$311!!!!“ .535’5‘ AS. .. 1. Es“ .93. 7.. {Azrwfivfi . and. 1.. a}... it... . _ due? .. . . :2...- 3:1 3 17.: ... . e2 .2: six . t... . . A ‘h »>v‘i. 1.3 .5»... IIA.-v3»..~io I!«Ir. .........n...nn..m.i Elihu?! “M11... 2 .. .3. 3 ... $.11 . 1. {623.1 ;. , $....u..£.ifi.§.§..rfi. flxnmxizbai “Sire... 59233.1... $311.1 1:13:11: 11.: 1. :23. 1...»? $3.15»: 5:. i . u i}. (1.1. 0"“ . .1. z , {a “w .w..... L. .5. 1......,F....u...".u.1; .4 .An:l' an a v.15 crufin . n: .1. .50.”. 3. 2.. “have. ‘3". .rl. .L .a. 1.4 . 1 .Y . .. 2.1.1:. .. e. 11...??? .. .. ammmfi .1... 5..., 1. «3.. if. It I 1.. 1.. 5...... i.h§.as::-.:. .th... 3V.§.3lu..o“¢.§'1§ 5C. a Run}... ......u._.........w. . 1.1.». 921‘... 1.19.3131... I. 311...}...{nfls I: .an 15:13. {tinijatcit \ .{....c......13 8:... . Ea \ o .1..«. .3113...) 3 3...: ¢ 1. 25.1.1.1. 3. 1:73. . :3. a. 3"!!!1 l 3’3 L12... ..a..\i? . mime - LIBRARY 9.00} Michigan State University This is to certify that the dissertation entitled PREDICTION OF RHEOLOGICAL PROPERTIES OF STRUCTURED FLUIDS IN HOMOGENEOUS SHEAR BASED ON A REALIZABLE MODEL FOR THE ORIENTATION DYAD presented by YoChan Kim has been accepted towards fulfillment of the requirements for the Ph.D degree in Chemical Engineering and Materials Science Qjawm Pm; Major Professor’ 5 Signature November 1St , 2006 Date MSU is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN Box to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2105 p:/ClRC/DaleDue.indd-p.1 PREDICTION OF RHEOLOGICAL PROPERTIES OF STRUCTURED FLUIDS IN HOMOGENEOUS SHEAR BASED ON A REALIZABLE MODEL FOR THE ORIENTATION DYAD By YoChan Kim A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering and Materials Science 2006 ABSTRACT PREDICTION OF RHEOLOGICAL PROPERTIES OF RIGID ROD FLUIDS IN SIMPLE HOMOGENEOUS SHEAR FLOWS BASED ON A REALIZABLE MODEL FOR THE ORIENTATION DYAD By YoChan Kim Non-spherical particles dispersed in a fluid have a tendency to align in shear flows because of particle-fluid drag. This phenomenon is opposed by rotary diffusion. At high concentrations and in the absence of hydrodynamic couples, self-alignment can also occur because excluded volume forces prevent the retum-to-isotropy of anisotropic states by rotary Brownian motion. The balance between microhydrodynamic and diffusive (i.c., Brownian and excluded volume) torques at the microscale has a direct impact on the rheological properties of rigid rod fluids (particulate suspensions and liquid crystalline polymers) at the continuum scale. Over the past sixty years, important characteristics of the microstructure associated with the foregoing alignment phenomenon have been quantified in terms of the low order moments Of the orientation density function governed by the rotary Smoluchowski equation. In this research, a closed model for the second order moment < pp > (orientation dyad) has been identified based on the condition that in the absence of an external field all realizable anisotropic states must relax to stable equilibrium states. A key step in the development of the new closure is the use of an algebraic pre-closure for the orientation tetrad < pppp > in terms of the orientation dyad < pp> that preserves the six-fold symmetry and contraction properties of the original orientation tetrad. In the presence of a simple shear flow, the microstructure and the rheological characteristics predicted for rigid-rod fluids agree with previous theoretical and experimental results for a wide range of Péclet numbers. In addition to the Péclet number (i.c., Pe :-:||Vg||/(6D;) ), the orientation director also depends on three other dimensionless groups: a tumbling parameter, A; an excluded volume coefficient, U; and, a dimensionless time t a 6 D: t. The rotary diffusion coefficient for dilute solutions, D; , is used to scale time. Unlike other closure models, the approach developed hereinafter predicts that all two-dimensional and three-dirnensional realizable anisotropic states relax to either a steady state (isotropic or anisotropic) or a periodic state, depending on Pe, A, and U. The model predicts the existence of shear thinning and shear'thickening phenomena, Newtonian plateau regions at low and high Péclet numbers, positive (and negative) first normal stress differences, and negative (and positive) second normal stress differences. For Pe = 0 , multiple equilibrium states exist for 4.72 S U S 5.00 . For Pe > O and initial directors located in the flow-deformation plane, the predominant feature for U < 25 is the existence of a unique nematic-like microstructure with a steady alignment of the director that becomes completely aligned with the velocity as Fe —) 00. For A < l and U > 25, tumbling and wagging of the director occur at low to moderate values of the Péclet number. If the initial director has a component in the direction of the vorticity, then director kayaking and director log-rolling may occur. The coexistence of stable anisotropic states (or texture) predicted by the model may provide an explanation of why micro defects occur during the processing of some structured fluids. To My Family for Love, Support, and Patience iv ACKNOWLEDGEMENT First of all, I would like to thank Professor Charles A. Petty for his support and encouragement throughout my Ph.D. program. I truly believe that I would not have completed my program without him. He has not only inspired me in many aspects from an academic perspective, but has also given me personal guidance in my twenties. His suggestion to participate in two internships helped me greatly. The NSF summer program in Japan helped me broaden my world perspective. The summer internship at Bechtel National, Inc. inspired me to become a better engineer. In addition, I want to acknowledge Professors Peter W. Bates, Andre Bénard, Krisnamurthy Jayaraman, and Michael E. Mackay for serving on my dissertation committee. I would also like to recognize Professor Jun-ichi Takimoto fi'om Nagoya University (now at Yamaguchi University) for giving me guidance during my research. I also want to recognize my gradate school colleagues Dr. Steven Parks, Chinh Nguyen, Hemant Kini, and Dilip Manda] for the inspiration of this research. I thank also Jon Berkoe, Kelly Knight, Lin Lorraine, and Brigette Rosendall of Bechtel National, Inc for their wonderful support. I gratefully acknowledge financial support of this work by the National Science Foundation and by Michigan State University. Finally, I want to express a very special thanks to my wife Seiyoun Jung and my parents for unlimited love and moral support. TABLE OF CONTENTS TABLE OF CONTENTS ................................................................................................... vi LIST OF TABLES ............................................................................................................. ix LIST OF FIGURES ............................................................................................................ x NOTATION ..................................................................................................................... xvi Chapter 1. INTRODUCTION 1.1 Motivation .......................................................................................................... l 1.2 Background ........................................................................................................ 4 1.3 Objective .......................................................................................................... 14 1.4 Outline .............................................................................................................. 16 2. SMOLUCHOWSKI EQUATION FOR RIGID ROD SUSPENSIONS 2.1 2.2 2.3 2.4 2.5 Introduction ...................................................................................................... 17 Jeffery’s Model for Rotary Convection: Tumbling Coefficient ...................... 18 Brownian Motion: Rotary Diffusivity ............................................................. 20 Excluded Volume Phenomena: Maier-Saupe Potential ................................... 21 Discussion ........................................................................................................ 23 3. MOMENTS OF THE ORIENTATION DISTRIBUTION 3.1 3.2 3.3 3.4 3.5 Introduction ...................................................................................................... 24 Orientation Dyad: Structure Tensor and the Order Parameter ......................... 26 Orientation Tetrad: Symmetry and Contraction Properties ............................. 28 Realizable Anisotropic States: Invariant Diagram ........................................... 29 Discussion ........................................................................................................ 32 4. EQUATIONS FOR THE MICROSTRUCTURE AND THE STRESS 4.1 4.2 4.3 4.4 4.5 Introduction ...................................................................................................... 33 Dynamic Equation for the Orientation Dyad ................................................... 33 Dynamic Equations for the Structure Invariants for Pe = O ............................. 34 Algebraic Equation for the Stress .................................................................... 35 Discussion ........................................................................................................ 36 5. CLOSURE FOR THE ORIENTATION TETRAD 5.1 5.2 5.3 Introduction ...................................................................................................... 38 Closure Models ................................................................................................ 40 Discussion ........................................................................................................ 44 6. REALIZABLE CLOSURE 6.1 6.2 6.3 6.4 6.5 Introduction ...................................................................................................... 45 Realizable Isotropic, Planar Isotropic, and Nematic States ............................. 46 Realizable Prolate and Oblate States ............................................................... 50 Realizable Planar Anisotropic Boundary ......................................................... 51 Discussion ........................................................................................................ 53 7. MICROSTRUCTURE IN THE ABSENCE OF AN EXTERNAL FIELD 7.1 7.2 7.3 7.4 Introduction ...................................................................................................... 56 Biphasic Phenomena ........................................................................................ 57 Relaxation to Isotropic and Anisotropic Steady States .................................... 62 Discussion ........................................................................................................ 65 8. MICROSTRUCTURE INDUCED BY HOMOGENEOUS SHEAR 8.1 8.2 8.3 8.4 8.5 8.6 Introduction ...................................................................................................... 72 Relaxation of Planar Anisotropic States for L/d = 00 ....................................... 73 Relaxation of Planar Anisotropic States for L/d g 12 ...................................... 76 Relaxation of Planar Anisotropic States for O S L/d < oo ................................. 93 The Effect of Tube Dilation on the Relaxation of Planar Anisotropic States ................................................................................................................ 96 Discussion ...................................................................................................... 100 vii 9. VISCOSITY AND NORMAL STRESS DIFFERENCES 9.1 Introduction .................................................................................................... 102 9.2 Rheological Properties: L/d = 00 .................................................................... 103 9.3 Rheological Properties: L/d s 12 .................................................................. 117 9.4 Rheological Properties: 0 _<. L/d < oo .............................................................. 129 9.5 Rheological Properties: Effect of Tube Dilation ........................................... 133 9.6 Discussion ...................................................................................................... 143 10. CONCLUSIONS ..................................................................................................... 145 1 1. RECOMMENDATIONS ........................................................................................ 154 APPENDICES APPENDDC A. Derivation of Moment Equations with Structure Tensor ...................... 161 APPENDIX B. Derivation of Moment Equations in Invariant Form ............................. 165 APPENDIX C. Normal Vectors in the Invariant Diagram ............................................. 169 APPENDIX D. Derivation of Various Closure Analysis in the Invariant Form ............ 172 APPENDIX E. Eigenvalues and Eigenvectors of the Orientation Dyad ........................ 182 APPENDIX F. Eigenvalues and Eigenvectors of the Orientation Dyad ........................ 189 APPENDD( G. Computational Code for Transient Calculations ................................... 189 LIST OF REFERENCES ................................................................................................. 199 viii LIST OF TABLES Table 7.1 Invariants and Order Parameter of the Equilibrium Structure Tensor on the Prolate Line ........................................................................................ 63 Table 7.2 Invariants and Order Parameter of the Equilibrium Structure Tensor on the Oblate Line ....................................................................................................... 64 Table 8.1 Invariants of the Equilibrium Anisotropic Tensor for Different Tumbling Parameters (F SQ-model; FTD = l; U = 0) ........................................................ 77 Table 8.2 Effect of l on the Tumbling Period (ESQ-model; FTD = 1; U = 27; Pe = 10) .......................................................... 98 Table 9.1 Viscous and Elastic Contributions to the Shear Viscosity and the First and Second Normal Stress Differences at Selected Values of U and Pe for L/d = 00 ........................................................................................................... 1 13 Table 9.2 Viscous and Elastic Contributions to the Shear Viscosity and the First and Second Normal Stress Differences at Selected Values of U and Pe for L/d g 12 .......................................................................................................... 124 Table 9.3 Legend for Figures 9.19 — 9.23 ....................................................................... 135 LIST OF FIGURES Figure 1.1 Anisotropic Orientation States (Parks, 1997) ..................................................... 3 Figure 1.2 Molecular Structure of LPC (Larson, 1999) ....................................................... 6 Figure 1.3 Schematic of NMR Spectra for PBLG in DMF (Abe and Yamazaki, 1987a; Robinson, 1966) ................................................... 8 Figure 3.1 Instantanous Orientation Vector Directed Along the Molecular Axis of a Rigid Rod ................................................................................................................ 25 Figure 3.2 Three Orientation States for the Orientation Dyad .......................................... 27 Figure 6.1 Closure Coefficient C2 ( IIIb) for the Realizable FSQ-model ........................ 54 Figure 6.2 The Excluded Volume Potential for the Decoupling Approximation and the FSQ-model ....................................................................................................... 55 Figure 7.1 Multiple Equilibrium States Predicted by the FSQ-model for C2 = constant ......................................................................................................................... 59 Figure 7.2 Multiple Equilibrium States Predicted by the FSQ-model for C2 = C2 (IIIb) ........................................................................................................................ 61 Figure 7 .3 Relaxation of the Microstructure due to Rotary Brownian Diffirsion (U=0; i=1/(6D§)) ....................................................................................... 66 Figure 7.4 Effect of Excluded Volume on the Relaxation of the Microstructure (F SQ-model ; FTD = l) ................................................................................... 67 Figure 7.5 The Effect of Tube Dilation on the Relaxation of the Microstructure (F SQ-model; U = 3; initial conditions: 11b (0) = 2/9; IIIb (0) = 0). ............... 68 Figure 7.6 The Effect of U and Tube Dilation on the Characteristic Relaxation Time tc (F SQ-model; (1(0) = 0.9). ................................................................................. 69 Figure 8.1 The Effect of U and Pe on the Steady State Director Angle for L/d = 00 (FSQ-model; FTD = 1; A = 1) .......................................................................... 74 Figure 8.2 The Effect of U and Pe on the Steady State Microstructure for L/d = 00 (FSQ-model; FTD = 1; 7» = 1) .......................................................................... 75 Figure 8.3 Instantaneous Microstructure for Director Tumbling for L/d s 12 (F SQ-model; FTD = l; I» = 0.987; U = 27, Pe = 10; 2(0)=§y_ey+gz§z) ............................................................................ 78 Figure 8.4 Microstructure for Director Tumbling in the Phase Plane for L/d s 12 (F SQ-model; FTD = 1; 7t = 0.987; U = 27, Pe = 10; 2(0)=gygy+ez§z) ............................................................................ 79 Figure 8.5 Instantaneous Microstructure for Director Wagging for L/d E 12 (F SQ-model; Fm = 1; X = 0.987; U = 27, Pe = 24; 2(0)=§ygy+e_zgz) ............................................................................ 81 Figure 8.6 Microstructure for Director Wagging in the Phase Plane for L/d s 12 (F SQ-model; FTD = 1; A = 0.987; U = 27, Pe = 24; 2 < 22 > (0) =§y9y +5225:z ......................................................................................................................... 82 Figure 8.7 Instantaneous Microstructure for Director Steady Alignment for L/d s 12 (F SQ-model; Fm = l; l = 0.987; U = 27, Fe = 95; 2(0)=ey_ey+gzgz) ............................................................................ 84 Figure 8.8 Microstructure for Director Steady Alignment in the Phase Plane for L/d a 12 A (F SQ-model; Fm = 1; A. = 0.987; U = 27, Pe = 95; 2(0)=ey_e_y+gzgz ............................................................................ 85 Figure 8.9 Phase Diagram of U and Pe for L/d s 12 (FSQ-model; Fm =1;l=0.987; 2<_'p_p>(0)=§;ye_y +9292) .................... 86 Figure 8.10 Instantaneous Director Log—rolling for L/d 5 12 (F SQ-model; U = 27; Pe = 95; Fm = l; X = 0.987; 2(0)=_e_x§x+gz_ez) ............................................................................ 87 Figure 8.11 Director Log-rolling in the Phase Plane L/d E 12 (F SQ-model; U = 27; Pe = 95; Fm = 1; X = 0.987; 2(0)=§x§x +2292) ............................................................................ 88 Figure 8.12 Phase Diagram of Shear-voriticity Plane Initial Condition L/d s 12 (FSQ-model;F1-D =1;7t=0.987; 2(0)=§xgx+§zgz) ................... 90 Figure 8.13 Instantaneous Director Kayaking L/d s 12 (F SQ-model; U = 27; Pe = 95; Fm = 1; 7L = 0.987) ...................................... 92 Figure 8.14 Microstructures for Director Tumbling and Steady State Alignment in the Phase Plane (FSQ-model; FTD = 1;)» = 0.5; U = 27, Pe = 95; 2(0)=§y§y +£2.92) ............................................................................ 94 Figure 8.15 Microstructures for Director Tumbling and Steady State Alignment in the Phase Plane (FSQ-model; Fm =1;X=0;U=27,Pe=95; 2(0)=§y_e_y+§z_ez ............................................................................ 95 xii Figure 8.16 Microstructure for Director Tumbling and Steady State Alignment in the Phase Plane (FSQ-model;F1-D =1;}t = 0; U= 0, Pe=95; 2(0)=§y§y+§z§z) ............................................................................ 97 Figure 8.17 The Effect of Tube Dilation on the Phase Diagram (FSQ-model; Fm = 1; A= 0.987; 2(0)=e_y +£7.92 .................... 99 9y Figure 9.1 The Effect of U and Pe on the Shear Viscosity for L/d = 00 (FSQ-model;Fm= 1; 1= 1) .......................................................................... 104 Figure 9.2 The Effect of U and Pe on the Viscous and Elastic Components of the Shear Viscosity for L/d = 00 (FSQ-model; Fm = 1; A = 1). ........................................................................ 106 Figure 9.3 The Effect of U and Pe on the Brownian and Excluded Volume Contributions to the Shear Elastic Component of Viscosity for L/d = 00 (FSQ-model;F1-D= 1; 7t= 1) .......................................................................... 108 Figure 9.4 Elastic Contributions to the Shear Stress (FSQ-model; 7» =1; Fm =1) ........................................................................ 110 Figure 9.5 The Effect of U and Pe on the First Normal Stress Difference for L/d = 00 (FSQ-model; Fm = 1; k= 1). ..................................................................... 112 Figure 9.6 The Effect of U and Pe on the Second Normal Stress Difference for L/d = 00 (FSQ-model;Fm= l;7\.= 1) .......................................................................... 115 Figure 9.7 The Contributions of Stress on the Time Averaged Normal Stress Differences (FSQ-model; FTD =1;X=1;U=27). ........................................................ 116 Figure 9.8 The Effect of Tumbling Parameter on the Shear Viscosity for U = 0 (FSQ-model; Fm = 1). ................................................................................ 118 xiii Figure 9.9 The Instantaneous Shear Viscosity for Director Tumbling (FSQ-model; Fm = 1; k = 0.987; U = 27, Pe = 10). .................................... 119 Figure 9.10 The Effect of U and Pe on the Time Averaged Shear Viscosity for L/d 5 12 (FSQ-model; Fm = 1; x = 0.987) ................................................................. 121 Figure 9.11 The Contributions of Viscous and Elastic Stresses on the Time Averaged Shear Viscosity for L/d E 12 (F SQ-model; FTD = 1; A = 0.987; U = 27). ................................................. 123 Figure 9.12 The Instantaneous First Normal Stress Difference for Director Tumbling (FSQ-model; Fm = 1; A = 0.987; U = 27, Pe = 10). .................................... 125 Figure 9.13 The Effect of U and Pe on the Time Averaged First Normal Stress Difference for L/d E 12 (FSQ-model; Fm = 1; A = 0.987). ............................................................... 126 Figure 9.14 The Instantaneous Second Normal Stress Difference for Director Tumbling with L/d 5 12 (F SQ-model; Fm = 1; k = 0.987; U = 27, Pe = 10). .................................... 128 Figure 9.15 The Effect of U and Pe on the Time Averaged Second Normal Stress Difference for L/d s 12 (FSQ-model; Fm = 1; A = 0.987) ................................................................. 130 Figure 9.16 The Effect of Tumbling Parameter on the Time Averaged Shear Viscosity (FSQ-model; Fm =1;U=27,Pe=10). ...................................................... 131 Figure 9.17 The Effect of Tumbling Parameter on the Time Averaged First Normal Stress Difference (FSQ-model; Fm = 1; U = 27, Pe = 10). ...................................................... 132 Figure 9.18 The Effect of Tumbling Parameter on the Time Averaged Second Normal Stress Difference (FSQ-model; Fm = 1; U = 27, Pe = 10). ...................................................... 134 xiv Figure 9.19 The Effect of Tube Dilation on the Time Averaged Shear Viscosity (F SQ-model; it = 0.987; U = 27 see Table 9.3 for legend). .......................... 137 Figure 9.20 The Contributions of Stress on the Time Averaged Shear Viscosity (FSQ-model; F196" in qu.(4.1) and (4.6); 7. = 0.987; U = 27). ................... 138 Figure 9.21 The Effect of Tube Dilation on the Time Averaged First Normal Stress Difference (F SQ-model; 7L = 0.987; U = 27 see Table 9.3 for legend) ............................ 139 Figure 9.22 The Effect of Tube Dilation on the Time Averaged Second Normal Stress Difference (F SQ-model; 7» = 0.987; U = 27 see Table 9.3 for legend). .......................... 141 Figure 9.23 The Contributions of Stress on the Time Averaged Normal Stress Differences (FSQ-model; F193 ; 7. = 0.987; U = 27). ...................................................... 142 Figure 10.1 Validation of the FSQ-closure ...................................................................... 147 Figure 11.1 The Effect of Particle Aspect Ratio on the Dimensionless Friction Coefficient ...................................................................................................... 156 Figure 11.2 Multiple Stable Steady States for Various Initial Orientation Dyads (U = 27, Pe = 95, A = 0.987) .......................................................................... 158 XV IIII> "w AijBkz FSQ I") NOTATION Second order tensor, or dyadic-valued operator Second order tensor, or dyadic-valued operator Index notation for tetradic-valued operator for dyadic valued operators A and 2, equivalent to aibjcddl Diameter of tube in tube dilation effect Anisotropic structure tenser Closure coefficient for FSQ-model Number of rigid rods in unit volume Rotary diffusion coefficient (units: 1/time) Rotary diffusion coefficient in dilute solution Diameter of rigid rod Unit vector in Hb and H11, direction Unit vector in the radius direction Unit vector in the vorticity direction Unit vector in the cross-flow direction Unit vector in flow direction Tube dilation effect Tube dilation effect in Doi’s model Fully symmetric quadratic Invariant form of dynamic moment vector in the invariant space xvi Dimensional tumbling frequency Dimensionless tumbling frequency Dimensional wagging frequency Dimensionless wagging frequency Hinch and Lea] l closure Hinch and Lea] 2 closure Unit dyadic tensor First invariant of 2 Index notation of unit dyadic tensor Second invariant of 2 Third invariant of E Boltzmann constant Length of rigid rod Number of species Dimensionless first normal stress Dimensional first normal stress Dimensionless second normal stress Dimensional second normal stress Newtonian plateau Normal vector xvii I'd I'U' Unit vector direct along rigid rod axis Unit vector in rigid rod rotation Neighboring rigid rod vector respect to p Convective flux of rigid rod rotation The first order moment Components of unit vector direct along rigid rod axis Component of unit vector direct along rigid rod axis, respect to eigenvector x1, x2, and x3 Orientation dyad The second order moment Index notation of orientation dyad Orientation tetrad The fourth order moment Péclet number Three spherical coordinate axes Dimensional strain rate tensor Dimensionless strain rate tensor Symmetry operator of arbitrary dyadic valued operator A and __B Temperature Dimensional time xviii Dimensionless time tt Dimensional tumbling period tt Dimensionless tumbling period U Nematic potential coefficient U Critical nematic potential coefficient of the phase transition U1 Upper critical nematic potential coefficient for biphasic region U2 Lower critical nematic potential coefficient for biphasic region Q Fluid velocity uz Scalar valued fluid velocity in flow direction l Dimensionless vorticity tensor )9, 2&2, x; Eigenvectors associated with 11,1, kpz, 21,3, respectively. x, y, 2 Three Cartesian coordinate axes z Arbitrary vector Other Symbols (1 Degree of orientation order parameter, structure parameter )2 Motion of the suspension with velocity 6 Jaumann derivative An Polymer contribution shear viscosity A110 Zero-shear rate viscosity At Change of time step AUMs Maier-Saupe excluded volume potential xix AUo AUMS :3) 1'lE TlS nv k Ale~pzfl~p3 1b1,1b2,lbs, 0 "rt 123 I'd Id lie-l Onsager nematic potential Maier-Saupe nematic potential Excluded volume potential Angle between the projection of molecular axis on the plane [x1, X3] and x1 axis Strain rate Total shear viscosity Elastic contribution of shear viscosity Solvent contribution of shear viscosity Solvent contribution of shear viscosity Tumbling parameter Three eigenvalues of < pp > Three eigenvalues of lg) Angle between rigid rod and y-direction Total stress tensor Elastic contribution of shear stress Elastic contribution of stress tensor Solvent contribution of stress tensor Viscous contribution of stress tensor Orientation density function Q Q) lfi) Subscript MAC. 3 MS Viscous drag coefficient Gradient dimensional velocity gradient dimensionless velocity gradient Ensemble average Function of symmetry operator Hand’s first order closure Second order quadratic closure Dimensional quantity Structure tensor component Convective contribution Elastic contribution Vector indices Maier-Saupe Body rotation Three spherical coordinate axes Solvent contribution Tube Dilation Viscous contribution Three Cartesian coordinate axes Material frame of reference (1.13.7 Superscript E S Vector indices Elastic contribution Solvent contribution Viscous contribution CHAPTER 1 INTRODUCTION 1.1 Motivation The statistical theory of rigid rod suspensions provides a means for understanding the microstructure and rheology of structured fluids (Doi and Edwards, 1978a, 1978b; Doi, 1981). The microstructure may be either isotropic or anisotropic. Under extreme conditions, a nematic phase may occur wherein the long axis of rod-like particles (or molecules) align in the same direction. The tendency for some fluids to develop nematic- like microstructures either spontaneously or under the influence of an external force field has a significant and practical impact on the rheological, optical, and material properties of structured fluids. Particle suspensions and liquid crystalline polymers may be either isotropic or anisotropic, depending on the local environment. The microstructure is often characterized statistically by low-order moments of the orientation distribution function, referred to hereinafter as the orientation dyad (second order tensor) and the orientation tetrad (fourth order tensor). The orientation dyad, < p p > , is symmetric and non- negative (i.c., realizable). The orientation vector B has unit length and is aligned with the principal axis of an axisymmetric ellipsoidal particle, cylindrical rod, or disk-like particle. The eigenvalues of <22 > are real and non-negative (i.c., 0 < 1p] < )‘pZ < 11,3 < 1). The “director” of the microstructure is defined as the eigenvector associated with the largest eigenvalue of the orientation dyad. For an isotropic material, the director has no preferred direction inasmuch as all the eigenvalues of are the sameOtpl = A = 11,3 =1/3). p2 Figure 1.1 illustrates the type of possible anisotropic states that can occur for axisymmetric ellipsoidal suspensions (see Kini et al., 2003; Nguyen et al., 2001a; Parks et al., 1998; Petty et al., 1999; Weispfennig et al., 1999). Each orientation state is parameterized by two nontrivial invariants of the structure tensor l_) (IIb = hf; Q) and 1111, = egg-2) ). The b-operator is defined as the anisotropic component of the orientation dyad (2=(—%l ). As noted on Figure 1.1, three-dimensional anisotropic states for which 0 < )‘p1 = 11,2 < 11,3 < 1 have quadratic forms with prolate surfaces (F -boundary) whereas three-dimensional anisotropic states for which 0 < hp] < lpz = 11,3 < 1 have quadratic forms with oblate surfaces (D-boundary). In the absence of external hydrodynamic forces, all stable and unstable equilibrium states are either on the prolate boundary or on the oblate boundary of the invariant diagram (Doi and Edwards, 1986; Kini, 2003). Two-dimensional planar anisotropic states are located on the B-boundary of Figure 1.1. These states are associated with structure tensors with one eigenvalue equal to zero and two unequal positive eigenvalues (lpl = 0, 11,2 ¢ 11,3 , 11,2 + Ap3= 1). Two-dimensional planar isotropic states (Point C on Figure 1.1) have one zero eigenvalue and two equal eigenvalues (lpl = 0, 11,2 = 11,3: 1/2). A fully-aligned microstructure forms an ideal nematic phase with 11 = 0, 7&2 = 0, and 7.3 = 1 (Point A on Figure 1.1). A q It’c? lug C. Planar Isotropic \. A. Nematic 2/3 2/9 B. Planar Anisotropic F. Prolate Set of Realizable Orientation States ' 2/9 . - 1 1/6 13 =< p p > ——l —l/3 . 3 D. Oblate HI = . b - b E. Isotropic b g “(2 = =) Invarinats of 2 Eigenvalues of < pp > Orientation State _- IIb IIIb M1 7~p2 7ups Notes A Nematic 2/3 2/9 0 0 1 B. Planar Anisotropic 11b = 2/9 +ZIIIb 0 21,2 1 - 21,2 Ap2=[0,1/2] c. Planar Isotropic 1/6 -l/36 0 1/2 1/2 D. Oblate 11b = 6(—IIIb/6)2/3 1—2xp2 rpz xpz rpz=[1/3,1/2] E. Isotropic o 0 1/3 1/3 1/3 F. Prolate 11;, = «rub/6)”3 xpl xpl 1—2xp1 xp1=[o,1/3] Figure 1.1 Anisotropic Orientation States (Parks, 1997) Microstructures associated with axisymmetric suspensions must fall either on the boundaries or within the bounded region of the (Hb , 1111, )-plane identified by Figure 1.1. Microstructures outside this domain are unphysical because at least one of the eigenvalues of the orientation dyad is negative. The realizability domain defined by Figure 1.1 stems directly from the algebraic properties of real, symmetric, non-negative operators and is a fundamental characteristic of any second-order moment of a distribution function. Appendix A shows how the boundaries depend on the invariants of the structure tensor. This model-independent result places an important theoretical constraint on allowable models for the orientation dyad. A practical consequence of Figure 1.1 is that it provides a means to identify a closure model for the orientation tetrad in terms of the orientation dyad, i.e., < pppp > = S(< pp >). This closure is developed in CHAPTER 5 and 6 below. 1.2 Background The primary objective of this research is to examine the influence of the low-order moments of the orientation distribution (microstructure) on the equilibrium and rheological properties of rigid-rod suspensions. Liquid crystalline polymers (LCPs), such as poly (g-benzyl-L-glutamate) in m-cresol and hydroxypropylcellulose in water are often represented as rigid-rod suspensions with a characteristic length L ~ 110 nm and a characteristic diameter (1 ~ 1.16-1.75 nm (Bibe and Armstrong, 1988; Larson, 1999; Walker and Wagner, 1994; Yousefi et al., 2003). The stiffness of LCPs stems from the presence of aromatic rings in the backbone of the polymer or from the a-helix structure due to hydrogen bonding (see Figure 1.2). The significant decrease in the shear viscosity of thermotropic and lyotropic liquid crystalline polymers (LCPs) during processing makes these materials commercially attractive. Specific end uses of LCPs exploit their low elongation resistance to cutting, favorable thermal properties, high resistance to wear, and high-strength, low-weight, and high-irnpact resistance (Collyer, 1992). The tensile moduli of LCPs in the solid phase may vary between 1-100 GPa, depending on the molecular orientation of the constituent polymers (Donald and Windle, 1992). Applications of LCPs are numerous and range from reinforced bulletproof vest to optical components in electrical devices (Collyer, 1992). Liquid crystalline polymers are generally manufactured by a stepwise polycondensation reaction in either a batch or a continuous process (J ansson, 1992). The polymer is mixed with various additives and extruded as a filament. Over half of the LCPs sold are reinforced with 30% — 40% glass fillers having polymeric sizing to produce a strong interface between the fiber and the matrix material (Clarke et al., 1997). Some LCPs are injection molded for special applications. Equilibrium Microstructure In the absence of an external field, a rigid rod suspension has an isotropic microstructure (i.e., IIb= 0 and 1111, = 0) at low concentration. As the concentration increases to a critical value, the microstructure undergoes a spontaneous transition to an anisotropic nematic-like state. Several methods have been developed to study this transition experimentally. For example, Robinson (1966) developed a birefiingence H o +C O—(IJ—N ,R = -[—CH2—CH21—|Cl—? CH2 R \ poly (y—benzyl-L-glutamate), PBLG oc-helix l / i d/2 3D view of PBLG Rigid Rod Figure 1.2 Molecular Structure of LCP (Larson, 1999) technique to study biphasic phenomena (i.e., coexistence of isotropic states and anisotropic states at the same concentration) for poly-y-benzyl-L-glutamate (PBLG) in dioxane solutions. This phenomenon was observed for PBLG aspect ratios from 10 to 100 and has been reported by other investigators using other methods (see Abe and Yamazaki, 1989b; Kubo and Ogino, 1979; Murthy et al., 1976; Sartirana et al., 1987). Abe and Yamazaki (1989a) developed a NMR technique to correlate the relative orientation of the (rt-helical backbone of PBLG rigid-rod molecules by exploiting a quadrupolar splitting phenomena related to the pendant side chain containing C-D and N-D bonds. If the PBLG solution is isotropic, the NMR spectrum has only one resonance peak. As the concentration increases and the microstructure approaches the biphasic region, quadrupolar splitting occurs. The split increases as the fluid becomes more anisotropic. In the biphasic region, the central peak corresponds to an isotropic microstructure and the split signal corresponds to a nematic-like microstructure. As the concentration increases firrther, the isotropic peak disappears while the peak-to-peak distance in the split increases. Figure 1.3 illustrates the observations reported by Abe and Yamazaki for PBLG in DMF (dimethylformamide) and in 1, 4-dioxane with aspect ratios of32,121, and 185. Non-Equilibrium Microstructure In a time independent external field, director tumbling of LCP solutions may occur at low shear rates, but direct measurements using optical methods are difficult. However, optical measurements of this phenomenon for lower molecular weight liquid crystal (LC) suspensions have been reported extensively (Bedford and Burghardt, 1996; Goa .aomosom 352 .%§§> poo 23 ”95 a 3mm é «€on Ezzuo ouoaooom 3 para 5:26on _ _ l noumbzooaoo awn; 3380 Z 295823 ,\ - .7 . < < - m 9 22235 |< < < m 235 - ,M oaobofl < 1 :osfifiooqoo Boa 88m oaobofl Burghardt and Fuller, 1991; Fuller, 1995; Larson, 1999). Rheological measurements on LCP solutions and melts have been used to study director tumbling. Erickson and others (see p. 453-456 in Larson, 1999) related this phenomenon to a phenomenological tumbling parameter A that couples the angular motion of the rigid rod to the strain rate of the flow field (see Eq.(10-3) p. 448 in Larson, 1999). For homogeneous shear flows and | X | > 1, the director attains a steady alignment relative to the flow direction in the shear (or deformation) plane (i.e., flow/cross-flow plane). On the other hand, if | 7t | < 1, the director rotates continuously in the shear plane (see p. 454 in Larson, 1999; Carlsson, 1982; Carlsson and Skarp, 1986). Because A is related to rheological Leslie-Ericksen coefficients, tumbling phenomenon has been studied indirectly for more than thirty years by measuring fluid properties of the suspensions (see Cladis and Torza, 1975; Gahwiller, 1973; Pieranski and Guyon, 1974; Skarp et al., 1981). Skarp et al. (1981) (also see Carlsson and Skarp, 1986; Clark et al., 1981) measured the Leslie-Ericksen coefficients for 4-n-octyl-4’-cyano-biphenyl (a thermotropic liquid crystal) over a range of temperatures (35 — 40 °C). They used an electromagnetic field to initially align the orientation director parallel and perpendicular to the flow direction. After removing the magnetic field, the anisotropic microstructure relaxes and the LE-coefficients were measured. The rheological data indicated that the magnitude of the tumbling parameter was less than unity, which implies director tumbling according to Ericson’s theory. This approach has also been applied to infer director tumbling in LCP solutions subjected to homogeneous shear flows with limited success (see Burghardt and Fuller, 1990; Larson, 1988). Rheological Properties The rheological characteristics of LCPs are important indicators of molecular orientation because time-dependent molecular conformation is strongly coupled with the flow (Walker et al., 1995). Some LCPs in simple shear flows show shear viscosity with strain rate (or stress) response curves with three distinct characteristics: 1) a shear thinning region at low strain rates (Region I); 2) a Newtonian plateau (Region II); and, 3) an additional shear thinning region at high strain rates (Region 111) (Walker and Wagner, 1994; Walker et al., 1995; Larson, 1999). Region 11 occurs for a wide range of strain rates because the molecular orientation and the conformation of the rigid-rod polymer solution are maintained. Shear thinning occurs at higher strain rates because the flow field distorts the microstructure by flow alignment. Clearly, flow alignment enhances the relative motion between phases (i.e. translational diffusion) with the result that the shear viscosity decreases (i.e., shear thinning). An anomalous shear-thinning region at low strain rates has only been observed for LCPs. However, not all LCPs show a Region I behavior. Walker and Wagner (1994) have shown that (1,4-phenylene-2,6-benzobisthiazole) (PBZT') has a Region I response only at relatively high concentrations. This phenomenon has not been firlly characterized experimentally because of inaccurate shear stress measurements at low strain rates (see Doraiswamy and Metzner, 1986; Larson, 1999; Walker et al., 1995). In addition, no clear theoretical explanation for this phenomenon has been identified. Another interesting characteristic of LCP solutions is the occurrence of a negative first normal stress difference (N1) at intermediate strain rates (see Baek et al., 1993; Chono et al., 1996; Kiss and Porter, 1980; Larson, 1999; Magda et al., 1991). At low 10 strain rates, N1 is positive; however, as the strain rate increases, N1 attains a maximum value and then decreases to zero and becomes negative. At higher shear rates, the first normal stress difference becomes positive again. This observation suggests that N1 is sensitive to changes in the microstructure as the strain rate increases. Beck et a1. (1993) have suggested that the transition from positive to negative values of N1 was associated with the orientation director changing from a stable periodic tumbling state to a stable periodic wagging state. This is consistent with the Leslie-Ericksen theory, which requires director tumbling for negative first normal stress differences (also, see p. 449 in Larson, 1999; Burghardt and Fuller, 1990). In addition to the negative first normal stress difference, direct oscillatory response of the viscosity also indicates molecular tumbling phenomenon. When the rate of shear is suddenly changed, the shear stress component of the deviatoric molecular stress shows an oscillatory response including a reversal in the strain rate (Burghardt and Fuller, 1991; Picken et al., 1991; Vermant et al., 1994; Walker et al., 1995). This relaxation response has multiple overshoots and undershoots that can be imposed with various shear rates before the steady state is reached against strain. Since this type of response is independent of strain rate, it must be due to changes in the microstructure rather than the flow properties. In addition, there is only one overshoot when the microstructure relaxes to a steady flow alignment state. These experimental observations support the hypothesis that multiple stress oscillations and director tumbling are correlated (Burghardt and Fuller, 1991; Larson, 1999). ll Theoretical Studies Liquid crystalline polymers have stiffness characteristics that are different from other polymers. LCPs have been studied for many years. Although “industria ” LCPs are not as rigid as “laboratory” LCPs, understanding the behavior of rigid rod suspensions would nevertheless provide valuable information and insight related to processing LCPs. Theoretical studies of LCP orientation phenomenon are primarily related to the Smoluchowski’s (S-) equation. The S-equation governs the distribution of orientation states. It is a partial differential equation that balances the accumulation of states subject to rotary convection and rotary diffusion in orientation space (see CHAPTER 2 below). The diffusive flow has two contributions: Brownian motion and the excluded volume phenomenon. Brownian motion tends to mix the LCP molecules randomly whereas the excluded volume effect tends to align the LCP molecules. The convective flux arises due to the torque on the LCP molecules in a shear field. There are several ways to study the S-equation. One approach is to develop a solution using spherical harmonics (Chaubal and Leal, 1997, 1999; Larson, 1990). Another approach is based on the method-of-moments. Developing an explicit expansion for the density function is a complicated process and requires significant computational resources. On the other hand, developing a solution based on low order moments of the density function requires a closure approximation (Marrucci, 1996). Doi (1981) developed a second order moment equation by integrating the S- equation with Maier-Saupe potential for the excluded volume. Doi used a quadratic closure approximation for the orientation tetrad (i.e., < p p p p > = < p p > < p p > ). Unfortunately, these approaches do not retain the six-fold symmetry and six-fold 12 contraction properties of the fourth order moment (see CHAPTER 3 and the development in CHAPTER 5). Hand (1962) used a first order closure approximation for < p p p p > , which satisfies six-fold symmetry and six-fold contraction. However, Hand’s closure is limited to microstructures near the isotropic state (see E on Figure 1.1). Later, Hinch and Leal (1976) introduced two closure approximations based on an anisotropic analysis of the orientation tetrad near the isotropic and nematic states (see Figure 1.1). The HL-closures predict phase transition from the isotropic to the nematic state, but predict unrealizable behavior for some situations (Chaubal and Leal, 1999). Later, Cintra and Tucker (1995) developed a new closure for the orientation tetrad based on an orthotropic operator. The approach assumes that the symmetry directions of the tetrad coincide with the orientation dyad. The closure coefficients of the tetrad are obtained by fitting the moment equation with the “exact” solution based on a spherical harmonic expansion of the orientation density function. The orthotropic closure is applicable to simple geometries and can provide valuable bench mark information. There are other closures that combine previous approximations (Tucker, 1988; Larson, 1999). For example, Tucker (1988) has combined Hand’s closure approximation at the isotropic state and the decoupling approximation at the nematic state. This superposition of two asymptotic closures is similar to the strategy employed by Hinch and Leak and is an example of a hybrid closure approximation. Larson (1990) used the decoupling closure for the excluded volume potential and the HL-closure for the convective flux in the same model. None of the foregoing closure approximations satisfy the symmetry properties of the orientation tetrad and the realizable condition on the 13 second order moment. In addition, none have firlly predicted the microstructure and rheological properties of LCPs. Recent research at Michigan State University has developed a representation of the orientation tetradic in terms of the orientation dyad that satisfies the six-fold symmetry and the six-fold contraction properties associated with < p > (see Parks et al., 1999; Parks and Petty, 1999a, 1999b; Petty et al., 1999; Imhoff, 2000; Nguyen, 2001; Kini, 2003; Mandal, 2004). This closure is incomplete and needs an appropriate closure coefficient C2 (11b , HIb) so that the orientation dyad is realizable for all conditions. Previous studies assumed that C 2 = 1/3 for all anisotropic states within the invariant diagram (Figure 1.1). This condition must be true at the nematic state, but it is not required elsewhere. However, it is noteworthy that C2 = 1/3 predicts biphasic phenomena (Kini, 2003; Nguyen, 2001) and tumbling phenomena (Nguyen, 2001). However, a “universal” value of C 2 = 1/3 causes unrealizable behavior for some physically allowable initial conditions. This is unacceptable and a resolution of the problem is developed in CHAPTER 6 below. Imhoff (2000) and Mandal (2004) identified a value for C2 by using solutions of the S-equation. Their best fitted C2 value was 0.37, but this choice of C2 is also unacceptable because it yields unrealizable results orientation dyad for certain initial conditions. 1.3 Objective This research addresses a long-standing fundamental problem related to the self-alignment and flow alignment of structured fluids. The approach, which builds on 14 the statistical theory developed earlier by Doi and many others (see, esp., Doi and Edwards, 1986; Bird et al., 1987a,b; Larson, 1999) provides new insights and understanding of the relationship between the microstructure and the phenomenological properties of structured fluids at the continuum scale. The objective of the research is to develop a closure for the orientation tetrad that yields a realizable model for the orientation dyad. The new approach is used to predict the microstructure and the rheological response of rigid rod suspensions to simple shear flows. By using the S-equation based on Doi’s theory (1981), an equation for the second order moment of the orientation density fimction can be developed that depends on the fourth order moment (i.e., the orientation tetrad). Although the method-of-moments has been employed for more than thirty years as a mean to study the microstructure of rigid rod suspensions, understanding the equation has been limited by the absence of a satisfactory closure model for the orientation tetrad. This is a significant theoretical deficiency that hinders the interpretation of rheological anomalies associated with the response of microstructure fluids to simple shear fields. To address this issue, this research presumes that the orientation tetrad can be approximated by using an algebraic closure, < pppp > = 3(< pp >) . The efficacy of this hypothesis will be evaluated for a class of microstructured fluids (i.e., rigid rod suspensions) in homogeneous shear flows. The aim of the research is to deve10p a realizable dynamic model for the orientation dyad for a wide class of complex engineering flows. 15 1.4 Outline The equilibrium relaxation of the orientation dyad in the absence of an external field (CHAPTER 7) and the non-equilibrium relaxation of the orientation dyad in the presence of an external field (CHAPTER 8) are addressed in this research. The S- equation in orientation space (CHAPTER 2) is used to develop an ordinary differential equation for the orientation dyad, < pp > (CHAPTER 4). The moment equation has three physical contributions: rotary Brownian motion, excluded volume phenomenon, and hydrodynamic interactions through particle/fluid torque. A Maier-Saupe potential is used for the excluded volume effect and Jeffery’s model is used for the hydrodynamic interactions (CHAPTER 2). In addition, the effect of tube dilation on the diffusive flux is examined. Doi’s stress model is used to predict the viscosity and the normal stress differences (CHAPTER 9). Once the realizable closure approximation is obtained, the ordinary differential equation is solved for the orientation dyad by using a fourth order Runge-Kutta algorithm with a dimensionless time step less than 0.0003 (see APPENDIX G). The moment equation has three independent dimensionless variables: U, Pe, and A. (see CHAPTER 2). For A _<_ l, the model is used to study the microstructure and rheology of rigid rod suspensions for a wide range of U and Pe (CHAPTER 7 and CHAPTER 8). Various initial value problems are examined with and without homogeneous shear. In CHAPTER 3, two metrics of the microstructure are defined to evaluate the results: 1) the order parameter a; and, the deviation of the director from the flow direction, 253 - g2 . l6 CHAPTER 2 SMOLUCHOWSKI EQUATION FOR RIGID ROD SUSPENSIONS 2.1 Introduction Smoluchowski’s equation (S-equation) governs the evolution of the orientation density function ‘I’( p, t; Z (X, b) for a suspension of rigid rod particles (see p.50 in Doi and Edwards, 1986). The fraction of particles with orientation vectors with angular coordinates between (0, 4)) and (0 + A0, 4) + Ad>) is given by ‘I’(p, t; 2(3, D)sin9AOA¢. In a frame of reference moving with the local velocity of the suspension, the S-equation is a balance equation for orientation states and can be written as: 6‘? 6 . (3)3 =-5E-(2‘I’). (2.1) . 6 . . . . . In the above equation, ‘6— rs a surface gradient operator on a sphere in orientation space P and the vector p is the angular velocity of the particle about its center of mass. The vector X represents the spatial position of a material fluid element at some arbitrary reference time; the spatial position of the material fluid element at time t is f; = )2 (X, t). The vector zLX, B is the motion of the suspension with velocity fl (3, D defined by A it; t) Eta—i] . (2.2) at X The Operator on the left-hand-side of Eq.(2.l) represents the substantial (or material) derivative of the orientation density function: 17 A (2.3) a: at E —T + —? X at i at The rotary flux of orientation states relative to a material frame of reference, p‘I’ , X. can be separated into a rotary convective flux pC‘I’ and a rotary diffusive flux lE—Ec )W; E‘I’EEC‘I’+@-2c)‘l’- (2.4) In this research, the rotary convective flux developed by Jeffery (1922) for ellipsoidal particles suspended in a homogenous shear field will be used for pc‘l’ (see Section 2.2 below). Doi’s model for the rotary diffusive flux (see Section 2.3 below) will be used for (p— p C )‘I’ . The rigid rods have the same density as the suspending fluid so gravity is unimportant. The S-equation given by Eq.(2.1) above assumes that spatial diffusion of the particles relative to the translational velocity is also unimportant. 2.2 Jeffery’s Model for Rotary Convection: Tumbling Coefficient Jeffery’s model is used for the rotary convective flux (Jeffery, 1922). Hydrodynamic drag causes the rotary motion of the suspended particles (Batchelor, 1976, 1982; Bibbo et al., 1985). A balance of angular momentum on an axisymmetric rigid rod yields the following equation for p C (see Jeffery, 1922; Parks et al., 1999): 5p re 5%», brawl-22H III!» '2 1, (2.5) 18 where 7» is a dimensionless tumbling parameter. ll Cl» and W are the rate-of-strain and vorticity tensors, respectively: §=§lf§+fitfl=§lv§+w2fl; (2.6a) i=§l€7§-(rrfl=§[v§—(vyTl. (26» For homogeneous shear, VQ = ye), 92 where 7: (6g):(vg)T =‘/2§:§T 423':ng =constant. (2.7) In this research, A is only a function of the particle aspect ratio, L/d. For axisymmetric particles (see Jeffery, 1922) L (592—1 ._ 5 +1 . (2.2) (g? +1 For large aspect ratio particles 71. i 1. For disk-like particles 7t -'—- 1. A typical tumbling parameter for slender rod-like particles is about 0.7, which is equivalent to L/d s 2.38 (see p.280 Larson, 1999; Bretherton, 1962; Trevelyn and Mason, 1951). The rotational period of a rigid rod can be related to the tumbling parameter (Jeffery, 1922; p.280 Larson, 1999). For )t = i 1, a prolate spheroidal particle and a disk-like oblate particle have infinite rotation periods (i.e., they are not rotating). Experimental evidence for particle rotation (or tumbling) in rigid rod suspensions has been given by Larson and many others (see p.280 Larson, 1999; Anczurowski and Mason, 1967a, 1967b; Frattini and Fuller, 1986). 19 2.3 Brownian Motion: Rotary Diffusivity Rotary Brownian motion is an important phenomenon in particle/fluid suspensions. This phenomenon has a direct impact on models for viscoelasticity, diffusion, birefringence and dynamic light scattering. The theory of rotary diffusion in concentrated suspensions is well described by the following hypothesis (see p.294, Doi and Edwards, 1986; Parks et al., 1999): as! ea ALI '—' =— —-‘I’— . 2.3 2 EC R" [op 62(kBT)] ( ) In the above equation, AU is an excluded volume potential and < DR > represents an average rotary diffusion coefficient. In general, < DR > depends on the phenomenon of tube dilation (see p.360 in Doi and Edwards, 1986; Kuzuu and Doi, 1993, 1994; p.520 in Larson, 1999), the particle aspect ratio L/d, the volume fraction of particles, and the temperature. In this research, the influence of tube dilation on the microstructure and rheology will be examined (see CHAPTER 8 and CHAPTER 9), but most of the applications will assume that < DR > is given by 3ka éD§a 1mg L3 (29) In the above equation, k]; and 718 represent, respectively, the Boltzmann constant and the solvent viscosity; T is the temperature. Dfi has units of 1/(time) and represents the rotary diffusion coefficient for dilute suspensions of rigid rods (see p.334 in Doi and Edwards, 1986; p.281 in Larson, 1999). For semi-dilute and concentrated suspensions, Eq.(2.9) is multiplied by a tube dilation factor Fm (< DR > = Fm D; ), which depends 2O on the invariants of the microstructure (see Figure 1.1). In this research, FTD = 1 for most of the applications. For the tube dilation examples (see Sections 8.5 and 9.5), Doi’s theory for FTD is used (see p.360 in Doi and Edwards, 1986): 1 Doi 11b ) (1- 3 (2.10) 2 Note that Eq.(2. 10) implies that < DR > —» 00 as the microstructure approaches a nematic state (see Point A on Figure 1.1). 2.4 Excluded Volume Phenomena: Maier-Saupe Potential The excluded volume potential introduced by Eq.(2.6) above accounts for the interaction of a rigid rod particle with neighboring particles. The main physical idea is that particles cannot occupy the same space at the same time. This phenomenon has important consequences that partly explain the self-alignment and the flow-induced alignment of particles. Doi (1981) and others have developed models for the instantaneous excluded volume potential by minimizing the Onsager free energy for rigid rod suspensions (11g et al., 1999; Onsager, 1949) with the result that (see p.359, Doi and Edwards, 1986): A kAI'JI‘ =-Upp:+{higherorderterms}. (2.11) B _- -_ The above equation stems from an expansion of the second virial coefficient of the Onsager nematic potential. The lead term is the so—called Maier-Saupe potential: A AUMS kBT =-Upp:. (2.12) 21 This model is used in the Doi theory for rigid rod suspensions and is also used hereinafter. The average of Eq.(2.12) shows that < AUMS > is proportional to the second invariant of the orientation dyad: < AU MS > = - ngT < p p > :< pp >. With < p p > = 1+ 2 , the average excluded volume potential can be expressed as l 3 (AUMS >=—UkBT(%+Hb). ‘ (2.13) Thus, at the isotropic state (i.e., Point E on Figure 1.1), < AUMS >= —-;:UkBT ; and, at the nematic state (i.e., Point A on Figure 1.1), < AUMS > = —UkBT . The nematic coeffiCient U is dimensionless and, as indicated above, compares the average excluded volume potential with kBT. For rigid rod suspensions (see p.66 in Larson, 1999), U is proportional to the concentration of particles and the excluded volume V}; = aoL. The parameter a0 is the diameter of a tube of length L that contains a single rigid rod. In this research, a0 is assumed to be independent of the local microstructure (i.e., 11b and 1111,). Tube dilation affects the nematic potential (and rotary Brownian diffusion) through the F TD-factor introduced by Eq.(2.10). 22 2.5 Discussion The S-equation with Jeffery’s model for rotary convection (see, Eq.(2.5)) and Doi’s model for the excluded volume potential (see, Eq.(2.12)) can be written as 6‘1’ 6 (57); +Pe-a— K—l amt-.1221 L§ 21M= ' ‘ (2.14) a 6 a +FTD—- —+U‘I’—(pp 2 62 52 " '- where t is a dimensionless time and Pe is Péclet number: taongi, Pea 1 d“}. (2.15) 0 d 6DR y The S-equation determines how the orientation density function changes with time over the surface of a sphere in orientation (phase) space (see Edwards and Beris, 1989). The relaxation of ‘1’ (p , t) from an initial state is controlled by four physical factors: 1) a rotational torque due to the antisymmetric component of the velocity gradient; 2) a rotational torque due to the symmetric component of the velocity gradient; , 3) a rotational torque due to Brownian motion; and, 4) a rotational torque due to the excluded volume phenomenon. A direct numerical (or analytical) analysis of Eq.(2.14) subject to arbitrary, but realizable, an initial condition has not been done. Some limited results have been reported for isotropic initial conditions (see Doi, 1981; Hand, 1962; Hinch and Leal, 1976; Petty et al.,1999; Tucker, 1988), but an understanding of Eq.(2.14) has primarily resulted from a study of the low order moments of ‘I' (_p , t). The moment method (see CHAPTER 3) will be used in this research. 23 CHAPTER 3 MOMENTS OF THE ORIENTATION DISTRIBUTION 3 .1 Introduction In this research, a rigid rod is approximated as an axisymmetric ellipsoid. A single orientation vector 2 defines the instantaneous orientation state of the rod in terms of the angular variable 0 and d) (see Figure 3.1): 2:9r = sin(0)cos(¢t)ex + sin (0)sin((l>)§y + cos(0)§z. (3.1) There is no distinction between the head and the tail of the rod. Therefore, the orientation density function satisfies the symmetry condition ‘1’(p, t) = ‘1’ (-p, t). The vorticity direction is ex; the cross-flow direction is gy; and, the flow direction is £2. The plane that contains the flow direction and the cross-flow direction is called the shear plane (or the deformation plane); the plane that contains the cross-flow direction and the vorticity direction is called the cross-flow/vorticity plane; and, the plane that contains the vorticity direction and the flow direction is called the vorticity/flow plane (see Figure 3.1). Measuring the orientation of individual rigid rods in a suspension is not practical. However, low-order moments of the density function ‘I’( p , t) can be measured. These moments provide an objective means to understand the complex behavior of rigid rod suspensions. For a suspension of axisymmetric particles, the first moment

is zero because of symmetry: 24 com Ema a mo mire. 330222 05 mac? @8085 H885 magneto msooqflqfimfi Tm 2:me Nx Em ml .1. Q X \4 $6 283 bmowtozaomémouo "x £2886 €092, “ T u o Nx >I 3o 1 ..... I 5 w e w o .couoohw Bomémeo 4A a c I o w v / 25m 32:363.“? m MW .MK A mm v o mo: «>890 BEE Eonm \ N o m _ . um. mm .m& n 3882890 £3826 Bow 4 25 (0,4), t) sin(0)d0d4)= 0. (3.2) ”I” All of the odd moments are zero. The zeroth moment (i.e., integration of the density 0:3 function over the sphere) is unity because the total number of particles is a constant. 3.2 Orientation Dyad: Structure Tensor and the Order Parameter The second moment of the orientation density function is I'd I'D 22] I pp ‘I’(0,4),t)sin(0)d0d4). (3.3) 0 O The second moment < pp > is also called the orientation dyad. < p _p_ > is a symmetric, non-negative operator, and its trace is unity because tr()a<_p-p>=1. (3.4) This dyadic-valued operator defines the microstructure of rigid rod suspensions. Figure 3 .2 illustrates an isotropic (or three-dimensional random) orientation state for which 1 . . = 3(gxgx +§y§y +§z§z) , isotropic state. When the rigid rod particles are randomly distributed in a two-dimensional plane (see Figure 3.2 and Figure 1.1), then -1 l ' tr ' tat ——?:(_e_y§_y +_e_z_e_z), panarrso oprcs e. When all of the rigid rods are pointing in one direction (see Figure 3.2), then < p p > = g e nematic state. 2—2’ 26 3er 83858 2: com moafim aozfiaoto 83H NM 8&5 Osman: . . . — ulnl IAmmV . ... - . . ofiobofl 853 Aumnm+ .T osmVInA mmv 1, J m 2858“ HNMNIIT %l %| +xIKIVIIA 27 The orientation dyad can be represented as the sum of an isotropic operator and an anisotropic operator: "H = +2. (3-5) wIH A scalar-valued order parameter a is often used to define the orientation state or the prolate (and oblate) boundary of Figure 1.1. This parameter defined in terms of the second invariant of the structure tensor b: 3 1/2 {311,} . (3.6) Note that: or = 1 at the nematic state; or = 0 at the isotropic state; and, a = —1/2 at the planar isotropic state (see Figure 1.1). 3.3 Orientation Tetrad: Symmetry and Contraction Properties The fourth order moment of the orientation density function is: 21! 1! _p p =j Jpp_ pp,‘I’(9 q), t) sm(0)d0d4) (3.7) o o This statistical property is called the orientation tetrad. Previous studies have used the following closure hypothesis for the orientation tetrad (Hand, 1962; Hinch and Leal, 1976; Doi, 1981; Tucker, 1988; Petty et al., 1999): 3(< pp >) (3.8) Eq.(3.8) is also used to support this research (see CHAPTER 5). Clearly, the orientation tetrad has six-fold symmetry. For example, with p = a = b = g = d , it follows that 28 _=. E25<éhsa> (3-9) E E E <_a.12<_19>- IO v V II» Io‘ IO ID- IO I19 1:3. 16:. I!» I0‘ Io‘lo V V (3.10) V V III III A A A A 10" It” IO- Eqs.(3.9) and (3.10) are important properties that should be retained by any closure based on Eq.(3.8) above. 3.4 Realizable Anisotropic States: Invariant Diagram The orientation dyad is a symmetric and non-negative operator (see Parks et al., 1999). The eigenvectors ii and the eigenvalues 7» pi of the orientation dyad < _p_p > are defined by: ~xi=hpixi . (3.11) The eigenvalues of < pp > are real and non-negative: 0 5 API 5 11,2 S 2,1,3 5 1. The eigenvectors of < p p > are: )‘1 r 5.1 = xxlgx +xyl§y +le§z 12 , x2 = xngx +xy2gy +x22§z (3.12) B 2 $3 = xx3§x +xy3§y +xz3§z The three invariants of < pp_ > are 29 Ip =I'I()=}"pl + Apz '1' )‘p3 _ _ 2 2 2 11p —tr(-)—).pl+1.1)2+1.p3 (3.13) _ _ 3 3 3 IIIp --t(-)—7.pl +1132 +1133. The orientation director is defined as the eigenvector associated with the largest eigenvalue of < pp > (i.e., x3 ). The structure tensor p , defined by Eq.(3.5), has three eigenvalues: bbl , hbzl, and 21,3. The invariants of p are 1b =h’fg)=lb1+7»b2+’~b3=0 11b =u(b-p)=1§2+1§2+1§3 (3.14) mb=tr< Bartlet“;- IIU‘ II The second and third invariants of p are non-zero. Figure 1.1 uses these invariants of the microstucture to identify all possible realizable orientation (see Parks et al., 1999; esp., Lumley, 1978). The orientation states within this designated region are realizable inasmuch as the eigenvalues of < pp > are real and non-negative. The eigenvectors of < pp > and 1:) are the same. The eigenvalues are related by 1 . _ )‘bi =lpi ~§ , 1-1,2,3. (3.15) For uniaxial alignment states (see point A of Figure 1.1), 1 l 2 Apl = O, 1.132: 0, and 11,3 =1 (xbl = -3, Abz =—§, lb3 :3). (3.163) For planar isotopic states (see Point C of Figure 1.1), A'pl = 0, 1.132: 1/2, and A133 =1/2 (hb1=-?13-, le =-;-, 71.133 =%). (3.161)) 30 For isotopic states (see Point E of Figure 1.1), 1p, =1/3, 1P2 = 1/3, and 1p, =1/3 (ltbl =0, 11,2 :0, 1b, =0). (3.16c) For planar anisotopic states (see Line B of Figure 1.1), 1 1 3 AP] = 0, A132: 1 -1133 , and A133 0‘13] =—§, )‘bZ =Ap2 —§, Ab3 =E-Ap2) (116(1) For axisymmetric oblate states (see Line D of Figure 1.1), _ _ 1 1 4 1 l (1131:3—271132, M32 =7~p2 -§, 1133 =7~p2 -§) (3.16e) For axisymmetric oblate states (see Line D of Figure 1.1), 1 hp]: xpz = leand )‘p3 =1-21p1 (05113153) 1 1 4 0.14.17, ”b2 =0” '3’ x... 721.1) (3.160 The planar anisotopic boundary of Figure 1.1 follows by substituting the planar anisotopic eigenvalues of 2 into Eq.(3.14): 11b = 2/9 + 2111b . (3.17) The prolate boundary of Figure 1.1 follows by substituting the axisymmetric prolate eigenvalues of p into Eq.(3.14): 2/3 111, = 6 (13151] . (3.18) The oblate boundary of Figure 1.1 follows by substituting the axisymmetric oblate eigenvalues of 2 into Eq.(3.14): 31 2/3 11b=6[ 1611b) . (3.19) 3.5 Discussion This research proposes to examine a longstanding moment closure problem related to the orientation tetrad of rigid rod suspensions, such as liquid crystalline polymer (LCP). The proposed approach is based on an analysis of the low order moments of the S-equation for the orientation density function. The moment equation for < pp > is unclosed inasmuch as it depends explicitly on < p p p p >. In CHAPTER 4, the moment equation for < pp > is presented. In CHAPTER 5 and CHAPTER 6, a realizable closure for

based on the hypothesis expressed by Eq.(3.8) is developed that satisfies all the symmetry and contaction properties defined by Eq.(3.9) and Eq.(3.10). 32 CHAPTER 4 EQUATIONS FOR THE MICROSTRUCTURE AND THE STRESS 4.1 Intoduction In principle, the S-equation given by Eq.(2.14) can be solved numerically for any initial condition, ‘I’(p, 0). The resulting solution can be used a posteriori to calculate the low-order moments that appear explicitly in the stess equation. This approach does not require an a priori closure model for moments. Although a direct numerical simulation of the S-equation for relatively simple flows and initial conditions provides useful predictions of statistical properties, this approach is not practical for complex flows or complex initial conditions. The method-of-moments provides an alternative means to study Eq.(2.14). Unfortunately, this approach is unclosed inasmuch as the dynamic equation for < pp > depends explicitly on < p p p p >. However, once an appropriate closure has been identified (and validated) the moment equation for < pp > can be used to study the relaxation of the microstucture of rigid rod suspensions from arbitrary anisotopic states. Thus, the main objective of this research is to develop an algebraic closure for the orientation tetad in terms of the orientation dyad. 4.2 Dynamic Equation for the Orientation Dyad An equation for < pp > follows directly from Eq.(2.14) by first multiplying the equation by p p and then integrating over the unit sphere: 33 6 6t >x+PellT-<22>+<22>_fl_l= —FTD[(< pp> —%l) —U(< pp> - < pp> — < pppp>:< pp>)] (4.1) +>~Pel§T° <22> +<22> -§-2 <2222>=§l The terms on the lefi-hand-side of Eq.(4.1) is the Jaumann derivative of < pp > , which represents the rate of change of < pp > relative to a tame rotating with an angular velocity proportional to the vorticity (see Bird et al., 1987b). The first bracket on the ri t-hand-side represents rotary diffusion due to Brownian motion and excluded volume phenomenon. The second bracket on the right-hand-side accounts for rotary convection due to fluid/particle drag. Clearly, Eq.(4.1) is unclosed due to the explicit appearance of the orientation tetad < p p p p >. Note that both of the bracket terms, which represents different physical phenomena, depend on < p p p p >. 4.3 Dynamic Equations for the Stucture Invariants for Pe = 0 For Pe = 0, Eq.(4.l) can be reduced to two coupled scalar equations for the stucture tensor invariants IIb and 1111, (see Figure 1.1). The following equations for Pe = 0 are derived in APPENDD( A and B: dcllltb =—211b+2FTDU|:§-Ilb+IHb—p:zp] (4.2) din: =—3IHb +3FTDU[§-1Hb +-;—11b2 _2::-(22)] (4.3) 34 The stucture tensor 2 (.=.< pp > —-13-£ ) appears explicitly in Eqs.(4.2) and (4.3). Once closed, the above coupled nonlinear autonomous (i.e., the independent variable does not appear explicitly in the differential equation) first-order, ordinary differential equations can be integrated from any orientation state in the realizable region defined by Figure 1.1. If U = 0, then the equations are closed and linear and can be integrated analytically (see Section 7.3). 4.4 Algebraic Equation for the Stess The deviatoric component of the total stess for a rigid rod suspension consists of three contributions: 1 = is +i" +35 (4.4) where is is the solvent contribution; iv is the viscous contribution; and, :E is the elastic contribution (Baek and Magda, 1994). is depends on the solvent viscosity ns and the stain rate of the flow (Newtonian fluid): S = Znsg . (45) 111-1) The microstucture of the suspension couples with the stain rate to produce an additional viscous contribution to i : v (4.6) II ch II (11> = ch <2222>= The viscosity coefficient cCR is given by 35 30k T ccR =——;B—— (4.7) 6DR Fm The viscous stess is due to the drag of the solvent on the rigid rod (see p.521 Larson, 1999) Doi (1981) deve10ped an elastic stess based on the Onsager free energy with the result that (see Han and Kim, 1993; Larson, 1988): :E = 3ckBT[( —%l)—U(< pp>~-:):l . (4.8) The first term on the right-hand-side of Eq.(4.8) represents the stess induced by rotary Brownian motion. The second term is the stess caused by the excluded volume phenomenon. For homogeneous shear flows, the viscosity of the suspension is given by A E .:.§z n= ’ . (4.9) 1’ The first and second normal stess differences are defined as follows Neel-3921,32,. (4.10) 1212:; '39 -§x-:-2x- (4.11) Y In a cone-and-plate viscometer, a positive N1 represents a force that pushes on the cone. A negative N1 represents a force that causes the cone to push on the fluid. 4.5 Discussion Eq.(4.l) provides the relaxation of the dyad. The solution governs depends on three dimensionless groups: Pe, U, and A. These groups are independent and account for different physical phenomena. A closure approximation for the orientation tetrad 36 < pp pp > in Eq.(4.1) is developed based on six-fold symmetry and six-fold contaction properties (CHAPTER 5). In CHAPTER 6, Eqs.(4.2) and (4.3) are used to develop a realizable closure for Eq.(4.1). Doi’s elastic stess model defined by Eq.(4.8), is similar to but not as complete as Ericksen, Leslie, and Parodi’s (ELP) stess. Doi (1981) noted that the ELP stess has a limitation for predicting nonlinear viscoelasticity, which is important in rigid rod suspensions (Doi, 1981). The Doi stess, even if it is not as general as the ELP stess, separates the elastic and viscous contributions of the stess in both isotopic and nematic phase tansition. However, Doi emphasizes that his model is incomplete because of physical and mathematical assumptions. The influence of the microstucture (i.e., < pp > and < p pp p >) on Eqs.(4.6) and (4.8) is the focus of CHAPTER 9. 37 CHAPTER 5 CLOSURE FOR THE ORIENTATION TETRAD 5 .1 Intoduction In this chapter, a closure stategy for the orientation tetad is intoduced based on the hypothesis that (5. 1) <2222 > = 3<< 22 >> Parks et a1. (1999) used the Cayley-Hamilton theorem of linear algebra (Frazer et al., 1960) and developed a representation for 3(< pp >) in terms of the following six independent tetadic Operators: I ; l < p p > ;

m—t 111—: IT) <22> <21)? ; <22>-<22> .1. : They proposed that S(< pp >) could be written as a linear combination of two tetradic operators with six-fold symmety and six-fold contaction (see Eqs.(3.9) and (3.10)): =C131()+C232(). (5.2) In the above representation, 31(< p p >) is first-order in < p p > and 32 (< p p >) is second-order in < pp >. The scalar coefficients C1 and C2 are functions of the eigenvalues of < p p > or, equivalently, the eigenvalues of the stucture tensor The closure coefficients C1 and C2 are not independent because p=- ;. le Eq.(5.2) must satisfy the six-fold contaction conditions (see Eq.(3.10): 38 tr<2222>=trlfi<<22>>l = C1 t1‘[31(< B B >)]+C2 t[32(< E E >)] =(C1+C2) The above result implies that C1 + C2 = 1. Parks, Petty, and Shao (see Parks et al., 1999) developed explicit expressions for Sl() and 32() by using the following two six-fold symmetic tetadic operators: S[g g] =AijAkt +AikAj£ +AieAjk (5-3) and 312, El=AijBkt + Aikle + AilBjk + BijAkl + BikAjt + BirAjk (5-4) In the above equations, the operators A and B are symmetric and may be 1 , < pp > , or ~. The operators Sl() and 32() are definedas follows: 1 5 31()E--3—5—S[I {HESRBP l] (5.5) 2 32():—:-3—5:S[ll] +S[] (5.6) 10 -—< >-< >1 . 3,122 22 .1 Eqs.(5.2), (5.5), and (5.6) define a preclosure for the orientation tetad < >2 I'U I'd 1'6 1'6 <2 pg>=[l—Cztnb.mb)]81(>+czmb.mb)32(). (5.7) 39 This representation satisfies all six-fold symmetry and six-fold contaction properties of the exact orientation tetad. A closure for the second-order coefficient C2 (11b , 1111,) will be identified in CHAPTER 6 based on the idea that solutions to Eq.(4.1) must be realizable (i.e.,

2gz 2 0 , g arbitary) for all initial conditions

(O) with invariants IIb(0) and IIIb(0) on the invariant diagram defined by Figure 1.1. Realizable solutions (unsteady, steady, or periodic) must be produced for any combination of the physical property groups of 1., U, and Pe: —1 5 AS 1; 0 5 U < oo; 0 5 Pe < 00. The development of C2(IIb, 1111,) in CHAPTER 6 is the primary accomplishment of this research. 5.2 Closure Models Hand’s Closure Hand (1962) studied the microstucture and rheology of rigid-rod suspensions near the isotopic state (see Figure 1.1) and assumed that Eq.(5.1) could be approximated as <£EBB>Hand=3ML (5.8) Which is Eq.(5.7) with C2 = 0. As demonstated in CHAPTER 7 (Pe = 0), this yields unrealizable orientation dyads for finite values of U and, thereby does not provide a basis for understanding the self-alignment phenomenon of rigid-rod suspensions. Doi ’s Closure (Decoupling Approximation) 40 Doi (1981) studied the microstucture and rheology of rigid rod suspensions near the nematic state (see Figure 1.1) and assumed that Eq.(5.1) could be approximated as Doi=- (5.9) Unfortunately, this closure does not satisfy the six-fold symmety and six-fold contaction properties of the orientation tetad. Although Bq.(S .9) has been widely used by many researches for more than fifty years, it misrepresents the fundamental symmety characteristics of the orientation tetad and, thereby, the symmetry characteristics of the orientation density function for rigid suspensions of ellipsoidal particles. This unphysical property occurs for all realizable orientation states, including the nematic-like states near Point A of Figure 1.1. Bq.(5.9) is not a good approximation for < p p p p > , comments to the contary in the cmrent literature notwithstanding (Doi and Edwards, 1986; Chaubal, 1997; Larson, 1999). Tucker '5' Hybrid Closure Tucker (see Tucker, 1988) used the following hybrid closure for < p p p p > to study the flow-induced alignment of rigid rod suspensions: < EBEB>Tucker= 27(det )31() (5.10) +[1—27(det)] . For isotopic states, Xpl = 21,2 = 711,3 = 1/3; therefore, 27(det) =1 and Eq.(5.10) asymptotically approaches Hand’s closure for orientation states near the isotopic state (see Point E, Figure 1.1). For the nematic state, 21,1 = 711,2 = 0 and 11,3 = 1; therefore, 27(det) = 0 and Eq.(5.10) asymptotically approaches Doi’s closure. For the 41 reason cited above, the hybrid closure also misrepresents the firndamental symmetry characteristics of the orientation density function and, thereby, does not provide an appropriate closure for the moment equation governing the behavior of < p p > . Hinch and Leal ’3 Closure Hinch and Leal (1976) developed the following closure for < p p p p > :< p p > , which appears in the moment equation for < p p > (see Eq.(4.1)): HL3=CII+°2<22>+C3 (511) where s——+-8—t[ pp> ]-—6—t[ ] 15 15 -- 5 -- -- 2 1 c2 =———t[ ] 5 5 - c __3 3" 5 0 =2 4'5 Note that the above result is symmetric and that t[< p ppp >:< p p >] =< pp >:< p p > , as required by Eq.(4.1). The foregoing closure stems from the idea that for Pe = 0 and U —) 0 (‘weak’ nematic stength), the microstucture asymptotically approached the isotOpic state. And, for Pe = 0 and U —) oo (‘stong’ nematic stength), the microstucture asymptotically approaches the nematic state. Unfortunately, as shown in CHAPTER 7, the Hinch and Leal closure yields unrealizable predictions for Fe = 0 and 42 finite values of U and, thereby, does not provide a physically acceptable closure for Eq.(4.1) (see APPENDIX D). Fully Symmetric Quadratic (FSQ-) and Orthotropic Closure Petty and Bénard together with their students (see Imhoff, 2000; Imhoff et al., 2000; Kim et al., 2001, 2002, 2003, 2004, 2005; Kini 2003; Kini et al., 2003, 2004; Mandal et al., 2003, 2004; Nguyen, 2001; Nguyen et al., 2001a, 2001b; Parks et al., 1999; Parks and Petty, 1999a, 1990b; Petty et al., 1999) have used Eq.(5.7) as a closure for the orientation tetrad by assuming that C2 is a universal constant. However, realizability at the nematic state (Point A or Figure 1.1) and at the planar isotopic state (Point C or Figure 1.1) requires C2 = 1/3 and C2 = 1/2, respectively (see CHAPTER 6). Clearly, C2 must depend or the local orientation state characterized by 11b and 1111,. Therefore, as demonstrated in CHAPTER 7, Eq.(5.7) with C2 = constant may yield unrealizable predictions for Pe = 0 and U < co and, thereby, does not provide a physically acceptable closure for Eq.(4.1). Cinta and Tucker (1995) also developed a closure for < p p p p > that satisfies all six-fold symmety and six-fold contaction properties of the exact orientation tetad. They related their closure coefficients to the local properties of the microstucture by “fitting” model predictions with results based on a direct numerical simulation of the S-equation for homogeneous shear flows and for homogeneous extensional flows. The general realizability of the resulting orthotropic closure has not been determined. 43 5.3 Discussion Hand’s closure satisfies six-fold symmetry and six-fold contaction properties, but it is not realizable for finite values of U. Doi’s (decoupling) closure predicts realizable solutions to Eq.(4.1) (see APPENDIX D), but does not satisfy six-fold symmety and six- fold contaction properties. Fully symmetric quadratic (F SQ) closure and the orthotropic closure satisfy six-fold symmety, and six-fold contaction. In CHAPTER 6, the FSQ- closure coefficient C2 is related to the local microstucture so that < pp > is also realizable for all rigid rod suspensions subject to simple homogeneous shear. CHAPTER 6 REALIZABLE CLOSURE 6.1 Intoduction In this chapter, a realizable closure model for the orientation dyad < pp > is identified for the relaxation of anisotopic microstuctures in the absence of an external field (i.e., Pe = 0). A closure for the orientation tetad < p p pp > , defined by Eq.(5.7), will be completed by developing an equation for C2(IIb, 1111,) based on the condition that all initial orientation states on the boundary of the realizable region (see Figure 1.1) must remain either on the boundary or be attracted by states within the realizable region. Clearly, this condition requires (1 '11 p- go. (6.1) all boundaries of Figure 1.1 D: H In the above inequality, the vector _11 is an outward pointing unit vector perpendicular to the local tangent of the realizable boundary. The components of the vector E are the invariants of the stucture tensor: The vectors g1 and m are orthogonal unit vectors on the two-dimensional invariant plane (i.e., 911 “9111 =0). For Pe = 0, FTD = 1, and defined by Eq.(5.7), it follows directly from Eq.(4.1) that (see Appendix D): 45 dIIb 7 3 54 2 ——=—211 +2U —II +—III -—-II C 6.2b dt b [35 b 7 b 35 b 2] ( ) dIIIb dt = —31111, + 3U[ 315111, +3-an -3311bmbcz] . (6.26) 14 35 Eqs.(6.2b) and (6.2c) govern the relaxation of all orientation states. The objective of this chapter is to relate the second-order closure coefficient C2 to the local invariants IIb and III), by using Ineq.(6.1). 6.2 Realizable Isotopic, Planar Isotopic, and Nematic States The excluded volume term in Eq.(4.1) is zero at the isotopic, the planar isotopic, and the nematic states. This can be seen by evaluating < p ppp >:< pp > at these three states. An eigenvector representation for the orientation dyad < p p > is 3 3 3 (t)=Z?~pi(t)zsi(t)ri(t)=ZZ(t)§i 9,- (6.3) Eq.(6.3) is a representation of < pp > using the fixed mutually orthogonal base vectors g1, 92, and Q3. The instantaneous orientation vector p can be expressed as 3 3 g = 2140);, = 213101310). (6.4) i=1 k=l which implies that 3 3 3 3 _2 = 22p.(t)p,(t)s.9, = 225. (0'5: (081mm). (6.5) i=1 3:1 k=1£=1 Therefore, the average of Eq.(6.5) shows that 46 ~ (t)~ (0) O, ifk¢t (66) < = . pk p‘ 7.1,, ifk=t Note that

=0_ implies that =0 and =0. Also, tr=Z=X=Zxk =1 (6.7) i k k Eqs.(6.3) and (6.5) can be used to represent the components of < p p p p >:< pp > in terms of the eigenvectors of < p p > with the result that KBE> E< 2222K B> < (2251133 Kifij )(ZZEkF! £185! ):(Z)"pm 35me ) > I'd (6.8) Eq.(6.8) holds for all orientation states (see Figure 1.1). Isotropic States For an isotopic state, < p p > = 1° + 1° + 1.0 - 1 Thus Eq (6 8) reduces to l 2 3 — 3 o , o o [<2222>=<22>]L.mpic 222031 15ij 13m >33? mij 922611311me pm)>?£? a? (6.9) =§ZZ33? 1 J lulu-t =11. isotopic - belt—- 47 Eq.(6.9) implies that the excluded volume effect in Eq.(4.1) is zero at the isotropic state. This conclusion holds for the FSQ-closure for any value of C2(O, 0). Thus, the isotropic state is a fixed point (i.e., steady state) of Eq.(4.1) for Fe = 0 and U 2 0. Nematic States For a nematic state, =x§x§ and A? :43;- =O,k‘; =1 (see Figure 1.1). Thus, Eq.(6.9) reduces to ~ ... (6.10) =< 22133133 > = ZZ< 155133153 >883). I 1' At the nematic state, ~ ~ ~ ~ 1, if i = . = 3 < PinP3P3 > = 0 otherwise (6.11) Therefore, [:'<22>-<2222>=<22>]| =(z§z§)~(§§x§)—(L§z§)=g. (6.13) nematic This conclusion holds for the FSQ-closure provided C2(2/9, 2/3) = 1/3. Unlike the isotropic state, the nematic state is not a fixed point (i.e., steady state) of Eq.(4.1) inasmuch as the Brownian motion term is non-zero: 48 1 2 1 [QP‘gl] =§l§£§ 73%? +L‘2’L‘2’) (6.14) nematic Planar Isotropic States For a planar isotropic state, =§ngg + xgxg) and 1‘; = 0, A3 = 1‘; =% (see Figure 1.1). Thus, Eq.(6.9) reduces to [< 2222 >:< BE >]Iplanar isotropic 1 ~ ~ ~ ~ =§ZZZ< Pi Pjpm Pm >505? m i j 1 ~ ~ ~ ~ =§ZZ< Pi 15(me pm)>§?§§’ (6'15) 1 J m 1 ~ ~ 0 o =EZZ< pi pj >25 Zfij 1 J _1 _1 o o o o _§ -z(lz§2+§353)' planarisotropic Eq.(6.15) implies that the excluded volume effect in Eq.(4.1) is zero at the planar isotropic state: [<22>'<22>-<2222>=<22>]| planar isotropic (6.16) (8353 +§§a§)-(x‘2’2s2 +x§z§)-%(§§l‘2’ +2313) =2 This conclusion holds for the FSQ-closure provided C2(1/6, —1/36) = 1/2. This result shows that the planar isotropic state is not a fixed point of Eq.(4.1) because the Brownian motion term is not zero [-ll] _... 3: (533.3 +£33.54). (6.17) planar isotropic 6.3 Realizable Prolate and Oblate States The prolate boundary is defined by (see Figure 1.1): III 2/3 11b = 6 ('61) , o s 1111, 5 2/9. (6.18) The outward pointing normal vector on the prolate boundary is given by Ep =+D¥I§H ++n¥n§nl (6.193) where up a — 1 (6.1%) 11 2/3 4 6 1+— — J 41111)) 1/3 2(3) 3 III nfn .=. + b (6.19c) 4 6 2/3 1+— _— 9 111,, Note that 111, pp =(ni’1)2 + (n‘l’n)2 =1. On the prolate boundary, the components of dE/dt are defined by Eqs.(6.2b) and (6.20) with 111, and HIb related by Eq.(6.18). It is noteworthy that Ineq.(6. 1) on the prolate boundary (F -line on Figure 1.1) reduces to 1:1 which implies that initial orientation states on the prolate boundary remain on the prolate z 0 , (6.20) prolate boundary boundary for all time without any additional conditions on C2(Ilb, 1111,). A similar result holds for the oblate boundary (see Figure 1.1): 50 =0. (6.21) 1:1 The analog of Eq. (6.18) for the oblate boundary (see Figure 1.1) is oblate boundary — — 5 111b s o (6.22) _mb 2/3 1 ’ 36 Hb=6( The outward pointing normal vector on the oblate boundary is E0 = +nfign + +ncfilgln (6.233) where o _ 1 1111 = - 2/3 (6.23b) 4 6 1+— —— J 9( 1111)) 1/ 2[__6_] 3 3 111 ° — b (6.23c) Note that no 3110 = (n a)? + (ni’n)2 =1. Eq.(6.21) implies that initial orientation states on the oblate boundary remain on the oblate boundary for all time without any additional conditions on C2(Ilb, 1111,). 6.4 Realizable Planar Anisotropic Boundary The planar anisotropic boundary is defined by (see Figure 1.1): 2 1 2 H =ZIII +—- , —— SIII _<_—. 6.24 b b 9 36 b 9 ( ) 51 The outward pointing normal vector on the planar anisotropic boundary is given by npa = +n {la 911 + “11131:; 9.111 (6.25a) where pa 1 nII = — (6.25b) 5 11pa = ——2— (6.25c) III \f5— Note that npa ~npa =(ng’la)2 +(n11’fi)2 =1. On the planar anisotropic boundary, the components of dE/dt are defined by Eqs.(6.2b) and (6.2c) with IIb and IIIb related by Eq.(6.24) above. Ineq.(6.1) on the planar anisotrOpic boundary has two contributions (see APPENDD( F): npa E - 51—:- = 1111): £1 + III d mb (1 t . . d t d t planar amsotropic pa =2121:1111){mum-8:511:+18<22<1+H1>1> 0) show that this important feature of the FSQ-closure is satisfied by a wide class of planar anisotropic states. Figure 6.1 illustrates the behavior of C2(IIIb) for —1/36 S III}, S 8/36. A major hypothesis for the theoretical results developed in CHAPTER 7, 8, and 9 is that the FSQ-closure coefficient defined by Eq.(6.28) above applies for all anisotropic states. Although the Doi-closure (see Bq.(5.9)) does not satisfy all the six-fold symmetry and six-fold contraction properties of < p 2 pp > , Eq.(4.1) nevertheless yields a realizable orientation dyad for Fe = 0 and < pppp > = < pp > . Figure 6.2 shows that how the excluded volume (EV) terms that appear in Eqs.(6.2b) and (6.2c) vary over the planar anisotropic boundary of Figure 6.1. Note that the EV-terms are zero for the Doi-closure and the F SQ-closure at the planar isotropic state and the nematic state. The Doi EV-term in Eq.(6.2b) is significantly larger than the FSQ EV-term in Eq.(6.2b). This indicates that the Doi-closure has a higher tendency for self-alignment. This conjecture is confirmed by the equilibrium calculations presented in CHAPTER 7 inasmuch as the Doi theory predicts biphasic phenomenon at smaller values of U. 53 388-0% 65 é 35 No 666E660 8:66 8 65mm 95 cm m N m m v m N F o _n v .u u T u u u u 0 o _ _ _ . _ u : to $35 38581.3 E: 8.: 3QO T 33% ofiobofiam Sum—Q 8m ND 68050an noewomom “ " moan: 030.583..» Sana . 3:5 + $3 1 Nu m .. No a + I E? w m «U m: 1 Nu m .. no IIIIIIIIIIIII _ _ _ _ m m: vd 8a u QE .3 1 DE mOHSm OBNEDH— N\— H ND | l l r. lllll m.O \ 63.. u 3E a: 1 pa mesa. cacao: 28% 54 388-0% 05 was soummeoamdm wanna—coon ofi Sm 398:5 0823/ wowfioxm 2E. m6 oSmE 50m m n o m w m N F o F- n( ..u cod . mod H x m. . o3 m. 9 p - 3d M .. owd a .d w . 9 . mm o W. A v>m 0mm... . m. 63:. . . . . ) $332.. Oman? . on o M 95:33 weasooow 1? . . ( 333$ wagssoon? ., , u U H u ‘ w ._ . . mmd L ovd 55 CHAPTER 7 MICROSTURCTURE IN THE ABSENCE OF AN EXTERNAL FIELD 7.1 Introduction In the absence of an external field, the relaxation of < pp > is governed Eq.(4.1) with Fe = 0: d 1 d-t— =FTD[—(-‘3‘l)+U(-z)). (7.1) The Fm -factor in the above equation accounts for the tube dilation phenomenon (see Eq. (2.6)). The first term on the right-hand-side of Eq.(7.1) represents rotary Brownian diffusion. The second term accounts for the excluded volume phenomenon. The dimensionless group U measures the relative importance of self-alignment and rotary Brownian motion. Experimental studies for rigid rod suspensions, such as lyotropic liquid crystalline polymers, show that a transition fi'om an isotropic state to an anisotropic state occurs at some critical concentration (Abe and Yamazaki 1989a, 1989b; Farhoudi and Rey, 1993; Kubo and Ogino, 1979; Murthy et al.,l976; Orwoll and Vold, 1971; Robinson, 1966; Sartirana et al., 1987; Srinivasarao and Barry, 1991). The equilibrium orientation state is isotropic for dilute solutions (U << Uc) and anisotropic for concentrated solutions (U >> Uc ). Uc is a critical value of the nematic coefficient that depends on the concentration of the dispersed phase. The objective of this chapter is to determine the effect of U on the steady state solutions of Eq. (7.1) for the FSQ-closure developed in CHAPTER 5 and CHAPTER 6. 56 7.2 Biphasic Phenomena The asymptotic solutions of Eq.(7.l) are steady equilibrium states for all U. Clearly, the Fm-factor does not affect the steady state solutions. In APPENDIX B and APPENDD( D, the following equations for the second and third invariants of the structure tensor 2 are derived from Eq.(7.1): d—HL- = -2FTDIIb + ZFTDU —7—Hb +3111}, —fiIIb2C2 :| (723) 35 7 35 dmb = -3FTDIIIb +3FmU[ 375-1111; +finb2 -%Hbmbcz]- (731’) The steady state solutions to Eq. (7.1) have two equal eigenvalues (see APPENDIX D). This means that the equilibrium states in the absence of an external field are either prolate states or oblate states (see Figure 1.1). Thus, application of Eq.(3.6) implies that the steady state solutions to Eqs.(7.2a) and (7.2b) can be represented 2 in terms of the order parameter or. defined by 11b = i— (a) . Since the equilibrium solutions of Eq.(7.2a) and Eq.(7.2b) are either prolate or oblate states, Eq.(3.6) can be represented in terms of a and 1111,: 111b = g (:t (1)3 . The positive sign is for prolate states and the negative sign is for oblate states (see Figure 1.1). It follows directly from Eq. (7.2a) that the order parameter is determined by the following algebraic equation: 0: l—U[l-+-l-a--3-§a2C2]. (7.3) 35 7 35 Eq. (7.3) has three solutions: 57 a=0 l/7i‘fil/7)2-(144/35)C2(-113-—%). (7.4) (72/35)c2 a: The solution a = 0 corresponds to the isotropic state. This steady state may be stable or unstable, depending on the value of U. If the steady state is on the prolate boundary of Figure 1.1, then the positive sign of Eq.(7.4) applies. For oblate steady states, the negative sign applies. Previous application of the FSQ-closure assumed that C2 was constant and independent of the microstructure (Imhoff, 2000; Imhoff et al., 2000; Kim et al., 2001, 2002, 2003, 2004, 2005; Kini 2003; Kini et al., 2003, 2004; Mandal et al., 2003, 2004; Nguyen, 2001; Nguyen et al., 2001a, 2001b; Parks et al., 1999; Parks and Petty, 1999a, 1990b; Petty et al., 1999). Figure 7.1 shows how U influences the steady state order parameter a for C2 = constant. For U < U1, all steady states are isotropic (a = 0). For U1 < U < U2 , three steady states exist: two stable and one unstable. The unstable state is on the prolate boundary. The region U1 < U 1/2, but it can not cover all possible orientation states inasmuch as the order parameter a S 0.78 for C2 5 1/2. Nevertheless, this result provides the possibility that the FSQ-closure coefficient C2 can be fitted to the 58 .3838 M NO Sm _oucEAvmm 05 3 @8235 83% gunfincm 032:2 fin oSwE D N _ 2: S D a F p n _r O®.OI 033:; ........... .I/l I.I..I.I.I.I.I.I.I.I.I.l ; h . 98- .k/ n u .l/ n u 3me §85> can on< 35m 2290 / /. m m I 5&8 3%an r ONd- 808598 0 M M N: III . D m: I 33.». 030.50% \ m and Iul 3.. .. ON C We ONO I0I a. 0.4.. 0 8m 2a 9a a . (I eduaolol s _ . .25 U , .. cod :\, mud I \ u omd \ mad I I I c . IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII .. co _. 59 experimental results with C2 = C2 (111, ,IIIb) (see Abe and Yamazaki, 1986 for the experimental data and CHAPTER 6 for the FSQ-closure coefficient). Analogous to Eq. (7.4), the order parameters for the decoupling-closure and the HLl-closure are (Chaubal et al., 1995): a = 0, -1— 1:31 ’1 — —§— for decoupling-closure (7.5a) 4 4 3U a: 0, %i%‘/49—2—:J—0 for HLl-closure. (7.5b) Figure 7.2 shows the comparison between decoupling-closure, HLl-closure, and the F SQ-closure developed in CHAPTER 6. The phase transition from the isotropic state to the anisotropic state appears in both closure approximations. The figure also shows the biphasic region for the decoupling-closure, HLl-closure, and FSQ-closure. The biphasic region has two stable states and one unstable state that can coexist for the same U. The isotropic state is a steady state solution to Eq. (7.1) for all three closures. The equilibrium orientation state is the anisotropic state if the initial condition on a is above the unstable state. When the initial condition on a is below the unstable state, the orientation state relaxes to the isotropic state. The locations of the unstable states were determined by solving Eq.(7.l) as an initial value problem. Although the qualitative trends of each closure are similar, the quantitative differences are significant. The decoupling approximation predicts the existence of prolate states for which a(U) —> g as U —-> 00 (see Figure 7.2). However, the biphasic transition point is lower than the other models (U2 = 3). The HLl—closure, however, has 60 .EE No I No 5 688-0% .5 .3 328$ .33 82:953.» .3232 2 22mm D 3 N? or m m v N o u q q _ q u q - omAul 2... .. ......... .32.... . e..- I .. .93- - x . , m m .. m mI 8&2 . 8.0- /./ u H 23.33 . ME I 00.9 \ \ \ 7.....2... . 3:50 .Omm II 88m oaobofl m u: mu: a 038mm: ._ 0N6 $2 “.55 ream o .. .. 65 33 2.. 655 S: a . 23 8 ms 968 . .. 5.2. e ._ an . / as. . \ s . 8 o \. v UflOum Omnfluw u\.\1 n IIIII . u . 8.0 ......................... I . 00... 033283: 61 the same U2 as the FSQ-closure, but the HLl-closure predicts unrealizable oblate states for U < 00. The FSQ-closure with C2 (1111,) defined by (6.12) satisfies both prolate and oblate realizability conditions. Therefore, the decoupling approximation and the FSQ- closure are realizable for all equilibrium steady states. Tables 7.1 and 7.2 show how U influences the order parameter and the invariants for the equilibrium states. 7.3 Relaxation to Isotropic and Anisotropic Steady States When U and PC are zero and Fm = 1, Eq.(4.1) reduces to Nguyen (2001) discussed various aspects of this equation. An analytical solution to Eq.(7.6a) is (t)=§l+((O)—%l)exp(—t). (7.6) Eq. (7.6a) becomes an isotropic state (< pp > = 1/3) as t —> 00. The structure tensor corresponding to < p p > (t), defined by Eq.(7.6b), is gm= 2(0)exp(—t) (7.7) The two invariants, IIb and 1111, of g are 111, = tr(b Q) = Kb (0) exp(—2 t) , and (7.8a) 1111, = tr(g 111:3) = mb (0) exp(—3 t). (7.8b) Eqs.(7.8a) and (7.8b) imply that 62 111. Table 7.1: Invariants and Order Parameter of the Equilibrium Structure Tensor on the Prolate Line mb 0.0221 0.0013 0.0682 0.0073 0.1640 0.0271 0.231 1 0.0454 0.2828 0.0614 0.3239 0.0752 0.3572 0.0872 0.5511 0.1670 . = (3/2 IIb)1/2 63 Table 7.2: Invariants and Order Parameter of the Equilibrium Structure Tensor on the Oblate Line Hb IIIb 0.0000 0.0000 -0.2744 0.0162 -0.3200 0.0352 -0.3510 0.0509 -0.3728 0.0637 -0.3 890 0.0741 a0 = — (3/211b)“2 3/2 M] (7,.) III. (t)=III.<0)[ IIb(0) Figure 7.3 shows that all planar anisotropic states relax to isotropic states on a time scale of order 4tc if U = O and PC = 0. If U > 0, the relaxation trajectories in the invariant plane depend on the closure approximation and the excluded volume potential model. When the initial state is planar isotropic (see Figure 1.1), solutions to Eqs.(7.2a) and (7.2b) remain on the oblate line. All other initial conditions relax to the prolate boundary. Figure 7.4 shows that transient solutions for U = 0 and U = 3. Both solutions relax to the isotropic state. However, the transient solution for U = 3 is attracted towards the nematic state before it reaches the isotropic steady state. When the rigid rod suspension is concentrated, the orientation state is anisotropic state (U > U 2 = 5). Tube dilation does not affect the steady state solutions. However, as illustrated by Figures 7.5 and 7.6, tube dilation makes the orientation state relax faster to the equilibrium state. In addition, the relaxation time increases as the orientation state is closer to U1 , and it decreases for higher U (see Figures 7.5 and 7.6). 7.4 Discussion The microstructures of rigid-rod particle suspensions predicted by the FSQ- closure are realizable for all values of U. For U > U1, the orientation structure parameter a increases as U increases. For U —> 00, a -> 1, which is the nematic state (see Figure 1.1). These predictions are consistent with other closure approximations and experimental observations. The decoupling approximation with the Maier-Saupe potential has U1 = 8/3 with 65 0.7 - 0.6 - 0.5 - planar anisotropic states 0.4 - Kb 0.3 - (a) relaxation of planar 0'2 ' anisotropic states 0.1 note: time increment between points = 0.09t 0'0 0.00 0.05 0.10 0.15 0.20 0.25 KL, 1. 0.9 0'8 (b) relaxation of the nematic state 0.7 0.6 < Pi P j > 0'5” 0.4 0.3 0'21. ——& 0.1 o 1 2 3 3, 5 6 7 = E/Ec Figure 7.3 Relaxation of the Microstructure due to Rotary Brownian Diffusion (U = 0; E =1/(6D0R) ). 66 .2 u Em . 38:.de 28258822 05 we cosmos—om 05 so 0823/ wouioxm mo “comm v.5 oSwE .25 mod mod rod o rod- Nod- mod- . . . . . 85 59o u 852” :3an «58805 08: .. mod \ . w ofio m n D . a: :m . 2d 5N : fl cud / u a . n a o E QN z .. mNd 83:8 3:5 67 0.25 0.20 Fm =1 11b 0.15 .‘ . 3 _2 l ------- F1196" = (17111,) , p. 360 Doi & Edward (1986) 0.10 - 0.05 . III], Figure 7.5 The Effect of Tube Dilation on the Relaxation of the Microstructure (FSQ-model; U = 3; initial conditions: IIb(O) = 2/9; IIIb(O) = O). 68 .85 u 88 888-8% 8 08C. “Reflexion 3658830 2: :0 8:25 093. 28 D we 80am oi. cs 853m 0 ‘— 4-«n D common 38an :17 f 8.0 . co... ‘. .- _..o u 00—. mm venue—e J T :2on 33.35 .. coo? 69 a = 1/4. The second critical U for the decoupling approximation is U 2 = 3 with a = 0 and 1/2. The HLl-closure predicts that U1 = 240/49 with a = 1/8, and U2 = 5 with a = 0 and 1/4. The FSQ-closure predicts that U1 = 4.73 with O. = 0.167, and U2 = 5 with a = 0 and 0.321 (see Table 7.1). de Genne (1974) estimated a value for U2 (5 4.55) from the orientation density function. Chaubal et al. (1995) also calculated the order parameter curve based on the density function. The HLl- and F SQ-closures agree better with the transition point than the decoupling approximation. However, the anisotropic transition fits the decoupling approximation better (i.e., the curve is much steeper in both the exact and Chaubal et al. solutions). The I-ILl-closure becomes unrealizable on the oblate boundary. Using other types of excluded volume potential model, qualitative difference with Chaubal et al. can be modified, but U2 is much higher than other approximations (see Ilg et al., 1999). The computational results are based on initial condition for < pp_ > with only diagonal components in the planar anisotropic state. When U = 0 and Pe =0, Brownian motion is the only driving force that makes the orientation state random. The nematic potential coefficient U influences the anisotropic orientation state as well as the equilibrium isotopic state. For U > U2 , all the steady states are either on the prolate or on the oblate boundaries (see Figure 1.1). In addition, relaxation experiments provide a means to determine the rotary diffusion coefficient DR Previous studies have estimated DR values by fitting the first normal stress difference N1 with computational results. Beak and Magda (1993) have reported that the DR coefficient for PBLG solutions is about 2 s'l. With DR ~ 2s‘1 , the time scale for relaxation to an equilibrium state in the 70 biphasic region may be as long as 10 — 100$ inasmuch as tc 5 10 to 100 (see Figure 7.6). Figure 7.6 may be used to design an experiment to estimate the rotary diffusion coefficient by measuring the relaxation time of rigid rod suspensions are a wide range of concentations inasmuch as to ~1/(6Dfi) . 71 CHAPTER 8 MICROSTUCTURE INDUCED BY HOMOGENOUS SHEAR 8.1 Intoduction The relaxation of < pp > (t) in the presence of a homogeneous shear field (7 = constant) will be examined in this chapter by solving Eq.(4.1) for Pe > 0. The velocity gradient for the rigid rod suspension is Vfi=tsysz- (8.1) ~ . 1 §=§M =§(§y§z +§z_e.y) (8.2) W-Wl'-1 83 =—= 7’5 sygz-gzsy). (.) The objective of the chapter is to explore the effect of the tumbling parameter 7L on the microstructure for a wide range of U and Pe. Appendix B shows how the director of < pp > can be calculated in terms of the components of < pp > . The motion of the director relative to the fixed flow direction (3 = quZ) will be examined for L/d = 00 (Section 8.2), L/d =12 (Section 8.3), and 0 _<_ L/d < 00 (Section 8.4). In Section 8.5, the influence of tube dilation on the microstucture will be discussed. The results of this chapter are based on the F SQ-closure for < p p p p > defined by Eqs. (5.11) and (6.12). Eq. (4.1) is integrated by using a fourth-order Runge-Kutta algorithm (see APPENDIX G). 72 8.2 Relaxation of Planar Anisotopic States for L/d = 00 For L/d = 00, 7. = 1 (see Eq.(2.4)). For this case, Eq. (4.1) with E = 1, Fm = l, and 8 Q defined by Eq.(8.l) above predicts that all realizable orientation states (see Figure 1.1) relax to steady states for U 2 0 and Fe > 0 (see CHAPTER 7 for Fe = 0). If the initial condition for the director (i.e., the eigenvector associated with the largest eigenvalue of < pp > ) is in the deformation plane, then the steady state is unique. Figure 8.1 shows the effect of U and Pe on the steady state angle between the flow direction ( u =uzgz) and the director (cost) = x3 o_e_z ). The director angle decreases monotonically as U increases for fixed Pe, and as Fe increases for fixed U. The results show that hydrodynamic coupling (l-parameter) and the excluded volume phenomenon (U-parameter) cause anisotropic steady states. For U S 3 and Pe —> 0, Figure 8.2 indicates that the microstucture is isotopic (i.e., IIb = 0 and UL, = 0). As Pe increases, anisotopic steady states occur. For fixed values of Pe, the microstucture becomes more nematic-like as U increases. This occurs because the excluded volume term in Eq.(4.1) mitigates the return-to-isotropy due to rotary Brownian motion. Thus, the flow-induced alignment mechanism and the excluded-volume effect simultaneously promote nematic-like microstuctures. The excluded volume mechanism makes the microstucture more prolate axisymmetric (see CHAPTER 7) whereas the hydrodynamic coupling of < pp > with g tends to make the microstructure more isotopic. 73 .2 u a x u at 888.3% 8 n 34 tom 2»: Houoohfl 33m 33on 05 no em can D we Hoobm 2D. 3w own—mi a om D 000? 00—. or _. D D D v.0 .fl 74 2%: _l....... ; o . xxx" 3 on _rooo“ ooooooooooo u u n u A on u .. “'1 74 .2 n & x .I. Em £03008de 8 M ED 8m egosbmoeomz 88m 38on 06 no cm 98 D we 80am 2D. Wm 8:me pH: wvod wmod mmod m 3.0 mood No .. N _ mm m H D . D m e\c o u D . o v .. 4 m .. . m 2 ._ 2.8.2 8e 888 . ofiogmfim 838a o t c a n. a . . / . o n a .w u on 0 \ m H D «a H Dam od- .o .. mod r _..o .. m...o .. Nd u mmd a: 75 8.3 Relaxation of Planar Anisotropic States for L/d E 12 The tumbling parameter k in Eq.(4.1) plays a significant role in the relaxation of the microstucture. This is anticipated by Eq.(2.2), which predicts that a single rigid rod in a steady shear flow will tumble continuously if | A | < 1 (see p.449 , Larson 1999). For U = 0 and l 5 Fe _<_1,000, Table 8.1 compares the invariants of the structure tensor for [M = 00 (A. = 1) and L/d = 12 (2. = 0.987). For 7» = 1, the steady state invariants approach the nematic state (see Figure 1.1) for large values of Pe. However, for Fe = 1,000 and 1. = 0.987, the microstucture of the suspension is less nematic (11b = 0.48, 1111, = 0.13). For 7t = 0.987 and U > 27, Eq.(4.1) predicts director tumbling for Pe a 10. For this case, Figures 8.3 and 8.4 show that an initial planar isotopic state (see Figure 1.1) relaxes to a periodic state characterized by director tumbling with a frequency of ft 5 6D°R1rl15. The initial condition for < pp_ > (0) is planar isotopic: 2<22><0)=@y2y +2.2.). (8.4) The eigenvectors of < pp > (0) are in the deformation (or shear) plane (see Figure 3.1). The director tumbles 180' because there is no distinction between the head and the tail of a rigid rod. Director tumbling occurs for two reason: 1) the excluded volume phenomenon mitigates rotary Brownian motion and, thereby, reduces the intrinsic diffusive torque on the microstucture; and, 2) the torque on the microstructure due to particle coupling with the shear rate is weakened by a reduced tumbling parameter (i.e., A = 0.987 < 1). As a consequence, the torque due to the antisymmetric component of the velocity gradient can 76 Table 8.1 Invariants of the Equilibrium Anisotropic Tensor for Different Tumbling Parameters (FSQ-model; Fm = 1; U = 0) 7. 1 7. = 0.987 P6 111, 1111, 11b 111b 1 0.0412 0.0017 0.0402 0.0016 5 0.1754 0.0249 0.1686 0.0232 10 0.2588 0.0488 0.2458 0.0447 50 0.4477 0.1202 0.4062 0.1029 100 0.5099 0.1475 0.4475 0.1201 500 0.6020 0.1905 0.4785 0.1335 1000 0.6240 0.2011 0.4801 0.1342 00 0.6667 0.2222 0.4805 0.1344 77 .A 26.8 + 3} u é A mm v N ”2 u 6.18 u a .58 u a x u 8.2 888-36 Q m EA Sm wEESPH .8885 8% 83035832 30282885 Wm 053m H odv Wmv 0.9» 934 93V mdv ode. vw.o - n b - Hr o . 1 8 8 8 cm o now $0 28M.a m J 14 g D .8 3o . .8 2.9. . .8 . cm omo .. .8. “mod H 3509 502509 808805 2:: 82. u om mad. FFIIFFJS .............. .. o2 wad Hr ...... ONF 8o . \ . o: m o® IV 86 .. +| .. \ .8: OO.—. I -----unnuuuuuuuunflnununuu IOQF 78 .A Nana + 88.8 n 80 A mm v N NE u 8 .8 n 0 .88 u .a x n Em 888-86 Q m 03 80 883 883 05 E 00:083. .8885 80 83008822 1w oSmE .000 0N0 0N0 3.0 9.0 00.0 00.0 00.0- p n p n n n 00.0 “no.0 .1. $500 58300 288805 080 .E . . 0_. 0 23 «:8 :3 23 83 . . . _ . onto - 0N0 «8.0 on: 8280 88 a: n 00.0 >\ 84.0 a: \ \ 8... . - 8.0 84.0 x\.u\ . DE 1... . 8.83 9.828 $82.0 882.0 I 00.0 . . . - _ «8:3 05—2280 \4 88:3 1 00.0 \ . a: 823° 2 0 5.0 vmwwnmd 79 sustain the phenomenon of director tumbling (see Eq.(4.1)). The dimensionless tumbling frequency ft aft/6D?l depends on U and Pe. For U > 27, Eq.(4.1) predicts that ft increases as Pe increases. This conclusion is consistent with previous theoretical and experimental studies related to director tumbling (see p.280 and p. 463, Larson, 1999). Figure 8.3 shows that the director rotates slowly when it is nearly aligned with the flow direction and rapidly rotates as it crosses the vorticity/cross-flow plane (see Figure 3.1). The temporal response of the eigenvalue associated with the director is also shown in Figure 8.3. It is noteworthy that 13 (t) is a minimum during the rapid tumbling phase of the motion. The invariants of the structure tensor for this example are shown on the invariant diagram, Figure 8.4. Clearly, Eq.(4.1) together with the FSQ—closure yields a realizable orientation dyad for director tumbling. Note also that the initial planar isotropic orientation state rapidly locks onto the tumbling orbit for ts 0.09/6Dfi . For A. = 0.987 and U = 27, Eq.(4.1) predicts that director wagging will occur for Pe = 24. This phenomenon is illustrated by Figure 8.5, which shows the angle between the fixed flow direction and the director. For this case, the wagging frequency ?w z 6D‘fin/l 1. Note that the director eigenvalue 13 is a minimum as the director wags about the flow director. A planar isotropic initial condition is also specified for this calculation and, as illustrated by Figure 8.6, the orientation state locks onto the periodic wagging state within t5 0.09/6Dfi. Eq.(4.1) predicts that f 5 f‘w/6Da —-> 00 for Pe E 30 and U a 27. As the Péclet number increases, the microstructure approaches the fully aligned state (Point A, Figure 1.1). The coupling between the strain rate dyadic and the 80 .8 £8. + 8888 u 80 A mm v N am u 98 .8 u D 53.88 n ,8 m8 u Em m8255.380 Q m 883 8‘8 wEwmmB 88885 .80 830858822 3028888888882: Wm 0888mm.» a 0.0m QNN 0.NN m._.N 0._.N 00.0 8 8 8 00.. OF.O I 8'55 IONI 0N0 u a a n 88.0 n 388828 888383 888080.805 088888 . .... .2- 000 . : Ill! Inuunnnnunnnnnnuunnnnnfl O 0.90 uuuunuun, cm 88.8 com- \ one - 2.; ll .9 an. I Owl . u a :9 mtenn .8 . 8o .. . ca .8 050 r a J 9 fi M = 4 on 8.88 . _ ,8 _ _ TN / .. 8 OO.O I \oc 4..uuuuuununuunoununuuuuuuuuuuuuuuunnonon.O.v -- .......... ““‘-‘-““‘ 00. _. om 81 1%.“... + 8.888 n 80 A mm v N an n 0.8 .2 n 88 53.8. n .8 88 u an 880858-80 Q m 03 .80 SSE 8083 05 E 08880me 88885 80 830858.822 0.0 8:03 .888 0N0 0N.0 08.0 0.0 00.0 00.0 00.0- F P b n n - 008° .40 800.0 a 8500 8833800 8880.85 080 u 0 _. .0 055.0 3.05.0 «502.0 035.0 s . . 05050.0 . 0N0 \ . 3050.0 \ . 2.2.3 .8: .. 00.0 . ..888 a: . ~85 o .. 0V0 000.0. 0.0.0. 03.0. - 8888.0 . . . 38.0 I 00.0 on: 82080 . 008.0 \ 888 . 00.0 \ . 83 . 0nd . 0:20 82 orientation dyad weakens as (11b,IIIb) —> Point A. Thus, for U = 27 and Pe > 30, the diffusive torque balances the hydrodynamic torque to produce a steady state microstructure. Figures 8.7 and 8.8 illustrate this phenomenon for 7t. = 0.987, U = 27, and Pe = 95. A planar isotropic initial condition with the director in the shear plane was also specified for the calculation (2 < pp_ > (O) = ey e + 92 e2 ). The director angle relaxes y rapidly fi'om its initial condition of 45' and fluctuates around the flow director. Eventually the director attains a steady state with a small negative offset from the flow direction. Figure 8.8 shows that the invariants of the structure tensor are close to the prolate axisymmetric boundary. It is noteworthy that for U = 27, Pe = 95, and 7t. = 1 (rather than 0.987), the director also attains a steady state, but with a small positive offset fi'om the flow direction (see results on Figure 8.1). Figure 8.9 gives a phase diagram for A = 0.987. The diagram was constructed by integrating Eq.(4.1), F”) = l, and the FSQ-closure from a planar isotropic initial condition (2 < p_p > (0) = eye + 9292 ). The asymptotic state depends on U and Pe. Y Three possible states were found: 1) steady states; 2) periodic tumbling states; and 3) periodic wagging states. The boundaries shown on Figure 8.9 were determined within (APe, AU) = (1, 1). For example, at Pe = 10, steady alignment occurs for U = 25 and director tumbling occurs for U = 26. For U = 27, Fe = 95, and 7L = 0.987, Figures 8.10 and 8.11 show how a planar isotropic state with an initial director colinear with the vorticity (see Figure 3.1) relaxes to its steady state. The initial condition for the orientation dyad is 2 < pp > (O) = ex ex + 9292. The eigenvalues and eigenvectors for this initial condition are 83 .8 8888 + 8888 n 88 A mm v N .3 u 8.8 .R n 88 .528 u .8 m8 u 8.88 88888-30 3 m 05 80 8880888082 .8880 888.85 800 83258822 38888888qu 0.0 88808.0 H 00.8 00.0 00.0 0Y0 0m 0 00.0 8.88 . . . . . m- O—‘80 _. H H H H H H H H H H H H H H H H H H H H H H H o.“ ............... l o 0N0 .. . u a In 800 0 n 8500 8802583 88880803 08: \ one . .88 ll. .. .2 3o . .. .2 00.0 . . cm 00.0 . mm 05.0 . 00 All 3. 00.0 . 00 8.88 . / .. 8 84 .8 88.88 + 888.8. n 88 A mm v N .3 u 8.8 .8 u 88 8883.88 u .8 m8 u 888.8 88888-30 N8 m 3.8 80.8 288880 8888888 080 888 888088888082 50080 88885 888.8 230858.832 0.0 088800.88 8888 mud 0N0 08.0 08.0 00.0 00.0 00.0.. r n n p F - 008° 8888 00.0 n 3500 883383 888080.803 28888 508.0 00.20 00.20 808.0 008.0 .. owd . . . . . $0.0 88.8. n 8888 2.2 n 888 . 3.3 . 00.0 . 00.0 .888 . 0.88.0 . 00.0 . 00.0 - 05.0 85 .8 N88 + 8888 u 88 A 88 v 8 m888.88 u .8 M8 u 8.88 882888-080 88 .u. 888.8 88 8.8 8a.W 8880 as8.888 38.8.8 8.8 8588.8 on 00 05 00 00 0.8.. 00 cm 0 _. 0 D n D n b bl - n o . 08. 888688808888 50888 D . 0m cm .I. D . 00 050083 0888383 86 .88N N8+8 -8 n 8888 A .88. v 8 m888.88 u .8 m8 u 8.88 888 n 8.8 888 u 88 m8885888888888 S m 888.8 888.8 088808-0884 88885 8852888888888 08 .0 2880888 8 00.8. 00.0 00.0 05.0 00.0 00.0 010 00.0 0N0 0.30 00.0 pl 0 n D I. D I D D D I ONoOI 800.0 n 8888888 888832,. 8880898888.: 0888 8.88 N- - - .. .. .. . . x 88 n W8” 8888.8 + 8... 8888.88 u 888888 .88. 88- 828888 / A 8988 888 o 88 n 8988 N888? + 8.8 8888.88 u 88088 .8888. 88- 82898 N8. 8.88 . 888.88 0 N8 A 8 H xmhfi 0 H fihfi Nmmmod + 88m 500.0 N Acov—m a8880.0 H 0800‘ 00.0 r 0 "8880.5 . 888.88 / 88 8.8 II 0N8.8mnw.~.m N ofomwoo+88N88No8888+ ~8.8n88.8m /% 888 .8888 - - - - I - -. m No8ov00.0+88o.8ov80.0+ 0A8 8 8.8 88 8888.88 "888 A 88. v 87 .8N8N8 + 8888 n 8888 A 88 v 8 m888.88 u .8 88 n em 888 u 8.8 888 u 88 8088-088: 88 m 888-8 88 2888.8 08888.8 288 88 888888-83 82885 88.8 288888 8888 0N0 0N0 m _..0 2.0 00.0 00.0 00.0- n - n n 00.0 - 35.0 H 350m GDOEOD «€080.85 08w“ - 0820 - 0N0 888.88 ".8888 88.8.88 .- .888 . 888.88 - 088.0 - 00.0 - 00.0 - 05.0 88 XI (0) = 0, 2&1 (0) = an 8 (OH/2 3.2 <0)=sz (8.5) 13 (0)=1/2 233 (0) =gx. The director 53 is initially aligned with the vorticity V xu =3 = wxex and as indicated by Figure 8.10, remains colinear with _w_ for the entire relaxation process. The steady state invariants of the structure tensor are Kb (00) = 0.492 and 111-, (00) = 0.138 (see Figure 8.11). It is noteworthy that this steady state is different from the one that deve10ps from a planar isotropic state with an initial director in the shear plane (see Figure 8.7 and 8.8). Thus, if U = 27., Fe = 95, and A = 0.987, Eq.(4.1) predicts the existence of two steady states: 1) 11b (00) = 0.546, H11, (00) = 0.165 (see Figure 8.8); and, 2) 11b (00) = 0.492, Int, (00) = 0.138 (see Figure 8.10). Figure 8.10 shows that g1 '92 = $2 -§y = 0.052 for t —+ 00. The rocking motion of the two eigenvectors _igl and 2&2 around the director g3 (see Figure 8.10) has been termed log-rolling (see p.450 Larson, 1999). The three eigenvalues associated with the steady state orientation dyad < pp > (oo) are ll (00) = 0.014, 282 (00) = 0.083, and 13(00) = 0.903. Thus, a relaxation process with the initial director colinear with the vorticity (i.e., 3 3 (0) ~ v_v) produces a final microstructure which is less nematic than a process with the initial director colinear with the cross flow direction (i.e., £3 (0) ~ 3x31). Figure 8.12 gives a phase diagram for k = 0.987 for a planar isotropic initial condition with the director in colinear with the vorticity (see Figure 3.1): 2 < pp > (0) = ex ex + gzgz. The 89 .8N8N8 + 8888 n 8888 A 88 v 8 m888.88 u .8 88 u e88 8.888-088: 88 m 888.8 88 88888880 8888888 28888 8888888888088 80 88888888 38.8.8 88 .8 2888.8 mm 00 _. 00 00 088 0N 0 D D n P n o 8880888888088 .8838 AIIII. 8888:8808 u 0_. .. 0N 8880888888888 8888888 All 8888:9882 08 n D ----------- D u 00 mm H D 88888me ...... . 8883883 .. 0v. 8882888808 .8 88 8888 All 8888 888-80 .8 88 .0 8 888 u D -------- .00 90 asymptotic states depend on U and Pe. Three possible states were found: 1) log- rolling/steady alignment; 2) periodic tumbling; and, 3) periodic wagging. Figure 8.13 shows the relaxation of an anisotropic microstructure with the initial director located in the vorticity/flow plane. The initial condition for the orientation dyad is (0)=-5—e 2 +—2-§ s +-5-e 9 +i(§ 2 +§ 9,.) (8.6) The dimensionless groups U, Pe, and 78 are the same as the relaxation process illustrated by Figures 8.8 and 8.10 (i.e., U = 27, Pe = 95, and 7t. = 0.987). The initial condition for the eigenvalues and eigenvectors for < p p > (0) are 2x-(0)=1/6, 51(0)=-§y, 12(0)=2/6, 8 (0)= J5 (Ex—£2) (8.7) x3(0)=3/6, 8 (0)= J5 (8+8). As indicated by Figure 8.13, the director x3 relaxes to a final steady state ( x3 (00) = g2 ) by executing a complex three dimensional motion wherein the “cross-flow” component of £3 (i.e., x3 - ey) first increases to a maximum and then decreases to a steady state (i.e., x3 (00) - 9y = 0 ) by a damped oscillation through the vorticity/cross-flow plane. The “vorticity” component of 53 (i.e., x3 ex) shows a monotonic decrease from its initial condition (33 (0)-§x = «[2- ) to its final steady state (x3 (ao)-gx < 0). This relaxation process, which has been termed director kayaking, produces a microstructure with invariants 11b (00) = 0.546, and Int, (00) = 0.165 (see p537, in Larson, 1999). Steady state component of < pp > (oo) are 91 8888.88 .- .8 88 n 8:88 888 u 8.8 88.8 u 88 m8288.88.08.88 N8 m 885 80.8 88888888an 88885 88888888888888 2 .w 08:85 8 00.0 0.8.0 00.0 0N0 9.0 00.0 8 8 8 8 8 . ON.OI 30.0 n 8888800 8833800 8880888980888 088888 00.0 8.8 .. 00.0 oooooolo-loooo cu. u u n u u n .I. u n u u u n u u 00.? II (D .-/ 8:7 92 < pp > (oo) = 0.031 exex — 0.001 exey — 0.001exez — 0.001eyex + 0.032 eyey - 0.004 eyez . (8.8) — 0.001ezex - 0.004 62 ey — 0.937 ezez The eigenvalues and eigenvectors associated with < pp > (00) are 9.1 (8) = 0.031, x1 (00) = 0.998 g, + 0.065 9y + 0.052 g, 12(8) = 0.032, 252(00) = 0.065 9,, - 0.098 _e_y — 0.004 g2 (8.9) 9.380) = 0.937, 103(00): —0.0019x — 0.004 gy + 0.997 g, 8.4 Relaxation of Planar Anisotropic States for 0 _<. L/d < 00 Figure 8.14 shows relaxation of a planar anisotropic state with (< _p_p > (0) =§gy§y +§gz§z) for A. = 0.5, U = 27, Pe = 95, and Fm = 1. The director and its eigenvalue are periodic. One eigenvalue relaxes to a steady state. The other two eigenvalues fluctuate and the director tumbles in the deformation plane with a period it i 0.08/6D°R. Figure 8.15 shows director tumbling for A = 0, U = 27, FTD = 1, and Pe = 95. The director tumbles 180' with period of it i 0.07/6Dfi. However, for this case the invariants of the structure tensor and the eigenvalues of < p p > relax to steady state values with a microstructure on the prolate boundary (11b (00) = 0.551, 111-, (00) = 0.167). For this case (A = 0), asymptotic solutions to Eq.(4. 1) split into two contributions: 1) solid body rotation; and, 2) a steady state microstructure that exactly balances the Brownian torque and the torque due to the excluded volume potential. It is noteworthy that Ill) (00) 93 .8 8888+ .88.me 80 A mm v N 30 H 00 Km H D 30 u .8 88 H 88.8.88 880008-0080 2888888 08888888 288 88 88828888888888 88m 88888080 888888 8888808883. 83885 888.8 80830888888082 3 .0 088.8888 8H: 0N0 0N0 m _..0 0 F .0 00.0 00.0 00.0- 8 00 0 w 0.0 who v.0 Nd < o 0 /<< .8 88.88.0388 -0_..0 . 9 00.0PM .0 . . 8 I ON C 8 0 . 88 . 0 "858 I end . .8. 88 - 888 .888 . .8. . 8888.88 7>8§ . 88 . . . 0.0 0.0 v.0 .11 0 r omd - < 1 . .888 N.8 _.8 . 8.88 88828 8.8828 8888.88 888.88 _ . - 00.0 . . - -8 , . v.88 \4 \ 88.88.88.888 / . _ . 88.88 .. on 0 888838883 888800880 8.‘ . 0.0 / 8 8 8888.88 .8838 H 8.888 .8888 .882 94 A 8888+ 88.8.0.“ 800 A mm. v m 800 u 00 Km n D 80 u .8 88 n Em xflEE-Ommv 88888888 088883 288 888 88828888008888 BSm 88888880 888888 8888808883. 88885 888.8 80830858888082 2.0 08:08.8 .888 08.0 08.0 08.0 08.0 00.0 00.0 00.0- . . . 8.88 8888.88 8828 88.88 83.88 883.88 .000 $8.88" 88 . 188.88 88888883 8888888 88883 888 88888888 .. 0N0 . 80.0 . 00.0 .88 - 088.0 - 00.0 «.0 .. 00.0 . v.0 . 0 . .. 05.0 . 0.0 um. mm -I. 308880 8.8 . 0.0 06:: P 8.8888 0 .8 8 A 95 and IIIb (00) are the same as shown in Table 7.1, for U = 27, Pe = 95, and A. = 0 and for as well as U = 27, Pe = O, A =1. For A = O, U = O, and Fm =1, Eq.(4.1) reduces to 6< > RE +Pe[lT-+-1]= I-< >. 8.10 at X = ‘32 ( ) .1. 3 For Pe = 95, Figure 8.16 shows the relaxation of the invariants of the structure tensor llc" ( < pp > — g l) to the isotropic state. The director tumbles with a period tt i 0.07/ 6D§ , which is the same dynamic response as k = 0, U = 27, and Fe = 95 (see Figure 8.15). Figures 8.15 and 8.16 support the idea that the asymptotic solutions to Eq.(4.1) for A = 0 splits into a solid body rotation and a steady state axisymmetric prolate microstructure on the F-boundary of Figure 1.1. For U = 27, Fe = 10, and Fm = 1, Table 8.2 shows that the dimensionless tumbling period decreases as 7L decreases. Larson noted that the tumbling period tt ~ L/d. Table 8.2 shows the relation of tt ~ (L/d)2/3 for O < k < 0.91, but as 7t —> 1, tt —-> 00. 8.5 The Effect of Tube Dilation on the Relaxation of Planar Anisotropic States Figure 8.17 shows the effect of tube dilation on the microstructure for A = 0.987. The phase transition between steady alignment and tumbling for fixed Pe is the same as Figure 8.9 (i.e., U = 26), but the tumbling and wagging regions are extended to larger values of Pe. This occurs because the diffusive flux becomes larger inasmuch as FTB‘ —+ 00 as 11}, -—> 3/2. 96 A NwNm+ 00.03.” A00 A mm v N ”mm H on .0 H D 00 .I.. & 00 H Em 000008-0m5 200E 9,803 200 000 00025032 80% >035 0:0 $00585. 08805 00.0 80300580002 00 .w 8009.10 D000 mNd 0N6 0.20 0rd mod cod mod- D P h b - oo-o \. 0 00000080 I 0.20 . 0N0 r 00.0 900 . 0v.0 . 00.0 . 00.0 QumTNT: . .000 . to . 0.0 97 Table 8.2 The Effect of K on the Tumbling Period (FSQ-model; Fm = 1; U = 27; Fe = 10) A. tt 5 6D§Et L 2 (—) -1 1 °° —1 s 9. a d 5 +1 0.987 14.883 (EV +1 0.985 8.232 1-2 ‘ _____________________________ 0.980 4.983 0.8 - 0.970 3.312 0.4 . 0.950 2.286 0.930 1.858 1 o - 0.920 1.720 _o_4 . 0.910 1.610 0.900 1.520 '0-8 ' ___________________________________ 0.800 1.068 -12 , r , i 0.500 0.913 0.01 0 1 1 10 100 0.600 0.790 W 0.400 0.687 0.200 0.641 0.000 0.628 100 - O 10 ‘ o 0 1c : I 1 " . 9 _________ 9 O ........... £22221"--- ---- 2/3 0.1 . 0.1 1 98 .0 9060 + 000m n 80 A 0.0 v N $3.00 1 .0 00 n “.0000 0 06068.30 000000005 8005 200 000 0000025 0000.01.00 03080.5 25. 5.0 80000.0 on 00000 _. 0000 _. 000 _. 00 _. or _. . q . 0 1 0 1 m .. 0 _. $8.0m . mum n0001M... 0w H 000% 08800000“ 00080.6. .. m _. . 0N D .. mN .. om \ 00500830380000 . mm 9000me 0008.000 99 8.6 Discussion The microstructure of the orientation dyad, which is induced by a homogeneous shear, has four independent variables: U; Pe; l; and, the initial state of the orientation dyad. Figures 8.2 and 8.15 show that the nematic potential U influences the axisymmetric orientation state (the prolate state) regardless of director tumbling. For Pe > 0 and A = 1, the microstructure is anisotropic and inside the invariant diagram. For large values of U and Pe, the orientation states are near the nematic state. The tumbling parameter k influences the periodic orientation state and its period. For k = 1, the orientation states are steady states. For A. < 1, periodic orientation states occur. Table 8.2 shows that reducing 7» reduces the tumbling period. Eq.(4.1) has four physical features that determine the orientation state: vorticity, Brownian motion, nematic, and strain contributions. The vorticity contribution tends to have periodic rotation, but the strain contribution hinders its motion. For U < 26, the nematic contribution is not large enough to reduce the strain contribution (steady alignment). For U > 26 with J. < 1, the director tumbling occurs at low value of Pe because the vorticity contribution becomes larger than the strain rate term. However, for large values of Pe, the strain rate contribution regains its strength (steady alignment) (see Figure 8.9). If the strain rate contribution is reduced by 1 (Figure 8.14 and 8.15), director tumbling occurs for even larger values of Pe. The HLl-closure approximation predicts tumbling phenomenon with k = 1. According to Chaubal (1995), the decoupling approximation used by Doi and others can also predict tumbling phenomenon (Chaubal and Leal, 1997, 1998; Chaubal et al. 1995) provided the flow field is modified. The results developed in this chapter shows that 100 Doi’s theory predicts director tumbling for homogenous shear provided 71. < 1 (see Figure 11.3). The log-rolling and kayaking phenomena have been predicted by other closure models (Chaubal et al. 1995; Faraoni et al. 1999; Larson and Ottinger 1991). FSQ- closure also predicts director tumbling, log-rolling, and kayaking. As demonstrated in this chapter, the realizable F SQ-closure also predicts log-rolling, and kayaking by the director as well as the existence of multiple steady states for Pe > 0. 101 CHAPTER 9 VISCOSITY AND NORMAL STRESS DIFFERENCES 9.1 Introduction For homogenous shear flow, the deviatoric stress has three nontrivial normal components and two non tr1v1al shear components: ;zgxgx , ;zeygy, ;zgzgz , __rzgygz = izgz 52),. In this chapter, the effect of the tumbling parameter it, the excluded volume coefficient U, and the Péclet number Pe on the viscosity, the first normal stress difference, and the second normal stress difference will be developed by using Doi’s theory for the stress (see Eqs.(4.8)) and the FSQ-closure for the orientation tetrad (see Eqs.(5.8) and (6.12)). The dimensional rheological properties are defined as follows: viscosity 1:9 9 05‘ iy. (90 Y zgz — Eygy) , (9.2) 1:12 5::(§y§y_§z§z)' (93) The objective is to access the impact of the realizable FSQ-closure on the rheological properties of rigid rod suspensions. 102 9.2 Rheological Properties: L/d = 00 Shear Viscosity Figure 9.1 shows the effect of the excluded volume coefficient U and the Péclet number, Pe = y/(6DR ) , on the steady state shear viscosity due to the viscous and elastic components of Doi’s stress for the realizable FSQ-closure (see Eq.(4.8)): _ A .. 6DR Tl Tls-(Tl Tls) 3ckBT =F(U,Pe) , osU 600 DR ). For large Pe, Figure 9.1 shows that the shear viscosity becomes independent of U for large Pe: Pe —900 -3/5 n—nsEF(U,Pe)——)O.18Pe (9.5) PBLG in m-cresol solution and PBZT in methane sulfonic acid solution, which are lyotropic LCPs, also show shear thinning phenomena at large strain rates with n 5 —1/3 and n .2. —3/4, resp. (see, p.286 and p.510 in Larson, 1999). For low values of Pe, the shear viscosity becomes independent of Fe (i.e., Newtonian plateau) but still depends on U. Figure 9.1 shows that lim F(U,Pe)=o.01+——O‘—25——. 96—30 1+1.17U (9.6) For dilute suspensions, DR is independent of concentration (i.e., DR —> D3) and U << 1. Therefore, Eq.(9.4) and Eq.(9.6) predict that Afio at c, which is consistent with experiments and other theoretical predictions for rigid rod suspensions and LCPs (see p.281 in Larson, 1999). 103 .00 u .0 m0 .1. 80 m026200.300 8 n 03 000 38085 00000 200 000 00 00000 D .00 Beam 2E. 0.0 8300.0 on 0000 000 00 0 .1 u " Sod 000.0 mm o on 008.0. 0.00 n wmn 0.8.00 2 a 4 ./,m m m 820 o o w IIIIII m 0 0:4 D u mam... _ m\m1_.|/// coed q n. n u m 0000 o o 9:0 /,/ooo 4 4 < 00000 0 [III 0 0 q III: 0 QQQQ Q Q o .................. 0 . 0100 00 0: 0+0 1 o 1|l|.||1 1:06 U ADV 0:4 m Amfl CV 8: 0 0 mg 6 .r to As the concentration increases, U becomes large and DR at: c“2 (see p.287 and p. 520 in Larson, 1999). Under these conditions, Eq.(9.4) indicates that Afio at c3 / U for "concentrated" suspensions. For "high" concentrations of PBLG in m-cresol (i.e., > 0.5 wt% ), Mead and Larson (see p. 290 in Larson, 1999) observed that the zero-shear rate viscosity increases with concentration as Afio 0: c3. Thus, Eq.(9.6) and the foregoing experimental observation imply that U becomes independent of concentration for semi- dilute and concentrated suspensions. For dilute suspensions, U at c (see p.66 in Larson, 1999) Figure 9.2 shows how the viscous and elastic contributions to the shear viscosity (i.e., Afi :— fi —fi s: fiv + fig) depend on Pe for U = 0 and U = 27. For the stress model used herein (see Eqs.(4.6) and (4.8) ), fiv and ma are defined as follows: fi 5‘7": " . (9-7) V Y Ilg and fi _fi _3ckBT[—U(-:)] (98) E=—.‘- '* - - 1 |l§|| In the above equations, cQR represents the viscous drag coefficient between the rod and the suspending fluid per unit volume of mixture and has units of viscosity (force- time/area). The parameter 3ckBT has units of energy per unit volume of mixture, and c CR is related to 3 ckBT and the rotary diffusion coefficient DR as follows 3ckBT c = . 9.9 CR 615, ( > 105 .00 n .0 m0 1 9.0 00688-30 8 n 0600 060 000885 080m 20 .00 3000000000000 232m 0000 3585 20 so 00 0000 D .00 80.0m— ofi. Nd 00:00.00 um on 000? 00.. 0.. 0. 000_. 00? 0.. F 5000.0 . 0000.0 106 . 50.0 Sit-Lt . 0.0.0 a=+>cumcucmc<\ Bachelor developed Eq.(9.9) for dilute suspensions of slender rods (see p. 284 in Larson, 1999). This parameter is used to scale the viscosity coefficients: TlV =fiV/cCR and n5 =fiE/CCR. For Pe > 50, Figure 9.2a and 9.2b show that the shear viscosity is primarily due to the viscous stress; however, for Fe << 5, the elastic stress becomes more important although the viscous contribution remains significant (Tlv is approximately 25% ofAn at Pe = 0.1). For U = 0, the elastic stress is due to Brownian motion. For U > 0, the excluded volume phenomenon mitigates rotary Brownian motion. With U = 27, Figure 9.3 shows that the elastic stress does not contribute significantly to the viscosity because rotary Brownian motion is balanced by counter diffusion due to the excluded volume potential. For low values of Pe, both the viscous and the elastic components of the viscosity are approximately independent of the strain rate (i.e., Newtonian plateau, NP); therefore, Eqs.(9.7) and (9.8) imply that in this region fi ..=—:,':P ”(llélll W [< pypz > —U (< pyp> - < ppz> — < pypzpp>z< 22>):le oc ll§||~ (9.11) For U = 0 and large Pe, Eq.(9.7), Eq.(9.8), and Figure 9.2a imply that lim at: ”gm—3’5 (9.12) Pe—wo - l/5 lim < pypz > at H S ||- (9.13) Pe—No - For U = 27 and large Pe, Eq.(9.7) and Figure 9.2b imply that lim °c|I§||’3’5, (9.14) Pe—mo 107 .2 H & 00 n Em 000008-0m00 8 u 03 000 30885 20 .00 0002000800 0035 00200 20 00. 3000050000000 20020.0, 0003830 0000 000000305 20 000 on 0000 D .00 08.0m— ofi. 0.0 oSwE oh 000? 00? or w 8 8 D 0 0020050000000 20000000, 0002800 / 0000005000008 0000000305 calm: 1? . 0010030 ”.0: L .moAfi .mPAV -NAV on 000V 00? 0? A0 u 30 0000005000008 0823 0002800 ......... m: n 0000055008 03000305 0103 m: W0- .000? rmoau -NAV 108 which is the same as Eq.(9.12). It also follows from Figure 9.2b and Eq.(9.5) that . —/ P119133 [—U(-:)]ocl|§|| 15. (9.15) 00 The effect of U and Pe on the elastic contribution of i : gzg is shown in Figure Y 9.4. If U E O for dilute and semi-dilute solution, then Doi’s shear stress for the elastic contribution can be approximated by the Brownian motion contribution of stress (see p.308 and p.338, Doi and Edward, 1986; and, Smyth et al., 1995). Thus, with U = 0, Eq.(9.8) implies that “352 |U=o = 3nk13T < Psz > (9-16) Figure 9.4 shows that the general trend of total elastic contribution, predicted by Doi’s shear stress depends on U and Pe. Smyth et al. (1995) used a birefringence method to estimate the elastic contribution to the shear stress for semi-dilute solutions of xanthan gum (L = 1440 nm and L/d 660) in fructose solvent (r13 = 0.483 Pa-s). They used Eq.(9.16) to relate the stress to the microstructure. For limited range of shear rates, their 1/3 experimental data indicated that $32 at (y) . Figure 9.4 shows that these experiments are consistent with the Doi theory with a realizable FSQ-closure provided U = 5 and 0.2 < Pe < 2 (i.e., 1.2 Dfi <7 < 12 D‘fi ). The strain rates in the Smyth/Mackay experiments covered the range ls-1 < ‘y < 20s_1. Therefore a combination of Doi’s theory and the experimental observations gives the following estimate D‘fi s 1.3 s_1. For U = 0, the elastic stress makes a significant contribution to the total stress for Pe < 100. 109 C H 80 00 n .0 000008-0m40 $000 03000 20 00 3000530000 003E 0.0 00:00.0 om 000—. 00—. 0.. _. v.0 5.0 n n n p u . p f - Foo.o .. ... . . \\\ ... .5... ....C.. \\\ L— - o \ n 1 \\\ 1 nur o o m 1 1 - coco... . . .0 N0 ....11 800000 o o o. o 0. Q0 .... m _|I.111111- o 2 LT .... . 0: 1.1 _ll..- one o // 0 1ml . _. \.1. duo. .. O O O O on... o o N? + M009 11.0" 3809 D >8— N.» 09 . l a N.» NvoAVN00E~mID0om 200000P .0_. 110 Normal Stress Differences Figure 9.5 shows that N is positive for 10’1 S Pe S 103 and 0 S U S 35. For large 3/5 values of Pe, Nl oc Pe . For small values of Pe, Nl oc Pe. These results are consistent with the experimental measurements of Zirnsak et al. (see p. 295 in Larson, 1999) for glass fibers suspended in Newtonian fluids and with experimental measurements of Kim and Han (see p. 513 in Larson, 1999) for thermotropic polyesters, OQO (phenylsulfonyl) 10. By contrast, at low Péclet numbers, Nl cc Fe2 for isotropic viscoelastic fluids (see p. 450 in Larson, 1999). For a fixed value of Pe, Figure 9.5 also shows that the first normal stress difference decreases as the excluded volume coefficient U increases. This occurs because counter diffusion due to the excluded volume effect balances rotary Brownian motion for sufficiently large values of U (see Table 9.1). For U —> 00 at low Pe, N1 is primarily determined by the viscous stress. This prediction is qualitatively supported by the thermotropic polyester experiments mentioned of Kim and Han (1993) above inasmuch as N1 increases with an increase in molecular weight. In an earlier study of PBLG solutions (lyotropic LCP), Robinson (1965) observed that the critical concentration corresponding to a transition from an isotropic to a nematic state decreases with an increase in molecular weight. Therefore, if U decreases as the molecular weight increases, it follows that Figure 9.5 is qualitatively consistent with trends observed for both thermotropic and lyotropic LCPs. 111 6.0 000 _. 00 .. .0 1 .0 m0 1 e0 m060.60.00.30 8 n 05 00.. 8:80.05 3000 .0800 Z 03. m 20 000 mm 0000 D .00 Beam 2.0. 0.0 00:00... 0.. N. 0.0 N000 000.0 000.0 mm mm m. G 1 6.00 _z v.0 0.. 00? 112 Table 9.1 Viscous and Elastic Contributions to the Shear Viscosity and the First and Second Normal Stress Differences at Selected Values of U and Fe for L/d = no Elastic Property Pe U Total Viscous . Excluded Total Browman Vl oume 0 0.0761 0.0374 0.0387 0.0387 0 5 15 0.0295 0.0272 0.0023 0.0180 —0.0157 n-ns 27 0.0122 0.0109 0.0013 0.0208 —0.0195 0 0.0019 0.0019 0.0000+ 0.0000+ 0 1,000 15 0.0028 0.0028 0.0000+ 0.0000+ 0.0000+ 27 0.0017 0.0017 0.0000+ o.0000+ 0.0000+ 0 0.8928 0.4011 0.4972 0.4972 0 5 15 0.5027 0.3965 0.1062 0.8337 —0.7274 27 0.3619 0.3160 0.0459 0.9040 —0.8581 N1 0 28.487 27.515 0.9717 0.9717 0 1,000 15 27.675 26.840 0.8351 0.9756 —0.1405 27 27.016 26.267 0.7488 0.9780 —0.2292 0 —0.0420 0.0420 —0.0840 —0.0840 0 5 15 —0.0159 0.0159 -0.0318 0.0037 —0.0355 27 —0.0071 0.0071 —0.0142 0.0037 —0.0179 N2 0 —0.0055 0.0055 —0.0110 -0.0110 0 1,000 15 —0.0294 0.0294 —0.0588 —0.0076 —0.0512 27 —0.0394 0.0394 —0.0788 —0.0059 —0.0729 113 Unlike N1, Figure 9.6 shows thath < 0 for 10‘1 3 Fe :103 and 0 s U s 35. As expected, the magnitude of the second normal stress difference is significantly smaller than the first normal stress difference (|N2| .<_|N1|/200). The calculations support the conclusion that INzl/INIIAO for Pe—>O and Pe—)oo . Note that for PeleO , |N2| SlNll/IOO for U = 35 and decreases to lNzlslNll/ZSO for U = 0. For a fixed value of U, the second normal stress difference increases in magnitude as Pe increases, but reaches a maximum for an intermediate value of Pe. For large values of U (i.e., U > 15), max |N2| occurs at Pe 5 200. For Pe E 10 and U > 15, |N2| decreases as U ll increases; however, for Pe _ 500, |N2| increases as U increases. Coincidently, for 10 3 Fe 5 500 and U > 15, the shear viscosity decreases significantly as Pe increases whereas the first and second normal stresses increase as Pe increases. Figure 9.7 shows the contribution of normal stress differences with FTD = 1, 7s = 0.987, and U = 27. For Pe > 10, the viscous contribution to N1 becomes dominant, though the elastic contribution is increasing (see Table 9.1). On the other hand, the elastic contribution to N2 is more important than the viscous contribution. Table 9.1 shows that the elastic contribution of N2 is nearly twice as larger as the viscous contribution 114 N2 .0 1 K x 1 as. 2688-840 8 u 05 80 8080005 3000 .488 Z 0:800 20 so 00 0:4 D mo Beam 2:. 0.0 8:03 on 000v 00.. 0F .. u u n 00.0- 1.. 00.0- o a n 444 00 o n. m 4 4 4 4 0000 _u 4 DD 4 o .0: 0..... G O G \0‘0‘01? .0119 .u 3 on . 4 o D x / << < .u be \ dd 0 n.\ \0. xx .1 00.0- 0 O n. x \0 O 4 xx .0 <4 / O Q 1% . 1 \hobs 4 8 o o\\\ 4 n 4 .o\ D .u ..0 0- 0.0019019 20.0. mm o 0000 n. n 4 23. a u o o 83- 2 4 H . 000.0 0 ...¢ 1 .......... oo o A8 T 0% NZ D 115 AR 1 D x 1 a m _ 1 e... £688-30 $285005 mmobm 0280 Z 0008034 08.... 20 no mmobm mo 20.50.0400 2.01 5.0 850.”. on. cm 000? 00F or F 000w 00.. or P T . . v.0- . . . m- -- 00.0- oE=.o> 025.88 .3242 \ .. 00.0- o. 0 5.4305 .1 m 0 \o ’0. .. 40.0- .2202 03000?» ENEBEm ./ 11 o_‘ \ -. No.0- NZ 2 I: .. 2 1111111111111111 II C .- Nod .. ow .. 3o -. mm 2583 .. mod L- on 116 9.3 Rheological Properties: L/d s 12 Shear Viscosity Figure 9.8 shows the effect of the tumbling parameter )1 on n — n5 for U = 0. This rigid rod contribution to the viscosity for k = 0.987 (L/d 5 12) is the same as X = l for Pe <100. However, for Pe > 100, a Newtonian plateau occurs for 71 = 0.987 and the shear thinning region continues for it = 1. Because Pe is high (see Eq. (4.1)), the diffusive flux becomes negligible, but it still makes a relatively significant contribution because the convective flux (hydrodynamic interaction part) is mitigated by 71. For example, Table 8.1 shows that the invariants of the structure tensor are nearly constant with 71 = 0.987 at high value of Pe. Hypothetically, if the microstructure could be forced to align more for k = 0.987, another shear thinning state may appear for larger Pe (Region 111) (see Larson, 1999 p. 509 — 511; Walker and Wagner, 1994; Walker et al., 1995). In the tumbling region (high U and low Pe), the shear viscosity has a periodic behavior because the microstructure does (see CHAPTER 8). Figure 9.9 shows an example of the instantaneous viscosity for director tumbling (I. = 0.987, U = 27, Pe = 10, and Fm = 1). The dimensionless frequency of the dynamic viscosity humbling is 0.21. Note that it has two local maximums and one local minimum per period. The shear viscosity shows instantaneous thickening when the director is not aligned with the flow direction. A period of shear thinning follows when the director rotates towards the cross flow direction (i.e., u x E ), but the flow and rigid rod resistance is not significant. The resistance to director rotation becomes significant when the director passes through the 117 .0 1 en. 0688-300 0 n D 000 @6085 00000 05 :0 0808340 05.0800. 00 Hootm 22. 0.0 050.0 000 —. 000.0 M 2a M on t ..00.0 s ..0.0 80—0 s ...0 118 .0: 1 60 .R 1 0 53.0 1 a 0 1 em 0688-30 00.58:... 000025 000 36083 82.0 msoocficfimfi 2.... 0.0 050.0 0 0 .04 0 m4 0 .04 00.0 00.0 a a 19. o... 2.0 I1_w U S m...0 .2 2:90 80 BE 1 .0 8.0 0N0 119 cross-flow/vorticity plane, which trigger another shear thickening episode occurs (see Figure 8.3). As the director rotates away from the cross flow direction, another shear thinning episode occurs. It is noteworthy that this prediction agrees qualitatively with experimental results reported by Gu and Jarnieson for thermotropic liquid crystals, 8CP at 36.6 °C (see p.465 in Larson 1999; Chaubal and Leal, 1999). The time averaged shear viscosity can be obtained by averaging the instantaneous shear viscosity over several time periods. Figure 9.10 shows the effect of U and Fe on the time averaged shear viscosity for A. = 0.987. As U increases, the Newtonian plateau region is extended to higher values of Pe because the tumbling region is extended for A. = 0.987 (see Figure 8.9). Shear thinning occurs in the wagging region. In the steady alignment region (high values of Pe), the shear viscosity is independent of Pe. In addition, shear viscosity becomes independent of U and Fe for 71 = 0.987 at high Pe because the orientation state becomes nearly constant (see Table 8.1). For some LCP experimental studies, the shear viscosity shows a shear thinning phenomenon at low strain rates (Region I), a Newtonian plateau region at medium strain rates (Region II), and a shear thinning region at high strain rates (Region III) (Walker and Wagner 1994, Walker et a1. 1995, and see Larson 1999 p. 509 — 511). Region I may be due to layers of different orientation states (texture affect) in LCP solutions (Marrucci, 1991), Region II is due to the director tumbling phenomenon, and Region II] is due to the shear aligning phenomenon (see Larson 1999, p. 509 — 511). Figure 9.10 shows that Region II coincides with director tumbling and Region III with the director wagging region (see Figure 8.9). Note, however, that additional Newtonia plateau occurs in the 120 $3.0 1 a x 1 Es. 0688-30 N. m 3.. 00.. 3.88..» 002.0 000000244 08.... 2.. 00 on. .000 D 00 Spam. 2.... 0. .0 0.00.”. on. 000.. 00w 0.. .. u u n .000 4.0.0 mm IQI 20.0 R 1m1 N.0.0 m . Lcl 0.0.0 0 IOI 8:4 D 050.... com. 000500.? 00080 00.000? 05383 M. m u ..0.0 _ 0.6 650E 68 .u S Dom.0+. 3.0- :5 m 8:4 1.0158: 1.20 121 steady alignment region at high Pe’clet numbers (see CHAPTER 11 for further discussion). For U = 27, Figure 9.11 shows that the elastic contribution to the shear viscosity is relatively small for all Péclet number. Indeed, the elastic contribution to the rigid rod suspension stress is almost negligible for Pe > 20, and the viscous contribution becomes independent of Pe (see Table 9.2). This shows that the Newtonian plateau region at high Fe is determined by the viscous contribution of the stress and that the local microstructure is insensitive to further increases in Pe. Normal Stress Differences Figure 9.12 shows the instantaneous first normal stress difference (N 1) for director tumbling. The frequency of tumbling is approximately 6 Dfi rt/15 when U = 27, Fe = 10, A = 0.987, and FTD = 1. When the director tumbles, N1 changes sign. N1 increases when the initial director is off-aligned from the flow direction (t s 44.3), and decreases when the director rotates towards the cross-flow direction (compare with Figure 8.3). N1 increases again when the director completes the rotation from the cross- flow direction to the flow direction. In a cone-and-plate viscometer, the plate is pushed apart when the director rotates away from flow direction because this causes an increase in N1. On the other hand, when the director aligns with the flow direction the plates push back on the fluid. Figure 9.13 shows the effect of U and Pe on the time averaged first normal stress difference for k = 0.987 and FTD = 1. When U is relatively small (steady alignment 122 .8 1 0 ”56.0 1 x 0 1 B... 0688-000 2 m .5 .50 00865 00000 0008344 08.... 0... :0 00000000 0.30:”. 000 000005 00 000.000.0000 0..... ...0 0000.... 0n. 000. oo. o. F . . . 588.0 . 508.0 \ . 38.0 00. m . 80.0 0. >0. .su. IIEIIJI h 5 . 00.0 >0+001$10104 .8 123 Table 9.2 Viscous and Elastic Contributions to the Shear Viscosity and the First and Second Normal Stress Differences at Selected Values of U and Fe for L/d -_'=. 12 Elastic Property Pe U Total Viscous . Excluded Total Browman Volume 0 0.0809 0.0560 0.0249 0.0249 0 5 15 0.0251 0.0238 0.0011 0.0126 -0.0115 11 _ TI 8 27 0.0228 0.0214 0.0014 -0.0004 0.0018 0 0.0099 0.0099 0.0000+ 0.0000+ 0 1,000 15 0.0123 0.0123 0.0000+ 0.0000+ 0.0000+ 27 0.0132 0.0132 0.0000+ 0.0000+ 0.0000+ 0 0.8728 0.3813 0.4915 0.4915 0 5 15 0.3586 0.2803 0.0190 0.8223 -0.8033 27 -0.0128 -0.0111 -0.0017 0.8699 -0.8716 N1 0 5.5053 4.6326 0.8727 0.8727 0 1,000 15 0.2563 1.9266 0.3297 0.8952 -O.5655 27 -0.3244 -0.2789 -0.0455 0.9046 -0.9501 0 -0.0470 0.0399 -0.0869 -0.0869 0 5 15 -0.0140 0.0102 -0.0242 -0.0245 0.0003 27 0.0003 -0.0001 0.0004 0.0196 -0.0192 N2 0 -0.0579 -0.0074 -0.0505 0.0505 0 1,000 15 -0.0537 0.0275 -0.0813 -0.0135 -0.0678 27 0.0094 -0.0056 0.0150 0.0014 0.0136 124 .2: n on .R n D 535 n a x u an ”388-0% @2383. .8885 you 8:98th mmobm 38.8 Z 3: m msoocfiafimfi 25. Ed gamma a 0.9V 09‘ 06¢ 9: 0.3V mév m- a . m- o.m m.v o.v man. 3 . _a. . O ~z . F 9.x ouzwmmoomv . N SPMu .m c . N 2 m 125 .Ead n a m _ n E ”_ouoadme S m 3 E 8:80me mmobm $8.82 BEE wowfio>< 08E. 2: so on 98 D mo 30th BE. 2d 85me om 89 2: S .r n # wimm?» v mEEEB 8.x 2:9..— oomv Eon—:93 €83 :d 2:5 08 v HM 84loam ww.m + mmd- m _2 E: .. m 126 region), N1 is always positive. However, N1 is negative in the tumbling region and remains negative for higher values of Pe. Figure 8.7 shows that the director angle for high U and Pe is negative. The negative angle indicates that the director is pointing downward so that the plate is pushed towards the local microstate. Larson (1999) explains that N1 is positive when the director is tumbling, and becomes negative when it is wagging. He suggests that the negative N1 in both the computational results and experimental results supports the idea of director tumbling for LCPs. Both results also show increase in N1 at high strain rates. Figure 9.13, however, shows that Nl,avg predicted by the Doi stress is independent of Pe at high values of Pe. However, N1,avg still depends on U. N1 remains negative as long as the director tumbles. A dependence of Pe only occurs when 7» = 1 (see Figure 9.1). As mentioned previously, a negative N1, which is independent of Pe, is caused by a constant orientation state. Figure 9.14 shows the instantaneous second normal stress difference (N 2) for director tumbling. The dimensionless frequency of tumbling is 0.21 when U = 27, Pe = 10, and FTD = 1. Analogous to the instantaneous N1 analysis, the sign of N2 can be related to the microstructure of the rigid rod suspension. N2 decreases to negative values when the director is close to the cross flow direction. However, when the director passes the cross flow direction (t E 44.3), it changes to positive values. Then, N2 slowly decreases to zero as the director approaches the flow direction (see Figure 8.3). 127 .82 n on .R u 3 ”$3 n a x u an €88-96 S m 3 as, mEEEPH 8885 Sm 8:80me mmobm 38.82 988m mzoocfiafimfi 2C. 3 .m 833m H 99. one mi 34 m- r .omZ .N m DEMu AmwoSmEoomv .N L m 128 Figure 9.15 shows the effect of U and Fe on the time averaged N2. When U is relatively small (steady alignment region), N2 is always negative. For U > 27 (where the director tumbles), N2 is always positive. At the high value of Pe, N2 becomes independent of Pe, but depends on U. As previously noted by Beak and Magda (1993), Beak et al. (1993), and Larson (see p.534, 1999), the results summarized by Figure 9.15 show that 1) N1 and N2 have opposite signs; 2) N2 has a local minimum where N1 has a local maximum; and, 3) [NZ] 5 |N1| /20 (see Table 9.2). 9.4 Rheological Properties: 0 S L/d < 00 Figure 9.16 shows the effect of the tumbling parameter A on the shear viscosity with Fm = 1. In this research, the tumbling parameter is related to the aspect ratio of a rigid rod by Eq. (2.3), which shows that 7t —-> 1 as L/d —-> co; and, A —> —1 as L/d —+ 0. If L/d = 1, then 2. = 0. The suspension viscosity 1]an — 113 2 A11an has a maximum at k = 0.100, which corresponds to L/d = 1.106. Note that the shear viscosity is not symmetric about A. = 0. As I k | —> 0, the frequency of tumbling increases, which causes the viscosity to increase (see Table 8.2). Figure 9.17 shows the effect of A on the first normal stress difference with Fm = 1. There are two local positive maxima: one at 7» = 0.750 and N1 = 0.1084; and, another at A = -0.965 and N1 = —0.0168. A local minimum is located at A = —0.600 and 129 52 u a x u an 3808.30 2 m as é 805805 mmubm 38.82 vacuum wowmco>< 083. 20 no mm 05 D we Hootm 2F 2 .0 830E om coo? cow or F 33.9; . 81$ ll.| +m0 0 m «2 a: S o- . o 0 ll '4 8.0 830$— 30 E983? >083 050me 2 .m 03:. 8“ M©Od mm IOI 000.0 mm Iml 03.0. 2 Lcl 30.0. 0 Iol 3383 l I 0W0- 00.0- 00.0 m>a.mz 00.0 3.0 3.0 130 .2: n on .R u D x u an €88-30 b.6085 Seam cowfio>< 083. 05 no 6068980 mam—085. mo Hoobm Bi. 3.0 830E & 0. r 0.0 0.0 #0 N0 0.0 N0. #0. 0.0. 0.0. 0. T n b D P D b b I b b - coco .... mod .. v0.0 131 ‘v xx x. .c ..s D 2 Av s. x . 00.0 311 _ 3A1?“ as .o H 4823.50 .. 0_..0 . Nfio . 3.0 .2: n on: .R u D x u E: €88de 855:5 305% 38.8 Z Hm: m 3089?. 08E. 05 so :Boafimm 05383. 00 86th BE. 5 _ .0 33020 < 0. _. 0.0 0.0 V0 N0 0.0 N0- #0. 0.0- 0.0- 0. T . P - . - p . n . p ”0.0.. . 00.0- ll \\ ;I A- It 0 )4 a II . 3.3-) c r ‘ .N0.0- ........................................... .86 .86 $32 .30 .000 \\ ,. KL 3250 8 Mis: /, . o I OF.O \ .. 00.0 . Nfio 132 N1 = —0.070. N1 is negative for both positive and negative values of 2.. However, N1 is positive for A = :1 (A = 1, N1 = 0.699; and, A = —1, N1 = 0.531). According to Larson (1999 p. 294), N1 oc (L/d)21n(L/d) for rigid rods (2. > 0). This relationship can be approximated by power law relationship, N1 oc (L/d)7'4. Figure 9.17 shows qualitative agreement with this theory until A s 0.7. N1 decreases with the tumbling parameter as l —-> 1. Figure 9.18 shows the effect of 2. on the second normal stress difference with FTD = 1. Notice that there are two local minima: X = 0.700, N2 = —0.025; and, X = —0.920, N2 = —0.007. A local maximum occurs at k = —0.500 (N2 = 0.012). N2 increases to positive values as X —* 21:1; however, N2 is negative for X = :tl (A = 1, N2 = —0.0127; and, A = —1, N2 = —0.5688). The sign of N2 is always opposite NLbut the magnitude varies with A. 9.5 Rheological Properties: Effect of Tube Dilation Shear Viscosity In this section, the effect of tube dilation on the rheological properties will be examined. This phenomenon directly impacts the rotary diffusion coefficient in Eqs.(4.1) and (4.6). The tube dilation coefficient FTD depends on the local microstructure through . —2 . the second invariant of the structure tensor: F196” = (I —%Ilb) . Table 9.3 defines how 133 .2: n on .R n D m: n em €88-30 855:5 $050 38.8 Z vacuum 0008952 08C. 20 no 83883 022083. mo Hoobm 23. w 0 .0 ear: .2 0. _. 0.0 0.0 V0 «.0 0.0 N0- #0. 0.0- 0.0- 0. T 00.0- .N0.0.. . 5.0- ........................................... ., .00.0 .. .. . Ed 9&2 u No.0 00.0 00.0 134 Table 9.3 Legend for Figures 9.19 — 9.23 . -2 FPB‘ = [l—g—Hb) ,Eq.(2.8) legend Eq.(4.1) Eq.(4.6) _a_ No No -A— Yes N0 + Yes Yes 135 the FTD-factor is applied in the results presented hereinafter. If no tube dilation is considered, then this is designated as “No” in the table. If tube dilation is included, then the affected equations are identified as “Yes.” Figure 9.19 shows the effect of tube dilation on the time averaged shear viscosity for 2. = 0.987 and U = 27. With tube dilation, the tumbling region is extended to larger values of Pe (cf. Figures 8.9 and 8.17). Tube dilation enhances the diffusive flux in the moment equation so the director has more freedom to rotate. Because of director tumbling, the Newtonian plateau is also extended. When the tube dilation is included in Eq.(4.1) and Eq.(4.6), shear thickening occurs near the tumbling and wagging transition region. Larson notes that shear thickening occurs because the increase in interparticle spacing makes it harder for the solution to deform (see p.273, Larson, 1999). In Figure 9.19, the shear thickening phenomenon may be explained as an interparticle spacing effect. However, Larson also mentions that there is no known direct relationship between tube dilation and shear thickening. Figure 9.20 shows the contribution of stress components on the time averaged shear viscosity when Fm = E??? in Eq. (4.1) and Eq.(4.6). Shear viscosity is nearly identical with the viscous contribution of the stress, which is consistent with 3. = l and A = 0.987 study without tube dilation. Normal Stress Differences Figure 9.21 shows the effect of tube dilation on the time averaged of the first normal stress difference with k = 0.987 and U = 27. Note that N] is independent of Pe 136 2803 5 .3 22$ 8... E u D 2&3 u 2 2085-30 36085 Scam 000652;»: 08:. 20 no :28an 00:01 mo 30th 22. 0:0 8:03 cm 0000 _. 000 _. 00 _. 0 w _. _.000.0 4 i . 5.0 ...u a 50.0 r _ a... E.» 2&3 Be Began 0036 050925 MEEEB L 0 _. .0 _.0.0 . 137 28 u a K88 n 2 28.0 88 2.0 sum 5 “m. : 888-30 008%; Seam wowfio>< 082. 05 co mmobm mo 28:25.5qu 2E. 0N0 850E on 82: 82 2: 2 P . . . . 88.8 >2 m: \ . 38.0 / . _.00.0 su _ Emu \ >:+mrnmclrn:< . _.0.0 138 6505 80 0.0 2an com mm H D m$0.0 n & 5255-00”: 8:80.05 8050 3582 85.0 00080>< 250- 20 co 25:25 33. mo 800mm 2:- 30 830E on 0000.. 000—. 00.. 0F _. 00.0—3- . q a ON.—-I E.» 8:5 83 50:50:“: 003% 050095 05383 00.3.- - . 00.7 00.9- - . 00.0- 00.07 - 8.8- . . 8.0- m2:2 00.0- - .. 0v.0- 00...r - - 0N0- 00.N- r 00.0 - P 00.0 139 for Pe >> 2,000. In addition, N1 with and without tube dilation have similar trends in the tumbling, wagging, and steady alignment regions. The inclusion of tube dilation in Eq.(4.1) causes the tumbling at high value of Pe. This influences prediction of the minimum of N1 and its magnitude of N1. If tube dilation is included in Eq.(4.1) and Eq.(4.6), N] has a local minimum in the wagging region. Tube dilation affects the magnitude of N1 at the local minimum significantly. Figure 9.22 shows the effect of tube dilation on the time averaged second normal stress difference with k = 0.987 and U = 27. Though N2 values are independent of large Pe value for all tube dilation cases (see Table 9.3), they have different trends related to tumbling, wagging, and steady alignment. With the tube dilation effect, signs of N2 values are not always opposite to N1. When BB? is only applied to Eq.(4.1), N2 is negative for low value of Pe. In the director wagging region, N2 changes from negative to positive. In the steady alignment region, N2 remains positive and becomes independent of Pe. When FPSi is applied to both Eq.(4.1) and Eq.(4.6), N2 is positive for low value of Pe and decreases to negative in the wagging region. In the steady alignment region, N2 remains negative and becomes independent of Pe. None of these results have been observed experimentally. Figure 9.23 shows the contribution of stress on the time averaged normal stress differences with EB? only in Eq.(4.1), A = 0.987, and U = 27. The viscous contribution 140 g 0.502 E Q 23. 03 R u D ”53 n K 0088-30 85505 $000 38.8 Z 0:080 0008030 082. 05 co noun—5 003.0 00 Beam 23. 00.0 850E um 82: 2: 2 F 00.0- n n 00.0- 8.0. .. .. mod- 0N0- .2 .. _.0.0- ‘l 9.0. 1.. 1.. 00.0 03.02 070- nu .. a - .. 1 uuuu n u 1.. rod 00.0.. 1.. .. No.0 00.0 .u n. mod E0 2:5 30 mod % Ecru—0% >086 050095 05553 I. V0.0 141 0000 _. pl G 02 u 0 ”53 u .0 m 0% 0 £85.30 80:80me 00800. 020.8 Z 0o080>< 08E. 20 :0 00000 .00 30035000 05. 00.0. 050E 000 _. b q 0:080» on 00 _. 0r b J “NH" U / 0823 000388 038$on l J L 00.0- - No.0- - 3.0- - 5.0 - No.0 - 00.0 - 00.0 - 00.0 - 00.0 on 0000? 000.. 00? 0? P u u 0 u m...- 0829» 000288 1: PI £583 .- 0.0- - .......... _ Z - 0 Essen - 0.0 / LI F 142 to N1 determines the total N1, while the sum of Brownian and the excluded volume contribution to the elastic component of N1 are very small. 9.6 Discussion Predictions of rheological properties, N1, N2, and A1] are coupled with the microstructure of the orientation states. These orientation states can be categorized by a tumbling parameter A. For )- = 1, all solutions are steady state. A11 has a Newtonian plateau and a shear thinning region, N1 is always positive, and N2 is always negative (see Figures 9.1, 9.5, and 9.6). For )- 3 0.987, a steady alignment region at small values of U (U < 27) is predicted. An, N1, and N2 have similar trends as the case with 7t. = 1, until the orientation state becomes independent of Pe at large values of Pe. Based on previous experimental and theoretical studies, negative N1 is caused by tumbling phenomenon, which is predicted by this research. For )- 5 0.987, N] becomes independent of Pe for large values of Pe. The effect of tube dilation in the moment equation adds additional diffusive flux in the orientation state, which gives an extended tumbling region. This phenomenon broadens the Newtonian plateau region. When the tube dilation effect is included in the moment equation and in the viscous contribution of the stress, shear thickening occurs near the tumbling/wagging transition region. Moreover, the magnitudes of the normal ' stress differences are increased significantly. 143 In general, An and N1 are mainly determined by the viscous contribution to the stress, while the elastic contribution, which includes rotary Brownian motion and the excluded volume effect, is very small. The second normal stress difference, however, is determined mainly by the excluded volume contribution. These results offer useful insights on how to modify the stress model. 144 CHAPTER 10 CONCLUSIONS In this research, the Smoluchowski equation (S-equation) for the orientation density fimction was used to study the self-alignment and the flow-induced alignment of semi-dilute and concentrated suspensions of ellipsoidal particles. A Maier—Saupe excluded volume potential and Jefi‘ery’s model for rotary convection in a homogeneous shear field were used to close the S-equation. Doi’s model for the rotary diffusion coefficient accounts for the influence of tube dilation on highly aligned suspensions. Low-order moments of the orientation density fimction were used to quantify the relaxation of the microstructure from initial anisotropic states. An unclosed moment equation for the orientation dyad < pp > was derived from the S-equation. An algebraic closure for the orientation tetrad < pppp > , which appears in the moment equation for < pp > , was developed based on the hypothesis that all planar anisotropic states are attracted by three dimensional anisotropic states. The objective of this chapter is to summarize the salient conclusions of this research. The following four complementary topics, which relate to the self-alignment and the flow-induced alignment of rigid rod suspensions, will be addressed: 1) closure for the orientation tetrad; 2) equilibrium states; 3) non-equilibrium states; and, 4) rheology of rigid-rod suspensions. 145 Closure for the Orientation T etrad This research has developed a new algebraic closure approximation (FSQ- closure) for the orientation tetrad < pppp > in terms of the orientation dyad < pp > with the following three characteristics: 0 the closure retains the six-fold symmetry properties of the exact orientation tetrad; o the closure retains the six-fold contraction properties of the exact orientation tetrad; and, o the second-order FSQ-closure coefficient, defined by 2 = 8*“ mb , —is IIIb s—8— (10.13) 18(1+9 1111,) 36 36 1111- Etr[2°2'2] (10.1b) ga—%l , (10.10) has the feature that all planar anisotropic states are attracted by three dimensional anisotropic states (see Figure 1.1). The FSQ-closure for < pppp > is < E 222 > = [1 — c2(1nb)]31(< £2 >)+c2( IIIb)32(< 22 >). (10.2) The tetradic operators 310 and 32 (-) are defined by Eqs.(5.5) and (5.6), respectively. Eqs.(10.1a) and (10.2) are significant discoveries that will have an immediate impact on the further development and understanding of the flow (and light scattering) properties of suspensions, liquid crystalline polymers, and other fluids with microstructure. Figure 10.1 compares the steady state solutions of the moment equation (see ~ Eq.(4.1)) with statistical properties calculated from solutions of the S-equation deve10ped 146 0E 00.0 00.0 00.0 00.0 0 b I n 003 .0003. 000 0.030 _. 00000000 D 000.0 M & ~H.« 3 u H0 :0 0080 0000000: 3 No.0- 00.0 . mod . 9.0 . m; . 8.0 . mud - 00.0 - 00.0 0580-30 3:0 30003, 2: mama _0 00 00.0 000.0 0000.0 D n n o E: .320. 05 8050 3 00000000 D 0030 00000.0 00 .. NA 147 by Cintra and Tucker (1994). Cintra and Tucker solved Eq.(2.14) numerically subject to a uniform initial condition (isotropic) with 7» = l, U = O, FTD/Pe = 9C1, and t* = tPe. The orientation dyad was calculated directly from the orientation density function. No closure approximation was required for < p_ppp >. It is noteworthy that solutions to Eq.(4.1) with the realizable FSQ-closure for < pppp > predicts qualitatively the same steady states as the direct numerical simulation for a wide variation in the dimensionless diffusion coefficient C1. Figure 10.1a shows a comparison of the F SQ-closure predictions of < pp > : ezez with the results of Cintra and Tucker (1994). Figure 10.1b compares the relaxation of the microstructure (i.e., IIb and 1111,) for the two approaches. The results show that the FSQ-closure with C2(IIIb) defined by Eq.(6.28) provides a qualitatively consistent prediction of the exact statistical properties. Note that for C1 —> 0, the steady state approaches the nematic state (see Point A on Figure 1.1); and, as C1 —-> co, the steady state approaches the isotropic state. Earlier application of the FSQ- closure with C2(IIIb) = constant (see Parks et al., 1999; Nguyen et al., 2001; Kini et al., 2003) predicted unrealizable behavior for small values of C1. For example, if C2 = 1/3, unrealizable steady states occur for C1 < 0.01. Clearly, the discovery of the closure coefficient C2(IIIb) defined by Eq.(6.28) is a major accomplishment of this research. 148 Equilibrium States The steady state (equilibrium) solutions to Eq.(4.1) with Pe = 0 are either isotropic, prolate anisotropic, or oblate anisotropic. All initial planar anisotropic microstructures are attracted (relax) to the prolate anisotropic boundary of Figure 1.1. If the initial state is oblate anisotropic, then the equilibrium state is oblate anisotropic. However, if an infinitesimal disturbance produces a microstructure within the invariant domain (see Figure 1.1) near the oblate line, the relaxed equilibrium state will be prolate anisotropic, not oblate. Thus, oblate equilibrium states are stable to disturbance that are oblate; however, they are unstable to arbitrary realizable disturbance. Figure 7.2 shows how the steady state order parameter or (a (3/211b)1/2) depends on the excluded volume coefficient U. The isotropic state (a = 0) is a solution to Eq.(4. 1) for all U, 0 S U < 00. For U < U1, or = 0 is the only steady state solution. For U1 < U < U2, three equilibrium solutions exist: two are stable (isotropic and prolate anisotropic) and one is unstable (prolate anisotropic). The existence of this so- called biphasic region has been confirmed by experiments and by other theories (see CHAPTER 1). For U > U2 = 5, Eq.(4.1) still predicts the existence of three equilibrium states for the same value of U: l) a stable prolate anisotropic state; 2) an unstable isotropic state; and, 3) a conditionally stable oblate state. For U —> co, the prolate state approaches the nematic state (Point A on Figure 1.1) and the oblate state approaches the planar isotropic state (Point C on Figure 1.1) 149 Non-equilibrium States The class of transient solutions to Eq.(4.1) with the FSQ-closure is sensitive to the tumbling parameter )- and to the initial conditions for < _p_p > (0). For 7L = l, Pe > 0, and U > 0, all planar anisotropic states with the initial director in the flow/cross-flow plane (see Figure 8.8) relax to three-dimensional anisotropic states. None of these states are on the prolate boundary. Figure 8.2 shows the distribution of the steady state invariants. The locus of steady states (111,, 1111,)SS is parameterized by U and Pe. For fixed Pe, the steady states approach the prolate boundary as U increases. Also, for fixed U, the steady states approach the nematic state as Pe increases. Increases in U and Pe have qualitatively similar effect on director alignment. Thus, Figure 8.1 suggests that an initially isotropic material could be processed to attain the same three-dimensional anisotropic microstructure by following different paths in the plane variations in U and Pe. For A. < 1, Pe > 0, and U > 0, the asymptotic solutions to Eq.(4.1) with the F SQ- closure may be either steady or periodic. Periodic solutions require a relatively large value of the excluded volume coefficient (e.g., U > 27). If the initial director is in the flow/cross-flow plane and if U > 27, then the theory predicts director tumbling for Pe < 20 and director wagging as Pe increases. For large Pe and U > 27, the director relaxes to a steady state alignment relative to the flow direction. Figure 8.9 summarizes the physical nature of the asymptotic states for different combinations of U and Pe. This phase diagram or microstructure map, provides the insight needed to attain (or to avoid) specific microstructures. Table 8.1 summarizes an unanticipated result related to the alignment 150 phenomenon for large values of Pe. Note that for A. < l, the steady state invariants approach a highly aligned state independent of Pe; whereas, for 2» = 1, the steady state invariants continue to approach the nematic state (see Point A on Figure 1.1). Apparently, the rotational torque due to the vorticity (~ Pe) is unable to counter the strong rotational torque due to diffusion because the complementary rotational torque due to the strain rate (~lPe) is weaker with A. < 1. For high Péclet numbers (Pe > 90) and for high excluded volume coefiicients (U > 27), the asymptotic response of the director for 7s. < 1 depends on the initial condition < pp > (O). For example, if the initial director is in the flow/cross-flow plane, then the director relaxes to a steady state, which has a negative offset from the flow direction (see Figure 8.7; it = 0.987; U = 27; Pe = 95). However, if the initial director is in the vorticity/flow plane, director log-rolling occurs (see Figure 8.10; A = 0.987; U = 27; Pe = 95) with a positive offset from the flow direction. On the other hand, if the initial director is in the vorticity/flow plane and the initial condition of < pp > (0) has an off-diagonal component, then director kayaking occurs (see Figure 8.13; l = 0.987; U = 27; Pe = 95) with a negative offset from the flow direction. This type of behavior has been reported by others based on direct simulations of the S-equation, but the comprehensive set of results given in CHAPTER 8 has not been previously developed using low-order moments to the S-equation primarily because a realizable theory for < p p > was unavailable heretofore. 151 Rheologt of Rigid Rod Suspensions The microstructure of a rigid rod suspension has a significant impact on the viscous and elastic components of the deviatoric stress (see Eqs.(4.5) and (4.8)) inasmuch as .v .~ ; °C-§ - (10.3) . 1 ;E at [( —§;)—U(.:):| . (10.4) The tumbling parameter A, which controls the behavior of < pp > and, thereby, < pppp > , also impacts the rheology of the suspensions. For X = 1, the shear viscosity has a Newtonian-like behavior for low Pe and a shear thinning behavior at high Pe (see Figure 9.1). As indicated by Figure 9.2, the viscous component of the suspension viscosity determines the shear thinning behavior at high Pe. At low Pe, the elastic and viscous components are equally important. Figure 9.4 shows how U and Pe influence the shear component of the elastic stress for it = 1(see Eq.(10.4) above; and, Eq.(4.8)). It is not surprising that for Pe > 1, the contribution of the elastic stress decreases inasmuch as the excluded volume contribution to 152 diminishes as the director aligns with the flow. Note that the stress . - - V model (Eq.(4.7)) wrth the FSQ-closure for < pppp > (see Figure 9.4) predicts that ‘Cyz 1/3 E 1/3 at Pe for U = 0 and Pe > 10. However, for U = 5 and 0.2 < Pe < 2, ryz oc Pe also. E Thus, in a flow/ stop experiment for which iyz can be measured (see Smyth et al., 1995), . it is important to cover a wide range of U (CC C) and Pe (cc 7) to compare with appropriate 152 theoretical models. For 0.05 wt% xanthan gum in a fructose solution, Smyth et al. (1995) . . / -l . — . . . observed that 152 at y] 3 for 1 s < y < 20 s 1. A companson of 121118 experimental result with the microstructure theory developed herein indicates that DR ~ 1.3 s.1 if U E 5. This estimate is consistent with the rotary diffusion coefficients for PBLG solutions (Baek and Magda, 1993; also see Figure 9.4). As anticipated by the microstructure results in Table 8.1, the shear viscosity predicted for it < 1 has a Newtonian plateau for Pe —+ co (i.e., it = 0.987, U = 0, and Pe > 500). Once again, this phenomenon implies that shear thinning is limited by the weaker rotational torque due to particle coupling with the strain rate. Finally, Figure 9.19 shows that tube dilation causes shear thickening. This unanticipated result occurs provided tube dilation afi‘ects rotational diffusion in the moment equation and hydrodynamic drag in the viscous stress equation. This interesting (and perhaps doubtful) result requires additional research (see p.337 Doi and Edwards 1986) 153 CHAPTER 11 RECOMMENDATIONS The F SQ-closure for the orientation tetrad needs additional development and validation to become a practical closure for Eq.(4.1). Although the traditional models for the rotary convective flux and the rotary dtfiitsive flux used to close the Smoluchowski equation could be improved, the purpose of this chapter is to identify specific problems that would support the use of Eqs.(4.1) and (4.4). Suspension Theory In the approach adopted in this research, the tumbling parameter A (see Eq.(2.5)), the excluded volume coefficient U (see Eq.(2.12)), and the Péclet number Pe (see Eq.(2.15)) were treated as independent dimensionless groups. In general, 7», U, , and CR depend on the aspect ratio L/d, the volume fraction of the dispersed phase av, and the local microstructure (11b and 1111,): A=Fl(-I§,(1v , 11b, HIb) U =Fu(%-, (1v , 11b, IIIb) =FU(£,avaHbaIHb) Dfi d PEEL—Fol: a 11 111) kBT gd’ V’ b’ b' 154 A systematic parametric study of Eq.(4.1) and (4.4) with the F SQ-closure should be developed for suspensions of prolate ellipsoids for which 1 S L/d < 00. An explicit dependence of L/d and av on the above groups should be used in the parametric study. The tube dilation model of Doi was used in this research to determine the effect of the microstructure on /D§ and DfiQR lkBT. Shear thickening occurs because tube dilation was included in both factors. The influence of tube dilation on the tumbling parameter A should also be developed, if appropriate. An explicit dependence of L/d on the tumbling parameter was used in this research to study oblate and prolate ellipsoidal suspensions. However, the effect of L/d on and CR were not studied. Figure 11.1 shows the variation of the rotary drag coefficient with L/d for prolate ellipsoidal suspensions (see p.292, Doi and Edwards, 1986; and Jeffery, 1922). This theory, and its generalization to oblate ellipsoidal suspensions, should be used in the proposed parametric study of Eqs.(4.1) and (4.4). F SQ-closure coefficient The FSQ-closure was developed based on the Cayley-Hamilton theorem, which led to the representation of < pppp > given by Eq.(5.7). The second order closure coefficient was selected so that planar anisotropic states are attracted by three- dimensional anisotropic states for all values of 2», U, and Pe. The results in CHAPTER 7, 8, and 9 validated that this approach produces realizable orientation states for simple shear flows. However, the universality of C2(IIIb) as a second-order closure coefficient 155 000005000 00000.0 0020060085 05 :0 000% 0000m< 200.80 00 0.00-mm 05. 0.: 080E 0D coo? cor or F D I I o . S . No 002 000300 20 60 .32 89 \ mum 5.2 .. 0o 0% .- NGEV - 050 0% -N 050063: 120 4| 7 s . . ._ 0m 7 S0 1 NAE-cv +050 TN 05: _ . to . 3 - 0.0 156 needs to be tested for other flows. C2(IIIb) also needs to be validated experimentally for simple shear flows. As demonstrated in CHAPTER 7 as well as in CHAPTER 8, the relaxation of the microstructure to an anisotropic state has a dynamic signature that is sensitive to the closure model for < pppp > . For example, the initial trajectory from a planar anisotropic state depends on the closure assumption and could be tested directly (see Figure 7.4). Stress relaxation experiments and/or optical experiments that directly measure the relaxation of the director could be designed to test the ad hoc assumption of replacing Ineq.(6.27) with Eq.(6.28). Texture (Pe > 0) Figure 11.2 shows that Eq.(4.1) with an FSQ-closure has multiple steady states for the same parameter set (7L, U, Pe). Note that for A. = 0.987, U = 27, and Pe = 95, two stable steady state solutions to Eq.(4. 1) occur with distinct microstructures: (11b, IIIb)1 = (0.4923, 0.1381) and (111,, 1111,); = (0.5463, 0.1648). This result was" discovered serendipitously by numerically integrating Eq.(4.1) from two different initial conditions (see Figure. 11.2). The first microstucture was obtained by the relaxation of an initial director in the flow/vorticity plane, whereas the second microstructure resulted fiom the relaxation of an initial director in the flow/cross-flow plane. The existence of multiple steady states for Fe = 0 has been known for many years (see biphasic region in CHAPTER 7). The existence of multiple steady states for Fe > 0 have not been systematically studied or discussed in the literature. Larson (p.515, 1999) and Marrucci 157 Sad n a .3 u on .R u a mega aouflECO 3:5 30?; 8m mouflm >3on 038m 03252 N: 23me E wwd :3 No 2.6 3.0 Cd 9.0 mto 36 2.6 . . . . . . . . _ m3 0 . md 9%; .838 u E: 55 II / NmNm+sm$n§A§vN .39 a: $25 .838 u 2: ea .2 as? xmxmustmmvm . mad . 5.0 158 (1999) have previously suggested that multiple steady state phenomenon may be the underlying cause for the existence of small strains associated with defects in the microstructure (texture). For very small strain rates, the viscosity would be high, but would decrease significantly for marginal increases in the strain rate inasmuch as the multiple steady states would be eliminated. Marrucci has hypothesized that this phenomenon may explain the Region I behavior observed for some liquid crystalline polymers. Thus, the results illustrated by Figure 11.2 together with the possibility that multiple steady states for Fe > 0 may relate to Region I behavior supports the suggestion that a comprehensive study of the possible multiple steady state solutions to Eq.(4.1) with the FSQ-closure should be conducted for 0 < U < co and 0 < Pe < 00. This could provide significant new insights and understanding on the origins of texture and other phenomena associated with structured fluids. 159 APPENDICES 160 APPENDIX A Derivation of Moment Equations with Structure Tensor 161 The objective of APPENDIX A is to find the moment equation in terms of the structure tensor 2. From Eq.(4.1), the derivative of dynamic equation with the structure tensor 2 for Fm = 1 is starting from: d b U( -- =—+< >< >—< >< > 5, PP 22 22 2222) (M) +1Pe(S-+§—2§) Where b-< >-lI = 22 3=’ Pe= Y , y= 2SzS, __ 6DR == S=_I-§, §:Vu+Vu , = Y: = 2 and 8 1 a . .r —— s +fi-V< +W-< >+< >-W A.2 5t 6DR[[6—t - HEB == (>92 22 = ( ) Once the orientation dyad < pp > is substituted with 2 from left hand side of Eq.(A.1) becomes: II €> A I'd I'd V + A I'd I'd V Ilé’ || Ila l H + O‘ \_/ + /'_\ be I H llr—a + "c" \_/ HS» .3 ll llé’ "0‘ + llc‘ "2 (A3) Once the orientation dyad < p p > is substituted with 2 , right hand side of the first line in Eq.(A.1) becomes: 162 III/J A I'd I’U V + /\ I'd I'd Combining Eqs.(A.4) and (A5), 3? Note]: 62 1 5t — 6DR 1 s—tr 6DR since =\i=’:=:\iIT =0 And 5t +XPe(§:+: = —t1-(2) Therefore, t = O, 3(2) =0 and t > 0, 121-(b) = 0. A Ilc‘ "6* . .4; U) 110‘ + IIO" 110‘ l m—a l ooh—a llo‘ n6: 1 Ilo‘ x In: m u: re v b 1 = -2+U(§2+2-2—22< pp_pp >)+XPe(§§+=S__-I_J+§'§-2 3_S_) (i:+:_:) 5 —tr 6BR Dt 6b tr[—=—]=-tr(2)+U(z<-pp> —<22>2<1_)E(B'B) >) _2<(E.E) um I'd (A.5) (A.6) (A.7) (A.8) I“ Note2: T T 5b 6b . . [ =) --1— [—3] +Tle+lzT E 6DR at --———-——l —a—bT+< >T-WT+W'< >T 6DR 6t= 39 ‘= =' 22 (A9) a) _— =—q—)—tb =U(TT—) +2.Pe(T:§T +§T :T —2:§) And 2T=2, “(2)20 S181, tr<§=0 bT satisfies the same equation as _b_. If 2T = Eat t = 0, then 2T =2 for t > 0. Therefore, "(2) and the symmetry of g are “conserved.” 164 APPENDIX B Derivation of Moment Equations in the Invariant Form 165 The objective of APPENDIX B is to find the moment equations in terms of the invariants IIb and 1111, of the structure tensor 2. General Second Invariant Dynamic Equation The second invariant dynamic equation is derived by taking trace of the dot product of the structure tensor from Eq.(A.6): +UJ + APeK]-= b —+UJ+>.PeK] (3.1) + IO‘ lé ’-1 IICIT The trace of the ordinary derivative is: u[d—(2~2)]=%[trbb tr2-( 2)]=‘”‘—-b— dt = —2IIb + 2U[ J. b ]+ 2m: (3.2) l—_‘ M ucr I-———l Note that 0 (E-2)=2=E=(2-2) because b is traceless tensor. Since gT=-g,(g-g7):g=—(g-g)z g: (31; 2) £50. Therefore,the vorticity does not influence dynamic equation of second invariant. The second and third terms in the second line of Eq. (3.2) is 166 2:242 2+2 2):2—2 <2222>2 3; (2.3) =§Hb+IHb—g:2 and 22-3-2 2+2 2) 2+(2-2) 2—22 <2222>2 =§§=2+§ (2 2)+§=(2'2)—2§ 2 (3.4) =2§=E2+2 1.2-2 <2222>]=2§; Combining Eqs.(B.3) and (BA) into Eq.(B.2), the dynamic equation for the second invariant becomes: c111b dt =—211b+2U[;:g]+2xPe[;zg] (13.5) ll'-—1 H =—211b+2U[ ;Hb+IHb—2::g ]+4kPe[§: General Third Invariant Dynamic Equation The third invariant dynamic equation is derived by taking trace of another dot product from Eq.(B.1): __‘Lb b-b)=i‘2 b-b+b d_2.b+h 2.92 dt "" dt=a== _ t: ""dt =—32-2 2+Uu-2-2+2 2 2+2-2 J]+>~PeK-2-2+2 22+2 2 2] —[W 2+2 229-222 W 2+2 2?) 2+2-2- W 2+2-.W__T) (3.6) 1 2 ,where J=§2+2 2—2 and §=§§+§ 2+1; __S_-2<_pppp> § Notethat 167 And since 2“=—2, (22.“ )=(2-2)=-(2-2)=(2-2H222 )= "2 '50. Therefore, the vorticity does not influence dynamic equation of second invariant. The second and third terms in the second line of Eq. (B.6) d t V (13.7) (B.8) =—111b+ IIbZ—b <2222>Qz 2) and 2 (2 2)=ZS b 2)+s (2 2 2)+2 (2°--2)—22:<2222>=(22) =22 [—2 2+2 2 2—<2222> (2 2)] (39) Notice that tr(2 - 2 2 - __1;): -;—11b2 by using Caley-Hamilton theorem. Combining Eqs.(B.8) and (B9) into Eq.(B.6), the dynamic equation for the second invariant becomes: dIII t" =-3111b +3U[%Illb +-12—11b2 —b:<2222>:(2.2 )] 1 (13.10) +6>~Pe§:[g22+222-<2222>=<2~2)] 168 APPENDIX C Normal Vectors in the Invariant Diagram 169 The objective of APPENDD( C is to find the outward pointing normal vectors in the boundary of the invariant diagram. Normal vector _n of the invariant diagram is defined as: E = 1111911 +En1§111 (C-l) In order to be realizable for the orientation state, the dot product of invariant dynamic equations (dIIb/dt and dIIIb/dt) and the normal vector must be less than zero: n-d—£50 orn-hSO (C.2a) dt where d—-f- = _de 9H +——dmb gm, 2 = di ’dt (C.2b) dt d t d t ,/(d§ /d t)-(d§ /d t) Normal vector on the planar anisotropic line The relationship of 111J and 1m, is IIb = 2111b + 2/9 (C.3) Therefore slope on the planar anisotropic line is 2. Based on Eq.(C.3) the normalized vector of the planar anisotropic line is: 2 1 Then, the normal vector becomes 1 2 Eplanar anisotropic = ‘J‘ggll _ TEEIII (C-S) Normal vector on the prolate line From Eq.(3.18), the slope of the prolate line is (4/6)(6/111b)“3 (06) Then, normalized base vector becomes: 170 (4/6X6/Inb)“3 1 = /3 911+ /3 E111 J1+(4/9)(6/111b)2 J1+(4/9)(6/111b)2 Eprolate Therefore, the normal vector on the prolate line becomes: —1 (4/6)(6/111b)“3 /3 E11+ ,3 $2111 J1+(4/9)(6/111b)2 J1+(4/9)(6/111b)2 Qprolate = Normal vector at oblate line From Eq.(3.17), the normalized base vector becomes: (4/6X—6/mb)“3 e + -1 -—II J1+(4/9X—6/IIIb)2/3 fi+(4/9)(—6/mb)2’3 Hoblate = 2111 Therefore, the normal vector becomes: —1 + -(4/6)(—6/mb)“3 E blt = $11 9111 0 ac J1+(4/9X—6/IIIb)2/3 ,/1+(4/9)(-6/1111,)2’3 171 (C7) (C8) (C9) (010) APPENDD( D Derivation of Various Closure Analysis in the Invariant Form 172 The objective of APPENDIX D is to find the moment equations in terms of the invariants 11b and 1111, of the structure tensor ‘9 for various closure approximations. With the invariants of moment equations, the realizability of orientation dyad < p p > is determined. Realizability of Decoupling Closure Decoupling approximation provides realizable orientation dyad without the external field (Pe = 0). The moment equation can be represented with invariants of Q, then the moment equation can be represented with scalar value of 111, and 1111,. The invariant form of decoupling closure double dot with dyad 2 from Eq.(A.6) is: "O" <2222>=2=<_2><22> (DI) Based on general the invariant dynamic equations (see Eq. (B5) and Eq. (B.10)), each the double dot product of decoupling closure becomes: 2 <22> >=b (D.2) gz:(=b.-g) Since :b==(2+ 5:1}: b= b: g=llb (D.3a) and:(2_b_)= [b+-;—I):(=b-g)=tr(Q-=b_-g)+-;-tr(l=)-=g) IIIb+;IIb, (D.3b) the invariant form from Eq.(D.2) becomes: IIO" 32=Hb2 (D.4a) and 173 HO" :<_p_p>:(b~b)=IIbIIIb+-:1;Ilb2 (D.4b) Therefore, the invariant forms of moment equations for the decoupling closure, when Pe = O are: d 11" = —2111, +2U[ 1111, +111b «nbz ] (11561) d t 3 and d mb -_- —3111b +3U[%1Hb +24le -111be ] (D'Sb) Realizability at planar anisotropic line Substituting 11., and III}, relationship at planar anisotropic line (111, = 21111) + 2/9) into Eq.(D.5a) and Eq.(D.5b), the invariant form of moment equations become: dH —b— : 4111b —3+U[—8mb2 +3111}, +1] (D-6a) dt 9 9 81 dIIIb 2 7 2 =-3111 +U —4III +—III +— . D-6b (it b i b 9 b 81] ( ) Using normal vector analysis (see Bq.(C.2)), df dIIb dIIIb o—:-—= . _ + $0 D07 9 '[dt 911 dt9111) ( ) Therefore dot product becomes: — + d d t J? t J3 t (D.8) Since the maximum III}, value is 2/9, the dot product is always negative. Realizability prolate and oblate line 174 Using the same method from Eq.(D.6), the invariant form of moment equations on the prolate line become: 2 2 4 (111 III “ III ‘ 111 ‘ —9—=—12 —‘l 3 +2U 2 J 3’+1111,-3 ——‘-’- 3 (D-9a) dt 6 6 6 2 dIII m ‘ III ‘ dtb =—3lIIb +3U 31-1111, +6[—€b-)3 —6IIIb [-6—bj3 (ng) Afier taking dot product with normal vector from Eq.(C.10) to Eq.(D.9a) and Eq.(D.9b), Eq.(D.7) is zero. Analogous to the prolate line, the oblate line provides the same result. Therefore, decoupling closure in prolate line and oblate line are realizable. Realizability of Hand’s Closure From section 6.2 mention, Hand’s closure provides realizable orientation dyad on the prolate and the oblate line. Based on general the invariant moment equations (see Eq. (B5) and Eq. (B.10)), each the double dot products of Hand’s closure are: 2K PinPkPl >Hand=2 = bij [- 3%(13'121 +Iik1j1 + Iilek) 1 (p.10) + 7(Pipj11d + PiPkIjl +PiP11jk + IiijPl +IikPjP1 + Iiipjpkl] bij 1 2K P1P ijP1 >Hand=(2'2) =bij [‘Efiijlkl + 111:1 ji + 1111 jk) 1 (D.11) + 7(Pin1k1 + PiPkIjl + PiPlek + IiijPl +11kPjP1 + 1111’ij )](biybjy) The right hand side from Eq. (D.11) becomes 2 4 —II +—III D.12 15 b 7 b ( ) 175 —__-.. .—. _p Substituting Eq.(D.]Z) into Eq. (B.5), the second invariant moment equation becomes: c111,, = —2Fm11b + ZFTDUI: 373-11}, +21%] (D.13) The right hand side from Eq. (D.11) becomes 3 2 2 —II +_m D.14 7 b 15 b ( ) Substituting Eq.(D. 14) into Eq. (8.10), the third invariant moment equation becomes: dIIIb dt = —3FTDIIIb +3FTDU[ %mb +li411b2] (D.15) Realizability at planar anisotropic line Substituting 11., and 1111, relationship at planar anisotropic line (111, = 2111b + 2/9) into Eq.(D.]3) and Eq.(D.]S), the invariant form of moment equations become: c111,, 4 2 29 ——=—4HI -—-+U -+——III -16 (it b 9 [9 7 b] (D ) dIII 1’ =--3111b +U[-—6IIIb2 +3111b +327] (D.17) Using normal vector analysis (see Eq.(C.2)), df dIIb (1111b ' n--—==n- —e + e SO .18 _ dt ( -11 dt -111] (D ) T d t Therefore dot product becomes: _p_-ii:— 2(- 2 +9mb )(-105 + 2(8 + 4SIIIb)U) (D1813) d t 9453 Therefore, Hand’s closure is not realizable at a certain U and 111;, (where 2(8+ 45111b )U is more than 105). Realizability prolate and oblate line 176 The invariant forms of moment equation of Hand’s closure on the prolate line become: _2_ 2 (111 In 111 ‘ b —12 ——b 3+2U 9 —‘l 3 +9111}, (D.19) dt 6 5 6 7 4 dIII " b =—3IIIb+3U ;mb +29%? (p.20) 1/ 3 Multiplying Eq.(D.]9) and Eq. (D.20) with _%[%] (see Eq.(C.lO)), Eq.(D.7) b becomes zero. Therefore, 11 3% = 0 at the prolate line. Analogous to the prolate line, the oblate line provides the same result. Therefore, Hand’s closure on the prolate line and the oblate line are realizable. Realizability of HLl Closure HLl closure provides realizable orientation dyad on the prolate and the oblate line for Fe = 0. Based on general the invariant moment equations (see Eq. (BS) and Eq. (B.10)), each the double dot products of in 111.1 closure are: l b:< :=b —b. 6< >-b-< >-< >< >:b = 2222> = 5-< <22 .. 22 22 22 2 (D21) +21< 22> =-2 22<22><22>=2>=2 b:< >: ——b: 6< >b< ->< >< >:b = <2222 £22)= < 22> 2 <>22 22 22 ._. (D22) +21<252>=2-21< 22> - 2<2>=2>=(2-2) Eq. (D21) becomes 1 2 2 §(4IIIb +3111, + 2111, ) (D23) 177 Substituting Eq.(D.22) into Eq. (8.5), the second invariant moment equation becomes: d H b = -2FT'D Hb + ZFTD U[: -1- 1111) + l 11b --Z 111,2] (13.24) (1 t 5 5 5 Eq.(D.22) becomes %(§mb + 2111,1111; +9113) (D25) Substituting Eq.(D.25) into Eq. (B.lO), the third invariant moment equation becomes: (1111,, dt = —3FTDIIIb +3FTDU[ £112, €111,111}, «Lg-611$] (D26) Realizability at planar anisotropic line Substituting 11b and 1111> relationship at planar anisotropic line (11), = 2111b + 2/9) into Eq.(D.24) and Eq.(D.26), the invariant form of moment equations become: dIIb 4 3 ——=—4111 ——+U— —2+9111 5+72111 .27 dt b 9 405[( bx 1.)] (D ) dIII b =-3111b +U[(—2+9IIIb)(l+9OIIIb)]. (D28) Using normal vector analysis (see Eq.(C.2)), dm" 2m] 5 0 . (D.29a) Therefore dot product becomes: . d_§ _ 2(— 2 + 9111b )(45 + 2(—2 + 9111b )U) _n — (D.30b) d t 405J§ Therefore, HLl closure is not realizable at a certain U and IIIb (where 2(—2+9IIIb )U is more than 105). 178 Realizability prolate line The invariant forms of moment equation of HM on the prolate line become: 2 2 4 (111 111 ‘ " III " b =—12 —b— 3 +2U 9 E’- 3 +1111b £3 ——b- 3 (D31) dt 6 5 6 5 5 6 4 5 (1111b 1 6 1111, 3' 72 1111, 3 =—3111 +3U —111 +— — —— — .32 dt b 5 b 5L 6) 5i 6) (D ) 1/ 3 Multiplying Eq.(D.3 1) and .Eq. (D32) with -%[-III—:] (see Eq.(C.lO)), Eq.(D.7) becomes zero. Therefore, 11 g—gt = 0 at prolate line. Analogous to the prolate line, the oblate line provides the same result. Therefore, decoupling closure on the prolate line and the oblate line are realizable. Realizability of FSQ Closure FSQ-model provides realizable orientation dyad on the planar anisotropic line if 8+45111b 18(1+9111b)' 2 In addition, FSQ-model provides realizable orientation dyad on the prolate and the oblate line, regardless of C2 without the external field The first term of FSQ-model is Hand’s closure and it is already proven in Eq. (D.18), Eq. (D.19), and (D20). Based on general the invariant moment equations (see Eq. (B5) and Eq. (3.10)), each the double dot products of in second term of FSQ closure are: 179 "O" a, 2 2J2():2=C2(llb,111b)bij [-3-§< pupB >(Iij1kl +Iik1jl + Iilek) + <13in >++) 2 _7(< PiPY ><13:yl’i'>llcl+< Pka >Ijl+111 + Iij <1)ka >+Iik +111 <13ij >)]bij (D.33) IIO‘ :32():(2-2) 2 =C2(nbalnb)bij [§(Iij1kl +12;le +Iu1jk) + (<19in >+ +) 2 —7(< piPy >< 13ij >Ik1+Ijl+< P791 >11k +Iij <1)ku >+Iik <1)ij >< p721 >+In <1)ij >< PyPk >)](biybjy) (D.34) Eq. (D.33) becomes 4 1 2 10 2 4 2( 2 ) — II +— +— II +—II +—III =-—7II +8111 +30111 D.35 35("3)21]"7"7"105b b b () Substituting Eq.(D.35) into Eq. (B.5), the second invariant moment equation becomes: de 7 3 54 2 —=—2F II +2F U —II +—III ——II C .36 (it TD b TD [35 b 7 b 35 b 2] (D ) Eq. (D.34) becomes 4 1 2 3 2 10 3 2 2 54 — II +— +— III +—II +—II III =—II +—III +—II III .37 35(b3lzlib7b7bb7b15b35bb(D) Substituting Eq.(D.37) into Eq. (B.5), the third invariant moment equation becomes: dIIIb dt 7 l 2 54 =—3F III +3F U —III +—II -—II III C D.38 TD b TD [35 b 14 b 35 b b 2] ( ) 180 Realizability prolate line The invariant forms of dynamic equation of FSQ-model on the prolate line become: 2 2 4 (111 111 ‘ 111 “ 111 ‘ ——b =—12 —b 3 +2U 9 —b 3 +3112, -—1944 —b 3C2 (D39) dt 6 5 6 7 35 6 4 5 (1111}, 1 18 111,, 3 1944 111,, 3 =—3111 +3U —111 +—— — -— — c .40 dt b 5 b 7( 6 i 35 i 6 j 2 (D ) 1/3 Multiplying Eq.(D.39) and .Eq. (D.40) with "it—T6) (see Eq.(C.10)), Eq.(D.7) b becomes zero. Therefore, 11% = 0 at prolate line. Analogous to the prolate line, the oblate line provides the same result. Therefore, FSQ-model closure on the prolate line and the oblate line are realizable. 181 APPENDIX E Eigenvalues and Eigenvectors of the Orientation Dyad 182 The objective of APPENDIX E is to find the eigenvalues and eigenvectors when the dyadic valued operator has zero components in pxpy, pxpz, pypx, and psz. Thus, the eigenvalues and the eigenvectors can be solved algebraically. If the dyadic valued operator is given as: ’Z‘-i = Ai Ki (E-l) , then it can be expanded to: Pxpx 0 0 xix xix 0 Pypy Pypz xiy = A'i xiy a (E2) 0 pzpy Pzpz xiz xiz where ZSi is the eigenvector, related to the eigenvalues Xi Therefore, Pxpx xix = Ai xix (E33) PyPy Xiy + Psz Xiy = Xi xiy (5313) mm Xiz + Psz xiz = 1i Xiz (E30) Based on Eqs.(E.3b) and (E.3c), x1 x, xzy, x22, 1(3),, and X32 are only non-zero component of eigenvector if all of non-zero components of the dyad from Eq.(E.2) and all of the eigenvalues are different each other. Using Eqs.(E.3b) and (E.3c), [(pzpz -?~i) (Pypy ->~l)-pzpy Pszin2 = 0 (E4) For i = 1, Eq.(E.3a) becomes: PxPx xlx = 11 x1x (55) Therefore, 11 is the same as pxpx. 183 The first eigenvector related to 2.1 becomes: 1 x]: xlx 0 (E6) 0 For i = 2, Eqs.(E.3a) and (E4) becomes: pypzx22 X2 = — (E73) y (pypy "'12) [(prz —>»2> (pypy —82)—pzpy pyszxay = 0 (271» Since X32 is non-zero component, the inside bracket in Eq.(E.7b) can be solved using quadratic equation. _ (pypy +pzpz)-\/(pypy +Psz)2 -4(pypy pzpz -pypz pzpy) 7»2 2 (E8) From Eq.(E.7a), the second eigenvector related to lg becomes: P0P £2 2 - (WP: :12) x” (E9) 1 For i = 3, Eqs.(E.3b) and (E4) becomes: x22 ___ _ Pzpyx22 (£103) (13sz - 13) [(22192 -?~3) (pypy -?~3)-pzpy pyszX3z = 0 (31%) Since X32 is non-zero component, the inside bracket in Eq.(E.lOb) can be solved using quadratic equation. 184 2 _ (Pypy +pzpz)+J(pypy +2222) -4(pypy pzpz -pypz Pzpy) 2 El] 3 2 ( ) From Eq.(E.7a), the third eigenvector related to 23 becomes: ( 0 i x 3 = 1 x 22 (E12) _ Pzpy \ (1)sz _ 13 ) ) The eigenvectors are orthogonal to each other, the dot product of each eigenvector itself becomes: 2121=2222=23°23=1 (E13) Therefore, XIX =1 (E.l4a) P P 2 xZ-xz = 1+ _X_i_ xgy =1 (E.l4b) pypy '32 P P 2 953 33 = 1+[—z—y——] x§z :1 (2.146) Psz"x3 Solving Eqs.(E.14b) and (E.l4c), x3, __. 1 (E.lSa) PP 2 1+ ___2L£L_. Pypy"12 and 185 x2 = (E.le) z p p 2 pzpz-M l 0 O 22 = - pypz 1 (E.l6b) (Pypy "'12) p P 2 1 1+ —y-z_._ pypy ->~2 ( N 0 53 = 1 1 (E.l6c) -_..EL— 1 Pzpy 2 + _— ( (Psz ’96)} pzpz —7L3 186 APPENDD( F Realizability in the Planar Anisotropic Line 187 The objective of APPENDIX F is to derive the inequality equation in Ineq.(6.27). In order to find the closure coefficient C2, Ineq.(6.1) is applied to the planar anisotropic line. The outward pointing normal vectors in the prolate line are n}? :1}? (RI) nffi =72§ (F.2) In the planar anisotropic line Eqs.(6.2b) and (6.2c) become dIIb 4 4 58 16 96 423 2 _=———4111 +U—+—III — —+—111 +—111 c M dt 9 b [45 35 b (105 35 b 35 b) 2] ( ) dIIlb 2 83 6 2 36 324 2 ——=—3111 +U——+—III +-111 — —111 —-—m c F.4 dt b [189 105 b 7 b (35 b 35 b) 2] ( ) Multiplying Bq.(F .1) with Eq.(F.3) and Bq.(F .2) with Eq.(F .4), Ineq.(6.1) becomes dF 4 2 n-—==-—+—m ‘dt 96 J? b 64 8 12 2 +U + H1 --———--111 (F.5) [9454/3 105J§ 743 b 16 24 216 2 —( + 111 ——111 c ]so 105J§ 3545 b 353 b) 2 After arrange Eq.(F .5) the inequality becomes 2(—2 + 9111,, ){105 + 2U[—8 -45111b +18C2(l+ 1111,)]}< 0 9453 (R6) 188 APPENDIX G Computational Code for Transient Calculations 189 The objective of APPENDIX G is to demonstrate the computation scheme using flow chart and list the computational code for MATLAB 12. [start] 1 time step (dt), number iteration (stop). shape factor (lam), U, Pe, solving order (order), < p p > (a_real), Identity tensor (1D), velocity gradient (Du) symmetric Tensor (s), vorticity tensor (w) l Dummy <32) (a) l n = number iteration (n) = stop yes 1110 yes r = solving order (n) = 4? i no anisotropic tensor (b), eigen value of (lamda), eigen vector of anisotopic tensor (1amdab) l second invariant (iid). third invariant (iiid), C3 (C2). C1 (C1), difi‘usion coefficient (Drbar) l <2£>'<££> (a), 2'2 ('bb) ‘2? § (a5) <2£>'V“(aDu), <22>§ <22> (aSa). <22>°V° <22>caDua). ”kw-9 (traceas). ”(-V” (traceDu), c (aaa) b- b b’Cbbb) l _ <2222>=< 22>(ppppddpp).< 132>1§ (ppppdds). §+ S’(saas), § +<_p>-§’(waaw) i <> 0 o 190 moment equation (dk) 1 yes third Runge-Kutta iteration (dk3) fourth Runge-Kutta iteration (dk4) stress tensor (tau), stress components (W, X, Y. Z) <32 > tensor (a), components (A, B, C, D, F), physical time (time) / print the results /<—-— 191 %clf clear all %time step dt=0.003; %number of time step stop=20000; %tumbling parameter lam=0.987; %nematic potential U=27; %Péclet number Pe=30; %1st order Euler order=1 %2nd order R-K order=2 %4th order R-K order=4 order=1; %Constant Diffusion Coefficient Dr=1 %Doi Tube Dilation Dr=2 Dr=l; dbar=l; %Decoupling approximation closure=1 %Hybrid approximation closure=2 %FSQ approximation closure=3 closure=3; %Doi stress st=l %Ottinger stress st=2 st=1; %Constant C2 const=1 %function of IIIb const=2 const=2; c2const=l/3; %Initial Condition a_real(1,1)=0;a_real(1,2)=O;a_real(1,3)=0; a_real(2,1)=O;a_real(2,2)=1/3;a_real(2,3)=0; a_real(3,1)=0;a_real(3,2)=0;a_real(3,3)=2/3; %Unit tensor; ID(1,1)=1;ID(1,2)=0;ID(1,3)=0; ID(2,l)=O;ID(2,2)=1;ID(2,3)=O; ID(3,1)=0;ID(3,2)=0;ID(3,3)=1; %Velocity gradient for simple shear flow; 192 Du(l,l)=O;Du(1,2)=O;Du(1,3)=0; Du(2,l)=O;Du(2,2)=O;Du(2,3)=l; Du(3,l)=O;Du(3,2)=O;Du(3,3)=0; for i=l:3 for j=l:3 s(i,j)=.5*(Du(i,j)+ou(j.i)); W(i,j)=.5*(DU(i,j)-DU(j,i)); end end a_test=a_real; a=a_real; for n=l:stop for r=lzorder b=a-l/3*ID; %Determine eigenvalues; lamda1=a(1,l); lamda2=(a(2,2)+a(3,3)-((a(2,2)+a(3,3))“2- 4*(a(2,2)*a(3,3)-a(2,3)*a(3,2)))“(1/2))/2; lamda3=(a(2,2)+a(3,3)+((a(2,2)+a(3,3))“2- 4*(a(2,2)*a(3,3)-a(2,3)*a(3,2)))“(1/2))/2; lamdab1=lamdal-l/3; lamdab2=lamda2-l/3; lamdab3=lamda3-l/3; %determine C2 iiid=lamdab1“3+lamdab2“3+lamdab3“3; iid=lamdabl“2+lamdab2“2+lamdab3“2; if const== C2=c2¢onst; end if const==2 C2=(8+45*iiid)/(18*(1+9*iiid)); end C1=l-C2; C3=0; C4=1-C1-C2-C3; c2(n)=C2; %Tube Dilation if Dr== Drbar=dbar; end if Dr== 193 if st==l Drbar=(l-3/2*iid)“(-2); end if st==2 Drbar=(l-3/2*iid)“(-l); end end %Determine S[I(a*a)]; aa=a*a; bb=(a-ID/3)*(a-ID/3); aS=a*s; aDu=a*Du; aSa=aS*a; aDua=aDu*a; traceas=trace(aS); traceaDu=trace= ppppddpp=cl*((— l/35+1/7*IIa)*ID+3/35*a+4/7*aa)+c2*(2/35*IIa*ID- 2/7*IIIa*ID+6/7*aaa-2/7*aa+39/35*IIa*a); 194 %:S ppppddS=Cl*(- 2/35*s+1/7*(2*a*s+2*s*a+traceas*ID))+C2*(4/35*s*IIa+a*traceas+2*a Sa-2/7*(trace(aa*s)*ID+2*aa*s+2*s*aa)); ppppddDu=Cl*(- 2/35*Du+1/7*(2*a*Du+2*Du*a+traceaDu*ID))+C2*(4/35*Du*IIa+a*tracea Du+2*aDua-2/7*(trace(aa*Du)*ID+2*aa*Du+2*Du*aa)); PPPPddSC2=(4/35*s*IIa+a*traceas+2*aSa— 2/7*(trace(aa*s)*ID+2*aa*s+2*s*aa)); end saas=s*a+a*s; waaw=w*a+a*(W'); if st== ot=l; end if st== ot=(6*(l-3/2*iid))“(-1/2); end %Moment Equation dk=Drbar*((ID/3-a)+ot*(U*(aa-ppppddpp)))+lam*Pe*(saas- 2*ppppddS)-Pe*waaw; % lst order Eular if order== a=a+dk*dt; end %4th order Runge-Kuttar if order==4 if r== dk1=dk; a=a_real+dk1*dt*0.5; end if r==2 dk2=dk; a=a_real+dk2*dt*0.5; end if r== dk3=dk; a=a_real+dk3*dt; end if r==4 dk4=dk; a=a_real+(1/6)*(dkl+2*dk2+2*dk3+dk4)*dt; end end 195 %Second order Runge-Kuttar if orde == if r== a(i,j)=a_real(i,j)+dkl(i,j)*dt; end if r== dk2(i,j)=dk(i,j); a(i,j)=a_real(i,j)+(1/2)*(dkl(i,j)+dk2(i,j))*dt; end end end %Decomposition of Moment Equation test1(n)=U*(aa(2,3)-ppppddpp(2,3)); test2(n)=U*(aa(2,2)-ppppddpp(2,2)); test3(n)=U*(aa(3,3)-ppppddpp(3,3)); test4(n)=Pe*(saas(2,3)-2*ppppddS(2,3)-waaw(2,3)); test5(n)=Pe*(saas(2,2)-2*ppppddS(2,2)-waaw(2,2)); test6(n)=Pe*(saas(3,3)-2*ppppddS(3,3)-waaw(3,3)); test7(n)=ID(2,3)/3-a(2,3); test8(n)=ID(2,2)/3-a(2,2); test9(n)=ID(3,3)/3-a(3,3); test10(n)=dk(2,3); test11(n)=dk(2,2); test12(n)=dk(3,3); test13(n)=Pe*(saas(2,3)-2*ppppddS(2,3)); testl4(n)=Pe*(saas(2,2)-2*ppppddS(2,2)); tetslS(n)=Pe*(saas(3,3)-2*ppppdds(3,3)); test16(n)=-Pe*waaw(2,3); testl7(n)=-Pe*waaw(2,2); test18(n)=-Pe*waaw(3,3); test19(n)=(dk(2,2)-(ID(2,2)/3—a(2;2))-(U*(aa(2,2)- ppppddpp(2,2)))-(-Pe*waaw(2,2)))/(Pe*(saas(2,2)-2*ppppddS(2,2))); test20 (n)=(dk(3, 3) - (ID(3,3) /3-a(3,3) ) - (U* (aa(3,3)- ppppddpp(3,3)))-(-Pe*waaw(3,3)))/(Pe*(saas(3,3)-2*ppppddS(3,3))); test21(n)=(dk(2,3)-(ID(2,3)/3-a(2,3))-(U*(aa(2,3)- ppppddpp(2,3)))-(—Pe*waaw(2,3)))/(Pe*(saas(2,3)-2*ppppddS(2,3))); if st==1 tau=b-U*(aa-ppppddpp)+Pe*ppppddS/Drbar; end if st==2 tau=b-U/((l-IIa)“(l/2))/3*(aa-ppppddpp)+Pe*ppppddS/Drbar; end %Decomposition of Stress N1btest(n)=b(3,3)-b(2,2); Nlntest(n)=-U*(aa(3,3)-ppppddpp(3,3))+U*(aa(2,2)- ppppddpp<2.2)); Nlhtest(n)=Pe*ppppddS(3,3)/Drbar-Pe*ppppddS(2,2)/Drbar; N2btest(n)=b(2,2)-b(1,l); 196 N2ntest(n)=-U*(aa(2,2)-ppppddpp(2,2))+U*(aa(1,1)- ppppddpp(1.l)); N2htest(n)=Pe*ppppddS(2,2)/Drbar—Pe*ppppddS(l,l)/Drbar; shearb(n)=b(2,3)/Pe; shearn(n)=-U*(aa(2,3)—ppppddpp(2,3))/Pe; shearh(n)=Pe*ppppddS(2,3)/Drbar/Pe; All(n)=all; A22(n)=a22; A33(n)=a33; A23(n)=a23; a_real=a; %Components of and Stress W(n)=tau(2,3); X(n)=tau(2,2); Ytn)=tau(1,l); Z(n)=taU(3,3); A(n)=a(111); B(n)=a(2,2); C(n)=a(3.3); D(n)=a(2,3); F(n)=a(3,2); %Time step time(n)=(n-l)*dt; %First Normal Stress Nl=Z-X; %Second Normal Stress N2=X-Y; end %Eigenvalue and Eigenvector eigenvalue_1=A11; eigenvalue_2=(A22+A33—sqrt((A22+A33).“2-4.*(A22.*A33-A23.“2)))/2; eigenvalue_3=(A22+A33+sqrt((A22+A33).“2-4.*(A22.*A33-A23.“2)))/2; eigenvector_32=1./sqrt(1+(A23./(B-eigenvalue_2)).AZ); eigenvector_23=l./sqrt(1+(A23./(A33-eigenvalue_3)).“2); eigenvector_33=(-A23./(A33—eigenvalue_3))./sqrt(l+(A23./(A33- eigenvalue_3)).“2); eigenvector_22=(-A23./(A22-eigenvalue_2))./sqrt(1+(A23./(A22— eigenvalue_2)).“2); eigenvalueb_1=eigenvalue_1-1/3; eigenvalueb_2=eigenvalue_2-1/3; eigenvalueb_3=eigenvalue_3-l/3; %Shear Viscosity shear=W/Pe; 197 %Invariants II=eigenvalueb_l.“2+eigenvalueb_2.“2+eigenvalueb_3.“2; III=eigenvalueb_l.“3+eigenvalueb_2.“3+eigenvalueb_3.“3; %Invariant Domain xIII=[-l/36:0.000l:2/9]; yII=2/9+2*XIII; nyI=[O:0.000l:1/6]; xxIII=-6*(nyI/6).“(3/2); ynyI=[O:0.000l:2/3]; xxxIII=6*(ynyI/6)-“(3/2); plot(XIII,yII,xxIII,nyI,xxxIII,ynyI,III,II,'+') 198 LIST OF REFERENCES Abe, A. and T. Yamazaki, 1989, “Deuterium NMR Analysis of Poly(y-benzyl L- glutamate) in the Lyotropic Liquid-Crystalline State: Orientational Order of the a- Helical Backbone and Conformation of the Predant Side Chain,” Macromolecule, 22, 2138 — 2145. Abe, A. and T. Yamazaki, 1989, “Orientational Order of the a—Helical Poly(y-benzyl L- glutamate) in the Lyotropic Liquid-Crystalline State Comparison of Theory with Experiments,” Macromolecule, 22, 2145 — 2149. Advani, 8G. and CL. Tucker III, 1987, “The Use of Tensors to Describe and Predict Fiber Orientation in Short Fiber Composites,” J. Rheol, 31 (8), 751 — 784. Advani, S.G., and CL. Tucker III, 1990, “Closure Approximations for Three- Dimensional Structure Tensors,” J. Rheol, 34 (3), 367 — 386. Advani, S.G., 1994, Editor, Flow and Rheology in Polymer Composites Manufacturing, Volume 10, Composite Materials Series, Elsevier Publishers, Amsterdam. Anczurowski, E. and S. G. Mason, 1967, “The Kinetics of Flowing Dispersions III. Equilibrium Orientations of Rods and Discs (Experimental),” J. Colloid and Interface Science, 23, 533 — 546. Asada, T., T. Tanaka and S. Onogi, 1985, “Rheology of Liquid Crystalline Solution of a- helix Polypeptides,” J Appl Polym Sci Appl Polym Symp 41 — 229. Batchelor, G. K., 1976, “Brownian Diffusion of Particles with Hydrodynamic Interaction”, J. Fluid Mechanics, 74,1 - 29. Batchelor, G. K, 1982, “Sedimentation in a Dilute Polydisperse System of Interacting Spheres, Part I: General Theory”, J. Fluid Mechanics, 119, 379 — 408. Back, Seong-Gi, J .J. Magda and S. Cementwala, 1993, “Normal stress differences in liquid crystalline hydroxypropylcellulose solutions,” J. Rheol, 37, 935 — 945. Back, S.-G. and J. J. Magda, 1993, “Rheological differences among liquid-crystalline polymers. I. The first and second normal stress differences of PBG solutions,” J. Rheol, 37, 1201 — 1224. Baek, S.-G. and J. J. Magda, 1994, “Rheological differences among liquid-crystalline polymers. II. Disappearance of negative N1 in densely packed lyotropes and therrnotropes,” J. Rheol, 38, 1473 — 1503. 199 Bedford, B. and W. R. Burghardt, 1996, “Molecular Orientation of a Liquid-Crystalline Polymer Solution in Mixed Shear-Extensional Flows,” Journal of Rheology, 40 (2), 235 — 257. - ' Becraft, M. L. and A. B. Metzner, 1992, “The Rheology, Fiber Orientation, and Processing Behavior of Fiber-Fiber Fluids,” J. Rheol, 36 (1), 143-174. Be'nard, A., Mandal, D. K., S. M. Parks, and C. A. Petty, 2002, “A Closure Model for Fluids with Microstructure”, Poster Session 3.3, WCCM V - Fifth World Congress on Computational Mechanics, Vienna, Austria, July 7-12. Bhave, A. V., R. K. Menon., R. C. Armstrong and R. A. Brown, 1993, “A Constitutive Equation for Liquid-Crystalline Polymer Solutions,” J. Rheol, 37(3), 413 - 441. Bibbo, MA. and RC. Armstrong, 1988, “Rheology of Semi-Concentrated Fiber Suspensions in Newtonian and Non-Newtonian Fluids,” p.105 in Manufacturing International ’88, the Manufacturing Science of Composites, Editor: T.G. Gutowski, Vol. IV, ASME. Bibbo, M. A., S. M. Dinh and R C. Armstrong, 1985, “Shear Flow Properties of Semiconcentrated Fiber Suspensions,” J. Rheol, 29 (6), 905 — 929. Bird, R.B., R.C. Armstrong and O. Hassager, 1987, Dynamics of Polymeric Liquids: Volume I, Fluid Mechanics, Second Edition, Wiley-Interscience, New York. Bird, R.B., C.F. Curtiss, R.C. Armstrongand O. Hassager, 1987, Dynamics of Polymeric Liquids: Volume 2, Kinetic Theory, Second Edition, Wiley-Interscience, New York. Brenner H, 1974, “Rheology of a dilute suspension of axisymmetric Brownian particles,” International Journal of Multiphase F low, 1 (2), 195 - 341. Bretherton, F. P., 1962, “The Motion of Rigid Particles in a Shear Flow at Low Reynolds Number,” J. Fluid Mech., 14, 284 — 304. Burghardt, W. R. and G. G. Fuller, 1990, “Transient Shear Flow of Nematic Liquid Crystals: Manifestations of Director Tumbling,” Journal of Rheology, 34, 959 — 992. Burghardt, W. R. and G. G. Fuller, 1991, “Role of Director Tumbling in the Rheology of Polymer Liquid Crystal Solutions,” Macromolecules, 24, 2546 — 255 5. Carlsson, T., 1982, “The Possibility of the Existence of a Positive Leslie Viscosity (12. Proposed Flow Behavior of Disk Like Nematic Liquid Crystals,” Molecular Crystals and Liquid Crystals, 89, 57-66. 200 Carlsson, T. and K Skarp, 1986, “Observation of the tumbling instability in torsional shear flow of a nematic liquid crystal with (13 > 0,” Liquid Crystals, l, 455 — 471. Chandrasekhar, S., 1943, “Stochastic Problems in Physics and Astronomy, Review of Modern Physics,” 15 (1), p.1 — 89. Reproduced in Selected Papers on Noise and Stochastic Processes, Editor: N. Wax, Dover Publications, New York. Chaubal, C. V., 1997, Theoretical Models for the Dynamics of Liquid Crystalline Polymers, Ph.D. Thesis, University of California, Santa Barbara. Chaubal, C. V., G. Leal and G. H. Fredrickson, 1995, “A Comparison of Closure Approximations for the Doi Theory of LCPs,” J. Rheol, 39 (1), 73 - 103. Chaubal, C. V. and G. Leal, 1998, “A Closure Approximation for Liquid—Crystalline Polymer Models Based on Parametric Density Estimation,” J. Rheol, 42(1), 177 - 201. Chaubal C. V, and L. G. Leal, 1997, “Smoothed particle hydrodynamics technique for the solution of kinetic theory problems. Part 1. Method,” J. Non-Newtonian Fluid Mech., 70, 125 — 154. Chaubal C. V, and L. G. Lea], 1999, “Smoothed particle hydrodynamics technique for the solution of kinetic theory problems. Part 2. The efi‘ect of flow perturbations on the simply shear behavior of LCPs,” J. Non-Newtonian Fluid Mech., 82, 25 — 55. Chono, S., T. Tsuji and A. Taniguchi, 1996, “Numerical Analysis of the Rheology of Polymeric Liquid Crystals (1St Report, Shear Flow Behavior),” Transactions of the Japan Society of Mechanical Engineers, Part B, 62, 600-607. Cintra, J .S. and CL. Tucker III, 1995, “Orthotropic Closure Approximations for Flow- Induced Fiber Orientation,” J. Rheol, 39 (6), 1095 - 1122. Cladis, P. E. and S. Torza, 1975, “Stability of Nematic Liquid Crystals in Couette Flow,” Physical Review Letters, 35, 1283 - 1286. Clark, M. G., F. C. Saunders, I. A. Shanks, and F. M. Leslie, 1981, “A Study of Flow Alignment Instability During Rectilinear Oscillatory Shear of Nematic,” Molecular Crystals and Liquid Crystals, 79, 195 — 22. Clarke, A. R., N. C. Davidson and G. Archenhold, 1997, “Mesostructural Characteristics of Aligned Fiber Composites,” p.230 — 292 in F low Induced Alignment in Composite Materials, Editor: T.D. Papathanasian and DC. Guell, Woodhead Publishing Ltd, Cambridge, England. 201 Collyer 1992, Liquid Crystal Polymers: From Structures to Applications, Elsevier Applied Science, London. Currie, PK, 1982, “Constitutive Equations for Polymer Melts Predicted by the Doi- Edwards and Curtiss-Bird Kinetic Theory Models,” J. Non-Newtonian Fluid Mech. 11, 53 — 68. De, S. K. and J .R. White, 1996, Editors, Short fibre-polymer composites, Woodhead Publishing Ltd., Cambridge. de Gennes, P. G., 1974, The Physics of Liquid Crystals, Oxford University Press, London. Dinh, SM. and RC. Armstrong, 1984, “A Rheological Equation of State for Semiconcentrated Fiber Suspensions,” J. Rheol, 28 (3), 207 — 227. Doi, M. and S. F. Edwards, 1978, “Dynamics of Rod-like Macromolecules in Concentrated Solution, Part 1,” J. Chem. Soc. Faraday Trans. 11, 74, 560 — 570. Doi, M. and S. F. Edwards, 1978, “Dynamics of Rod-like Macromolecules in Concentrated Solution, Part 12,” J. Chem. Soc. Faraday Trans. II, 74, 918 - 932. Doi, M., 1981, “Molecular Dynamics and Rheological Properties of Concentrated Solutions of Rodlike Polymers in Isotropic and Liquid Crystalline Phases,” Journal of Polymer Science, 19, 229 — 243. Doi, M. and SF. Edwards, 1986, Theory of Polymer Dynamics, Oxford University Press, London. Donald, A. M., A. H. Windle,1992, Liquid Crystalline Polymers, Cambridge University Press, New York Doraiswamy, D. and A. B. Metzner, 1986, “The rheology of polymeric liquid crystals,” Rheologica Acta,25, 580 - 587 Edwards, BJ. and AN. Beris, 1989, “Flow Induced Orientation in Monodomain Systems of Polymeric Liquid Crystals,” J. Rheol, 33 (3), 537 — 557. Edwards, B]. and H. C. Ottinger, 1997, “T ime-Invariance Criteria for Closure Approximations,” Physical Review E, 56 (4) , 4097 - 4103. Faraoni, V., M. Grosso, S. Crescitelli, P. L. Maffettone, 1999, “The rigid-rod model for nematic polymers: An analysis of the shear flow problem,” J. Rheol, 43 (3), 829 — 843. 202 Farhoudi, Y. and AD. Rey, 1993, “Shear Flows of Nematic Polymers. I. Orienting Models, Bifiircations, and Steady State Rheological Predictions,” J. Rheol, 37 (2), 289-314. Feng, J. and LG. Leal, 1997, “Simulating Complex Flows of Liquid-Crystalline Polymers Using the Doi Theory,” J. Rheol, 41(6), 1317-1335. Feng, J., C. V. Chaubal and LG. Leal, 1998, “Closure Approximations for the Doi Theory: Which to Use in Simulating Complex Flows of Liquid-Crystalline Polymers?,”J. Rheol, 42(5), 1095-1119. Folgar, F. and CL. Tucker III, 1984, “Orientation Behavior of Fibers in Concentrated Suspensions,” J. Reinf. Plast. Campos, 98-119. Frattini, P. L. and G. G. Fuller, 1986, “Rheo-Optical Studies of the Effect of Weak Brownian Rotations in Sheared Suspensions,” J. Fluid Mech., 168, 119 — 150. Frazer, R.A., W.J. Duncan and AR. Collar, 1960, Elementary Matrices and Some Applications to Dynamics and Diflerential Equations, Cambridge. Goettler, LA, 1970, “Controlling Flow Orientation in Molding of Short-Fiber Compounds,” Modern Plastics, 47, 1- 140. Guell, DC. and A. Benard, 1997, “F low-Induced Alignment in Composite Materials: Current Applications and Future Prospects,” p.l — 42, in Flow Induced Alignment in Composite Materials, Editor: T.D. Papathanasian and DC. Guell, Woodhead Publishing Ltd., Cambridge, England. Gahwiller, CH., 1973, “Direct Determination of the Five Independent Viscosity Coefficients of Nematic Liquid Crystals,” Molecular Crystals and Liquid Crystals, 20, 301 — 318. Han, C. D., and S. S. Kim, 1993, “Transient rheological behavior of a thermotropic liquid-crystalline polymer. 1. The start-up of shear flow,” J. Rheol, 37, 847 - 866. Hand, G.L., 1962, “A Theory of Anisotopic Fluids,” J. Fluid Mech., 13, 33 — 46. Happel, J. and H. Brenner, 1965, Low Reynolds Number Hydrodynamics with Special Applications to Particulate Media, Prentice-Hall, Englewood Cliffs, NJ. Harlen, CG. and D.L. Koch, 1992, “Extensional Flow of a Suspension of Fibers in a Dilute Polymer Solution,” Phys. Fluids, 4(5), 1070 —— 1072. Harlen, 0G. and D.L. Koch, 1993, “Simple Shear Flow of a Suspension of Fibres in a Dilute Polymer Solution at High Deborah Number,” J. Fluid Mech., 252, 187 - 207. 203 Herczynski, R. and I. Pienkowska, 1980, “Toward a Statistical Theory of Suspensions,” Ann. Rev. Fluid Mech., 12, 237 — 269. Hinch, E. J. and L. G. Leal, 1972, “The effect of Brownian motion on the rheological properties of a suspension of non-spherical particles,” Fluid Mech., 52, 683 — 712. Hinch, E.J. and LG. Leal, 1975, “Constitutive Equations in Suspension Mechanics. Part 1. General Formulation,” J. Fluid Mech., 71, Part 3, 481 — 495. Hinch, E. J. and L. G. Leal, 1976, “Constitutive Equation in Suspension Mechanics. Part 2. Approximate forms for a suspension of rigid particles affected by Brownian Rotations,” J. Fluid Mech., 76, 187 — 208. Holmstrom, S. and S. T. Lagerwall, 1977, “Shear Flow in the Presence of Elastic Fields,” Molecular Crystals and Liquid Crystals, 38, 141 — 153. Imhoff, A., 2000, Diplomarbeit: Validation of Closure Models for Fiber Induced Alignment of Fibers, Mechanical Engineering, Michigan State University and University of Aachen, 2000. Imhoff, A., S. Parks, C. Petty, and A. Benard, 2000, “Validation of a New Closure Model for Flow-Induced Alignment of Fibers”, in Proceedings of the International Mechanical Engineering Conference and Exhibit, Symposium on CAE and Related Innovations for Polymer Processing, IMECE '00, Orlando, Florida, November 5-10. Ilg, P., I. V. Karlin and H. C. Ottinger, 1999, “Generating moment equation in the Doi model of liquid-crystalline polymers,” Physical Review E, 60(5), 5783 — 5787. 130, Y., D.L. Koch and C. Cohen, 1996a, “Orientation in Simple Shear Flow of Semi- Dilute Fiber Suspensions 1. Weakly Elastic Fluids,” J. Non-Newtonian Fluid Mech., 62, 115 — 134. Iso, Y., C. Cohen and D.L. Koch, 1996b, “Orientation in Simple Shear Flow of Semi- Dilute Fiber Suspensions 2. Highly Elastic Fluids,” J. Non-Newtonian Fluid Mech., 62, 135 - 153. Jeffery, GB, 1922, “The Motion of Ellipsoidal Particles Immersed in Viscous Fluid,” Proc. R. Soc. Land A, 102, 161 — 179. Jansson, J. F., 1992, “Applications of LCP Materials,” Liquid Crystal Polymers: From Structures to Applications, edited by A. A. Collyer, Elsevier, London. Kacir, L., M. Narkis and O. Ishai, 1975, “Orientation Short Glass-Fiber Composites. I Preparation and Statistical Analysis of Aligned Fiber Mats,” Polymer Engineering Science, 15, 525. 204 Kacir, L., M. Narkis and O. Ishai, 1975, “Orientation Short Glass-Fiber Composites. II Analysis of Parameters Controlling the Fiber/Glycerine Orientation Process, ’Polymer Engineering Science, 15, 532. Kacir, L., M. Narkis and O. Ishai, 1977, “Orientation Short Glass-Fiber Composites. III. Structure and Mechanical Properties of Molded Sheets,” Polymer Engineering Science, 17, 234. Kamal, MR. and A. T. Mutel, 1989, “The Prediction of Flow and Orientation Behavior of Short Fiber Reinforced Melts in Simple Flow Systems,” Polymer Composites, 10(5), 337. Kennedy, P., 1995, “Flow Analysis of Injection Molds,” Hanser Publishers, New York, NY. Kim, SS. and C. D. Han, 1993, “Effect of molecular weight on the rheological behavior of thermotmpic liquid-crystalline polymer,” Macromolecules, 26(24), 6633 — 6642. Kim, S. and S. J. Karrila, 1991, Microhydrodynamics: Principles and Selected Applications, Butterworth-Heineman, New York. Kim, Y-C, C. T. Nguyen, S. M. Parks, C. A. Petty, D. Mandal and A. Bénard, 2001, “Rheological and Alignment Properties of Liquid Crystalline Polymers”, Poster Session on Fluid Mechanics, Annual AIChE Meeting, Reno Hilton, Reno, NV, November 4-9. Kim, Y-C, C. T. Nguyen, S. M. Parks, D. Mandal, A. Bénard, and C. A. Petty, 2002, "Microstructure of Liquid Crystalline Polymers Induced by Simple Shear", Symposium on Manipulation of Nanophases by External Fields, Annual AIChE Meeting, Indianapolis Convention Center, Indianapolis, November 3 — 8. Kim, Y-C, H. Kini, D. Mandal, A. Bénard, and C. Petty, 2003, "Realizable Closure for the Orientation Tetrad for Rigid Rod Suspensions", Symposium on Suspensions, 56th Annual Meeting, American Physical Society: Division of Fluid Dynamics, 2003, East Rutherford, NJ, November 23-25. Kim, Y-C, C.A. Petty, and A. Bénard, 2004, "Equilibrium Microstructure of Complex Fluids", Symposium on Self-Assembly in Solutions 1, AIChE Annual Meeting, Austin, 7-12 November 7-12. Kim, Y-C, C.A. Petty, and A. Benard, 2005, “Microstructure of Multiphase Fluids in Homogeneous Shear Flows”, Symposium 144:Poster Session in Fluid Mechanics, Exhibit Hall A, Cincinnati Convention Center, AIChE Annual Meeting, October 31. 205 Kini, H., Y-C Kim, S. M. Parks, C. A. Petty, D. Mandal and A. Benard, 2002, "Flow Induced Microstructure of Composite Materials with Clay Fillers", Symposium on Particle Technology, Poster Session, Annual AIChE Meeting, Indianapolis Convention Center, November 3 — 8. Kini, H. K. 2003, Equilibrium Microstructure for Liquid Crystalline Polymers, Master of Science Thesis, Michigan State University. Kini, H., Y.C. Kim, C. T. Nguyen, A. Bénard, and C. A. Petty, 2003, "Flow Induced Microstructure in Composite Materials", Proceedings of 141' h International Conference on Composite Materials, ICCM14, CD-ROM, San Diego, July 14-18. Kiss, G. and R. S. Porter, 1980, “Rheo-Optical Studies of Liquid Crystalline Solutions of Helical Polypeptides,” Molecular Crystals and Liquid Crystals, 60, 267-280. Koch, D. L., 1995, “A Model for Orientational Diffusion in Fiber Suspensions,” Phys. Fluids, 7(8), 2086-2088. Kubo, K and K. Ogino, 1979, “Comparison of Osmotic Pressure for the Poly(y-benzyl- L-glutamate) Solutions with the Theories for a System of Hard Spherocylinders,” Mol. Cryst. Liq. Crsyt., 53, 207 — 228. Kuzuu, N. and M. Doi, 1983, “Constitutive Equation for Nematic Liquid Crystals under Weak Velocity Gradient Derived from a Molecular Kinetic Equation,” Journal of the Physical Society of Japan, 52(10), 3486-3494. Kuzuu, N. and M. Doi, 1984, “Constitutive Equation for Nematic Liquid Crystals Under Weak Velocity Gradient Derived fiom a Molecular Kinetic Equation. II. —Leslie Coefficients for Rodlike Polymers— ,” Journal of the Physical Society of Japan, 53(3), 103 1-1040. Lamb, H., 1932, Hydrodynamics, Dover Publications, New York. Larson, R. G., 1990, “Arrested Tumbling in Shearing Flows of Liquid Crystal Polymers,” Macromolecules, 23, 3983 — 3992. Larson, R. G. and H. C. Ottinger, 1991, “Effect of Molecular Elasticity on Out-of-Plane Orientations in Shearing Flows of Liquid-Crystalline Polymers,” Macromolecules, 24, 6270 — 6282. Larson, R. G., 1988, Constitutive Equations for Polymer Melts and Solutions, Butterworths, Boston. Larson, R. G., 1999, The Structure of Complex Fluids, Oxford University Press, New York. 206 Leal, LG, 1980, “Particle Motions in Viscous Fluid,” Ann. Rev. Fluid Mech., 12, 435 — 476. Leighton, D. T. and A. Acrivos, 1987, “The Shear-induced migration of particles in concentrated suspension,” J. Fluid Mech. 181, 415-439. Lipscomb II, G.G., M.M. Denn, D.U. Hur and D.V. Bogar, 1988, “The Flow of Fiber Suspensions in Complex Geometries,” J. Non-Newtonian Fluid Mech., 26, 297-325. Magda, J. J., S-G Baek, K L. DeVries and R. G. Larson, 1991, “Shear Flows of Liquid Crystal Polymers: Measurements of the Second Normal Stress Difference and the Doi Molecular Theory,” Macromolecules, 24, 4460-4468. Maier, W. and A. Saupe, 1959, “Eine einfache molecular-statistische Theorie der nematischen kristallinfliissigen Phase,” Z. Naturforsch, 13A, 564-566. Mandal, D, S. M. Parks, C. A. Petty, and A. Bénard, 2001, "Discussion of a Closure Model for Fiber Orientation", Proceedings of the Seventh Annual Meeting of the Polymer Processing Society, Montreal, Canada, May 21-24. Mandal, D., Bénard, A., Petty, CA, 2004, “Modeling Flow-Induced Orientation of Fibers Using a New Closure Model”, International Conference on Flow Processes in Composite Materials, Newark, DE, July 7-9. Mandal, D. K 2004, Simulation of Flow-Induced Fiber Orientation with a New Closure Model Using the Finite Element Method, Ph.D. Dissertation, Michigan State University. Marrucci, G. and P. L. Maffettone, 1990, “Nematic Phase of Rodlike Polymers 1. Prediction of Transient Behavior at High Shear Rates,” J. Rheol., 34, 1217-1230. Marrucci, G., 1991, “Tumbling Regime of Liquid-Crystalline Polymers,” Macromolecules, 24, 4176-4182. Marrucci, G., 1996, “Theoretical aspect of the flow of liquid crystal polymers,” p.31 — 48, in Rheology and Processing of Liquid Crystal Polymers, Editor: D. Aciemo and A. A. Collyer, Chapman & Hall, London. Milliken, W.J. and R.L. Powell, 1994, “Short-Fiber Suspensions,” p.53 — 83, in Flow and Rheology in Polymer Composites Manufacturing, 10, Composite Materials Series, Editor: S.G. Advani, Elsevier Publishers, Amsterdam. Mondy, L. A., H. Brenner, S. A. Altobelli, J. R. Abbott and A. L. Grahman, 1994, “Shear-Induced Particle Migration in Suspensions of Rods,” J. Rheol., 38(2), 444 — 452. 207 Murthy, S. N., J. R. Knox, and E. T. Samulski, 1976, “Order Parameter Measurement in Polypeptide Liquid Crystals,” Journal of Chemical Physics, 19, 4823 — 4839. Nguyen, C. T. 2001, Microstructure of Liquid Crystalline Polymers in Simple Shear F lows, Master of Science Thesis, Michigan State University. Nguyen, C. T., S. M. Parks, A. Bénard and C. A. Petty, 2001a, “Prediction of Low-Order Orientation Statistics for Alignment of Liquid Crystalline Polymers in Homogeneous Shear,” Proceedings of 13th International Conference on Composite Materials, ICCM13 CD-ROM, Beijing, June 25-29 Nguyen, C. T., Y-C Kim, S. M. Parks, C. A. Petty, D. Mandal, and A. Benard, 2001b, “Flow—Induced Alignment of Liquid Crystalline Polymers”, Symposium on Polymer Processing and Rheology 111, Annual AIChE Meeting, Reno Hilton, Reno, NV, November 4-9. Odijk, T., 1986, “Theory of Lyotropic Polymer Liquid Crystals,” Macromolecules, 19, 2313 — 2329. Orwoll, R. D. and R. L. Vold, 1971, “ Molecular Order in Liquid Crystalline Solutions of Poly (y-Benzyl-L-Glutamate) in Dichloromethane,” J. Am. Chem. Soc., 93, 5335 -— 5338. Onsager, L., 1949, “The Effect of Shape on the Interaction of Colloidal Particles,” Ann. N. Y. Acad. Sci, 51, 627 — 659. Papathanasiou, T.D., 1996, “Microstructure Evolution During Molding of Particulate Reinforced Thermoplastic Composites,” International Polymer Processing, 11(3), 275 - 283. Papathanasiou, T.D., 1997, “Flow-Induced Alignment in Injection Molding of Fiber- Reinforced Polymer Composites,” p.112 — 165, in Flow-Induced Alignment in Composite Materials, Editors: T. D. Papathanasiou and DC. Guell, Woodhead Publishing Ltd., Cambridge, England. Papathanasiou, T. D. and DC. Guell, 1997, F low-Induced Alignment in Composite Materials, Woodhead Publishing Ltd, Cambridge, England. Parks, S. M. 1997, Relaxation Model for Homogeneous Turbulent F lows, Ph.D. Dissertation, Michigan State University. Parks, S.M., K Weispfennig, C. A. Petty, 1998, “An Algebraic Preclosure Theory for the Reynolds Stress”, Physics of Fluids, 10(13), 645 - 653. 208 Parks, S. M. and C. A. Petty, 1999, “Prediction of Low-Order Orientation Statistics for Flow-Induced Alignment of Fibers and Platelets”, Fundamental Research in Fluid Mechanics: Particulate and Multiphase Flow 1, AIChE Annual Meeting, Dallas, TX, October 31- November 5. Parks, S. M., C. A. Petty, and S. M. Shao, 1999, “Flow Induced Alignment of Fibers,” in Proceedings of 12th International Conference on Composite Materials, ICCM12/T CA, Paris, July 5-9. Petty, C. A., S. M. Parks, and M. Shafer, 1999, Flow-Induced Alignment of Fibers in the Absence of Fiber-Fiber Interactions, paper presentation, Symposium on Suspensions, APS/DFD, New Orleans, LA, November 21-23. Phan-Thien, N., 1995, “Constitutive Equation for Concentrated Suspensions in Newtonian Liquids,” J. Rheol., 39 (4), 679 — 695. Phan-Thien, N. and R. Zheng, 1997, “Macroscopic Modelling of the Evolution of Fibre Orientation During Flow,” p.77 — 111, in Flow Induced Alignment in Composite Materials, Editors: T.D. Papathanasian and DC. Guell, Woodhead Publishing Ltd., Cambridge, England. Picken, S. J., J. Aerts, H. L. Doppert, A. J. Reuvers, and M. G. Northolt, 1991 “Structure and rheology of aramid solutions: transient rheological and rheooptical measurements,” macromolecules, 24(6), 1366 — 1375. Pieranski, P. and E. Guyon, 1974, “Two Shear-F low Regimes in Nematic p-n- Hexyloxybenzilidene-p’-aminobenzonitrile,” Physical Review Letters, 32, 924 — 926. Rahnama, M., D.L. Koch and E.S.G. Shaqfeh, 1995, “The Effect of Hydrodynamic Interactions on the Orientation Distribution in a Fiber Suspension Subject to Simple Shear Flow,” Phys. Fluids, 7 (3), 487-506. Rahnama, M., D.L. Koch and C. Cohen, 1995, “Observations of Fiber Orientation in Suspensions Subjected to Planar Extensional Flows,” Phys. Fluids, 7(8), 1811 - 1817. Ranganathan, S. and S.G. Advani, 1997, “Fiber-Fiber and Fiber-Wall Interactions During the Flow of Non-Dilute Suspensions,” p.43 - 76, in Flow Induced Alignment in Composite Materials, Editors: T.D. Papathanasian and DC. Guell, Woodhead Publishing Ltd., Cambridge, England. Rey, A.D., 1997, “Theory and Simulation of Shear Flow-Induced Microstructure in Liquid Crystalline Polymers,” p.203 — 229, in Flow Induced Alignment in Composite Materials, Editors: T.D. Papathanasian and DC. Guell, Woodhead Publishing Ltd., Cambridge, England. 209 Rey, AD. and M.M. Denn, 2002, “Dynamical Phenomena in Liquid-Crystalline Materials,” Annual Review of Fluid Mechanics, 34, 233 — 266. Robinson, C., 1966, “The Cholesteric Phase in Polypeptide Solutions and Biological Structure,” Mol. Cryst, l, 467 — 494 . Russo, PS, 2001, “Diffusion of Tabacco Mosaic Virus Through Solutions of Dextran,” see web site www.chem.lsu.edu/web/f2_1cultvbiofls.html. Sartirana, M. L., M. Marsano, E. Bianchi, 1987, “Order Parameter in Polymer Liquid Crystal 2. Poly(y-Benzyl-L-Glutamate) in Dioxane,” Mol. Cryst. Liq. Cryst, 144, 263 — 274. Servais, C. A., J. E. Manson and S. Toll, 1999, “Fiber-Fiber Interaction in Concentrated Suspensions: Dispersed Fibers,” J. Rheol., 43(4), 991 — 1004. Servais, C., A. Luciani and J. E. Manson, 1999, “Fiber-Fiber Interaction in Concentrated Suspensions: Dispersed Fiber Bundles,” J. Rheol., 43(4), 1005 — 1018. Shaqfeh, E. S. G. and D. L. Koch, 1990, “Orientational Dispersion of Fibers in Extensional Flows,” Phys. Fluids A, 2(7), 1077 — 1093. Skarp, K and T. Carlsson, , 1978, “Influence of an Electric Field on the Flow Alignment Angle in Shear Flow of Nematic Liquid Crystals,” Molecular Crystals and Liquid Crystals, 49, 75 — 82. Skarp, K., S. Carlson,rT. Lagerwall, and B. Stebler, 1981, “Flow Properties of Nematic 8CB: An Example of Diverging and Vanishing (13,” Molecular Crystals and Liquid Crystals, 66, 199 — 208. Soize, C., 1995, The Fokker-Planck Equation for Stochastic Dynamical Systems and Its Explicit Steaay State Solutions, World Scientific, River Edge, NJ. Srinivasarao, M. and G. C. Barry, 1991, “Rheo-optical Studies on Aligned Nemactic Solutions of a Rodlike Polymer,” Journal of Rheology, 35(3), 379 — 397. Stover, C.A., D.L. Koch and C. Cohen, 1992, “Observations of Fibre Orientation in Simple Shear Flow of Semi-Dilute Suspensions,” J. Fluid Mech., 238, 277-296. Szeri, A. J. and L. G. Leal, 1993, “Microstructure suspended in three-dimensional flows,” J. Fluid Mech., 250, 143 — 167. ' Smyth, S. F. and M. E. Mackay, 1994, “The viscous stress contribution to lyotropic hydroxypropylcellulose solutions in the biphasic and liquid-crystalline regions,” J. Rheol., 38(5), 1549 - 1558. 210 Smyth, S. F ., C. Liang, M. E. Mackay, and G. G. Fuller, 1995, “The stress jump of a semirigid macromolecule after shear: Comparison of the elastic stress to the birefiingence,” J. Rheol., 39(4), 659 — 672. Tetlow, N., A. L. Graham, M. S. Ingber, S. R. Subia, L. A. Mondy and S. A. Altobelli, 1998, “Particle Migration in a Couette Apparatus: Experiment and Modelling,” J. Rheol., 42(2), 307 — 327. Trevelyn, B. J ., and S. G. Mason, 1951, “Particle motions in sheared suspensions. I. Rotations,” J Colloid Science, 6, 354 — 367. Tucker, C.L., 1988, “Predicting Fiber Orientations in Short Fiber Composites,” p.95 in Proc. Manufacturing Int ’1, AMSE Publication, New York.. Tucker, C.L., 1991, “Flow Regimes for Fiber Suspensions in Narrow Gaps,” Journal of Non-Newtonian Fluid Mechanics, 39(3), 239 — 268. Tucker 111, CL. and S.G. Advani, 1994, “Processing of Short-Fiber Systems,” p.147 —202, in F low and Rheology in Polymer Composites Manufacturing, 10, Composite Materials Series, Editor: S.G. Advani, Elsevier Publishers, Amsterdam. Tucker, C. L. and P. Moldenaers, 2002, “Microstructural Evolution in Polymer Blends,” Annual Reviews of Fluid Mechanics, 34, 177 - 210. Vermant, J., P. Moldenaers, S. J. Picken, J. Mewis, 1994, “A comparison between texture and rheological behavior of lyotropic liquid crystalline polymers during flow,” J Non- Newtonian Fluid Mech, 53, l — 23. Wahl, J. and F. Ficher, 1973, “Elastic and Viscosity Constants of Nematic Liquid Crystals from a New Optical Method,” Liquid Crystals, l, 455 - 471. Walker, L. M. and N. J. Wagner, 1994, “Rheology of region I flow in a lyotropic liquid crystal polymer: The effects of defect texture,” Journal of Rheology, 38, 1525 — 1547. Walker, L. M., N. J. Wagner, R. G. Larson, P. A. Mirau and P. Moldenaers, 1995, “The rheology of highly concentrated PBLG solutions,” Journal of Rheology, 39, 925 — 952. Weispfennig, K, S. M. Parks, and C. A. Petty, 1999, “Isotropic Prestress for Fully Developed Channel Flows”, Physics of Fluids, 11(5), 1262-1271. Williams, M., 1975, “Molecular Rheology of Polymer Solutions: Interpretation and Utility,” AlChE Journal 21, 1-25. 211 Yousefi, H., G. Wiberg, M.-L. Skytt, J .J . Magda and U.W. Gedde, 2003, “Development and relaxation of orientation in sheared concentrated lyotropic solutions of hydroxypropylcellulose in m-cresol,” Polymer, 44, 1203—1210. 212