473.5: . . unnamfiflnmwfil: 1h? .. m... if .. z .E "a. 5 J V 1 ‘15.. . vhnirfir 31: .z .03 . ‘ .5 .1»: r y Y 53... I. #1:“! 5.. fun. 33.15%: i . 39.84..» hit a Ewing,” .Hz... . ‘ [1m x? 3 “flaififi .. g fimwfi. : , . . . x i , V ‘ ‘ .1? THESIS ”is LIBRAva * Michigan State University This is to certify that the dissertation entitled Pure Azimuthal Shearing Deformations for an Extended Class of Elastic Materials that are Characterized By a Microstructurally Motivated lntemal Balance Principle presented by Hasan Demirkoparan has been accepted towards fulfillment of the requirements for the Ph. D. degree in Mathematics Wj Mum M j r Professor’s Signature ' 06/13/2005 ‘ 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 2/os'—"cm__—c:/crn momma-p.15" fi.‘ Pure Azimuthal Shearing Deformations for an Extended Class of Elastic Materials that are Characterized By a Microstructurally Motivated Internal Balance Principle By Hasan Demirkoparan A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 2005 ABSTRACT Pure Azimuthal Shearing Deformations for an Extended Class of Elastic Materials that are Characterized By a Microstructurally Motivated Internal Balance Principle By Hasan Demirkoparan In the event of certain substructural reconfiguration in solids, it is sometimes the case that the notion of simple deformation X —+ x must be broadened so as to incorporate a more detailed kinematic description. This typically involves new kinematic variables in addition to x = x(X). Each new kinematic variable is in turn associated with an additional balance principle of the same tensor order as the new kinematic field variable. Gurtin’s theory of configurational forces and the Ericksen model of liquid crystals are examples of such theories. Recently a new constitutive framework for treating finite deformations when con- ventional elastic behavior is modified by the action of an additional microstructural balance requirement was proposed by Pence and Tsai. This framework offers inter- esting possibilities for the development of singular surfaces that can be interpreted as locations of concentrated microstructural rearrangement. This thesis investigates the solution of the pure azimuthal shear problem in this new framework. To my parents and my wife, iii ACKNOWLEDGMENTS I am very grateful to have known Prof. T. J. Pence both as a person and as a mathematician. It has been a privilege to work with him, and this thesis would have not been possible without his guidance. I would like to thank to the other thesis committee members, Professors C. Y. Wang, C. Wei], H. Tsai and K. Promislov for their advise and helpful discussions. iv TABLE OF CONTENTS LIST OF TABLES ....................... vi LIST OF FIGURES ............................................................................. vii 1 Introduction ................................................................................... l 2 .Pure Azimuthal Shear Problem for a Neo-Hookean material ........ 5 2.1 Introduction .................................. 5 2.2 Pure Azimuthal Shear Problem for a Neo-Hookean material ....... 5 3 The Internally Balanced Material ................................................. 14 3.1 Formulation of the Problem for the Internally Balanced Material ..... 14 3.2 Solution for Small k (6: << 0.") ....................... 21 3.3 Numerical Solution .............................. 27 3.4 Explicit Solution when d = a‘ (k = 1) ................... 39 3.5 Alternative Derivation from Energy ..................... 46 3.6 Direct Methods for Minimization when k = 1 ............... 56 4 A More Generalized Class of Material Models ............................. 74 4.1 Extended Material Model .......................... 74 4.2 Existence and Uniqueness for the Extended Model ............. 78 4.3 Numerical Solution for the Extended Model ................ 89 4.4 Regular Perturbation Solution for the Extended Model .......... 94 . 4.5 Boundary Layer Type Solution for the Extended Model .......... 100 4.6 Twist, Energy and Torque in the Extended Model ............. 114 5 Conclusions and Discussions ............ i ............................................. 118 A g(r) given by (3.115) is a strong local minimum of (3.168) if [till 5 ‘Ilm .............................................................................................. 120 B Alternate Proof of Theorem 4.2.3 ................................................ 127 BIBLIOGRAPHY ............................................................................... 134 3.1 3.2 4.1 4.2 LIST OF TABLES Normalized energies for A1 and Any for several 1/10 values when R.- = 1, Ro=4andtb,-=O ............................ 61 Normalized minimum energies A4(nmin) for several 1,00 values when R.- = 1,Ro=4andt/J,-=0 .......................... 65 N umerically computed values of 1" and do for one term patching for several t/Jo ando‘rvalues when R,=1, Ro=4 and 114:0 .......... 110 Numerically computed values of 7“, do and (II for two term patching for several we and 61 values when R,- = 1, R0 = 4 and z/Ji = 0 ...... 111 vi 2.1 3.1 3.2 3.3 3.4 3.5 3.6 LIST OF FIGURES Deformation of an originally vertical line segment for different amount of twist we and w.- = 0. Specifically shown are we = .75413, 1.50826, 3.0165, 5.49779 corresponding respectively to 43.20° , 86.41°, 129.62°, 315° for a neo—Hookean material in standard nonlinear elasticity theory. Note that R,- = 1 and Re 2 4 are used and multiples of arcccs(Rf/Rg) = 1.50826 are selected in order to provide correspon- dence with results of the more general theory that is the object of this thesis .................................... Comparison of normalized torque-twist relation when k = 0 and k = 0.02 for R..- = 1 and Re = 4. The dashed line shows the case k = 0.02. The normalized torque is M / (47r h 61 R3123) .............. Graphs of g(R) for five values of we: we 2 1, we = 2, we = 3, we = 5 and we=10 when R;=1, Re=4, w,=0 and k=0.1. For we=5 and we = 10 there is an apparent discontinuity at R,- -- 1 ....... Graphs of p(R) for five values of we: we =1, we = 2, we = 3, we = 5 and we =10 when R,- =1, Re 2 4, 103': 0 and k = 0.1. The top curve corresponds to we 2 1 and the middle one corresponds to we = 2. The graphs of p(R) are almost identical for we = 3, we = 5 and we = 10 and so generates an essentially common lower curve .......... Graphs of g(R) for five values of we: we = 1, we = 2, we = 3, we = 5 and we=10 when R,=1, Re=4, wi=0 and k=0.5. For we=2, we = 3, we = 5 and we 2 10 there is an apparent discontinuity at R, = ................................... Graphs of p(R) for five values of we: we :..— 1, we 2 2, we = 3, we = 5 and we = 10 when R.- =1, Re = 4, w, = O and k = 0.5. The top curve corresponds to we 2 1. The graphs of p(R) are almost identical for we = 2, we = 3, we = 5 and we = 10 and so generates an essentially common lower curve ........................... Graphs of g(R) for five values of we: we = 1, we = 2, we 2 3, we = 5 and we=10 when R,=1,Re=4,w,=0 and kzl. For we=2, we = 3, we 2 5 and we = 10 there is an apparent discontinuity at R,- = ................................... vii 13 26 31 31 3.7 Graphs of p(R) for five values of we: we = 1, we = 2, we = 3, we = 5 and we =10 when R, =1, Re =4, w,- =0 and k = 1. The top curve corresponds to we = 1. The graphs of p(R) are almost identical for we = 2, we = 3, we = 5 and we = 10 and so generates an essentially common lower curve ........................... 3.8 Graphs of g(R) for five values of we: we = 1, we = 2, we 2 3, we = 5 and we=10 when I?,-=1,Re=4,w,~=0 and k=2. For we=2, we = 3, we = 5 and we 2 10 there is an apparent discontinuity at R,- = 1 ................................... 3.9 Graphs of p(R) for five values of we: we = 1, we = 2, we = 3, we = 5 and we =10 when R, = 1, Re = 4, w,- = 0 and k = 2. The top curve corresponds to we = 1. The graphs of p(R) are almost identical for we = 2, we = 3, we = 5 and we 2 10 and so generates an essentially common lower curve ........................... 3.10 The lower curve is the graph of \Ilext(k) and the upper curve is the nu- merically estimated \IlemUc) when R,- = 1, Re = 4. Note that \IleritUc) appears to take its minimum when k = 1 ................ 3.11 Graphs of the normalized total energy, normalized F} and normalized E“ as afunction of k when R.- =1, Re=4 and we—w, =1 ....... 3.12 The deformation of an initially vertical line segment for a material with k = l for we = .75413 and we = 1.50826 = \Ilmee. These two values of we were also used in Figure 2.1 for the neo—Hookean material. Here Ri=1, Re=4 and 114:0 ....................... 3.13 Comparison of the deformation of an initially vertical line segment for a material with k = 1 to that of the neo—Hookean material (k = 0) 3.14 Graph of g(R) when k = 1 for the following ten values of we, we = 0.150826, 0.30165, 0.452475, 0.6033, 0.754125, 0.90495, 1.055775, 1.2066,1.357425, 1.50826 = \Ilmee. Here R,- = 1, Re = 4 and w,- = 0 . 3.15 Graph of the family A3(n) with R,- = 1, Re = 4, 111i: 0 and we 2 4 for n = 0 (the diagonal line), and 1,10, 100,1000,10000 .......... 3.16 Graph of the family A4(n) with R,- = 1, Re = 4, w,- = 0 and we = 4 for n = 0 (the diagonal line), and 1,10, 83.1,100, 1000, 10000. The dashed line shows the case n = 83.15 which results the minimum energy among the family A4(n) ..................... 3.17 Graph of A1 and A4(83.15) when R,- = 1, Re = 4 and we = 4 ...... 3.18 Normalized energy versus twist for R,- = 1, Re = 4, w,- = 0 and varying we. Here \I/e,ex = 1.50826. The energies corresponding to A1, A2, A3,ee, A4,ee and Any are normalized by dividing 27rhci. The nor- malized energy associated with A4Inmie] is also indicated by stars for we = 2, 2.5, 3, 3.5, 4 ............................ 4.1 Numerical solution when R, = 1, Re = 4, w,- = 0 and we = 0.452478 for three values of 61, d = 0.1,0.01,0.001. Here |we — 'lbz'l = 0.3\Ilmee, . . . 4.2 Numerical solution when R,- = 1, Re = 4, w,- = 0 and we = 1.5 for three values of 07, (‘1 = 0.1,0.01,0.001. Here |we — 31M 2 0.995\Ilmee, viii 34 35 43 45 45 63 64 66 66 91 92 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 Numerical solution when R,— — 1, Re —- 4, w,- — 0 and we- -— 3 for three values of 01,0: 01, 0.01, 0001. Here |we— 1M“ — 1. 989‘11mex Numerical solution when R,-- — 1, Re: 4 and t/Ji- — 0 and we— - 3 for three values of a, a- —— 0.1, 0.01,0.001 and three values of we, we: 0.3\Ilmee, 0.995\Ilmee, 1.989‘Ilmee .......................... One-term, two-term perturbation solutions and numeric solution with w,- = 0 and we = 0.150826 when 51 = 0.1. Here R,- = 1, Re 2 4 and \I’max = 1.50826 so that we = 0.1‘Ilmee ............... One-term, two-term perturbation solutions and numeric solution with w. = 0 and we = 1.3 when a = 0.01. Here R,- =1, Re = 4 and \Ilmee. = 1.50826 so that we 2 0.862\Ilm ................ One-term, two—term perturbation solutions and numeric solution with 1M: 0 and we = 0.97 when 6 = 0.1. Here R,- =1, Re = 4 and ‘I’max = 1.50826 so that we = 0643me ................ One-term, two-term perturbation solutions and numeric solution with wi=0andwe=l.3whend=0.1. HereR,=l,Re=4and \I’max = 1.50826 so that we = 0.862\Ilmee,. Since we > \Ilem, the two term perturbation solution involves decrease of g(R) when R = R; One—term outer solutions with R.- = 1, Re = 4, w,- = 0 and we = 3 for the following six trial values Jo, 3., = 1,2.25,4,6.25,9,12.25. Here do = 1 is on the far left and do = 12.25 is on the far right. The curves order themselves sequentially in do. The envelope of vertical tangency is shown as a dashed line ......................... One-term patched, one—term regular perturbation and numeric solutions when R,- =1, Re = 4, w,- = 0, we = 1.5 < \Ilmee z1.50826 and ft = 0.1. The patching point is at r = 1.220537 ............ One-term patched, two—term patched and numeric solutions when R,- = 1, Re = 4, w,- = 0, we = 1.5 < \Ilmee z 1.50826 and (1 = 0.1 The patching point is at r = 1.220537 for the one-term patched solution and the patching point is at r = 1.114934699 for the two-term patched solution .................................. Two—term patched and numeric solution when R,- = 1, Re = 4, w.- = 0, we = 3 > \Ilmee, z 1.50826 and 6: = 0.1. The patching point is at r = 1.2336459 ............................... One-term patched, two-term patched and numeric solution when R.- = 1, Re = 4, w,- = 0, we = 10‘Ilmex and dz = 0.01. The patching point is at r = 1.246533 for the one-term patched solution and the patching point is at r = 1.245361 for the two-term patched solution ...... Normalized energy E / (27rhd) versus twist when R,- = 1, Re = 4 for the four values of it, (t = 0,0.001,0.01, 0.1 are depicted as the curves E1, E2, E3, E4 respectively ........................ Normalized torque M / (27rhd) versus twist when R, = 1, Re = 4 for the four values of (3, d = 0,0.001,0.01, 0.1 are depicted as the curves Ml, M2, M3, M4 respectively ....................... ix 93 93 97 98 99 100 104 111 113 113 115 116 CHAPTER 1 Introduction In the event of substructural reconfiguration in solids, it is sometimes the case that the notion of simple deformation X -+ x must be broadened so as to incorporate a more detailed kinematic description. This typically involves new kinematic vari- ables in addition to x = x(X). Each new kinematic variable is in turn associated with an additional balance principle of the same tensor order as the new kinematic field variable. The continuum mechanical theory known variously as “continua with microstructure” [9] or “multi-field theory” [18] gives insight into the role of such addi- tional balance equations. Gurtin’s theory of configurational forces provides a related type of description [12]. A well-known example of such a theory is the Ericksen model of liquid crystals [10]. In the liquid crystal theory, the additional field is a (unit) vector that describes the liquid crystal orientation. A new constitutive framework for treating finite deformations when conventional elastic behavior is modified by substructural reconfiguration has recently been pro— posed by Pence and Tsai [22]. Energy minimization then leads to an additional balance principle that is associated with the richer kinematics. In this thesis, we investigate the solution of the pure azimuthal shear problem in this new framework. Azimuthal shear is most simply conceived as a twisting deformation of a circular cylinder due to an imposed difference in twist \11 between in the inner and outer radii. The constitutive model has, among other things, a natural parameter k 2 0 such that the neo—Hookean material in standard nonlinear elasticity is retrieved in the limit I: -—> 0. In Chapter 2 the solution of the pure azimuthal shear problem is summarized for a neo-Hookean material in standard nonlinear elasticity theory. The methodology and results of the pure azimuthal shear problem for a neo-Hookean material are well-known and will serve as a basis for comparison with the new results described in this thesis. The new content of the thesis is divided into two parts as presented in Chapters 3 and 4. The first part of this thesis, Chapter 3, is concerned with a material model such that the stored energy to be minimized is the sum of two terms. One term can be viewed as penalizing macroscopic elastic deformation and the strength of this penalization is described by a positive constant 6:. The second term can be viewed as penalizing elastic substructural reconfiguration and its strength is described by a positive constant a‘. The parameter k is the ratio 6/01“. In the first section of Chapter 3 the pure azimuthal shear problem in this new framework is formulated. It is confirmed that if the parameter k goes to zero, then the pure azimuthal shear problem in the new framework reduces to the standard pure azimuthal shear problem for a neo—Hookean material. Since the governing field equations are complicated for a general k, it does not seem feasible to get a closed form solution for arbitrary k. Hence, in Section 3.2, the pure azimuthal shear problem is studied for small Ii: by a perturbation expansion about the base neo-Hookean solution. The torque- twist relation is also presented for small k. An interesting aspect of the torque-twist relation as determined by the perturbation analysis is that torque is not an increasing function of twist for all values of twist if k 7e 0. The pure azimuthal shear problem is then solved numerically for any k in Section 3.3. When k ¢ 0 the numerical solution reveals evidence of a threshold value for twist at which smooth solutions are no longer available. In Section 3.4 the pure azimuthal shear problem is solved analytically in the special value k = 1 and a restriction in the form [‘11] S \Ilmex on the imposed twist \I' is obtained for a classically smooth solution. With the exception of k = 0 and k = 1 it does not seem likely that analytic solutions can be constructed explicitly. In Section 3.5 the pure azimuthal shear problem is derived by minimizing an appropriately reduced stored energy function which acknowledges the symmetries inherent in the solution. In particular, we prove that the exact solution that is derived in Section 3.4, is indeed an absolute minimum for the energy integral when k = 1. In Section 3.6 direct methods of calculus of variations are used to investigate solution possibilities for arbitrarily large twist ‘1! when k = 1. A minimizing curve is obtained among a collection of test curves for the parametric energy integral when the twist is arbitrarily large. Indeed we prove that this minimizing curve is an absolute minimum of the parametric energy integral in the class of piecewise-smooth curves such that, once restricted to the interior, their parametrization reduces to a function. The importance of such a result is that if we seek minimizers of the nonparametric energy integral for arbitrarily large twist \11 in the space of piecewise-smooth functions, then the global minimum involves a discontinuity at the inner radius. A more generalized material model is considered in the second part, Chapter 4. The strain-energy function for the new material model augments the strain-energy function of the first part by inclusion of an additional term. This additional term can be viewed as penalizing the overall deformation, as opposed to penalizing specific partial deformations. The strength of this penalization is described by a positive constant (1. Thus there are now three terms to the stored energy with respective strengths 6:, a” and a. The pure azimuthal shear problem subject to the same boundary conditions is studied in this extended material model when d = a" (k = 1) . Our interest is in the extent to which a > 0 gives different qualitative behavior from the or = 0 treatment of Chapter 3. In the first section of Chapter 4, the pure azimuthal shear problem is formulated for the extended material model when d = of. Moreover the resulting boundary value problem is expressed as an integral equation. In addition two modified boundary value problems are also defined. These second boundary value problems for the pure azimuthal shear problem are useful for obtaining existence and uniqueness results. The principle of contracting mapping is then used to prove an existence and unique- ness result for the pure azimuthal shear problem in Section 4.2. In Section 4.3 the pure azimuthal shear problem is solved numerically in the extended material model. In Section 4.4 a regular perturbation procedure is used to investigate the solution of the pure azimuthal shear problem when the material parameter a/ci is small. It is observed that if [‘11] > \Ilmex, then the regular perturbation cannot be used. Further if [\IJI > \IImax or [\III = (1 — e)\II,e,ax with 0 < 6 << 1, then severe requirements are placed on the smallness of a/d in order to obtain a useful approximation near the inner radius. In Section 4.5 a boundary layer type analysis is developed in order to get a useful approximation in such cases. In particular, inner and outer solutions are obtained and a patching method is proposed for obtaining the full solution. This patching method is found to give satisfactory agreement with the numerical solution. The relation between torque, twist and energy in the extended material model is studied in Section 4.6, which is then followed by a concluding discussion in Chapter 5. CHAPTER 2 Pure Azimuthal Shear Problem for a Neo-Hookean material 2.1 Introduction This section provides preliminary discussion from the well known conventional theory of hyperelasticity in order to provide context for this thesis. Readers familiar with the pure azimuthal shear problem for a neo—Hookean material can, if they wish, proceed directly to Chapter 3. 2.2 Pure Azimuthal Shear Problem for a Neo- Hookean material In standard nonlinear continuum mechanics it is the case that the continuity equation, the principles of linear and angular momentum, and the energy equation hold for any material regardless of its constitution. However, unless the body can be regarded as rigid, these equations are in general insufficient to determine the motion produced by given boundary conditions and body forces. They need to be supplemented by a further set of equations, known as constitutive equations, which characterize the material composition of the body. Such a set of constitutive equations usually serves to define an ideal material, and much of the work on modern continuum mechanics has been concerned with the formulation of constitutive equations to model as closely as possible the behavior of real materials. We consider a continuous body which occupies a connected open subset of a three- dimensional Euclidean point space, and we refer to such a subset as a configuration of the body. We identify a specific configuration as a reference configuration and denote this by B,.. Let points in B, be labeled by their position vectors X relative to some chosen origin and let BB, denote the boundary of 8,. Now suppose that the body is deformed quasi-statically from B, so that it occupies a new configuration, B with boundary 88. We refer to B as the current or deformed configuration of the body. The deformation is represented by the mapping X : B, —-> B which takes points X in B, to points x in B, where x is the position vector corresponding to the point X in B... The mapping x is called the deformation from B, to B. In this section we review the well known treatment of this deformation in the context of conventional continuum mechanics and hyperelasticity. In such a treatment the equilibrium equations in the absence of body forces are given by div 0' = 0, (2.1) where a' is the Cauchy stress tensor and div is the divergence operator with respect to current coordinates. The constitutive equation of an elastic material is given in the form a' = C(F), (2.2) where G is a symmetric tensor-valued function defined on the space of deformation gradients F. Here F = Gradx or in component form 6113 Ea — 8X07 (2.3) where i and a 6 {1,2,3}. In general the form of G depends on the choice of reference configuration and G is referred to as the response function of the material relative to reference configuration 8,. A material whose constitutive law has the form (2.2) is referred to as a Cauchy elastic material. From the point of view of both theory and applications a more useful concept of elasticity, which is a special case of Cauchy elasticity, is hyperelasticity (or Green elasticity). In this theory there exists a stored energy function (or strain-energy function), denoted W = W(F), defined on the space of deformation gradients such that the total stored energy is given by E = [W(F) dV. (2.4) Then for an unconstrained material 8W _ _ —1 T a — (3(F) — J 8F F , (2.5) where J = detF. The Jacobian J of the deformation gradient F determines the resulting volume change in the deformation. In this setting (2.1), (2.5) are the Euler- Lagmnge equations associated with minimization of (2.4). For a volume preserving (isochoric) deformation J = 1. (2.6) A material for which (2.6) is constrained to be true for all deformation gradients F is said to be incompressible. The modification of (2.5) appropriate for incompressibility is _ 8W where I is the identity tensor and the scalar p is an arbitrary hydrostatic pressure, that is formally the Lagrange multiplier associated with the constraint (2.6). 7 A standard model for an incompressible material is given by the neo—Hookean stored energy density W in the form W = $1! (11 - 3): (2-8) where u > 0 is the shear modulus of the material in the reference configuration, 11 = tr(B) and B = F FT, the left Cauchy-Green tensor of the deformation, and tr denotes the trace. More detailed explanation of the above concepts, including restric- tions upon W motivated by observer invariance and material symmetry requirements, can be found in any standard nonlinear elasticity book [2], [3], [11], [21]. Consider a nonlinearly elastic thick-walled circular cylindrical tube whose natural (unstressed) configuration is defined by 0 1 in this limit, which by virtue of (3.35) gives that q(r) ——> 01“. 3.2 Solution for Small k (3 << oz“) At the end of the last section it was shown that the present theory with k = 0 reduces to the problem for the neo-Hookean type material of standard nonlinear finite elasticity. In this section we will study the pure azimuthal shear problem for small It by a perturbation expansion about the base neo—Hookean solution. If k = 0, then the neo—Hookean solution applies. From equations (2.34) and (2.35) this solution is given by c 90(R) = 3% + 020, (349) and dcio P0(R) = if + C30, (3.50) where C10, C20, C30 are the constants computed from boundary conditions (2.17), (2.18) and given by (2.36), (2.37), (2.39) respectively. Note that instead of c1, c2, c3, the constants C10, c20, C30 are used respectively to emphasize that these are the constants corresponding to zeroth order solution. The subscript o is used to denote the zeroth order approximations for the unknown functions and tensor quantities. Moreover from (3.46) the base neo—Hookean value of 13 is 1 rgo' 0 B0 = rgo’ 1+ (rgo’)2 0 , (3-51) 0 O 1 21 and from the remark after (3.48) (10 = a‘. (3.52) Assume that there is a small k solution in the form of following expansions for g(R), p(R), q(R), B(R) for the given problem: g(R) = 90(R) + 23.11 k" 974R), (3-53) p(R) = p0(R) + 23:, k" MR), (3.54) q(R) = <10 + 233:1 *5" (MR). (3.55) B(R) _—. 130(3) + 23:, k" 13,,(12), (3.56) where go(R), p0(R), go and 130(R) are as given above. Note that by using the expansion (3.53) the left cauchy-green tensor, B, given by (2.12), can be written as: - 1 r(90'+kgl’+...) Ol B: r(go'+kgl’+...) 1+r2(go'+kgl’+...)2 0 , (3.57) _ O 0 1 _ or equivalently, ' 1 rgo 0 - ' 0 rgl 0 - B = rgo’ 1+ r2(go’)2 O + k rgl ’ 2r2go’gl’ 0 + . .. (3-58) _ O 0 1 _ _ O 0 0 _ giving that F 1 rgo’ 0 ' Bo = rgo’ l + r2(go ')2 0 a (3-59) _ 0 O 1 _ and " T 0 7‘91 I 0 B1 7°91, 27290,91, 0 - (3.60) 0 0 O .. .J Substituting from expansions (3.53)—(3.56) into the internal balance (3.34) gives (03+kQ1'l'...) k(I§0+kl§1+...)2+ (130+kf31+...) at —(Bo+kB1+...) =0, (3.61) or equivalently, * ‘ 2 ‘ Cll ‘ 2 (Bo — B0) + k(B0 - B1+ B1+ 5B0) + 0(k ) = 0. (3.62) The 0(1) terms of (3.62) give 130 = BO, (3.63) which was already evident from (3.51) and (3.59), while the 0(k) terms give 133 + g 130 + 131 — B1 = 0. (3.64) Therefore 131 = B1 — 1‘33 — g 130. (3.65) By using (3.59), (3.60) and (3.63) -1-(ql/a*) - ”(9032 (904; 0 f3. = ... (13,)... 0 , (3-66) _ ‘1 — (QI/O‘) where (Bar. = my — 27‘90’ — #6033 — g-mg (3.67) 23 (B1).» = 212903. ' — 1 — 37360)? — r4(go'>4 — £3; —- gear (3.68) and the other components of 131 follow from the symmetry of 131. To find the complete solution up to the order k , it is necessary to find ql, 91 and p1. If the com- ponents of I30 and 131 , given by (3.59) and (3.66)—(3.68) respectively, are substituted into (3.23), then it follows that 391 1p 1+ k( — 3 —— - r1190 )2) + 0(k2) = 1, (3.69) from which it is immediate that t (3 + ”(90 ')2)~ (3-70) (11: If go ’ (r) is computed from (3.49) and used in (3.70), then ql is found as 4% (11: _a. (1 _,_ W) (3.71) where cm is the constant given by (2.36). By using the reduced equilibrium equations (3.32) and (3.33) it follows that 91 and p1 are given by 83 20 —c _ 10 + 10 11 — 9—73 2T2 + C21, (3-72) 91 and = (1101120 — 3C10C11) & + 4C‘i'0é pl 37‘4 T + C31. (3.73) The constants cu, C21, C31 are determined by requiring gl(R,v) = 91(Ro) = 0 and 0,,(R0) = 0,9,. Hence 16 (R? + R3123 + R3) at. 011 = 2 C10 + 9 R? R3 1 (3-74) 8 R? + RS‘ c3 C2. = ( 9 R, R; 10, (3.75) _4 ,3 _ , * C31 :: C10[ ('10 + R303 C11 19C10)](10 (376) 312:; 24 The azimuthal component of the traction on the inner and outer surfaces are respec— tively given by A A a (I 09,.(Ri) = QC“)? — [CCU-RT? + 0(k2), (3.77) 67 Cl 06r(Ro) = —2010§2' + ICC”? + 0(k'2). (3.78) O 0 By using (3.78) the Torque can be computed from 27r M = h / 123094123719. (3.79) 0 Then Torque- Twist relation is given by 47rhéR3R§ M: 123-123 11,- 4 7r h d R? R3 [9(R3 — R3)2 + 8(R;‘ + R? R3 + R3) \112] 1: RE - R? 9 (R? - REV \II + 00:2). (3.80) Note from equation (3.80) that M (-\II) = —M(\IJ) by the terms thus far computed. The torque can be normalized by dividing 4 7r h d R? R3. The normalized Torque- Twist relation up to 0(k) as given by (3.80) is displayed in Figure 3.1. Here R,- = 1, R0 = 4 and the solid line shows the neo—Hookean case k = 0. The dashed line shows the case k = 0.02. Figure 3.1 shows that increasing k = d/a" results in a relative softening. Perhaps the most interesting aspect of (3.80) is that, to within the terms thus far computed, M is not an increasing function of \I’ when k > 0 for all \II. By using (3.80) it is immediate that dM ”magma—1) W = (R? — R2) (3'81) +32rhc‘1RfRfi(R;-‘+R?R§+R3)k\ll2 3(R? _ R2)3 + 0(k2). 25 53.0.0.0 aroma Normalized Torque O -0.1 -0.2 -0.3 l 0‘15 0 5 Amount of Twist Figure 3.1: Comparison of normalized torque-twist relation when k = 0 and k = 0.02 for R, = l and R0 = 4. The dashed line shows the case 16 = 0.02. The normalized torque is M/(4 7r h 67 R2123) On the basis of the terms thus far computed one can solve dM/dIII = 0 for ‘11 and hence M takes its extremum values at \/§(R§ - REM/1 - k \II = :t‘IJext(k), \Ilcxt(k) :2 fiJMR? + 12ng + R3). (3.82) Also note from (3.82) that lift 11....(k) = 00, (3.83) and that % _ V3023 ‘ R?) (3.84) dk — _ J33 \/1'—""'E k (fun? + R3123 + 123,) Restricting attention to small It, and in particular taking 0 < k < 1 it follows that dWext/dk < 0. It will indeed be shown analytically in Section 3.4 that loss of resistance occurs for k = 1 at sufficiently large ‘11. On the other hand, this thesis 26 will show that the subsequent decrease in M with \11 for \II > \Ilext is not found analytically. Presumably this decrease is an artifact of the expansion procedure that would be corrected by invoking a \11 dependent criterion for determination of the number of terms necessary for a uniformly valid approximation. These issues will be clarified in the following development. 3.3 Numerical Solution In this section we present a numerical solution procedure for the pure azimuthal shear problem which does not require an assumption of small k. The solution reveals evidence of a threshold value for \11 at which smooth solutions no longer are available. We begin with some additional reduction of the governing equations. Post- multiplying (3.34) with 13—1 gives H3 + p1 = B 13—1. (3.85) Since the product of two symmetric matrices need not be symmetric, there are five nontrivial components of (3.85) as follows k B... + p = 3.41%.. — mm], (3.86) k 3.. = Baby's... — 19.9], (3.87) k 3.. = Bums... — (1 + 99933.9], (3.88) 4 Bee + p = B..[(1 + (r9933... — 193.9], (3.89) k 13.. + p = .i. (3.90) B... 27 The values 13,, and 1399 can now be obtained from (3.87) and (3.88) respectively in terms of 13.9 and 13”, providing . k )3. B,, = [ , +1] f, (3.91) 82.2 7‘9 . k B 9 z A 1 I 2] __T_. . Baa [Bzz + + (79 ) rg' (3 92) Note, if B" as determined from (3.91) and Bag as determined from (3.92) are both substituted into (3.86) and (3.89) respectively, then the resulting equations are equivalent. Therefore only four of the five equations in (3.86)—(3.90) are independent. In addition Bu can be obtained uniquely from (3.90) in terms of p by using the positive-definiteness of B; A _ 2 B2,: M V” “L4,“. (3.93) 2k The constraint (3.23) yields . . . 1 B3,, = B... B99 — B . (3.94) Taken together (3.91)—-(3.94) indicates that Bra can be found as a function of g’ and p. Indeed it is given by Mr 9’ \/2k2(r2g’2 + 2) + (k3 — 1)p + (/4k + p2(1+ k3) By using (3.93), (3.95) in (3.91), (3.92) the nontrivial components of 13 can be found {3.0 = (3.95) in terms of g’ and p. In principal (3.86), or equivalently (3.89), now supply an equation relating p and g’ . To get a numerical solution observe first from (3.32) and (3.95) that 2k1‘2g' 2 d? 2k2(7‘2g’2 + 2) + (k3 —1)p+ W(1 + k3) _ 57 _ Secondly upon substituting from (3.91), (3.92), (3.93), (395) into (3.86) and manip- (3.96) ulating the resulting equation it is found that 193p6 — (k3 —1)k?(1"2g'2 + 3)p5 + [1 — (r29'2 +1)ic:3 + l~:6](r2g'2 + 3)kp4+ 28 (1 + 5k3 — 5k6 — k9)p3 + 209- 1) 218(ng ’2 + 3));2 + (k3 — 1)4 = 0. (3.97) Let k be fixed and choose a trial value for (11. Then at a given radial location 7‘ = r, the equations (3.96) and (3.97) can be solved simultaneously for p, E p(r,) and gj E g’(1'J-). Once g;- is known the tangent line approximation can be used to estimate g(r) at a nearby point 1' provided that g(rj) is also known. This is the strategy we will follow. At issue is the proper value of d1. Divide the interval [R4, R0] into N equal subintervals and label the end points as R,- = r0, r1, . . . ,rN = R0. One of the boundary conditions in (2.17) can be used as starting point, say g(ro) = 2,0,. Then by the tangent line approximation 901) % 9(7‘0) + (9'00) AT), (398) where Ar = (R0 — R,)/N . In general is g,- = g(r ~ g( m) + g’(rl) Ar), (3.99) J l=0( for j = 1,2,... ,N. Since g(RO) = 1120, it is therefore necessary to choose d1 so that = Z(g'(n) Ar). (3.100) (=0 That is, d1 can be used as a shooting parameter. Using this numerical scheme it is found that all and the overall solution curves are robust with respect to finer and finer discretization (increase in N). However it is found that the R = R, end point value of g’ ; that is, g’(R,-), is very sensitive to change in N when 2110 is sufficiently large. For example take k = 0.1, R,- = 1, R0 = 4, and «[2,- = 0. Then for 1P0 = 2 it is found that d1 = 2.84365 for N = 300 and that d1 = 2.86951 for N = 3000. The associated numerical values of g’(R) at R,- = 1 are found to be 9’ (1) = 8.54856 and g’ (1) = 9.05256 respectively. On the other hand for ’l/Jo = 10 it is found that (11 = 3.1621814432 for N = 300 and that (11 = 3.162276607355 for N = 3000. The associated numerical value of g’(R) 29 at R,- = 1 are found to be g’(1) = 736.36329 and g’ (1) = 7101.95319 respectively. We also find for we = 2 that the numerical value of the derivatives at R = 1.01 are given by g’(1.01) = 7.56906 and g’ (1.01) = 7.94672 respectively for N = 300 and N = 3000. For we = 10 it is found that g’(1.01) = 23.48574 and g’(1.01) = 23.50641 respectively for N = 300 and N = 3000. Therefore only g’(R,-) is very sensitive to change in N when we is sufficiently large. In general, if j 75 0, then g; values are quite stable to increase in N even for large we values. In particular for sufficiently large we , the numerically computed value of g’(R.,-) increases without apparent bound as N gets larger. We speculate that this is numerical evidence for a lack of smooth solutions at R = R, when we is sufficiently large. For k = 0.1 Figure 3.2 shows the graph of g(R) for five values of we: we = 1, we=2,we=3,we=5andwe=10whenRi=1,Re=4,w,-=0from the numerical procedure outlined above. For we = 5 and we 2 10 the numerical procedure does not converge to a finite value for g’(R,-) and the associated graphs for g(R) show an apparent discontinuity at R = R,. Figure 3.3 shows the corresponding p(R) for the same parameter values. The graphs of p(R) are ordered from top to bottom for we = 1, we 2 2 and we = 3 and they are almost identical for we = 3, we: 5 and we = 10. For we = 1 and we = 2 we find that p(Re) = 0.89962 and p(Re) = 0.89903 respectively. For we = 3,5, 10 we find that p(Re) = 0.89883. However p(R1-) = 0.790, 0.490, 0064,0005, 0.001 for we = 1,2,3, 5,10 respectively. For k = 0.5 Figure 3.4 shows the graph of g(R) for five values of we: we = 1, we=2, we=3, we=5 and we=10 when R,=1, Re=4, w,=0. For we=2, we 2 3, we = 5 and we = 10 the numerical procedure does not converge to a finite value for g’(R,.-) and the associated graphs for g(R) show an apparent discontinuity at R = R,. Figure 3.5 shows the corresponding p(R) for the same parameter values. The top curve corresponds to we 2 1. The graphs of p(R) are almost identical for 30 Figure 3.2: Graphs of g(R) for five values of we: we = 1, we = 2, we = 3, we = 5 and we=10 when R,=1, Re=4, ¢i=0 and 19:01. For we=5 and we==10 there is an apparent discontinuity at R,- = 1 p(R) Figure 3.3: Graphs of p(R) for five values of we: we = 1, we = 2, we = 3, we = 5 and we =10 when R,- =1, Re = 4, w,- = 0 and k = 0.1. The top curve corresponds to we = 1 and the middle one corresponds to we = 2. The graphs of p(R) are almost identical for we = 3, we = 5 and we = 10 and so generates an essentially common lower curve 31 9(R) Figure 3.4: Graphs of g(R) for five values of we: we 2 1, we 2 2, we = 3, we = 5 andwe=10whenR,-=1, Re=4,wi=0andk=0.5. For we:2, we=3, we = 5 and we = 10 there is an apparent discontinuity at R,- = 1 we = 2, we = 3, we = 5 and we 2 10 and so generates an essentially common lower curve. For we = 2,3,5,10 we find that p(Re) = 0.49934 and for we = 1 we find that p(Re) = 0.49955. However p(R,-) = 0.3252,0.0042,0.0012,0.0005, 0.0002 for we = 1, 2, 3, 5, 10 respectively. For k = 1 Figure 3.6 shows the graph of g(R) for five values of we: we = 1, we = 2, we = 3, we = 5 and we =10 when R,- =1, Re 2 4, w,- = 0. Theassociated graphs for g(R) again show an apparent discontinuity at R = R,- when we = 2, we = 3, we = 5 and we = 10. Figure 3.7 shows the corresponding p(R) for the same parameter values. Note from the scale that this numerical solution gives 0 < p(R) << 1 for R > R,- = 1. The top curve corresponds to we 2 1. The graphs of p(R) are almost identical for we = 2, we = 3, we = 5 and we = 10 and so generates an essentially common lower curve. For each we we find that p(Re) = 0.000422356. However p(R,) = 0.0001, 0.000001,0.0000003,0.0000001,0.000000056 for we 2 1, 2,3, 5, 10 respectively. 32 p(R) 0.1 -------- -------- -------- é -------- 1 -------- é ------- « Figure 3.5: Graphs of p(R) for five values of we: we 2 1, we = 2, we 2 3, we = 5 and we = 10 when R, = 1, Re = 4, w,- = 0 and k = 0.5. The top curve corresponds to we = 1. The graphs of p(R) are almost identical for we = 2, we = 3, we = 5 and we = 10 and so generates an essentially common lower curve 1o . . , , . g(R) Figure 3.6: Graphs of g(R) for five values of we: we = 1, we = 2, we = 3, we = 5 and we=10 when R.-=1, Re=4, w,=0 and 1:21. For we=2, we=3, we=5 and we 2 10 there is an apparent discontinuity at R,- = 1 33 x10 p(R) Figure 3.7: Graphs of p(R) for five values of we: we = l, we = 2, we = 3, we = 5 and we 2 10 when R,- = 1, Re = 4, w, = 0 and k = 1. The top curve corresponds to we 2 1. The graphs of p(R) are almost identical for we = 2, we = 3, we = 5 and lite = 10 and so generates an essentially common lower curve For k. = 2 Figure 3.8 shows the graph of g(R) for five values of we: we 2 1, we 2 2, Lie 2 3, we 2 5 and we 2 10 when R,- = 1, Re = 4, w,- = O. The associated graphs for g(R) again show an apparent discontinuity at R = R,- when we = 2, we = 3, we = 5 and we = 10. Figure 3.9 shows the corresponding p(R) for the same parameter values. The lower curve corresponds to we 2 1 and graphs of p(R) are almost identical for we 2 2, we = 3. we 2 5 and 'ci've = 10 and so generates an essentially common top curve. For are = 2.3.5.10 we find that p(Re) R3 —0.9987 and p(Re) = -—0.9991 when we 2 l. H(;)we\'er p(R) = —0.6505,—0.0080.—0.0025,—0.00ll,—0.00IO for Le = l. 2. .3. 5 10 respectively: It is tempting to seek a correlation between the issue of an apparent discontinuity at R = R, for sufficiently large L-‘e and the previously obtained \I/ex,(k) in Section 32 This gives rise to the notion of a smooth solution region in (we. A‘)-space for .34 g(R) Figure 3.8: Graphs of g(R) for five values of we: we 2 1, we = 2, we = 3, we = 5 and we=10 when R,=1, Re=4, ¢i=0 and k=2. For we=2, we=3, we=5 and we = 10 there is an apparent discontinuity at R,- = 1 p(R) Figure 3.9: Graphs of p(R) for five values of we: we = 1, we = 2, we = 3, we 2 5 and we = 10 when R.- = 1, Re = 4, w, = 0 and k = 2. The top curve corresponds to we = 1. The graphs of p(R) are almost identical for we = 2, we = 3, we = 5 and we = 10 and so generates an essentially common lower curve 35 fixed R,, Re, w,. Using R,- = 1 and Re = 4 in (3.82) gives q, (k)._15\/3\/1-—k ext -— fim . Let w, = 0. Then the curve (3.101) can be used to estimate the boundary for (3.101) the proposed smooth solution region in (we,k)—space. On the basis of following the argument it is found that smooth solution region is in fact somewhat larger than that predicted by (3.101). Based on the numerical solutions presented above it can be assumed for a given k that there is an \Ilem(k) for which g’(R,-) —> oo. Equivalently there is a critical value of shooting parameter d1, say de,,t(k), for which g’(R,-) —* 00. Indeed it can be seen from (3.96) that R? dcrit (k) = «I; - Then for a given Re and N equation (3.96) with (11 = dem in conjunction with (3.102) and (3.97) can be solved simultaneously for pj E p(r,) and g]. E g’(rJ-) as before. In principle \Ilerit(k) can now be approximated by (3.100). However the sensitivity of g’(R.-) to N gives difficulty. Since p,- (j = 0,...,N) has no such sensitivity, an alternative is to find a good fitting polynomial for p(r). Once a closed form approximating function for p(r) is available, equation (3.96) can be solved for g’ (7‘) giving \/4 1.23% + (k3 — 1) d? p(r) + (1 + k3) d; ./4"7c + p(_r)2 I :2}: gm ,/2kr6—2k2d§r2 (3.103) Note that since we > w,- it is the positive root of (3.103) that is of interest. If the positive root of (3.103) is integrated numerically from R,- to Re with d1 = cm, one then obtains an estimate for \IlemUc). By using the outlined method, the maximum amount of twist versus 1: is depicted in Figure 3.10. Note that \IIemUc) apparently takes its minimum when k = 1. The graph of \Ilext(k) is also depicted in the same figure. Note that \IlextUc) is defined if 36 010) Maximum Twist A N 0’ I Figure 3.10: The lower curve is the graph of \IJextUc) and the upper curve is the numerically estimated \Ilerit(k) when R,- = 1, Re = 4. Note that \IlemUc) appears to take its minimum when k = 1 k E (0, 1] and that it gives a reasonable estimate for the critical amount of twist only if k is sufficiently small. For k = 0.1, R,- = 1, Re = 4, w,- = 0, we find that \Ilerit(0.1) = 3.030996. Note in comparison that ‘Ilext(0.1) = 1.667811 < ‘Ilem(0.1). While (3.101) could likely be improved by computing more terms in the perturbation expansion, the central issue of an apparent threshold angle we remains. Namely, both the perturbation expansion and the numerical solution procedure give evidence that there is a maximum amount of twist above which there is no smooth solution for the pure azimuthal shear problem under consideration. With respect to the later development of Section 6, it is useful to remark that the numerical estimate for \Ilem(1) is 1.50826 when R = 1, Re = 4. Note that the energy density (3.6) is separable in i1 and If. Hence the potential 37 2.5 2 5 51.5 9, a g 3 1 I E : O 1 Z : 0.5 ; . l 1 0o 05 1 15 2 1: Figure 3.11: Graphs of the normalized total energy, normalized E and normalized E“ asafunctionofkwhenR,=1, Re=4 and we—w,=1 energy functional (3.5) can be decomposed into two parts E=E+E2 where A E=/%(i1—3)dX is the energy corresponding to elastic part of the deformation and E‘ =/“7(1;—3) dX is the energy corresponding to pre-elastic part of the deformation. (3.104) (3.105) (3.106) Let we — w,- be fixed. For given Ri, Re and k the numerical solution presented in this section can be computed numerically to find (3.105) and (3.106). This numerical computation reveals the following relation between E and E‘ for a fixed amount of twist. If 0 < k < 1 then R > E" and if k > 1 then E < E‘. Moreover 1.3} = E“ when k = 1 meaning that the energies corresponding to elastic and pre-elastic part of the deformation are equal. 38 For we — w,- = 1, Figure 3.11 shows the numerically estimated graphs of the normalized total energy, normalized E and normalized E“ as a function of I: when R,- = 1, Re = 4. Energies are normalized by dividing whd. 3.4 Explicit Solution when (31 = a* (k = 1) In this section we will show for k = 1 that the governing field equations for azimuthal shear can be solved explicitly up to integration constants. Then using the boundary conditions we will show that the pure azimuthal shear problem for k = 1 has a unique solution provided that the twist \II is not too large thereby verifying the notion of a threshold twist as described above. Recall from Section 3.3 that (3.97) supplies an equation relating p and g’ . However it is only for k = 1 that ensuing algebraic simplifications as given next will take place. That k = 1 is special in this regard could have been anticipated by the discussions in Pence and Tsai. If (3.97) is used with k = 1, then p is a root of the equation 64 p4 [102 - (019')4 + 20932 - 3)] = 0- (3107) Hence p = 0, :l:\/(rg’)4 + 2(rg’)2 — 3. (3.108) However if components of B are computed from (3.91), (3.92), (3.93), (3.95) by using either p = i\/(rg’)4 + 2(1‘g’)2 — 3, then one finds that the component equations (3.36), (3.37), (3.38) of the internal balance equation (3.34) are not satisfied. Thus thase roots are spurious and p = 0 is the only solution possibility for the internal balance equation (3.34). This is consistent with the uniqueness result for the solution of internal balance. For p = 0, it follows from (3.91), (3.92), (3.93), (3.95) that I B — ——T—g——— (3.109) To _ 4 + (rg’)2’ 39 B... = ——-——., (3.110) 4 + (79)) A I 2 oo = m) (3.111) W B... =1, (3.112) and it is verified that (3.34) holds. Now equations (3.109) and (3.32) give I d ——’3—— = --§, (3.113) 4 + (79’)2 7" which in turn provides 2d1 ' = —————. 3.114 9 T T, _ d‘i ( ) After integration and replacing r by R it is found that 4 - d2 g(R) = arctan R d2 1 + d2, (3.115) 1 where d2 is another constant. By applying the boundary conditions (2.17) and elim- inating d2 it follows that R4—d2 R‘.‘—d2 ° 1—arctan ' 1 d1 = (w. - 10.). (3-115) arctan 5.3 Equation (3.116) can be solved for (11 if and only if the twist \II as given by (2.44) obeys R2 |\IJ| S \Ilmex, \Ilmex = arccos(F’2). (3.117) By restricting the boundary values as indicated in equation (3.117) the constants d1 and d2 are found to be R-2R2 sin‘II d = ' ° , 3.118 1 V/R;1 + Rf, — 2R,.2R§ cos ‘11 ( ) 124—3? d? (3.119) d2 = w,- — arctan 40 Note that cos \I'mex = Rf/Rg implies sin \Ilmex = «Rf, — Rf/Rg whereupon the asso- ciated values of d1 and (12 are (d1)ma.r : R321 (d2)max : 1192'- (3.120) In particular (3.120)1 shows that (d1)mee = dem(1) as given by (3.102). In addition \Ilmex = arccos(1/16) = 1.50826 when R,- = 1, Re = 4 which is consistent with \Ilem(1) as given at the end of Section 3.3. Substituting for B", R99, 9’ from (3.110), (3.111), (3.114), into the equilibrium equation (3.33) gives d . a d —(r ’)2 211: = 05(71271‘9/7) + ;(W). (3.121) This equation simplifies to 2’2_ dr _ 0. (3.122) Using (3.122) and replacing r by R gives p(R) = d3, (3.123) where d3 is another constant. This constant can be computed from the boundary condition (2.18) with the help of (3.24): d (31 3 = fi 123— ct? — 00 (3.124) rr? where d1 is the constant given by (3.118). Since p = q/a“ and p = 0 it is immediate that q = 0- (3.125) Thus it has been shown that if k = 1 and \Il obeys (3.117) then B is given from (3.22), (3.109)-(3.112), g is given by (3.115), p is given by (3.123) and q is given by (3.125), where the constants d1, d2, d3 follow from (3.118), (3.119), (3.124). Note that (3.125) is consistent with the k = 1 numerical solution given in Section 3.3. 41 We remark that (3.114), (3119) give 2R2R2 ' \II g'(R) = = 08‘“ . (3.126) R\/R4(R‘,-‘ + Rf, — 2133123 cos\II) — RfRf, sin2 \1' Thus |g’(R)| is decreasing in R with , 2R3, sin \II , 2R22 sin \II 9 (Re) — RARE _ Re cos\II)’ (3.128) Observe also that 9,031) —* 00 as ‘1’ -+ ‘I’max- (3.129) The normal traction on the inner surface R = R,- is & 61 Urr(R’i) : 'fii V R? _ (if _ fl V R: _ df + 0197-1 (3°130) and we observe again that an.(R.-) < 09,. By using (3.24) and (3.32) the shear traction on the inner and outer surfaces are 0,0(Re) = 61—. (3.131) The Torque-Twist relations follows from (2.44), (3.118), (3.131) and is given by M = Me = —M.- where _ 27rhc‘erRg sin\II \/R§+R3—2R,?R§cos\ll’ M g 111...... (3.132) Note that g—Agfiwm) = 0. (3-133) Thus \Ilmex is similar to Wen as discussed in Section 3.2. More generally, the notion of a threshold value for \II has been verified. To compute the total stored energy E first note from (3.22) that, i1 = 3.. + Bag + 13... (3.134) 42 .3- 4 44 -2 o 2 4 X Figure 3.12: The deformation of an initially vertical line segment for a material with k = 1 for we = .75413 and we = 1.50826 = \Ilme... These two values of we were also used in Figure 2.1 for the neo—Hookean material. Here R,- = 1, Re = 4 and w.- = 0 By using If = tr B“ = tr C“ , where C" 2' (F‘)T F" is the right Cauchy-Green tensor of the pre—elastic part of the deformation and F’ = F4 F, one finds (3* = [fi‘1 F]T [13“1 F] = FT 3'1 F. (3.135) Then by (2.11) and (3.22) it follows that I 2 A _ ’ A A (1 + (rg )A)B,:, 27:9 Br9 + 390 + .1 . (3.136) Brr899 — 830 B 2‘ At this point an interesting observation following from (3.109)—(3.112), (3.134), 1;: (3.136) is that the values of i1 and If are equal for k = 1. This common value is given by 1“1 = I; = ,/4 + (rg’)2 +1. (3.137) The total stored energy E for this deformation follows from (3.5), (3.6), (3.136) as R0 E: nh(d+a’)/ (1/4-11- (rg’)2 —2)rd7‘. (3.138) R. 43 Evaluating this integral with the aid of (3.114), (3.118) and using a“ = (it (since k = ci/a‘ = 1) gives E=27rhd(\/R§+Rg—2R3Rgcos\II+Rf—R§). (3.139) Note as a result of (3.137) that the energies corresponding to elastic and pre—elastic part of the deformation are equal for k = 1; i.e., /%(If —3)dX = [SJ—(1} —3)dX '1‘? (3140) The work of the torque M follows from (2.48) and (3.133) as v W=27rhd(\/R3+Rg—2R§R§cos\II+Rf—Rg). (3.141) By direct comparison of (3.139) and (3.141) it is again found that W = E. (3.142) We conclude this section by sketching three graphs for the deformation when k = 1. The first is analogous to Figure 2.1 which we recall corresponds to the neo—Hookean material (k = 0). Consider again the deformation of a line segment occupying {(R,9)|1 S R g 4, 6 = 7r/2} in the cross-section R,- = 1, Re = 4 in the reference configuration. It now follows from (3.117) that the solution (3.115) holds only if |\IJ| S ‘Ilmex = arccos(ilg) 2 1.50826. Suppose w,- = 0 so that deformation is governed by we. Figure 3.12 shows the deformed location of the line segment for two values we: we 2 0.5‘Ilmex = 075413 and we = \Ilmee. = 1.50826. Observe that 9’(R.-) = 00 for 111.. = \Ilmax. For we = .75413 and we = 1.50826 = \I’mex Figure 3.13 compares the defor- mation of an initially vertical line segment for a material with k = 1 to that of the neo-Hookean material (19 = 0). Since it is possible to twist the neo—Hookean material beyond \Ilmex , the deformation of the initially vertical line segment for the neo—Hookean material is also depicted for two more values of we > \Ilmex. 44 Figure 3.13: Comparison of the deformation of an initially vertical line segment for a material with k = 1 to that of the neo-Hookean material (k = 0) 1.6 . . e. . . 1.4 --------- r -------- r ...... : ........ . ........ . ....... - 1.2 ........................................ i 1 ........ .............. 1 ........... 1 ....... _ 5.50.8 - -- E- ----------------- : ------- - 0.6» - ......... 0.4, - L ------ .1--- ------- i -------- 1----—.. 0.2 ----- ........ ........ . ....... q 0 " 1 1 i 1 1 1 15 2 2: 3 35 4 Figure 3.14: Graph of g(R) when k = 1 for the following ten val- ues of we, we = 0.150826,0.30165,0.452475,0.6033,0.754125,0.90495,1.055775, 1.2066,1.357425, 1.50826 = \Pm. Here R,- = 1, Re = 4 and w.- = 0 45 Figure 3.14 shows the graph of g(R) for ten values of we: we 2 \I/mex = 1.50826, we = 0.9\IImax = 1.357425, we = 0.8\Ilmax = 1.2066,. . ., we 2 0.1\Ilmax = 0.150826. It is again observed that g’(R,-) -—1 00 as we —+ \I/mex. 3.5 Alternative Derivation from Energy In this section it will be shown that the formulation for the pure azimuthal shear problem can be obtained directly by minimizing an appropriate energy functional that acknowledges the a—priori symmetries of azimuthal shear. This will serve as a basis for later developments and in particular will clarify the interpretation of the results of the previous section for the case w > \Ilmex. Consider the deformation given by (2.10) for a material whose energy function is given by (3.6) and subject to constraint (3.7). Building the constraint into the minimization with Lagrange multiplier 'r gives the functional H = H(k) = 0*] g(r; -— 3) +311 — 3) — T(det 1‘3 — 1) dV. (3.143) 9 . A A A A Upon use of the symmetries g = g(r), 1' = T(r), B" = Br,(r), Bra = 8,9(7‘), Egg = 1399(7‘), E... = E... (r), the functional H becomes . 3° 1 . k 4 . H = 27rha / [—(I1 — 3) + —(11 — 3) — 7(det B -— 1)] rdr, (3.144) R, 2 2 with f1 and I f given by (3.134) and (3.136) respectively. Thus (3.144) may be written as H03 3m 81:01 Bee, 3221 73919,) = Re A A A A 7rho£t A A(r, B", Bro, B99, B..,T,g,g’)dr, (3.145) where A : [k(Brr + BOO + Bzz _' 3) _ 2’l—(BzzBrr BOO — 3221330 —' 1) 46 1 '28..—2 '8. 8 1 +(( +(1‘9)) TAQ 9+ 90+ _3)],._ . . . 3.146 BrrBBQ — 372-9 822 ( ) The Euler-Lagrange equations for minimization of (3.145) are 6A _ 6A _ 6A _ 6A _ 0 aBrr 7 837-9 , 8890 ’ 6322 , 8A d 8A 8A _=0,___.._=0, 3.147 ()7 dr 89’) 8g ( ) The first five equations in (3.147) give k + 332(27'9’31—9390 - Ego _ 839 - (T9,)2B39 — 2TBZZBQQ : 0, (3.148) A 8.. [((rg')2 + 1)8.. 8.. + 8.. 8.. — r g' (83. + 8.. 8.9)] (3.149) + 2 T 8.9 = 0, k + 834898.83 — BE. — BE. — 099283.) — 2TB..8.. = 0. (3.150) I: 83, —- 278.. — 1 = 0, (3.151) 8.. (8.. 89., — 83.) — 1 = 0. (3.152) The last equation in (3.147) admits an immediate first integral by virtue of 6/1/89 = 0, the result being ‘ 3 I ‘ 2 ‘ J B..(r g B... — r Bra) = E, (3.153) where d is an integration constant. Note that equation (3.152) is equivalent in form to (3.7). Moreover, if p = —2r, then equations (3.39) and (3.151) are also equivalent. It remains to show that (3.148), (3.149), (3.150), (3.153) provides the same completions as (3.36), (3.37), (3.38), (3.39) for suitably chosen d1 and (I. Although we do not 47 have the proof of equivalency of these two systems for general [6, such equivalency can be demonstrated for both 19 = 0 and k = 1. Let us first consider the neo-Hookean limit case k = 0. By taking k = 0 and p = —27', equations (3.148), (3.149), (3.150) reduce to 832(27‘9'31—0300 — B30 — BEG — (T9,)2830) + P322369 = 0, (3-154) A Bzz[((7'g’)2 + ”En-Bra + 3993,49 — Tg’(éfg + Brréggfl - pBrg = 0, (3.155) 8342.48-89 — BE. — Bf. — (rg’12BE.)+ p8..8.. = o, (3.156) and (3.153) reduces to r3g'Err — r2E.g = 0. (3.157) Moreover (3.151) gives 8.. = —. (3.158) Now substituting from (3.158) into (3.154)1-(3.156) gives after some manipulation that p2399 = [Tg’Bro — 39912 + 839, (3.159) P2316 = [(0932 + 1)BrrBr9 + BOOBrO — T9'(Bfo + BwBoall, (3-160) pZErr = [rg’Err —- Erg]2 + 133,. (3.161) Now, (3.159) and (3161) are quadratic equations for Egg and B... respectively and using the positive—definiteness of B can be solved uniquely in terms of Br6 and p. After substituting Egg and B... in (3.160), the resulting equation is (2 Bro — TQ’PZ) \/p4 + 431'!) (T 9,192 — BT59) 2-1-21'2(g’)2 — (3.162) 48 Tg’lp4 + 43w (7.9, P2 '— BN9” 2 + 2 7'2 (g’)2 :0, which can be solved for Bra in terms of p. By using the resulting expressions for A B", Bro, B99 as functions of p in (3.152) one obtains p3 +Px/P4 (1 +2139”) 2 + 2139’? = 1, (3.163) which has root p = 1, (3.164) and so retrieves the neo—Hookean limit (viz.(3.45).) If other roots of (3.163) are used to find the nontrivial components of B, it can be shown that the internal balance equation (3.34) is not satisfied. This is consistent with the uniqueness result for the solution of internal balance equation (3.34). It then follows that (3.148)—-(3.l53) again lead to (3.46). If (3.46) is used in (3.157), then (3.157) reduces to an identity. If (3.46) is used in (3.145) with k = 0, then H is a functional of 9’ only. Indeed it is the same energy integral (2.4) as given in the standard neo—Hookean case. By solving the Euler-Lagrange equations associated with this integral, one obtains g(R) given by (2.34). This shows that the solution of the Euler-Lagrange equations associated with vanishing first variation to the functional (3.144) for k = O indeed retrieves the k = 0 neo—Hookean result. Next we will consider the case k = 1 which corresponds to a = 01*. By taking k = 1 and p = —27', equations (3.148), (3.149), (3.150) reduce to 1 + 332(27‘9'Br9396 — 5’39 - 339 - (T9,)2BEB) + szzBBG = 0, (3.165) B..[((rg')2 + Diana, + 6998., — rg'aéfg + 6.309)] — p119 = o, (3.166) 1+ B:2(27"9’37'1‘891'9 _ BEG — Bgr _ (T9,)283r) + péZVZBTT : 0’ (3'167) 49 and (3.153) does not change. An explicitly constructive solution procedure for obtain- ing p from (3.165)—(3.167) is not obvious. However it follows by direct verification that a solution of (3.165)—(3.167) is given by p = 0 in conjunction with components of B given by (3.109)—(3.112). If (3.109)—(3.112) are substituted into (3.153), then resulting expression reduces to an identity provided that cl = 0. This shows that the Euler-Lagrange equations associated with the functional (3.144) for k = 1 retrieves the k = 1 result for the internally balanced elastic material model. By using 16 = 1 with components of B given by (3.109)—(3.112) in (3.145) we obtain 12.. H1 2 27rhé /R.- (rm — 27‘) (11". (3.168) Here the subscript 1 in H1 is notation to indicate that k = 1. It therefore follows for k = 1 that if |\II| S \Ilmax then g(r) as given by (3.115) is an extremal of (3.168) obeying (2.17). Indeed (3.113) is the first integral of the Euler-Lagrange equation for (3.168). The formal theory of calculus of variations can now be used to prove that g(r) given by (3.115) is in fact a strong local minimum of (3.168) if |\II| S \Ilmax (See Appendix A). Thus minimization of (3.168) is central to the k = 1 theory, suggesting that (3.168) can be used to investigate solution possibilities for |\II! > \Ilmax. This will be the focus of the next section using direct methods of the calculus of variations. We will close this section by proving that g(r) given by (3.115) is indeed an absolute minimum of (3.168) if Nil 3 \Ilmax. We proceed on the basis of a convexity argument using standard procedures of the calculus .of variations as reviewed next. Consider the basic problem of calculus of variations: Find a function y(:r) on the interval (a, b] that minimizes the definite integral b lel = / F($,y($),y’(r)) dx, (3.169) 50 subject to y(a) = A, y(b) = B. (3.170) In general, the minimum value of J [y] depends on the eligible functions specified by the problem. Hence, the basic problem is stated more precisely as to find 1;};ng | W) = A, W?) = B}, (3-171) where S is the collection of eligible functions specified by the problem. Any g(x) E S will be called as a test function. Given that y’ appears in the integrand F (m,y(:c),y’ (123)) of J [y], it is rather natural to use S = Cl, that is the space of smooth test functions. However for many problems, it is more useful to work with continuous and piecewise-smooth test functions, denoted by D1. The following are well-known from the calculus of variations [1], [28]. Definition 3.5.1 A test function is called an admissible test function if it also sat- isfies all the prescribed constraints on y(2:) such as the end conditions. Let g(a‘) be an admissible test function for minimizing J [y] and introduce the nota- tion {W(z) := F(:I:, g(x), 37(3)), 131,,(x) := 6F(:c, g(x), 3)’(a:))/0y, etc. Theorem 3.5.2 A C1 solution 3) that minimizes J [y] must be a solution of the Euler-Lagrange equation Ry - —-(F.y') = 0- (3.172) Proof: See [28], page 13. D 51 Definition 3.5.3 A smooth test function which. satisfies (3.172) is called a smooth extremal for the basic problem (3.169)-(3.170). Definition 3.5.4 An admissible test function y(:z:) is said to be a weak local min- imum for the basic problem {3.169)~(3.170) if J[y(:c)] Z J[3)(a:)] for all admis- sible functions of the type y($) 2 10(3) + €h(:r) with |h(:1:)[ S 1, |h’(:r)[ S 1, h(a) = h(b) = 0 and e > 0 sufiiciently small. Definition 3.5.5 An admissible test function 32(23) is said to be a strong local min- imum if 3 e > 0 such that J[y($)] 2 J[y(2:)] for any admissible y satisfying lit/(~73) - 3?($)| < 6 in [a,b]- Weak and strong local minima can also be defined by using weak and strong neigh- borhoods of y($) as described next Definition 3.5.6 For a given 6 > 0, the set of all piecewise-smooth test functions satisfying [31(1?) - 9($)l < 6 (3-173) for each at E [a, b], is called a strong e-neighborhood of g(x) and denoted by Us(6,i]($)) . Definition 3.5.7 For a given 6 > 0, the set of all piecewise-smooth test functions satisfying the inequality (3.173) and |y’(:r) - i?’(x)| < 6 (3.174) for each :1: E [a, b], is called a weak e-neighborhood of g(x) and denoted by UW(€ag($)) 52 Definition 3.5.8 An admissible test function y(:r) is said to be a strong local mini- mum if 3 e > 0 such that J[y(a:)] 2 J[y(:r)] for all y E U5(e,y(:r)). Definition 3.5.9 An admissible test function y(:r) is said to be a weak local mini- mum if El 6 > 0 such that J[y(a:)] Z J[y(:r)] for all y E Uw-(e,y(:c)). If S 2 D1; i.e., the set of all piecewise-smooth functions in [a, b], then the Euler- Lagrange equation (3.172) alone is not sufficient to characterize the minimization. The characterization needs to be modified as follows. Theorem 3.5.10 In order for the admissible piecewise-smooth test function g of the basic problem (3.169)—(3.170) to render the integral J [y] a weak local minimum, it is necessary that A EN) = c + / Eye) dt, (3.175) for some constant c. Proof: See [28], page 38. D Equation (3.175) is called the integral form of the Euler-Lagrange equation. Corollary 3.5.11 At any point :1: where g' is continuous, it is necessary that 9 satisfies the Euler-Lagrange equation (3.172). Definition 3.5.12 A piecewise-smooth test function g(x) that satisfies (3.175) is called an extremal for the basic problem (3.169)-—(3.170). 53 It is more effective and convenient to work with the Euler-Lagrange equation (3.172). Hence, it is important to know whether the extremals for a basic problem are smooth and the locations of discontinuities of their derivative when they are not smooth. These discontinuities are known as corners of extremals. Information on discontinu- ities can be obtained from the so—called Erdmann’s corner conditions. Definition 3.5.13 A corner point of a test function y(:1:) is a point :r where the left derivative and right derivative of y(:r) exist at :z: (and are denoted by y’_ and y’+ respectively) but they are not equal. Theorem 3.5.14 If y(:1:) is an extremal, then 13],,(1‘) is continuous at a corner point. Proof: See [28], page 45. [:1 Theorem 3.5.15 If g(z) is an extremal, then the quantity (F—g'fiw is continuous at a corner point. Proof: See [28], page 46. [:1 Theorem 3.5.16 (Hilbert ’3 Theorem)Suppose that g is an extremal, 2:0 is not a corner point, and 13‘ y.y,(a:0) 51$ 0. Then 7) is C2 and (man (Riot/1L (F... — F.) = 0. (3.176) holds in some neighborhood of .130. Proof: See [28], page 48. [:1 Equation (3.176) is sometimes called the ultradifferentiated form of the Euler- Lagrange equation. 54 Definition 3.5.17 A scalar valued function f : R" ——> R is said to be convex if f(/\w+ (1 — /\)2) S Af(w)+(1— /\)f(z) (3.177) Vw,z€ R", and VA 6 [0,1]. Theorem 3.5.18 If F (x,y, y’) is differentiable in its arguments and convex in y and y’ for each x, then any admissible extremal of the basic problem (3.169)-— (3.170) renders J [y] a global minimum and conversely. Proof: See [28], page 108. CI Theorem 3.5.19 A scalar valued function of one variable f (y) is convex if and only if f” is everywhere non-negative. Proof: See [28], page 105. E] This completes the quick review of the appropriate machinery from the calculus of variations. We may now on this basis rapidly show that g(r) as given by (3.115) is indeed an absolute minimum. For (3.168) the Lagrangian F is F(r,g,g') = r\/4 + (rg’)2 — 2r. (3.178) From (3.178) it follows that (92F 4r3 8—9’5 = (4 + (T9,)2)3/2 > 0. (3.179) Hence F (r,g,g’ ) = F (r,g’ ) is convex in g’ for each fixed r by Theorem 3.5.19. If [‘11] S \Ilmax, then it follows by the Theorem 3.5.18 that g(r) given by (3.115), (3.118), (3.119) is an absolute minimum of (3.168) obeying (2.17). 55 3.6 Direct Methods for Minimization when k = 1 At the end of the last section it was shown that if [‘1'] is restricted by [W] S \I/max where \Ilma,‘ is given by (3.117), then g(r) given by (3.115) provides an absolute minimum for (3.168) among functions obeying (2.17). The significance of this result is that minimization of (3.168) subject to (2.17) governs the k = 1 problem. In this section we will investigate the minimization of (3.168) subject to (2.17) for arbitrarily large twist \II. Since g’(R,-) -> 00 as W —> \Ilmax, it is conjectured that minimizers might involve vertical line segments. Since such a vertical line segment cannot be considered as a function of R, it is expedient to use the parametric form of the energy integral (3.168) for further investigation. Consider the basic problem of calculus of variations in parametric form: Find a piecewise-smooth curve C = {(x,y) = (x(t),y(t)) : t1 S t S t2} that minimizes the definite integral t—2 J0: / f(t,x,y,x,y)dt,= / f(t,x,y,x,y)dt, (3.180) C t—l subject to WHILE/(’51)) = (331,91): ($(tzlay(t2)) = ($2ay2)' (3-181) The dots indicate derivatives with respect to the parameter and f is a given function, continuous in all its arguments when (x(t),y(t)) lies in a given region G; x, 3] have arbitrary values and are not zero simultaneously. The function f cannot be completely arbitrary because the integral under consideration must depend only on the curve C and not on any particular parametric representation of the curve. By the discussions given in [1], p 39—40 this independence of parametric representation will hold if and only if: (i1) t does not appear explicitly in the integrand function f; that is, f(ta$.y.i.3)) = f($.y.i3.3'l)- (3-182) 56 (i2) The integrand function f (x, y, x, 3)) must be positively homogeneous and of the first degree relative to the second pair of arguments; that is, f (x. y. mi. me) = m f (:v. y, i. 3)) Vm > 0- (3-183) As a result of homogeneity If the common ratio in (3.184) is denoted by K = K (x, y, x, y), then fix = 92 K, fey = ~33? K, fyg = 1132 K. (3.185) Suppose an admissible curve C: x = at). y = @(t) (t. s t 3 6) (3186) gives the functional JC a weak relative minimum. Then functions (13(t), (p(t) must satisfy the Euler-Lagrange equations in parametric form f. = A + f ME) d5. f. = B + f me as (3.187) in which A and B are constants. If arc length 3 is used as parameter, then d :i: = i = cosw, y = "g 2: sinw, (3.188) ds ds where w is the angle between the tangent to the curve and the x-axis. Along each smooth arc of a curve at d _ . __ = __ ... = 3.1 9 dsf. f. o, d812, f. o, ( 8 ) and at a corner point 3 = so fi(3(i) " I'd-(36) 2' 0) fy(5(i) " 131(36): 0' (3.190) 57 (Here 53 and 3; denote the one-sided limits lime—.3; and limsfisa respectively.) A smooth solution of (3.189) is called an extremal and (3.190) express the Weierstrass- Erdmann corner conditions. Note that (3.190)1 is equivalent to corner condition given by Theorem 3.5.15 and (3190);; is equivalent to corner condition given by Theorem 3.5.14 for the nonparametric case. If parametric forms of r = x(t) and g = y(t) are used, then the energy integral (3. 168) becomes Hp[x(t),y(t)] = 27rhoz [t2 E W — 2x] 36 dt, (3.191) t1 where the subscript p in Hp indicates that the parametric form of the energy is used, if : dIII/dt, 3] : dy/dt, ($(t1),y(t1))=(R1,2,/Ji) and ($02), 3102)) = (R0, 2110)’ Note that the Lagrangian of (3.191) can be written as f(x, y, x, y) = x 43':2 + x2y2 — 22:31. (3.192) Homogeneity of f in the last two variables is immediate since f(x, y, mi, my) = x \/4(m:i:)2 + x2(my)2 — 2x(mx) (3.193) = x \/4m2x2 + x2m2y'2 — 2xmx = m[x\/4:i:2 + x2 '2 — 2x33] : mf($7y3j:7y.)' Thus condition (i2) is satisfied. Moreover, condition (i1) is satisfied since t does not appear explicitly in the Lagrangian (3.192). Therefore the value of the energy integral (3.191) depends only on the curve C not on any particular parametric representation of the curve. 58 The Euler-Lagrange equations (3.187) are [ 4x1i: t and x3y \/4i2 + $2312 where 51 and 52 are constants. Note that (3.195) is equivalent to (3.113) and by = 52, (3.195) taking the derivative with respect to t in (3.194) the first Euler— Lagrange equation becomes d [ 423x . x2y2 — — 2x] = 4x2 + 2:2 '2 + — 2a. 3.196 dt 43';2 + 33292 y /4:i:2 + $292 ( ) From (3.195) and (3.196) it follows that y(t) = constant and curves described by (3.115) are extremals whereas x(t) = constant are not extremals. An extremal consisting of an arc of (3.115) and a straight line segment y(t) = constant will not satisfy the Erdmann’s Corner Conditions (3.190). For [\IJI S ‘Ilmax, we consider the composite curve consisting of the curve obtained from (3.115) by vertical translation (with d1 = R? and d2 = 1b,») in conjunction with the line segment x = R,- between y = w,- and y = «to — \Ilmax. This composite curve will be called A1. Thus A1 consists of a vertical translation of the \P = \Ilmax extremal so as to meet the x = R0 end condition y(Ro) = 1110, along with an x = R,- vertical segment so as to connect to the x = R,- end condition y(R,) = 212,-. Since this vertical segment is confined to the boundary, it need not satisfy Euler-Lagrange condition (3.195) and (3.196). We begin with some direct comparisons. In particular, it will now be shown that Al as described above provides a minimum to (3.191) among a collection test curves: AnH, A2, A3(n), A4(n) that we now define. 59 In the remainder of this section we will take 11), = 0 and 1/10 > 0. For (b0 S \Ilmax, we note that A1 can be parameterized as follows mt): R‘ if t6 [O’Ri] (3.197) t if te [R.,Ro], (7pc— \Ilmax) t/R1 if t E [0,R1] (wo — \I’max) + arctan \/(t4/R§) — 1 if t E [R,-, R0], x(t) = Then the corresponding energy can be computed from (3.191) as leAI] = Ed + E3 (3.198) where Ed -_— when? (a, — rm], E, = 277hci [, /Rg — 123+ R3 — R3]. (3.199) Note that first term Ed is the energy corresponding to vertical segment x = R,- and second term E, is the energy corresponding to the curve obtained from (3.115) by vertical translation. The associate matching requires that d1 = R? and d2 = 0 in (3.115). Since \Ilmax = arccos(Rf/Rg), the energy expression (3.198) may also be written as Hp[A1] = 2nhd [RE (the — arccos(fi—‘ZD + VR: — R? + (R,2 — 123)] (3.200) As we —> \IJmm = arccos(Rf/Rg) the energy associated with the discontinuity Ed goes to zero as expected. Let Any be the curve corresponding to neo—Hookean solution (2.34), (2.36) and (2.37) with w, = 0. Then it has the parametrization an(t) =t ift E [R,-,Ro], (3.201) _ '"RgRg'wo + $0 R3 _ (R?) — R?) V (R?) - Rf) ynHU) 1ft 6 [Ri,Ro]. 60 [ we [I N. Energy for A1 ] N. Energy for Any ] 2 1.4605 1.7056 2.5 1.9605 2.4835 3 2.4605 3.3440 3.5 2.9605 4.2720 4 3.4605 5.2570 Table 3.1: Normalized energies for A1 and An” for several we values when R.- = 1, Re 2 4 and w, = 0 Then the energy for the present case k = 1 using the k = 0 neo—Hookean deformation can be computed from (3.191) as 33103 leAnHl =27Thé{Ri2—RZ+R3 \/1+ (R?_R2)2 _ R4 (t2 12.2sz R2 «p R4 w? 2 o o i o 0 o 0 o o R*‘\/1 + (162,.2 — 123;)2 +(123— R3)210g [123— 123+ 1+ (12,?- Rg)2l -- .22 (fi—%) x—e :—er “0) R3R3¢o2 16g [31L + \/1+ (Raft/23 ] } Note that He[A,,H] z 5.2570(2nhd) when Ri = 1, Re 2 4 and we = 4. In the following table, Table 3.1, the normalized minimum energies for A1 and An” are presented for several we values when R,- = 1, Re = 4 and w,- = 0. The energy in this table is normalized via division by 2 77 h a. It is consistently verified that A1 < An”. We now construct a broader class of comparison functions. Let A2 be the straight line connecting (Re 0) to (Re, we). It is given by the parametrization x2(t) =t ift e [Rene], (3.203) 920) = ”(t—1,25% ift e [R,-,Re]. 61 The corresponding energy is computed from (3.191) to be He[A2] = 2 77 ha [R3 — 173]. (3.204) {4 (R. — R.)2 + RS 12313” — [4 (R. — R.)2 + R? 212213/2] “Ra-302413 . Let A3(n) be the family of curves given by the parametrization: +2nhd] 3:3,..(t) =t ift e [Rene], (3.205) y3,,,(t) = we [1+ (arctan —T%:—:—f§-)-)/arctan n] ift E [Iti,Re], with n 2 0 and not generally an integer. Figure 3.15 shows 6 curves of this family for the case R,- = 1 and Re = 4 (implying \Ilmex 2 1.50826) and we = 4, namely n = 0,1, 10,100, 1000, 10000. Note that in the limit as n —+ 0 these rectifiable curves tend to A2, and as n —> oo tend to the step function t if t e R,,Re x3,ee(t) = [ 1 (3.206) Re if t e [Re,Re+we], - 0 if t e [ReRe] 93,000) = (t—Re) if t e [Re,Re+we]. By numerical computation it is found that Hp[A3(n)] increases with n. The corre- sponding energy as n —> 00 can be computed from (3.191) as ”11.520 Hp[A3(n)] = 2 77 hd R3 we. (3.207) Let A4(n) be the family of curves given by the parametrization: x4,,,(t) = t if t 6 [Re Re], (3.208) 2%))/arctann] ift E [RaRo]. y4,n(t) = we [(arctan R0 62 G(R) Figure 3.15: Graph of the family A3(n) with R.- = 1, Re = 4, w,- = 0 and we = 4 for n = 0 (the diagonal line), and 1, 10, 100,1000, 10000 with n 2 0. Figure 3.16 shows 7 curves of this family for the case R,- = 1 and Re = 4 and we = 4, namely n = 0,1,10,83.1,100,1000, 10000. The dashed line shows the case n = 83.1, the significance of which will shortly be made clear. In the limit as n —-+ 0 these rectifiable curves tend to A2, and as n —> oo tend to the step function . “f 0, . $4,000): R' 1 t6[ R] (3.209) t if t e [R,-,Re], «at if t6 [0.1%] we if t e [ReRe]. 314,00“) = In this case the corresponding energy as n —> 00 can be computed from (3.191) as lim He[A4(n)] = 2 7r 66. R? we. (3.210) Numerical calculation shows for this family of curves that the corresponding energy He[A4(n)] initially decreases with n after which He[A4(n)] subsequently increases 63 .5 0| Figure 3.16: Graph of the family A4(n) with R,- = 1, Re = 4, w,- = 0 and we = 4 for n = 0 (the diagonal line), and 1, 10, 83.1, 100, 1000, 10000. The dashed line shows the case n = 83.15 which results the minimum energy among the family A4(n) with n. Let nmie = nmie(R,-, Re, we) denote the transition value of n between these two behaviors. We find for example nmin(1,4, 4) z 83.15 and the corresponding minimum energy among A4(n) is Hp[A4(83.15)] z 3.70242 (2rhé). In Table 3.2 the transition value of n and corresponding normalized minimum energies among A4 (71min) are presented for several we values when R,- = 1, Re = 4 and w,- = 0. Once again the energy in this table is normalized via division by 2 77 h 6. Taken together the family of curves A3(n) with n decreasing, the single curve A; and the family of curves A4(n) with n increasing form a larger continuous family of curves with minimum He at A4[nmie(R,-,Re,we)]. We find on the basis of extensive numerical calculation that the energy He at A4[nmie(R,-, Re, we)] is still greater than the energy corresponding to A1 for the same values of R,- , Re and we. In particular comparison of Tables 3.1 and 3.2 shows this for a few selected cases. Figure 3.17 for R, = 1, Re = 4 and we 2 4 compares the 64 [ 212.. [[ n [Normalized Energy] 2 18.08 1.54436 2.5 28.14 2.10070 3 42.52 2.64498 3.5 61.00 3.17782 4 83.15 3.70242 Table 3.2: Normalized minimum energies A4(nmie) for several we values when R,‘ = 1, R0 = 4 and $1 = 0 graph of A1 with the graph of A4(83.l5) which gives the minimum energy in this extended family. Consequently, for a given Re and fixed we 2 ‘Ilmex the minimum among the computed energies remains Hp[A1]. Formally, A1 and hence He[A1] is defined only if we 2 \Ilmex. However it is natural to extend A1 to 0 S we S \Ilmex via (3.115) and this is understood in what follows. The graph in Figure 3.18 shows normalized energies E1 , E2, E3, E4 and E,” as a function of the amount of twist we corresponding to A1, A2, A300. 114,00 and An”, respectively, when R,- = 1 and Re = 4. The normalized energy associated with A4['nmin] as given in Table 3.2 is also indicated by stars for we = 2, 2.5, 3, 3.5, 4. If we E [0, \Ilmex] then the torque-twist relation is given by (3.132). If we 2 \Ilmex then we note from (2.48), (2.49) that it is enough to compute BW/Bwe which can be computed from (3.200). Hence the torque-twist relation for A1 is given by 27rhéR§R§sinwe if 0322,3me M = JR? +R;‘,—-2R,2 R3 coswe (3-211) 27th§6 if we 2 am The conspicuous feature of (3.211) is that the twisting moment associated with this extended deformation A1 is monotonically increasing with we on 0 S we S \Ilme,x after which the twisting moment remains constant. The physical interpretation is 65 G(R) 1 ' ' ' I 0-5 --------------- J" - — (dashed line) A4(83,1) _ —— (solid line) A1 0 ..... 4 """"" 1 ______________________________ Figure 3.17: Graph of A1 and A4(83.15) when R,- = 1, Re = 4 and we = 4 0| A Normalized Energy to (D Amount of Twist Figure 3.18: Normalized energy versus twist for R,- = 1, Re 2 4, w,- = 0 and varying we. Here \I/max = 1.50826. The energies corresponding to A1, Ag, A3,ee, A4,ee and Any are normalized by dividing 2 77 h (‘1. The normalized energy associated with A4[nmie] is also indicated by stars for we = 2, 2.5, 3, 3.5, 4 66 that conventional macroelastic resistance to twist occurs until we 2 \Ilmex after which substructural rearrangement concentrates at the singular surface R = R,- so as to relieve any further macroelastic resistance. This novel effect is not present in the conventional theory of hyperelasticity, although it is reminiscent of other phenomena in the conventional theory such as cavity formation and shear banding at critical load values [5], [14], [19], [23], [27]. It will now be proven that Al given by (3197) provides a weak local minimum for the parametric energy integral (3.191) subject to (x(tl),y(t1)) = (R,,0) and ($(t2), y(t2)) = (Re, we) in the family of piecewise-smooth curves lying in the closed region a = {(x,y)|R.- s a: 3 120,0 3 y s 7).}- To this end we begin with a technical detail. Namely we note that Al given by (3.197) is a differentiable curve only if t 75 1. However the location of non- differentiability at t = l is an artifact of the parametrization. If A1 is reparameterized in terms of arc length, then it is also differentiable at the location associated with the original t = 1. To this end let 5 be the arc length parameter. Then A1 can be parameterized by R,- if s E [0,we— \I'mex] ‘3 h(s) if s E [we—\Ilmex,se], (3.212) s ifs E [0,we — \Ilmex] we — \Ilmex + arctan \/(h4(s)/Rf) —— l ifs 6 [we - ‘Ilmex, se], where se is the total arc length and h(s) 0 By using (3.197), it follows that ( 3 213) )can be written as ’1(3) 4 s— — we — \Ilmex +/ \/1-l~::7 4R dt. (3.214) -—R4 t2 The integral in (3.214) does not have a readily explicit antiderivative hence there is no readily explicit form for h(s). However using (3.214) it can be seen that parametriza- tion of A1 given by (3.212) is a C1 parametrization. Proposition 3.6.1 The A1 given by {3.197) is a weak local minimum of {3.191) sub- ject to (x(t)),y(t1)) = (Re-,0) and (x(tg),y(t2)) = (Re,we) in the class of piecewise- smooth. curves lying in the closed region G = {(x.y)]R,,- S x S Re, 0 S y S we}. Proof: Let 6 2 0 be a given small nonnegative number. Consider the curve defined by 5(7) = 51(7) + e ((7), (3.215) W) = 910) + 6 720). where ((0) 2 77(0) 2 0, C(Re) = 77(Re) = 0 and ((t) Z 0 if 0 S t S Rf. Here x1(t) and y1(t) are given by (3.197); i.e., (x(t), y(t)) is a small perturbation of (x1(t), y1(t)) that is confined to lie in the closed region G . Then the variation in H,e as a result of this perturbation is defined by AH. = Hp[x1(t) + . ((03.0) + e 1200] - H.930, 91(1)], (3216) or equivalently R0 AHP=(27rhd) {f[x1+6C,y1+677,x'1+eC,y°1+67'7]— (3.217) 0 f[xla 111,151: 91]} dt) where “dot” denotes d/dt. By using the Taylor Series expansion for the first term in the integrand, (3.217) becomes R0 AHP = (27fhd) {f[$1,y1,£61,?j1]+ (3.218) 0 68 of of of . of. . . €[8_xl-C+ Can-*- 5.13—1—(+— Cyln 7”] +6"°2[l + — flzlaylixlvyllldt’ Hence (3.218) can be written as 2 AHP=(2rh&)[661+%621+...], (3.219) where 61 is the first variation given by 3° 3f 3f (9-f (9f 61 2 —~ + — +— +-——— dt. 3.220 /0 [Bxlc (93/177 6231 8y1 7}] ( ) Since A] has a C1 parametrization, ('3 f /0x1 and a f /0y‘) are differentiable with respect to t. Hence integrating by parts in (3.220) is permissible, which in turn gives ‘51: logici'ayfl l R0 at (8f 8f 6f /o “of 6231 dt(8x1)lC+ leg—g a—gh)]n}dt. The boundary terms in (3.221) vanishes by virtue of ((0) 2 77(0) = 0, ((Re) 2 + (3.221) 0 q(Re) = 0 so (3.221) reduces to 61 = /0R0{[§—xfl -— Big—bl (+ [8%]; — 52(6‘3—i)] 1]} dt. (3.222) Note that the integral in (3.222) is naturally written as a sum of the two integrals associated with the different parts of the composite curve A1: 32' . 61=/0 {[gxil—%(g—1{1)]c+ [g—yfl—ac-i; 62%)]77} dt. (3.223) Ro I +fR, {[§£*%(%)]C+[§i—% %)]n]dt. The second integral in (3.223) is identically zero as the portion of A1 restricted to [R.-,Re] is the solution of the corresponding Euler-Lagrange equation for (3.168). Moreover since the Lagrangian (3.192) is independent of y, it follows that (3.223) reduces to 61 2 Am { [g1- — %(66_1f1)] C— [1(3le 7)} dt. (3.224) 69 Since x1, y1 are given by (3.197), it follows that the integral in (3.224) simplifies to Rt 61 = / 2w. — 91......) ((7) dt. (3.225) 0 In view of the restriction C (t) 2 0, it is immediate that 61 2 0. (3.226) Consequently for a sufficiently small 6 2 0, the sign of AH? is determined by the sign of 61 which, in turn, is positive by virtue of (3.226). Thus AH. = lesvl(t) + e C(t).y1(t) + . 0(01 —- lercl(t).y1(t)l 2 o, (3.227) which shows A1 is a weak local minimum of (3.191). This completes the proof of Proposition 3.6.1. C] We conclude this section by proving that A1 is indeed an absolute minimum of (3.191) in the class of piecewise-smooth curves such that, once restricted to the interior of G , their parametrization reduces to a function y = f (x) The importance of such a result is that if we seek minimizers of (3.168) when we > \Ilmax in the space of piecewise-smooth functions g(R) then the global minimum involves a discontinuity at R = R,- of magnitude [[9]] = we — \Ile,ex such that g(R) on R,- < R S Re is given by a vertical translate of the we = \Ilmex solution. The energy Ed associated with the discontinuity is given in (3.199)1 and the energy Ee associated with the translation is given in (3.199)2. Proposition 3.6.2 Let y(x) be any given function E D1([Re-,Re]) with y(R.) = 12:“ and y(Re) = we. Let y(x) be the function given by g(x) = (a. — 9,...) + arctan (TE-)4 — 1 x e [17,, Re]. (3.228) Then HIE/(513)] _ H1 [g(xll 2 R22 (271' lid!) ("/90 _ ‘I’max — 121'), (3229) 70 where H1[y(x)] is given by (3.168) (x and y(x) are used respectively instead of r and g(r)) and \IImax is given by (3.117). Proof: Since the Lagrangian of (3.168), F (x, y’) = x 4 + (xy’)2 — 2x, is convex in y’ for a fixed x E [R,-, Re] it follows that F(Ivyl) _ F(SE, 9’) Z 1231’ (y, _ ill)? (3230) where F.y’ = Ey’ (3.231) y=f1 Integrating (3.230) from Re- to Re gives R0 R0 / F(x.y’> —F(x.7)dx 2 f F... (y'—0’>dz. (3232) Hi Ri whereupon R0 A fills/(13)] — Haw] 2 (2.7...) f _ 13./(y’— :7) dz. (3233) By integrating by parts from (3.233) R0 — (3.234) Ri H.000] — H.399] 2 omega, 0 — 27>] [201,372.10 — .0) dz}. Ri Since y(x) is the solution of the Euler—Lagrange equation for (3.168), d[13‘ y,]/ dx = 0. Hence equation (3.234) reduces to R0 H.090] — H.129» 2 (2 «h a) [Ry] (y — 27)] (3235) Rt. Moreover y(Re) = y(Re) = we, so (3.235) reduces to H1[y(:r)l—H1[39(:v)l2-(27129) [F37 mas—3(a)». (3.236) x=R1 71 Now F.y’ == x3y’/(/4+ (73y’)2 implies F y, = RE. Using this result in (3.236) in conjunction with the boundary values y(E) : w, and y(R.) = (1110 - \I/mex), it follows that (3.236) becomes H1[y(:v)l - H1 [31(7)] 2 R? (2 77 h a) (we — Wm... — 17,-). (3.237) This completes the proof of Proposition 3.6.2. El Theorem 3.6.3 Let A1 be the curve given by (3.197). Let C be any curve given by the following parametrization x(t) = Rd 2ft 6 [O’R‘] (3.238) t z‘ft e [R.,R..]. (1/21 tl/Ri 2f t 5 [0,52] 900) 2'f t E [Rifle]. y(t) = where cp(t) is a piecewise-smooth function with go(R,-) = w,- and (p(Re) = we. Then le$(t).y(t)l — lefliiU). 3110)] Z 0- (3-239) Proof: Note that R0 le$(t).y(t)l=27rhé 0 f[$(t).y(t).i¢(t).i/(t)ldt, (3-240) .. Rt . . Ha I =2rha( 0 fl:v(t).y(t),x(t).y(t)ldt+fm le.y($).y($)ld$). where F [x, y(x), y’ (x)] = x(/4 + (xy’)2 — 2x is the nonparametric Lagrangian. Simi- larly R i le$1(t).yl(t)l = ma 0 711230330330,7.0)] dt (3.241) 72 R0 +277hd/ F[x,y1(x),y’1(x)] dx. Ri Therefore lel‘U). y(t)] - lex1(t). 910)] = Ri 27Thé {flxayaxayl —f[$11ylaaflay.1]}dt+ 0 R0 27% R. {F [$.y($),y’($)l-F [audits/{(006133- By Proposition 3.6.2, equation (3.242) gives le$(t).y(t)l - le$1(t).yl(t)l 2 Ri (3.242) (3.243) 277hé f[x, y, 6:, g] — f[x1,y1,x'1,y'1] dt + R?277h&(we — wee. — 773,-). 0 However after computing the integral in (3.243) it is found that le$(t)a y(t)] — le$1(t). 3110)] Z 2M5! R? [1171'— (w. — ‘I’maxfl +Ri2 2717161 [(wo "' \I’max — 127)], or simply lem(t)ay(t)l — HPl$1(t)v y1(t)l Z 0 This completes the proof of Theorem 3.6.3. Cl 73 (3.244) (3.245) CHAPTER 4 A More Generalized Class of Material Models 4.1 Extended Material Model We now extend the model in Chapter 3 along lines suggested by Pence and Tsai [22]. In particular, we consider an extended material model whose strain-energy function is given by a a" W—§(11—3)+3 (If - 3) + (i1 - 3). (4-1) MIQ> where a Z 0, 07" Z 0, d 2 0. The form (4.1) augments the previous form (3.6) by inclusion of the additional 11 -— 3 term. Our interest is in the extent to which a > 0 gives different qualitative behavior from the a = 0 treatment of Chapter 3. The boundary value problem for pure azimuthal shear with boundary conditions (2.17) is again the focus of study. For the extended material (4.1) the minimization procedure again yields (39) where now the Cauchy stress (3.10) is given by a=aB+dB—pI. (4.2) 74 In polar coordinates (3.9) again reduces to (2.14)—(2.16). The interstitial balance is still given by (3.11) namely a‘B—ciB —qB=0. (4.3) Here p is the hydrostatic pressure and q is the Lagrange multiplier associated with the constraint (3.4) and B is still the left Cauchy-Green tensor given by (2.12). When 61 = 01* the pure azimuthal shear problem reduces to the following first order nonlinear ordinary differential equation - , d arg’+—-arg “— (4'4) where d is an integration constant. This result can be obtained as follows. Note from Section 6 that if (i = oz“ (k = 1) then p = 0 or equivalently q = 0. Therefore the nontrivial components of B are still given by (3.109)—(3.112). The Cauchy stress tensor (4.2) can thus be expressed in terms of r and g’ (r) One now finds from (2.14)—(2.16) that p = p(R) whereupon (2.16) is satisfied automatically and (2.14) yields (4.4). The remaining equation (2.15) then gives dp . d 1 I 2 (37(902 —=2a— —— —ar —-———, 4.5 d7 d7( 4+ (r992) (‘9 ) \/4+ 0.7)? ( ) from which p = p(r) can be computed once g(r) is known. Attention shall henceforth be restricted to this case 6. = a“. Note that (4.4) with a 2: 0 and d1 = d/c‘r retrieves (3.113) of the previous development. Equation (4.4) when written as a polynomial in g’ is quartic and hence can be solved for g’ . It will be shown that if a > 0 then equation (4.4) subject to boundary conditions (2.17) has a unique solution. Hence it is necessary to pick the correct root of the quartic polynomial. Even though it is possible to distinguish the correct root from extraneous ones, it does not seem feasible to integrate this root to get an explicit 75 form of g(r). Therefore the solution of (4.4) subject to (2.17) will be investigated by alternate means. Note that equation (4.4) can be written as 9': .1. [i _ 5‘9, ] (4.6) a 7‘3 4+ (79’)2 or equivalently _ ,1 '1 3 " 9’(€) g(r)—g(awafeu 0Lfl+(€g,(€»,de (4.7) By integrating the first integral in (4.7) and using the first boundary condition in (2.17), g(Re) = w,, equation (4.7) can be written as d r2—R.2 61 ' T' = i ——2——2t -— F , I d, 4.8 g(> Mam”, a]... (mom ( ) where F , I : g’(€) . 49 (69(5)) \/4+(£g'(o)2 < ) By using the other boundary condition in (2.17), g(Re) = we, the constant (1 can be computed as F(€,9'(€)) 616- (4-10) 2aRzR2 2737R2R2 3" d=(w..—w.) * ° 7 f 63—6.? + fez—R? ...- Therefore boundary value problem (4.4), (2.17) is equivalent to the following integral equation 90‘) = 1111' + (100 — 1Pi)P(7‘)+ r (3 RO (Pm—1) f F(r,g'0. 78 Since it is more convenient to deal with integrals than derivatives, we will use the integral equation form (4.8) of the first boundary value problem (4.4), (2.17). Then the principle of contraction mapping will be used to prove existence and uniqueness. Let S be a normed linear space with the norm denoted by ||..|| A sequence of elements yn of S is a Cauchy sequence if for every 5 > 0, there exist N such that Hyn — ymll < 6 whenever n, m > N. S is a complete space if every Cauchy sequence in S converges to a limit element in S. An operator T mapping a normed space S into itself is said to be a contraction mapping on the space if there is a number A, 0 < /\ < 1, such that for all 11:, y E S IlTx - Tyll S A llx - yll. (4-25) The following theorem is well-known. Theorem 4.2.1 Principle of Contraction Mappings: Every contraction mapping T defined on a complete normed linear space S has one and only one fired point (i.e., y = Ty has exactly one solution). For any point yo 6 S, the sequence of iterates yn = Tyn—i = = T"y0 converges to the fixed point y and n A Ilyn - 31” S 1_ ,\ llyr - ml (426) Proof: See [4], page 26. Cl Lemma 4.2.2 If A _ . 2 2,3 (12" RR‘) } < 1, (4.27) max {go—l (R0 — R.) then the first boundary value problem (4],), {2.17) has one and only one solution. 79 Proof: Let So be the space of continuously differentiable functions on [R4, R0] with the norm ||g(r)||=maX{&I;1gRo lg(r)| 1.2%. '33)“) (4.28) where z(r) is the positive weight function given by 2(r)— Rg— 21R. (4.29) The space So is complete with this norm, and convergence of a sequence gn(r) in this norm to g(r) implies that g(r) is continuously differentiable and that gn(r) -—> g(r) and g;(r) —-> g’ (r) uniformly. Let T be the operator defined by the right side of (4.14). Observe that 3TgIr)=(zl.— w.)P’(r)+ 3]R° HI (re F(t’égI )) d4 (4.30) where Hr(r,€) = BIT/8r. Since P’ (r), g’ (r) and Hr(r,{) are continuous functions in r, the derivative of Tg(r) with respect to r as given in (4.30) is continuous in 1'. Thus T maps So into So. To see when T is a contraction mapping consider T91()-T92( =3] HIro Hégi(€))—F(€,g’2(€)))d€. (431) Hence a Re _ |T91(r)-Tg2(r)IS-] lH(r,€)l |F(€,gi(£))-F(€,9§(€))ld€. (4-32) 0‘ R.- Since F (5 , g’ (£ )) is Lipshitz in g’ with Lipshitz constant 1 / 2, inequality (4.32) gives ‘ ’ d R0 ‘ 1 I I ITgi(r)-ng(r)| .<_ 3 ] IHIr,oI 5 lam—92mm, (4.33) Rf which implies that |T91(r)-Tg2(r)l< 55-min —g.I r)” )]R°z( €)|H(r€)|d€ (4.34) 80 Since 0 < P(r) < 1 It IS clear that |H(r,{)| S 1 for every r E [R2 R ] Therefore inequality (4.34) provides A R0 IT’91( )— Tg2I )I< :2— ]3z zI4) 44 ”91(7‘) — 92(1')||- (435) Computing the integral in equation (4 35) gives ITg2I ) Tg2I )I s 3 (R2 — a)? Il91( ) 92( )II (4.36) Similarly it can be shown that 1 d - — mlng91(T)"T92( )} < 1 a “0 - 3 ) —3] zI4) IH.Ir.4)I 44 ”91(7") g2( )II (4.37) and by using (4.13) it is clear that 2 R2 I 2r,I 4)I< _ M22122) (4-38) for every r E [R.-,Ro]. Thus 1 d — — Tl 379919) — T92(T)}l S 1 52 R3 “0 33-3 2.3 (12% R2) ] zI4) 44 “91(7‘) —g2(r)ll (439) Computing the integral in (4 39) gives d - 1 d R§(Ro— R2-—)2 Idr{ g2I )— T22 r()<}l 3m 3 23022 R2r)ug2I )— 92(T)ll- (440) After Simplification equation (4 40) becomes “ 1 & R42)(Ro — R4) 3(1)—{T91(d)- Tg2Ir)}| 3 $5 3 ,3 ( 122+ R.- l|91( )— 92(r)ll-(4-41) Note that 1 R3 — R? 22213322135} = R3 (4'42) 81 Consequently (4.28), (4.36), (4.41) and (4.42) together give IITQIU‘) - T920)” S (4.43) ‘ 6 RO— - 2 max {3304, — 1).-)2. 3%} H.910") — g2(r)ll- Hence, if - a _ _ 2 M=maX{§93(R2—R2)2,3E°R1—B‘)—}<1 (4.44) then T is a contraction mapping on So. This completes the proof of Lemma 4.2.2. C] The inequality (4.27) provides a restriction that can be interpreted in various ways. For example if d > 0, R0 > R..- > O are given then (4.27) requires sufficiently large positive (1. Alternatively if a > 0, d > O, R,- > 0 are given then (4.27) requires that R0 is sufficiently close to R2. It can also be shown upon modifying the norm (4.28) with a different positive weight function that it is possible to weaken the inequality (4.27) thereby making the result less restrictive. However such modification does not seem to give an existence and uniqueness result that applies to arbitrary a > 0, d > 0, R0 > R.- > 0. As we now show, such a result is obtainable by bootstrapping the result with a development that first obtains an existence and uniqueness result for the second boundary value problem. Theorem 4.2.3 The Second boundary value problem, (4.4), (4.15) has a unique so- lution on [R,-, R0] for every R, > R,- > 0. Proof: Recall that the second boundary value problem (4.4), (4.15) is equivalent to initial value problem (4.19), (4.20). Let (r = 04/43 and d: d/éz. Then (4.19) is (4.45) 82 where d follows from (4.17). Let 2 _3 , 9’ _ i H(r29(7‘)2g(7‘))— 9+ 43339,), 33- (446) Then initial value problem (4.19), (4.20) is equivalent to following initial value prob- lem H (r2g(r)2g'(r)) = 02 g(Ri) = 4/22. (4-47) Let U = (0,00) x (—oo,oo) x (—oo,oo). Clearly U is an open convex region of the three dimensional Euclidean-space and H is defined and continuous on U. Moreover 0H/6g and 8H/Bg’ exist and are continuous on U. Additionally H(R.-,g(R2-),g’(R,-)) = 0 has a solution in U and the Jacobian J = BH/ag’ is not equal to zero at this point. Hence by the existence and uniqueness theorem for implic- itly defined initial value problems (for example [20] pages 44-47) there is an interval containing R,- such that the initial value problem (4.47) has a unique solution on this interval. Let the maximal solution domain for g(r) be the interval (a, b). Clearly R,- E (a, b). It will now be established that b 2 R, by showing that the alternative as- sumption b < R, leads to a contradiction. Note that (4.45) can be written as 2107 9’ : — —- —— ——-———-—— . 4.48 g 3 33 4+ (33),) ( ) Integrating (4.48) from R,- to r provides 1 ’ J 9’(8) = - + -- - - (18. 4.49 g(r) g(R2) 3 ] (33 4+ (8938”,) ( ) Let R1 and R2 be any two numbers with a < R,- < R1 < R2 < b. Then 1 R” J g’(8) (R —(R 2:] —— __ds 4.50 9 2) 9. 1) a 121 (33 4+ (sg’(s))2) ( l 83 Since __ 9’0") _ 9’(T) i _1_ ' |,.2. «“3933”,! S |T3l+l\/4+(rg’(r))2| S R§+R2 (4.51) equation (4.50) implies that |g(R2) — g(R1)I < I3 + i)1IR2 — HI. (4162) _ R? R.- _ Assume that b is finite. Since it > 0, values of g(r) for r near b and less than b satisfy a Cauchy condition and thus there exists lim,._..b g(r). Call this limit 9;, and let gncw(r) be a solution of the initial value problem H(T, g(T), 9,0.» = 0) g(b) = gb' (453) Since H is independent of g(r), it satisfies a global Lipschitz condition in 9. There- fore the solution gnew(r) is an extension of g(r). Since the domain of this extension contains b, the domain (a, b) is not maximal. This contradiction comes from the assumption b < R0. Hence the maximal solution domain for g(r) can be extended to any R, < 00. Of course this means that the second boundary value problem, (4.4), (4.15) has a unique solution on [R, R0] for every R0 > R2- ) 0. This completes the proof of Theorem 4.2.3. D Note that an alternative proof of the above theorem is presented in Appendix B. Similarly it can be shown that Theorem 4.2.4 The Second boundary value problem, {4.4), (4.16) has a unique so- lution on [R2, R0] for every R, > R,- > 0. To get the main result of this section, the following standard result from the theory of ordinary differential equations will be now utilized. 84 Theorem 4.2.5 Uniqueness and Continuability of Solutions of Initial Value Prob- lems: Consider the following initial value problem y”(t) + f (t2 y(t), v’(t)) = 02 (4-54) 900) = 3102 (4.55) 3/(150) = 316- (456) If f (t,y(t),y' (t)) is continuous and satisfies a Lipschitz condition in a domain D which contains the point (to,y0,y6) then there is only one solution of (4.54), (4.55), (4.56). Furthermore, this solution can be uniquely continued arbitrarily close to the boundary of D . Proof: See [4], page 10. D Theorem 4.2.6 The second order nonlinear boundary value problem (4.4), {2.17) has a unique solution on [R2, R0] for every 0 < R.- < R, < 00. Proof: According to Lemma 4.2.2 there is some finite length A such that all first boundary value problems for (4.4) on any interval of length less than A has a unique solution. Let n be an integer such that (R,- — Ro)/2" = A < A. Divide the interval [R2, R0] into 2" equal subintervals. Then by Lemma 4.2.2 on each subinterval all first boundary value problems for (4.4) have unique solutions. Now if it can be shown that first boundary value problem for (4.4) has a unique solution on an interval whose length is 2A then the same argument can be repeated for intervals whose length are 4A, 8A, etc. until 2" A = (R.- — R0). This in turn would establish that the first boundary value problem (4.4) has a unique solution on the interval [R2, R0]. Without 85 loss of generality we may assume that n = 1. Let Rm be the midpoint of the interval [Rb R2,]. Then by Theorem 4.2.3 the second boundary value problem (4.4) subject to 90%) = Ila, 9’01".) = m (4.57) has a unique solution. Moreover by Theorem 4.2.4 the second boundary value problem (4.4) subject to 9’(Rm) = m 90%) = 2122 (4.58) also has a unique solution. To continue the proof of Theorem 4.2.6 we now consider the following claim Claim 4.2.7 The solution gl(r, m) of (4.4), (4.57) and its derivative g,’(r, m) = agz/8r(r, m) are each strictly increasing in m. Namely, if 1712 > m1, then 91(r2m2) > 9le m1) (4.59) on [122,122.22] and 91(73 m2) > 9103 ml) (4.60) on (R2,le- Proof of Claim 4.2.7: Suppose m2 > m1. Clearly (4.59) is true for those r 6 (R2, Rm] and sufficiently close to Rm. If 9i(Rn2m2) = 95(Rmm1) (4-61) for some R, E (R2, R22], then the second boundary value problem (4.4) subject to g(Ri) = 1% g,(Rn) = 9l(Rn2m2) (4-62) has a unique solution by Theorem 4.2.3. This implies that g,(r,m1) E g,(r,m2) for every r E [R.-,R,2]. Moreover since f (r, g’ ) is Lipshitz, then every initial value problem has a unique solution by the Theorem 4.2.5. Therefore g,(r, m1) E g,(r, mg) 86 for every r E [Rm Rm]. In other words, branching of the solution at r = Rn is not possible. Hence gz(r,m1) E g,(r,m2) for every 1' E [R2,Rm]. But this implies that m2 = gf(r,m2) = gf(r,m1) 2 m1 contrary to the assumption m2 > m1. Thus g,’ (r, m) is a strictly increasing function in m. Moreover since g1(R2-,m1) = g)(R.-,m2) = 2]),- and gf(r,m2) > g; (r, m1) on [R2, Rm], we have g,(r, m) is also an increasing function in m for every r E (Ri, Rm]. This completes the proof of Claim 4.2.7. [I] I The proof of Theorem 4.2.6 now continues. It similarly can be shown that the solution g,(r, m) of (4.4), (4.58) and its derivative g;(r, m) are respectively decreasing and increasing functions of m. In particular g1(Rm,m) is an increasing function of m and g,.(Rm,m) is a de— creasing function of m. Since g,(r, m) and g,(r, m) are monotone functions of m without jump discontinuities, they are continuous in m on [R2,RNJ and [erRo] respectively. Therefore every first boundary value problem (4.4) subject to and every first boundary value problem (4.4) subject to g(Rm) : 2Ibma g(Ro) = $0 (4.64) have continuous solutions in m. To continue the proof of Theorem 4.2.6 we now establish the following claim Claim 4.2.8 If every first boundary value problem (4.4), (4.63) and every first boundary problem (4.4), (4.64) have continuous solutions in m, then these solutions both map the real line R onto itself. Proof of Claim 4.2.8: Let g;(Rm,m) and g,(R,,,,m) be the solutions of first boundary value problems (4.4), (4.63) and (4.4), (4.64) respectively. Consider the map h : R —-> R defined by m —) gz(Rm,m). Let 1]) be any real number. To show surjectivity we need to show that there exists m E R such that h(m) = gl(R2,2, m) = 87 w. Since the first boundary value problem (4.4) subject to g(Ri) = 112,-, g(Rm) = 1,!) has a unique solution let m be the value of the derivative at Rm. For such an m it is clear that h(m) = 91(Rm, m) = 1b. Therefore h(m) = g,(R,,,,m) is onto. Similarly g,(Rm, m) is onto. This completes the proof of Claim 4.2.8. CI The proof of Theorem 4.2.6 now continues. As a result of Claim 4.2.8, there must be exactly one number mo for which 91(Rm, mo) = gr(Rm, mo)- (465) By definition 91'(Rm, mo) = mo = g£(Rm, mo). (4.66) For m = mo, the function g(r) defined by g(r) = g(r’m) r E [R4, Rm] (4.67) gr(7'2m) T E [RmaRol has a continuous derivative at r = Rm. Hence it is the only solution of the nonlinear boundary value problem (4.4), (2.17). This completes the proof of Theorem 4.2.6. El Theorem 4.2.6 establishes the existence of a unique classically smooth solution for the pure azimuthal shear problem in the extended material model (a > 0, d = a" > 0) for all R4, R0, 1b., 1110. This contrasts with the lack of a classically smooth solution in the precursor model (a = 0, d = (1* > 0) whenever R2 I212. — WI > ‘I’max = arccos(fi’g). (4.68) Recall that in this case we found that a portion of the deformation localized at R = R.-. Thus the expectation for the extended theory is that a —+ 0 should in some sense retrieve the localized deformation studied in Section 3.4, suggesting a boundary layer type phenomena for sufficiently small a > 0 whenever we — 1M becomes sufficiently large. 88 4.3 Numerical Solution for the Extended Model In this section we present a numerical solution procedure for the pure azimuthal shear problem in the extended material model. Note that for a fixed or > 0 the pure azimuthal shear problem is governed by (4.4) and boundary conditions are still given by (2.17). Note also that for a fixed (3: > O (4.4) can be written as 3 T3 9' J (4 69) g 4 + (rg’)2 where a = a/a, J = d/a. (4.70) Differentiating (4.69) to eliminate J gives g”(7‘) + f (7", 9’0"), 6) = 0, (471) where I - — g_l 27.2 9’2 f(r’g’a)- r [3+4+a[4+r2g’2]3/2l' (4'72) For a given it the numerical solution of (4.71) subject to (2.17) can be found by various “shooting”, “parallel shooting”, “collocation” or “finite-difference” methods [16], [17]. Here we employ the finite—difference method to numerically solve (4.71) subject to (2.17). Let R,- = 7'0 < r1 < - -- < rn_1 < 7“,. = R0 be an equally spaced partition of the interval [Rb R0]. The derivativas in (4.71) are replaced by suitably chosen difference quotients defined in terms of the points 1“,. A typical choice is to use the centered difierence approximations 9'01)“ g(rj+1)2';zg(Tj—1), (4.73) 9.0,], g g(r)“) - 29g» + 903-1). (4.74) 89 When these approximations are used in (4.71), we obtain a system of equations of the form g(r)-+1) -29(rj) +9634) = f(r,- g(r)-+1) _9(7‘j-1)). 4. h2 2h ( 75) for j =1,2,...,n — 1 where g(ro) = 11),, 903,) = 2,00. In (4.75) the quantities g(rl), g(rg), . . . , g(rn_1) are unknown since any solution g(r) of (4.71) will probably not satisfy (4.75) exactly. We think of the system (4.75) as an approximation to the equation in (4.71) and hope that for a sufficiently small spacing, h = (R0 — R;)/n, the solutions 90",) of (4.75) are good approximations to the exact values of g(rj). Since the function f (r, g’ , o?) is nonlinear in g’ , we are faced with solving a non- linear system of n — 1 equations in n -— 1 unknowns. Newton’s method for several variables can be used to determine g(rl),g(rz), . . . ,g(rn..1). Note that Newton’s method applied to (4.75) takes the form Gk+1 = Gk — J(Gk)—1P(Gk) (476) where Gk = [g(k)(r1),g(k)(rg),...,g(")(rn-1)]T and where the 1"" component of the vector function P(G) is given by 1).-(G) = g(r...) — 29m) + you-..) — I22 f(r,, 9""J+1)2';f("j-1)). (4.77) Since BPj/ag(n) is zero except for Z: j — 1,j,j +1, the (n — 1) x (n — 1) Jacobian Matrix J (Gk) is tridiagonal; and it is relatively easy to apply Newton’s method to (4.75) even for large n. The starting vector Go could have components obtained by a linear interpolation of the boundary conditions: (0) .=I1_Z_Rl_' RO‘TJ' g (’3) R.-R.“’°+R.-R.- There are several programs to solve boundary value problems based on finite- z/a. 133' s n - 1. (4.78) difference method. MATLAB’s bvp4c is one such program and it is well explained in 90 0.5 0.4- 0.3- E ‘5 0.2- / .’ 0-1 ' ----- (dotted line) a=0.1 —- (dashed line) a=o.o1 — (solid line) $0001 0 x r 1 1.5 2 2.5 3 3.5 4 R Figure 4.1: Numerical solution when R,- = 1, R0 = 4, (D,- = 0 and” 1120 = 0.452478 for three values of 67, d =2 0.1, 0.01,0.001. Here [1/20 — 4),] = 03me [25]. MATLAB’s bvp4c is utilized to solve (4.75) subject to (2.17) for several values of 5:. The specific problem that we consider involves R,- = 1, R0 = 4, 212,- = 0 whereupon we consider the effect of increasing 1110 for various 67 > 0. The limit (31 —> 0 should then retrieve the results of Chapter 3 wherein we recall that \Ilmam z 1.50826 for a geometry with R0 = 4R,-. In the first case we investigate the solution of (4.75) when R, = 1 , R0 = 4, «[2,- = 0 and 1P0 = 0.452478. Values it = 0.1, Er = 0.01 and (I = 0.001 are used in Figure 4.1 and corresponding graphs are depicted by dotted, dashed and solid lines respectively. Note that 2/10 — w, = 0.452478 << \Ilma,x = 1.50826. In this case the three graphs are almost identical. In the second case we investigate the solution of (4.75) when R.- = 1, R0 = 4, w,- = 0 and 2/20 = 1.5. Note that \IIm,x 2 1.50826 when R,- = 1 and R0 = 4. Hence for this case 2/20 x \Ilmx. 57 = 0.1, (r = 0.01 and 67 = 0.001 are used in Figure 4.2 and the corresponding graphs are depicted by dotted, dashed and solid lines respectively. 91 g(R) : ----- (dotted line) 22:0.1 [’ — — (dashed line) a=0.01 — (solid line) a=0.001 1 1.5 2 2.5 3 3.5 4 R Figure 4.2: Numerical solution when R,- = 1, R0 = 4, w,- = 0 and 7110 = 1.5 for three values of a, a = 0.1, 0.01,0.001. Here [2,00 — 7,0,] = 0995me As 6 tends to zero we observe that graph of the g(R) moves upward meaning that smaller 67 gives rise to more angular displacement on the interior R,- < R < R0. In the third case we investigate the solution of (4.75) when R,- = 1, R0 = 4, 21),- = 0 and we = 3. Values c1 = 0.1, d = 0.01 and (r = 0.001 are used in Figure 4.3 and corresponding graphs are depicted by dotted, dashed and solid lines respectively. Note that 2120 — 1/2, = 3 > \Ilmax = 1.50826. Hence for this case 7,00 > ‘Ilmax. As suggested by the previous material model, 67 = 0, the graph of g(R) converges to a vertical line segment near 7‘ = R,- as Er -—> 0. In Figure 4.4 the previous three graphs are superimposed. For 7,120 > \Ilmax, as suggested by previous material model, (i = 0, the graph of g(R) converges to a vertical line segment near 7‘ = R,- as 67 -—> 0. 92 0 5 3' ----- (dotted line) a=0.1 . - ' — — (dashed line) G=0.01 — (solid line) a=0.001 1 .5 2 2.5 3 3.5 4 R ‘ Figure 4.3: Numerical solution when R,- = 1, R0 = 4, w, = 0 and 7,120 = 3 for three values of Er, G- = 0.1,0.01,0.001. Here [1170 — 42,] = 1.989‘1/max ----- (dotted line) a=0.1 — — (dashed line) a=0.01 — (solid line) a=0.001 Figure 4.4: Numerical solution when R, = 1, R0 = 4 and 212,- : 0 and 1110 = 3 for three values of (i, 6: = 0.1,0.01,0.001 and three values of 1120, $0 = 0.3\Ilmax, 0.995\Ilm, 1.989‘I’max 93 4.4 Regular Perturbation Solution for the Ex- tended Model In this section it will be shown for small 57 > O that a regular perturbation can be used to get an approximate solution of (4.4) subject to (2.17) for the case [2110 —- 1b,] < \Ilmax. If [1,00 — 70,-] > \IJmax the technique presented here does not apply. Further if [1170 — 4),] = (1 — e)\II.m,x with 0 < 6 << 1, i.e [1110 — 2b,] close to but less than \Ilmax, then severe requirements are placed on the smallness of Cr in order to obtain a useful regular perturbation solution approximation near R = R4. Let and Then equation (4.69) becomes FWMfl+mMfl+~J 6773(gf,(r) +5zg’1(r) +...) + \/4+T2(96(7‘) +c'i9’1(7‘)+---)2 By expanding in 6: equation (4.81) provides T3 96 3 r I 491 + a r + { , }] + O a? 4 + r2(g(’,)2 [9° (4 + r2607)” ( ) =4+a$+omw The 0(1) equation is r396 _ - 4 + 73(96)? 0' 94 mm) MW) (4.81) aw) aw) Note that by renaming the function and constant in the equation (4.83) it can be seen that (4.83) is identical with (3.113). Hence the solution is 4__ ’2 7‘ d0 J02 go(r) = arctan + 60. (4.84) Since attention is restricted to [1/20 — 112,] < \I/nm,x = arccos(R? R3) , it follows that do and 50 can be computed uniquely from the boundary conditions (2.17) as in Section 3.4. Indeed do and 50 are given by (viz. (3.118) and (3.119)) — R?R2 sin ‘11 d = . ° , 4.85 0 «R? + Rf, -- 2R3R3 cos ‘11 ( ) R4 — d2 50 = 21;,- — arctan 1 dz 0 . (4.86) 0 At this point an important observation is that if [we — 4),] > \I/mM then there is no 90(7') for which boundary conditions (2.17) are met. Hence the technique under consideration does not apply. Moreover from equation (4.84) 2 do I 7. = _ _ , 4.87 90( ) 1" mg ( ) and using (4.85) in equation (4.87) gives , 2R3, sin‘II 90(Ri) = (4-88) R,(R§ cos \IJ — RE)’ where ‘11 = [(00 — 112,]. Therefore 96(R4') —> 00 as [2120 -—¢,] —> ‘Ilmax. Hence for fixed (r if We — 71),] = (1 — e)\IIImam with 0 < 6 << 1 then the first term, 5173 g’ (r) , in equation (4.69) becomes large as r —-> 17,-. Thus the smallness of a for a useful approximation must correlate with the closeness of [1110 — $1] to \Ilmax, as discussed at the end of this section. Deferring these issues for the present, the 0(a) equation associated with (4.82) is 47.39! 3 I 1 T90+ 2 123/2 [4+r(go)l 95 which can be solved for g] , _ (J, - 7‘3 96) l4 + 7‘2(94’1)2l3/2 _ 4. 91 4 T3 ( 90) Substituting from (4.87) into (4.90) gives after some simplification J1 T3 2 (1.05 T' , = 2 [_.___ _ ____] 4.91 91 [7'4 _ ng/Q [7'4 —d2]2 ( ) Therefore 1r2 d ,/ 4 — 2 —J 2 , 91(7‘) = - 108 2+ ———-—°]— d1 T ”—2 or +51. (4.92) 2 —d0 7'4 — By using boundary conditions 91(R’i) = 0? 91(R0) = 07 (493) constants (11 and 51 can be determined uniquely: _ R3 + Jo 133+ 30 ‘10 R3 J" R? dl— (0.5 log RQ—do —0.5 log 123—620] 113—113 _R4_ :202) 1 1 __T_____=______ , 4.94 /(,/Rg—d3 «Rf-d3) ( ) and _ — 2 2 51: d1 _ d017,, _110g R +d_o] (4.95) \/Rg-d§ Rg-dd ng do Consequently the perturbation expansion solution of (4.4) subject to (2.17) with We — 111.] < \Ilmax is given by g(r) = 90(1) + 5491(1) + 0(642) (4.96) where 90(7), 91(7‘) are given by respectively (4.84) and (4.92). This procedure can be continued to any order Er" by further expansion of (4.82) to obtain a first order ODE for 9,, containing dn. Each such ODE for n > 1 is to be solved subject to gn(R,) = 9,,(Ro) = 0. Thus truncation of the perturbation expansion at order n 96 0.16 . . . . - 0.14- / 0.12- - 0.1 . 1 A / "I 030.08- / 0.06 - / 0.04 - —- — (dashed line) one-berm 0.02 ~ ----- (dotted line) two-term — (solid line) numeric 1 1.5 2 2.5 3 3.5 4 R Figure 4.5: One-term, two-term perturbation solutions and numeric solution with w,- = 0 and 1P0 = 0.150826 when 64 = 0.1. Here R,- = 1, R0 = 4 and \Ilm =1.50826 so that 11),, = 0.1\I'max results in an expansion that exactly satisfies the boundary conditions (2.17), but only solves the ODE (4.4) to 0(51"). Figure 4.5 shows the graph of one-term, two-term perturbation solutions and nu- meric solution with w,- = 0 and 1110 = 0.150826 2 0.1\II,,W when 61 = 0.1, R,- = 1, R0 = 4. These three graphs look like almost identical. However if any part of the graph is enlarged then it becomes apparent that two-term perturbation solution, shown by the dotted line, is better than one-term perturbation solution, shown by the dashed line, as expected from general perturbation theory. Figure 4.6 shows the graph of the one-term, the two-term perturbation solutions, and the numeric solution with 1b,- = 0 and do = 1.3 = 0.862‘1’max when 6 = 0.01, R,- = 1, R0 = 4. These three graphs again look like almost identical. However if any part of the graph is enlarged then it again becomes apparent that two-term perturbation solution, shown by the dotted line, is better than one-term perturbation solution, shown by the dashed line, as expected from general perturbation theory. 97 / / l A; — —— (dashed line) one-term 0 - ----- (dotted line) two-term -— (solid line) numeric 1 1.5 2 2.5 3 3.5 4 R Figure 4.6: One—term, two-term perturbation solutions and numeric solution with $4: 0 and 11),, =1.3 when 6: = 0.01. Here R,- =1, R0 = 4 and \II,,m =1.50826 so that 1110 = 0-862\I”ma.:1: As mentioned earlier the smallness of 67 for a useful approximation is expected to correlate with the closeness of [190 — 11),] to \Ilmx. Numerical investigation of (4.96) gives useful insight about this correlation. In particular we note that the correction gl(R) involves 91 ’(R) < 0. Thus the two term expansion for g’(R) will involve g’ (R) < 0 at some R if 54 is sufficiently large. This breakdown first occurs at R = R.-. For fixed R5, R0, 1/2, and 57 there is a critical value of 1&0, say \Ilcm, such that if 1120 2 \Ilcrit then 90,014) + €791,014) S 0, (4.97) which is contrary to the physical expectation that g(R,-) is an increasing function of 1%. Hence (4.97) provides an analytic expression to relate smallness of 64 with 1,00. Indeed by solving (4.97) numerically it has been found that \IJm, = 1.0887 2 0.722‘11max when R,- = 1, R0 = 4, 11’),- = 0, d = 0.1 and that ‘I’Cm = 1.301 = 0.862‘1’max when R,=1, R0 = 4, w,- = 0, d = 0.01. 98 0.8 - 0.6 ' ' i g(R) 0.4 - 02* — — (dashed line) one-term ~ ----- (dotted line) two-term — (solid line) numeric 1 1.5 2 2.5 3 3.5 4 R Figure 4.7: One-term, two-term perturbation solutions and numeric solution with w,- = 0 and we = 0.97 when 64 = 0.1. Here R,- =1, Re = 4 and \Ilem =1.50826 so that we = 0.643\Ilmee Since \Ilem = 1.0887 when R, = 1, Re 2 4, w,- = 0 and 57 = 0.1 the values of we = 0.97 and we = 1.3 are selected in the following graphs to emphasize the change in the nature of perturbation expansion. It can be seen from Figure 4.7 that the two-term perturbation solution, shown by the dotted line, is better than the one-term perturbation solution, shown by the dashed line, as expected from general perturbation theory. However as we increases through \I’em the value of the derivative g’(R,-) in the two term expansion becomes negative. For we = 1.3 this gives a graph for the two term expansion involving initially decreasing g(R). (see Figure 4.8). As a result the two—term , perturbation solution becomes a worse approximation to the numeric solution (shown by the solid line) than the one-term perturbation solution. On the other hand recall from Figure 4.6 that two—term perturbation solution is better than one-term perturbation solution when (r = 0.01 and we 2 1.3. 99 1.5 '. — — (dashed line) one-term i: ----- (dotted line) two-berm — (solid line) numeric 1.5 2 2.5 3 3.5 4 R Figure 4.8: One-term, two-term perturbation solutions and numeric solution with w.- = 0 and we = 1.3 when it = 0.1. Here R,- =1, Re = 4 and Warm = 1.50826 so that we = 0.862\Ilmex. Since we > \Ilem, the two term perturbation solution involves decrease of g(R) when R = R,- 4.5 Boundary Layer Type Solution for the Ex- tended Model In this section the case [we — w,] > \Ilmex will be examined by means of perturbation theory for small 64. This has the added benefit of clarifying the proper treatment of the case [we — w,] = (1 — e)\IImam with 0 < 6 << 1. As discussed at the end of previous section if [we — w,] > ‘I’max then there is no go(r) that satisfies the given boundary conditions so a regular perturbation treatment is not possible. Moreover if [we — w,] = (1 — e)\IIe,ax with 0 < 6 << 1 then the standard perturbation procedure does not yield a useful approximation if Er is not sufficiently small. To get a useful approximation in such cases we will proceed as follows: From (4.69) and (4.88) it follows that there is a region near 7' = R,- in which g(r) changes rapidly thus motivating a boundary layer type analysis. Indeed from (4.69) one can see that if g’(r) is of order 1/5 100 then the first term in (4.69) cannot be neglected while computing the first term of the perturbation expansion. i.e., there is a boundary layer at r = R,. A boundary layer correction (inner solution) at r = R,- will be constructed by using the stretching transformation (4.98) This inner solution will be denoted ((7). Note that away from 7‘ = R._-, the first term in (4.69) is small, whereupon a standard perturbation can be used to get an approximate form of the outer solution. This outer solution will be denoted C(r). However there will be only one boundary condition, namely g(Re) = we, for the outer solution to meet. The additional integration constant generated by the outer solution will be determined by patching the outer solution with the inner solution so that G000 = (00") (4-99) and go '(7‘) = (60"), (4-100) for some 7‘ E [R4, Re] where 00(7) is the first term of the outer solution and (g(r) is the first term of the inner solution. Indeed it will be shown that if 64 and d are suitably restricted, then f can be chosen uniquely. The geometric meaning of (4.99) and (4.100) is that the one—term outer and one-term inner solutions can be pieced up together smoothly; i.e., they have a common tangent at the point of contact. The terminology “patching” is used to indicate that the inner and outer solutions are reconciled at a single point, as opposed to some form of matching over a region [7], [8], [13]- We remark that the phenomena of interest are not associated with a singular perturbation in the usual sense. Namely, the small parameter 54 does not annihilate higher order derivatives of the governing equations in the limit 64 —’ 0. 101 If more terms are used in the outer and inner solutions, then patching can proceed as follows. For each new term that is added to the inner and outer solutions there will be one new integration constant. Now 7" and the constants (including the new one) can be chosen so that derivatives of the inner and outer solutions match up to one more order at f . For example if a two—term inner and outer expansions are used we require that (Co + 01W) = (C0 + (1)0) (4-101) and (90' + 91’)(7‘) = (<6 + (0(7‘) (4-102) and (90 ” + 91")(5) = ( 6' + (1')”), (4-103) where G1 and C1 are the second terms in the outer and inner solutions respectively. Note that f and all constants will be computed independently from any previous determination involving less terms in the expansions. In this sense this method is not iterative for computing patching constants. Technically, this process can be imple- mented for any order. The procedure is illustrated with the two term form of the outer solution C(r) = 00(7) + 6 G1(7‘) + C(62). (4.104) As usual we impose that 000%) = 1.00, (4.105) and 0,02.) = 0, j 2 1. (4.106) 102 The outer solution C(r) must satisfy (4.69). By using the expansions (4.80) (for d) and (4.104) in (4.69), the 0(1) equation is given by 3 I T 00 = d‘e. (4.107) 4/4 + (70;)2 Thus, r4 — 1,2 — Go(r) = arctan 82 + Co, (4.108) 0 where _ R4 _ '2 Co = we — arctan 0&2 0. (4.109) Hence it is found that the first term of the outer solution is given by (4108) up to constant do which will be computed by patching. Note for a given do that the maximal domain of the one—term outer solution, Go(r), is [\/d?o,oo). Moreover the derivative of Go(r) at r = \/d?o is infinite. Therefore as do varies by virtue of a change in boundary conditions, the domain of the first term of the outer solution also changes. Indeed by using (4.108) and (4.109) the leftmost endpoints of the one-term outer solutions has the following coordinates as do varies between R? and R3,: (7‘, g) = (\fdro we — arctan W ). (4.110) The graphs of the one-term outer solutions are depicted in Figure 4.9 for trial values do = 1, 2.25, 4, 6.25, 9, 12.25. Note that as the trial value do increases the graph of the one-term outer solutions move to the right. The dashed line gives (4.110) at Wthh point each one-term outer solution has a vertical tangent. The second term of the outer solution is now computed as follows. By using the expansions (4.80) and (4.104) in (4.69), the 0(64) equation is given by 40', _ 4?, G" + [4 + r2(G;,)2) 3/‘2 ‘ 5’ (4.111) 103 2.2 """"" |" """" r """"" r --------------- r ....... ..1 00 2r -------------------------- -/ ................... .. 18 -------------------- 7 -:- ----------------- * ------- *' /'/ ' 1.6 ---------- ,—-< .................................. _ F/ 1.4 ‘ 1 ‘ 1 . 1 15 2.5 3 35 4 I’ Figure 4.9: One-term outer solutions with R,- = 1, Re 2 4, w,- = 0 and we = 3 for the following six trial values do, do = l,2.25,4,6.25,9,12.25. Here do = 1 is on the far left and do = 12.25 is on the far right. The curves order themselves sequentially in do. The envelope of vertical tangency is shown as a dashed line which can be solved for G"1 , z (41 — Ga) [4 + ”(C-‘02P” 1 4 r3 ' (4.112) Computing G], from (4.108) and then substituting into (4.112) gives after some sim- plification d-1 7'3 2 (IO 1'5 G' = 2 [_____ _ _____—]. 4113‘ 1 [r4 — 4413/2 94 — 4812 ( ’ Therefore 2 - - ./T 275.4; 2 - (11(7) :1 1044]” +80 _ d1 7' 0 0’ +01, (4.114) 2 r2—do r4—dg where Cl is an integration constant and is determined on the basis of boundary condition (4.106) yielding (4.115) Note that (4.115) is the same as (4.95). Hence (4.114) is also same as (4.92). Hence two-term outer solution is given by 7‘4 al2 C(r) = arctan J? 0 + 00 (4.116) 0 1 2 d +é[-logl:—_—0 JI,/—_'r4_dg_aor2 2 r2—do r4—dg +Cll’ where 00 and Cl are given by (4.109) and (4.115) respectively. The boundary layer equation (inner equation) can be found by substituting (4.98) in (4.69) ((5) + \/4 + (R1. +2 8):? (<’<2>)2 ___ W (4.117) Let ((2) = (0(2) + 2 (1(7) + 0(22) (4118) be the form of the inner solution. Since d is a constant, the same value must apply to both inner and outer expansion. Thus d is again given by expansion (4.80). For the first term of the inner solution we set 51 = 0. The resulting equation is , 1 J (0(8) + E = 1%., (4.119) Thus, d— 1 (0(8) = (5‘;- — R7) 8 + 214-, (4.120) where the boundary condition g(R,) = (0(0) = 1,1),- is used to determine the integration constant. Note that by using (4.98) first term of the inner solution is given by Co“) = (1% “ i) (r 2310+ 1121. (4.121) 105 The second term of the inner solution is computed as follows. By using the expansions (4.80) and (4.118) in (4.117), the 0(a) equation is given by , d1 {id-03 s Q®=§—7?+E' (um Therefore d— d— 2 2 _ (1(8)— 18 3 03 5 +01. (4.123) — 71: ‘ TR? + fi— By using the boundary condition (1(0) = 0 it is found that CI = 0. Again by using (4.98) the second term of the inner solution can be written as C1(r) = (éiRg — 3%) (3:52—31)2 +21%, (T LR‘). (4.124) Consequently the two-term inner solution is given by ((2) = w. + (1% - i) (7" La) (4.125) _ 2 _ The patching procedure will be demonstrated for the one term case where ((r) is given by (0(1") in (4.121) and G(r) is given by Go(r) in (4.108) with Co given by (4.109). At issue then is the determination of the single constant do which appears in both (4.108) and (4.121). Indeed it will be shown that if d is sufficiently (small then do is uniquely determined by patching the one-term inner and outer solutions so that (4.99) and (4.100) hold for some 1" E [Rh R0]. To begin for given R47, R0, 1,0,- and 1,00, let a — 3 1 — —-1— (4 126) "m R? m mR1’ ' and R3 d” = m 2 (4.127) 106 Where m = (100 - W/(Ro - R1)- amaz given by (4.126) is a consequence of the requirement that G6(Re) = m which in turn provides a natural upper bound for a values. Note for an arbitrary we — w,- that amaa: given by (4.126) may not be positive. However if |we — 1M > \Ilmx or |we — wil = (1 — e)\Ilmex with 0 < 6 << 1, then amex > O. The above restriction on we — w.- once used in (4.127) also implies that R? g (1*. Finally it is easy to see from (4.127) that d" 3 R3. Therefore R? g d“ 3 R2 if |we — 1(le > (1 — e)\IImax for sufficiently small 6 > 0. The following claim establishes that do can be chosen uniquely so that one-term inner and outer solutions can be patched on the basis of (4.99), (4.100). Claim 4.5.1 If 0 < a _<_ amaz) then there exists a unique do 6 [R2, d‘] and a unique 1" E [R,-,Re] such that Go(7") = co(f) and go’('F) = (60‘) for this unique 1* e [R1, Re]. Proof of Claim 4.5.1: This claim will be proven in two steps. Step 1: For each do 6 [R2,d‘] there is a unique 7" E [\/d—_o,Re] such that 06(1‘) = Th, where in is the slope of the line passing through (Roi/21') and (7", Go(f)). Let do 6 [R2, d*] and let S : [\/d?o, Re] -—2 R“ be the function defined by 5(2) = G60“) (7‘ - R1) - 00(2) + 2124-, (4-128) where the dependence of Go(r) on do follows from (4. 108), (4.109). Equivalently Jor—( R1) . . S r): -——-——— we — ,- 4.129 < . T 2W —( w) ( > — arctan r4 — .3 + arctan R3 _ .0 63 402 ' Clearly S (r) is a continuous function for r E [Vt-1:, Re]. Since Go(\/dTo) —- w. and \/d-_—o — R,- are finite and lim er—o + go ’(7‘) = + 00, it follows that lim S(=) +00. (4.130) do 107 Also note that do S (1* gives lim S(r) = 230 (R0 — R2) — r—vRa R0‘/R3—dg Hence by the intermediate value theorem there exists 7" 6 [V do, Re] such that S (7‘) = (2120 — 212.) s 0. (4.131) O. This proves that there is a line passing through (R), wi); that is, tangent to Go (7') at (7‘, 00(7)) . Note that _—2do 3r4—d02 "' r2 (7.4_J02)3/2 5'0) (r —- R.) < O. (4.132) Hence S (r) is a strictly decreasing function. Thus 2? must be unique. Step 2: Let 0 < d g (1*. Then there exists a do 6 [R2, d‘] such that the one—term inner solution, (o('r) , coincides with the tangent line whose existence is guaranteed by step one. Moreover such a do is unique. We introduce two families of rays, each family of which is centered at (124,102)- The first family is the one—term inner solutions (4.120) where S is given by (4.98). The slope Bw/BR of any member of this family is m1(do) = agfla — i) (4.133) This family is parameterized by do. The second family consists of the rays passing through (Rafi/'11) that osculate with the one-term outer solution at r = 7“ E [V do, Re] (the existence of such a family is shown in step 1). This family is also parameterized by do. The slope Bw/BR of any member of this family is m2(0lo) = if]: + (4.134) 1 -4 _ ‘2 (arctan _ 0 — arctan ° 0 ), M2.- 4.2 .2 108 where r = r is point of contact. Let U : [R2, d‘] ——> R“ be the function defined by U(d0) = "MW—0) — m2(d_ol- (4-135) Clearly U (do) is a continuous function on [R2, d*]. Since m1(R,2) = 0 and m2(R,2) = 00 it follows that U(R,2) < 0. (4.136) Conversely . 1 d" 1 U(d ) _ E(F€§ —- E) — (4.137) 1 , F4 - (d*)2 [133— (d*)2 f _ Ri (we — wi + arctan W — arctan W). Since 1" ——> Re as do ——> d*, (4.137) gives It _ 1 d2 1 11/)O _ 1&1: U(d)—E(-fi§-E)—Ro_&, (4.138) or equivalently * l d* 1 where the inequality is a consequence of the restriction 0 < 67 3 am. Hence again by the intermediate value theorem there exists a do 6 [R2,d‘] for which U (do) = 0, i.e. m1(do) = m2(do). This proves that the one-term inner solution (o(r) and the one-term outer solution Go(r) are tangent to each other at (f, Go(f)) for this choice of do. Moreover since U’(do) = (4.140) ah? _ (‘77:173) (fi‘fl) >0. 109 L we [ a | f do 1.5 0.1 1.220537 1.259226 1.5 0.01 1.055384 1.057055 1.5 0.001 1.012634 1.012155 3 0.1 1.289941 1.613179 3 0.01 1.089118 1.180059 3 0.001 1.027634 1.055379 15.0826 0.1 1.713057 2.929204 15.0826 0.01 1.246533 1.553191 15.0826 0.001 1.080849 1.168164 Table 4.1: Numerically computed values of 7" and do for one term patching for several we and a values when R,- =1, Re=4 and 1111‘ =0 such a do must be unique. This completes the proof of Claim 4.5.1 demonstrating that do can be chosen uniquely so that one-term inner and outer solutions can be patched on the basis of (4.99), (4.100). D The Table 4.1 shows the values of r' and do for one term patching for several we and a values when R,- =1, Re = 4 and w, = 0. Figure 4.10 shows the one-term patched solution (dotted line), the one—term regu- lar perturbation solution (dashed line) and the numerical solution (solid line) corre- sponding to the values R,- = 1, Re = 4, w.- = O, we = 1.5 and 6: = 0.1. Note that \Ilmax z 1.50826; hence [we — 1M = 0.995\Ilmex. This suggests when |we -— w,| z \I'max that this new construction gives a satisfactory result even with patching only one-term inner and outer solutions. The patching procedure will be demonstrated for the two term case where C (r) is given by (4.125) and 0(7) is given by (4.116). At issue then is the determination of the constants do and d1, which appear in both (4.116) and (4.125), and the patching location 7" E [R,-, Re]. The constants do, d1 and 7" are computed numerically so that (4.101), (4.102) and (4.103) hold. 110 — (solid line) numeric — — (dashed line) one-term poMbalion ~ ----- (dotted line) one-term patched 1.5 2 2.5 R 3 3.5 4 Figure 4.10: Oneterm patched, one-term regular perturbation and numeric solutions when R,- =1, Re = 4, wi = 0, we =1.5 < \Ilmex % 1.50826 and (I = 0.1. The patching point is at r = 1.220537 [ we 1 a 72 do | all ] 1.5 0.1 1.114935 0.769605 6.607128 1.5 0.01 1.007296 0.932538 16.428066 1.5 0.001 1.000533 0.984481 35.746076 3 0.1 1.233646 1.202777 8.082971 3 0.01 1.081715 1.094041 18.477852 3 0.001 1.029278 1.042663 39.633449 15.0826 0.1 1.586166 2.056238 26.629071 15.0826 0.01 1.245361 1.442835 50.379749 15.0826 0.001 1.093927 1.173599 87.871407 Table 4.2: Numerically computed values of 1", do and d1 for two term patching for several we and E! values when R,- = 1, Re = 4 and w.- = 0 111 1.5 g(R) 0.5 - ' 1 J — (solid line) numeric ----- (dotted line) one-term patched — — (dashed line) two-term patched 1 1 .5 2 2.5 3 3.5 4 R Figure 4.11: One-term patched, two-term patched and numeric solutions when R,- = 1, Re = 4, w,- = O, we = 1.5 < \Ilm z 1.50826 and ('1 = 0.1 The patching point is at r = 1.220537 for the one-term patched solution and the patching point is at r = 1.114934699 for the two-term patched solution Table 4.2 shows the values of 1‘, do and d1 for two term patching for several we and d values when R,- =1, Re = 4 and 1121:: 0. Figure 4.11 shows the one-term patched solution (dotted line), the two-term patched solution (dashed line), and the numeric solution (solid line) corresponding to the values R.- =1, Re = 4, 1111: O, we =1.5 and ii = 0.1. It is clear from Figure 4.11 that two-term patched solution is better than the one-term patched solution. Figure 4.12 shows the one-term patched solution (dotted line), the two-term patched solution (dashed line), and the numeric solution (solid line) correspond- ing to the values R,- = 1, Re = 4, w,- = 0, we = 3 and (i = 0.1. Note that we = 3 > \IImax z 1.50826 and hence regular perturbation cannot be used. However two-term patched solution (dotted line) gives an excellent approximation to numerical solution (solid line). The graphs of three solutions are enlarged around the patching point for the one term patched solution. 112 0 5 _ — (solid line) numeric ' ' ----- (dotted line) one-term patched l — -— (dashed line) two-term patched 1 1 .5 2 2.5 3 3.5 4 R Figure 4.12: Two—term patched and numeric solution when R.- = 1, Re = 4, w. = 0, we = 3 > \Ilmee z 1.50826 and (1 = 0.1. The patching point is at r = 1.2336459 16 r . . . r . V '1 12» l ._ 2 ,/ 1o- : g(R) ONAOQ — (solid line) numeric ----- (dotted line) one-term patched — — (dashed line) two-term patched 1 1 .5 2 2.5 3 3.5 4 R Figure 4.13: One—term patched, two-term patched and numeric solution when R,- = 1, Re 2 4, w,- = O, we = 10\I1max and a = 0.01. The patching point is at r = 1.246533 for the one-term patched solution and the patching point is at r = 1.245361 for the two-term patched solution 113 f‘w‘r We close this section by depicting the graph of the one-term patched solution (dotted line), the two—term patched solution (dashed line), and the numeric solution (solid line) corresponding to the values R,- = 1, Re = 4, w, = 0, we = 10111mam and 6: = 0.01. The two-term patched solution (dashed line) gives an excellent approxi- mation to the numerical solution (solid line). The graphs of all three solutions are enlarged around the patching points. 4.6 Twist, Energy and Torque in the Extended Model The relation between energy and twist in the extended model can be obtained by using the numerical solution discussed in Section 4.3. Recall that the strain-energy density in the extended material model is given by (4.1). Hence when (37 = 07* the potential energy functional (3.4) becomes R0 A E = nh/ [0(11 — 3) + 61(1f + 11— 6)] r dr. (4.141) R,- By using 11 = 3 + (rg’)2 and If 2 i1 = 4+ (rg’)2 + 1 it follows that (4.141) becomes Re E = 71h / [a (rg')2 + 3(2 4 + (rg')2 — 4)] r dr. (4.142) R.- Letting a = 01/61, it follows from (4.142) that He ______ E = nhd/ [51(7‘9')2 + (2\/4 + (rg’)2 -— 4)] r dr. (4.143) R, Divide the interval [R4, Re] into N equal subintervals and label the end points as R,- = ro,r1, . . . ,rN = Re. For given 1/14, we and 64 the numerical solution presented in Section 4.3 provides the values of g, E g(rj). Once 93-, j = 0, . . . , N are known the centered difference approximation (4.71) can be used to estimate 93- _=-: g’(rj). Then 114 0| A Normalized Energy 1» u Amount of Twist Figure 4.14: Normalized energy E/(27rhd) versus twist when R, = 1, Re = 4 for the four values of 67, 5: = 0, 0.001, 0.01, 0.1 are depicted as the curves E1, E2, E3, E4 respectively (4.143) can be integrated numerically by using the standard composite Simpson’s rule. The normalized energy (energy /27rhd) is computed by the outlined method for several values of we and (1 when R; = 1, Re = 4 and w,- = 0. The normalized energy versus twist graph is depicted in the following figure for the values of 51 = 0.1, 0.01, 0.001. Note that since 51 = 0 corresponds to the material model discussed earlier in Part I it is possible to sketch the normalized energy versus twist and then compare with the extended material model. The energy versus twist graph is labeled as E1, E2, E3, E4 for 51 = O, 0.001, 0.01, 0.1 respectively. It is clear from Figure 4.14 that the stOred energr is an increasing function of 52 for a fixed amount of twist. This is to be expected since increasing 6: at fixed 61 represents an increase in energy penalization which would, in general, stiffen the material response. Similarly the relation between torque and twist in the extended material model can be obtained as follows. Torque M acting on the outer surface is given by virtue 115 2.5» ............. ............ .5 ............ 5 5 5 M4 . 2 ------------ 4 3 I I 1 E = - = 1315' "M --------- a Q E = = 1.112 E 11 ------------ :-- —r—#—T— fl 2° ; : M1 0.5 L' """ é ---------- A ------------------------- ..l 0 1 1 1 0 1 2 3 4 AmountofTwist Figure 4.15: Normalized torque M / (27rhd) versus twist when R,- = 1, Re = 4 for the four values of 64, d = 0,0.001,0.01, 0.1 are depicted as the curves M1,M2,M3,M4 respectively of (2.42), (4.2), (4.4) as 271' M = h / 1523094120) (19 = (4.144) 0 2n 2‘ I h R3 [5 Re g’(Re) + 2R” (12") ] O \/4 + (R09’(Ro))2 Equivalently 3 I M = 2mm R3 g’(Re) + R29 (R2) 1. (4.145) \/ 4 + (Rog’(Ro))"’ Since the torque acting on the outer surface R = Re depends on 9’ (Re), the value of g’ (Re) must be estimated numerically. By using the numerical solution given in Section 4.3 and one sided centered difference the value of 9’ (Re) is estimated for fixed R=R,-, R=Re, Er, w,=0, and we. The normalized torque (torque / 27rhci) versus twist graph is depicted in Figure 4.15 for the values of ii = 0.1, 0.01, 0.001. Note that since 51 = 0 corresponds to the 116 earlier material model it is possible to sketch normalized torque versus twist and then compare with the extended material model. The torque versus twist graph is labeled as M1, M2, M3, M4 for 54 = 0, 0.001, 0.01, 0.1 respectively. It is clear from the Figure 4.15 that torque is also an increasing function of 64 for a fixed amount of twist. 117 CHAPTER 5 Conclusions and Discussions In this concluding chapter the main work of this thesis is discussed and summarized. A preliminary discussion from the well known conventional theory of hyperelastic- ity is presented at the outset. Then the solution of the pure azimuthal shear problem in this context is presented in order to provide context for this thesis. In Chapter 3 the pure azimuthal shear problem is formulated in the new framework that is proposed by Pence and Tsai. This theory contains a material parameter k = 64/04“ where d can be viewed as a macroscopic stiffness and 04"“ can be viewed as a substructural stiffness. It is confirmed that if k —+ 0, then the pure azimuthal shear problem in the new framework reduces to the standard pure azimuthal shear problem for a neo-Hookean material. Since the governing field equations are too complicated for a general k, the pure azimuthal shear problem is studied for small I: by perturbation method and a numerical solution is presented for any k. The exact solution is also presented when k = 1 and a necessary restriction [\III _<_ \llme,x for a smooth solution is determined. A variational formulation of the problem is also presented and direct methods of the calculus of variations are used to investigate minimum energy configurations for arbitrarily large twist \II when k = 1. In Chapter 4 a more generalized material model is considered. It is characterized by an additional stiffness parameter oz. The first order nonlinear ODE governing the 118 pure azimuthal shear problem is derived by using equilibrium and internal balance equations. An existence and uniqueness result for the pure azimuthal shear problem for any 01* = d: > 0, a > O and total twist \II > 0 is obtained by using the integral equation and initial value problems corresponding to the governing nonlinear ODE. A numerical solution is presented for any 04/43 and regular perturbation is used to investigate solution possibilities when 04/61 is small for three qualitatively different ranges of \II: either \I' < \Ilmex or \I' > \Ilmex or \I1 = (1 — e)\IImex with 0 < 6 << 1. A boundary layer type analysis is used to get a useful approximation for the latter two ranges of \II. In particular using the properties of inner and outer solutions a patching method is proposed. This patching method gives satisfactory agreement with the numerical solution. The relation between torque, twist and energy in the extended material model is also studied. The pure azimuthal shear problem is complex in this new framework in either ma- terial model. The internal balance equation is a second order tensor equation. Even though it has been proved that the internal balance equation subject to requirement det E = 1 has a unique solution the proof is not fully constructive. The nonlin— ear governing equations offer many conceptual, analytical and numerical challenges. In this thesis we have advanced in all of these areas and examined in depth many interesting phenomena such as occurrence of singular surfaces, boundary layer type behavior and uniqueness and regularity of the solution. .119 Appendix A g(r) given by (3.115) is a strong local minimum of (3.168) if Consider the basic problem of calculus of variations: Find a piecewise—smooth function y(z) on the interval [a, b] that minimizes the definite integral b an = / F(4,y(4>.y' 44. (4.1) subject to W) = 4. y(t)) = 13. (A2) The following are well-known from the calculus of variations [28]: Theorem A.0.1 For a general Lagrangian F (11:,y, y’), the Legendre condition Flo/yr _>_ 0 for all :1: in [a, b] is a necessary condition for 3) to minimize J [y] Proof: See [28], page 88. Cl Let [[h] E / [Ryy(:r)h2(a:)+2Ryy1(:r)h(:r)h'(:r)+Rylyr(x)h'(x)2] dx. (A3) 120 Then it follows by direct calculation the Euler-Lagrange Equation associated with I [h] is [P(r) h'(a:>1' — 0(4) h(m) = o. (4.4) where A A A P(x) : Fan/(3’3): QC”) = F.yy($) " [FM/(17W: (A-5) The differential equation (A4) is called Jacobi ’s accessory equation. Definition A.0.2 A point e > a is said to be conjugate to point a relative to g(r) if there is a nontrivial extremal h(m) of I [h] which vanishes at c as well as at a, i.e., h(a) = h(c) = 0. Because P(r) and Q(:r) are defined for a particular g(r), conjugacy is relative to a given extremal g. Theorem A.0.3 Let J [y] attain a weak local minimum with the admissible extremal 37(2) and let the strengthened Legendre condition Rory: > 0 hold for all :1: in [a, b]. Then there are no points in (a, b) conjugate to point a relative to 3). Proof: See [28], page 91. II] Let E($7y7vvw) = F($,y,’lU) _ F(r,g,v) _ (”LU — v)F,yI(:I:,y,v). (A6) The quantity E (2:, y,v,w) is called the Weierstrass excess function. Theorem A.0.4 If, for the admissible extremal 1), J [g] is a strong local minimum for J [y] , it is necessary that E(x. 19(4). 17(1). w) _>_ 0. (A.7) for all x in [a, b] and for all [w] < 00. At a corner point, this equality is to hold for both 37+ and 3):. 121 Proof: See [28], page 116. D The strengthened form of the above necessary conditions are useful in developing sufficiency conditions for minimizing J [y]: (51) g(x) is a piecewise—smooth (admissible) extremal. i.e., it is a solution of Euler- Lagrange Equation where it is smooth and satisfies the Erdmann’s corner con- ditions at a corner point. (32) Strengthened Legendre Condition: F yryr > 0 on [a, b]. (33) Strengthened Jacobi Condition: y(x) has no conjugate points to a in (a, b]. (54) Strengthened Weierstrass Condition: E(x,u,v,w) = F(x,u,w) — F(x,u,v) — (w — v)Fyr(x,u,v) Z 0 for all (u,v) near (9,37) and all [w] < 00. Theorem A.0.5 If F is C4 in its arguments and the admissible extremal g(x) sat— isfies (51), (s2), (s3), and (34), then J[y] is a strong local minimum. Proof: See [28], page 128. C] Using the above theorem it will be proven that g(r) given by (3.115) is a strong local minimum of (3.168) if |\II| g \IJmex. For (3.168) the Lagrangian F is F(r,g,g’) = r\/4 + (rg’)2 — 2r. (A.8) It is clear that F is C4 in its arguments. By using the Erdmann’s Comer Conditions (Fe: and F — 32’ F e: must be continuous even at the corner points of 3)(x)), it will be shown that any extremal of (3.168) must be smooth; i.e., there is no corner point. From (A.8) 8F _ r3g’ 5? — \/4+ (rg’)2' (A.9) 122 Let p = g; and q = g’_ be the values of g’ at the left and right side of a corner point respectively. Then by the first Erdmann’s corner condition 3 3 _23_ = __2'9__, (4.10) (/4+ (rp)2 \/4-l—(rq)2 Since r 71$ 0, P q . ___ = —, A.11 \/4+r2p2 «4+qu2 ( ) By taking the squares of both sides in (All) 2 2 P _ q 4+r2p2 — 4+r2q2’ (A'IZ) which leads to p2 = q2. (A.13) If (A.13) is used in (All), then (A.13) simplifies to p = q. (A.14) Therefore any extremal of (3.168) must be smooth. Note that in general the second Erdmann’s corner condition is necessary to find the values of p and q. But in this case it is not necessary to check the second Erdmann’s comer condition. In Section 3.4 it was shown that g(r) given by (3.115) is the only solution of the Euler-Lagrange Equation if |\Il| S ‘Ilmax when k = 1. Therefore (31) is satisfied. From (A.9) it follows that 62F 4r3 _ = _ .1 Since r > 0, on [R1, Re] (A.15) gives that r2F 4 3 d—— = r > 0. (A.16) 89’? (4 + 7.29/2)3/2 Hence (s2) is satisfied. Note from (A.16) that an extremal g(r) must be at least C2 by Hilbert ’5 Theorem. 123 Jacobi’s accessory equation (A.4) for the current problem may be written as 4r3 , (4 + 7.2942)3/2 h (T) = 0,1, (A.17) where a1 is a constant and h(r) is a nontrivial function that vanishes at both c and R,; i.e., ME) = h(c) = 0. From (A.17) it follows (11 (4 + r2g’2)3/2 h’(T) = 4r, (A.18) Consequently (3.114) in (A.18) yields h'(r) = ___204173' , (A.19) (.4 — 402/2 where d1 is the constant given by (3.118). Hence h(r) = 7:43: + 42, (A20) where a2 is another constant. By using h(R,) = 0 from (A.20) it follows that 01 a Z . A21 2 44742 ‘ ’ Whereupon (A.20) can be written as 1 1 h r = a —— — —— . A.22 (> .( .___R:,_d, ,__T,_d,) < ) Now if h(r) = 0, then either a1 = 0 giving h(r) E 0 or 1 1 . _— = ___, A.23 4—42—42 472—41 ( ) giving r = Re. Therefore there is no conjugate points to R,- in (Re, Re]. Consequently (33) is satisfied. One important remark is that the notion of being conjugate is defined relative to an extremal. Since there is no admissible extremal when [‘11] > \I/max the outlined procedure cannot be used in that case. 124 It remains to show that Strengthened Weierstrass Condition (54) is also satisfied. For the current problem the Weierstrass excess function E (A.6) becomes 3 E(r,u, v, w) = r\/4 + (rw)2 — “/4 + (rv)2 — (w - v) r v (A24) 4 + (rv)2, or equivalently E(r,u,v,w) = r(\/m — %). (A.25) Note that (v — w)2 2 0, (A26) for all |w| < 00. Then (A.26) can be written as v2 + w2 2 2vw. (A27) Multiplying both sides of (A27) with 4r2(> 0) yields (4r2)v2 + (4r2)w2 _>_ (4r2)2vw. (A.28) Adding 16 + r4v2w2 to both sides of (A28) leads to 16 + 4r2v2 + 4r2w2 + r4v2w2 Z 16 + 8r2vw + r4v2w2. (A.29) Note that (A.29) can be written as (4 + r2142) (4 + r2102) 2 (4 + r2vw)2, (A.30) or equivalently (4 + r2102) > (4 + r2vw)2 _ m7- (A.31) By taking square root of both sides of (A31) 125 (4 + r2vw) 4 .1. y 2 > ___—___, A.32 (ru) '— (/4+ (rv)2 ( ) or equivalently (4 + r2vw) 2 _ __ 4 + (rw) 4 + (rv)2 _ 0 (A33) giving that E(r,u,v,w) Z 0 for all [w] < 00. This completes the proof of (34). Consequently g(r) given by (3.115) is a strong local minimum of (3.168) if [\IJI S \Ilmax. Cl 126 Appendix B Alternate Proof of Theorem 4.2.3 In Section 4.2 we presented a proof of the Theorem 4.2.3 based on maximal solution intervals for first order initial value problems. Here we present an alternative proof based on best existence and uniqueness result for first boundary value problem for second order nonlinear ODEs. By eliminating d, equation (4.4) can be written as 9"(7‘) + K239201045!) = 0, (31) where " 2 l2 2‘" 9 ] (B2) + 4d+a[4+r2g’2]3/2 Lemma B.0.6 f(r,g’, 62,61) is a Lipschitzian in g’ on U E [R,-, Re] x R X R+ x R+. i.e., there exists a nonnegative constant L such that |f(r.gi.a.c‘r) -f(r.9§.a.d)| .<_ L lgi -9£|, (33) for every (r,g[,a,oz) and (r,g§,a,(3z) in U. Proof: Note that if f (r, g’ ,a,d) has bounded partial derivative 8 f /8g’ in U then it is a Lipschitzian with lipschitz constant is given by , . L = sup 8f(r, 9 ,la, a) . (B4) (r,g’,a,d) E U 89 127 Now (8.2) gives af(7',gl,01,51) _ 3 +24arg’2[a+a ‘/4+,.2 9:2] 89’ _ ;+ [4d +a(4+r29’2)3/2]2 ' (B.5) Note from (B.5) that 8 f /dg’ is an even function of g’ and 8 f 3 9' 'Blzlleoo— 89’- _. (B6) By using the limit in (8.6) and the fact that 8 f / 69’ is continuous on U it can be proven that 8 f / 89’ is bounded on U. For a given a and 61, it is enough to find the supremum of (9 f / 89’ over the infinite strip 9 = {(r,sf)lr e [Ri,Rola g’ e (-oo,oo)}- For each fixed r E [Ra R0] and e > 0, there exists N > 0 such that 3f 3 a—g’—;| <6 (13.7) whenever |g’l > N, or equivalently §_ (if r €<6_g’<— 3+6 (B.8) whenever |g’| > N. Since 7" E [R4,Ro], 3 8f 3 E—é N. Let m1 = |(3/Ro )——e| and m2: (3/R.) +6. Then (B. 9) implies that whenever lg’ I > N and r 6 [R13 R0], where M1 = max{m1,m2}. Consider the region = {(7‘.g’)|7‘ E [R;,Ro], g' E [—N,N]}. Since 90 is a compact set and df/Bg’ is a continuous function, there exists a positive number M2 such that 128 whenever (r, g') E wo. Let L = max{M1,M2}. Then affix 9’, a, d) 09’ < L (8.12) for every (r, g’ ,a, d) in U. Thus f(r,g’,oz,d) is a Lipschitzian in g’ on U. This completes the proof of Lemma B.0.6 E1 Lemma B.0.7 Consider the following differential equation w”(r) + §w'(r) = 0, (B.13) subject to W(Ra') = 211,- > 0, w'(R,-) 2 mi > 0, (B.14) where L is the lipschitz constant given by Lemma 3.1 and /\ is a positive real number less than 1. Then the solution w(r) of (B.13), (3.14) and its derivative w’ (r) are strictly positive functions on [Rh 00). Proof: Clearly the solution of (B.13) satisfying (B.14) is given by i A i ._ w(r) = 112,: + A: ~— Tm eL(R‘ r)”. (B.15) Note that w’(r) = m,- eL(R‘—T)/A > 0 (BIG) for every r E [a,oo). Since w’(r) > 0 on [IL-,oo) and w(Rl-) = wi > 0, it follows that w(r) > 0 on [R,~,oo). Hence w(r) and w’ (r) are strictly positive functions on [R4, 00). This completes the proof of Lemma B.0.7. [:1 Proof of Theorem 4.2.3: Note that the second boundary value problem (B.1), (4.15) is equivalent to follow- ing integral equation Ra g(r) 2 l(r) +/ H(r,s) f(s,g'(s)) ds for R.- _<_ r S R0, (B.17) R4. 129 and R0 g’(r) = mo + R.- H,(r, s) f(s,g'(s)) ds, (B.18) where l(r) = 1%“ + mo (r — R4), (B.19) and H (r, s) is the Green’s function given by _ . -< < < H(r,s)= (3 R“) E’s—LR) (B20) (T_R1-) ILST'SSSROa and H( )—3H ) (B21) TT38 —87' (T’S' . Note that both H and H, are nonnegative functions. Let S be the space of continuously differentiable functions on [R,-, R0] with the norm llg(r)|l=max{ max 'gm' max @933}, (3.22) 3.51530 w(r) ’ 3.95120 w’(r) where w(r) is the positive weight function given by Lemma B.0.7. The space S is complete with this norm, and convergence in the norm of gn(r) to g(r) implies uniform convergence of both gn(r) to g(r) and g;,(r) to g’ (r) Let T be the operator defined by the right side of (B.17). In order to see that T maps S into S, we need to verify that Tg(r) is continuously differentiable whenever g(r) is. Observe that r — r R" iii?) Tg( +3 Tg( ) =mo+/& Hr(r,s) f(s,g’(s))ds, (B23) and that this limit function is continuous in r. To verify that T is a contraction mapping, consider |T91(7') " T920“ wb") )I—i—Rorss’s—s’ss Swm/R. Ha )|f( ,gl< >> f( ,92( ))Id . (13.24) 130 By using the Lipschitz condition (B3) equation (B24) becomes lT91(7‘)‘T92(7‘)| < 1 MT) _ w(r) R0 [34 H(r,s) L was) —g;l ds (B25) or equivalently |T91(r) - T92(r)| 1 R" , lgi(8) - gé(8)| w(r) g w(r) [Rt H(r,s) Lw (s) w’(s) ds, (B26) and hence by (B22) ITQIU") — T92(T)| < (B 27) “1(7") — ' 1 3° H(r, 8) L w'(8) 613 “910") - 920‘)”- w(r) R. Note that since w(r) is a solution of (B.13) with w(R,-) > O and w’(R,) > 0, it is also a solution of the following integral equation Ra L w(r) = w(R,) + w'(R,-) (r — R.) + H(r, s) X w'(s) ds. (B28) R.- Since w(R,-) + w'(R,-) (r — 12,-) 2 O, for every r E [R1330] equation (B28) implies Ra L w(r) 2 / H(r,s) — w’(s) ds, (B29) R.- A and since w(r) and /\ are positive, inequality (B29) may be written as 1r) Ana H(r, s) L w'(s) ds 3 A. (B30) w Consequently substitution from (B30) into (B27) gives |T91(7") - T92(7‘)| w(7‘) S A ”910") - 92(7")||- (B31) Similarly from (B.18) we obtain |%(T91(7‘) - Tg2(r))| w’(r) g (13.32) R, $5 [R- mas) magic» — f(s,gé(8))| ds, 131 which by virtue of (B3) and (B22) yields Isagm - T920)» < 1 R0 , f Hras) Lw (s) ds “gm — 92m“. Now (B28) supplies R0 L w’(r) = w’(R,-) +/ Hr(r, s) — w'(s) ds. R,- A Since w’ (1%,) > 0 for every r E [R4, R0], equation (B34) implies 30 L w’(r) > / H,(r,s) — w’(s) ds. R; A Since 211’ (r) and A are positive, inequality (B35) may be written as 1 fRoH (r s)Lw’(s)ds R,- > O. This completes the proof of Theorem 4.2.3. [I] 132 BIBLIOGRAPHY 133 BIBLIOGRAPHY [1] Akhiezer, N.I., The Calculus of Variation, Blaisdell Publishing Company, New York-London, 1962 [2] Antman, S.S., Nonlinear Problems of Elasticity, Applied Mathematical Sciences, 107, Springer Verlag, New York, 1995 [3] Atkin, R.J. and Fox N., An Introduction to the Theory of Elasticity, Longman Mathematical Texts, London, 1980 [4] Bailey, P.B., Shampine, LP. and Waltman, P.E., Nonlinear Two Point. Boundary Value Problems, New Academic Press, New York-London, 1968 [5] Ball, J .M., Discontinuous equilibrium solutions and cavitation in nonlinear elas- ticity, Philosophical Transactions of the Royal Society of London. Series A, (1982) vol. 306, pp. 557-610 [6] Beatty, M.F. and Jiang, Q., On Compressible Materials Capable of Sustaining Axisymmetric Shear Deformations. Part 2: Rotational Shear of Isotropic Hy- perelastic Materials, Quarterly Journal of Mechanics and Applied Mathematics, (1997) vol. 50, pp. 211-237 [7] Bender, CM. and Orszag, S.A., Advanced Mathematical Methods for Scientists and Engineers, McGraw-Hill Book Company, New York, 1978 [8] Bush, A.W., Perturbation Methods for Engineers and Scientists, CRC Press, Boca Raton, 1992 [9] Capriz, G., Continua with microstructure, Springer Tracts in Natural Philosophy, 35, Springer Verlag, New York, 1989 [10] Ericksen, J .L., Liquid crystals with variable degree of orientation, Archive for Rational Mechanics and Analysis, (1991) vol. 113, pp. 97-120 134 [11] Fu, Y.B. and Ogden, R.W., Nonlinear Elastic Deformations, Cambridge Univer- sity Press, Cambridge, 2001 [12] Gurtin, M.E., Configurational forces as basic concepts of continuum physics, Applied Mathematical Sciences, 137, Springer Verlag, New York, 2000 [13] Holmes, M.H., Introduction to Perturbation Methods, Texts in Applied Mathe- matics, 20, Springer Verlag, New York, 1995 [14] Horgan CO. and Polignone, D.A., Cavitation in nonlinear elastic solids: a re- view, Applied Mechanical Rewiev, (1995) vol. 48, pp. 471-485 [15] J iang, X. and Ogden, R.W., On Azimuthal Shear of a Circular Cylindrical Tube of Compressible Elastic Material, Quarterly Journal of Mechanics and Applied Mathematics, (1998) vol. 51, pp. 143-158 [16] Johnson, L.W. and Riess, R.D., Numerical Analysis, Addison-Wesley Publishing Company, London, 1982 [17] Keller, H.B., Numerical Solution of Two Point Boundary Value Problems, Society for Industrial and Applied Mathematics, Philadelphia, 1976 [18] Mariano, P.M., Multifield theories in mechanics of solids, Adv. Appl. Mech, (2001) vol. 38, pp. 1-93 [19] Muller S. and Spector, S.J., An existence theory for nonlinear elasticity that allows for cavitation, Archive for Rational Mechanics and Analysis, (1995) vol. 131, pp. 1-66 [20] Murray, F.J. and Miller, K. S., Existence Theorems for Ordinary Differential Equations, Robert E. Krieger Publishing Company, New York, 1976 [21] Ogden, R.W., Nonlinear Elasticity Theory and Applications, Dover Publications, New York, 1997 [22] Pence, T.J. and T sai, H., An Extended Theory of Elastic Material Response Un- der Prescribed Swelling, Research notes (2003) with publications in preparation [23] Pence, T.J. and Tsai, H., On the cavitation of a swollen compressible sphere in finite elasticity, International Journal of Non-Linear Mechanics, (2005) vol. 40, pp. 307-321 135 [24] Rivlin, R.S., Large Elastic Deformations of Isotropic Materials. VI Fhrther Re- sults in the Theory of Torsion, Shear and Flexure, Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences, (1949) vol. 242, pp. 173-195 [25] Shampine, L.F., Kierzenka J. and Reichelt M.W., Solving Boundary Value Prob— lems for Ordinary Differential Equations in MATLAB with bvp4c, available at ftp: / / ftp.mathworks.com.pub / doc / papers / bvp / 2000 [26] Tao, L., Rajagopal, K.R. and Wineman, A.S., Circular Shearing and Torsion of Generalized neo-Hookean Materials, IMA Journal of Applied Mathematics, (1992) vol. 48, pp. 23-37 [27] Triantafyllidis, N. and Bardenhagen, S., On higher order gradient continuum theories in 1-D nonlinear elasticity. Derivation from and comparison to the cor- responding discrete models, Journal of Elasticity, (1993) vol. 33, pp. 259-293 [28] Wan, F .Y.M., Introduction to the Calculus of Variations and Its Applications, Chapman and Hall, New York, 1995 136 [I[[[l]][i[[l[][j[[[[[[171]]