‘ [wag 2;;- ‘ fig‘gc‘gfiag %* 4 ._ ‘ . ’_ qf'r‘zfii“ - v ‘1 : V @%§g§;5$*=‘#= ‘ ‘ ‘ i‘t Fit ‘3 v ‘2 '2 s" e“ 1 fififikf KP” LA I . . . “ha 5 ~: > ‘ '; .J.‘ . '2, “f. - 4 2:. ' .Ai. " 7. ' . .t‘égfiflt"9§ . I ' -‘ ‘61 I , I rfiZp» V f ' , “fi-‘fit‘égggi E t 7 ‘ ~. " W . . ., " ‘ h. 3f. I , “3 ' ‘ -~ '1 . 4' “' '5 " fl . .u '1 . ‘ '43 ‘r-‘ -. H ”7. ,. : .1 , . - , é-rk _ ‘ 1% . hi; . «wig, N 2 . *‘ ‘ -‘ - ., .. '1‘ . . .. . , . f, a W _ gfiflg .' ’ ‘ ~ > .V 13? ', ' , 5w; ,‘ f“ 2.. j. Wfifi‘é‘ \ .. I <. .8 :3 7 ‘ E ' h! l '1 .' I I” {3.5; “ ~ ‘ ‘v gl‘:%;k§§? A ‘ . . ~. ‘ 1' ‘ ."iau‘u."‘f: 'i" ' ‘I‘I' ‘ A ‘ , ”1 . “5. ~ . «g ‘ ‘A . «at; A" I , w 33",“ l'iaygfififi mi ‘zizftw % r :ss’t‘ ; n‘. W; ‘ ',-""'1'5.‘T‘f m :g‘ _-;-$§:;;Yg,, 5', {22.15% . J -.A-:w ..,.:v‘\..‘..- . J3? flak 1%. f 2 ”dirk: {inuc ' a; r‘ v I ‘ 31‘- 318:5 " 'Wf-‘u . 5% M3 k 2. ‘ : ' e . L :, '~ ‘. j {'1 ,.' .* aha; . ‘ ‘ 1 ~ ~ r’ , ~. 4w 3 , , ~‘ ‘ ‘73 1.". i1. Eb 33“.: ".5 ~. , J: : , - '» L: ‘ T ‘- ~ .. fig; «. I "22% _ g Jig, ' g1. .v V‘JIIS' ‘. w . , * ~ -« «gm A, 6 ”‘3‘ "I: ' a?!” . ‘ A ‘ 1‘17 . . . . ,’ “535.) ‘ ”1‘ 4 . . _ . L ‘ 5:,» . u »‘ . > V 2' ' .. - ' ‘1 1 I 3' . 4“ C~ ":éh‘fibv ‘2} ,7 I téfifirfiu" MtCHIGAN STATE UNTEF‘SITY LIBRARIES NW \“HWWHH um NM W i H Hi, i 3 1293 01399 4417 \ LIBRARY Mlchlgan $tate University THESS This is to certify that the thesis entitled FINITE ELEMENT SIMULATION OF SWAGING IN THIN-WALLED CYLINDERS presented by Eric James Leaman has been accepted towards fulfillment of the requirements for Master's degree in Mechanics Mam Major professor Date March 28, 1995 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution CE II RETURN BOX to mow-rhi- chockoutm your mood. Motown duo. FLA To AVOID FtNES mum on or W‘W MSU |IM W WHO p.301 FINITE ELEMENT SIMULATION OF SWAGIN G IN THIN-WALLED CYLINDERS BY Eric James Leaman A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Materials Science and Mechanics 1995 ABSTRACT FINITE ELEMENT SIMULATION OF SWAGIN G IN THIN-WALLED CYLINDERS By Eric James Leaman The sheet metal forming operation of swaging can produce wrinkles in a thin- walled cylinder. In order to alleviate wrinkling, an understanding of the deformations and stress state was necessary to isolate those variables controlling the formation of wrinkles. In this investigation, a finite element model for simulating the swaging process was devel- oped using the commercial finite element code MARC. Revealed in the analysis was the presence of tensile circumferential stresses on the inner surface of the cylinder in the wrin- kling region. These stresses, in conjunction with the transcending of critical circumferen- tial strain levels were hypothesized to be responsible for wrinkling. A proposal was made to increase the cylinder wall’s effective moment of inertia by the introduction of a circum- ferential stiffening rib. ACKNOWLEDGEMENTS I would like to acknowledge the Faculty and Students of the Department of Mate- rials Science and Mechanics for their guidance in completing this project. In particular, I would like to thank Mr. Mike Wood and Dr. James Lucas for their contribution of the mechanical properties, and also Mr. Yue (Jeffery) Qiu for drawing Figure 1.1 of the text. I would like to acknowledge the financial support of the United States Can Company under grant 61-8502. 1 would like to thank the personnel at MARC Analysis Research Corpora- tion for their assistance in using their product. I must, also extend my gratitude to my research advisor, Dr. Ronald C. Averill, for his guidance and support throughout this project. Although simple words are inadequate, I must say thank you to my wife Laurie for all of her encouragement, support, and sacrifices that she has made during the comple- tion of this work. Table of Contents List of Tables vii List of Figures x Chapter 1 Introduction 1 1.0 Introduction 1 1.1 Problem Statement 5 1.2 Literature Review 7 1.3 The Present Study 25 Chapter 2 Modelling of the Swaging Operation 28 2.0 Introduction 28 2.1 Geometric Model of Cylinder and Forming Tools 29 2.1.1 Preliminary Element Selection 33 2.1.2 Evaluation of Element Performance for a Linear Problem 35 2.1.3 Discussion of Results for a Linear Problem 37 2.1.4 Nonlinear Evaluation of Element Performance 43 2.2 Boundary Conditions and Contact Constraints 45 2.2.1 Boundary Conditions 45 2.2.2 Initial Stresses 47 2.2.3 Contact Conditions 50 iv 2.3 Material Model 2.3.1 Mechanical Properties 2.3.2 Plasticity Model 2.4 Solution Parameters and Analysis Options 2.4.1 Solution Method 2.4.2 Analysis Options Chapter 3 Numerical Simulation of Swaging 3.0 Introduction 3.1 Simulation of the Swaging Process 3.2 Final Element Selection 3.2.1 Forming Analysis Results Using Element 1 3.2.2 Forming Analysis Results Using Element 116 3.3 Convergence of Forming Problem 3.3.1 Mesh Refinement Study 3.3.2 Discussion of Mesh Convergence Study Data 3.3.3 Comparison of Stress Trends Resulting from Mesh Refinement 3.3.4 Convergence Study on Number of Load Increments 3.4 Results from Forming Simulation 3.4.1 Strain and Stress States in the Critical Region 3.4.2 Assessment of Strain Hardening and Initial Stress Effects on Strain and Stress States Chapter 4 Discussion of Wrinkling 4.0 Introduction 52 52 54 57 57 59 60 60 61 66 67 67 69 71 73 81 91 91 91 93 102 102 4.1 Existence of Critical Strain 4.2 Bending Effects 4.3 Introduction of Stiffening Rib Chapter 5 Summary and Future Work 5.1 Summary 5.2 Future Work APPENDIX A] Mesh Refinement Study Data A.2 Analysis Results without Strain Hardening REFERENCES vi 103 106 108 122 122 123 128 137 142 Table 2.1 Table 2.2 Table 2.3 Table 2.4 Table 2.5 Table 2.6 Table 2.7 Table 3.1 Table 3.2 List of Tables MARC Elements Predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius=l.0, thickness=0.l, and length=100.0. All values are normalized to the closed form solution. Predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius=5.0, thick- ness=0.05, and length=300.0. All values are normalized to the closed form solution. Predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius=100.0, thickness=l.0, and length=200.0. All values are normalized to the closed form solu- tion. Nonlinear test case results corresponding to the predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius: 100.0, thickness=1.0, and length=100.0. All values are normalized to the closed form solu- tion. Mechanical Properties of Double-Reduced Tinplate Steel, Type L. Stress versus Plastic Strain for Double-Reduced Tinplate Sheet Steel, Type L. Percentage Change in Axial Strain Values in the Critical Region. (%) Percentage Change in Circumferential Strain Values in the Critical Region. (%) vii 34 38 39 42 53 56 76 76 Table 3.3 Table 3.4 Table 3.5 Table 3.6 Table A.1 Table A2 Table A3 Table A4 Table A5 Table A6 Table A7 Table A8 Table A9 Table A. 10 Table A. 11 Stress Values for 250 vs. 500 incs for the First Stage in the Critical Region. Stress Values for 250 vs. 500 incs for the Second Stage in the Critical Region. Effects of Strain Hardening on the Global Minimum and Maximum Strain and Stress Values. Effects of Initial Stresses on the Global Minimum and Maximum Strain and Stress Values. Axial Strain on the Outer Surface in the First Bending Region. (x10’3) Axial Strain on the Mid-Plane Surface in the First Bending Region. (x10'3) Axial Strain on the Inner Surface in the First Bending Region. (x10’3) Circumferential Strain on the Outer Surface in the First Bending Region. (x10‘3) Circumferential Strain on the Mid-Plane Surface in the First Bending Region. (x10’3) Circumferential Strain on the Inner Surface in the First Bending Region. (x10’3) Axial Stress on the Outer Surface in the First Bending Region. (szi) Axial Stress on the Mid-Plane Surface in the First Bending Region. (szi) Axial Stress on the Inner Surface in the First Bending Region. (szi) Circumferential Stress on the Outer Surface in the First Bending Region. (szi) Circumferential Stress on the Mid-Plane Surface in the First Bending Region. (szi) viii 92 92 94 95 128 128 129 129 129 130 130 130 131 131 131 Table A. 12 Table A. 13 Table A. 14 Table A. 15 Table A. 16 Table A. 17 Table A. 18 Table A. 19 Table A.20 Table A.21 Table A.22 Table A.23 Table A.24 Circumferential Stress on the Inner Surface in the First Bending Region. (szi) Axial Strains on the Outer Surface in the Critical Region. (x10'3) Axial Strains on the Mid-Plane Surface in the Critical Region. (x10'3) Axial Strains on the Inner Surface in the Critical Region. (x10’3) Circumferential Strains on the Outer Surface in the Critical Region. (x10‘3) Hoop Strains on the Mid-Plane Surface in the Critical Region. (x10’3) Hoop Strains on the Inner Surface in the Critical Region. (x10—3) Axial Stresses on the Outer Surface in the Critical Region. (szi) Axial Stresses on the Mid-Plane Surface in the Critical Region. (szi) Axial Stresses on the Inner Surface in the Critical Region. (szi) Circumferential Stresses on the Outer Surface in the Critical Region. (szi) Circumferential Stresses on the Mid-Plane Surface in the Critical Region. (szi) Circumferential Stresses on the Inner Surface in the Critical Region. (szi) 132 132 132 133 133 133 134 134 134 135 135 135 136 Figure 1.1 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 Figure 2.7 List of Figures Wrinkling Behavior Exhibited in Swaged Cylinder. Relationship Between Modelled Surface and Full Cylinder. Geometry of Forming Tools with Reference to the Cylinder Axis. (Not to scale) MENTAT Model of Cylinder and Forming Tools. Geometric Model with Applied Boundary Conditions. Cross Section of Cylinder with Stress Resulting from Pure Bending. Stress versus Plastic Strain for Double Reduced Tinplate Sheet Steel. Newton-Raphson Solution Method. Figure 3.1(a) First Stage of Swaging at the Zero‘h Increment. Figure 3.1(b) First Stage of Swaging at the 50th Increment. Figure 3.1(c) First Stage of Swaging at the 100th Increment. Figure 3.1(d) First Stage of Swaging at the 250th Increment. Figure 3.1(e) Release of the First Stage Tool. Figure 3.1(0 Tool Moved into Position in First Increment of Second Stage. Figure 3.1(g) Second Stage of Swaging at the 352nd Increment. Figure 3.1(h) Final Increment of the Second Stage. 30 31 32 46 48 56 58 62 62 63 63 65 65 Figure 3.1(i) Release of Second Stage Tool. Figure 3.2 Figure 3.3 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 Figure 3.9 Figure 3.10 Figure 3.11 Figure 3.12 Figure 3.13 Figure 3.14 Figure 3.15 Figure 3.16 Figure 3.17 Figure 3.18 Swaging Analysis Using Element 116 with a 150x2 Mesh. Sampling Locations for Convergence Study of Forming Analysis. Axial Strain on the Outer Surface in the Critical Region vs. Mesh Size Circumferential Stress on the Outer Surface in the Critical Region vs. Mesh Size. Axial Strain on the Inner Surface in the Critical Region vs. Mesh Size. Circumferential Strain on the Inner Surface in the Critical Region vs. Mesh Size. Axial Stress on the Outer Surface in the Critical Region vs. Mesh Size. Circumferential Stress on the Outer Surface in the Critical Region vs. Mesh Size. Axial Stress on the Inner Surface in the Critical Region vs. Mesh Size. Circumferential Stress on Inner Surface in the Critical Region vs. Mesh Size. Computational Time vs. Mesh Size. Axial Stress in the Critical Region for the Second Stage of the 150x8 Mesh. Circumferential Stress in the Critical Region for the Second Stage of the 150x8 Mesh. Axial Stress in the Critical Region for the Second Stage of the 300x8 Mesh. Circumferential Stress in the Critical Region for the Second Stage of the 300x8 Mesh. Axial Stress in the Critical Region for the Second Stage of the 150x2 Mesh. Circumferential Stress in the Critical Region for the Second Stage of the 150x2 Mesh xi 66 68 74 74 75 75 78 78 79 79 80 83 84 85 86 87 Figure 3.19 Figure 3.20 Figure 3.21 Figure 3.22 Figure 3.23 Figure 3.24 Figure 3.25 Figure 3.26 Figure 3.27 Figure 3.28 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.8 Figure 4.9 Figure 4.10 Deformations of the 800x2 Mesh at Increment 303. Deformations of the 150x2 Mesh at Increment 303. Deformations for the 800x2 Mesh at Increment 373. Deformations for the 150x2 Mesh at Increment 373. Axial Strains in the Critical Region at the 503rd Increment. Circumferential Strains in the Critical Region at the 503rd Increment. Axial Stresses in the Critical Region at the 503rd Increment. Circumferential Stresses in the Critical Region at the 503rd Increment. Axial Stresses in Critical Region at 503rd Increment, Calculated without Initial Stresses. Circumferential Stresses in the Critical Region at the 503rd Increment, Calculated without Initial Stresses. Euler Buckling Column. Shell Element Subject to Bending and Resultant Moments. Circumferential Stress versus Strains for the First Stage of Forming. Simple Beam Subjected to a Bending Moment. Plate Subjected to a Bending Moment. Initial Geometry of Cylinder with Added Stiffening Rib. Formed Shape of the Cylinder with Added Stiffening Rib. Final Shape of the Cylinder with Added Stiffening Rib after Elastic Snapback. Axial Strains in the Critical Region of the Cylinder with Added Stiffening Rib. Circumferential Strains in the Critical Region of the Cylinder with Added Stiffening Rib. xii 89 89 90 9O 96 97 98 99 100 101 104 104 109 114 114 115 116 117 118 119 Figure 4.11 Figure 4.12 Figure A.1 Figure A.2 Figure A.3 Figure A.4 Axial Stresses in the Critical Region for the Cylinder with Added Stiffening Rib. Circumferential Stresses in the Critical Region for the Cylinder with Added Stiffening Rib. Axial Strains in the Critical Region Without Work Hardening. Circumferential Strains in the Critical Region Without Work Hardening. Axial Stresses in the Critical Region Without Work Hardening. Circumferential Stresses in the Critical Region Without Work Hardening. xiii 120 121 138 139 140 141 Chapter 1 Introduction 1.0 Introduction The current thrust in the manufacturing community is toward the conservation of resources and energy, while at the same time maintaining product quality and responding to the demands of the consumer. This philosophy has its roots in the latter days of the stone age and has evolved over the next six thousand years to its present state. The pri- mary concern of early man was simply to obtain food to eat. The tools necessary to kill were fashioned by chipping flint into weapons. Eventually flint tools began to take on spe- cific roles, such as axes, knives, or borers. Toward the end of the Stone Age man discov- ered the ease with which some materials could be worked into shape and metal forming was born. A couple of millennia later casting was found to make the working of metals 2 more efficient by eliminating some of the hammering necessary, but, because the castings were of copper, a soft material, it was still advantageous to use stone. Bronze improved the utility of cast tools, but shortcomings were still present. For instance, bronze Chisels could not cut stone. Thus a new material was needed and iron was found to satisfy those needs, giving us the iron age (Parr, 1967). In each of these cases, societal needs for improved tools and weapons required new materials and methods of production to be developed. Many discoveries were made by pure luck and others were the result of invest- ments in manpower, raw material, and time. The experiences of those working in metal forming built up over time and eventually a group of artisans was established. The needs of the United States government in the late 18th century made it neces- sary to obtain a large stand of arms. The time and the gunsmiths necessary to fill the need was scarce. Eli Whitney was driven by these two circumstances to develop a new system of manufacturing whereby he would “form the tools so the tools themselves shall fashion the work and give to every part its just proportion - which when accomplished will give expedition, uniformity, and exactness to the whole” (Green, 1956). The process Whitney was proposing was currently being used to produce goods, but not on the scale of com- plexity that the musket required. The lock on a musket required an exact fit of the parts in order to work. This being the case, the standard of the day called for crude parts to be made and then assembled by a craftsman who filed and fitted the parts together to produce the working lock, a very time consuming task. Whitney’s plan would eliminate the need for filing, and introduce the system of interchangeable parts. This evolution in manufac- turing was consistent with the goals of speeding up the manufacturing process, reducing the labor necessary, and improving the quality in musket production. But the system still 3 required craftsmen to assemble the parts and tools needed to do the work. These two diffi- culties prevented the delivery of the arms on time, but the system was successful nonethe- less. Finally, Henry Ford brought mass production to its present form. He recognized the need for a low cost, high quality vehicle that almost anyone could afford. To do so required the marriage of five components: precision, standardization, interchangeability, synchronization, and continuity. While all of these existed individually, Ford brought them all together in the moving assembly line where the work moved to the worker who per- formed a single operation. Production costs were once again driven downward by this new manufacturing concept and the joys of driving were delivered to a wider audience. Thus the advent of mass production had arrived on the scene, with its roots stemming from Whitney’s system of interchangeability. However, elaborate planning, expensive tooling, and exact synchronization of the parts moving through the plant were still required. While each of these evolutionary steps has brought about greater efficiency and reduced cost in manufacturing, societal requirements are for even greater cost reductions, quicker product development times, and environmental impact awareness. To accomplish these mandates manufacturers have turned to a concept known as simultaneous or concur- rent engineering. This represents a fundamental change in the manner in which products are designed and manufactured. Previously a sequential process was followed. A need was defined, a part designed to meet the need, a means of manufacturing the part devised, and the process was tested. Rarely, though, is the cycle correct the first time. One or more of the components in the design cycle may have to be altered, thus forcing a retrial of the process. Trial and error is expensive and not consistent with the current thrust of conserva- tion. Concurrent engineering is “a more co-operative, simultaneous, computer-aided” approach to product development (Butman, 1991). Design and manufacturing engineers are voicing their concerns to each other at the initial stages. This parallel approach gives rise to a more efficient design cycle, but methods of analyzing the integrity of the part and the manufacturing process are also essential. Although trial and error methods served this function in the past, the decline of skilled tradesmen and the need for rapid results render this scheme insufficient. What is needed is' a means to do the analysis without the physical product or process. Numerical methods are readily adept at performing this function. Numerical methods consist of building mathematical models that are representa- tive of the product and tools of the process. In this way testing is done on the computer rather than on the plant floor.The finite element method is one such numerical scheme available for the analysis of continuous structures that are currently being manufactured. The approach of the method is to model the structure by discretizing the geometry with a group of elements. The equations that govern the behavior of each of these individual ele- ments are then assembled to produce a system of algebraic equations of the form Ki = F (1.1) where E is the stiffness matrix of the entire structure and is a function of the elemental deformation modes and the material properties, 35 is a vector of the displacements and rota- tions (or the degrees of freedom, DOF) that describe the deformations of the elements, and [—7 is a vector of the forces at discrete points within each element. By substituting the applied loading along with the DOE which are fixed on the boundary into Equation 1.1, 5 the unknown displacements, rotations, and forces may be solved for in each element and thus the global response of the structure may be predicted. The results obtained, however, are rarely exact because of the approximations necessary to manipulate the governing equations of the elements into the form of Equation 1.1. Even so, solutions accurate for engineering purposes can usually be obtained, though the large number of equations nec- essary to obtain theses results may demand the use of large scale computing machinery. In the above paragraph it was assumed that the stiffness matrix, 7?, and the force vector, 7‘, were independent of the displacement vector, 3. In many practical situations this is not true and the analysis becomes nonlinear. Such is the case when analyzing a manu- facturing process. One such operation is sheet metal forming where a set of rigid tools introduces plastic deformations into a relatively simple geometry to produce the more complex configuration of the final product. The analysis of such an operation is nonlinear because of the large deformations required, the relationship between stress and strain, and the contact of the rigid tools and the blank. Not only does this add to the complexity of the finite element analysis, but it also renders nearly impossible any analytical solution. Therefore, the means to analyze sheet metal forming are either the trial and error method, which has been shown to be costly and contrary to the requirements of simultaneous engi- neering, or the development of computer models where part and process evaluation may be done for relatively little cost on the computer. 1.1 Problem Statement A typical example of a metal forming operation is swaging or necking wherein the diameter of a cylinder is reduced by the action of a set of rigid dies. When using this pro- Figure 1.1 Wrinkling Behavior Exhibited in Swaged Cylinder. cess on thin-walled cylinders the possibility then exists for the formation of wrinkles in the necked-in region, as shown in Figure 1.1. The wrinkles are the result of a build up of compressive circumferential stresses in the cylinder. When these stresses exceed some critical value, any disturbance, whether due to material imperfections or the manufactur- ing process, will cause the cylinder to buckle (or bifurcate) to an adjacent equilibrium position (the wrinkled configuration). One possible solution to the problem is to increase the thickness of the cylinder wall. However, arbitrary use of thicker material increases the cost associated with the purchasing of raw materials which in the end is passed on to the consumer or absorbed by the manufacturer, neither of which is desirable. In addition, other solutions may exist that will eliminate the misuse of resources. In other words, the increased wall thickness places greater demands on the environment both in terms of raw materials for steel production and waste management at the end of product life. However 7 an investigation of the stress state and kinematics of the cylinder during the swaging pro- cess may lead to an alternate solution that is consistent with the goal of conserving resources and energy. Trial and error techniques coupled with the die makers experience will not produce the desired information, and this also inherently conflicts with the current thrust in manufacturing as already explained. No analytical means of evaluating the pro- cess is available. The nonlinearities due to large deflections, plasticity, and contact between the cylinder and tools associated with the. problem render the solution analyti- cally intractable. Thus, a finite element simulation of the necking process is sought in order to achieve the stated objectives. With the stress states and deformations of the cylin- der in hand, changes, whether in the material, the forming path, and/or the final geometry, with an engineering basis may be proposed to eliminate the wrinkling phenomena. 1.2 Literature Review Early attempts to predict the success of a forming operation centered on the use of forming limit diagrams (FLD). The formation of the FLD is through experimental proce- dures wherein the sheet blank is etched with a grid pattern of circles. The blank is then stretched over an unlubricated punch and the deformation of the circles in the region of failure are observed. The shape of the etchings will have changed from circular to ellipti- cal. The major strain is calculated by measuring the major axis of the ellipse and the minor strain from the minor axis. After a series of tests the delineation between a successful for- mation and one that fails is established (Kalpakjian, 1991). While experimental construc- tion of FLD’s has been successful, numerical prediction of the curves requires a failure criterion. Toh (1989) proposed a method of predicting the forming limit curves of sheet 8 materials using an incremental rigid-plastic finite element method and a new limit crite- rion. In the analysis of a hemispherical punch stretching a circular blank with various cir- cular cutoffs and various friction conditions at the tool-sheet interface, the material was assumed to obey Hill’s anisotropic yield criterion and its associated flow rule. The result- ing set of load versus displacement curves were then used to derive a critical slope condi- tion which could subsequently be used as a criterion to determine the corresponding critical major and minor strains. This approach has been shown to give reasonable predic- tion of the FLD’s, but this is only a method of predicting the failure of the material. To understand the underlying behavior of a structure, simulation of the forming process is necessary. Simulation requires a sophisticated finite element formulation, especially concem- ing those components giving rise to non-linearities. These include the large deformations that are usually introduced to the workpiece. The changes in the configuration of the workpiece should be permanent and as such the material experiences plastic deformations, another source of nonlinearity. Finally, the deformations are produced by a tool or set of tools that come into contact with the workpiece, thus introducing nonlinear boundary con- ditions into the analysis. While many considerations are necessary in the analysis of metal forming operations, these components draw the most attention. Keck, et a1. (1990) summarized results obtained for various sheet metal forming operations that were simulated using an implicit finite element code. In general, a variety of element types are used, but all are formulated based on the principle of virtual displace- ments and an updated Lagrangian description for the incremental calculation of the defor- mation process. 9 Use was made of the rate form of the elastic-plastic material model. However with the assumption of small strains, a hypoelastic constitutive law may be implied, because the strains may be decomposed into their elastic and plastic components. This required a yield function for the material which, in general, was assumed to conform to von Mises yield criterion. But in the case of a membrane element formulation, Hill’s quadratic yield criterion was used, thereby taking into account the normal anisotropy. The rate form requires the integration of the constitutive relationship which was performed using either a radial return method or a return mapping algorithm. The treatment of the changing contact conditions of the problem were handled in the following manner: The dies were treated as rigid bodies and in order to fulfill the con- tact conditions a penalty formulation was adopted. Penetration of the tool was checked only at nodes that were specified as contact nodes. Then a penalty factor was calculated for each possible contact node from the stiffness of the workpiece perpendicular to the tool surface and scaled by a user defined factor. The resulting contact friction was mod- elled with a modified Coulomb friction model. This overcame the problem of singularities arising in the first derivative of the Coulomb friction model at vanishing relative motions and enables the linearization of the friction law in connection with the Newton-Raphson iteration scheme. The equilibrium of each incremental loading of the structure was obtained by a dynamic solution scheme. In general, the N ewton-Raphson iteration method was used, but in cases where the rate of convergence reached some predetermined value, the so called BFGS method (named for Broyden Fletcher, Goldfarb, and Shanno, see for example, Zienkiewicz, 1989) was employed. The Newton-Raphson method was reintroduced upon a deceleration of the convergence rate. 10 The finite element formulation proposed was applied to the analysis of hydrostatic bulging of a circular sheet of annealed aluminum, hydrostatic bulging of rectangular sheets of anisotropic mild sheet steel, an axisymmetric stretch forming process, and the deep drawing of a rectangular cup. In each case the results showed good agreement with previous experimental work. The authors point out, however, that the formulation pre- sented is limited to materials which have accurately identified constitutive models and material parameters. Emphasis is thus placed on the accurate description of the material behavior through experimental techniques before proceeding with more complex analyses than those performed. A more complicated analysis was conducted by Lee, et a1. (1991) on a stretch- drawing process particularly used in the automotive industry. The aim was to develop a finite element method which resulted in acceptable computational times, while retaining the ability to model the key features of a metal forming analysis. Chief among these being the contact algorithm which is generally regarded as a major issue for numerical stability. As such the authors implemented a membrane line element consistent with a shell formu- lation for the contact formulation. Additionally, to aid in the reduction of computational time, the number of degrees of freedom within the shell elements was reduced. The model was formulated using an updated Lagrangian description of motion. In addition, the shells used to describe the sheet metal were assumed to be thin and therefore, the through the thickness shear deformations could be neglected. “With a full bending the- ory being considered, the strain expressions contained the second derivative of the incre- mental in-plane displacements. The degrees of freedom therefore consisted of the change in axial displacement, the change in transverse deflection, and their first derivatives. But, 11 by assuming a linear variation of the incremental axial displacement, its derivative was no longer required, resulting in a net reduction in the number of DOF in the model. There- fore, the interpolation functions used in the finite element formulation were linear for inplane displacements and Hermite cubic for the transverse deflection. These interpolation functions were then used in the principle of virtual work, which when linearized along with a geometric constraint for tool contact resulted in the stiffness matrices and force vectors due to inertia, contact forces, and internal reactions. When evaluating the resulting algebraic equations the through the thickness integration was reduced to single point quadrature rule in order to reduce computational time, while retaining the key features of the bending model. In regard to the material model of the formulation, the response was assumed to be rigid visco-plastic so that incompressibility could be applied using small natural strains. In addition, normal anisotropy was taken into account through the use of Hill’s yield crite- rion. Finally, a power law hardening relation with strain rate sensitivity of the form a = oo+K(a0+eo+A£)"(é/y)m (1.2) was implemented, where 00 is the yield stress, K is the material strength parameter, a0 is the pre-strain, n is the strain hardening exponent, m is the strain rate sensitivity index, and Y is the base strain rate. Using this expression a wider variety of sheet metals could be considered. Finally, the detection of contact between the tool and the workpiece was indicated by satisfaction of the following, 12 gnEn-( —x3) = 0 , (13) where n is the normal between the tool and the workpiece and x" and .15s are position vec— tors of a point on the sheet and it nearest point on the tool. However, because 1:" and x‘ are unknowns the equation must be linearized and solved iteratively. This was a general gap condition. In the event that the tool moved only in the vertical direction, a vertical gap parameter would be used. In either case, when the tool and sheet metal were in contact external forces would be transmitted through the tool to the sheet metal and therefore a consistent contact formulation was required. The methods used above show good agreement with other numerical results as well as with experimental data. However, it was found that when the strain reaches the n value of the strain hardening equation (1.2), the model failed to converge. This difficulty was due to a loss of definiteness in the stiffness matrix and is termed softening. Rebolo, et a1. (1990) are members of a team developing a general robust finite ele- ment code for the analysis of sheet metal forming problems. As such the techniques which they employ have been proven to be effective in the solution of these types of analyses. The focus of the discussion was again toward the element types used, the modelling of the contact condition, and the frictional model used in the metal forming analyses. Little con- sideration was given to the constitutive model as it appeared to the authors that the con- ventional elasto—plastic models together with small strain linear elasticity, a smooth flow surface, and isotropic hardening provided adequate representation of the material behavior 13 during forming. In discussing the types of elements which are available, the authors point out that the use of continuum elements is prohibitively expensive. These computational difficulties can be overcome by the use of structural elements, such as membrane or shell elements, but these elements introduce the inability to handle two-sided contact due to their lack of ability to accommodate thickness stresses. In contrast, the introduction of assumed strain and stress elements allow for the advantages of structural elements, while at the same time providing for two-sided contact. The formulation presented was based on a membrane element which employed finite strain plasticity. Generally this type of formulation would require the introduction of an additional rotation tensor to account for the stresses and strains. However, a more direct method was offered by calculating the strain measures from the polar decomposition of the incremental deformation gradient, the incremental rotation vector, and the incremental plastic strain. The stress was then a function of the elastic strain. The advantage of this for- mulation was the ease in which it may be implemented in an updated Lagrangian frame- work. A simplified approach to the detection of contact and the use Lagrange multiplier techniques to impose the no penetration constraint were employed. A measure of overclo- sure, quite similar to that proposed by Lee, et al., was formulated by calculating the dis- tance along the normal from the closest point on the rigid surface to the contact node on the deformable body. When the overclosure is greater than or equal to zero, contact is detected and a Lagrange multiplier was introduced which provided the contact pressure. Finally, when contact was identified a Coulomb friction model was employed to 14 account for the sticking or slipping stresses generated within the model. Again this was done by introducing Lagrange multipliers. In this situation, however, the parameters could take on different roles. In the event that sheet metal and the die stuck together the Lagrange multiplier was used to enforce the constraint that the shear strain was zero. Con- versely for a sliding condition, the Lagrangian took on the value of the frictional stress. Within each of these formulations the technique used for the detection of contact was carried out in similar fashions. A normal vectorfrom the tool(s) to the workpiece was constructed and when the magnitude of the vector goes to zero contact was assumed. Once contact has been realized, a constraint must be added to the governing equations to pre- vent penetration of the tool into the material. In general this constraint was enforced in one of two manners, either a Lagrange multiplier was used or a penalty function parameter was used. The Lagrange multiplier enforces the contact condition exactly, but at the expense of additional DOF in the solution. Also, a singular stiffness matrix can result from this formulation. Both of these difficulties can be avoided by using the penalty function, but satisfaction of the contact condition is not exact and the solution is highly sensitive to the choice of the penalty factor. In each case the contact constraint is enforced in a “point- wise” sense. A method of enforcing the contact constraint in an average sense over the element boundary was proposed by Simo and Taylor (1985). This work was conducted within the context of bilinear isoparametric interpolation of the displacement field, thereby allowing the assumption that the contact pressure was constant over each contact segment, and dis- continuous across segments. The contact constraint was therefore enforced in an average sense over the contact segment, rather than in a pointwise manner. This resulted in the 15 average gap being a kinematic variable. The introduction of the average contact constraint was accomplished by using a perturbed Lagrangian formulation of the total potential energy functional. By defining the gap function in the usual manner, a modified total potential energy functional for the con- strained system may be written as 1 2 2 A A 1 2 ”8(9 ,LI ,A) = A2111 (9 )‘i’Ag-z—EA (1.4) where the first term on the right hand side is the potential energy, the second term is the contribution of the Lagrangian constraint, and the last term has the form of a penalty term. As this penalty term goes to infinity, 8 -* co, the standard Lagrangian formulation is recov- ered. The real value in adding the additional ‘penalty term’ to the functional was that the possibility of a non-positive definite stiffness matrix was eliminated. The perturbed Lagrangian functional could then, after taking the first variation, be used to formulate a finite element solution to the contact problem. Yet the perturbed Lagrangian formulation, as with the other classical approaches, is not without faults. In each method only those constraints that were active during the iteration contributed to the incremental equations. Undesirable consequences may arise from this approach. For example, the quadratic convergence of the Newton-Raphson method may be seriously affected if there are frequent changes in the active set. By introducing a method in which all the inequality constraints are enforced directly by means of additional equations together with the equilibrium equations to form a system of nonlinear equations in the displacements and the contact tractions, the New- 16 ton-Raphson method may be used to overcome these difficulties. Just such an algorithm was introduced by Eterovic and Bathe (1991). The goal was to derive a set of finite ele- ment equations that would explicitly account for the presence of the contact conditions. This was done by formulating the virtual work for a continuous body in a total Lagrangian fashion. To this statement was added the virtual work done by the contact tractions over the virtual relative displacements between the contacting bodies. A second governing equation was derived from the contact conditions. Thus, using the two equations in con- junction resulted in a complete variational formulation of the motion. These equations were then cast into a finite element model which contained n+4m unknowns, where n is the number of nodal dof and m is the number of contact nodes. But the elegance of the model was that all the inequality constraints associated with the contact conditions were contained within the finite element equations. But this gives rise to a highly nonlinear set of equations, and therefore a robust solution technique was paramount. Riccobono (1992) presented a comparison of two methods which are used to ban- dle the nonlinearities presented by the plasticity model in metal forming problems. The first was a matrix method proposed by Prager and Hodge in which the actual strain rate field of the analysis was the one that satisfied the following: 2 1. . J = 7—560! /§eijeinV— ITividS (1.5) v s, where S is the surface and T,- are components of the surface traction. In this formulation, the incompressibility condition is enforced using Lagrange multipliers. Thus the problem 17 is defined as finding the minimum to a functional which combines (1.5) and the Lagrangian constraint of the incompressibility condition. The second method used the infinitesimal strain increment form of (1.5). Lineariz- ing the yield surface by a set of tangent planes and using Gaussian numerical integration, the functional was written as, 4 n, n, J = Jic 2 2115.5]. 2 (1).,- 2 Fidui (1.6) j=I i=1 i=1 where for the axisymmetric state, 21cerJ. is the volume associated with the integration point j, and u,- the number of the nodes on which the forces F ,- are applied. Thus, using this new functional along with two sets of restraints (one which states that the infinitesimal strain increments derived by the velocity field equals the one derived by imposing the nor- mality law at each integration point and the other stating that the die never penetrates the blank) the linearization of the yield function is complete. A comparison of the two methods revealed that the first formulation is based on an iterative process and uses a non-linear system of equations, while, on the other hand, the second formulation is given in a closed form and thus eliminates any chance of conver- gence problems. This also meant that the matrix method may allow the tool to penetrate the blank during a solution step, but the linearization technique enforces the non-penetra- tion restraint explicitly. Therefore the linearization method may provide superior perfor- mance to the matrix method, but the increased number of unknowns associated with the formulation may indeed cancel any advantages gained. Finite element models are usually formulated for fully three dimensional or possi- 18 bly two dimensional spaces when assumptions of plane stress or plane strain are made. This may lead to a prohibitively expensive analysis in the case of a nonlinear metal form- ing problem. In cases when the geometry, properties, and loading are constant in one direction, for example an axisymmetric problem, the model may be simplified by model- ling the mechanical fields in that coordinate with a traditional, trigonometric Fourier series. In general this method fails to satisfy the anisotropic form of the governing equa- tions for laminated composite structures. Padovan (1974) developed a quasi-analytic finite element procedure for axisymmetric anisotropic shells and solids by using the exact func- tional representation for the circumferential variable dependency. In shell theory for anisotropic materials the appearance of the shearing terms in the stiffness matrix cause the traditional, trigonometric Fourier series approach to break down when trying to obtain a solution for the equilibrium equations. One possible resolution is to transform the equilibrium equations into an orthotropic form, but the analytical solution is intractable due to the distortion of the boundary conditions. Therefore, it was proposed to cast the displacement vector into a complex Fourier series expansion. Applying the new formulation for the displacement and force vectors in the equilibrium equations yielded a complex differential equation with the real and imaginary parts of the displacement vector being coupled for the fully anisotrOpic case. The difficulty in obtaining an analytical solu- tion lead to the development of a finite element solution. The formulation of the finite ele- ment equations was the same as a traditional approach with the exception that the displacement and force vectors were in the complex, Fourier expansion form. This lead to a complex stiffness matrix for solution. The procedure outlined was tested against problems with known solutions and it was found that the complex form of the equations gave very good approximations in all 19 cases. Given the accuracy of the results and the difficulties in using the traditional Fourier series, the complex Fourier series representation appeared to have some advantages, espe- cially in that the significant effects of material anisotropy were revealed by its use. Two constraints on the use of the semi-analytic, or finite strip method are that the geometry of solids of revolution be axisymmetric and that the material properties in the circumferential direction be constant. To broaden the scope of the method Sedaghat and Hermann (1983) proposed relaxing the first constraint and removing the second altogether. The formulation started with the removal of the circumferential or theta dependence of the total potential energy of the system. This was done by expanding the displacements and body forces with a trigonometric Fourier expansion. However the elastic coefficient matrix may also be theta dependent. This difficulty was handled by decomposing this matrix into two components called the base properties and the deviation properties respec- tively. The base properties are theta independent and are in most cases found by averaging the material properties in theta coordinate. The deviation properties are simply the devia- tion from the base values. From the decomposed elastic matrix the stress-strain relationship was in the form of (IN = De”+be”” (1.7) where the first term was independent of theta and yielded the usual uncoupled equations and the second term was theta dependent and contributed to the load matrix in a manner similar to initial stresses. The finite element model could then be constructed from the 20 total potential energy of the system. The variation of the material properties aided in the formulation of the semi-ana- lytic method for non-axisymmetric geometries. By idealizing the geometry and then vary- ing the material properties a semi-analytic formulation could be brought about. For example a semi-infinite plate with a hole loaded by internal pressure was idealized by a circular plate of radius p, where p was chosen such that the value of stress in the x direc- tion was small. If the hole is near the top edge of the plate, the edge effects would be accounted for by varying the properties within the idealized plate. Results from a finite element model formulated using the idealized geometry and the varying material proper- ties were in good agreement with the closed form solution for the problem. Thus it appears that the scope of the Fourier series formulation had been broadened. Attention will now be turned to analyses involving the formation of wrinkles in sheet metals. Currently the most popular method of predicting wrinkle failures is the Yoshida Buckling test. However, this method, as well as other empirical techniques, have proved to be inadequate for observed trends. In particular the local curvature of sheet metal during forming has been shown to have a significant effect on the conditions for wrinkling, but the empirical methods fail to account for this parameter. Neale and Tugcu ( 1990) concentrated on some of the basic features of wrinkle for- mation, in particular those occurring within the context of the plastic buckling theory for thin plates and shells. Wrinkling can be viewed as a plastic buckling process in which a wavelength of the buckles in one direction is very short. This buckling is local and depen- dent on the local curvature, the thickness of the sheet, on the material properties, and the local stress state. 21 The basic element considered had constant radii of curvature R1 and R2 and con- stant thickness t, the state of stress was assumed to be a uniform membrane state (no bend- ing), and the investigation was limited to those regions of the sheet that were not in contact with the die. Simplifications arose by exploiting the fact that the short-wavelength wrinkling modes were shallow and could be analyzed by the Donnell—Mushtari-Vlasov (DMV) shallow shell theory, which restricted the analysis to modes where the characteris- tic wavelength was large in comparison to the thickness, but small compared to the local radii of curvature. To determine the critical stress state for buckling, a so called “bifurcation func- tional” was developed. For all admissable displacement fields the condition that the sec- ond variation of the bifurcation functional is greater than zero ensures buckling does not occur. However, buckling becomes possible when the bifurcation functional is zero for some non-zero field. The analysis was done by considering three fields in the bifurcation functional and then integrating over some local region, S, which was separated from the rest of the sheet. Because the integration is carried out over a local region, the boundary or continuity conditions are relatively unimportant. The bifurcation model may also be written as a function of the buckling displace- ments and a coefficient matrix, M. When the determinant of the matrix M goes to zero, buckling was again predicted. Also, by minimizing the determinant with respect to some waveform parameters and setting it equal to zero, the critical stresses associated with buckling could be found. Because of the complexity of the equations, closed-form solu- tions to the problem are difficult to obtain, therefore numerical solutions are generated by means of the Newton—Raphson technique. This gave rise to an understanding of how the material flowed in the plastic regime. The most widely used constitutive relations in buck— 22 ling are based on the J2 flow theory or the J2 deformation theory of plasticity. In addition, the material behavior has been assumed to be isotropic and the power-law hardening rule used. A parametric study was done to determine the influence of material properties and geometry of the critical stress state for local buckling. Material constant values of mild steel and low-hardening steel were used and the results were plotted in terms of principal compressive stresses and angle between the principal stress axes and the principal axes of curvature. When comparing the flow theory results to the deformation theory results, the deformation theory gave a more conservative prediction for the onset of wrinkling than that of the flow theory and as a result was the preferred theory. Also, the results showed that the critical stresses for the onset of wrinkling decreased as the strain hardening expo- nent decreased and the critical stresses were also shown to decrease with the thickness of the sheet. Chan (1993) presented a method for locating the bifurcation points on the general- ized force-displacement curve. This method was to be carried out in conjunction with any technique which could traverse the limit point using the following equation, [AFf + Alf/AF] = [KT] f[Auf + Afon] (1.3) . . k . where A)»: rs the load parameter vector, AF: rs the vector of unbalanced forces, Au 1. is the residual displacement vector, AF is the reference load vector parallel to the applied load vector, and A12 is the vector of conjugate displacement. The superscript is the load increment and the subscript is the iteration within the load increment. The load parameter, 23 Alf, was chosen such that the equilibrium error was minimized and the load increment was chosen using the constant arc-length method. This procedure has shown to have good convergence and reliability against divergence. A change in equilibrium state was detected by checking, De:(K’;)-Der(l<’;") - Axisymmetric - Linear displacement interpolation - 1 point Gaussian integration - Simpson’s rule integration through the thickness 89 l U, V, - Axisymmetric - Quadratic displacement interpolation - 1 point Gaussian integration - Simpson’s rule integration through the thickness 35 integration points has been reduced to one and hour glass control has been added. (More information on the meaning of these two aspects of element 116 will follow.) The second group of elements included element 1 and element 89. Both elements are also axisymmetric and isoparametric, but, in contrast to the previous group, the kine- matics are based on first order shear deformation shell theory (FSDT). The FSDT assumes that planes originally straight and normal to the midplane remain straight but not necessar- ily normal during deformation. This assumption allowed for inclusion of the shear defor- mations which may arise. The displacement field was then be approximated using three DOF at each node, axial and radial displacements and a right hand rotation in the plane. Again the behavior of the elements is governed by the displacement field interpolation and number of integration points along the length. Element 1 uses a linear shape function with one integration point and element 89 uses a quadratic interpolation and a two point inte- gration rule. For both of these elements the integration through the thickness is carried out using Simpson’s rule based on a user specified number of points or shell layers determined by the type of problem to be analyzed. At this time, the number of layers was set to three. 2.1.2 Evaluation of Element Performance for a Linear Problem Given this set of eligible elements, the validity of each may be verified through the comparison of results obtained for simple problems using these elements to those from closed form solutions to the same problems. Therefore, each element was used to model a test case of an infinitely long cylinder subjected to a circumferentially distributed line load of unit magnitude. The data required for modelling the test case includes Young’s modulus and Poisson’s ratio of the material and the radius and the thickness. Because some finite 36 length must be declared in order to use the finite element method, a finite cylinder was modelled with sufficient length that the displacements near the ends were neglible. Addi- tionally, it was appropriate to apply simply supported boundary conditions to the end points of the finite cylinder. Two sets of geometric data were used to generate the model, one being a moderately thick-walled cylinder: R = 1.0 in. h = 0.1 in. L = 100.0 in. (2.2) where R was the radius, h was the thickness, and L was the length of the cylinder. The second set of data was representative of a thin-walled cylinder with the following geometry: R = 5.0 in. h = 0.05 in. L = 300.0 in. (2.3) The material properties were the same for each of the cases: E = 6.895x1010 Psi v = 0.3 (2.4) where E was Young’s modulus and v is Poisson’s ratio. After obtaining approximate linear solutions from MARC, the results were normalized to the closed form solution developed for a thin-walled cylinder based on FSDT (Reddy, 1984) given by 37 QR B (2.5) u3 (max) = - 25h where u 3(max) was the maximum radial deflection, Q was the magnitude of the circumfer- entially distributed line load, R was the radius of the cylinder, E was Young’s Modulus, h was the thickness, and B was a parameter defined by 2 1/4 (3 = [ill—SJ] . (2.6) R h Using a load of 1.0/21: , the maximum deflection for the thick—walled cylinder was calculated to be -4.8072x10‘11 in., while the thin-walled cylinder had a maximum deflection of -2.9745x10'10 in. The normalized MARC results for the thick cylinder are shown in Table 2.2 and the results for the thin cylinder are in Tables 2.3. In all cases the advantage of symmetry was taken to reduce computational times. 2.1.3 Discussion Results for a Linear Problem A discussion of the results obtained from the MARC analysis of the test problem requires an examination of the assumptions that were involved in the formulation of the elements and the closed form solution. The closed form solution was based on FSDT. Thus any result obtained using Equation 2.5 was approximate. Since the normalizing solution was based on shell theory, one would expect elements 1 and 89 to have better agreement with the closed form solution because of the consistency in formulation. Elements 10, 28, and 116, being based on a continuum theory may not converge exactly to Table 2.2 Predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius = 1.01m, thickness = 0.1 in., and length = 100.0 in. All values are normalized to the closed form solution. Mesh. Size Element (wrth Element 10 Element 28 116 Element 1 Element 89 symmetry) 50 x 1 0.4458 0.6689 0.6555 1.0321 1.0285 50 x 2 0.4462 0.6683 ------------------------------ 100 x 1 0.6020 0.8499 0.8782 1.0321 1.0285 200 x 1 0.7666 0.9810 0.9776 1.0321 1.0285 200 x 2 0.6027 0.9817 -------------------------- 400 x 1 0.8861 1.0063 1.0021 1.0321 1.0285 600 x 1 0.9208 1.0103 1.0009 1.0321 1.0285 600 x 2 0.8559 1.0215 ------------------------------ 750 x 1 0.9312 1.0142 1.0193 1.0321 1.0285 1000 x 1 0.9387 1.0180 1.0301 1.0321 1.0285 the results given by Equation 2.5 for thick shells, but should approach the same results for thin shells. Examination of the results for elements 1 and 89 in Table 2.2 showed good correlation with the closed form solution, with the error for each being approximately 3 percent. In comparison, substantial mesh refinement was required for the continuum elements to converge. Using the 50 x l mesh size as a baseline, elements 28 and 116 required an increase of about four times as many elements, while twenty times the number of elements were necessary for convergence of element 10. Despite the need for an 39 Table 2.3 Predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius = 5.01m, thickness = 0.05 in., and length = 300.0 in. All values are normalized to the closed form solution. Mesh Size Element (wrth Element 10 Element 28 116 Element 1 Element 89 Symmetry) 50 x 1 0.1833 0.2786 0.4101 1.0032 1.0032 50 x 2 0.1833 0.2786 ------------------------------ 100x1 0.2499 0.4365 0.6695 1.0032 1.0032 200 x 1 0.3438 0.6986 0.8760 1.0032 1.0032 400 x 1 0.4762 0.9116 0.9659 1.0032 1.0032 750 x 1 0.6235 0.9839 0.9921 1.0032 1.0032 750 x 2 0.6267 0.9850 ---------- ---------- --------- 1250 x 1 0.7494 0.9981 0.9989 1.0032 1.0032 1500 x 1 0.7896 0.9998 1.0001 1.0032 1.0032 2000 x 1 0.8396 0.9996 1.0012 1.0032 1.0032 2000 x 2 0.8562 ---------- --------- -------- ---------- increased number of elements, 10, 28, and 116 did provide acceptable solutions to the test problem. Examination of the results for the thin-walled case should provide a better understanding of the behavior of the elements for the diameter reduction problem as it too is thin-walled problem. In Table 2.3 the shell elements again quickly converged to the result of Equation 2.5. Also as before, the mesh had to be refined in order for elements 10, 28, and 89 to approach convergence, with elements 28 and 116 continuing to yield results nearly corresponding to the analytical result. On the other hand, element 10 required a far greater 40 number of elements to converge and the final results obtained (200x2 mesh) were still in error by approximately 15 percent. The large error may be a result of the element experiencing locking and/or the size aspect ratio of the element. Locking may be understood by considering the case of a thin beam subject to pure bending. For a situation such as this the transverse shear and membrane energies should be negligible throughout the element. However, the element approximation may not be able to satisfy the kinematic constraints of vanishing shear and membrane strains and also adequately represent the bending strains. As a result, the element may be too stiff and not converge. The most prevalent method for eliminating locking is to use fewer Gauss integration points than the minimum required for exact integration of the energy quantities, thereby relaxing the constraints in the element and allowing the bending energy to control the deformation (Averill and Reddy, 1990). One drawback to this solution technique is that a singular or nearly singular stiffness matrix may result which will cause large deformations to appear without any change in strain energy, a so-called zero energy or hourglass mode. Control of these modes may be attained by adding a small amount of stiffness, akin to adding springs, to their contributions in the coefficient matrix of the model. However, insufficient stiffness will not prevent the spurious energy modes, while too much will prevent viable modes from being found. Element 116 has been derived using both reduced integration and hourglass control, thereby presenting a viable alternative to element 10 which exhibits locking. While element 116 mandated more elements for convergence than the standard mesh, by a factor of 8, the increase in mesh size was manageable and the error for a mesh of this size was less than four percent. Another solution to the problem of element locking is to increase the number of 41 DOF within the model. Two ways of doing this were available: a) increase the elements in the mesh or b) use an element with a higher order interpolation scheme. As can be seen from the results in Tables 2.2 and 2.3, the mesh refinement technique seemed to be working, though a converged solution was not yet attained using element 10 for these two cases. Use of element 28 increased the DOF in the analysis by increasing the interpolation order. Once again the table shows that a greater number of elements were needed to model the cylinder (about fifteen times more) than the benchmark, but the converged result was less than two percent in variance with the normalizing result. The size aspect ratio, the length of the element divided by the height, also contributed to the failure of element 10 to converge. In the thick-walled case the 50 x l mesh resulted in an aspect ratio of 10, meaning the element length was 10 times longer than its height. Thus, to achieve the desired deflection the element was required to bend while experiencing very little shear. This was in contrast to the case of the 100x] mesh, where the minimum aspect ratio was one half and shearing was the dominant mode of deformation. So, the mesh of 1000 x 1 converged to the closed form solution because the shear was not forced to vanish. The maximum and minimum aspect ratios in the thin case were 60.0 and 1.5 respectively. In both cases the bending energy was governing the deformation of the elements and, thus, locking was a consequence. Keeping the aspect ratio close to one or less, the phenomena of locking may be relieved. In order to confirm this hypothesis a third test case was run. In this final case the same material properties as before were used and the geometry was as follows: 42 R = 100.0 in. h = 1.0 in. L = 200.0 in.. (2.7) This geometry allowed greater flexibility in creating meshes with the aspect ratio of one. The analytical solution was again calculated using Equation 2.5 and found to be 1.4835x10'll in. Because elements 1 and 89 did not exhibit any difficulties in converging, only elements 10, 28, and 116 were used in modelling this second thin walled case. Of course both elements 28 and 116 did converge to the normalizing solution, but the hope was that by maintaining the aspect ratio as close to one as possible, these elements would converge faster. The results from the MARC analysis are given in Table 2.4. As expected element 10 did converge to the normalizing solution. Also, as was expected, elements 28 and 116 both converged to the analytical solution at a faster rate. Actually, they converged with the initial mesh. The results obtained indicated that the effect of the aspect ratio was significant and should be kept to a value of 1.0 or less. Table 2.4 Predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius = 100.01n., thickness = 1.0 in., and length = 200.0 in. All values are normalized to the closed form solution. Mesh Element 10 Element 28 E16111“: nt == _ = = 100 x 1 0.8989 1.0041 1.0046 200 x 1 0.9717 1.0049 1.0043 400 x 4 0.9961 1.0050 1.0048 800 x 4 1.0000 1.0050 1.0048 43 2.1.4 Nonlinear Evaluation of Element Performance The ability of element 10 to represent the kinematics of a thin structure was still in question and the decision was made to test this element in a nonlinear situation. So a nonlinear test case was devised using the same geometry as given in Equation 2.7, but with I;100.0 and the following material properties: 1511 = 3.230x107 Psi 1522 = 2.72x107 Psi 1533 = 2.72x1o7 Psi v12 = 0.3 v13 = 0.3 v23 = 0.3 7 7 7 (2.8) G12 = 1.242x10 Psi Gl3 = 1.242x10 Psi 023 = 1.242x10 Psi 0), = 7.785x10“ Psi where Em are the Young’s moduli, Gm are the shear modulus and 0'), is the yield stress. The subscripts given refer to the direction in which the property is valid. The 1 direction was defined to be parallel to the axis of the cylinder, 2 was in the circumferential coordinate, and 3 was parallel to the radial direction. In order to keep the computational times to a minimum plasticity was not included in this test case. A compressive circumferential line load of 4.5x106 pounds was applied to one end of the cylinder in the model. This load was chosen as it would provide for approximately the same amount of diameter reduction as in the swaging process under study. Also, because this problem was to be solved nonlinearly, Equation 2.5 was no longer valid and no other closed form solution was readily available. Hence the converged result of element 28 was chosen to serve as the normalizing solution for this case due to its previously demonstrated ability to converge. Table 2.5 contains the values obtained for this validation problem. Table 2.5 Nonlinear test case results corresponding to the predicted maximum deflection due to a circumferential line load in a simply supported cylinder of radius = 100.01n., thickness = 1.0 in., and length =100.0 in. All values are normalized to the converged solution of element 28. Mesh Element 10 Element 28 100 x 1 0.3078 0.9451 200 x 2 0.6322 0.9811 400 x 4 0.9498 1.0000 800 x 8 0.9537 ----- The data attests to the fact that element 10 will indeed converge in a nonlinear analysis. This, coupled with earlier information from the other test cases, indicates that any of the elements tested would be suitable for use in the forming process, but other information helped to eliminate some elements from further consideration. One of the major features of the MARC finite element code is the ease in which the user can define contact bodies and conditions. Potential contact bodies are identified by the user, and the code internally defines the contact constraints. However, this algorithm does not detect contact at midside nodes, thus eliminating higher order elements from contact analysis. Elements 28 and 89 were, therefore, not eligible to be used in the analysis of the forming problem as the topology of these elements includes midside nodes. This left elements 10, 116, and 1 as the remaining possibilities. Each of these elements had capabilities as well as insufficiencies for use in forming problems. One drawback to element 116 was the reduced integration technique which provides only one Gauss point 45 within the element. Because the strains and stresses (and, hence, the elasto-plastic material properties) will vary from the outer surfaces inward, a single integration point through the thickness would not adequately predict the variation of these values through the thickness of the cylinder. It then becomes necessary to increase the number of elements in the transverse direction to gain more Gauss points. In contrast, Element 1 allows the user to input the number of points to use for Simpson’s integration rule through the thickness, eliminating the need for multiple elements in the radial direction. The final decision of which element to use was left until each could be used in the necking analysis. 2.2 Boundary Conditions and Contact Constraints 2.2.1 Boundary Conditions Some values of the primary degrees of freedom must be known on the boundaries of the cylinder in order to constrain rigid body motions. Therefore, the next step in the modelling process was to establish boundary conditions for the model. For the solution of the diameter reduction problem, it was appropriate to fix the axial displacements on the end opposite the necked region to zero. In doing so the continuity requirements between the modelled and unmodelled portions of the cylinder were satisfied. Since transverse dis- placements are constrained by the axisymmetric response of the cylinder and none of the elements under consideration have the ability to deform in the circumferential direction, this was the only boundary condition needed for the solution of the forming problem. The applied boundary conditions are illustrated in Figure 2.4 by the arrows located on the left hand end of the cylinder. 46 aunt: Figure 2.4 Geometric Model with Applied Boundary Conditions. 47 2.2.2 Initial Stresses Prior to the actual forming process, the cylinder had a pre-existing state of stress due to the rolling of a flat sheet into a cylinder. These effects were also included in the model. It was assumed that the bending was the result of equal yet opposite moments being applied to the edges of the plate resulting in a state of pure bending and, hence, an exact elastic-plastic solution could be obtained. The effects of anticlastic curvature were, however, neglected. For elastic or plastic deformations, the circumferential strain in the cylinder after bending was obtained from geometric considerations as: 599 = g (2.9) where R=1 .294 in. was the mean radius of curvature of the cylinder and 2 was the distance from the midplane of the cylinder. Of course, the maximum strain occurred at 2 = i; and was equal to 0.0027, which exceeded the yield at: O’ .39 = 0.00227 = am (2.10) E00 where (type was the yield stress in the circumferential direction, E99 was the elastic mod- uli in the same direction, and Eype is the yield strain. While some plasticity would be evi- dentin the cylinder, the entire cross-section would not yield due to bending. Rather, there would exist an elastic core near the midplane of the flat sheet and yielding would occur 48 near the edges of the plate, with the plastic zone increasing toward the midplane as curva- ture was increased. In the final configuration, the elastic core extended a distance 2e, away from the midsurface found by: O' YPB . ze, = 3.3-R = 0.00294 in. (2.11) and illustrated in Figure 2.5 in which the cross section of the cylinder is shown with the distribution of stresses resulting from bending the flat sheet. Determination of the limits of the elastic core was important in that MARC does Plastic Region N I Plastic Region Figure 2.5 Cross Section of Cylinder with Stress Resulting from Cylindrical Bending. 49 not allow the initial stresses to exceed the elastic yield limit of the material. Being that the initial stresses must be input at the Gauss points of each element, it was important to ensure that the stress at these points did not exceed the yield stress. Thus, the coordinates of the Gauss points in reference to the inner diameter of the cylinder needed to be deter- mined. Using this information the } coordinate of each was also found with 2 = y(GP) -- ° (2.12) where y(GP) was the coordinate of the Gauss point measured from the inner diameter of the cylinder and h was the cylinder thickness. Initial stresses were determined in the following manner. From the bending strain found in Equation 2.9, the stresses at 2 = i; were found using the one dimensional Hooke’s Law, 91090 = 500300 (2°13) where fine was the calculated yield stress in the circumferential direction. Finally, assuming a linear distribution of stress through the thickness, and using the results from Equations 2.12 and 2.13, the stress at the Gauss points was found using: 6 a (GP) = #292 (2.14) 50 where 0(GP) is the stress at the Gauss point. In the event that the Gauss point fell within the plastic zone, the initial stress value was assumed to take on the value of stress at the elastic/plastic interface. An alternate method of calculating the initial stresses would have been to assume a linear distribution through the elastic region of the cylinder and having the stress in the plastic zone maintain the stress value at the elastic/plastic interface. Employing this scheme would have resulted in a more accurate model of the initial stresses and a 16 per- cent increase in their magnitude. The results to be presented later will show that either approach does not significantly impact the final results of the model. In either case, the input of this data into the model results in circumferential stresses being present at the out- set of analysis. 2.2.3 Contact Conditions Using MARC, the user is able to define a deformable body through the definition of a set of finite elements and a rigid body is defined as a set of geometrical entities, such as lines, circles, splines, or surfaces. In many commercial finite element codes, contact is defined by placing gap- friction elements between the two bodies that may come into contact during the analysis. The difficulty in this method is that the user is required to know a priori which bodies come into contact. While this approach is valid in MARC, another option is also available to automate the contact definition procedure. The basic thrust behind the option is the definition of bodies rather than gap elements. All the required information to enforce non- penetration is contained on the surfaces of these boundaries. In the case of a deformable body, which is defined by a set of finite elements, all the nodes at the boundary become a 51 set of candidate nodes for contact. For rigid bodies only the surfaces of potential contact need to be defined by a set of geometric entities, such as lines. As the analysis is performed, the contact condition is checked by calculating the vector product of the nodal displacements and the normal of the rigid body. If this result is less than or equal to the distance between the bodies contact is detected and constraints are applied to the contact nodes to prevent penetration. Because the distance between the deformable and rigid bodies is a calculated quantity, some round off error will be expected and lead to some penetration. To account for this phenomenon, a contact tolerance is provided to allow nodes to go an incremental distance below a surface and yet still be considered in contact. This may be a user defined quantity or assigned by the algorithm. For a continuum element, the default is 1/20 the smallest element edge and for a shell element it is 1/2 the element thickness. Finally the rigid body motion of the dies is given in the contact option by defining an instantaneous velocity. By explicitly integrating the velocity over time the motion path of the tools is provided (MARC, 1994). For the swaging analysis the elements constituting the cylinder were classified as the deformable body and the forming tools declared as separate rigid bodies. The action of the forming process is such that the tools move toward the cylinder, so each rigid body defined was given a series of velocities to simulate their displacements. At the present time, displacements rather than velocities will be discussed with assumption that the time integration limits have been chosen properly to achieve the correct velocities. The first stage forming process was defined by applying an axial displacement of -0.3729 inches to the body defining the first stage tool. This die was then pulled off the cylinder by using a release option provided in MARC to decouple the nodes of the deformable body from the 52 rigid die in conjunction with a second rigid body translation of 0.4271 inches in the axial direction. Simulation of the swaging process was completed by the motion of the second tool. Initially the position of the tool relative to the cylinder was quite significant and would prevent the analysis from finding a converged solution. Therefore the first movement of the second stage was to bring the tool to a point just before contact would be detected. The required displacement was -0.1060 inches. Forming was done by then moving the die -0.2558 inches axially. This amount of movement was sufficient to bring the ramp of the tool into a position parallel with the previously formed angle in the cylinder. The process was completed by again using the release option with the 0.4271 inch translation of the tool. 2.3 Material Model The response of the cylinder to the action of the forming is also dependent on the mechanical properties of the material used to construct the cylinder. Among these quantities are properties to describe the elastic response of the cylinder, but more important are the characteristics which dictate the behavior after the elastic limit has been exceeded. 2.3.1 Mechanical Properties An elasto-plastic model was developed to describe the behavior of double-reduced tinplate sheet steel conforming to A.S.T.M. A-623 Type L, the material of choice in this analysis. The model began with the specification of the mechanical properties which are given in Table 2.6 (Wood, 1994). The in-plane values for Young’s moduli and yield stress 53 were determined by averaging the results obtained from five samples used in standard tensile testing procedures. Given the thickness of the material, current experimental procedures were inadequate to measure the transverse material properties. Table 2.6 Mechanical Properties of Double-Reduced Tinplate Steel, Type L. Property Axial, x Circumferential, 0 Transverse, z Young’s Modulus. 3.2311107 in. 2.721(107 in. 2.72x107 in. E (psi) Poisson’s ratio, v 0.3 0.3 0.3 Shear Modulus. 1.242x107 in. 1.24211107 in. 1.24211107 in. G(psi) Yield Stress. 7.785x104 in" 6.544x104 in. 6.544x104 in. oyp(psr) Therefore, the shear moduli were calculated using, E with E being given by the axial value of Young’s modulus. This resulted in isotropic values for the shear moduli, in addition to the given isotropic values of Poisson’s ratio. This assumption of isotropy in the shearing values will not lead to significant error. In fact, transverse shear effects are usually neglected in a structure with a radius to thickness ratio of greater than 20 (Dym, 1974). In this case the radius to thickness ratio is in excess of 180. 54 2.3.2 Plasticity Model The definition of the plasticity model requires the yield stresses, a method for determining the onset of yielding (a yield surface), a definition of the hardening curve, and finally the effects of hardening on the yield criteria. The yield stresses are given in Table 2.6 and were found using the same tensile tests as used in finding the elastic properties. Since the cylinder was in a state of multi-axial loading during the simulation, a multi-axial yield criterion was required. Several models exist for this purpose, most notably among them is the von Mises yield criterion wherein the material is assumed to have yielded when the distortion energy is equal to the measured uniaxial yield stress (Mendelson, 1968). For an anisotropic material, the relationship of distortion energy to yield stress is given as 2 2 2 2 2 2 2 + 304123 + 3a5‘t31 + 3a6t12 where am corresponds to the axial yield stress, oij are the calculated stresses referred to the coordinate axes of the cylinder (11 - axial, 22 - circumferential, and 33 - transverse), ti]- are the calculated shear stresses in the cylinder during analysis. The coefficients aa, 01:1 to 6, account for directional variations in the yield stress from the axial and may be found by defining the Hill’s yield ratios YRDIRl, YRDIR2, and YRDIR3 in conjunction with the following 55 a1:———l———+___1.__ l YRDIR22 YRDIR32 1011911912 ,2=___1_,__1_ 1 YRDIR32 YRDIRlz YRDIR22 ,3._._1_.,__1__ 1 (2.17) YRDIR12 YRDIR22 YRDIR32 _ 2 _ 2 _ 2 a4 - ———2 a5 — ——-5 a — ———2 YRDIR3 YRDIR2 YRDIRl where O' 0' 1 0' YRDIRl = Y”: YRDIR2 = "’R YRDIR3 = 332 (2.18) OYP GYP OYP I I X For double-reduced steel, YRDIR1=1.0, YRDIR2=0.8405, and YRDIR3=0.8405 were used. After the material points in the cylinder reach the plastic region of the stress-strain curve, the material will begin to harden. The amount of hardening the material experiences must also be described through a mathematical model. Isotropic hardening, used in this model, allows the yield locus to retain its center point throughout the analysis, but the size of the locus will grow with increasing strain hardening. Inputting a table of plastic strain versus stress into the MARC data deck allowed the program to internally account for these effects. Table 2.7 contains the measured values for the double reduced tinplate sheet steel. When this data was plotted in a stress versus plastic strain curve, as in Figure 2.6, it was observed that the curve was nearly flat. This would indicate that the strain hardening will not have a significant impact on the results obtained for the model. Proof of this hypothesis will be left until the model has been completed and the analyses run. 56 Table 2.7 Stress versus Plastic Strain for Double-Reduced Tinplate Sheet Steel, Type L. Stress Plastic Strain i — 7.7854x104 Psi 0-0 7.902411104 Psi 6.4000x10'4 7.9727x104 Psi 1.3700x10'3 8.0079x104 Psi 2. 190011103 8.0109x104 Psi 3.10001110-3 Stress (Psi) (a) # UI to 1 0 0.5 1 1.5 2 2.5 3 Plastic Strain x 104 Figure 2.6 Stress versus Plastic Strain for Double Reduced Tinplate Sheet Steel. 57 2.4 Solution Parameters and Analysis Options 2.4.1 Solution Method As previously mentioned, because the problem under consideration is nonlinear, the Solution path is also nonlinear. In this case a method of tracing this curve was required in order to obtain solutions. Therefore, the Newton-Raphson method, which is one of several techniques, was employed to approximate a solution to the resulting nonlinear algebraic equations. The advantage of the Newton-Raphson method is that it provides a quadratic rate of convergence to the solution (Zienkiewicz, 1989). In order to understand the Newton-Raphson method, consider a fictitious load-deflection curve as shown in Figure 2.7. The desired load is given as PC. In cases where the entire path must be known, such as in plasticity, the curve may be approximated by loading the structure incrementally. Within each load step the algorithm ensures the equilibrium by reaching the solution point in an iterative sequence which is terminated when some convergence criterion is satisfied. Accordingly the motion of the forming tools was applied incrementally. Recalling the definitions of the tool motions in Section 2.2.3, the first stage tool has two separate translations, one onto the tool and the second a releasing action. In the Newton-Raphson solution the initial tool movement was divided into 250 equal load steps and the second was done in 2 unequal steps. The first increment of the releasing action was very small in order to ensure that the motion started in the correct direction and the second increment completed the releasing motion. For the second tool, three movements were given: bringing the tool near the cylinder, the forming operation, and the releasing action. The initial movement of the second stage tool was performed in one load step, the forming was 58 Load Deflection UA UB UC U1 U2 U1 U2 Figure 2.7 Newton-Raphson solution method. again divided into 250 equal increments, and the releasing action also performed in two increments, one small and the other large enough to complete the requirements. Within all of the load increments a maximum of 20 iterations were allowed, but this was set as upper bound. After each iteration, the maximum displacement increment was divided by the maximum displacement change (for the load increment) within the model to calculate a convergence tolerance. When this tolerance was found to be less than or equal to 0.10 the increment was said to have satisfied the convergence criterion and the next load increment was begun. In the event that convergence was not achieved, upon reaching the end of the 20th iteration the program terminated giving a message that the convergence criterion was 59 not met. This leads to some comments about the number of load increments used. 250 increments was found to be a lower limit for convergence reasons. When a smaller number of increments was used, the load steps were too large for the algorithm to attain a converged result within the specified number of iterations. Using a greater number of increments could result in a more accurate solution, with the trade-off being increased solution time. Therefore, the initial simulations were done with 250 increments, but the effects of increased load steps on the solution were also investigated. 2.4.2 Analysis Options The large deformations that are inherent in the process of swaging the cylinder required a reference frame from which the displacements may be measured. In addition, the particular measures of strain and stress that are to be used must also allow for the large deformation effects. MARC offers a set of options to account for this phenomenon, of which three were used in combination in the simulation model. These were the large displacement, finite strain plasticity, and updated Lagrange options. The first two cards account for the large deformations and strains as their names imply. The last option, updated Lagrange, gives a reference frame for the measurement of deformations and calculation of the stiffness matrix. Here the mesh is connected to the material throughout the analysis and is updated at the end of each load increment. By using these three options in combination the entire spectrum of large deformation were taken into account in the model. And with that the definition of the model for simulating the swaging process was completed. The forming process may now be analyzed and an understanding of the deformations and stress state of the cylinder during forming can be sought. Chapter 3 Numerical Simulation of Swaging 3.0 Introduction The model for the simulation of the swaging process has been described. Using this model preliminary deformations of the cylinder as it progresses through the process may be observed. The deformations are preliminary in that up to this point two rather important details have not yet been determined. The element that is to be used is yet to be finalized, with the remaining choices being elements 10, 116, and 1. In addition to the element decision, the size of the mesh is also a lingering question to this point. Upon making a decision about these components, the final deformations and stress state of the cylinder during and after the necking process may be analyzed. The information revealed may then be used to ascertain possible causes for the formation of wrinkles in the cylinder 60 61 and ultimately may lead to candidate solutions. 3.1 Simulation of the Swaging Process Now that a model has been defined, the swaging process can be simulated. Although, it should be noted that the results shown are preliminary until final selection of an element and mesh size. Thus, a baseline mesh of 150 elements along the length and 2 elements through the thickness, or a 150 x 2 mesh, was used in concert with element 10 to produce the initial simulation shown in Figures 3.1 (a) through (i). In Figure 3.1(a) the th increment of the simulation, before the forming cylinder and tools are shown at the zero operation had begun. As the first stage tool progressed forward, the leading edge of the cylinder was bent down and traveled along the tool ramp toward the “alley” as shown in Figure 3.1(b), the 50th increment of the simulation. Figure 3.1(c) depicts the leading edge of the cylinder as it was just being bent back into the “alley”, increment 100, so that the diameter reduction of the first stage could be achieved. By the 250th increment, the leading edge had slid down the “alley” until the end of the tool was reached, as illustrated in Fig- ure 3.1(d). Finally the first stage was completed as the tool was removed in the 252“d increment of analysis, as seen in Figure 3.1(e). In Figure 3.1(i) the second stage tool has moved to a position near the cylinder so as to prevent numerical difficulties in attaining convergence, increment 253. The tool then pushed the cylinder’s leading edge down until the “alley” was again reached at the 352“d increment of analysis, as shown in Figure 3.1(g). The advancement of the second tool was discontinued when the angled portion of the cylinder was parallel to the ramp of the tool 62 Figure 3.1(a) First Stage of Swaging at the Zeroth Increment. M Figure 3.1(b) First Stage of Swaging at the 50th Increment. 63 .0 ‘ i 08 Figure 3.1(c) First Stage of Swaging at the 100th Increment. ii '. ' ‘ [to Figure 3.1(d) First Stage of Swaging at the 250th Increment. “. ital Figure 3.1(e) Release of the First Stage Tool. e e u a a 0 2.1“ :w 1m Figure 3.1(f) Tool Moved into Position in First Increment of Second Stage. 65 £133 igir 8.24 L ”I Figure 3.1(g) Second Stage of Swaging at the 352nd Increment. : ~ 0 :SJM m #4 L. ”1 Figure 3.1(h) Final Increment of the Second Stage. 66 iii?" l1” Figure 3.1(i) Release of Second Stage Tool. which occurred at the 503rd increment and is shown in Figure 3.1(h). The forming process was completed in the 505th increment when the second stage tool was removed from the cylinder, as illustrated in Figure 3.1(i). 3.2 Final Element Selection Three elements were still available for use in the modelling of the swaging process, elements 10,116, and 1. Each of these was used the model previously'discussed. The results from each of these analyses were then evaluated to determine which provided the most accurate solution at the best computational cost. 67 3.2.1 Forming Analysis Results Using Element 1 After establishing the baseline with element 10, element 1 was used in the simula- tion of the forming process to determine if the efficiency which was exhibited in the test cases still held true. In review, Table 2.1 indicates that element 1 is a two-noded, isopara- metric, and axisymmetric element which employs Simpson’s rule integration through the thickness. As such, when substituting element 1 into the model, the number of layers within the element needed to be defined. For the metal forming analysis to be conducted 11 layers were specified. This was equivalent to using 11 Gauss points through the thick- ness which was attractive in that an accurate representation of the material effects should be the result. However, when the analysis was run, the model consistently failed to pro- vide a solution. Either the solution failed to converge or the stiffness matrix of the model became singular (or nearly singular). Several attempts were made to overcome these diffi- culties, but no resolution was found. Therefore the decision was made to abandon further use of element 1. 3.2.2 Forming Analysis Results Using Element 116 The second element that was used in the model was element 116. Recalling Table 2.1, element 116 is a four-noded, isoparametric, and axisymmetric element employing reduced integration and hourglass control. In the previous studies of Section 2.1.1, this element also provided good convergence characteristics within a reasonable amount of computational time, and the same was hoped of the swaging simulation. Once again, however, the analysis deteriorated in all cases with element 116 providing the kinematical model for the cylinder. Inspection of Figure 3.2 reveals the presence of hourglass modes 68 INC: I! 808: 0 TIIE:3.9200-01 FREQ:0.0000-o-00 am 10M Figure 3.2 Swaging Analysis Using Element 116 with a 150x2 Mesh. 69 within the model at the 96th increment of the analysis. Thus it may be concluded that the stiffness that was added to the element in order to prevent zero-energy modes from occurring was not sufficient for this particular model. Thus, alternative approaches of prevention were tested. The first run using element 116 only had two elements through the thickness. It was believed that this may not provide enough stiffness in the presence of the plasticity model, nonetheless, increasing the number of through the thickness elements did not eliminate the hour glass modes. Therefore, a‘second attempt was made using a mixed model with two elements of type 10 at the leading edge of the cylinder and the remaining elements being type 116. This again proved to be a fruitless effort. As a result, element 116 was also no longer considered a viable element for the analysis at hand. 3.3 Convergence of Forming Problem Element 10 was left as the only available element to discretize the model of the cylinder. \Vrth an element now selected, the mesh size needed to be finalized by conducting another convergence study similar to that of Section 2.1.1. Again, for the case of the forming process no tractable analytical solution is available, and experimental studies of the problem do not provide an adequate description of the internal stress state of the cylinder. Therefore, to gain confidence in the finite element solution, the model must be evaluated until a converged solution is found. For the swaging simulation, convergence must be achieved in two manners. The mesh used to model the geometry should be refined in both the axial and radial directions to gain monotonically convergent results. Several reasons exist for the need to smooth the model in both coordinates. The cylinder initially has regions of plastic stress 70 due to the rolling of the flat sheet into a cylinder at the inner and outer surfaces as discussed in section 2.2.1. These plastic regions will grow radially inward as the cylinder is subjected to the forces of the forming dies. In order to capture the material effects of the growing plastic region, the model should have a large number of Guass points through the thickness of the cylinder cross section. The convergence of the calculated values in the radial coordinate would indicate the accuracy of the material model. In addition, the number of elements in the axial direction were increased to capture the path dependency involved. This was done in conjunction with a study on the number of load steps that were necessary for reliable results. The number of load used to model the forming process will also contribute to the path dependency of the results obtained from the model. This is because the physical process of swaging the cylinder is carried out in two smooth, continuous motions. However, in the finite element model of the process, the tools were given a velocity which was divided into a number of equal load steps. By this method, the displacement of the rigid dies is not a continuous function of time, but rather a piecewise linear approximation. Of course a greater number of time steps taken will increase the smoothness of the tool displacement which affects the path dependency of the cylinder. Larger time steps will move the tool a greater distance along the axis of the cylinder and will result in greater incremental displacements. Therefore it is recommended that a convergence study on the number of load steps also be conducted. 71 3.3.1 Mesh Refinement Study A study of the elemental convergence of the first stage was conducted to determine which mesh would provide for the most accurate results. For this convergence study, the baseline mesh remained 150 x 2. The elements were then increased through the thickness of the cylinder up to eight. The number of axial elements were then expanded to 300 and finally to 400. In the case of 300 elements the through the thickness elements were the same as for the 150 meshes. For the meshes containing 400 elements, analyses with 2 and 3 elements through thickness were not run because it was already determined that more elements through the thickness were needed to account for the material variations in that direction. Upon completion of each analysis, the values of axial and circumferential strain and stress were collected on the outer surface, mid-plane, and inner surface of the cylinder at two points within the mesh. The first of these points, labeled ‘Point A’ in Figure 3.3, was within the first bending region and the second was in the critical region, ‘Point B’ in Figure 3.3. Also it should be noted that the data was gathered at the 240th increment of the analysis and not the final 250th increment as specified in the model. The reason for this was that some of the models exhibited slight penetration through the end of the tool, which resulted in extraneous axial stresses in the cylinder. Therefore, to make a fair comparison between the models, the 240th increment was used to avoid the penetration problem. The strain and stress data resulting from the convergence study can be found in the Appendix Tables A.1 through A.24. For compactness, the discussion that follows will refer to that data, but specific data values are not mentioned. 72 THE : LN FREQ : 0.000900 Figure 3.3 Sampling Locations for Convergence Study of Forming Analysis. 73 3.3.2 Discussion of Mesh Convergence Study Data The analysis of the gathered data began with an evaluation of the strain values, where very little variation resulted from increasing the number of through-the-thickness elements. In Figures 3.4 through 3.7, the strain values versus mesh size were plotted for the critical region and appear to contradict this assessment. While the data for those meshes with 150 elements and 400 elements along the length were in good agreement with one another, meshes with 300 elements along the length exhibited large variations in the strain values with increasing mesh refinement through-the-thickness. This is, however, misleading due to the scaling used for the generation of these plots. In fact, an examination of the percentage change in the strain resulting from refining the mesh size would be more revealing as has been done in Tables 3.1 and 3.2. The observation could then be made that the changes in strain magnitude were actually less than four percent for each component on both of the cylinder surfaces. The exception was the axial strain on the outer surface of the cylinder. Here the values varied by up to 350 percent, however, the magnitude of strain in this region was an order of magnitude smaller than the other strain measures. Thus, this strain component was not dominant in the swaging simulation and therefore did not alter the earlier conclusion. Similar conclusions may be drawn for the strains in the first bending region. In any case, these small changes in strain were not surprising because the displacements in the first stage of the forming process were controlled by the geometry of the tool. The route followed by the cylinder would not be significantly altered by an increase in the number of elements through the thickness. 74 3 0 a t 3 w 3 -1 '5 O l E -2 P — 150 Elements I; - - 300 Elements '2 3 ~- ' - 400 Elements 4 l- _5 A 1 l l l l 2 3 4 5 6 8 Elements Through the Thickness Figure 3.4 Axial Strain on the Outer Surface in the Critical Region vs. Mesh Size. -0.0615P a. - - - - ‘9 I \ I 1 -0052 X ‘ 8 ' I ‘, g ’I g- - .. ‘ w A l 4% -1‘\~ k A A 3 m v 1 ‘ If; v 5 ~4 1 7 04.0525- . I 1 ‘ / 7.! X 7’ ‘E ‘ I D g -o.osa- ‘. ,’ ”S —150Elements “ , ,s_. - - aooElements 1 ,’ Q-0.0535» r-~-400Elements ‘, ,’ \ t” 4 _o.m‘ l J 1 J 4 1 2 3 4 5 6 B EIemntsThroummeThldmess Figure 3.5 Circumferential Stress on the Outer Surface in the Critical Region vs. Mesh Size. 0.044 - .09 5% .o 9 d I Axial Strain - Inner Surface 0 . 0 § 2 .0 r 0.037 - Figure 3.6 Axial Strain on the Inner Surface in the Critical Region vs. Mesh Size. 75 —-— 150 Elements - - 300 Elements - - ~ - 400 Elements vol 4 5 6 8 Elements Through the Thickness -0.0517 - - r ‘ ‘1 4.1518 " IT— 1 \ I \ 8 4.051% ,' 1 —150Elements E [I X - - 300 Elements ‘3 .0052- , “ ----4OOE|ernents g 1 ' 4.0521 - ‘1 g 1 a \ 3 4.0522) 1‘ S 1 3.‘ -o.0523 - 1 E l E H- - — -‘ l \ o -O.m24 " ‘3‘ \ v \ 3 fi- ‘k '\ \ .. — - -+ \ I ' ‘1'- ----- -o.0525L \ ‘ , x' ‘4 -O.m26 ' ' ‘ 2 3 4 5 6 8 Elements Throum the “midmess Figure 3.7 Circumferential Strain on the Inner Surface in the Critical Region vs. Mesh Size. Table 3.1 Percentage Change in Axial Strain Values in the Critical Region. (%) Elements Through the Thickness Cylinder Axial Surface Increase 2 3 4 5 6 8 150 to 91.46 267.71 286.70 81.60 34.98 30.70 3 300 5 30010 444.70 -348.90 -21.83 400 1: afi- 150 to 0.70 14.05 -0.95 -140 1.53 0.92 S 300 t: 5 300 to 2.90 3.34 0.60 400 Table 3.2 Percentage Change in Circumferential Strain Values in the Critical Region. Elements Through the Thickness Cylinder Axial Surface Increase 2 3 4 5 6 8 150 to 0.06 -0.26 1.16 1.22 -2.69 -0.06 3 300 5 300 to -1.03 -1.07 2.59 400 150 to 0.06 -0.23 1.24 1.30 -0.09 -0.06 s 300 t: E 300 to -1.09 -1.13 -0.06 400 77 The formation of wrinkles in the cylinder is directly attributable to the build up of stresses and strains in the critical region. Therefore attention was also focused on the calculated stress values obtained from the convergence study where it was found that the trends contrasted those of strains. The change in stress states due to an increase in elements along the length was found to be minimal, but a change in the number of elements in the radial coordinate had a distinct impact on the stress values as shown in Figure 3.8 through 3.11. For the cases of 150 and 300 elements along the length, these plots show the stresses found on the outer and inner surfaces in the critical region versus the number of elements through the thickness. The effect of increasing the axial elements had very little consequence on the stress values obtained from the model as evidenced by the close agreement between the two curves in each of the plots. In contrast the multiplication of the number of elements through the thickness seemed to have a dramatic effect. As discussed in Section 2.2.1, the cylinder enters the forming process with initial stresses which induce zones of plasticity near the outer and inner surfaces of the cross- section. During the forming of the cylinder these plastic regions grow radially inward and the resulting stress gradients were adequately captured only by increasing the number of elements (thereby increasing the Gauss points) through the thickness. Accordingly, as the mesh was refined through the thickness, the values of stress began to converge. However, as can been seen in Figures 3.8 through 3.11, a converged solution was not attained with the meshes used. 78 I 01 O 1 3 i Axial Stress (szi) - Outer Surface 1 -250 .— / — 150 Elements “300 ” - - 300 Elements -350 . 4m 1 1 l 1 J I 2 3 4 5 6 8 Elements Through the Thickness Figure 3.8 Axial Stress on the Outer Surface in the Critical Region vs. Mesh Size. —150Elements -- 300Elements 2'? l. 1. s 1.3 erential Stress (szi) - Outer Surface 1 -250. E 300. § 6 _350 _ -400- 2 3 4 5 6 8 Elements Through the Thickness Figure 3.9 Circumferential Stress on the Outer Surface in the Critical Region vs. Mesh Size. 79 350~ 300~ O 0250* E 3 a) 2 .E 200' 1 6 a 5150- 8 2 a 2100- 50- — 150 Elements - - 300 Elements 0 1 l I l 1 1 2 3 4 5 6 8 Elements Through the Thickness Figure 3.10 Axial Stress on the Inner Surface in the Critical Region vs. Mesh Size. i i ii 3 Circumferential Stress (szi) - Inner Surface 1? o 1 l l l l 2 3 4 5 6 Elements Through the Thickness @1- Figure 3.11 Circumferential Stress on Inner Surface in the Critical Region vs. Mesh Size. 3.5 1" to 01 CPU Time (seconds) in 0.5 80 x 10 _ , ,+ ,+’ ’ ’ / .. / I r: , _ , — 150 Elements / - - 300 Elements ’ 4t 2 3 4 5 6 8 Elements Through the Thickness Figure 3.12 Computational Time vs. Mesh Size. 81 The question still remains as to which mesh should be used in the simulation of the forming simulation. Two final analyses of the convergence data provided the answer. First of all, was the rather minimal increase in strain and stress values obtained from increasing the number of elements along the length computationally efficient? This query was answered by plotting the CPU time (the amount of time the computer took to solve the finite element equations) versus the mesh size, as was done in Figure 3.12. The CPU times were obtained by performing each analysis on a Hewlett-Packard Series 9000 model 715 Unix workstation running at 75 MHz with 32 megabytes of RAM and 1 Gigabyte of disk swap space. In Figure 3.12 the meshes containing 300 elements were observed to require an average of 1.57x104 CPU seconds (4.4 CPU hours) more CPU time to complete than those with 150 elements. Therefore, the 150 element meshes appeared to be more attractive, but final judgement was reserved until a comparison of the results was made. 3.3.3 Comparison of Stress Trends Resulting from Mesh Refinement Finally, the axial and circumferential stresses in the critical region for three meshes; 150x8, 300x8, and 150x2, were plotted to see the variations present. By zooming in on the critical region, the stress gradients were magnified to produce Figures 3.13 through 3.18. In each of the figures, the color bar indicates that compressive stresses were present on the outer surface, while the inner surface was subjected to tensile stresses. Comparing the stresses for the 150x8 mesh in Figure 3.13 and 3.14 to those of the 300x8 mesh in Figure 3.15 and 3.16 the only evident improvement was in the lengthwise gradients, but the gains were minimal. This information coupled with the run time data lead to the conclusion that 150 element meshes should be used for any subsequent analysis 82 of the forming process. In looking at the results from the 150x2 mesh in Figure 3.17 and 3.18 the trends remained the same, but the contours were more banded than in the previous plots. This was a result of only two elements through the thickness being used. As such, the stresses can only be linearly interpolated between four points rather than sixteen, thus yielding poorer gradations of stress. But the trends still show compressive stresses on the outer surface and tensile stress on the inner. This appeared to indicate that the 150x2 mesh would be appropriate for the simulation, provided the conclusions are to be based on the stress trends and not the absolute values. Earlier, concern was expressed about the presence of locking in meshes where the element aspect ratio was greater than 1.0. The 150 x 2 mesh has an aspect ratio of 1.33 and as a consequence locking should be a concern in the second stage of the forming. In order to judge the effects of locking an 800 x 2 mesh, aspect ratio = 0.25, was run. The displacements from that analysis were plotted in Figure 3.19 for increment 303 and in Figure 3.21 for increment 373. Analogous plots were constructed for the 150x2 mesh and are shown in Figures 3.20 and 3.22. The results were very nearly the same in both meshes, thereby indicating that the 150 x 2 mesh was not overly stiff. A possible explanation for the non-existence of locking could be the complicated state of plasticity in the model, but an in depth evaluation for this phenomenon was not performed. But, the absence of locking certainly solidified the argument to use the 150 x 2 mesh in the swaging simulation. 83 ...... L uterine!“ I Figure 3.13 Axial Stress in the Critical Region for the Second Stage of the 150x8 Mesh. 84 4M 4m "M L “We!“ 1 Figure 3.14 Circumferential Stress in the Critical Region for the Second Stage of the 150x8 Mesh. 85 per I‘M“... 1 Figure 3.15 Axial Stress in the Critical Region for the Second Stage of the 300x8 Mesh. 86 " L ”Genet..- 1 Figure 3.16 Circumferential Stress in the Critical Region for the Second Stage of the 300x8 Mesh. 87 or“ ,. M V 4.01m“ Id! tdcerlpd ”as 1 Figure 3.17 Axial Stress in the Critical Region for the Second Stage of the 150x2 Mesh. 88 1m “Met...- 1 Figure 3.18 Circumferential Stress in the Critical Region for the Second Stage of the 150x2 Mesh. 89 a: .. .... ee ii.‘ Figure 3.19 Deformations of the 800x2 Mesh at Increment 303. i .. ee Figure 3.20 Deformations of the 150x2 Mesh at Increment 303. 90 11:: Figure 3.21 Deformations for the 800x2 Mesh at Increment 373. -sn - m ' O . 1'7... . 'm . y: L psi Figure 3.22 Deformations for the 150x2 Mesh at Increment 373. 91 3.3.4 Convergence Study on Number of Load Increments The last component of the mesh that needed to be finalized was the number of increments required. Some indications of path independence were expressed above. To prove this was the case, both stages of the 150x2 mesh were used in an analysis with 500 load steps for each stage of forming instead of the usual 250. Once again, very little variation in the strains was evident. The stress data in the critical region for the first stage was collected in Table 3.3 and the second stage in Table 3.4. The data indicated that the largest change for any of the values was less than 3.0 percent, and therefore it could be said that the increase in the number of increments used to perform the analysis made no change in the results. As such, the simulation of the swaging process could confidently be carried out using the 150x2 mesh and 250 increments. 3.4 Results from Forming Simulation 3.4.1 Strain and Stress States in the Critical Region Once a mesh was settled upon, results from the simulation were analyzed to determine possible mechanisms for wrinkling.The results from the final analysis are shown in Figures 3.23 through 3.26. In each of the plots the cylinder was in the final increment of the forming process (increment 503) and the data was collected in the critical region of the cylinder. The first contour plot, Figure 3.23, consists of the axial strains. As before, the color bar on the left hand side designates the dark red (almost a purple) region on the outer surface of the cylinder to be in a state of tension which increases in magnitude to the inner surface that is yellow. Intuitively, one would think the outer surface should be in compression as a thin-walled structure in bending would exhibit. However, the severe 92 Table 3.3 Stress Values for 250 vs. 500 incs for the First Stage in the Critical Region. Number Oxx 03“ Cu 0’06 0'90 969 of Ines (outer) (mid) (inner) (outer) (mid) (inner) 250 -288.52 8.38 272.58 -366.38 -64.87 208.42 500 -283.95 8.14 267.11 -361.82 -65.36 204.21 % change 1.58 -2.86 -2.01 1.24 -0.75 -2.02 Table 3.4 Stress Values for 250 vs. 500 incs for the Second Stage in the Critical Region. Number (rxx ox, (1,,x 699 699 099 of Incs (outer) (mid) (inner) (outer) (mid) (inner) 250 #26348 :-4.12 240.95 477317 =--—--¥-63. l7IE 175.29 500 -263.56 -4.23 240.22 -276.62 -63.04 174.92 % change -0.03 -2.67 -0.30 0.32 0.21 0.21 state of bending to which the cylinder was being subjected has created membrane forces which overwhelm the bending forces resulting in tensile axial strains. Figure 3.24 shows the circumferential strains as increasing in magnitude moving left to right, corresponding to increased necking. Each value of strain was approximately constant through the thickness, analogous to a cylinder undergoing a uniform diameter reduction. The axial stress distribution displayed in Figure 3.25 indicates that the outer surface of the cylinder is in compression, as shown by the blue region, and the inner surface is in tension, as 93 shown by the red zone. In this case the membrane forces were as dominant as in the case of strains and so the bending forces were significant enough to create the compressive stresses on the outer surface. Finally, the stress gradients in the circumferential direction are given in Figure 3.26 and once again the outer surface was blue, or compressive, and the inner surface was red indicating tension. Recalling the circumferential strains were found to be uniformly compressive through the thickness, the presence of tensile circumferential stresses was initially surprising. These will be explained in the sequel 3.4.2 Assessment of Strain Hardening and Initial Stress Effects on Strain and Stress States The effects of strain hardening and the initial stresses were examined by carrying out additional analyses, once without the work hardening table and the other without the initial stresses. In Table 2.5 the global maximum and minimum values of strain and stress are shown for the strain hardening models. From the data tabulated it can be seen that the hypothesis of Section 2.3.2 was indeed correct, the strain hardening had very little effect on the outcome of the analysis. Further comparison can be made by contrasting the contour plots of the model with strain hardening, Figures 3.23 - 3.26, to those without strain hardening which may be found in the Appendix, Figures A] through A4. The effects of the initial stresses were also evaluated. In Section 2.2.2 two methods of calculating the initial stresses were given. At that time it was stated that either method would not significantly affect the final results obtained from the model. Table 3.6 contains the global minimum and maximum strain and stress values from two models, the first being the model run with the initial stresses calculated by the first method (linear distribution through the thickness) and the second a model analyzed with no initial 94 model are not dependent on either method. and Stress Values stresses. From the data it may be concluded that the impact of initial stresses on the strain and stress values was minimal. More importantly the trends of stress distribution did not change, which can be seen by comparing the plots in Figures 3.25 and 3.26 to those of Figures 3.27 and 3.28. As previously stated, the model used does not yield a converged solution, and any explanations offered for the formation of wrinkles in the cylinder must be based on the trends of stress. Therefore, despite the fact that an alternate method of calculating initial stresses was available, the conclusions drawn from the results of the Table 3.5 Effects of Strain Hardening on the Global Minimum and Maximum Strain 5xx(m3X/ 899(max/ oxx(rnax/ 099(rnax/ Model . rmn) rmn) mm) mm) With Strain 4.686x10‘2 1.210x10'3 3.357x105 2.768x105 Hardening -1.568x10'2 -8.7O7x10'2 -3.811x105 -3.481x105 Without 4.685x10'2 1.207x10'3 3.39211105 2.862x105 Stain Hard- ening 1574;110-2 -8.703x10'2 -3.863x105 -3.508x105 0.0 -0.2 1.0 3.3 % Change 0.4 0.5 -1.4 0.8 95 and Stress Values Table 3.6 Effects of Initial Stresses on the Global Minimum and Maximum Strain 2.2mm 86991:“ “MW stew Model . mm) mm) mm) mm) 4.686x10'2 1.210x10'3 ' 3.357x105 2.768x105 With Initial Stresses -1.568x10’2 -8.707x10’2 -3.811x105 -3.48lx105 4.471x10'2 1.089x10'3 3.463x105 2.867x105 Without Ini- “al Stresses -1.563x10'2 -8.705x10'2 3923:1105 -3.597x105 4.8 -11.1 3.1 3.5 % change 3.2 0.0 -2.9 -32 96 £171.43 4M V I“ I post id Gen. ofTeU w 1 Figure 3.23 Axial Strains in the Critical Region at the 503rd Increment. 97 4041.02 of” Y arch-ea l per 31 On. of Ted huh 1 Figure 3.24 Circumferential Strains in the Critical Region at the 503rd Increment. 98 “To.“ ‘ Idcondfinee 1 Figure 3.25 Axial Stresses in the Critical Region at the 503rd Increment. 99 Figure 3.26 Circumferential Stresses in the Critical Region at the 503rd Increment. income!“ 1 Figure 3.27 Axial Stresses in Critical Region at 503rd Increment, Calculated without Initial Stresses. 101 V Mstfl'ho-fl I ”We!“ 1 Figure 3.28 Circumferential Stresses in the Critical Region at the 503rd Increment, Calculated without Initial Stresses. Chapter 4 Discussion of Wrinkling 4.0 Introduction An assessment of the forming operation reveals that two basic modes of deformation are responsible for the diameter reduction of the cylinder. The first is a uniform diameter reduction and the second is the bending needed. to get the leading edge of the cylinder into the “alley”. While these two modes are, in fact, coupled due to the nonlinear nature of the problem, to understand their contribution to the formation of wrinkles, two arguments will be made. The first is based on the existence of some critical circumferential strain in the cylinder. The second will reveal the contribution of the bending energy to the formation of wrinkles. 102 103 4.1 Existence of Critical Strain For illustration purposes, consider a straight column under an axial load, as shown in Figure 4.1. From strength of materials it is known that this column has a critical load, PCR, which, when exceeded, will cause the column to snap into a new configuration upon the application of some small lateral load. This critical load may be found using Euler’s equation, - —, (4.1) where E is Young’s modulus of the material, I is the moment of inertia of the column, and L is the length of the column. The moment of inertia for a rectangular column is given as, I = 1—2bh , (4.2) where b is the width of the column and h is the thickness. Substituting Equation 4.2 into 4.1 and knowing that the area of the column is given as: A = bh , (4.3) Equation 4.1 may be written as: 104 1. Figure 4.1 Euler Buckling Column. Mxx '80 Figure 4.2 Shell Element Subject to Bending and Resultant Moments. 105 it h A PCR = (4.4) 12L2 ' Finally the critical load may be written in terms of stress, which in turn may be related to a strain by the following equations, PCR = oCRA em = £80,. (4.5) Thus, the critical strain relationship is found by, em = 12(2):. (3.6) From Equation 4.6 it may be seen that the critical strain of the column with a constant length, L, is a function of the thickness, h, squared. So as the thickness of the column is increased the critical load of the column grows in a quadratic manner. The issue of wrinkling occurs in a cylinder under external pressure. For this geometry, the critical strain is proportional to the square of the thickness, h, over the radius, R, of the cylinder. Thus, if the induced strain is equivalent in the two cylinders, the thicker cylinder may not exceed its buckling strain, whereas the thin-walled cylinder under consideration could. From the viewpoint of a cylinder subject to a uniform diameter reduction the importance of the wall thickness is realized. However, the cylinder was also 106 loaded in bending, which gives rise to a more complicated state. 4.2 Bending Effects Isolating the applied bending loads, a second approach is proposed to explain the results obtained for the variation of circumferential stresses as shown in Figure 3.26. The geometry to consider for this explanation is that of a cylindrical shell element depicted in Figure 4.2. Imagining that this shell was participating in the forming process, then as the leading edge of the shell was entering the “alley” it could be considered to be subjected to a bending moment M H = M , shown in blue. The moment-curvature relations are given 35, M = D (xxx+v1cee) , (4.7) M99 = D (lc66 +vxxx) where xxx and K99 are the changes in curvature and D is the flexural rigidity given by 3 12(1—v2) The changes in the circumferential curvature were assumed to be neglible in comparison to xxx so let, K99 = 0. So K99 was eliminated from both equations in Equation 4.8, yielding, 107 Mn = D(Kxx+0) . (4.9) M69 = D (0+vrcxx) Examination of the second of Equations 4.9 reveals the existence of a significant resulting circumferential moment, M99 = Dvrcn, (3.10) shown in red in Figure 4.2. M99 is responsible for the regions of tensile circumferential stress in Figure 3.26, while in the presence of uniform compressive circumferential strains, Figure 3.24. As the cylinder was initially bent toward the “alley”, corresponding to Point A in Figure 3.3, it was being loaded. Then, as it was bent back to go into the “alley”, Point B in Figure 3.3, it was unloaded. The action of unloading the cylinder allows for the tensile stresses while at the same time producing compressive strains. This phenomena is demonstrated in Figure 4.3 which shows a history plot of the leading edge node on the inner surface as it travels through the second stage of the forming process. As can be seen at the zeroth increment the strains were zero and the stresses took on the value of the initial stresses. The first load step of the analysis takes both the strains and stresses into the compressive region and by the one hundredth load step the cylinder has reached the maximum values of strain and compressive stress. This corresponds to the point at which the leading edge was ready to enter the “alley”. Subsequent load steps take the leading edge down the “alley”, and the 108 stress strain curve enters the region of tensile stresses and compressive strains. It has been shown that some critical strain exists within the cylinder. Upon exceeding that strain, any disturbance, whether due to the process or material imperfections, will cause the cylinder to buckle. However, the portion of the cylinder in the “alley” region of the tool does not exhibit the wrinkling phenomena. Thus, it is hypothesized that the wrinkling is due to the combined effects of diameter reduction and severe bending. Additionally, the resultant moment, M99, is responsible for the wrinkles protruding inward rather than outward. While the simple models used up to this point present possible explanations of the wrinkling problem, they do not, however, allow the prediction of the onset of wrinkling. They only provide a means to understand the phenomenon, and, at the same time, the suppositions made allow the possibility to present some recommendations. 4.4 Introduction of Stiffening Rib Based on the level of understanding of the mechanics involved in the formation of wrinkles during the swaging process, a recommendation was proposed as a possible means of alleviating the wrinkling phenomenon. Because the stability of the cylinder was governed by a complex relationship between stress, material properties, and geometry, techniques for increasing the level of stability could be derived from one or more of these factors. After evaluation of several different solution methods, the addition of a circumfer 109 and Carpetsueeeutoow) M1 mm T i / / / / / / 4.0“ mmurwsuungw; Figure 4.3 Circumferential Stress versus Strains for the First Stage of Forming. 110 ential stiffening rib presented itself as the most viable option. In Section 4.2, the cylinder was shown to have a critical circumferential strain above which the cylinder could buckle. In addition the severe bending necessary to accomplish the diameter reduction also contributed to buckling. The combination of these two mechanisms is believed to give rise to wrinkling. The diameter to which the cylinder is reduced is an integral feature of the finished product and cannot be altered. Therefore, the options would be to reduce the compressive/tensile stress gradients or increase the cyl- inder’s resistance to localized bending associated with the wrinkling mode. Stress reduc- tion could be achieved by actually thinning the wall thickness of the cylinder in localized zones. But this would then weaken the resistance of the cylinder to axial loading, for example in the press fitting of a top, and as such this option did not seem to be acceptable. Also, a means of thinning the material in a high volume production setting could not be established. Thus, stress reduction did not appear to be practical. As a result attention was then focused on increasing the cylinder’s resistance to bending. Again, several methods were available to increase bending stiffness. An inspection of the equation which relates the changes in curvature to the bending moments shed light on these possibilities, (4.11) 7: ll Ell: Equation 4.11 indicates that with a constant bending moment either the Young’s modulus, E, or the moment of inertia, I, could be increased to result in smaller changes in curvature. Because the material used produced other desirable characteristics, the decision was to lll attempt to increase the moment of inertia. This concept is reinforced by considering a beam in bending. In Figure 4.4(a) a bending moment is applied to the wide, flat face of the beam, whereas in Figure 4.4(b) the bending moment is applied to the thin edge of the beam. Intuitively, the beam of Figure 4.4(a) should bend more easily than the one in Fig- ure 4.4(b) because of the increased vertical thickness of the beam in Figure 4.4(b). This is confirmed by Equation 4.12 which gives the moment of inertia for the beam, I = 1—2bh . (4.12) where b is the width of the beam and h is the vertical thickness. Thus, Ia for Figure 4.4(a) is smaller than 1b for Figure 4.4(b). Substituting Ia and lb into Equation 4.11 individually, it can be seen that the curvature for the second beam is indeed smaller than that of the first beam. In a flat sheet, a stiffening rib can be used to increase the effective moment of iner- tia (see Figure 4.5). The addition of the central depression in the plate will act as a stiffen- ing rib in the plate and give an increased resistance to bending. These same principles could be applied to the cylinder. For the cylinder, the moment of inertia is a function of both the wall thickness and the (current, or instantaneous) cylinder radius. Increasing the wall thickness has already been deemed undesirable and a global change to the diameter is not feasible. So, a local change in the geometry was proposed to increase the moment of inertia. This would produce the desired increase in bending stiffness, while, at the same time, possibly enhancing the aesthetics for commercial applications. 112 Figure 4.6 shows a model of the cylinder with the addition of a circumferential stiffening rib. The placement of the rib was such that upon the completion of the swaging process the rib would approximately be at the midpoint of the ramped portion into the “alley”. The axialwidth of the stiffening rib was chosen to be one third of the overall length of the ramped portion of the cylinder and the depth was arbitrarily chosen to be one quarter of the width of the rib length. The initial stresses induced by the formation of the stiffening rib were not calculated as the primary purpose of the following numerical results was to observe the effect of the stiffening rib on the forming paths and the final geometry. Figures 4.6 through 4.8 show the initial geometry of the cylinder with the stiff- ening rib, the formed geometry, and the final geometry upon tool release. The final geom- etry has a characteristic stairstep shape. However, despite the apparent large change in geometry, the strain and stress distribution was not significantly affected, as evidenced by comparing Figures 4.9 through 4.12 to Figures 3.23 through 3.26 which did not include the stiffener. Nevertheless, the objective behind the stiffening rib was not reduction in stress and strain level, but an increase in the stability of the critical region. Introduction of the stiffening rib into the final geometry of the cylinder may be achieved in several manners. The simplest would be to use rollers or a stamping operation on the flat sheet before bending into the shape of the cylinder. A consequence of this may be difficulties in obtaining strong weld seams in the swaged zone of the cylinder. The stiff- ening rib may also be produced in the cylinder after the welding process has been finished. One possibility is to use a clamp-like tool with appropriately shaped complimentary roll- ers on opposite sides. The clamp can be closed, rotated around the circumference of the cylinder to create the desired shape, and then removed. Finally, the stairstep shape may be 113 realized by using a multi-stage forming process similar to that of this study to gain addi- tional “steps” in the critical region. For example in the current tooling the amount of diam- eter reduction produced by the first stage tool could be decreased and the translation of the second stage truncated. This would require only small modifications to the current tooling. 114 M _ C\. M (a) Figure 4.4 Simple Beam Subjected to a Bending Moment. M M \q_\‘\] a) (b) ( Figure 4.5 Plate Subjected to a Bending Moment. 115 Figure 4.6 Initial Geometry of Cylinder with Added Stiffening Rib. 116 INC: 503 “I” SIB: 0 w “:11“ FREOHMOOOo-N Figure 4.7 Formed Shape of the Cylinder with Added Stiffening Rib. 117 iiii i!” \ m\\ . L... 0 1 L Figure 4.8 Final Shape of the Cylinder with Added Stiffening Rib after Elastic Snapback. 118 lab! taco-monarchs. 1 Figure 4.9 Axial Strains in the Critical Region of the Cylinder with Added Stiffening Rib. 119 Ii" it" ~‘-'. ‘ -1 w W m 4.1m am one a“ v L_J lebi ace-parental» 1 Figure 4.10 Circumferential Strains in the Critical Region of the Cylinder with Added Stiffening Rib. 120 “tee-potatoes 1 Figure 4.11 Axial Stresses in the Critical Region for the Cylinder with Added Stiffening Rib. 121 41M “Carpet..- 1 Figure 4.12 Circumferential Stresses in the Critical Region for the Cylinder with Added Stiffening Rib. Chapter 5 Summary and Future Work 5.1 Summary The stated objectives of the project were to determine the deformations and the stress state of the cylinder during the swaging operation. Each of these goals has been accomplished through the use of a numerical simulation of each step in the manufacturing process. Simple examples were then used to interpret the results obtained from the analy- ses. From the explanations offered, a hypothesis of the mechanism responsible for the for- mation of the wrinkles was presented. Finally, candidate solution methods were proposed, with the most reasonable being partially studied using in the finite element model. Throughout this simulation the cost was in man time to develop the model and to subse- quently run the model on the computer. At no time was any physical simulation per- formed, thereby yielding a cost reduction in raw materials for test construction of the 122 123 cylinder and the forming tools. In addition, any follow up work or additional proposals can be performed using the developed model as a basis. Shortcomings have, however, been demonstrated within the model. The use of element 10, which is based on a continuum formulation, has proven to be inherently expensive for the type of analysis being done, as pointed out by Rebolo, et al. (1990). In the initial convergence studies done to confirm the most appropriate element for use in the simulation, element 10 required a large number of elements to reach the exact solutions sought, while the shell elements converge to the closed form solution quickly. Again, in the convergence studies performed to determine mesh size, conver- gence was slow and actually never achieved using element 10. Although acceptable solu- tions were found in each case, the computational cost involved was considerable. The conclusion must be drawn that an alternative should be found. The model developed does not predict the formation of wrinkles in the cylinder. This does not diminish the results obtained up to this point, but only presents the opportunity for further work. 5.2 Future Work It is proposed to develop a new finite element based on nonlinear shell theory and a novel formulation that appears to be ideally suited for analysis of sheet metal forming processes. The theory is based on a displacement field that allows a cubic distribution through the thickness of the inplane displacement components and satisfies the shear trac- tion boundary conditions on the top and bottom surfaces of the shell. For cylindrical shells undergoing axisymmetric deformations, the displacement field takes the form, 124 2 3 z 42 (1W “x (xx) = u (x) 4.wa (x) '1' Zk51+3—h2[k52—WX(x) -5] 2 3 (5-1) uy (x,z) = v (x) + 2% (x) + Zzku "' 3411-2 [1‘42 _ ‘15”) — Y%] uZ (x,z) = w(x) where ux, try, and uz are the displacements of a point in the shell in the x (axial), y (circum- ferential), and z (thickness) directions respectively, and, 1 1‘ 1" 1 1‘ “I" k =_ 5 _ 5 k =_ 5 + 5 51 2( (N) (1)] 52 2[ (N) (1)] Q55 Q55 Q55 Q55 t b 1 T4 T4 t b (5.2) k“ = 2[Q(N) 'Qm) k42 = 2[Q(N) +Q11) 44 44 44 44 1 b . . t4 , I; and 1;, 1:5 are the shear tractrons on the top and bottom surfaces of the shell in x and y directions, respectively and Q 3) , Q :4") and Q5? , Q55?) are the transverse shear stiffnesses in the first and Nth layer of the shell in the yz and xz planes, respectively. The shear tractions must be known, and may be due to friction between the tool and the cylin- der. In the current analysis friction was ignored, so 1251 = k52 = k4, = 1:42 = 0. The nonlinear shell theory based on the above displacement field allows for trans- verse shearing effects and a nonlinear distribution through the thickness of deformations due to bending. Such higher order effects are usually neglected in the analysis of tirin- walled structures. However, these effects may be important for nonlinear deformations, 125 and the inclusion of these effects allows the satisfaction of shear tractions on the top and bottom surfaces of the shell, yielding a more accurate physical description of the problem. This element, once formulated, could be used in the current analysis to provide a very accurate, efficient, and robust model for thin cylinders undergoing axisymmetric deforma- tions. The use of the new element would remove the current barriers to reaching a con- verged solution, but does not allow for the prediction of wrinkling in the swaging process. To accomplish this, a method similar to that proposed by Adams (1993) could be employed. Once a converged solution to the axisymmetric forming problem has been found the analysis could be rerun, seeking the onset of wrinkling. This could done by checking the solution after each stage of the forming process, to determine if circumferen- tial buckling has occurred. Tracing the determinant of the characteristic system of equa- tions for cylindrical shells to determine when (or if) it goes to zero would indicate that the cylinder has buckled or wrinkled. The characteristic system of equations could be formulated from the non- axisymmetric form of the governing equations of the nonlinear shell theory discussed above. The circumferential variations of the deformations, in this case, could be approximated using a Fourier series method. In this method, the problem is analyzed for several independent harmonics which are superimposed to give a final solution. Because of the independence of the harmonics only those harmonics which influence the buckling of the cylinder need be analyzed. Use of this method is possible because the circumferential direction (the rolling direction) is also a principal material direction, and the cylinder geometry is axisymmetric. Therefore, each of the five degrees of freedom may be represented by a Fourier series in y, 126 ny uinPl. (x) cos? Ms .MN I u(x.y) = 2 17:,(x)COS'-'R-y- = n=1 3 II N. ~ ll - 0° m 2 v(x, y) = 2 17;!(x)sin'% = 2 z PEP, (x) sin-{1R3 n = 1 n = Ii = I co m 3 w(x, y) = 2 uTn(x)c0s’% = 2 2 EN, (x) cos? (5.3) n = 1 n = 11': I _ — 22 ._. — . '2’ ‘11,“: 1’) - 2 Wxn(x)005 R WxinPl (x) cos R n=I 11MB 3M5 ny "Pl. (x) cos 7?- MNEMe :5 W,(x. y) = 2 \7; (x) 603% = n=1 3 N - ll 1... where m is the number of Fourier harmonics, P,- are the linear Lagrange interpolation functions, and N,- are the quadratic Lagrange interpolation functions. In the above equations the DOF are approximated in the circumferential direction by the product of a trigonometric function and an amplitude term, with the axial variation of the amplitudes approximated by the finite element method. Also of note is the approximation w“. In this case, an interdependent interpolation scheme is employed to allow for greater accuracy without an increase in computational cost. The third DOF in the element approximation of Wu would be eliminated before assembly by imposing a constraint on the variation of transverse shear strain (see Averill, 1994 and Tessler, 1983) Due to the form of the displacement approximations in Equation (5.3), integration in the thickness and circumferential directions can be performed analytically, so that the model is reduced to one dimension (along the axis of the cylinder). The final model is given as, 127 16.1: = p” (5.4) J j where n is the Fourier harmonic, K the global stiffness matrix, d the Fourier series amplitude, and F the force matrix. The characteristic system of equations is obtained for each harmonic to measure the stability of the current configuration. When the determinant goes to zero for a given harmonic, buckling is predicted to occur, with a buckling pattern wavelength related to n. The development of the axisymmetric, nonlinear shell element is currently being con- ducted by other researchers. Upon completion of the development, the element can be implemented into the current model for solution of the axisymmetric forming problem. The solutions from this analysis must then be coupled with the proposed wrinkling analy- sis formulation to complete the solution. APPENDIX 128 A.l Mesh Refinement Study Data In Section 3.3.1, an elemental convergence study was conducted on the first stage of the forming process. In this study a series of meshes were analyzed using the developed model. From those analyses data was collected on the outer, mid-plane, and inner surfaces of the cylinder at locations within the first bending region and the critical region. Tables A.1 through A.24 contain all of that data collected. Table A.l Axial Strain on the Outer Surface in the First Bending Region. (x10'3) Elements 2 3 4 5 6 8 150 14.3964 15.2937 15.2429 15.2084 15.2017 15.2091 300 14.2752 15.6635 15.5345 15.6663 15.6150 15.7637 400 ----- ----- 15.6079 15.7318 15.8140 ' ----- Table A.2 Axial Strain on the Mid-Plane Surface in the First Bending Region. (x10'3) Mesh 2 3 4 5 6 8 150 0.838835 1.233395 0.997778 0.98309 0.953442 0.923100=H 300 0.645666 1 .45242 1.04027 1 .00798 0.990998 0.884100 400 ----- mu 1 . 12246 1 . 158305 1 .080070 ----- 129 Table A.3 Axial Strain on the Inner Surface in the First Bending Region. (x10'3) Mesh 2 3 4 5 6 8 150 43.2124 43.1500 43.5015 43.4648 43.4725 43.5070 300 43.4868 43.0982 43.7462 43.9155 43.9076 44.1870 400 ---------- 43.6394 43.6624 43.8760 ------ Table A.4 Circumferential Strain on the Outer Surface in the First Bending Region. (x10'3) Mesh 2 3 4 5 6 8 150 -6.98879 -6.93845 -6.91135 -6.96230 -6.97252 -6.64920 300 -6.9025 8 -6.8201 1 -6.70422 -6.66099 -6.68490 -6.63974 400 ---------- -6.85317 -6.85937 -6.81014 ----- Table A.5 Circumferential Strain on the Mid-Plane Surface in the First Bending Region. (x10'3) Mesh 2 3 4 5 6 8 150 -6.92393 -6.87485 -6.84736 -6.899825 -6.90983 -6.91335 300 -6.83654 -6.75465 -6.63764 -6.59537 -6.6l930 -6.57391 400 ---------- -6.78859 -6.79609 -6.74601 ---—- 130 Table A.6 Circumferential Strain on the Inner Surface in the First Bending Region. (x103) Mesh 2 3 4 5 6 8 150 6.89687 6.84650 -6.82324 6.87584 -6.88715 6.89121 300 6.80847 6.72467 6.68394 6.56921 6.59467 6.54976 400 ----- 6.76480 6.77228 6.72332 ----- Table A.7 Axial Stress on the Outer Surface in the First Bending Region. (szi) Mesh 2 3 4 5 6 8 150 240.268 187.504 158.886 138.823 126.935 111.188 300 240.424 191.156 162.397 144.685 132.495 115.812 400 ---------- 157.724 140.057 128.731 ----- Table A.8 Axial Stress on the Mid-Plane Surface in the First Bending Region. (szi) Mesh 2 3 4 5 6 8 150 ' 0.17758 -4.8125 W 8.37254 -10.2651 -11.9106 300 1.14157 -4.7234 -65.3091 -8.6686 -10.8890 -13.7535 400 ---------- 45.8308 -6.54295 -8.965.84 ------ 131 Table A.9 Axial Stress on the Inner Surface at in the First Bending Region. (szi) Mesh 2 3 4 5 6 8 T50 -268.912 -208.256 476.320 453.901 440.543 422.824 300 -271.538 -213.056 480.558 459.968 446.372 427.406 400 ---------- 471.748 452.187 439.352 ----- Table A.10 Circumferential Stress on the Outer Surface in the First Bending Region. (szi) Mesh 2 3 4 5 6 8 150 155.628 106.460 787539—385953— 47.0817f 31.7309 300 156.503 109.640 81.7260 64.5423 52.8537 362061 400 ----- 75.4760 58.2574 47.2932 ----- Table A.11 Circumferential Stress on the Mid-Plane Surface in the First Bending Region. (szi) Mesh 2 3 4 5 6 8 == 1: 3;: 150 -55.8569 -56.7963 -62.6269 -6l4201 -64.9001 -657393 300 -58.4541 -58.0932 -65.3584 -63.0065 -67.8407 -69.l656 400 ---------- 65.8639 63.9364 68.2242 --..- 132 Table A.12 Circumferential Stress on the Inner Surface in the First Bending Region. (szi) Mesh 2 3 4 5 6 8 150 -256.441 — 198.253 —165.400 - 145.595 -131795 -114095 300 -258.999 -199.263 -167.479 - 147.434 -132.657 -116.286 400 ----- ----- - 172.804 7151.758 - 138.334 ---- Table A.13 Axial Strains on the Outer Surface in the Critical Region. (x10'3) Mesh 2 3 4 5 6 8 150 -4. 18674 -1.04515 -0.278518 0.101525 0.336483 0.644162 300 -.35759 1.75286 0.14918 0.551638 0.517526 0.929506 400 ----- ----- -0.333748 0.122887 0.424808 ----- Table A.14 Axial Strains on the Mid-Plane Surface in the Critical Region. (x10'3) Mesh 2 3 4 5 6 8 150 17.8305 1W T9696 :- 18.9899 19.0050 18.9928= 300 18.2751 23.7753 18.8816 18.8507 19.3794 19.2492 400 ----- ----- 19.2907 19.3346 19.3919 ---- 133 Table A.15 Axial Strains on the Inner Surface in the Critical Region. (x10‘3) Mesh 2 3 4 5 6 8 150 38.8382 38.3533 37.7740 37.6701 37.6560 37.5003 300 39.1136 44.6235 37.4199 37.1498 38.2427 37.8484 400 ---------- 38.5379 38.4350 38.4716 ----- Table A.16 Circumferential Strains on the Outer Surface in the Critical Region. (x103) Mesh 2 3 4 5 6 8 150 62.3648 -52.3186 -52.—3061 -52.3227 62.3256 62.3274 300 62.3355 62.4560 61.7063 61.6914 63.7710 62.3589 400 ----- 62.2466 62.2544 62.4124 ----- Table A.17 Hoop Strains on the Mid-Plane Surface in the Critical Region. (x10'3) Mesh 2 3 4 5 6 8 150 -52.4591 -52.4024 -52.3916 -52.4067 -52.4104 -52.4121 300 -52.4281 -52.5357 -51.7063 -51.7540 -52.4607 -52.4428 400 ---------- -52.3256 -52.3312 -52.4950 ----- 134 Table A.18 Hoop Strains on the Inner Surface in the Critical Region. (x10'3) Mesh 2 3 4 5 6 8 150 -52.4937 -52.4353 -52.4211 -52.4368 -52.4393 -52.4409 300 -52.4615 -52.5579 -51.7774 -51.7640 -52.4881 -52.4710 400 ----- ----- -52.3482 -52.3537 -52.5213 ----- Table A.19 Axial Stresses on the Outer Surface in the Critical Region. (szi) Mesh 2 3 4 5 6 8 150; -360.213 -246.750 -197.342 -171.537 -153.706 -127.704 300 -354.848 -232.841 —208.305 - 177.613 -155.114 -130.090 400 ----- ----- -216.718 -186.542 - 164.428 ----- Table A.20 Axial Stresses on the Mid-Plane Surface in the Critical Region. (szi) Mesh 2 3 4 5 6 8 I = r I 150 12.007 4.2907 2.18321 1.0688 -0.00944 -2.0748 300 12.5069 4.06294 5.9629 4.2178 1.6014 0.0673 400 ---------- 6.5923 4.3753 3.7378 ----- 135 Table A.2] Axial Stresses on the Inner Surface in the Critical Region. (szi) Mesh 2 3 4 5 6 8 150 335.940 226.770 180.082 157.769 140.193 117.602 300 331.314 266.730 197.465 168.970 144.993 122.361 400 ---------- 201.974 174.301 154.099 ----- Table A.22 Circumferential Stresses on the Outer Surface in the Critical Region. (szi) Mesh 2 3 4 5 6 8 150 -394. 100 -284.650 -233.855 -204.445 - 185.684 - 160.038 300 -397.440 -282.371 -235.476 -204.863 -189.751 - 162.522 400 ----- ----- -238.453 -209. 154 - 188.076 ----- Table A.23 Circumferential Stresses on the Mid-Plane Surface in the Critical Region. (szi) Mesh 2 3 4 5 6 8 150‘ 68.99% 63.8572 68.5703 67.9680 -70.7456 20.9200 300 69.7367 20.7880 67.0310 66.2861 21.3535 22.1101 400 136 Table A.24 Circumferential Stresses on the Inner Surface in the Critical Region. (szi) Mesh 2 3 4 5 6 8 150 253.175 155.333 109.250 83.8303 66.7538 44.0962 300 247.643 204.532 119.618 91.9736 65.4983 43.4282 400 ---------- 123.298 96.5859 76.1358 ----- 137 A.2 Analysis Results without Strain Hardening The effects of strain hardening on the analysis results were evaluated in Section 3.4.2. It was determined at that time the effects were minimal at most. This may be con- formed by comparing the plots that follow in Figure A.1 - A.4 to those of Figures 3.23 - 3.26. Doing so will show that while the magnitudes of the strain and stress values have small incremental changes, the trends remain constant. Therefore, strain hardening effects are categorized as minimal. 138 you Incend‘l’ehl“ I Figure A.1 Axial Strains in the Critical Region Without Work Hardening. 139 4‘1.“ 4“ ‘1 ado-garment. 1 Figure A.2 Circumferential Strains in the Critical Region Without Work Hardening. I‘M“... 1 Figure A.3 Axial Stresses in the Critical Region Without Work Hardening. 141 1.51 “001.01“ 1 Figure A.4 Circumferential Stresses in the Critical Region Without Work Hardening. REFERENCES 142 Adams, G. G., (1993),“Elastic Wrinkling of a Tensioned Circular Plate Using von Karman Plate Theory”, Journal of Applied Mechanics, 60. pp. 520 - 525 Averill, R. C. and Reddy, J. N., (1990), “Behaviour of Plate Elements Based on the First Order Shear Deformation Theory”, Engineering Computing, vol. 7, pp. 57-74 Averill, R. C., (1994) “Static and Dynamic Response of Moderately Thick Laminated Beams with Damage”, Composites Engineering, vol. 4, pp. 381-395 Butman, J ., (1991),Car Wars, Grafton Book, London, p 11 Chan, S. L., (1993), “A Non-Linear Numerical Method for Accurate Determination of Limit and Bifurcation Points”, International Journal for Numerical Methods in Engineering, vol. 36, pp. 2779 - 2790 Dym, C. L., (1974), Introduction to the Theory of Shells, Pergamon Press, Oxford, p. 26. Eterovic, A. L. and Bathe, K. J ., (1991), “On the Treatment of Inequality Constraints Arising form Contact Conditions in Finite Element Analysis”, Computers and Structures,vol.40, pp. 203-209 Green, C. M., (1956), Eli Whitney and the Birth of American Technology, Little, Brown and Company, Boston, p. 120 Kalpakjian, S., (1991), “Manufacturing Processes for Engineering Materials”, Second Edition, Addison-Wesley Publishing Company, Reading, MA, pp. 461-462 Keck, P., Wilhelm, M., and Lange, K., (1991), “Application of the Finite Element Method to the Simulation of Sheet Forming Processes: Comparison of Calculations and Experiments”, International Journal for Numerical Methods in Engineering, vol. 30, pp. 1415-1430 Lee, J. K., Cho, U. Y., and Hambrecht, J., (1991), “Recent Advances in Sheet Metal Forming Analysis”, Advances in Finite Deformation Problems in Materials Processing and Structures, ASME 1991, AMD vol. 125 MARC User’s Guide Volume A: User Information, (1994), MARC Analysis Research Corporation, Palo Alto, CA Mendelson, A., (1968), Plasticity: Theory and Application, Robert E. Krieger Publishing Company, Malabar, FL. Mentat H Command Reference, (1994), MARC Analysis Research Corporation, Palo Alto, CA 143 Neale, K.W. and Tugcu, P., (1990), “A Numerical Analysis of Wrinkle Formation Tendencies in Sheet Metals”, International Journal for Numerical Methods in Engineering, vol. 30, pp. 1595-1608 Padovan, J., (1974), “Quasi-Analytical Finite Element Procedures for Axisymmetric Anisotropic Shells and Solids”, Computers and Structures, vol. 4, pp. 467 - 483 Parr, G., (1958), Man, Metals and Modern Magic, American Society for Metals, Cleveland and The Iowa State University Press, Ames, Iowa Rebolo, N ., Nagetaal, J. C., and Hibbitt, H. D., (1990), “Finite Element Analysis of Sheet Forming Processes”, International Journal for Numerical Methods in Engineering, vol. 30. PP. 1739 - 1758 Reddy, J. N., ( 1984), Energy and Variational Methods in Applied Mechanics, John “Wiley and Sons, New York Riccobno, R., (1992), “A Comparison Between Two Different Models for Metal Forming FEM Analysis”, Computers and Structures, vol. 44, pp. 429-433 Sedaghat, M. and Herrmann, L. R., (1983), “A Nonlinear, Semi-Analytical Finite Element Analysis for Nearly Axisymmetric Solids”, Computers and Structures, vol. 17, no. 3, pp. 389 - 401 Simo, J. C., Wriggers, P., and Taylor, R. L., (1985), “A Perturbed Lagrangian Formulation for the Finite Element Solution of Contact Problems”, Computer Methods in Applied Mechanics and Engineering, vol. 50, pp. 163 - 180 Tessler, A. and Hughes, T. J. R., (1983), “An Improved Treatment of Transverse Shear in the Mindlin-Type Four-N ode Quadrilateral Element”, Computer Methods in Applied Mechanics and Engineering, vol. 39, pp. 311-335 Toh, C. H., (1989), “Prediction of the Forming Limit Curves of Sheet Materials Using the Rigid-Plastic Finite Element Method”, International Journal of Machine Tools and Manufacture, vol. 29, pp. 333-343 Wood, M. D., (1994), “The Effect of Microstructure and Texture on the Mechanical Properties of Double-Reduced Tinplate Steel”, Master’s Thesis, Michigan State University, East Lansing, Michigan Zhang, L. C. and Yu T. X., (1988), “The Plastic Wrinkling of an Annular Plate Under Uniform Tension on Its Inner Surface”, International Journal of Solid Structures, vol 24, no. 5, pp. 497 - 503 Zienkiewicz, O. C. and Taylor, R. L., (1989), The Finite Element Method, McGraw-Hill, New York