTRAN code (main program and all related subroutines) use the VAX/VMS FORTRAN statement: INCLUDE 'INCLUDE.FOR' The contents of this file will then be included in all FORTRAN programs and subroutines; i.e., IDIM and JDIM need no longer be be defined in a PARAMETER statement; variables need no longer be be passed as a subroutine argument or in a COMMON statement (i.e., common block) in the main programs or subroutines. All variables and arrays in this file are defined in the main program or in the subroutines where they occur. Declaration of Variables; Double Precision for All Real Variables: PARAMETER (IDIM-2001,JDIMI51) ! define space dimensions of arrays PARAMETER (IDIM10I(IDIMI1)/10) PARAMETER (KDIMIIZO) PARAMETER (KDIM8I8) ! define time dimension of arrays REAL*8 RE, ST REAL*8 RFS REAL*8 va REAL*8 H, H2, TH REAL*8 X(IDIM), Y(JDIM), XR(IDIM) REAL*8 F(IDIM,JDIM), Q(IDIM,JDIM) REAL*8 FS(IDIM,JDIM), QS(IDIM,JDIM) REAL*8 FOLD(IDIM,JDIM) REAL*8 PI REAL*8 KT REAL*8 PHI REAL*8 TA(KDIM), PHIA(KDIM) REAL*8 TKK, KTP REAL*8 C1, C2 REAL*8 G1, 62 REAL*8 EPSILON, EPSILONF, EPSILONQ REAL*8 ERF, ERQ 000 0000 187 REAL*8 U(IDIM,JDIM), V(IDIM,JDIM) REAL*8 A(IDIM), D(IDIM), C(IDIM), D(IDIM), W(IDIM) REAL*8 A1(IDIM), A2(IDIM), A3(IDIM), BB(IDIM), WW(IDIM) REAL*8 AI(IDIM), BI(IDIM), CI(IDIM), DI(IDIM), WI(IDIM) INTEGER M, N, 1041, NM1, NC, M1, M2 INTEGER ITERA(KDIM), KKA(KDIM) INTEGER ITMAx, ITER, ITMXF, ITERF, ITMXQ, ITERQ EQUIVALENCE (A,A1,AI), (B,A2,BI), (C,A3,CI), (D,BB,DI), & (W,WW,WI) Common Blocks: COMMON / NAME010 / RE COMMON / NAME011 / ST COMMON / NAMEozo / M, N COMMON / NAME021 / MM1, NM1 COMMON / NAME023 / NC COMMON / NAME025 / M1, M2 COMMON / NAME028 / M81, M52, NS COMMON / NAMEO30 / H COMMON / NAME031 / H2 COMMON / NAME032 / TH COMMON / NAME039 / PI COMMON / NAME040 / RFS COMMON / NAME041 / BKV COMMON / NAME045 / C1, C2 COMMON / NAME051 / KT COMMON / NAME053 / 61, G2 COMMON / NAME054 / PHI COMMON / NAME056 / ITERA, KKA COMMON / NAME058 / TA, PHIA COMMON / NAME058 / TKK, KTP COMMON / NAME060 / ITMAx, ITER COMMON / NAME061 / ITMXF, ITERF COMMON / NAME062 / ITMxo, ITERQ COMMON / NAME070 / EPSILON COMMON / NAME071 / EPSILONF COMMON / NAME072 / EPSILONQ COMMON / NAME075 / ERF COMMON / NAME076 / ERQ COMMON / NAME080 / x, Y, XR COMMON / NAME091 / F, Q COMMON / NAME093 / Fs, QS COMMON / NAME095 / FOLD COMMON / NAMEloo / A, B, C, D, W COMMON / NAME111 / U, v ******************i**********t**************************************** BIBLIOGRAPHY BIBLIOGRAPHY Abarbanel, S., Bennett, 8., Brandt, A., and Gillis, 1., 1970, “Velocity Profiles Of Flow at Low Reynolds Number”, ASME Journal of Applied Mechanics, Vol. 37, pp. 2-4. Abdulla, N. N., 1987, “Laminar Flow Separation in a Constricted Channel”, Ph.D. Dissertation, Department Of Mechanical Engineering, Michigan State University, East Lansing, MI. Allen, D., and Southwell, R. V., 1955, “Relaxation Methods Applied to Determine the Motion, in Two Dimensions, Of a Viscous Fluid Past 11 Fixed Cylinder”, Quarterly Journal of Mechanics and Applied Mathematics, Vol. 8, pp. 129-145. Ames, W. F., 1977, Numerical Methods for Partial Difl’erential Equations, 2nd Ed., Academic Press, New York. Anderson, C. R., 1989, “Observations on Vorticity Creation Boundary Conditions”, Mathematics Aspects of Vortex Dynamics, E. C. Caflisch, ed., pp. 144-159, SIAM. Anderson, D. A., Tannehill, J. C., and Fletcher, R. H., 1984, Computational Fluid Mechanics and Heat Transfer, Hemisphere Publishing, New York. Aref, H., 1986, “The Numerical Experiment in Fluid Mechanics”, Journal of Fluid Mechanics, Vol. 173, pp. 15-41. Armaly, B. F., Durst, R, Pereira, J. C. F., and Schdnung, B., 1983, “Experimental and Theoretical Investigation of Backward-Facing Step Flow”, Journal of Fluid Mechanics, Vol. 127, pp. 473-496. Arpaci, V. 8., and Larsen, P. S., 1984, Convection Heat Transfer, Prentice-Hall, New Jersey. Aziz, K., and Hellums, J. D., 1967, “Numerical Solution of the Three-Dimensional Equations of Motion for Laminar Natural Convection”, Physics of Fluids, Vol. 10, pp. 314 -324. Backus, J. W., 1989, Private Communications, Department of Mechanical Engineering, Michigan State University, East Lansing, MI. 188 189 Barakat, H. 2., and Clark, J. A., 1966, “Analytical and Experimental Study of the Transient laminar Natural Convection Flows in Partially Filled Liquid Containers”, Proceedings of the Third International Heat Transfer Conference, Vol. 2, pp. 152- 162, AIChE, New York. Benjamin, A. S., and Denny, V. B., 1979, “On the Convergence of Numerical Solutions for 2-D Flows in a Cavity at Large Re”, Journal of Computational Physics, Vol. 33, pp. 340-358. Birkhoff, G., and Varga, R. S., 1959, “Implicit Alternating Direction Methods”, Transactions of the American Mathematical Society, Vol. 92, pp. 13-24. Birkhoff, 6., Varga, R. S., and Young, D., 1962, “Alternating Direction Implicit Methods”, Advanced in Computers, F. L. Alt et al., ed, Vol. 3, pp. 189-237, Academic Press, New York. Bodoia, J. R., 1959, "Ihc Finite Difference Analysis of Confined Viscous Flows”, Ph.D. Thesis, Department of Mechanical Engineering, Carnegie Institute of Technology, Pittsburgh, PA. Bodoia, J. R., and Osterle, J. F., 1961, “Finite Difference Analysis of Plane Poiseuille and Couette Flow Developments”, Applied Scientific Research, See. A, Vol. 10, pp. 265-276. Bontoux, P., Gilly, B., and Roux, B., 1980, “Analysis of the Effect of Boundary Conditions on Numerical Stability of Solutions of Navier-Stokes Equations", Journal of Computational Physics, VOL 36, pp. 417-427. Boussinesq, J., 1891, “Sur la martiere dont les vitesses, dans un tube cylindrique dc section circulaire, evasé 8 son entree, se distribuent depuis cette entree jusqu’aux endroits 00 se trouve établi un régime uniftrme”, Comptes Rendus Hebdomadaires des Seances de l'Académie des Sciences, Vol. 113, pp. 9-15; also pp. 49-51. Bozeman, J. D., and Dalton, C., 1973, “Numerical Study of Viscous Flow in a Cavity”, Journal of Computational Physics, Vol. 12, pp. 348-363. Brandt, A., and Gillis, J., 1966, “Magnetohydrodynamics Flow in the Inlet Region of a Straight Channel”, Physics of Fluids, Vol. 9, pp. 690-699. Briley, W. R., 1970, “A Numerical Study of Laminar Separation Bubbles Using the Navier- Stokes Equations”, Report 1110614-1, United Aircraft Research laboratories, East Hartford, CT. Briley, W. R., 1971, “A Numerical Study of laminar Separation Bubbles Using the Navier- Stokes Equations”, Journal of Fluid Mechanics, Vol. 47, pp. 713-736. Burggraf, O. R., 1966, “Analytical and Numerical Studies of the Structure of Steady Separated Flows”, Journal of Fluid Mechanics, Vol. 24, pp. 113-151. Campbell, W. D., and Slattery, J. C., 1963, “Flow in the Entrance of 8 Tube”, ASME Journal of Basic Engineering, Vol. 85, pp. 41-46. Caughey, D. A., and Iyer, R. K., 1989, “Diagonal Implicit Multigrid Calculation of Inlet Flowfields”, AIAA Journal, Vol. 27, pp. 110-112. Chow, C.-Y., 1979, An Introduction to Computational Fluid Mechanics, John Wiley, New York. Christiansen, E. B., and Lemmon, H. B., 1965, “Enhance Region Flow”, AIChE Journal, Vol. ll,pp. 995-999. 190 Collins, M., and Schowalter, W. R., 1962, “laminar Flow in the Inlet Region of a Straight Channel”, Physics of Fluids, Vol. 5, pp. 1122-1124. Collins, M., and Schowalter, W. R., 1963, “Behavior of Non-Newtonian Fluids in the Inlet Region of a Channel”, AIChE Journal, Vol. 9, pp. 98-102. Daube, 0., and Ta Phoc Lee, 1979, “A Mixed Compact Hermitian Method for the Numerical Study of Unsteady Viscous Flow Around an Oscillating Airfoil”, Proceeding of the Third GAMM Conference on Numerical Methods in Fluid Mechanics, E. H. Hirschel, ed., Notes on Numerical Fluid Mechanics, Vol. 2, pp. 56-65, Vieweg and Sohn, Braunschweig. Davis, R. T., 1967, “Laminar Incompressible Flow Past a Semi-Infinite Flat Plate”, Journal of Fluid Mechanics, Vol. 27, pp. 691-704. Dennis, S. C. R., and Chang, G.-Z., 1969, “Numerical Integration of the Navier-Stokes Equations for Steady Two-Dimensional Flow”, Physics of Fluids, VOL 12 (Supplement 11), pp. 11- 88-11-93. Dennis, S. C. R., and Smith, F. '11, 1980, “Steady Flow Through a Channel with a Symmetrical Constriction in the Form of a Step”, Proceedings of the Royal Society of London, Series. A, Vol. 372, pp. 393-414. Deshpande, M. D., Giddens, D. P., and Mabon, R. F., 1976, “Steady laminar Flow Through Modelled Vascular Stcnoses", Journal of Biomechanics, Vol. 9., pp. 165-174. Douglas, J., 1955, “On the Numerical Integration of 32u/3x2 + azu/ayz . au/at by Implicit Methods", Journal ofthe Society oflndustrial and Applied Mathematics, Vol. 3, pp. 42455. Douglas, J., and Rachford, H. H., 1956, “On the Numerical Solution of Heat Conduction Problems in Two and Three Space Variables”, Transactions of the American Mathematical Society, Vol. 82, pp. 421-439. Douglas, J., and Gunn, J. B., 1964, “A General Formulation of Alternating Direction Methods: Part I. Parabolic and Hyperbolic Problems”, Numerische Mathematik, Vol. 6, pp. 428-453. Durst, F., and Pereira, J. C. F., 1988, “lime-Dependent Lamina Backward-Facing Step Flow in a Two-Dimensional Duct", ASME Journal of Fluids Engineering, Vol. 110, pp. 289-296. Duck, P. W., 1982, “Oscillatory flow inside a square cavity”, Journal of Fluid Mechanics, Vol. 122, pp. 215-234. Fairweather, G., and Mitchell, A. R., 1967, “A New Computational Procedure for A.D.I. Methods”, SIAM Journal of Numerical Analysis, Vol. 4, pp. 163-170. Fargie, D., and Martin, B. W., 1971, “Developing laminar Flow in a Pipe of Circular Cross- Section", Proceedings of the Royal Society of London, Series. A, Vol. 321, pp. 461-476. Feliss, N. A., Smith, M. C., and Potter, M. C., 1977, “An Experimental Investigation of Inoompressible Channel Flow Near Transition”, ASME Journal (1' Fluids Engineering, Vol. 99, pp. 693-698. Fasythe, G. E., and Wasow, W. R., 1960, Finite-Diference Methods for Partial Difl'erential Equations, John Wiley, New York. 191 Friedman, M., 1970, “Flow in a Circular Pipe with Recessed Walls”, ASME Journal of Applied Mechanics, Vol. 37, pp. 5-8. Friedman, M., 1972, “Laminar Flow in a Channel with a Step”, Journal of Engineering Mathematics, Vol. 6, pp. 285-290. Friedmann, M., Gillis, J., and Liron, N., 1968, “lamina- Flow in a Pipe at Low and Moderate Reynolds Numbers", Applied Scientific Research, Vol. 19, pp. 426- 438. Frankel, S. P., 1950, “Convergence Rates of Iterative Treatments of Partial Differential Equatiom”, Mathematical Tables and Other Aids to Computation, Vol. 4, pp. 65-75. Fromm, J. B., 1963, “A Method frr Computing Nonsteady, Incompressible, Viscous Fluid Flows”,ReportLA-2910,Los Alamos Scientific laboratory, Los Alamos, NM. Fromm, J., 1964, "The Time Dependent Flow of an Incompressible Viscous Fluid”, Fundamental Methods in Hydrodynamics, B. Alder et al., ed., Methods in Computational Physics, Vol. 3 , pp. 345-382, Academic Press, New York. Fromm, J. B., 1972, “Numerical Solution of the Navies Stokes Equations at High Reynolds Numbers and the Problem of Discretization of Convective Derivatives”, Numerical Methods in Fluid Dynamics, J. J. Smoldern, ed., AGARD Lecnlre Series No. 48, Advisory Group for Aerospace Research and Development, North Atlantic Treaty Organization, pp. 6.1-6.44. Fromm, J. B., and Harlow, F. H., 1963, “Numerical Solution of the Problem of Vortex Street Development”, Physics of Fluids, Vol. 6, p. 975-982. Ghia, U., Ghia, K. N., and Shin, C. T., 1982, “High-Re Solutions for Incompressible Flow Using the Navier-Stokes Equations and a Multigrid Method”, Journal of Computational Physics, Vol. 48, pp. 387-411. Gillis,J.,andBrandt,A.,l964,“1heNurnericallntegrationoftheEquationsofMotionofa Viscous Fluid”, Report SR-1,AFEOAR 63-73,AirForce EmopeanOffrceof Aerospace Research. Goldstein, S., 1965, Modern Developments in Fluid Dynamics, Vol. 1, Dover Publications, New York. Gasman, A. D., Pun, W. M., Runchal, A. K., Spalding, D. B., and Wolfshtein, M., 1969, Heat and Mass Transfer in Recirculating Flows, Academic Press, New York. Greenspan, D., 1965, Introductory Numerical Analysis of Elliptic Boundary Value Problems, Harper and Row, New York. Greenspan, D., 1967, “Numerical Studies of Two Dimensional Steady State Navies-Stokes Equations for Arbitrary Reynolds Number", Report No. 9, Computer Sciences Department, University of Wisconsin, Madison, WI. Greenspan, D., 1968, “Numerical Studies of Viscous, Incompressible Flow for Arbitrary Reynolds Number”, Report No. 11, Computer Sciences Department, University of Wisconsin, Madison, WI. Greenspan, D. J., 1969a. “Numerical Studies of Steady, Viscous, Incompressible Flow in a Channel with a Step”, Journal of Engineering Mathematics, Vol. 3, pp. 21-28. 192 Greenspan, D. J., 1969b, “Numerical Studies of Prototype Cavity Flow Problem”, The Computer Journal, Vol. 12, pp. 88-93. Greenspan, D. J., 1974, Discrete Numerical Methods in Physics and Engineering, Academic Press, New York. Gupta, M. M., and Manohar, R. P., 1979, “Bormdary Approximations and Accuracy in Viscous Flow Computations”, Journal of Computational Physics, Vol. 31, pp. 265-288. Harlow, F. H., and Fromm, J. E., 1965, “Computer Experiments in Fluid Dynamics”, Scientific America, Vol. 212, pp. 104-110. Hanbeck, R. W., 1964, “Laminar Flow in the Entrance Region of a Pipe”, Applied Scientific Research, See. A, Vol. 13, pp. 224-232. Hwang, C.-L., and Fan, L.-T., 1963, “A Finite Difference Analysis of Laminar Magneto- Hydrodynamic Flow in the Entrance Region of a Flat Rectangular Duct”, Applied Scientfic Research, Sec. B, Vol. 10, pp. 329-343. Jenson, V. G., 1959, “Viscous Flow Round a Sphere at Low Reynolds Numbers (<40)”, Proceedings of the Royal Society of London, Series A, Vol. 249, pp. 346-366. Kawaguti, M., 1961, “Numerical Solution of the Navier-Stokes Equations for the Flow in a Two-Dimensional Cavity”, Journal of the Physical Society of Japan, Vol. 16, pp. 2307-2315. Khosla, P. K., and Rubin, S. G., 1974, “A Diagonally Dominant Second-Order Accruate Implicit Scheme", Computers and Fluids, Vol. 2, pp. 207-209. Krrmae, B., 1985, “Computational Fluid Dynamics: Its Present Status and Future Direction”, Computers and Fluids, Vol. 13, pp. 239-269. Km, A., and Yajnik, K. S., 1980, “Internal Separated Flows at Large Reynolrk Number”, Journal of Fluid Mechanics, Vol. 97, pp. 27-51. Kuskova, T. V., 1967, “Difference Method for Calculating the Flow of a Viscous Fluid” (in Russian), Sbornik Robot Vychislitel’nogo Tsentra Moskovshogo Gosudarstvennogo Universiteta, Vol. 7, pp. 16-27; also 1969. APPlied Mechanics Reviews (Review No. 380), Vol. 22, p. 61. Kwon, 0., Pletcher, R. H., and Delaney, R. A., 1988, “Solution Procedure for Unsteady Two- Dimensional Boundary Layers”, ASME Journal of Fluids Engineering, Vol. 110, pp. 69-75. Lsnghmr, H. 1... 1942, “Steady Flow in the Transition Length of a Straight Tube”, ASME Journal of Applied Mechanics, Vol. 9, pp. ASS-A58. Leal, L. G., and Acrivos, A., 1969, “Structme of Steady Closed Streamline Flows within a Bormdary Layer”, Physics of Fluids, Vol. 12 (Supplement 11), pp. II-105 -II-l l3. Lew, H. 8., and Fung, Y. C., 1970, “Entry Flow into Blood Vessels at Arbitrary Reynolds Number", Journal of Biomechanics, Vol. 3, pp. 23-38. Li, L. C., and Ludford, G. S. S., 1980, “The Overshoot in Entry Flow”, Archives of Mechanics, Vol. 32, pp. 741-746. 193 Lilly, D. IL, 1965, “On the Computational Stability of Numerical Solutions of Time- Dependent Non-Linear Geophysical Fluid Dynamics Problems", Monthly Weather Review, Vol. 93, pp. 11-26. erdgren, T. S., Atabek, B. H., and Chang, C. C., 1961, “Transient Magnetohydrodynarnic Duct Flow”, Physics of Fluids, Vol. 4, pp. 1m6-1011. erdgren, T. S., Sparrow, E. M., and Starr, J. B.,1964, “Pressure Drop Due to the Entrance Region in Ducts of Arbitrary Cross Section”, ASME Journal of Basic Engineering, Vol. 86, pp. 620- 626. Lynch, R. E., and Rice, J. R., 1968, “Convergence Rates of ADI Methods with Smooth Initial Error”, Mathematics of Computation, Vol. 22, pp. 311-335. Mallinson, G. D., and de Vahl Davis, 6., 1973, "The Method of False Transient for the Solution of Coupled Elliptic Equations”, Journal of Computational Physics, Vol. 12, pp. 435-461. McDonald, J. W., Denny, V. E., and Mills, A. F., 1972, “Numerical Solutions of the Navier- Stokes Equations in Inlet Regions", ASME Journal of Applied Mechanics, Vol. 39, pp. 873-878. Mei, R., and Plotkin, A., 1986, “A Finite-Difference Scheme for the Solution of the Steady Navier-Stokes Equations", Computers and Fluids, Vol. 14, pp. 239-251. Mills, R. D., 1965, “Numerical Solutions of the Viscous Flow Equations for a Class of Closed Flows”, Journal of the Royal Aeronautical Society, Vol. 69, pp. 714-718; also correction p. 880. Mitchell, A. R., 1969, Computational Methods in Pam'al Dfi‘erential Equations, John Wiley, New York. Mitchell,A.R.,andFairweather,G., 1964,“Im}rovedFormsoftheAlternatiugDirection Methods of Douglas, Peaceman, and Rachford for Solving Parabolic and Elliptic Equations”, Numerische Mathematik, Vol. 6, pp. 285-292. Mohanty, A. K., and Asthana, S. B. L.,1978, “laminar Flow in the Entrance Region of a Smooth Pipe, Journal of Fluid Mechanics, Vol. 90, pp. 433-447. Maihara, H., and Cheng, R. T., 1973, “Numerical Solution of the Viscom Flow in the Enhance Region of Parallel Plates”, Journal of Computational Physics, Vol.11, pp. 550-572. Morris, D. J., 1975, “Solution of the Incompressible Driven Cavity Problem by the Alternating-Direction Implicit Method”, Numerical Studies of Incompressible Viscous Flow in a Driven Cavity, NASA SP-378, National Aeronautics and Space Administration, washinsm D.C., pp. 47-60. Nallasamy, M., 1986, “Numerical Solution of the Separating Flow Due to an Obstruction”, Computers and Fluids, Vol. 14, pp. 59-68. Orszag, S. A., and Israeli, M., 1974, “Numerical Simulation of Viscous lncornpressible Flows”, Annual Review of Fluid Mechanics, Vol. 6, pp. 281-318. Pan, F., and Acrivos, A., 1967, “Steady Flows in Rectangula' Cavities”, Journal of Fluid Mechanics, Vol. 28, pp. 643-655. 194 Pao, Y.-I-1., and Daugherty, R. J ., 1969, “Tune-Dependent Viscous Incompessible Flow Past a Finite Flat Plate”, Report D1-82-0822, Boeing Scientific Research Iaborataies, Seattle, WA. Paris, J., and Whitaker, S., 1965, “Confined Wakes' A Numerical Solution of the Navier- Stokes Equations”, AIChE Journal, Vol. 11, pp. 1033-1041. Peaceman, D. W., and Rachford, H. H., Jr., 1955, "The Numerical Solution of Parabolic and Elliptic Differential Equations”, Journal of the Society of Industrial and Applied Mathematics, Vol. 3, pp. 28-41. Pearson, C. E., 1964, “A Coruputational Method for Time-Dependent Two-Dimmsional Incompressible Viscous Flow Problems”, Report SRRC-RR-64-17, Sperry Rand Research Center, Subdtu'y, MA. Pearson, C. E., 1965a, “A Computational Method Fa Viscous Flow Problems", Journal of Fluid Mechanics, Vol. 21, pp. 611-622. Pearson, C. E., 1965b, “Numerical Solutions for the Time-Dependent Viscous Flow Between Two RotatingCoaxialDisks”, Journal of Fluid Mechanics, Vol. 21,pp. 623-633. Peyret, R., and Taylor, T. D., 1983, Computational Methods for Fluid Flow, Springer-Verlag, New York. Peyret, R., 1971, “Sun 1e calcul numerique des 6coulements stationnaires d’un fluide visqueux incompressible" Comptes Rendus Hebdomadaires des Seances de l'Academie des Sciences, Série A, Vol. 272, pp. 1274-1277. Prandtl, I... and Tietjens, O. G., 1934, Applied Hydro- and Aeromechanics, Dover Publications, New York. Richtmyer, R. D., lid Morton, K. W., 1967, Dw'erence Methods for Initial-Value Problems. 2nd Ed., Intaacience Publishers, New York. Roache, P. J, 1975, “The LAD, NOS and Split NOS Methods for the Steady-State Navier- Stokes Equations", Computers and Fluids, Vol. 3, pp. 179-195. Roache, P. J, 1982, Computational Fluid Dynamics, Revised Ed., Hermosa Publishes, Albuquerque, NM. Roidt, M, and Cess, R. D., 1962, “An Approximate Analysis of laminar Magnetohymodynamic Flow in the Entrance Region of a Flat Duct”, ASME Journal ofth Mechanics, Vol. 29, pp. 171-176. Rubin, S. G., and Harris, J. E., 1975 “Incompressible Viscous Flow in a Driven Cavity”, Numerical Studies of Incompressible Viscous Flow in a Driven Cavity, NASA SP-378, National AeronauticsandSpace Administration, Washington, D.C., pp. 1-6. Rubin, S. G., and Khosla, P. K., 1977, “Polynomial Interpolation Methods for Viscous Flow Calculations”, Journal of Computational Physics, Vol. 24, pp. 217-244. Runchal, A. K., Spalding, D. B., and Wolfshtein, M., 1969, “Numerical Solution of Elliptic Equations for Transport of Vorticity, Heat, and Matter in Two-Dimensional Flow”, Physics (1‘ Fluids, Vol. 12 (Supplement 11), pp. II-21-II-28. 195 Samuels, M. R., and Chmchill, S. W., 1967, “Stability of a Fluid in a Rectangular Region Heated from Below”, AIChE Journal, Vol. 13, pp. 77-85. Schiller, L., 1922, “Die Entwicklrmg der laminaren Geschwindigkeitsverteilung und ilue Bedeutung fur Zhhigkeitsmessungen” Zeitschrifl filr angewandte Mathematik und Mechanih, Vol. 2, pp. 96-106. Schlichting, H., 1934, “Lmninare Kanaleinlaufstrolnung", Zeitschrift fiZr angewandte Mathematik und Mechanik, Vol. 14, pp. 368-373. Schlichting, H., 1979, Boundary-Layer Theory, 7th Ed., McGraw-Hill, New York. Schmidt, F. W., and Zeldin, B., 1969, "Laminar Flows in Inlet Sections of Tubes and Ducts”, AIChE Journal, Vol. 15, pp. 612-614. Son,J.S.,andHanratty,T.J.,1969,“NtunericaISolution£tn'theFlowArormdaCylinderat Reynolds Numbers of 40, 200 and 500”, Journal of Fluid Mechanics, Vol. 35, pp.369-386. Smith, F. T., 1977, “Upstream Interaction in Channel Flows”, Journal of Fluid Mechanics, Vol. 79, pp.631-655. Smith, F. T., 1979, "The Separating Flow Through a Severely Constricted Syrrunetric Tube”, Journal of Fluid Mechanics, Vol. 90, pp.725-754. Sparrow, E. M., Anderson, C. B., 1977, “Effect of Upstream Flow Processes on Hydrodynamic Development in a Duct”, ASME Journal of Fluids Engineering, Vol. 99, pp. 556-560. Sparrow, E. M., Hixon, C. W., and Shavit, G., 1967, “Experiments on Laminar Flow Development in Rectangular Ducts”, ASME Journal (1 Basic Engineering, Vol. 89, pp. 116124. Sparrow, E. M., Lin, S. H., and Lundgren, T. S., 1964, “Flow Development in the Hydrodynamic Enhance Region of Tubes and Ducts”, Physics of Fluids, Vol. 7, pp. 333-347. Ta Phoc Loc, and Daube, 0., 1981, “Higher Order Accurate Numerical Solution of Unsteady Viscous Flow Generated by a Transversely Oscillating Ellipt'u: Cylinder”, Vortex Flows, W. L. Swift et al., ed., pp. 155-171, ASME, New York. Taylor, P. J., 1970, “The Linear Stability of No-Slip Boundary Conditiom in the Numerical Solution of Nonsteady Fluid Flows”, Journal of Computational Physics, Vol. 6, pp. 268-387. Taylor, T. D., and Ndefo, E., 1971, “Computation of Viscous Flow in a Charurel by the Method of Splitting”, Proceedings of the Second International Conference on Numerical Methods in Fluid Dynamics, M. Holt et al., ed., Lecture Notes in Physics, Vol. 8, pp. 356-364, Stringer- Verlag, Berlin. Taylor, T. D., and Ndefo, B., and Masson, B. S., 1972, “A study of Numerical Methods for Solving Viscousandlnviscid Flow Problems”, Journal of Computational Physics, Vol. 9, pp. 99-119. Thom, A., 1933, “The Flow Past Circula' Cylinders at Low Speeds”, Proceedings of the Royal Society of London, Series. A, Vol. 141, pp. 651-669. Thom, A., and Apelt. C. J., 1961, Field Computation in Engineering and Physics, Van Nostrand, London. 196 Thompson, J. F., 1969, “Optimized Acceleration of Convergence of an Implicit Numerical Solution of the Time-Dependent Navier-Stokes Equations”, AlAA Journal, Vol. 7, pp. 2186-2188. Torrance, K. E., 1968, “Comparison of Entire-Difference Computations of Natmal Convection”, Journal of Research of the National Bureau of Standards, Vol. 72B, pp. 281- 301. Tuann, S.-Y., and M. D. Olson, 1978, “Review of Computing Methods for Recirculating Flows”, Journal of Computational Physics, Vol. 29, pp. 1-19. Van Dyke, M., 1970, “Entry Flow in a Cinnuel", Journal of Fluid Mechanics, Vol. 44, pp. 813-823. Varga, R. S., 1962, Matrix Iterative Analysis, Prentice-Hall, New Jersey. Vrentas, J. S., Duds, J. L., and Bargeron, K. G., 1966, “Effect of Axial Diffusion of Va'ticity on Flow Development in Circular Conduits: Part 1. Numerical Solutions”, AIChE Journal, Vol. 12, pp. 837-844. Wachspress, E. L., 1966, Iterative Solution of Elliptic Systems, Prentice-Hall, New Jersey. Wang, Y. L., and Longwell, P. A., 1964, “Laminar Flow in the Inlet Section of Parallel Plates”, AIChE Journal, Vol. 10, pp. 323-329. Wiginton, C. L., and Dalton, C., 1970, “Incompresm’ble laminar Flow in the Entrance Region of a Rectangular Duct”, ASME Journal of Applied Mechanics, Vol. 37, pp. 854-856. Wiginton, C. L., and Wendt, R. L., 1969, “Flow in the Entrance Region of Ducts”, Physics of Fluids, Vol. 12, pp. 465-466. Wilkes, J. 0., and Chmchill, S. W., 1966, “The Finite Difference Computation of Natural Convection in a Rectangular Enclosrue”, AIChE Journal, Vol. 12, pp. 161-166. Wilson, 8., 1969, “The Development of Poiseuille Flow”, Journal of Fluid Mechanics, Vol. 38, pp. 793-806. Wilson, 8. D. R., 1971, “Entry Flow in a Channel. Part 2”, Journal of Fluid Mechanics, VOL 46, pp. 787-799. Woods, L. C., 1954, “A Note on the Numerical Solution of Foruth Order Differential Equations”, The Aeronautical Quarterly, Vol. 5, pp. 176-184. Yanenko, N. N., 1971, The Method of Fractional Steps: The Solution of Problems of Mathematical Physics in Several Variables, Springer-Verlag, New Yak. Young, D., 1954, “Iterative Methods for Solving Partial Difference Equations of Elliptic Type”, Transactions of the American Mathematical Society, Vol. 76, pp. 92-111.