., .1 Z. 5 . . :5 ‘r 1:1... Eva}. V .1 . . 1, 0.5.3.?» -1 4h. 2.-. This is to certify that the dissertation entitled Shape and Laminate Optimization of Fiber-Reinforced- Polymer Structures presented by Jun Wu has been accepted towards fulfillment of the requirements for the Department of Civil and Ph'D' degree In Environmental Engineering Major Profess 3 Signature 5/2/04 Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cz/CIRC/DateDuepGS-pts Shape and Laminate Optimization of Fiber-Reinforced-Polymer Structures By Jun Wu A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Civil and Environmental Engineering 2004 ABSTRACT Shape and Laminate Optimization of Fiber-Reinforced-Polymer Structures By Jun Wu Fiber Reinforced polymer (FRP) composites, or advanced composite materials, which have shown outstanding mechanical characteristics, such as high strength-to-weight and. high stiffness-to—weight, and high chemical and environmental endurance compared to conventional materials, have been adapted from aerospace and defense industries to civil infrastructure applications. The nature of FRP composites makes them strong and stiff in planes along fiber orientations but rather weak through their thickness. Thus, they are most efficient when used under global in-plane stress demands. Due to their high material costs, a way to employ their high material stiffness and strength is to use them in structural forms in which the structural efficiency is maximized by shaping the structural geometry to achieve mainly in-plane behavior under the applied loads. Therefore, an approach is needed to develop the structural forms where the in-plane properties of FRP composites can be used efficiently. An integrated approach is developed for the shape and laminate optimization of FRP structures. The integrated approach is implemented in a two-level uncoupled procedure: shape and laminate—property optimization followed by laminate design optimization, where structural shape and laminate design are optimized simultaneously with the aim of maximizing structural stiffness. Examples of optimizing laminated FRP shell structures are provided to validate the procedure. The performance of the integrated approach is evaluated by comparing a shape-and-material optimized FRP shell with two shape-only optimized FRP shells. The numerical results show that the proposed integrated approach is reliable and provides a useful tool for designing optimized laminated FRP structures. The integrated approach is also used to aid the rational implementation of FRP composites in civil infrastructure by developing innovative bridge design concepts. Innovative FRP composite membrane—based bridge systems were explored through analytical studies of the integrated approach. Two types of bridge systems are developed, FRP membrane beam bridges and FRP membrane suspension bridges. Both bridge systems consist of a shape-and-laminate optimized FRP membrane/shell carrying the in— plane tensile and shear forces together with a conventional reinforced concrete deck providing the live load transfer. The analytical studies of optimizing the FRP membrane bridges provide the initial development of innovative systems that use FRP composites in their inherent behavioral characteristics for new high-performance structures. Results from this work demonstrate that FRP composites can be used with higher efficiency in new structural systems as long as their advantageous properties of directional strength, lightweight, and tailored properties are properly considered in the design process. The work further discusses the feasible implementation of the optimum design for composite membrane-based bridges in practical engineering construction. The work also provides insight to further development and applications of the integrated Optimization approach. ACKNOWLEDGMENTS I would like to thank my advisor, Professor Rigoberto Burguefio for his excellent academic guidance, continuous encouragement, and patient support throughout this work. It has been a very enlightening and fruitful experience to work with Dr. Burguefio. My sincere thanks are also extended to the other members of my Ph.D. committee, Professor Ronald S. Harichandran, Professor Charles R. MacCluer, and Professor Amit H. Varma. I have my deep gratitude for Mr. Graig Gunn’s selfless help and tremendous assistance during my studies at Michigan State University. I am also grateful for all the help granted to me throughout the course of my work by the faculty and staff members, colleagues and many friends. I would like to acknowledge the financial support by Michigan State University through the Department of Civil and Environmental Engineering and an Intramural Research Grant Program award from the Office of the Vice President for Research and Graduate Studies. Grateful thanks should also go to my parents. Their love, encouragement and confidence in my abilities are the backbone of my persistence in the endeavor. Most of all I would like to thank my wife, Yue. Her endless love is the source of my dedication through the completion of this research. iv TABLE OF CONTENTS LIST OF FIGURES _ -- - viii LIST OF TABLES -- - -- xii Shape and Laminate Optimization of Fiber-Reinforced-Polymer Structures ............ 1 1 Introduction - 1 1.1 General ................................................................................................................ 1 1.2 FRP composites for civil infrastructure .............................................................. 1 1.2.1 Introduction to FRP composites .................................................................. 2 1.2.2 Mechanical properties of FRP laminates .................................................... 2 1.2.3 FRP composites in civil structures .............................................................. 4 1.3 Outstanding issues and motivation ..................................................................... 6 1.4 Shape and material optimized FRP structures .................................................... 7 1.4.1 Shape resistant structures and optimal shape finding ................................. 8 1.4.2 FRP laminates and optimal laminate design ............................................... 9 1.4.3 Shape and laminate optimization of FRP structures ................................. 10 1.5 Objective and scope of the dissertation ............................................................ 11 References ..................................................................................................................... 13 2 Laminated Fiber Reinforced Polymer (F RP) Composites - 14 2.1 Constitutive relations for FRP laminates .......................................................... 15 2.2 Stacking sequence of laminates ........................................................................ 22 2.3 Constrained relations of lamination parameters ............................................... 24 2.3.1 Relation between in—plane lamination parameters .................................... 24 2.3.2 Relation between out-of—plane lamination parameters ............................. 27 2.3.3 Relation between in-plane and out-of-plane lamination parameters ........ 29 2.4 Summary ........................................................................................................... 34 References ..................................................................................................................... 35 3 Integrated Optimization Process for Laminated FRP Structures ..................... 36 3.1 Form finding and structural shape optimization ............................................... 37 3.1.1 Form finding of tension membrane structures .......................................... 37 3.1.2 Form finding of shell structures ................................................................ 41 3.1.3 Structural shape optimization ................................................................... 45 3.2 Material optimization of laminated FRP composites ........................................ 52 3.3 Integrated shape and FRP laminate optimization ............................................. 55 References ..................................................................................................................... 58 4 Shape and Laminate-Property Optimization of FRP Structures 60 4.1 Formulation of the shape and laminate-property optimization problem .......... 60 4.1.1 Design variables ........................................................................................ 60 4.1.2 Objective function ..................................................................................... 61 4.1.3 Constraints ................................................................................................ 63 4.2 Algorithm of shape and laminate—property optimization .................................. 64 4.2.1 Linearization of objective and constrain functions ................................... 64 4.2.2 Definition and determination of a searching direction ............................. 65 4.2.3 Definition of the constrained steepest descent function ........................... 67 4.2.4 Determination of the step size ................................................................... 68 4.3 Implementation of shape and laminate—property optimization ......................... 69 References ..................................................................................................................... 71 5 FRP Laminates Design Optimization 72 5.1 Formulation of laminate design optimization ................................................... 72 5.1.1 Design variables ........................................................................................ 72 5.1.2 Objective function ..................................................................................... 74 5.1.3 Constraints ................................................................................................ 74 5.2 Algorithm for laminate design optimization ..................................................... 74 5.2.1 Design representation by genetic coding .................................................. 75 5.2.2 Definition of a population size .................................................................. 76 5.2.3 Definition of fitness values and a selection scheme ................................. 76 5.2.4 Implementation of genetic operators ........................................................ 77 5.3 Implementation of laminate design optimization ............................................. 78 References ..................................................................................................................... 80 6 Shape and Laminate Optim'nation of FRP Shells 81 6.1 Shape and laminate optimized FRP shell .......................................................... 81 6.2 Shape-only optimized laminated FRP shells .................................................... 86 6.3 Comparison of three optimal shells .................................................................. 89 6.4 Buckling analyses of optimized laminated FRP shells ..................................... 95 6.5 Summary ........................................................................................................... 97 References ..................................................................................................................... 98 7 FRP Composite Membrane-Based Bridge Systems- -- -- 99 7.1 FRP composite membrane beam (CMB) bridges ........................................... 100 7.1.1 Bridge system description ....................................................................... 100 7.1.2 Integrated optimization of CMB bridges ................................................ 103 7.1.3 System characteristics of optimal CMB bridges ..................................... 115 7.1.4 Post optimality analyses .......................................................................... 121 7.2 FRP composite membrane suspension (CMS) bridges ................................... 127 7.2.1 Bridge system description ....................................................................... 128 7.2.2 Integrated optimization of CMS bridges ................................................. 130 7.2.3 System characteristics of Optimal CMS bridges ..................................... 138 7.2.4 Post optimality analyses .......................................................................... 142 7.3 Discussion of FRP membrane-based bridge systems ..................................... 150 7.3.1 Comparison of FRP membrane-based bridge systems ........................... 150 7.3.2 Discussion of the objective function ....................................................... 154 7.3.3 Selection of the design variables ............................................................. 158 vi 7.4 Summary ......................................................................................................... 159 References ................................................................................................................... 16 l 8 Conclusions and Future Research Needs ..... 162 8.1 Conclusions ..................................................................................................... 162 8.2 Future research needs ...................................................................................... 163 8.2.1 About the integrated approach of shape and laminate optimization ....... 163 8.2.2 About applications of the integrated approach ....................................... 165 8.2.3 About the engineering and construction of optimal FRP composite structures ................................................................................................ 166 vii LIST OF FIGURES Figure 1—1 FRP elastic modulus as a function of fiber orientation .................................... 4 Figure 1—2 Applications of FRP composites in civil infrastructures .................................. 5 Figure 1—3 Shape resistant structures [Otto, 1969] ............................................................. 8 Figure 2—1 Laminated FRP composites ............................................................................ 18 Figure 2—2 Feasible domain of in—plane lamination parameters ....................................... 27 Figure 2—3 Feasible domain of out-plane lamination parameters ..................................... 28 Figure 2—4 Boundary determination of VID for a given {V1A, V3A} .................................. 31 Figure 2—5 Boundary determination of V30 for a given {V1A, V3A, V10} .......................... 32 Figure 2—6 Feasible domain of combining in-plane and out-plane lamination parameters ................................................................................................................................... 33 Figure 3—1 Shape resistant structures ............................................................................... 36 Figure 3—2 Soap film analogy method and tension membrane structures and [Hildebrandt and Tromba, 1983] .................................................................................................... 38 Figure 3—3 Hanging method and shell structures and [Isler, 1991, 1994] ........................ 43 Figure 3—4 Structural shape optimization [Ram, 1992] ................................................. 45 Figure 3—5 Structural shape optimization process ............................................................ 47 Figure 3—6 Two-level approach of shape and laminate optimization ............................... 56 Figure 4—1 Shape and laminate-property optimization ..................................................... 69 Figure 5—1 Typical symmetric and balanced laminate ..................................................... 73 Figure 5-2 Design representation by genetic coding ........................................................ 76 Figure 5—3 Crossover operator .......................................................................................... 77 Figure 5—4 Mutation operator ........................................................................................... 78 Figure 5—5 Laminate design optimization ........................................................................ 79 viii Figure 6—1 Geometry, loading, and control points for shell structures ............................ 82 Figure 6—2 History of objective function - FRP Shell 1 .................................................. 84 Figure 6—3 Histories of partial geometric design variables — FRP Shell 1 ....................... 84 Figure 64 Histories of laminate—property design variables — FRP Shell 1 ..................... 85 Figure 6—5 Second-level optimization process — FRP Shell 1 .......................................... 86 Figure 6—6 History of objective function —- FRP Shell 2: [45°/—45°/—45°/45°]65 .............. 87 Figure 6—7 Histories of geometric design variables — FRP Shell 2: [450/-450/-450/450]63 ................................................................................................................................... 88 Figure 6—8 History of objective function — FRP Shell 3: [0°/90°/90°/0°] ........................ 88 Figure 6—9 Histories of geometric design variables — FRP Shell 3: [0°/90°/90°/0°] ........ 89 Figure 6—10 Global and local coordinate systems ............................................................ 91 Figure 6—11 Structural responses of initial and optimal shells ......................................... 92 Figure 6—12 Membrane force distribution of optimal shells ............................................ 94 Figure 6—13 Bending moment distribution of Optimal shells ........................................... 94 Figure 7—1 Composite membrane beam (CMB) bridge ................................................. 100 Figure 7—2 Computational model of the CMB bridge .................................................... 101 Figure 7—3 Geometric key points for CMB bridge shape optimization .......................... 104 Figure 7—4 History of objective function — CMB Model A ............................................ 106 Figure 7—5 History of geometric design variables — CMB Model A .............................. 107 Figure 7—6 History of laminate-property design variables — CMB Model A ................. 107 Figure 7—7 History of objective function — CMB Model B ............................................ 108 Figure 7—8 History of geometric design variables — CMB Model B .............................. 108 Figure 7—9 History of laminate-property design variables — CMB Model B ................. 109 Figure 7—10 History of objective function — CMB Model C .......................................... 109 Figure 7—11 History of geometric design variables — CMB Model C ............................ 110 Figure 7—12 History of laminate-property design variables — CMB Model C ............... 110 Figure 7-13 Laminate design optimization for 8—layer laminates .................................. 113 Figure 7—14 Laminate design optimization for 48-layer laminates ................................ 114 Figure 7—15 Section force (SFl) distribution ................................................................. 118 Figure 7—16 Section force (SF2) distribution ................................................................. 118 Figure 7—17 Shear section force (SF3) distribution ........................................................ 119 Figure 7—18 Load case for maximum torsion in the strength limit state ........................ 126 Figure 7—19 Load case for maximum bending in the strength limit state ...................... 126 Figure 7—20 Concept for Composite Membrane Suspension (CMS) Bridges ................ 128 Figure 7—21 Geometry of the CMS bridge ..................................................................... 129 Figure 7—22 Loading and geometric key points for the CMS bridge ............................. 131 Figure 7—23 Histories of geometric design variables — CMS Model A .......................... 133 Figure 7—24 Histories of laminate-property design variables — CMS Model A ............. 133 Figure 7—25 History of objective function — CMS Model A .......................................... 134 Figure 7—26 Histories of geometric design variables — CMS Model B .......................... 134 Figure 7—27 Histories of laminate-property design variables — CMS Model B ............. 135 Figure 7—28 History of objective function — CMS Model B .......................................... 135 Figure 7—29 Second-level optimization process of Model A ......................................... 137 Figure 7—30 Second-level optimization process of Model B .......................................... 138 Figure 7—31 Coordinate systems ..................................................................................... 139 Figure 7—32 Loading pattern of spaced line loads .......................................................... 145 Figure 7—33 Section force (SFl) distribution in CMS optimal designs .......................... 149 Figure 7—34 Section force (SF2) distribution in CMS optimal designs .......................... 149 Figure 7—35 Structure material distribution of relative stress index ............................... 153 Figure 7—36 Performance of different objective functions for optimal CMB bridge ..... 157 xi LIST OF TABLES Table 2—1 A*, B*, and D* matrices in terms of lamina invariants and lamination parameters ................................................................................................................. 21 Table 6—1 Optimal laminate designs of FRP shells .......................................................... 86 Table 6—2 Summary of strain energies and displacements of optimal shells ................... 89 Table 6—3 Summary of structural responses of initial and optimal shells ........................ 91 Table 6—4 Buckling analyses of Shell 1 ............................................................................ 96 Table 6—5 Buckling analyses of Shell 2 ............................................................................ 96 Table 6—6 Buckling analyses of Shell 3 ............................................................................ 97 Table 7—1 Load case specification for CMB bridge optimization .................................. 101 Table 7—2 Dimensional constraints for geometric design variables ............................... 105 Table 7—3 Initial designs of CMB Bridges ..................................................................... 106 Table 7—4 Optimum results of shape and material-property optimization ..................... 111 Table 7—5 Optimal laminates from laminate design optimization process ..................... 115 Table 7—6 Section forces for the optimal CMB bridge ................................................... 117 Table 7—7 Section moments for the optimal CMB bridge .............................................. 117 Table 7—8 Buckling eigenvalues and eigenvectors of the optimal CMB bridge ............ 120 Table 7—9 Structural response of optimal CMB bridge with different optimal laminates ................................................................................................................................. 122 Table 7—10 Geometric variations of optimal CMB bridges ............................................ 123 Table 7—11 Laminate stresses of the optimized CMB bridge under different limit states ................................................................................................................................. 125 Table 7—12 Stress indices of the optimized CMB bridge under different limit states.... 125 Table 7—13 Laminate stresses of the optimized CMB due to different extreme effects. 127 Table 7—14 Stress indices of the optimized CMB due to different extreme effects ....... 127 xii Table 7—15 Load cases specification for CMS bridges .................................................. 130 Table 7—16 Coordinates constraints for geometric design variables .............................. 132 Table 7—17 Initial designs of CMS Bridges .................................................................... 132 Table 7-18 Optimal results of the first-level optimization for the CMS bridges ........... 136 Table 7—19 Section forces for optimal CMS bridges ...................................................... 140 Table 7—20 Section moments for optimal CMS bridges ................................................. 140 Table 7—21 Buckling eigenvalues and eigenshapes of the optimal CMS bridges .......... 142 Table 7—22 Structural response of optimized CMS bridges for different limit states 144 Table 7—23 Stress indices of optimized CMS bridges for different limit states ............. 144 Table 7—24 Structural responses of the optimized CMS bridges subject to spaced line loads under the service limits state ......................................................................... 146 Table 7—25 Stress indices of the optimized CMS bridges subject to spaced line loads under the service limits state ................................................................................... 146 Table 7—26 Section forces of optimal CMS bridges subject to loading pattern changes 147 Table 7—27 Section moments of optimal CMS bridges subject to loading pattern changes ................................................................................................................................. 147 Table 7—28 Stress indices distribution of three bridges .................................................. 152 Table 7—29 Statistics of relative stress index distribution .............................................. 153 Table 7—30 Optimal CMB bridge designs with different membrane thickness constraints ................................................................................................................................. 155 Table 7—31 Structural response with different membrane thickness .............................. 156 Table 7—32 Evaluations of alternative objective functions ............................................. 157 xiii 1 Introduction 1.1 General Fiber reinforced polymer (FRP) composites have been adopted from aerospace and defense industries to civil infrastructure due to their unique properties, such as high strength-to-weight and stiffness-to-weight ratios, and high chemical and environmental endurance. However, to date, the use of FRP composites for civil infrastructure is mainly for structural rehabilitation and strengthening. The use of FRP composites to fabricate primary structural components has been limited, primarily due to high material costs and designs based on shapes suitable to conventional materials. The layered and fiber dominated structure of FRP composites makes them most efficient when used under in— plane stress demands. The shape resistance concept that structures carry loads by shaping structural geometries to achieve in-plane or membrane resultants is introduced in this chapter, followed by a brief discussion of shape finding for shape resistant structures. Laminate Optimum designs are then discussed to improve in—plane behavior of shape resistant structures by tailoring the material properties of FRP composites property. The rationale to find material-adapted structural shapes and tailor shape-adapted material properties for efficient use of FRP composites in new construction is thus recognized. The aim of this research is then presented so as to develop an approach that accomplishes these integrated design tasks. 1.2 FRP composites for civil infrastructure This section provides an overview of the mechanical characteristics of laminated FRP composites and a brief review of current applications in civil infrastructure. 1.2.1 Introduction to FRP composites Fiber reinforced polymer (FRP) composites are materials in which a reinforcement constituent of fibers is surrounded by a continuous and uniform matrix constituent of polymer. The fibers are generally strong and stiff to provide a primary load-carrying capacity. The polymer matrix holds the fibers together, protects the fibers from damage, and allows loads to be distributed among individual fibers. FRP composites are typically used in laminates, which consist of thin layers of fiber- reinforced material fully bonded together. Each individual layer, or lamina, contains an arrangement of fibers embedded within a thin layer of polymer matrix material. Depending on the arrangement of the fiber constituent, an individual layer of a laminated composite may be constructed by a number of forms. A typical form is to align unidirectional continuous fibers in a given direction. As it will be discussed in Section 1.2, this arrangement provides a unique feature of the material properties of FRP laminates, which can be altered by varying the percentage of layers in the laminate with different orientations. Therefore, FRP laminates composed of unidirectional continuous- fiber composite layers are commonly used in designs of high-performance structures and are thus primarily considered in this research. 1.2.2 Mechanical properties of FRP laminates The properties of FRP composite layers, such as stiffness and strength, strongly depend on the directional nature of the fibers. The fibers used in FRP composites usually have much higher stiffness than the polymer matrix. Consequently, unidirectional FRP composites have different stiffness properties along the fiber direction and perpendicular to the fiber direction. The stiffness of FRP composites in the fiber direction is governed by the fibers and the stiffness perpendicular to the fiber direction is mostly dominated by the polymer matrix. Such materials, in which material properties have two mutually perpendicular planes of symmetry, are referred to as orthotropic. The stiffness of Orthotropic FRP laminates can be fully described by four elastic stiffness properties, which are referred to as the engineering constants, i.e., two Young’s moduli, E1 and E2, along the fiber and transverse to the fiber directions respectively, and the shear modulus 612 and Poisson ratio v12 in the plane of the layer. Typical stiffness properties of various unidirectional FRP composites are given in Table 1.1. Table 1.1 Properties of various fiber-reinforced cmosite layers 1 E2 G12 (GPa) (GPa) (GPa) 300/5208 Graphite/Epoxy 181 10.3 7.17 0.28 AS4/3501 Graphite/Epoxy 138 8.96 7.10 0.30 B(4)/ 5 505 Boron/Epoxy 204 18.5 5 .59 0.23 Kevlar49/Ep Aramid/Epoxy 76 5.50 2.30 0.34 Scotchply 1003 Glass/Epoxy 38.6 8.27 4.14 0.26 Material Constituents v12 It needs to be pointed out that FRP layers are highly dependent on the fiber orientation as shown in Figure 1—1, which demonstrates how the elastic moduli are strongly influenced by deviation of the fiber orientation. In addition, FRP laminates, which are constructed of unidirectional-fiber layers stacked at different orientations, inherit the properties of the individual layers and thus lead to high in-plane but rather low through-thickness properties. However, material properties of an entire laminate in different directions along the plane of fibers can be adjusted to maximize the utility of the directional nature of the material properties by varying the number, orientation, and stacking sequence of the layers. 80 _ _ 80 4 I 70 -} .3 7o - x - 60 :r 9 ‘: 50 A " t cu I . 0.50 j» +5»: . ~; 50 g +Eyy r m +ny I 2 40 T y ~— 40 3 j " 'o . O 30 :— “‘ 30 E - 20 i— -_ 20 i 10 i __ 10 _—"'" “"’ ‘ e c e 4 c + _ 0....»4..:....i...ji....4.....i....L.'..i,'.jo 0 1o 20 30 4o 50 60 70 so 90 Fiber Angle (Degree) Figure 1—1 FRP elastic modulus as a function of fiber orientation 1.2.3 FRP composites in civil structures Modulus (GPa) Traditionally, FRP composites have been used extensively in aerospace and consumer sporting goods where their high stiffness and strength-to—weight characteristics were first exploited. These properties, together with their high chemical and environmental endurance and non-magnetic properties compared to conventional materials, have increased interests in the use of FRP composites for civil infrastructure applications [Bakis et al., 2002]. Particularly in the rehabilitation of existing structural systems, advanced composites have shown significant promise in recent laboratory and field applications. The seismic retrofitting of bridge columns (Figure l—2a) with carbon fiber wraps or pre-formed jackets has been demonstrated to be technically just as effective, and in some circumstances more economical, as conventional steel jacketing. The benefits of light weight and high strength also make FRP composites attractive for the flexural and shear strengthening of existing concrete structures (Figure 1—2b) [Teng et al., 2002; Hollaway and Leeming, 1999]. c)Rehabrlrtaion Figure 1—2 Applications of FRP composites in civil infrastructures Structural rehabilitation with FRP composites includes not only repairing and strengthening of aging infrastructure but also replacement of substandard structural components such as new bridge decks [Black, 2000] (Figure 1—2c). The advantages of employing FRPs in new bridge decks include its corrosion resistance, their reduced weight, and ease of installation. Decks made from composite materials can be customized to dimensions of traditional decks and can avoid modifications or replacement to the existing substructure. Another application of using FRPs in structural components is reinforcing bars (rebars) fabricated from either glass-fiber or carbon-fiber reinforced plastic composites [Chaallal and Benmokrane, 1996]. Due to corrosion of steel reinforcement in concrete structures, deterioration of concrete results in costly maintenance, repairs and shortening of the structure’s service life. FRP rebars have thus shown to be a viable alternative to steel reinforcement and prestressing tendons in areas where the use of steel can lead to a limited life span due to the effects of corrosion. Finally, the use of all—FRP composite designs for major structural components in new construction has also become of great interest. Two design and construction systems for short- and medium-span bridges consisting of tubular FRP members in combination with conventional materials have been developed and applied to two vehicular bridges in California [Burguefio, 1999]. 1.3 Outstanding issues and motivation In spite of the many applications of FRP composites in civil infrastructure cited above, their use is predominantly in structural rehabilitation and strengthening [Bakis et al., 2002, Karbhari and Seible, 2000]. Construction of new structures composed in their majority by FRP composite materials is still limited to a few highly subsidized demonstration projects. Most of the projects are unique and wide acceptance by the civil engineering community and commercialization has not yet occurred. The limited use is partially due to a high cost of FRP composites compared to the conventional materials such as concrete and steel. Another reason is using FRP composites for primary structural members in shapes copied from those used for conventional materials (i.e., linear shapes such as I-section) does not take advantage of the predominant in—plane stiffness and strength of laminated FRP composites [Burguefio, 1999; Keller, 2002]. In order to make composite structures competitive with metallic counterparts, the overall cost of producing FRP composites needs to be reduced. Advances in manufacturing techniques, such as pultrusion, resin transfer molding, filament winding and the automated or semi-automated manufacturing of large components, has significantly reduced costs. A more efficient way to use FRP composites is in conjunction with conventional structural materials rather than for individual component replacement or complete FRP composite designs [Burguer'io, 1999]. This requires new structural concepts and systems that combine the dominant characteristics of concrete in compression and steel in inelastic deformation capacity with the superior mechanical characteristics of directional strength and stiffness in the direction of the composite fibers. Therefore, efficient use of laminated FRP composites requires developing structural concepts and systems that employ them under in-plane stress demands [Burguefio and Wu, 2002]. 1.4 Shape and material optimized FRP structures The development of FRP structures working under in-plane structural demands requires combining shape finding methods and material-property tailoring. The concepts of shape resistant structures and material tailoring are thus introduced next, followed by a brief discussion of the integrated optimization approach for laminated FRP structures presented in this dissertation. 1.4.1 Shape resistant structures and optimal shape finding A genre of structural systems that can be used to improve the efficiency of laminated FRP composites is called shape-resistant structures (Figure 1—1), which carry design loads in large spans mainly through in-plane or membrane resultants acquired by shaping the material according to the applied loads. Membrane structures and thin-shells are two types of shape resistant structures in which the material is efficiently used under in-plane stress demands developed by the structural shapes. Arch Bridge Shell Structure Figure 1-3 Shape resistant structures [Otto, 1969] Traditional design methods using trial-and-error are not applicable for the design shape-resistant structures. The methods typically used for the design of shape-resistant structures are called form-finding or shape optimization. These methods, which range from experimental to diverse numerical approaches, have been successfully developed for determining the optimal shapes and form finding of shell/membrane structures. However, they can not be readily applied to the optimal shape design of laminated FRP composite structures since they apply primarily to structures made from isotropic materials, maintaining the same material properties during the form finding process. Designs for FRP structures obtained through these methods are thus often far from optimal because other competitive material properties cannot be explored. 1.4.2 FRP laminates and optimal laminate design As previously mentioned in Section 1.2.2, the material properties of FRP laminates can be altered by varying the number, fiber orientations, and the stacking sequence of layers in a laminate. This advantage of tailoring the material properties provides the possibility of improving the structural performance of a shape resistant structure by adjusting the properties of FRP laminates to strengthen the in—plane behavior of the shape resistant design. Evolutionary design optimization methods have been successfully developed and applied for optimal laminate designs requiring discrete changes of fiber orientations and the stacking sequence. However, these methods cannot be directly incorporated with shape optimization procedures, which depend on continuous changes of the structural geometry. Material optimum designs are thus isolated from the shape- finding process of structural resistant designs. Therefore, development of efficient designs for FRP structures in infrastructure requires a general approach accomplishing simultaneous shape and material optimization. 1.4.3 Shape and laminate optimization of FRP structures From the previous discussion it is understood that finding an efficient structural design Of a laminated FRP structure that meets all the requirements for a specific application should be achieved not only by shaping the geometric configuration of the structure, but also by tailoring the material properties, It is thus considerably more complex to find efficient designs for laminated FRP structures than for those made of isotropic materials. Several researchers have investigated the optimum structural design of FRP structural components. For example, Aref [2001] investigated an approach using genetic algorithms to minimize the weight of structural components by simultaneously changing the cross- sectional shape and ply orientations of the FRP laminates. In another effort, Qiao et al. [1998] developed a global approximation method to optimize material architecture by volume fractions Of cross-plies and the cross-sectional area of laminated FRP beams. In spite of these and other developments, the methods developed thus far for the optimization Of FRP laminate structures are only applied to what can be termed as sizing optimization of standard linear shapes with limited laminate optimization of thickness and/or orientation of limited number of plies. Although research on the optimum design of FRP laminates and components has been under continued investigation, it is much less developed than shape optimization algorithms for conventional structures. The optimum design of an FRP laminate structure involves both the shape optimization Of the structure and the material Optimizations of the laminate. Thus, a general approach is required to simultaneously solve for structural shape and material properties thus improve the performance of laminated FRP structures. 10 1.5 Objective and scope of the dissertation The main Objective Of the research reported in this dissertation was to develop an implement a general analytical approach for the development and optimal design c material-adopted shapes for laminated FRP composites in civil structures. The propose analytical approach was evaluated and validated by Optimizing FRP shell structures an applied to develop innovative bridge systems using FRP membrane/shell elements. The dissertation outlines the developed analytical procedure for simultaneous] finding the Optimal shape and optimal laminate design of FRP structures. The integrate approach serves as an analytical tool to aid in the design of laminated FRP composit bridge systems and provides insight to the concept of efficiently using FRP laminates b taking advantage Of their directional strength and material tailorability. Chapter 1 provides with an introduction to FRP composites, particularly laminate FRP composites and their mechanical prOperties. Current applications of FRP composite in civil structures are briefly reviewed. The research motivation is discussed, followed b the research scope. Chapter 2 presents the fundamental relations that govern the linear elastic propertie of FRP laminates. The chapter then discusses the computation Of elastic properties 21 functions of variables that can be changed during the design process, followed by th formulation Of constraints between these variables. Chapter 3 reviews the existing techniques and their limits for form finding c membrane structures and shape Optimization of shell structures. The concept of structurz shape Optimization is then introduced with a brief discussion of requirements for materiz 11 optimization of FRP laminates. Finally, an uncoupled integrated approach is proposed for shape and material optimization. Chapter 4 and Chapter 5, respectively, describe the optimization problem formulation, the chosen algorithm and its corresponding implementation for each Optimization level in the integrated approach. Chapter 6 focuses on the investigation of the performance and stability of the integrated approach by investigating the optimization Of FRP shells. The proposed approach is evaluated by studying and comparing the structural behavior between a shape-and-material Optimized shell and two shape-only Optimized shells. Chapter 7 proposes two types of FRP composite membrane-based bridge systems that effectively employ laminated FRP composites. The developed integrated approach is implemented and evaluated through analytical studies on these two bridges. Chapter 8 summarizes and concludes the current research efforts for the integrated shape and laminate Optimization and provides recommendations for future research. 12 References Aref, A.J. (2001). A genetic algorithm-based approach for design optimization of fibe reinforced polymer structural components. Mechanics and Materials Summer Conferencl 2001 sponsored by ASME, ASCE, SES, San Diego, CA. Bakis, C.E., Bank, L.C., Brown, V.L., Cosenza, E., Davalos, Lesko, IF, A., Machid J .J ., Rizkalla, SH. and Triantafillou, TC. (2002). Fiber-Reinforced Polymer Composite for Construction — Sate-of—the—Art Review, Journal of Composites for Construction: 6(2: 73-87. Black, S., (2000). A survey Of composite bridges, Composites Technology: 6(2), 14-18. Burguefio, R. (1999). System characteristics and design of modular fiber reinforce: polymer (FRP) short- and medium-span bridges, Doctoral Dissertation, Univ. 0 California, San Diego, California. Burguefio, R. and Wu, J. (2002, September). Development of an FRP Membrane Bridg System. IABSE Symposium: Towards a Better Built Environment — Innovatior Sustainability, Information Technology, Melbourne, Australia. Chaallal, O. and Benmokrane, B. (1996). Fiber-Reinforced Plastic Rebars for Concret Applications. Composites: Part B 27B, 245-252. Hollaway, LC. and Leeming, M.B., (1999). Strengthening of reinforced concret. structures: using externally-bonded FRP composites in structural and civil engineering Boca Raton, FL: CRC Press. Otto, F, (1969). Tensile structures: design, structure, and calculation of buildings 0 cables, nets, and membranes. Cambridge, Mass: MIT. Press. Karbhari, V.M. and Seible, F. (2000). Fiber Reinforced Composites — Advance Materials for the Renewal Of Civil Infrastructure. Applied Composite Materials: 7, 95 124. Keller, T. (2002). Overview of Fiber-Reinforced Polymers in Bridge Constructior Structural Engineering International: Journal Of the International Association for Bridg and Structural Engineering: 2, 66-70. Qiao, P., Davalos, J .F. and Barbero, E.J. (1998). Design optimization of fiber-reinforce plastic composite shapes, Journal of Composite Materials: 32(2) 177-196. Teng, J .G., Chen, J F Smith, ST. and Lam, L., (2002). FRP strengthened RC structure: New York: Wiley. 13 2 Laminated Fiber Reinforced Polymer (FRP) Composites Laminated FRP composites are a class Of composite materials where the layers of unidirectional fiber—reinforced composites are staked at different orientations. By varying volume fractions and the fiber orientations Of layers in a laminate, material properties of the entire laminate can be adjusted to meet the requirements of a design. FRP laminates have a wide range of application in structural design, especially for high—performance structures that have stringent requirements of specific (property divided by weight) stiffness and strength. In this research, FRP laminates are the only type Of FRP composite considered in the composite structures. Applicable structural designs require meeting certain response quantities such as displacements, stresses, buckling loads and natural frequencies. These response quantities depend on the constitutive behavior of materials, which are determined by the elastic properties such as Young’s modulus, shear modulus, and Poisson’s ratio. For laminated FRP composites, the constitutive behavior is determined not only by the these elastic prOperties but also by the parameters Of fiber orientations and stacking sequence that can be altered in order to improve structural performance. The Objective of this chapter is to discuss how the sectional stifi‘ness calculations of FRP laminates depend on the parameters that can be changed during a design process. The chapter begins by presenting the stress-strain relation that governs linear elastic responses of an orthotropic lamina. The section stiffness of FRP laminates composed of orthotropic laminas oriented at angles is then formulated based on the constitutive relations and the lamination stacking sequence. In addition, the formulation of the section stiffness is given in mathematical terms of lamina invariants and lamination parameters. 14 This alternative formulation of the section stiffness requires establishing the relationships between lamination parameters when altering fiber orientations in a stacking sequence. Thus, the feasible domain of out-plane lamination parameters is explored with respect to the in-plane lamination parameters at the end of this Chapter. 2.1 Constitutive relations for F RP laminates The stress-strain relation for a three—dimensional anisotropic linear material, also known as Hooke’s law, is expressed in the following tensor form: 023': ijmnéinn (2-1) Because of symmetry, 03:0},- and £;,,,,=a.m, there are only 21 independent material constants in the C matrix, in which Cijmn= 11”,": gm. SO, Eq. (2.1) can be rewritten in the matrix form: r N r N r i 0.11 C11 C12 C13 C14 C15 C16 811 0.22 C22 C23 C24 C25 C26 822 03 C33 C3 C15 C36 8 3 i 3>=< 4 ‘ <3 f. (2.2) 0' 23 C44 C45 C46 823 031 C55 C56 831 (0'12 , isym C66 J .812 i In the case of a three—dimensional orthotropic material, such as a unidirectional fiber- reinforced lamina, there are two perpendicular planes Of symmetry that define two principal axes of material properties. These principal axes correspond to the direction of the fibers and the directions perpendicular to the fibers, denoted by subscripts 1, 2 and 3, respectively. The stress-strain relation in the principal material directions of an orthotrOpic lamina is given by Eq. (2.3) 15 ro-lli C11 C12 C13 0 0 0 \ r811‘ 022 C22 C23 0 0 0 822 0' C 0 0 0 e 1 33 i=4 33 l1 ‘3 y. (2.3) 023 C44 0 0 £33 031 C55 0 £31 1012, tsym C66, i512, Due to the thinness of typical lamina, all layers of the FRP laminates are assumed to behave in a plane stress state in the 1-2 principal material plane so that 0'33 2 0, 023 = 0 and 0,3 = 0. The strain—stress relation is then simplified to Eq. (2.4) 011 Q1] Q12 0 £11 022 2 Q12 Q22 0 822 , (2-4) 012 0 O Q66 812 where the Qij are the reduced material stiffness coefficients which are given in terms of four independent engineering material constants in principal material directions as: E E Q11 _ 1 v Q22 _ 2 , 1_V12V21 1 V17V21 v E V E Q12 _ 12 2 _ 21 l , (2.5) 1_V12V21 1—V12V21 Q66 : G12 ' Since an orthotropic FRP layer can be generally oriented at a certain angle with respect to a structural coordinate system of x-y-z, the stress—strain relations (Eq. (2.4)) in the material coordinate system must be transformed to the structural coordinate system. The transformation of stresses and strains can be accomplished by 0.11 an 81] xx 0'22 = T 0'yy and 6‘22 = T 8», , (2.6) 0'12 ny 812 8” 16 where T is a transformation matrix. Assuming that the 1-2-3 axis originally coincides with the x—y-z axis and considering a rotation 49 about the 3 (or z) axis, the transformation matrix is given by: cos2 (9 sin2 (9 2cost93inr9 T = sin26’ c0326 -—2cosr9$int9 . (2.7) . . 7 . ’3 ——cos€s1n6l cost9srn6 cos“ 6 —sm“ 6’ With Eq. (2.6) and Eq. (2.7) substituted into Eq. (2.4), the strain-stress relation in the structural coordinate system is transformed to: 0.1x 8)“ Q1 1 Q12 Q16 8.0 -1 — —" '— 0), =T QT e), = Q, Q,, Q,, a), . (2.8) axy 813’ Q1 6 Q26 Q66 8.x): Q}!- are called the transformed reduced stiffness coefficients, which can be expressed in a simpler form by introducing the lamina invariants [Tsai and Hahn, 1980] as: Q11 =U,+U,cos26+U,cos46 (2.9a) 5, = U, — U, c0346 (2%) 5,, = Ul — U, cos 2.9 + U3 cos 4e (2.9c) 5, = éU, sin 26 + U, sin 46 (2.9d) 5,, = éU,sm26 — U,Sin46 (2%) 5,, = US — U, cos46’ (2.91) where the Us are the lamina invariants defined as [Tsai and Hahn, 1980]: 1 U1 = g(3Qll + 35,, + 25,, + 4Q,,) (2.10a) l7 1 U2 25(Q11 _Q22) (Z-IOb) 1 U3 = §(Qr1 + Q22 — 2Q12 _4Q66) (Z-IOC) U4 :%(Q11 +Q22 +6Q12 _4Q66) (210(1) 1 Us : §(Q11 + Q22 _ 2Q12 +4Q66) (2-106) When n orthotropic layers oriented at different angles are stacked together, a set of additional assumptions are needed in order to derive the equations that govern the constitutive behavior of the laminate. Each layer is assumed to be perfectly bonded to the adjacent layers so that the laminated layers deform in unison without experiencing any discontinuity in displacements. Furthermore, Classical Kirchhoff plate theory is assumed for the bending behavior of the laminate so that the strains in the out-of—plane direction Z are neglected and the w displacement in direction z is constant through the thickness Of the laminate. These two assumptions, along with the assumption of the layers being in a state of plane stress, define what is known as classical lamination theory (CLT), which will be applied herein to formulate the constitutive behavior of the laminates. Z ll hk/:Zk+l'Zk ‘1— \ 1/ \_ l/ he, © 9L A ———,j \ ® OK i/ \ ZkiIZL L+l ) /7’ —3(Zf — Z134 )} ’ (2-17b) k=l h h h " . , 1 2 7 7 l2 V21A.B.Dl : Zsm 29(k){_ (Z1 _ Zk—l )2 Tizi _ Z1.2—1) ‘Tizi _ Zi—1)}’ (2-170) k=1 h h h " , 1 2 2 2 12 , V31A,B.Dl : ZCOS 46m{— (Zk _ Zk—l )’ 7(Zk — Zk—l )7 —3(ZI: ’— Z2-1)}’ (217(1) M h h h " . 1 2 2 2 12 , V4lA.B,D} : Zsrn 46(k){_(zk _ Zk—l ) __2(Zk _ ZH ) _3(Zf — 224)}. (2.176) k:1 h ll 11 The A; , B; ,D; coefficients in terms of the lamina invariants Us and the lamination parameters V’s are summarized in Table 2—1. Table 2—1 A*, B*, and D... matrices in terms of lamina invariants and lamination parameters V0{A,B,D} VllA,B.D) V2{A,B,D} V3]A,B.D} V4{A,B,D) {AiirBierfii U1 U2 0 U1 0 {A;..B;..D;.} U1 -U2 0 u. 0 {Aferfzerzi U4 0 0 U1 0 {A;,,B;,,D;,} U5 0 0 U1 0 {£6,812.01} 0 0 U2 0 2U3 {ester} 0 0 U2 0 -211, The lamina invariants US are independent Of the ply orientation and are determined only by the property of a single layer with respect tO the material coordinate system. The lamination parameters V’s are related to the fiber orientations and stacking sequence of 21 the layers with respect to the structural coordinate system. Thus, as long as the properties Of the FRP lamina are known, the section stiffness of an FRP laminate can be determined by lamination parameters representing the fiber orientations and the stacking sequence. 2.2 Stacking sequence of laminates In the previous section, the section stiffness matrix describing the elastic responses of a laminate was presented. According to the section stiffness matrix, it should be noted that coupling effects may exist. These coupling effects are easily introduced by random stacking sequences. However, these coupling effects can hamper the effective use of FRP laminates, and their source and effect should be clearly understood. The [B] matrix represents the coupling effects between in-plane and out-Of-plane deformations. Then, according to Eq. (2.17), the [B] matrix vanishes when the stacking sequence Of a laminate is restricted to be symmetric with respect to the laminate mid- plane. According to Eq. (2.14), for symmetric laminates, the in-plane responses and out- of—plane responses can be solved independently. The in-plane resultants are only related to the mid—plane strains and the moment resultants are only related to the section curvatures. The stiffness terms A16 and A26 in the in-plane stiffness matrix creates a shear-and- extension coupling effect. This coupling effect will result in in-plane normal stress resultants by shear deformations. The shear-and-extension coupling can be eliminated by balancing layers of Off-axis fiber orientations. Balanced laminates require that for each layer with a negative orientation angle 6there is a layer with a positive orientation angle 6. Although the layers in a balanced pair do not need to be placed adjacent to each other, 22 the distance of two balanced layers will determine the value Of the bending—twisting coupling terms of D16 and D26. The bending-twisting coupling terms D16 and 026 exist for all laminates that have off- axis fiber orientations. The out-Of-plane bending—twisting coupling effect can not be eliminated even when balanced laminates are used. However, balanced laminates in which layers Of positive and negative Off—axis angles are placed adjacent to one another will reduce the effect of out-Of-plane bending-twisting coupling. Furthermore, for a certain laminate thickness, the bending—twisting terms, 016 and D26, vanish with an increasing number Of the layers. This behavior will be proved in the following section dealing with constrained relations between lamination parameters. Designs with consideration of coupling effects are generally avoided in the use Of FRP laminates, unless the coupling behavior is sought for pseudo-active structures. Thus, only balanced and symmetric laminates, which have the least-pronounced coupling characteristics, were investigated in the current research. Coupling of the in-plane and out-Of-plane responses, denoted by the [B] matrix, vanishes for symmetric laminates. According to the Eq. (2. 17), the in-plane shear-extension coupling defined by A16 and A26 vanishes when the laminates are balanced. The out-of—plane bending-twist coupling is considered zero on condition that the number of layers is large enough. Based on the above-mentioned assumptions of the staking sequences, the section stiffness Of laminates can be simplified to: 23 N, All A,, 0 0 0 o a: N, A,, A,, 0 0 0 0 5;? N“, i 0 O A,, 0 0 O 8:; , . = . i (2.18) Mx 0 O 0 D11 D12 0 KX M, 0 0 0 1),, 1),, 0 K, [M 1,, 1 0 0 0 0 0 1),, - ,K, 2.3 Constrained relations of lamination parameters Use Of the lamination parameters to define the section stiffness of laminates requires the definition of an allowable domain for each lamination parameter so that they meet the coupled constraints in relation with other lamination parameters. The constrained relationship between the lamination parameters V1,, V3,, V10, V20, V3,, and V4,, and their feasible domain are established in the following sections. 2.3.1 Relation between in-plane lamination parameters As discussed previously, in—plane shear-extension coupling effects vanish for balanced and symmetric laminates. Then, according to Eq. (2.18), the in-plane section stiffness is determined by: A11 A12 0 A: A,, A,, 0 (2.19) 0 0 A,, Recalling Table 2.1and Eq. (16), the components Of the stiffness matrix A can be determined in terms of lamination parameters and lamina invariants by: 24 i A1 1 Ah U1 V1 A V3A 1 A *, v v , 22 =h —-—> Geometric key points Computer-Aided Geometric Design L Update the design by mathematical algorithms Geometric key points Sensitivity Analyses Figure 3—5 Structural shape optimization process 47 3.1.3.1 Computer aided geometric design In a shape Optimization problem, the design variables are geometric parameters defining a structural shape. The generality Of a structural shape requires as many geometric design variables as possible. However, considering the efficiency of computational approaches, the number of design variables should be restricted as far as possible. CAGD techniques [Bohm et al., 1984] can assure the generality of a structural shape defined by fewer design variables. The coordinates of the key nodes governing the structural shape are thus typically chosen as the design variables. The shape of an entire structure can then be generated by appropriate shape functions in terms of the coordinates of the key nodes. A finite element mesh is then defined discretize the resulting shape for conducting the finite element analyses. 3.1.3.2 Finite element analysis Finite element analyses used in structural optimization provide the evaluation of structural responses required by the objective and constraint functions, and also by the design sensitivity analyses. The finite element method provides a numerical procedure for analyzing continua and structures that are too complicated to be solved satisfactorily by analytical methods. In most of practical optimization problems, structures are modeled by the finite element method to calculate their response to applied loads. 3.1.3.3 Design sensitivity analysis The design sensitivity analysis (DSA) is a fundamental requirement for structural shape optimization. Design sensitivity analysis supplies gradient information on objective and constraint funcions with respect to the design variables for formulating the 48 Optimization problem. Considering that finite element analyses are used, two major techniques, numerical and analytical, can be used to calculate structural sensitivity derivatives for finite element modeled structures [Adelman and Haftka, 1986]. Numerical techniques using finite difference methods are straightforward methods in which derivatives are calculated by a finite difference approximation. The accuracy of the derivatives obtained by this technique is depended on the selection of the perturbation step size for the finite difference. Analytical techniques, such as direct differentiation methods and adjoint variable methods, are reliable methods for most kinds of applications at the cost of a higher programming and computational effort. They can be more efficient than numerical techniques if only parts of the structure are affected by shape variants. However, especially for the sensitivity analyses in general shape Optimization, analytical techniques will not be able to calculate structural sensitivity derivatives when structural responses implicitly depend on the design variables. In general, the techniques employed for design sensitivity analyses rely on the algorithm to be chosen to solve the optimization problem. 3.1.3.4 Mathematical algorithm Given a constrained optimization problem, design updates and the convergence to an optimum design rely on a mathematical algorithm. Most of mathematical algorithms are based on the general iteration: x(k+1) = x(k) + akdu) (k+1) where x00 is the design variables vector Of the kth iteration, x is the design estimate of the (k+1)th iteration, on, is a step size, and da‘) is a search direction. 49 Depending on techniques used to update the designs, mathematical algorithms for solving constrained Optimization problems are broadly classified as transformation methods and direct methods [Belegundu and Arora, 1985]. Transformation methods were developed to solve a constrained optimization problem by transforming it into an unconstrained problem whose solution converges to a solution of the original problem. The basic concept of transformation methods is to construct a transformed function by adding a penalty for constraint violations to the objective function. Transformation methods include the penalty (exterior) and barrier (interior) function methods as well as multiplier (augmented Lagragian) methods. Due to the limitation inherited from the transformed functions, direct methods, also known as primal methods, were developed to directly solve the original constrained Optimization problem. Many methods such as the method of feasible directions, the gradient projection method and the constrained steepest descent method have been developed and successfully used for engineering design problems. Compared to other methods, the constrained steepest descent (CSD) method is robust and effective for solving nonlinear constrained optimization problems. Direct methods are generally based on the following four basic steps: 1. Linearize the objective and constraint functions about the current desing estimate; 2. Define a subproblem formulated by the linearized objective and constraint functions to determine a searching direction; 9° Solve the subproblem to give a search direction in the design space; 9“ Calculate a step size to minimize a descent fimction in the search direction. 50 The objective and constraint functions are linearized by the design sensitivity analyses previously discussed. The searching direction is determined by a subproblem based on the linearized Objective and constraint functions of the current design. In numerical practices, the subproblems using nonlinear programming have a better algorithm performance and optimum convergence rate than those using linear programming. The descent function plays a very important role in CSD methods for constrained problems. The basic concept is to compute a step size along the search direction such that the descent function is reduced. Therefore, the descent function is used to monitor the progress of the algorithm towards an optimum point. For constrained problems, the descent function is generally constructed by adding a penalty for constraint violations to the current value of the objective function. Several descent functions have been proposed and applied successfully in practical optimization designs [Han, 1971; Powell, 1978; Schittkowski, 1981]. Based on a descent function, a step size is determined by searching for a minimum Of the descent function along the desirable direction in the design space. The step size determination is also known as a one-dimensional search, or line search, problem. Several numerical techniques have been employed for the step size calculation, such as equal interval search, golden section search and polynomial interpolation. However, such exact line search techniques can be inefficient for constrained optimization methods. Therefore, an inexact line search is preferred to determine a step size for practical implementation of constrained problems. 51 All of these classical mathematical algorithms make use of derivatives of the objective function and constraint functions with respect to the design variables to construct an approximate model of the initial optimization problem. These algorithms rely on the gradients derived by linearizing the original functions to perform the optimization processes. Such gradient-based algorithms can not be generally applied on the optimization of FRP laminates, which typically involve discrete design variables in the Objective function and constraints. The rationale behind this limitation brings about the concept and technique presented here for material optimization of laminated FRP composites as discussed in the next section. 3.2 Material optimization of laminated FRP composites Laminated FRP composites consist of thin layers of fiber-reinforced material fully bonded together. Each individual layer, or lamina, contains an arrangement of unidirectional fibers embedded within a thin layer of polymer matrix material. The material properties of the entire laminate can be adjusted to meet the requirements of a design by varying the percentage of layers in the laminate, their individual orientations and the stacking sequence of the layers. Therefore, finding an efficient structure design can be achieved by tailoring the material prOperties. FRP laminates are thus the most commonly used type of composite material in the design Of hi gh-performance structures. From the above statements, and as presented in Chapter 3, it follows that the design space for the material optimization of laminated FRP composites is typically a set of discrete parameters that define the material elastic properties. Such optimization problems in which design variables are restricted to take only discrete/integer value are 52 referred to as integer programming (IP) problems. Classical gradient-based algorithms for structural optimizations requiring the gradient information of objective and constraint functions can not generally be employed to solve IP optimization problems. In recent years, there has been considerable interest in exploring new optimization methods that do not rely on derivatives of the Objective function and constraints to solve this type of problems. For example, the branch-and—bound algorithm [Lawler and wood, 1966; Tomlin, 1970] was advanced to solve linear integer programming problems. Due to the robustness and reliability to achieve global optimums in nonlinear integer programming problems, genetic algorithms [Holland, 1975] and simulated annealing methods [Kirkpatrick et al., 1983] have emerged as strong contenders against classical gradient- based algorithms. These methods belong to a genetic category of stochastic search techniques. Such evolution-based algorithms rely on the selection of fittest “individuals” in a “population” to obtain an optimum design by implementing the operations based on the principles of natural genetics. Unlike classical gradient-based algorithms that move from one point to another in the design space, such algorithms work with a population of designs. By keeping the solutions that have the potential of being the optimum in the population, such algorithms generally result in a global or near-global Optimum rather than converging to a local optimum. The evolution-based algorithms, such as genetic algorithms have been successfully applied to the material design of FRP composites [Giirdal et al., 2000; Liu el al., 2000]. However, most of the research on FRP optimum design for use in high-performance structures has focused only on material optimization, that is, altering material properties by arranging the stacking sequence and fiber orientations. Such optimization is isolated 53 from the structural shape, which also influences the mechanical behavior of a structure. This approach may thus lead to designs that make limited use of the dominant in—plane strength and stiffness of FRP laminates. On the other hand, the structural shape optimization procedures discussed in Section 1.4.1 were primarily focused on Optimum designs of structures made from isotropic materials. Optimum designs obtained by traditional shape optimization methods maintain the same material properties during the form finding process. These designs are often far from optimal because other competitive material properties cannot be explored. Therefore, efficient structural designs can be best obtained through shape efficient structures constructed with tailored laminated FRP composites. Finding an efficient structure design that meets all the requirements for a specific application can then be achieved not only by shaping the geometric configuration of the structure but also by tailoring the material properties. It follows that not only continuous parameters (defining the shape Of the structure) but also discrete parameters of lay-up and angles (defining the elastic properties of the material) should be taken into account in an optimization procedure. For the optimal design of FRP composite structures, the design space is thus a mixed set of discrete and continuous design variables. Gradient-based algorithms, which depend upon sensitivity analyses requiring continuity and derivative existence, are thus not applicable. Although genetic algorithms have been used with some success in problems where a mixed set of integer, discrete, and continuous variables are presented, they generally require more iteration steps to arrive at an optimum design. The random search in genetic algorithms requires more finite element analyses for evaluation of the objective and constraint functions, which decreases 54 computational efficiency and may result in a slower convergence compared to gradient- based algorithms. Therefore, an efficient optimization process needs to be explored for the combined shape and laminate optimization Of FRP composite structures. 3.3 Integrated shape and FRP laminate optimization From the above discussions, shape optimization problems and FRP laminate optimization problems are generally highly nonlinear. However, gradient-based and genetic-based algorithms, which are robust and reliable to solve shape optimization problems and laminate Optimization problems respectively, can not be directly applied to solve optimum designs of shaped-optimized laminated FRP structures. Therefore, an integrated approach that utilizes the computational advantages offered by both algorithms is proposed to accomplish the combined task of shape and material optimization. Due to difficulties to solve the entire problem through a single optimization procedure, the optimization task is thus to be implemented in a two-level uncoupled approach (Figure 3— 6): Level 1: Shape and laminate-property optimization: to achieve an optimal shape with optimal stiffness properties; Level 2: Laminate design optimization: to design an FRP laminate that has the same stiffness properties as those obtained in the first step. 55 <_§@ Level-1 3 Shape and Laminate-Property Optimization (Gradient based algorithm) Level-2 fl Laminate Design Optimization (Genetic algorithm) 3 End Figure 3—6 Two-level approach of shape and laminate Optimization The alternative formulation of the laminated composite material stiffness derived in terms of lamina invariants and lamination parameters (see Section 2.1) allows implementation of the proposed optimization scheme. First, the lamination parameters (Vim, 0)) that determine the sectional stiffness of laminates are combined with geometric parameters that define the shape of the structure. The optimum set of continuous design variables (sectional stiffness and geometry) is found so as to maximize the stiffness of the structure. As a second and final step, a set of stacking sequences of variable fiber orientations is then searched to achieve the desired lamination parameters, as obtained from the first step. The two-level optimization approach combines shape optimization of the structural geometry with material optimization for the FRP laminate design. In addition to the computational efficiency provided by gradient-based algorithms, diverse laminate designs can be achieved by implementing genetic-based algorithms in the proposed integrated approach. The proposed approach takes advantage of decoupling the two 56 different Optimization processes and corresponding optimization Objectives with respect to their own set of design variables. According to the two levels employed in the integrated approach, each optimization level involves an independent problem formulation by identifying and defining design variables, objective functions and constraints, and requires the implementation of the corresponding algorithm. The following two chapters describe in detail the procedure for each optimization level. 57 References Adelman, HM. and Haftka, RT. (1986). Sensitivity analysis of discrete structural systems. AIAA Journal: 24(5), 823-832. Arora, J .S. (1989). Introduction to Optimum Design. New York: McGraw-Hill. Arora, J.S. and Haug, E.J. (1979). Methods of design sensitivity analysis in structural optimization. AIAA Journal: 17, 970-974. Barnes, M. (1994). Form and stress engineering of tension structures. Structural Engjreering Review: 6, 175-202. Bletzinger, K.U. (1997) Form finding of tensile structures by the updated reference strategy. IASS International Colloquium -— Towards the new millennium, University of Nottingham Bohm, W., Farin, G. and Kahmann, J. (1984). A survey of curve and surface methods in CAGD. Computer Aided Geometric Design: 1(1), 1-60. Bonet, J. and Mahaney, J. (2001). Form finding of membrane structures by the updated reference method with minimum mesh distortion. International Journal of Solids and Structures: 38, 5469-5480. Booker, L. (1987). Improving search in genetic algorithms. In Genetic alggrithms and simulated annealing, Davis, L. (ed.). Los Altos, CA: Morgan Kaufmann Inc., 61-73. Brew, J .S. and Lewis, W.J. (2003). Computational form finding of tension membrane structures — Non-finite element approaches (part 1-3). International Journal for numerical methods in engineering: 56(3), 651-698. Cook, R.D., Malkus, D.S. and Plesha, ME. (2001). Concepts and applications of finite element analysis. NewYorK: Wiley. Giirdal, Z., Haftka, RT. and Hajela, P. (1999). Design and optimization of laminated composite materials. New York: Wiley. Haftka, RT. and Adelman, HM. (1989). Recent developments in structural sensitivity analysis. Structural Optimization: 1, 137-151. Han, SP. (1979). A globally convergent method for nonlinear programming. Journal of ggtimization theory and applications: 22, 297-309. Haug, E. and Powell, OH. (1972). Finite element analysis of nonlinear membrane structures, Proc 1971 IASS Pacific Symposium Part II on Tension and Space Structures, Tokyo and Kyoto: 165-175. 58 Hildebrandt, S. and Tromba, A. (1983). Mathematics and optimal form. New York: Scientific American Library. lsler, H. (1991). Generating shell shapes by physical experiments, IASS Bulletin: 34, 53- 63. Isler, H. (1994). Concrete shells derived from experimental shapes. Structural Engineering international: 3, 142-147. Lewis, W.J. and Lewis TS. (1996). Application of formian and dynamic relaxation to the form finding of minimal surfaces, Journal of the International Association for the Shell and Spatial Structures: 37(122) 165-186. Linkwitz, K. (1999). Formfinding by the “direct approach” and pertinent strategies for the conceptual design of prestressed and hanging structures. International Journal of Space Structures: 14 (2), 73-87. Liu, B., Haftka, R.T., Akgiin, MA. and Todoroki, A. (2000). Permutation genetic algorithm for stacking sequence design of composite laminates. Computer Methods in Applied Mechanics and Engineering: 186, 357-372. Powell, M.J.D. (1978). Algorithms for nonlinear functions that use Lagrange functions. Mathematical Programming: 14, 224-248 Ramm, E., (1992). Shape finding methods of shells. IASS Bulletin: 33, 89-99. Ramm, E., Bletzinger, K. U., and Kimrnich, S., (1991). Strategies in shape optimization of free form shells. Nonlinear comgrtational mechanics — state of the art, P. Wriggers, W. Wagner (eds): 163-192. Ramm, E., Bletzinger, K.U., and Reitinger, R., (1993). Shape optimization of shell structures. IASS Bulletin: 34, 103-121. Schek, H]. (1974). The force density method for form finding and computation of general networks. Computer Methods in Applied Mechanics and Engineering: 3, 115- 134. Suzuki, T. and Hangai, Y. (1991). Shape analysis of minimal surface by the finite element method, Int. IASS Symposium on Spatial Structures at the Turn of the Millenium, Copenhagen: 103-110. 59 4 Shape and Laminate-Property Optimization of FRP Structures This optimization level seeks the most efficient structural geometry and laminate sectional stiffness subject to constraints such as structural performance, geometrical dimensions, and laminate failure criteria. This chapter firstly presents the formulation of optimization problem by identifying the design variables and the objective function used for this level of optimization. The algorithm used. to solve the optimization problem is given first, followed by its implementation. 4.1 Formulation of the shape and laminate-property optimization problem Solving optimization problems involves transcribing a verbal description of an optimization problem into a well-defined mathematical statement. Therefore, this optimization level starts by identifying design variables, an objective function and the corresponding constraints, which are presented in the following three sections. 4.1.1 Design variables The design variables are independent parameters chosen to describe the design of a system in mathematical terms. Proper formulation of an optimization problem always starts by identifying the design variables for the system. In this optimization level, the design variables are the continuous geometric parameters defining the structural shape and the continuous lamination parameters defining the section stiffness properties of the FRP laminate. The generality of an optimal shape requires as many geometric design variables as possible. However, in order to maximize the efficiency of the computational algorithm, 60 only key dimensional points that define the geometric shape of a structure are chosen as the shape (geometric) design variables. The entire structure is then constructed by CAGD techniques by interpolating the key points. The laminate coupling effects, which are introduced by random stacking sequences of angle plies (see Section 2.2), are generally avoided for effective use of engineering materials. Therefore, symmetric and balanced laminates, which have the least- pronounced coupling characteristics, are considered in this research. The in-plane and out-of—plane coupling defined by the [B] matrix, and the in-plane shear-extension coupling, defined by the A16 and A26 terms, vanish due to the use of symmetric and balanced laminates. In order to reduce the effects Of out-of-plane bend-twist coupling, defined by the D16 and D26 terms, 48-layer symmetric and balanced laminates are used, in which the lamination parameters V20 and V40 are considered to be equal to zero. Thus, only the lamination parameters VIA, V3A, V10 and V30 (see Section 2.2) [Giirdal et al., 1999], which fully define the in-plane and out-of-plane section stiffness properties, are chosen as the material-property design variables. 4.1.2 Objective function The objective of the shape optimization procedure to determine the layout of a shape resistant structure is to maximize structural stiffness as measured by the structural stiffness matrix. However, the two-order tensor of the structural stiffness matrix can not be used to evaluate the objective function, which requires a scalar value. Therefore, the structural strain energy, a scalar that quantifies the structural stiffness, is calculated instead in the objective function. 61 Considering a finite-element-discretized structure without initial strains and stresses, the total strain energy, U, can be stated as 1 l U = Ejlef [Elieldv =§{D}T [K 1D} (4.1) V where [E] is the material stiffness, {D} is the global degree-of-freedom (d.o.f.) vector of the entire structure, and [K] is the structural stiffness matrix. The global d.o.f. vector {D} can be solved by the equilibrium equation as: {D}=[Kl“{R}. (4.2) where {R} is the consistent global nodal load vector. By substituting Eq. (4.2) into (4.1), the strain energy can be evaluated as: u = $11164 {at min-11111} =§1rt11<1111<111<111n (43> =11er [Kl“{R}. 2 According to Eq. (4.3), for a given statically equivalent load {R}, the total strain energy U decreases as the structural stiffness [K] increases. Furthermore, the strain energy contribution by different strain types can be determined by recalling Kirchhoff’s theory, which is assumed in classical lamination theory. Thus, the strain in a plate under a state of plane stress is defined by the addition of the mid-plane strain {60 } and the linear variation due to bending curvatures {K} as: {e} = {5" }+ z{K}. (4.4) Substituting Eq. (4.4) into Eq. (4.1) one obtains: 62 U: l (1601+ 2181 19111808 zinildv j({s~ i [QKSO }+ 2z{8" }T [gym any (grain. (4.5) It is thus shown in Eq. (4.5) that, within a unit volume, the strain energy contributed by the bending curvature and axial-bending coupling are quadratically and linearly proportional to the section thickness, respectively. Therefore, minimization of the structural strain energy implies minimizing bending and axial-bending effects, thus maximizing in-plane response and consequently increasing the system stiffness. 4.1.3 Constraints As previously stated, the objective of the shape and laminate-property optimization is to maximize structural stiffness. However, maximizing the stiffness without restricting material use is meaningless since stiffness can be improved if the amount of material is increased. Thus, the mass of FRP composite used is restricted. Instead of using the equality constraint that the volume of FRP laminates remains the constant during the Optimization process, an inequality constraint is chosen to limit the volume of FRP laminates within a certain level. This can be implemented by constraining the thickness of the laminate and the structural geometry. Among the several function constraints that need to be satisfied for a laminated FRP structure, service and failure constraints are the two most important design criteria for engineering applications. These design criteria can be investigated by the structural responses of maximum structural deflection and maximum section stresses as evaluated by finite element analyses. 63 4.2 Algorithm of shape and laminate-property optimization In the shape and laminate—property Optimization of laminated FRP structures, the objective function and constraints are nonlinear functions with respect to the design variables. The constrained steepest descent (CSD) method is a simple, yet effective, gradient-based algorithm used for solving this type of nonlinear problems [Arora, 1989]. The CSD method is based on four basic steps as presented in the following. 4.2.1 Linearization of Objective and constrain functions For shape and material-property optimization, the objective and constraint functions are implicit functions of the design variables. The solution strategy employed by the CSD method involves solving an approximate problem obtained by linearizing the original objective and constraint functions. The derivatives of objective and constraints functions used to construct the approximate problem are derived by Taylor series expansions. The approximate optimization problem using the first-order Taylor series expansion is formulated as: Objective: n f :ZCldl’ or f :CTd (4'63) i=1 Subject to: n Za,d,sb,, j=ltom; or ATdsb (4.6b) i=1 n 2,1,6,- =ej, j=1to p; or NTd=e (4.6c) 1'21 64 where e,- is the negative of the j"’ equality constraint function value at the current design x , e, = —h,(x(k’); bj is the negative of the j"' inequality constraint function value at the current design it“), b, =—gj(xa"); Cj is the derivative Of the objective function with respect to the i"’ design variable, c]. = af(xa‘))/ 8x1; n, is the derivative of the j”' equality constraint with respect to the i"' design variable, n, 2 8h]. (x‘k’ )/ axl; a, is the derivative of the j”' inequality constraint with respect to the i"' design variable, a, 2 8g 1. (x‘k) )/ 8x, . The derivatives obtained by a linear Taylor series expansion can be numerically evaluated by finite difference approaches. In this optimization level, the forward difference technique is used to evaluate the numerical differentiation of a multi-variable function fix) with respect to variables x,- , which is defined as: 8f _ = limit f(x1,...,x, +Axi""’xit)—f(x1"“’xi""’xn) ax; ax,—>O Axl- (4.7) where Ax,- is a small perturbation in the variables xi. The accuracy of function gradients depend on the selection of the perturbation Ax. Value selection of the perturbation is presented in detail by Gill, Murray and Wright [1981]. In general, a perturbation of 1% of the current value, which works fairly well for most of optimum designs, was chosen for this optimization problem. 4.2.2 Definition and determination of a searching direction In order to solve the constrained optimization problem, the CSD method requires a desirable searching direction towards the optimum design. A searching direction is determined by solving an subproblem. Due to its high efficiency and quadratic 65 convergence rate, a subproblem using a quadratic programming (QP) formulation is chosen to determine the searching direction. The QP subproblem employs a Hessian matrix constructed by a quadratic objective function and linear constrains, which are defined in the following: Minimize: f=ch+%-dTHd Subject to: NTd = e ATd s b (4.8) where H is the Hessian matrix, which represents the second-order derivatives of the Objective function with respect to the design variables. For large-scale optimization problems it is difficult and inefficient to calculate the second-order derivatives matrix. However, due to the usefulness of incorporating the Hessian matrix into the optimization algorithm, the Quasi-Newton method was developed to approximate the Hessian matrix by making use of information obtained from previous iterations. In the Quasi-Newton method, an approximate Hessian matrix is updated by using design changes and the gradient vector of the previous iteration. Several updating procedures have been developed [Gill et al., 1981]. The modified BFGS (Broyden- Fletcher-Goldfarb-Shanno) method [Powell, 1978] is implemented to numerically approximate the Hessian matrix in this optimization level. 66 4.2.3 Definition of the constrained steepest descent function For shape and laminate-property optimization the objective is to minimize the strain energy of a structure. Achievement of this optimum requires a reduction of the objective function value in each iteration. A descent function is used to ensure the reduction in optimization processes towards the minimum. For constrained Optimization problems, the descent function needs to take into account violations of the constraints by adding a penalty for constraint violations to the current value of the objective function. Several descent functions [Han, 1977; Powell, 1978] have been proposed for constrained optimization problems. Pshenichny’s descent function [Pshenichny and Danilin, 1978, Arora, 1989, 1997] is chosen as the descent function of the CSD method in the shape and laminate-property optimization due to its wide use in engineering optimization problems. Pshenichny’s descent function (I) at point xfl‘) is defined as 1 :2: 20% 1 Optimal Lamination Parameters 0. VM = -0.127, V3,, 2 -0.084 2 v”, = -0005, v,,, = -0799 0 15% n m _ E :1 Optimal Laminate Design 10% [:46°/i43°/i45°/1~45°/:48°/:39°/ 51 i66°/i2°/:t88°/0°2/ i89°li90°2]S , °o :L 5/ : Error=0.157% o%‘....i.e..:!—--F-%I............... 0 1 000 2000 3000 4000 5000 6000 Iteration no. Figure 6—5 Second-level optimization process — FRP Shell 1 Table 6—1 Optimal laminate designs of FRP shells Laminate designs Error [: 46/: 43/: 46/: 45/: 39/: 47/: 66/ : 2/ : 88/02/902 /: 87], 0.610% [: 46 /: 43/: 46/: 45 /: 46/: 39/: 66/ : 2/ : 88/02/ : 89/: 87] S 0.445% [: 46/: 43/: 46/: 45/: 46/: 39/: 66/ : 2/ : 88/02/902/112 87] 5 0.427% [: 46/: 43/: 46/: 45/: 46/: 39/: 66/ : 2/ : 88/02/ : 89/9021s 0.300% [: 46/: 43/: 46/: 45/: 46/: 39/: 66/ : 2/ : 88/02/902/902JS 0.287% [: 46/: 43/: 45 /: 45/: 48/: 39/: 66/ : 2/ : 88/02/ : 89/902], Q 157% 6.2 Shape-only optimized laminated FRP shells The optimum result above (Shell 1) is compared to the shape optimization of two laminated FRP shells (Shell 2 and Shell 3) with constant material properties. These two shells have the same geometry and loading of the shell analyzed in the previous example. 86 The geometric design variables used here were the same as before, see Figure 6—1. Guided by the experiments of hanging models [Ramm et al., 2000], two laminate designs were chosen. Shell 2 used a [450/-450/-450/450J65 laminate with the lamination parameters {V]A, V3A, V10, V30}={0, -1, 0, —1} while Shell 3 used a [00/900/90°/0°] laminate with the lamination parameters {VI/b V321, V11), V3D}={0, l, 0.75, 1}. Therefore, the laminate-property design variables of the lamination parameters determining the stiffness matrix of each shell are constant throughout the optimization procedure. The Optimization process histories for Shell 2 and Shell 3 are presented in Figure 6—6 and Figure 6—7, and Figure 6—8 and Figure 6—9, respectively. It should be noted that the obtained optimal shapes for the two shells using different laminate designs are different (see Table 6—2), although the initial shapes were the same. 105; 1043* A ei Z , V >. 9 0103“ c 0 .E m h H U) 1023- 1014:111. 0 5 10 15 20 25 30 35 40 Iteration no. Figure 6—6 History of objective function — FRP Shell 2: [45°/-45°/-45°/45°]63 87 Dimension (m) .o s: 0') 00 P & IJAII 5 1o 15 20 25 30 35 Iteration no. 40 Figure 6—7 Histories of geometric design variables — FRP Shell 2: [45°/-45°/-45°/45°los 105, 1043 103? Strain energy (N-m) _| O M i 1 I 101 1 I 1 1 1 1 1 1 1 fifTrlyffiililllfl’IIII I'lll|l'll[ltll[ 111111111 [1 5 10 15 20 25 30 35 40 45 Iteration no. Figure 6—8 History of objective function — FRP Shell 3: [0°/90°/90°/0°] 88 50 (m) P Q ll_11111 Dimgnsion a» C h 1 T 3 N 3 u 3 b 3 or p N 1111 l \ \ / 0- 4m 0 5 10 15 20 25 30 35 4o 45 50 Iteration no. Figure 6—9 Histories of geometric design variables — FRP Shell 3: [00/900/900/00] 6.3 Comparison of three optimal shells The structural response of all three Optimal shells regarding strain energies, displacements and section resultants is summarized in Table 6—2 and Table 6—3, respectively. Table 6—2 Summary Of strain energies and displacements of optimal shells Strain Energy Max Deflection (N-m) (mm) Shell 2 M 72.76 1.23 Shell 3 r...‘ 92.98 1.59 89 Optimal Shape According to Table 6—2, the optimal shell obtained through the shape and laminate Optimization (Shell 1) has the maximum structural stiffness, that is, the smallest vertical displacement and the smallest strain energy among all three shells. Shell 2, constructed by the laminate of [45°/-45°/-45°/45°]6s, was found to be better than Shell 3, constructed by the laminate of [0°/90°/90°/0°] with respect to the structural stiffness. The inherent nature of shape Optimization for membrane structures and thin shells is to minimize bending stresses thus ensuring in-plane only response. Achievement of this Objective was evaluated by comparing the section resultants between initial and optimal structures for all three shells. As shown in Table 6—3, the bending moments are reduced significantly after the optimization process. Meanwhile, membrane compressions are developed in all three Optimal shells compared to the zero in-plane response of the initial structures. Contours of the force (SF1) and moment (SMl) section resultants (see Figure 6—10 for element axis orientation) are also compared for the initial and Optimal structures in Figure 6—11. Figure 6—11 makes evident that bending behavior is minimized while the in-plane behavior becomes dominant. Additionally the results in Figure 6—11 show that the section resultants seek to be uniformly distributed throughout the entire structure. 90 Local coordinate systems Global coordinate system Figure 6—10 Global and local coordinate systems Table 6—3 Summary of structural responses of initial and optimal shells Shell 1 Shell 2 Shell 3 Initial Optimum Initial Optimum Initial Optimum SFl Max 20 235.195 :0 192.815 :0 412.599 (kN/m) Min :0 -744.289 =0 -755147 =0 -820.294 SF2 Max 2:0 241.325 :0 192.990 zO 415.226 (kN/m) Min 2:0 -945.860 :0 -953.566 20 -1012.41 SMl Max 0287 0.329 108.288 0.395 123.407 0.355 (kN-m/m) Min -24.957 -0.368 -24.932 -0.283 -64.117 -0.171 SM2 Max -0.158 0.298 108.288 0.387 114.982 0.195 (kNvm/m) Min -54.077 -0.478 -24.932 0358 -41.885 -0.067 SF 1 , SF2: Direct membrane force per unit width in local l-, 2-axis of elements SMl, SM2: Bending moment force per unit width about local 2-, 1-axis of elements 91 Membrane Force (SFl) Bending Moment (SMl) Initial Optimal Initial Optimal 23x10'12kN 235.0 kN 220.0 kN-m/m 20.09 kN-m/m s—3><10'12 kN $0.09 kN-m/m $20.0 kN-m/m . Shell 1 Shell 2 Shell 3 i i i i i i i i l i i l i SMl: Bending moment force per unit width about local 2-axis of elements Figure 6—11 Structural responses of initial and optimal shells Images in this thesis/dissertation are presented in color. 92 It should be noted that stress leveling within the shell is not an optimization criterion to achieve maximum structural stiffness. For example, the shape-and-material optimal shell (Shell 1) attained the minimum strain energy and had the lowest membrane force and highest bending moment specific values (i.e., value at a given location) than the other two shape-only optimal shells. However, this does not conflict with the objective to minimize the strain energy of the entire structure by reducing bending moments. The in- plane structural performance can be best evaluated by comparing the stress-state in individual elements, which are summarized in Table 6—3. Thus, as shown in Figure 6—12, 90% of Shell 1 is subject to membrane forces ranging between —45 and 0 kN/m, compared to 79% and 84% for Shell 2 and Shell 3, respectively. The efficiency of Shell 1 is more obviously shown in Figure 6—13 regarding the bending moment SMl. Thus, although Shell 3 has the smallest range of bending moments among all three optimal shells, only 32% of Shell 3 is subject to bending moments of small magnitude, i.e., between —30 and 30 kN-m/m, compared to 74% and 47% for Shell 1 and Shell2, respectively. According to the analyses described above, the stiffness of bending structures is in fact improved by achieving the objective of minimizing the structure’s strain energy. In addition, the structural performance of laminated FRP shells can be further enhanced by the integration of shape and material optimization. Therefore, the shape and material objective function of minimizing strain energy leads to structural geometries and material properties that are dominated by in-plane response with minimum bending demands. 93 2500 22 60% . [SSW : JESheII . flamefl 2000 5: .A 500 -- \ Ele_r_nent Count 0 O O .21 30 45 60 Max Min -60 -45 -30 -15 0 15 Force Range (kN/m) Figure 6—12 Membrane force distribution of optimal shells 45% IShell1 ll EShe ‘ 40% @Shell i 35% 5 30% 4 25% 5 20% , 15% 10% f 5% . ~ ,, 10% Min -150 -120 -90 —60 -30 0 30 60 90 120 150 Max Moment Range (N-m/m) Figure 6—13 Bending moment distribution Of Optimal shells 94 Structure Fraction Structure Fraction 6.4 Buckling analyses of optimized laminated FRP shells As it is well known, shape-resistant structures carry their design loads primarily by axial or membrane action, rather than by bending action. When subjected to compression forces as in the case of shells, these stiff thin structures are sensitive to buckling. Considering that a structural stability criterion was not implemented in the presented integrated Optimization approach, the structural buckling performance was evaluated and compared among the Optimized shells. Eigenvalue buckling analyses are generally used to provide useful estimates of the critical buckling loads and collapse mode shapes of stiff structures. To evaluate the buckling response of the optimized shells, the unloaded optimal shells are chosen as the base state and buckling loads are calculated relative to the base state of the structure. The optimization load-generating load case of 6.9 kN/m2 applied in the gravity direction is used gain to introduce the linear perturbation. The subspace iteration extraction method is used to explore the eigenvalues and eigenvectors of the first five buckling modes. The results for Shell 1, Shell 2, and Shell 3 are summarized in Table 6—4, Table 6—5, and Table 6—6, respectively. As judged by the values for the first buckling eigenvalue given in Table 6—4 through Table 6—6, structural buckling capability is improved by the shape and FRP laminate optimization along with maximizing structural stiffness. The general-optimized shell (Shell 1) has the highest buckling capacity (4.: = 8.05 ), compared to Shell 2 (4,2 -- 4.29) and Shell 3 (if = 3.49 ). On the other hand, although the buckling mode shapes vary significantly in character, it can be seen that the first three eigenvalues are closely spaced. The series of closely spaced eigenvalues indicates that the optimized shells are 95 imperfection sensitive. In addition, the smaller spacing of eigenvalues for Shell 1 further implies Shell 1 is more buckling-sensitive compared to Shell 2 and Shell3. Therefore, it can be concluded that a stiffer structure will increase buckling resistance but suffer from higher buckling sensitiveness. Table 6—4 Buckling analyses of Shell 1 . Buckling shapes Mode Eigenvalue Side view Axis 1-3 Side view Axis 2-3 Base State _ fir ti 3 8.88 W m 4 13.38 m n 5 13.49 n m Table 6—5 Buckling analyses of Shell 2 Buckling shapes MOde Elgenvalue Side view Axis 13 Side view Axis 2—3 Base _ r 3 ‘ State L1 L2 1 4.29 2 4.37 3 4.82 4 6.70 5 9.46 96 Table 6—6 Buckling analyses of Shell 3 Buckling Shapes (eigenvectors) MOde Eigenvalue Side view Axis 13 Side view Axis 2-3 if}: 1L 1:2 1 3.49 h 2 3.53 n 3 3.64 r fl 4 4.63 n 5 11.55 m 6.5 Summary In this chapter, the Optimization procedure of the integrated approach was evaluated and validated by optimizing FRP shells. It was illustrated in the comparison of a general- Optirnized Shell and shape—only optimized FRP Shells that the structural performance of a Shape resistant structure can be improved by using FRP laminates with Optimized material design. Stability evaluation of the optimized shells showed that buckling capacity should be an essential constraint of the integrated approach for shape resistant structures subject to compression. 97 References Hibbitt, Karlsson & Sorensen, Inc. (2001). ABAQUS standard user’s manual Version _6_.2. lsler, H. (1991). Generating Shell shapes by physical experiments, IASS Bulletin: 34, 53- 63. lsler, H. (1994). Concrete shells derived from experimental Shapes. Structural Eggineeringintemational: 3, 142-147. Ramm, E., (1992). Shape finding methods of Shells. IASS Bulletin: 33, 89-99. Ramm, E., Kemmler, R. and Schwarz, S., (2000). Form finding and optimization of shell structures, In: Proc IASS-ICAM 2000. Fourth International Colloquium on Computation of Shell and Spatial Structures: Chania—Crete, Greece. Ramm, E. and Mehlhom G., (1991). On Shape finding methods and ultimate load analysis of reinforced concrete Shells. Engineering Structures: 13, 178-198. 98 7 FRP Composite Membrane-Based Bridge Systems Laminated FRP composites have been widely applied in strengthening and rehabilitating aging bridge systems and proved to be efficient. However, the application of laminated FRP composites for an entire structure is still limited. Although technical advances in manufacturing FRP composites have significantly reduced the manufacturing cost, their high material cost and the lack of proven design philosophies compared to conventional materials still confines the use of laminated FRP composites in primary structural components of bridge systems. However, the use of FRP composites in conjunction with conventional structural materials has Shown that technical efficiency can be achieved within competitive economical constraints [Burguefio, 1999; Seible et al., 1999]. Furthermore, it was proved that the use of the high in-plane strength and Stiffness of FRP composites can be maximized by material—adapted concepts of employing Shape resistant structures of membrane/shells. Therefore, FRP composites membrane-based bridge systems are proposed as analytical studies for the application of laminated FRP composites in bridge systems. The use of Optimized laminated FRP composites in bridge systems is conceptualized in two bridge designs that can effectively employ the in-plane stiffness of FRP laminates. Both bridge types consist of a laminated FRP membrane together with a conventional reinforced concrete deck, in which the laminated FRP membrane provides the in-plane strength and the concrete deck provides the live load transfer and resists the majority Of the compression forces. The developed optimization approach will be implemented and evaluated through analytical studies on these two types of FRP membrane-based bridges. 99 7.1 FRP composite membrane beam (CMB) bridges Based on the concept of girder—slab bridges, an FRP Composite Membrane Beam (CMB) bridge (Figure 7—1) is considered as a case study for the proposed optimization approach. Because of the high Stiffness provided by double-curvature surfaces, the system uses a hyperbolic paraboloid laminated membrane working compositely with a conventional reinforced concrete deck as schematically shown in Figure 7—1. The proposed system behavior is such that the FRP laminate is to carry the in-plane tensile and Shear forces while the concrete deck carries the compressive force. The developed Shape and FRP laminate Optimization approach was thus used to obtain the optimal Shape and laminate design of the membrane element for the CMB bridge system. Abutment Membrane Abutment Figure 7—1 Composite membrane beam (CMB) bridge 7.1.1 Bridge system description The proposed CMB Bridge for optimization studies is a 10.06 m long simple-span bridge girder with a 2.44 m wide compression flange (Figure 7—2). The flange is a 203.2 mm thick concrete deck with a density of 2402.8 kg/m3. In order to reduce difficulties of finite element modeling, the abutments were not modeled in the finite element analyses. Rather, the abutment stiffness was taken into account in the computational model by constraining the plane section motion that it imposes on the rotations and displacements 100 at both ends of the bridge (Figure 7—2). The laminates used in the CMB bridge systems are assumed to be from medium-performance carbon/epoxy FRP lamina, of which the lamina orthotropic properties are taken as: E / 1 = 88.253 GPa, E22 = 48.884 GPa, 012 = 4.557 GPa, 013 = 4.557 GPa, G23 = 0.456 GPa, and v12 = 0.513. Truck Load Figure 7—2 Computational model of the CMB bridge Table 7—1 Load case specification for CMB bridge ommization . . . . Design Criteria lelt State Load Case Specrficatlon Stress Deflection Optimization l.0>
<(LL + 1.33xTL) 0.250'u L/800 Strength 1.25xDL + l.75x(LL + 1.33xTL) 0.506u - According to the AASHTO Bridge Design Specification [AASHTO, 1998], the simple-span CMB bridge is to be designed to load combinations that include dead load (DL), live lane load (LL) and concentrated loads from a three-axle truck (TL). The design 101 lane load consists of a uniformly distributed pressure Of 3.06 kPa over each lane while the design truck load consists of three concentrated loads spaced at 4.27 m (Figure 7—2). The load case used in the integrated Optimization procedure takes into account the dead and uniform live load with unit, or service load factors (Table 7—1). In addition to the optimization load case, the service and strength load cases (Table 7—1), in which the loads are combined with different load factors also need to be considered [AASHTO, 1998]. The influence of these two load cases on the Optimal CMB bridge design was studied through post optimality analyses. Two design criteria, namely stress and deflection are used for the CMB bridge designs. The stress criterion chooses a percentage, or stress index, of ultimate material failure for each limit state. The prediction of the ultimate material failure varies with different failure stress criteria, which define different failure surfaces surrounding the origin in three-dimensional space {01], 0'22, 0'12}. The stress index is used to measure the proximity to the failure surface. The material failure determinations by different stress criteria are all depended on the stress limits of the individual lamina measured along the material directions. Following knowledge gained through experiments on carbon/epoxy FRP laminates, 1% and 0.5% of the Young’s modulus of the corresponding direction respectively are commonly used as the tensile and compressive stress limits, while 0.5% of the shear modulus was chosen as the Shear limit. Therefore, the tension and compression stress limits along the l-axis (fiber-longitudinal direction) and 2-axis (fiber- transverse direction) were chosen to be: X2 = 0.883 GPa, XC = -O.442 GPa and Y2 = 0.489 GPa, YC = -0.244 GPa, respectively, and the shear stress limit in the 1-2 plane was taken as S = 22.9 MPa. 102 In addition, the maximum structural deflection of U800 [AASHTO, 1998] was chosen as another design criterion for the service limit state. Although the AASHTO [1998] bridge design Specifications recommend the deflection limit Of U800 to apply for live load only, this limit is used conservatively in this work together with the system dead load. 7.1.2 Integrated Optimization of CMB bridges Based on the above-mentioned geometry, properties, and loading of the CMB bridge system, the integrated Optimization approach was applied to design CMB bridges by sequentially employing the shape and laminate-property optimization and the laminate design optimization procedures. 7.1.2.1 Shape and material-property optimization The shape design variables for the CMB bridge are the width and depth of the FRP membrane at the middle (W2, [12) and ends (w), h) of bridge component (Figure 7—3). These variables are used to define the hyperbolic paraboloid shape of the membrane through a CAGD algorithm. The laminate design variables are the lamination parameters {V,A, V3A, V10, V30} controlling the in-plane tension stiffness and out-plane bending stiffness. 103 Depth (h2), . * / Thickness (t) . ..» ” Width (w2) Wldth (w l) ,,/‘ // / Figure 7—3 Geometric key points for CMB bridge shape optimization It is well known that structural stiffness is improved when the amount of material increases. Therefore, the mass of materials needs to be constrained during shape optimization when the Objective is maximizing structural stiffness. An inequality constraint, instead of using an equality constraint, is used to limit the mass of materials used in the structure. Considering that the geometry of the concrete deck remains constant during the optimization process, the material constraint is thus implemented by setting a lower and upper bound for the geometric design variables that are controlling the geometric dimensions and the thickness of laminated FRP composite (Table 7—2). Maximum stresses evaluated at the top and bottom surface of the laminate and the maximum deflection of the entire structure are two additional constraints taken into account in the structural design criteria for the integrated optimization procedure. 104 Table 7—2 Dimensional constraints for geometric design variables W1 W2 ’1] [12 1 Bound (mm) (mm) (mm) (mm) (mm) Lower 609 609 609 305 6.35 Upper 2438 2438 1219 1219 19.05 The structural design achieved by an optimization process is normally obtained for a single load case. Due to the random nature of the truck loading position the optimization load case for the CMB bridge was taken as the combination of the dead load and live lane load with unit (service) factors (Table 7—1). This assumption was found to be conservative in the post optimality analyses (Section 7.2.4) as truck loading governs the demands in Short-Span bridges. In the finite element analyses, the strains on the FRP laminates and concrete are expected to remain in the linear elastic range. Furthermore, the approximation of formulating the structure stiffness matrix in the reference (original) configuration is expected to have an error of 10'3 order compared to unity because of the small rotations and displacements under the loads. Thus, material linear elasticity and geometric linearity were used in the structural finite element analyses. The stability of the proposed optimization approach was studied by searching for the Optimal CMB bridge design from three different starting or initial designs. The initial designs considered are listed in Table 7—3. The optimization histories of the first-level shape and laminate-property optimization for the CMB bridges are shown in Figure 7-4 through Figure 7—12. 105 Table 7—3 Initial designs of CMB Bridges Initial design Model A Model B Model C {WI’WZ’hlyw2,t} {1625.6, 1016, {1625.6, 1016, {1219.2, 1219.2, (mm) 609.6, 609.6, 19.05} 609.6, 609.6, 19.05} 812.8, 508, 6.35} {V1A,V3A,V,D, V30} {0,1,0.75,1} {O,-1,0,-1} {0,1,0.75,1} Laminate Design [0°/90°]s [:45°]g [0°/90°]3 O _l_. 10 15 20 Iteration no. Figure 7—4 History of Objective function — CMB Model A 106 25 2500 4 i i 1 2000 r E @1500 ~— I: .2 (D 51000 .5. D 500 ~- - [+w, +w2 +11, +11, +t(x20)J 0.1..li...14...lr.li . O 5 10 15 20 25 Iteration no. Figure 7—5 History of geometric design variables — CMB Model A 1.5 1__ parameters s: 01 O 1 1 1 1 l l Lamination 5': 0| J+V1A +V3A +V10 +VQ‘J .1,5 . = . . , . . . . , . . , . , . . . . . 4 0 5 10 15 20 25 Iteration no. Figure 7—6 History of laminate-property design variables — CMB Model A 107 ( Dimension mm) 800 _ 700 f— AGOO :— E _ ' 1 E500 :— > i m .1 L- 0) 400 -- c LIJ ,4 .E 300 —— / /. a . we” a: ‘ JV] .2}/ (D 200 3- K/ 12”” w/‘r‘ fl, : / 201h 100 17 3 ¢ ¢ c A. e e e e e g g I o . : . . . . . . . . . . . =1 0 5 10 15 20 Iteration no. Figure 7—7 History of objective function - CMB Model B 2500 2000 ~- 1500 ~— 500 r _ P—w, +w2 +|11 +112 +t(x20) ] 0 l I I I i l I I r i ' ' ' I I l I I T I 0 5 10 15 20 Iteration no. Figure 7—8 History of geometric design variables — CMB Model B 108 1.5 0.5 : Lamination parameters O 'o. NV ' ‘ J+V1A +V3A +V10 +V30J I I T I j 5 10 Iteration no. 20 Figure 7—9 History of laminate-property design variables — CMB Model B 01 05 O O O O 1p11111_LJI Strain energy (N-m) 8 8 O O 0 5 10 Iteration no. 109 Figure 7—10 History of objective function — CMB Model C 2500 2000-- A E £1500“ : .2 in 5 1000 *5 .E D 500 ‘5 q / {Va—w, +W2 +111 —o—h2 +t(x20)J o . . . . l . . . . : . . . . : 0 5 10 15 Iteration no. Figure 7-11 History of geometric design variables — CMB Model C 1.5 1- vkegfigkvkvgrefike . _ IIII - . . . . .2 0 _ 905. Y r E . V 10 h 8 c 0‘ .2 up 2 ---0.5~ E 10 .1 -1— 4 J—o—VM —I—v3A +V10 +V£J 0 5 10 15 Iteration no. Figure 7-12 History of laminate-property design variables — CMB Model C 110 AS summarized in Table 7—4, the shapes for all three CMB bridges (see Table 7—3) converged to a Similar geometry with negligible differences. The optimization processes starting from three different initial attempts also resulted in the same in-plane lamination parameters (see Table 7—4). However, the out-plane lamination parameters were different. Table 7—4 Optimum results of shape and material-property optimization Optimal design Model A Model B Model C Stra‘“ Energy 83.85 83.85 83.85 (N-m) {W1 W2 h) W2 1} { 1359.0, 2184.4, { 1365.2, 2184.4, { 1352.6, 2222.4, (mm) ’ 984.3, 1219.2, 1022.4, 1219.2, 997.0, 1219.2, 19.05} 19.05} 19.05} {V1A, V321, V11), V30} {081,031,421} {0.81,0.3l,-0.8,0.7} {O.81,0.31,0.75,1} In fact, it was noted that the lamination parameters governing out-plane bending stiffness of membranes remain unchanged beyond certain iteration steps. This behavior does not significantly disturb the solution since the strain energy minimization process leads to a system dominated by in-plane stress resultants, i.e., bending resultants become negligible. Therefore, the out-plane lamination parameters, which determine the flexural stiffness, have a smaller influence on the optimization process when the Structure becomes dominated by in-plane behavior. 7.1.2.2 Laminate design optimization Assuming that only the in—plane lamination parameters govern the sectional stiffness of the FRP laminate, the optimal lay-up of the laminate can be solved by an alternate method presented by G‘Lirdal et al. [1999]. This method allows finding the optimal lay-up for a 4-layer laminate with only two distinct orientation angles for a given volume 111 fraction. By this method, the lay-up corresponding to the optimal section stiffness, which is defined by the optimal lamination parameters of {VI/1, V3A }={0.81, 0.31} is found to be [l7°/19°]g. This result is later compared to the solution found by the developed laminate design optimization. Symmetric and balanced laminates with certain number of distinct fiber orientations, which are defined by the design variables, were determined by the laminate design optimization process. In order to be able to compare the results with the previous optimal laminate design, optimal 8-layer symmetric and balanced laminates that have two distinct orientation angles were searched by the laminate design optimization algorithm. Furthermore, as it was discussed in Section 2.3, the bend-twist (0,6 and 026 terms) coupling effect is minimized by increasing the number of plies in a symmetric and balanced laminate. Therefore, optimal 48-layer symmetric and balanced laminates (an arbitrarily selected large number of plies), which have twelve distinct fiber orientations, were also investigated. In the second-level optimization, an optimal laminate is achieved by minimizing the difference of the lamination parameters of the searched laminates from the Optimal lamination parameters derived from the first-level optimization. Therefore, the optimum criterion of the laminate design optimization is an allowable value of discrepancy. For the current research, a maximum error of 1% was considered acceptable for the optimum laminate design. 112 1 2% 10% 3— // on o\° 1 Optimal Lamination Parameters VM = 0.81, v3, = 0.31 Discrepency CD .\° Optimal Laminate Desigr 4% 4 [:17°/¢19°]S 2% Error = 0.22% _. O l 1 l l l o IIIIIIIII | IIIIIIIIIIIIIIIIIII r IIIIIIIII irrrrrrrrtrrrrrrrrrr Iteration no. Figure 7—13 Laminate design optimization for 8-layer laminates The optimal 8-layer laminate was found to be composed of [i17°/i'19°]g with an error of 0.22% with respect to the optimal in-plane section properties from the first-level optimization. The resulting laminate design is thus the same as that previously obtained neglecting out-of—plane behavior through Gurdal et al.’s [1999] method. The optimization history of the objective function (lamination parameter error minimization) is shown in Figure 7—13. The optimal 48-layer symmetric and balanced laminate was found to be [il7°/i21°/i'19°/i25°/i21°/i5°/ir22°/i12°/i25°/il4°/ i17°li12°]3 with an error of 0.73%. Its optimization history is shown in Figure 7—14. 113 16%~r 14% y >‘12% {— g -’ . . . 8,10% _’ OJgtzmal Lamination. Parameters e _; VIA=0.81, V3A=0.3l U 8°/o :’ g . Optimal Laminate Design 6% 3- [:17°/:2 1°/:19°/-1_-25°/:21°/:S°/ - :220/112°/~_~25°/¢14°/ i17°/:12°]S 4% {~ Error = 0.73% 2% 7 _ : _ 00/0 1 1 1 1 i 1 1 1 1 1L 1 1 1 1 i 1 1 1 V 11 T I i I i r 1 1 1 1L 1 r l r 0 500 1000 1500 2000 2500 3000 3500 Iteration no. Figure 7—14 Laminate design optimization for 48-layer laminates The lamination parameters and the section stiffness of the three optimal laminate are compared and summarized in Table 7—5. As shown in the table, all three optimal laminates have the same in-plane section properties and in-plane lamination parameters while the major terms of the out—of-plane section stiffness {D11, D12, D22, D66} and the out-plane lamination parameters {V113, V30} are only slightly different. However, the bend-twist coupling terms (D16 and D26) of the out-plane stiffness are different for all three laminates. Thus, as expected, balanced and symmetric laminates have less bend- twist coupling effects than symmetric-only laminates. It also can be noted that the bend- twist coupling terms can be further reduced by increasing the distinct groups of balanced fiber orientations. 114 Table 7—5 Optimal laminates from laminate design optimization process Lamination O timal laminate parameters A matrix D matrix p {VI/r, V34, V10. (103 kN) (kN-m) V20, V30, V40} {0,809,0308, ’362 6.01 0 1 [43.6 6.67 10.8) 4—layer [17°/l9°]s 0824,0566, 20.7 0 25.0 -4.41 0358,0933} L 6.98, 79!) {0,809,0308, "36.2 5.96 0 “ ”43.6 6.67 4.07) 8-layer [i17°/i19°]3 0824,0215, 20.7 0 25.0 —l.58 0358,0352} _ 6.98, l 7.91 l 48-layer i21°/i5°/i22°/i12°/ 0.800, -0.037, 20.8 24.4 0'23 i25°li14°lil7°li12°1s 0.231.-0.056} 893- [il7°/i21°/i19°/i-25°/ {0.803, 0.310, |736.1 5.96 0] ”42.1 7.80 —0.68“ 7.1.3 System characteristics of optimal CMB bridges The system characteristics of the optimized CMB bridge systems were investigated by studying their structural behavior for the optimal load case and their buckling sensitivity. Due to the slight geometric variations existing in the optimum designs, the CMB bridge chosen for the system evaluation studies has a shape with average dimensions of the obtained optimal bridges (see Table 7—4). The optimal 8-1ayer laminate design is used for FRP membrane in system characteristic studies. The variations to the Optimum design by using an average geometry and the 8-layer laminate with non-zero bend-twist coupling will be studied in post Optimality analyses. 7.1.3.1 Structural behavior of optimal CMB bridges As shown in Figure 7—4, Figure 7—7 and Figure 7—10, the optimal geometry Of the FRP membrane is that of a hyperbolic paraboloid. A hyperbolic paraboloid is obtained by translating a parabola along the longitudinal direction of another parabola. This shape 115 obviously follows as a direct consequence of the moment diagram of the loaded CMB bridge to assure in-plane behavior of the FRP membrane. The results illustrated in Table 7—6 and Table 7—7 clearly show that the FRP membrane of the Optimized CMB bridge is dominated by an in-plane stress-state by comparing the in-plane and out-of—plane section resultants. In addition, the FRP membrane also provides a shear force resistance, which reaches maximum values at both ends of the bridge structure. However, the bar chart in Figure 7—15 shows that 91% of the FRP membrane is subjected to longitudinal tension ranging from 0 kN/m to 133.9 kN/m, while only 9% of the membrane is subjected to longitudinal compressions with values les than 25 kN/m. Figure 7—16 reveals that the transverse section stress resultants of the FRP membrane due to the Poisson effect are mainly compression forces with values close to zero. The shear membrane force distribution shown in Figure 7—17 further indicates that the high level of shear forces is a local behavior and most of the FRP membrane is subjected to a low level shear forces, i.e., values close to zero. Therefore, the FRP membrane is primarily subjected to longitudinal tensile stresses rather than transverse stresses. According to Table 7—6 and Table 7—7, although the concrete deck is subjected to higher out-of—plane section resultants than the FRP membrane, the concrete deck is still dominated by in—plane membrane forces. Furthermore, the deck is mainly subjected to longitudinal compression (Figure 7—15), which introduces a small level of transverse tension due to the Poisson effects (Figure 7—16). Therefore, the optimized FRP CMB bridge behavior is such that the FRP membrane carries the in-plane longitudinal tensile and shear forces while the concrete deck carries the compressive forces. 116 .828 N; :80— 8 £83 “.8: 5m 8qu 80808 mcumwsrw ”m2w 280820 mo 85: .-m :82 80% 583 :8 Mom 088 80808 w8caom ”N2m .H2m No.0- mo. 0. 2. o- 82 285802 8.0 x52 \ .80. \oo m4. 0. .o_\H.M 82 zoom nmd mm v on. m 8“2 . . .._ A8\ 8 -Zvc moeom cocoom 8N2 82 m2m 8N2 82 N2m V82 82 22m 088.5 m20 8880 2: 88 $5808 conoom 515 28g .288 NA :32 8 883 88 “on 889 889808 88m ”am 95820 mo warm .-2 :80— 8 883 88 Ba 088 889808 8080 ”mum .H mm Tmimm- 05.2- 3.07 82 889802 mm.Nm ON; bwdg 82 868 8.8 8: . H . . . . A8203 880 m 888m x32 52 mum x32 52 Nmm x32 52 7mm . 0885 m20 38880 2t 88 888 882.com elm 0383 117 cozomhl “COCOQEOO finDuO-fiufl /o /o /o /o /o /o /o O O 0 O 0 O 0 O O 0 5 0 5 0 5 0 o/ o/ 4 3 3 2 2 1 4| 5 0 ___r~_ L.~_ ,_.L b8.» _ — pphlhk _ _ _ 2 a 2 . _ e _ n m k .m r m e M a ID 1 N725}. \ \\\\\ ... \ .s. \ \. \ \ \ xx. \ \.\ \ .x \ .\ \ \.\ 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 7 6 5 4 3 2 1 8:00 8082m— -100 -75 -50 -25 25 50 75 100 133.9 Force Range (kN/m) Figure 7—15 Section force (SF 1) distribution -105.2 c0308.... 80000800 83085 % % % % % % . 0 0 0 0 0 0 o/ 6 5 4 3 2 1 0 p L _ h p p p _ _ e _ . p _ p l. _ 2 u _ m m H. k m r m e M I I w\\\\\t\\\\\\\x\ _, _ h p N F _ p _ _ _ qq—fidqui—l—ufi_-—H~Jfij_ gall. _—q—u~—rlq_~——u—_q—uu_ 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 9 8 7 6 5 4 3 2 1. 1 8:00 EOEOE 12.6 12 3 0 -3 -12 -13.8 n .m w. b .n d m) [2 NF kS (( C «in .narm Rm ea 6 ms 06 F4 7 m U .wo F 118 700 [.-_ _ _. {I Deck f 40% H] Membrane I r; 35% l J 1 600 500 {— 30% 25% «b O O 1 / / z / I r, ,z I I _r ,v - ' ' '2 .- . l _/ ',. /,/ // f. 2 ,; .1. /. . ./. / ,{ I. I. 1. ./ ' /' .t' r‘ r’ / z z ./ ' " I' r' ,1 // / ' / / 1' 1' 1' v / . . ‘ -' / _/ I I, I . .4 r, 4' .‘ 1 ‘ I ' ' 1‘ 1 1 l / \ -2 20% Element Count 0) O O 1 1_L 1 ~— 15% \\\ : \\ 10% Structural Component Fraction f 5% . .. l i : - "0% -1o 0 10 20 30 52.3 Force Range (kN/m) Figure 7—17 Shear section force (SF3) distribution 7.1.3.2 Stability evaluation As previously discussed, shape optimization of a shape resistant structure will improve their structural stiffness. Such optimized structures become stiffer and carry their design loads primarily by axial or membrane action rather than bending. This leads to a response that usually involves very small deformations prior to buckling. Therefore, shape optimized structures are stability sensitive. The stability of the optimized CMB bridge was investigated by eigenvalue buckling analyses, which are generally used to estimate the critical buckling loads and potential collapse shapes of stiff structures. Considering that the material properties of the optimal 8-layer laminate were only slightly different to the optimal 48-layer laminate, the laminate design of [il7°/i19°]s was employed to reduce the computational cost of the buckling analyses. The buckling 119 analysis of the optimal CMB bridge considered a perturbation uniform load of 6.9 kN/m2 on the concrete deck acting in the gravity direction. The method of subspace iteration eigenvalue extraction [HKS, 2000] was used to predict the first five eigenvalues and the corresponding eigenmodels for the optimal CMB bridge as summarized in Table 7—8. Table 7—8 Buckling eigenvalues and eigenvectors of the optimal CMB bridge Mode Eigenvalue Eigenmodel 1 -76.909 2 —76.909 3 -77.137 4 -77.137 5 —79.923 As shown in Table 7—8, negative eigenvalues were obtained from the buckling analyses. Such negative eigenvalues indicate that the optimized CMB bridge would buckle if the perturbation load were applied Opposite to the gravity direction. However, the first five buckling modes show that the structure has closely spaced eigenvalues, which indicates that the structure is imperfection sensitive. Therefore, further consideration to buckling stability is required for CMB bridges if the applied loading can act in a direction opposite to that of gravity. 120 7.1.4 Post optimality analyses Structural designs by Optimization methods are obtained with optimal parameters achieved regarding certain optimization conditions. For example, the optimum design defined by the optimal geometric and material parameters for the CMB bridge is achieved with respect to an Optimization load case. However, the CMB bridge might not be constructed exactly to the Optimal geometry and the optimal laminate specified by the optimal design variables may not be easily achieved in practice. Furthermore, multiple load cases that must be resisted with corresponding performance limit states are typically considered for the design of bridge structures. The performance of an optimized structure could thus be compromised due to changes in structural geometry, loading and material properties. Therefore, the response of the optimum solution to the CMB bridge system due to geometric imperfections, loading alterations and material property changes (bend- twist coupling effects) was investigated. The CMB bridge model used in the post optimality analyses was constructed by averaging dimensions of the three optimized CMB bridges previously obtained (see Table 7—4). 7.1.4.] Influence of bending twisting coupling effect of laminate properties The influence of material coupling effects to the optimal CMB model was investigated by analyzing CMB bridge systems with different optimal laminates for the FRP membrane that varied only in their coupling stiffness terms (i.e., 016 and D26). The previously obtained optimal 8-layer and 48-layer laminates, which varied in the out-plane material properties (see Table 7—5), are thus used in this analysis. The structural responses of the two CMB models constructed by these two balanced-and-symmetric laminates are compared in Table 7—9. 121 Table 7—9 Structural response of optimal CMB bridge with different Optimal laminates CMB In-plane section Out-of—plane section Max Strain membrane resultants (kN/m) resultants (kN-m/m) Deflection Energy model SFl SF2 SF3 SMl SM2 3M3 (mm) WW 8- Max 133.87 1.20 52.33 0.15 0.06 0.02 layer Min -1051 -13.79 —52.31 -O.13 -008 -002 48- Max 133.88 1.29 52.31 0.15 0.06 0.02 layer Min -1047 -13.76 -52.31 —013 -007 -002 SFl, SF2: Direct membrane force per unit width in local l-, 2-axis of elements. SF3: Shear membrane force per unit width in local 1-2 plane. SMl, SM2: Bending moment force per unit width about local 2-, l-axis of elements. SM3: Twisting moment force per unit width in local 1-2 plane. 1.692 84.51 1.692 84.51 According to Table 7—9, both of the material-modified CMB bridges have the same strain energy while maximum deflection and the in-plane section resultants in the membrane vary slightly. This can be explained by the in-plane dominated behavior of the optimized CMB bridge. The optimized CMB bridge is primarily subjected to in-plane longitudinal tension rather than out-of—plane section resultants. Furthermore, the small magnitudes of the bend-twist coupling terms (D 16 and D26) compared to the other major bending stiffness terms (D11 and D22) also partially reduce the bending—twisting coupling effect. Therefore, material property changes in bend-twist coupling can be neglected for the structural behavior of the optimal CMB bridge. 7.1.4.2 Influence of geometric variations Table 7—10 summarizes the geometric variations of the CMB bridge using average shape dimensions with respect to the three optimal CMB bridges in terms of geometric design variables. It can be noted that the maximum variation of 1.94% occurs for h 1 (the height at the end of the bridge) by using average dimensions. Consequently, the strain energy of the CMB bridge with average dimensions has an 0.7% increase (Table 7—4 and 122 Table 7—9). Although the CMB bridge combining several slightly-different optimal designs by averaging shape geometries can be still considered an optimum design in the current application, it can be anticipated that adequate performance of optimum designs can be compromised by geometric imperfections. Table 7—10 Geometric variations of optimal CMB bridges W1 W2 I11 I22 I (mm) (turn) (turn) (Hun) (turn) Model A 1359.0 2184.4 984.3 1219.2 19.05 Model B 1365.2 2184.4 1022.4 1219.2 19.05 Model C 1352.6 2222.4 997.0 1219.2 19.05 Average 1358.9 2197.1 1001.2 1219.2 19.1 Variation 0.46% 1.00% 1.94% 0.00% 0.00% Variation = (Standard deviation of a dimension) / (Average dimension) 7.1.4.3 Influence of loading alterations Structural designs require a structure that meets design criteria (i.e. limit states) with respect to different load cases. However the CMB bridge was optimized with respect to a single load case. Other load cases, i.e., the service load case and the strength load case, are required to comply with structural design requirements for different limit states [ASSHTO, 1998]. Therefore, the performance of the optimized CMB bridge due to these two load cases was studied with respect to the corresponding design criteria. Considering that most unidirectional laminated composites behave in a brittle manner, local failures generally can lead to complete fracture and total loss of load- carrying capacity. Therefore, the first-ply failure theory, which assumes failure of a composite laminate to take place when failure initiates in the most critical layer of the laminate, is used to evaluate the material strength of laminates. Several stress criteria have been developed to predict the material failure of FRP lamina under different limit states [Jones, 1999]. In addition to the maximum stress criterion implemented in the 123 integrated optimization procedure, other three commonly used stress criteria were employed to calculate the ultimate strength of laminates in the study of load alternations. The criteria used were the Tsai-Hill [Tsai, 1968], Tsai-Wu [Wu, 1974] and Azzi-Tsai- Hill [Azzi and Tsai, 1965] models. These models are more accurate in comparison to the maximum stress criteria since they consider polynorrrialcombination of the stress state rather independent relations [J ones, 1999]. The structural responses and the stress indices of the optimized CMB bridge with respect to the different limit states are summarized in Table 7—11 and Table 7—12 respectively. As expected, the strength limit state has the maximum influence on the optimized CMB bridge as it results in maximum in-plane stress demands and a maximum deflection. With regards the deflection criterion, the maximum deflection obtained in the service limit state reached 5.237mm, which is 41.65% deflection limit (U800) of 12.573 mm. It should be pointed out that the maximum stress index reached 5.39% of the material failure, that is 21.54% of the allowed service stress limit (Table 7—1). Therefore, the deflection criterion governed the design of the Optimum CMB bridge under service loading. Furthermore, it can be noted that the stress indices are below the design limits proposed for each limit state (Table 7—1). Although the optimized C1VIB bridge satisfied the design criteria specified in Table 7— 1 for all limit states, the maximum stress criterion has the lowest stress index, which underestimates the stress level of laminates compared to the other failure stress criteria. Therefore, a more accurate material strength criteria should be applied in the stress constraint in future optimization studies. 124 Table 7—11 Laminate stresses of the optimized CMB bridge under different limit states Load case (N/Siirlmz) (N/Erzrinz) (NS/i312) Max $23M Optimization 11:44: 2:: if; -8233 4.692 3...... lit: 32:23 431?}. $32 -5537 Streng‘“ ii: 322:? i213: -3133 -8-649 S11, S22: Direct stress in local 1-, 2-axis of elements. S12: Shear stress in local 1-2 plane. Table 7—12 Stress indices of the optimized CMB bridge under different limit states Load Case MStrs (%) TsaiH (%) TsaiW (%) AzziT (%) Optimization 1.73 2.17 2.03 2.05 Service 4.31 5.39 5.02 5.07 Strength 7.01 8.76 8.17 8.24 In addition to the loading alteration by load cases corresponding different limit states, the structural response of the optimized CMB bridge due to the extreme loading effects introduced by the loading pattern of moving trucks in the strength limit state was also explored. Two extreme truck loading positions were considered. One loading pattern was introduced to create maximum torsion by placing the truck loads at the edge of the optimized CMB bridge (Figure 7—18). A second loading position sought to develop the maximum mid-span moment by placing the two axel loads (142.4 kN) in the middle of the bridge (Figure 7—19). The structural responses and stress indices for the Optimized CMB bridge due to these two critical loading patterns are summarized in Table 7—13 and Table 7—14. 125 Truck Load Live Load Live Load Figure 7—19 Load case for maximum bending in the strength limit state As shown in Table 7—13 and Table 7—14, the maximum structural response is developed due the load case of maximum torsion. The maximum deflection reached 11.82 mm and the maximum stress demand achieved about 21% of the material failure stress state (Table 7—14), which is 2 times greater than the maximum stress level of the 126 optimized CMB bridge under strength loads. Also, the transverse stresses (S22) are obviously increased due to the twisting behavior introduced by the truck loads (Table 7— 13). Although the optimized CMB bridge is still within the stress limit for the strength limit state (see Table 7—1), the structural responses under the critical load patterns are clearly different than those obtained for the optimum load case. Therefore, structural Optimum designs with respect to a typical load pattern should be designed with strength reserves to allow possible increases of structural response due to non-optimal loading alterations while in service. Table 7—13 Laminate stresses of the optimized CMB due to different extreme effects S S S Max deflection Load case (N/rrlrinz) (N/rrzrinz) (N/rrlrinz) (mm) Max 33.96 12.94 3.03 Strength Min -26.71 -l6.52 -319 '8'65 Maximum Max 48.74 23.54 4.07 -1 1.82 Torsion Min -38.36 -50.43 -4.94 Maximum Max 44.83 18.07 4.13 _9 50 Bending Min -37.27 -22.73 -4.22 ' S“, S22: Direct stress in local 1-, 2-axis of elements. 312: Shear stress in local 1-2 plane. Table 7—14 Stress indices of the optimized CMB due to different extreme effects Load Case MStrs (%) TsaiH (%) TsaiW (%) Azzit (%) Strength 7.01 8.76 8.17 8.24 Maximum Torsion 20.63 21.14 21.44 21.14 Maximum Bending 9.30 11.94 10.76 10.83 7.2 FRP composite membrane suspension (CMS) bridges Inspired by conventional cable suspension bridges, an FRP composite membrane bi- suspended (CMS) bridge (see Figure 7—20) was conceptualized as a second case study for the developed optimization approach. In CMS bridge systems, a deck system (either a 127 conventional concrete slab or an FRP panel) is assumed to be placed on top of the FRP membrane such that it transfers the applied live loads through internal diaphragms or “bulkheads.” Thus, the in—plane stiffness of the FRP membrane is employed in two—way tension under uniform loading from the deck system. A conceptual depiction of a three— span CMS bridge is shown in Figure 7—20. Membrane Deck Diaphragm ‘. Figure 7—20 Concept for Composite Membrane Suspension (CMS) Bridges 7.2.1 Bridge system description The CMS bridge proposed for Optimization studies is a 60.96 m long three-span bridge with a 10.36 m wide concrete deck to accommodate two 3.66 m traffic lanes and essential shoulders (Figure 7—21). The concrete deck is 304.8 mm thick with a density of 2400 kg/m3 for normal concrete. The loads acting on the bridge include the self-weight of the concrete deck, and the live lane and truck vehicular loads. The live lane load, which acts over the 3.05 m wide lane, is assumed to be distributed over the entire surface of the deck, resulting in a pressure of 3.06 kN/mz. The total truck loads introduced by two trucks for two lanes are also considered to be ideally distributed over the FRP membrane through diaphragms as a pressure of 1.014 kN/mz. The laminates used for this bridge system are based on typical medium—performance carbon/epoxy material system for use 128 in civil structures. Thus the lamina orthotropic properties are taken as: E“ = 155.025 GPa, E22 1' 10.335 GPa, 012 = 6.89 GPa, 013 = 6.89 GPa, (323 = 0.689 GPa, and v,2 = 0.3. ~ L = 60.96 m (a) Side view N N r\ l\ 0 El 0 ..l N N 18. —m. ooooooo i—aofi ———————————————— + ———————————————————— 2+ 15 at r .11 1 5.54 m C. N A B | L_ L = 60.96 m ll (b) Plan view (c) Cross view Figure 7—21 Geometry of the CMS bridge The CMS bridge is subjected to dead loads (DL), two lanes of uniformly distributed lane loads (LL) and two three-axe] truck loads (TL) [AASHTO, 1998]. Two load cases using different combinations of the applied loads, which are specified in Table 7—15, were selected to analyze the CMS bridges for service and strength limit states. As previously used for CMB bridges, the tension and compression stress limits based on the material properties for l-axis (fiber-longitudinal direction) and 2-axis (fiber- transverse direction) were chosen as X1 = 1.55 GPa, XC = -O.775 GPa and YT = 0.103 129 GPa, YC = —0.052 GPa respectively, and the shear stress limit in 1-2 plane is taken as S 234.45 MPa. The ultimate material failure 0;, is then determined by combining the stress limits measured on the material directions according to a stress criterion. Table 7—15 Load cases specification for CMS bridges Design Criteria Load Cases Specification Stress Strength 0250'u L/800 Service Limit Deflection State Service load l.0>
<(LL + 1.33xTL) case Strength load 1.25xDL + 1.75X(LL + 1.33xTL) case 0.500Ll — Strength 7.2.2 Integrated optimization of CMS bridges Based on the above-mentioned geometric, material, and loading definition for the bridge systems, the integrated optimization approach was applied to design FRP CMS bridges starting with the shape and laminate-property optimization and then followed by the laminate design Optimization. 7.2.2.1 Shape and FRP laminate-property optimization Taking advantage of structural symmetry, the hyperbolic profile of the CMS membrane can be controlled by four key points as shown in Figure 7-22. The shape design variables were therefore the six degrees of freedom of the four geometric key points. The membrane system was completely defined by spline interpolation of the key points. The positions of the pier supports (Figure 7—21 (a)) are constant throughout the optimization process. The rational of this assumption will be discussed in Section 7.3.3. The laminate design variables are the lamination parameters {VI/r, V3A, V10, V312} controlling the in-plane tension stiffness and out-plane bending stiffness. 130 Geometric key points Figure 7—22 Loading and geometric key points for the CMS bridge As applied in the integrated optimization of CMB bridges, an inequality constraint was employed to limit the used mass of the membrane FRP laminate. Therefore, a material constraint was implemented by setting a lower and upper bound for the coordinates of the shape key points (Table 7—16). Maximum stresses monitored at the top and bottom surface of the FRP laminate and maximum deflections of the complete system constituted two additional constraints required to enforce the structural design criteria in the integrated optimization procedure. Instead of only using dead load and live lane in the optimization load case, the load case for the service limit state (Table 7—15), which includes truck loads, was chosen as the optimization load case for the CMS bridges. 131 Table 7—16 Coordinates constraints for geometric design variables Coordinate P1 (m) P2 (m) P3 (m) P4 (m) bound x y z x y z x y z x y 2 Lower -6.1 —6.1 -10.4 -6.1 ~104 -6.1 Upper 30.48 0.0 0.0 0.0 0.0 0.0 30.48 -7.8 0.0 0.0 -7.8 0.0 Two typical boundary conditions for suspension bridges were considered in the CMS bridge analyses (Table 7—17) to evaluate the effect of boundary conditions on the optimized shape. Model A was provided with simple supports at the positions of the piers while Model B featured extra roller supports at both ends of the bridge. Table 7—17 Initial designs of CMS Bridges Initial design Model A Model B Boundary conditions { 21,22, Z3, 24, y3, y4} (m) {0, 0,0,0, -10.4, -104} Laminate design [O°/90°]s {V1A,V3A,VrD,V3D} {01.0-75.1} The integrated optimization approach was implemented as previously described and the performance and stability of the obtained results was assessed by studying a single initial design (geometry and laminate layup) for each CMS bridge model (Table 7—17). Both integrated optimizations were initiated with a plane FRP membrane constructed with a cross-ply laminate, i.e., a [0°/90°]s. The first-level optimization history of the two CMS bridges is shown in Figure 7—25 and Figure 7—28, respectively. 132 -2 -— ‘v ‘0; A_4 __ E . v C .2 7‘ a) -6 -— ffffffi . . . c _ Q) 4 g - D Iteration no. Figure 7—23 Histories of geometric design variables — CMS Model A 1 .1 1| T U) . h 7'- 1 G) 0.5 4-0 . a) E \ I (U l \/\, h . m - , l , O 7 4 i t: . N Is g - ...... _l'0.5 “ ..... I+V1A +V3A +V10 +Vao l -1 rrrrrrrrr i rrrrrrrrr i 111111111 i 111111111 i 11111111 ri11111T111i 111111111 0 10 20 30 40 50 60 70 Iteration no. Figure 7—24 Histories of laminate-property design variables — CMS Model A 133 107 ; 106 - ’15 2 ii.- j, , _ ‘ . Optimal Shape ‘x’ 1 05 $49195:gé‘méfigwfifwwmnmkk 1:} 1;" , ;«:07 2.9:“ .. >5 Plan View 8’ a) 5 c 10" —- 8 7 150 View Elevation View H (I) 103 102 1 . . . . . 0 10 20 30 40 50 60 70 Iteration no. Figure 7—25 History of objective function — CMS Model A 0 .2 ; ’E‘ -4 c .9 w -6 c 0.) .§ : D -8 . -10 . 1:”21 +Z+Zo;z“:* +Y .12 . . .. . ,; .Wr 1 #44 0 10 20 30 40 50 60 70 Iteration no. Figure 7—26 Histories of geometric design variables — CMS Model B 134 parameters Lamination I .° 01 Strain energy (kN-m) P a: O . i . ‘ v1 A VJA v1 D VJD ‘ 0 10 20 30 40 50 60 70 Iteration no. Figure 7—27 Histories of laminate-property design variables — CMS Model B 107; 105. Optimal Shape Plan View —I O 0! IsoView ..x O p. I l Elevation View A O u i 1o2 Iteration no. Figure 7—28 History of objective function — CMS Model B 135 The optimal results obtained from the first-level optimization are summarized in Table 7—18. It can be seen that although both models have a similar geometric shape, the optimized objective of Model B has a strain energy value 20% lower than that of Model A. The improvement in minimizing the strain energy came from changes in the laminate properties due to the different boundary conditions. In Chapter 6 it was demonstrated that structural optimization can be further improved by material optimization. Optimum laminates were thus also determined for the CMS bridges using the second-level optimization so as to achieve the obtained optimal lamination parameters. Table 7—18 Optimal results of the first-level optimization for the CMS bridges Optimum design Model A Model B Strain energy (kN-m) 183.26 149.93 {2 z z z } (m) {-6.05, -6.10, —0.43, -0.37, {—6.10, -6.10, -O.22, -0.30, 1’ 2’ 3’ 4’ Y3’ W -8.20, —7.99} -779, -7.85} {V1A,V3A,V1D,V3D} {0.138, 0.033, —0.317, —0.533} {0.155, 0.106, —0.309, -0.494} It should be noted in Figure 7—28 that the objective function experienced an abrupt decrease after a relative smooth development in the iteration history of CMS bridge optimization. The occurrence is thought to be inherent from the gradient properties of the objective function with respect to the design variables for this optimization case. 7.2.2.2 Laminate design optimization In the second level optimization, the genetic optimization procedure was employed to design optimal laminates with a maximum error of 1% with respect to the optimal lamination parameters obtained from the first-level optimization. The genetic iteration histories for the laminate designs of the two CMS bridge models are shown in Figure 7—29 and Figure 7—30, respectively. The optimal laminate for Model 136 A was found to be [i51°/i57°/i60°/i63°/i500/i670/i620/i21°/0°g]3 with a minimum error of 0.807% with respect to the Optimal lamination parameters. The optimal laminate for Model B was [i53°/i51°Ii64°2/i53°/i62°/i64°/i4°/0°8]5 with a minimum error of 0.643%. 35% _ 30% 25% ‘ 5‘ : . . . c 20% -. Optimal Lamination Parameters Q) - t - 35:: (2.-23;. h A, :- _ , :- _ 8 15% __ ID 30 E Optimal Laminate Design 10% j [:51°/1-57°/i60°/i63°/i50° : /i>67°/i62°/fl1°/0°8]S % 3~ 5 : Error=0.807% 0%4Tr:r'rrruriur-riwrrviwrrwirrrui o 1000 2000 3000 4000 5000 6000 Iteration no. Figure 7—29 Second-level optimization process of Model A 137 40% _ 35% f- 30% f— 325% - , , . c Optimal Lamination Parameters d) B VID : -Oo309, V3D : '0.494 (D D 15% " Optimal Laminate Design 100/ [fly/:5 1°li64°2/-_t5 3° ° /:62°/:64°/1-4°/0°8]S 5% i' Error = 0.643% 0%- fl % mfimfi rrrrrrrrr i vvvvvvvvv r 111111111 . .......... 0 1 000 2000 3000 4000 5000 6000 Iteration no. Figure 7—30 Second-level optimization process of Model B 7.2.3 System characteristics of optimal CMS bridges The system characteristics of the optimized CMS bridges was investigated by studying the structural response of the bridge systems under the optimal load case and the stability of the optimized CMS FRP membrane. The influence of boundary conditions on the optimum designs was evaluated by comparing the results from both models. 7.2.3.1 Structural behavior of optimal CMS bridges The section forces and moment resultants for the CMS bridges are given in Table 7— 19 and Table 7—20, respectively. From these tables it is clearly seen that the FRP membrane of both CMS bridges is dominated by in-plane stress demands. In addition, a further review of the section forces along the local 1- and 2-direction (Figure 7—31) 138 shows that the FRP membrane is subjected to two-way tension, as the relative magnitude of both of these stress fields is similar. Therefore, the behavior of FRP CMS bridges is such that the FRP membrane acts as a membrane structure carrying distributed loads primarily through bi-directional tension. Figure 7—31 Coordinate systems 139 .233 N; :82 E 523 ES 5Q 080m EoEoE wfiumgh 638 95820 mo was; .-N :82 Sosa £23 :2: Ba 088 :88po mEmEom “NEm 42m ow.m- mvd- 946- 82 m #0602 ow.m 0N6 Nw.N x32 Dev- exam- mmw- £2 < $602 3:4 wfim V©.N 5&2 II as..- 2.: I_I as..- ze II gaze 3; gas E2 mEm 52 NEW 52 22m . mowW/HEL mEO 385% com 8562: cocoom omln 2an .283 N; :82 E 523 E5 on 088 0:89:58 Scam ”Em $5820 .8 germ i :82 E 523 can So 088 285an 8th ”mum 4 mm NNdET NbSK- . hdew- E: m 3on NNdHOH hmdovm Sudfl MN 5&2 34.057 o_.:©- L ov.m~:- 52 < 3602 wmdafi mnNnvm mmwfimm 5&2 ll Ezé E73 880m 850m mmm x32 82 mum xwz E2 Em . mowEE mEO BEES com 888 cowoom elm 28d. 140 7.2.3.2 Stability From above-discussed results, the optimized CMS bridges are shape resistance structures in which high structural stiffness is primarily gained by in-plane membrane actions rather than flexural actions. As previously addressed, such structures are buckling sensitive and usually require only small deformations prior to experiencing buckling. Therefore, the stability of the optimized CMS bridges was investigated through buckling sensitivity analyses. Eigenvalue buckling analyses determined with ABAQUS [HKS, 2001] were used to predict critical buckling loads and potential collapse modes of CMS bridges. The loading of the service limit state was employed as the perturbation load required for the buckling analyses. The method of subspace iteration extraction was chosen to obtain the first four eigenvalues and corresponding eigenshapes, which are summarized in Table 7—21. According to the results shown in Table 7—21, negative eigenvalues are reported for both CMS bridge models. Such eigenvalues imply that buckling of the optimal CMS bridges would happen only if the perturbation load were applied opposite to the gravity direction. Further comparison of the two models shows how the different boundary conditions will influence the buckling modes of the structures. Model A has obvious buckling deformations at the free ends of the bridge while buckling of Model B will primarily happen at the positions where the columns support the FRP membrane. Considering that Model B is stiffer due to its lower strain energy, it is understood that Model B is more sensitive to buckling and, therefore, has higher and more closely spaced eigenvalues than Model A. 141 Table 7—21 Buckling eigenvalues and eigenshapes of the optimal CMS bridges Mode Model A Model B Eigenvalue Eigenshape Eigenvalue Eigenshape Base State ‘ ' 1 —0.037 —0.2526 2 -O.114 -0.3151 3 -0.120 —0.3152 4 —0.125 -0.3153 7.2.4 Post optimality analyses Optimal structural designs are achieved by satisfying certain optimization paramters and constraints such as loading and boundary conditions. Change of such conditions can alter the optimum designs whose response may violate the constraints originally satisfied. Therefore, it is necessary to evaluate the influence of such changes on the behavior of optimum designs through post optimality analyses. The following analyses focus on the influence of loading and boundary—condition changes on the performance of the previously-obtained optimal CMS bridges. 142 7.2.4.1 Influences ofloading alternations Since the load case for the service limit state was used to achieve the optimal CMS bridges, the constraints of maximum deflection and maximum stresses were verified for the strength limit state load case on the optimal CMS bridges. A first—ply failure analysis was applied to predict the material ultimate strength of the laminated FRP membrane. In addition to the independent maximum stress criterion implemented in the integrated optimization procedure, the coupled polynomial stress failure theories by Tsai-Hill [Tsai, 1968], Tsai-Wu [Wu, 1974] and Azzi-Tsai-Hill [Azzi and Tsai, 1965] were also used in a first-ply-failure criterion. These models are readily available in ABAQUS to estimate the stress failure of FRP laminates. The structural responses and the stress indices of the optimized CMS bridges with respect to the different limit states, see Table 7—15, are summarized in Table 7—22 and Table 7—23, respectively. As shown in the tables, although the maximum stresses and maximum deflection are increased from the service limit state to the strength limit state, the stress indices for both models of the optimized CMS bridges indicate that the load case associated to the service limit state is still the critical load case for the bridge system. It should be noted that the stress index evaluated by the maximum stress criterion employed in the optimization process predicted that the stress level in the laminates was within the allowable limits and the material strength. However, as seen in Table 7—23, the bi—axial stress failure criteria indicate that the stresses on the FRP laminate exceed the service stress limits. Therefore, multiple stress failure criteria should be considered as stress constraints during optimization processes. 143 Table 7—22 Structural response of optimized CMS bridges for different limit states SH $22 812 Max (16116011011 MOdel Load case (N/mmz) (N/mmz) (N/mmz) (mm) . Max 525.08 39.22 20.96 A semce Min -133.91 -15.08 -11.41 40030 Max 715.88 53.48 28.57 Strength Min -182.57 20.55 -1555 54559 Service Max 613.59 46.30 19.57 139.95 Min -187.88 ~19.06 —15.52 Max 836.54 63.12 26.68 Strength Min -256.15 -2599 —21.17 19”” Sn, S22: Direct stress in local l-, 2-axis of elements. S12: Shear stress in local 1-2 plane. SP1, SP2: Minimum, maximum principal stresses Table 7—23 Stress indices of optimized CMS bridges for different limit states Model Load Case MStrs (%) TsaiH (%) TsaiW (%) AzziT (%) Service 23.47 32.38 29.23 30.73 Strength 32.01 44.14 39.84 41.88 Service 20.78 27.20 25.30 25.83 Strength 33.86 45.70 52.90 45.70 MStrs: Maximum stress theory failure measure TsaiH: Tsai-Hill theory failure measure. TsaiW: Tsai-Wu theory failure measure. AzziT: Azzi-Tsai-Hill theory failure measure. In the optimization procedure, the design loads applied on the FRP membrane were ideally considered to be uniformly distributed over the area that the concrete deck projects onto the FRP membrane. However, in practice, it is conceptualized that the concrete deck weight and live loads will be transferred to the FRP membrane through diaphragms or “bulkheads.” Therefore, spaced line loads (Figure 7—32), which simulate the pattern of the loads transferred from the diaphragms, were applied on the FRP membrane as a non-optimal loading pattern. The structural response of the optimal CMS bridges subject to spaced line loads at the service limit state will be used to verify the design criteria and is compared to the structural behavior of the optimal CMS bridge subjected to the distributed loading. 144 Spaced line loads Membrane Diaphragm Figure 7—32 Loading pattern of spaced line loads The structural response of the optimal CMS bridges subject to the spaced diaphragm line loading pattern is summarized in Table 7—24, Table 7—25, Table 7—26 and Table 7— 27. By comparing the section forces and moments, it is shown that the response of both models still remains governed by in-plane stresses. However, it is easily seen in the figures of Table 7—26 and Table 7—27 that the distribution of the section forces and moments in the FRP membrane has obvious changes along the loading pattern, while the magnitude variation of the section forces and moments varied only modestly. A comparison of Table 7—22 and Table 7—24 shows that compressive forces noticeably increased in the FRP membrane of Model A after the loading pattern was altered. This lead to the stress index of the FRP membrane of Model A (see Table 7-23 and Table 7— 25) to be almost twice than the stress demand limits chosen for the service limit state. However, the stress level and stress indices of Model B had smaller variations compared to Model A due to changes in the loading pattern. It can be further noted for Model A that, although its section moments remained small compared to the in-plane forces, the magnitudes of the section moments increased considerably for the non-optimal loading pattern as compared to values obtained under 145 the uniform load distribution (Table 7—26 and Table 7—27). Accordingly, the structural strain energy under the non-optimal load pattern is more than two times that of the Optimal value previously obtained (Table 7—25). The increase in strain energy is caused by the increase in the section moments throughout the FRP membrane, which can be clearly seen in the figures and values given in Table 7—27. Consequently, optimal Model A, obtained by a uniform loading pattern, might not be considered as an optimal shape for the loading pattern of spaced line loads. On the other hand, the strain energy of Model B increased only moderately over the optimal strain energy value. Table 7—24 Structural responses of the optimized CMS bridges subject to spaced line loads under the service limits state Stress Model S1 1 S22 S 12 SP1 SP2 (N/mmz) (N/mmz) (N/mmz) (N/mmz) (N/mmz) A Max 439.21 25.99 27.65 17.43 439.29 Min -303.67 -2002 -26.61 —304.75 -18.82 B Max 506.56 36.96 19.74 36.71 506.74 Min —135.39 -1340 —10.02 —135.75 -1153 811. 822: Direct stress in local 1-, 2-axis of elements. Sn: Shear stress in local 1-2 plane. SP], SP2: Minimum, maximum principal stresses Table 7—25 Stress indices of the optimized CMS bridges subject to spaced line loads under the service limits state Stress Index Max Strain M d 1 ' 0 e MStrs (%) TsaiH(%) TsaiW (%) AzziT(%) (1623:?“ fit?) A 40.10 58.83 68.73 58.83 427.74 414.20 B 23.76 26.65 28.65 25.95 139.65 193.20 MStrs: Maximum stress theory failure measure TsaiH: Tsai-Hill theory failure measure. TsaiW: Tsai-Wu theory failure measure. Azzit: Azzi-Tsai-Hill theory failure measure. 146 .0829 NA :82 8 £23 :8: Hon 080m 80808 M85831 “92m 880820 mo 88; .-m 302 309m £23 85 SQ 080m 80808 wEvaom ”N35 435 3M- mme- wow. 82 mfiowoz $4.“- mmd- N34- 82 1:352 mwd 344m mvd 522 II €8-va II 958. vaL II ES. 20: 828 88% 82 mEm 8&2 82 92m 522 82 :2 mowgno 803mm @552 8 Loo .33. mowers. mEU 385% mo $80808 cocoom mm! 2an .283 m; :82 8 823 :8: 5m 088 0889808 808m ”mum 880820 We megm .-fi :82 8 823 88: Ba 8.8% 055808 885 ”mum . H mm $43- ER- 3%. £2 - L 8282 :42 398 L 338 a: 82:- 0.24- .. 082- e: L E682 N3: oammm . 9:48 8.2 i €20: €20: €20: 8868 888 382 £2 Em 38: a: 9% a: a: Em . mowqgo 8.5th WERE 8 80.33 mowers mEU E885 mo 888 803.com cmln 2an 147 The summarized results point out that the performance of an optimum design can be significantly altered by changes in loading patterns. However, Optimal CMS bridge Model B was found to be less sensitive to loading pattern changes than Model A (their difference arising from their boundary conditions). Thus, the optimal CNHB bridge of Model B would behave better under changes in loading patterns during service. 7.2.4.2 Influence of boundary conditions As previously discussed, the optimal CMS bridge designs show different performance detriments due to changes in loading patterns. The change in performance due to non- optimal loading patterns is largely originated from the difference in boundary conditions. Therefore, it is necessary to investigate the influence of boundary conditions on the optimum design of CMS bridges. This can be done by comparing the in-plane behavior of the two CMS bridge models under the service limit state. The membrane forces of Model A and Model B are compared in Figure 7—33 and Figure 7—34. According to Figure 7—34, both models are primarily subjected to in-plane tensile forces and have an almost equal force distribution in the transverse direction of the bridge. However, although both models are subjected to tension and compression in longitudinal direction, their in—plane structural behavior, demonstrated by the distribution of transverse section forces, is different. Figure 7—33 shows that about 85% of Model B is subjected to membrane forces with 300 kN/m or less while 15% is subjected to membrane forces higher than 300 kN/m. In contrast, about 30% of Model A has membrane forces over 300 kN/m. This is the reason that Model A has a higher optimal strain energy than Model B. The variation in section force distribution is caused by the difference in boundaryconditions between the two models. These results, combined the 148 canon.“— 3.3035 o/o % o o 5 O o/ o/ 1 1 5 o e L p b L _ \\\\\ E2 Model A _ BModel B 4- 20% Ezoo EoEoE Force Range (kN/m) Figure 7—33 Section force (SFl) distribution in CMS optimal designs 5:08“... 3.32:5 ____________________________ Model A Model B - § Z § V/////////% 1800 _ 1600 3— 1400 —} X a M m. .w 0 0 d 3 .nla .m t p w 0 .. m )C mm 0]“ 5N0 kfi. (..w e. Mm O .1 3.0 R3 eF S om( 60c C Fm f n 0 O 5 .U ... m S M W _ 3 7 . e r u .We F 149 issues discussed in the system characteristics and the load variation studies, lead to the conclusion that Model B has a more effective and stable in-plane behavior than Model A. This improved in-plane behavior is achieved by the addition of supporting boundaries at both ends of Model B (Table 7—17), which strengthens the dual hanging concept of the CMS bridge system. 7 .3 Discussion of FRP membrane-based bridge systems In this section, the conceptualized FRP composite membrane-based bridge systems (CMB, CMS—A, and CMS-B) are evaluated by comparing their structural behavior in terms of material use efficiency. In addition, the selection of an objective function that could further improve the integrated shape and laminate optimization approach is also discussed. 7.3.1 Comparison of FRP membrane-based bridge systems As shown through the studies of system characteristics, the performance of both types of composite membrane-based bridges is improved by their in-plane behavior as tension structures after the integrated optimization. However, efficient performance under in- plane behavior is gained through different schemes for both composite membrane-based bridges. The system characteristic studies of the optimal CMB bridges showed that the structural stiffness used by CMB bridges to carry external loadings is obtained by developing a force couple between the longitudinal tension forces in the FRP membrane and the longitudinal compression in the concrete deck. The in-plane section resultants of 150 the FRP membrane and the concrete deck in the transverse direction are due primarily to the Poisson effect and are two orders of magnitude less than the longitudinal in-plane resultants. Therefore, CMB bridges achieve mainly unidirectional in-plane behavior. However, the system characteristic studies for the optimal CMS bridges indicated that the FRP membrane in optimal CMS bridges are subject to in-plane longitudinal and transverse tensions. The structural stiffness of CMS bridges is thus acquired by achieving the bi-directional membrane behavior of membrane and shell structures. The two types of composite membrane-based bridges presented in this work differ not only in their structural behavior, as previously discussed, but also in their efficiency in material use. As it is well known, efficient use of materials requires achieving not only in—plane or membrane resultants but also a uniform stress distribution throughout the structure. The distribution of stress demands in a generally anisotropic structure can be represented by strength safety factors of the elements in a finite element model of the structure. The strength safety factor of an element is evaluated by the stress index of that element with respect to a specific failure criterion. However, due to the different structural geometry, loading and material volume used for CMB and CMS bridge systems, their structural response is different even for the same limit state. Therefore, a concept of relative stress indices is introduced to evaluate and compare the efficient use of the FRP membrane. Thus, the stress indices of individual elements are normalized by the maximum stress index for the complete structure (clearly applied to each structure separately). Table 7—28 summarizes the distribution of normalizing stress indices defined by the Tsai-Wu failure theory for the FRP membranes of all membrane-based bridges under the service limit state load demands. The structural 151 fraction distributions, defined as the number of elements within a given relative index range divided by the total number of elements in the structure, for all three bridges with respect to the relative stress indices are compared in Figure 7—35. The average and the standard deviations of the three curves are given in Table 7—29. Table 7—28 Stress indices distribution of three bridges CMB CMS - Model A CMS — Model B Normalized Fraction Normalized Fraction Normalized Fraction Stress # EL Stress # EL Stress # EL Index 317;; Index :75}: Index ‘21:; 0.00% 0.00% 0.00% 0.00% 0.00% 0.00% 6.67% 0.00% 6.25% 26.73% 6.25% 22.16% 13.33% 0.33% 12.50% 28.52% 12.50% 30.73% 20.00% 1.31% 18.75% 20.41% 18.75% 22.47% 26.67% 2.47% 25.00% 10.70% 25.00% 13.69% 33.33% 6.78% 31.25% 5.83% 31.25% 4.27% 40.00% 19.75% 37.50% 2.27% 37.50% 2.97% 46.67% 24.08% 43.75% 1.70% 43.75% 1.77% 53.33% 19.31% 50.00% 1.94% 50.00% 1.14% 60.00% 14.14% 56.25% 1.22% 56.25% 0.31% 66.67% 3.81% 62.50% 0.19% 62.50% 0.09% 73.33% 3.33% 68.75% 0.06% 68.75% 0.03% 80.00% 2.42% 75.00% 0.13% 75.00% 0.06% 86.67% 1.67% 81.25% 0.00% 81.25% 0.00% 93.33% 0.56% 87.50% 0.06% 87.50% 0.06% 100.00% 0.06% 93.75% 0.00% 93.75% 0.19% 0.00% 100.00% 0.25% 100.00% 0.06% 152 35% ‘ + CMB + CMS-Model A 30% : +CMS-Model B 1L 25% N o o\° I .1 Structural Fraction 3 ,\° 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.( Normalized Strength Index Figure 7—35 Structure material distribution of relative stress index Table 7—29 Statistics of relative stress index distribution Relative Stress Index CMB CMS —- Model A CMS - Model B Average 50.07% 14.51% 17.77% Standard deviation 13.1 1% 10.43% 11.06% Evaluation of the data presented in Figure 7—35 and Table 7—29 indicates that a higher average of relative stress index implies a higher average stress level for that structure. Thus, CMS bridges have a higher strength reserve in terms of FRP laminate capacity than CMB bridges of the same structural size and material use when the structures are subjected to an equal loading. On the other hand, a smaller standard deviation of the stress index distribution indicates a large portion of the FRP membrane is located within a smaller range of the average stress index. From this point of view, CMS bridges have a better structural performance of uniform stress distribution than CMB 153 bridges. Therefore, it can be concluded that CMS bridges utilize the FRP material more efficiently than CMB bridges. 7.3.2 Discussion Of the objective function As proved in Section 4.1.2, structural stiffness is maximized by achieving in-plane structural response, which is implemented by minimizing the structural strain energy in the integrated shape and laminate optimization approach. On the other hand, structural stiffness can be also improved by adding material. Therefore, structural geometric dimensions and the thickness Of the FRP membrane were constrained tO limit the use Of laminated FRP composites. However, it is difficult to distinguish the individual contribution from structural shape and material quantity towards improved structural stiffness by minirm'zing the structural strain energy. For example, strain energy can be further decreased with increased FRP materials even though the structural response be already in-plane dominant. This limitation Of the objective function could result in an over-designed optimum as shown in the post optimality analyses for CMB bridges (Table 7—12 and Table 7—14). The results from Section 7 .1.4 showed that the stress indices Of the optimal CMB bridges were far below the allowable stress levels for all limit states. An over-bound constraint on the FRP membrane thickness is the most possible reason. Therefore, an integrated Optimization employing a narrow-bound constraint on the membrane thickness was investigated for the Optimum design Of CMB bridge. The effort Of a narrow bound constraint on the FRP membrane thickness was implemented by evaluating the optimal design for the CMB bridge with the upper bound on thickness reduced from 19.05 mm to 9.525 mm. The newly Optimized CMB bridge 154 (Model D) is summarized in Table 7—30. It can be noted that the FRP membrane for Model D achieved essentially the same Optimal geometric shape and the same Optimal laminate properties compared tO the previously Obtained Optimum designs (see Table 7—3). Also, it is noted that the FRP membrane thickness still reached the upper bound of the thickness constraint. Due tO the reduced thickness (50% reduction) the structural strain energy was almost doubled. However, this increase Of the strain energy does not indicate that the in-plane structural behavior was impaired since the FRP membrane still achieved a close Optimal geometric shape. In fact, the increase Of the strain energy is caused by the increase Of in-plane stress resultants due to the thickness reduction. Therefore, strain energy minimization with geometric constraints can not avoid over-designed Optimal structures although the structural stiffness is properly improved by achieving membrane response. Table 7—30 Optimal CMB bridge designs with different membrane thickness constraints Optimal design Model A Model B Model C Model D 3m“ Energy 83.85 83.85 83.85 161.90 mm {w W k W t} {1359.0, {1365.2, {1352.6, {1327.3, 1’ 2' 1’ 2’ 2184.4,9843, 2184.4, 1022.4, 2222.4, 997.0, 2174.2, 990.6, (mm) 121921905} 121921905} 121921905} 121929525} {081,031, {081,031, {081,031, {081,031, {VIA,V3A,V10,V3D} _1’ 1} -03, 07} 0.75, 1} 0.75, 1} According to the Optimization experience Obtained above, Optimal CMB bridges with an average Optimal geometry were investigated to study the structural response with respect to the FRP membrane thickness (Table 7—31). It can be Observed that, while the structure remained under in-plane demands, reductions in membrane thickness increased 155 the structural strain energy while the maximum deflection and stress index approached the prescribed design criteria values, which is also indicative Of a more efficient design. Table 7—31 Structural response with different membrane thickness Thickness Strain Energy Deflection Stress Index T (m) SE (kN-mm) D (m) SI (%) 19.050 84.51 1.692 34.56 15.875 100.33 1.966 41.16 12.700 127.33 2.464 51.13 9.525 165.07 3.226 66.15 6.350 235.12 4.765 95.17 Several alternative Objective functions that could seek to maximize structural stiffness and minimize material use were investigated: 0 Objective function f1: Minimize (Strain Energy)><(Thickness) 0 Objective function f2: Minimize (Strain Energy) / (Stress Index) 0 Objective function f3: Minimize (Strain Energy)><(Thickness) / (Stress Index) The performance Of these Objective functions evaluated with Optimal CMB bridge designs is summarized in Table 7—32 and graphically represented in Figure 7—36. It can be seen that the proposed Objective function (f3 = SE-T/SI) linearly decreases with reduction Of the membrane thickness. Conversely, as shown in Table 7-31, the decrease in the membrane thickness leads to stresses and deflections that are closer to the design criteria. Therefore, the Objective Of maximizing structural stiffness while minimizing the material can be achieved by implementing a function that incorporates minimization Of the structural strain energy and the membrane thickness as well as maximization of stress levels. 156 Table 7—32 Evaluations Of alternative Objective functions f1=SE~T f2=SElT f3=SE-T/SI (kN-mmz) (kN-mm) (kN-mmz) 1610.0 244.5 4658.5 1592.7 243.8 3869.6 1617.1 249.0 3162.8 1572.3 249.5 2376.9 1493.0 247.1 1568.8 5000 - +1, = SE-T (kN-mmz) ‘ +12: SEISI (kN-mm) : +f3=SE-T/SI(kN-mm2) 4000 “r c .2 j 2 3000 ~~ 3 LL 0 .2 8 2000 ~— 3 . o a 1000 a : I 4 f I J 0 r 1 . . r . 4 5 10 15 20 Membrane Thickness (mm) Figure 7—36 Performance Of different Objective functions for Optimal CMB bridge In addition, the Optimum designs for FRP composite membrane-based bridges are currently Obtained under an Optimization load case taking into account a single loading effect and a single loading pattern. However, as evaluated in the post optimality analyses of both bridges, the structural behavior Of the Optimum designs can be significantly under non-Optimization loading effects and patterns. Therefore, multiple loading effects and loading patterns should be considered simultaneously in a limit state in order tO reduce 157 significant variations in the response Of an Optimum design. This can be implemented by a general Optimization load case in which different loading effects and patterns are combined with different weight factors. For example, the extreme loading effects Of maximum torsion, maximum mid-span bending and the normal load effect might be considered with a uniform weight factor Of 1/3 in the Optimization load case for the Optimum design Of CMB bridges. Similarly, the Optimization load case for CMS bridges might combine the loading patterns Of uniform and spaced line loads with a weight factor Of 1/2. Instead Of using a general Optimization load case, multiple loading effects and patterns can be taken into account by using a multi-Objective Optimization approach. Multi-Objective Optimization consists Of defining different Objective functions assigned with weight factors with respect to their respective loading effects and patterns. The Optimum design that compromises among the different loading effects and patterns is Obtained by Optimizing the multi—Objective functions with respect to their own set Of constraints. 7.3.3 Selection Of the design variables It should be noted that the structural performance Of CMS bridges (Section 7.2) will vary with the position Of the pier supports. Therefore, the pier position should be taken into account as one of the geometric design variables in Optimizing CMS bridges. However, preliminary Optimization efforts revealed that the Objective function of minimizing the strain energy varied discretely with respect to changes Of the pier positions. This resulted in premature Optimization solutions at local Optimum points. 158 Thus, the pier positions Of CMS bridges were currently fixed at constant locations throughout the Optimization process. Moreover, it should be noted that, for all Of the Optimization examples presented in this dissertation, the geometric key points that determine the structural shape were selected tO achieve a higher computational efficiency. A structural shape is then generated by interpolating the geometric coordinates Of the key points. Therefore, the structural shapes that can be Obtained are limited by the number Of geometric key points and their degrees Of freedom. For example, the number Of geometric key points selected for shape finding Of FRP membranes in the bridge systems can only determine the surfaces high up to a 2nd-Order curvature by interpolation techniques. The accuracy Of Optimal shapes Obtained by structural shape Optimization for free form finding can be improved by increasing the number Of the geometric key points. However, free form finding is achieved at the cost of computational efforts due to the increase Of design variables. In addition, general free-forms, although efficient, may not be practically feasible. The determination Of the number Of design variables therefore requires considering the computational efforts, the accuracy Of the Optimal shape sought, and the ease Of implementation. Thus, in spite Of the power provided by these mathematical Optimization tools, it is still the design’s judgment that dictates the successful design Of an efficient structural form. 7.4 Summary In this chapter, the integrated shape and laminate Optimization approach was applied to the design and Optimization Of two innovative and newly proposed FRP composite 159 membrane-based bridge systems. The characteristic studies for both bridge systems show that the high in-plane stiffness and strength characteristics Of FRP laminates are utilized to maximize the performance and efficiency of the bridge systems due to the effectiveness provided by shape resistant membrane structures. Furthermore, based on the Observation Of over-designed Optimal CMB bridges, several alternative Objective functions, in addition to the minimization Of structural strain energy, were investigated to achieve in-plane structural response with minimum material use. The Objective function Of minimizing the strain energy and the membrane thickness while maximizing the membrane stress level seems tO have the best performance for CMB bridges and is thought tO be a viable Objective function for maximizing structural stiffness with minimum material use. The use Of Objective functions that take intO account multiple loading effects and patterns were discussed, and the importance behind the selection of geometric design variables was noted to place the current work and its results into context. 160 References AASHTO—American Association Of State Highway and Transportation Officials (1998). AASHTO LRFD bridge design specifications. Washington, DC: The Association. Azzi, V.D. and Tsai, SW. (1965). Anisotropic strength Of composites. Experimental mechanics: 5, 283-288. Hibbitt, Karlsson & Sorensen, Inc. (2001). ABAQUS standard user’s manual Version 6;. Jones, RM. (1999). Mechanics of composite materials. Philadelphia, PA: Taylor & Francis. Seible, F., Karbhari, V.M., and BurguefiO,R. (1999). Kings Stormwater and 1-5/Gilman Bridges. Structural Engineering International: Journal of the International Association for Bridge and Structural Engineering: 9, 4, 250-253. Tsai, SW. (1968). Strength theories of filamentary structures. In Fundamental Aspects Of Fibeer Reinforced Plastic Composites Schwartz, RT. and Schwartz H.S. (eds), 3-11. New York: Wiley Interscience. Wu, EM. (1974). Strength and Fracture Of composites. In Composite Materials, Broutman, LI. and Krock, R.H. (eds) Vol. 5, 191-247. New York: Academic Press. 161 8 Conclusions and Future Research Needs 8.1 Conclusions The presented integrated Optimization approach Offers a procedure for simultaneously finding the Optimal shape and Optimal laminate design Of FRP structures. The integrated approach was investigated by analytical studies on the development Of FRP shell structures that combine FRP laminates and shape resistant structures for improved FRP structural designs. The approach provides an efficient tOOl to determine maximum stiffness designs by Optimizing not only the geometry Of the membrane or shell but also the material properties and the composite laminate design. The integrated Optimization approach can also serve as an analytical tOOl to aid the rational implementation Of laminated FRP composites in civil infrastructure by developing innovative design concepts. The developed concepts were that Of membrane- based FRP composite bridge systems. The bridge systems consist Of a shape-and-material Optimized laminated FRP membranes that carry the in-plane tensile and shear forces while using a conventional reinforced concrete slab or FRP deck system to provide the live load transfer. The dual-level Optimization approach can thus support the analytical studies required for initial development Of innovative systems that use FRP composites in their inherent behavioral characteristics for new high-performance structures. During the efforts Of integrated shape and material Optimization, it was Observed that an over-bounded constraint Of material use can lead to an over-designed Optimum achieved for maximizing structural stiffness. Several feasible Objective functions using combination Of the structural strain energy, the membrane thickness, and stresses were investigated. The Objective function Of minimizing the strain energy and the membrane 162 thickness while maximizing the membrane stress level seems tO have the best performance and is thought tO be a viable Objective function for maximizing structural stiffness with minimum material use. The work has provided insight tO the concept that FRP laminates can be used with higher efficiency in new structural systems as long as their advantageous properties Of directional strength, light weight, and tailored properties are properly considered in the design process. 8.2 Future research needs 8.2.1 About the integrated approach Of shape and laminate Optimization As it is well known, shape-resistant structures, such as thin shells subject to compression forces, are sensitive to buckling. While not considered in the presented work, structural buckling performance can be easily included in the proposed integrated Optimization procedure. This can be done by two distinct approaches. One is to consider a determined buckling capacity as a constraint in the shape and laminate Optimization procedure, thus achieving the Optimization Objective Of maximum stiffness while satisfying the required buckling resistance. Another approach is to search for the maximum buckling capacity along with maximum stiffness in a multi-Objective Optimization process, where the different design Objectives can be weighted to maximize structural performance. Therefore, the presented integrated approach can further evolve into a multi-Objective decoupled approach for both shape and laminate optimization Of FRP lightweight structures by adding appropriate Objectives and constraint functions. 163 A gradient algorithm is used in the first step Of the integrated approach to achieve the Objective Of maximizing structural stiffness. Due to the expectation Of unaffordable computation demands by the large-scale optimizations, the numerical technique Of finite difference methods was employed to supply the required gradients information tO advance the gradient algorithm. The main disadvantage Of calculating gradients by finite difference methods is the uncertainty in the choice Of a perturbation step size. A perturbation that is either tOO large or tOO small will result in inaccurate gradient evaluations and misguide the search direction Of the algorithm. The techniques from analytical methods have a better performance, in terms Of robustness and accuracy, tO derive gradient information than finite difference methods. However, the differentiation Of Objective and constraint functions in analytical methods requires the structural stiffness matrix, which will change completely even for small shape changes. This makes gradient derivation computationally costly. Therefore, future reSearch is needed to provide the proposed integrated approach with an efficient and accurate technique for sensitivity analyses. In the current approach, the shape and laminate Optimization processes were decoupled into a two-step procedure, in which laminate designs in the second-step Optimization are based on the lamination parameters from the first step. This requires that feasible larrrination parameters be Obtained during the first-step optimization. Thus, the constrained relationship between the lamination parameters is very important and has to be solved for the first-step Optimization. At present, only the constraints between the essential lamination parameters, {V,A, V3A, V10, V30}, required by the current approach were numerically established (with certain assumptions) and applied in the Optimization 164 process. However, this might not be enough for general applications. A competitive approach that can avoid the formulation Of explicit constraints and implicitly satisfy the feasibility of laminate designs is to choose fiber orientations as material design variables. For this alternative approach, the integrated Optimization has to be implemented with genetic algorithms in which geometric and material design variables are both treated as discrete variables. Although the volume of computations needed for sensitivity analyses can be avoided by using genetic algorithms, the computational effort will probably not be reduced because Of the random search approach Of genetic algorithms. The formulation Of this alternative approach and its computational efficiency can maximize the broad implementation Of the developed method. 8.2.2 About applications Of the integrated approach Although not explored in the current research effort, the integrated approach can be applied for the Optimum design Of different lightweight structures seeking to maximize structural stiffness, such as domes and membrane roof systems. Furthermore, with appropriate formulation of Objective and constraint functions, the integrated optimization approach can be employed for many diverse applications. For example, a potential application is vibration tuning Of structures. Structures subject to dynamic forces or excitation at their supports might need to be designed to avoid resonance or other negative vibration related problems. By relating the Objective function to the fundamental vibration frequency, the integrated approach can yield Optimum designs Of structures that can avoid vibration problems. 165 The concept employed in the current approach to consider continuous and discrete design variables separately in a two decoupled procedure can also be applied to other Optimization applications, for example, for smart structures made using piezoelectric materials. Piezoelectric materials are the materials that can convert electric energy into mechanical energy and vice versa. They produce an electrical response when mechanically stressed (sensors) and high motion can be Obtained when an electric field is applied to them (actuators). Composite structures integrated with piezoelectric sensors and actuators Offer potential benefits in a wide range of engineering applications such as vibration suppression and shape control. The development and efficiency Of this type of structures rely on Optimizing the location of the actuators and sensors that control the behavior of adaptive structures. This type of optimization applications involves continuous design variables of structural geometric shapes and discrete design variables that define the position of actuators and sensors. Therefore, the current integrated approach can be directly applied to investigate this type Of Optimization problems. 8.2.3 About the engineering and construction Of Optimal FRP composite structures It needs to be noted that the fiber orientations Of the Optimal laminates achieved in the previous Optimization examples are based on the local structural coordinates of the finite elements describing the FRP membrane, which vary with the normal of the membrane surface. Therefore, the construction Of an FRP membrane with a laminate design in which the layout is specified based on local coordinates requires further study. In addition, the construction of bridges like those presented in this work requires providing an effective membrane/slab connection to assure composite behavior between 166 the FRP membrane and the concrete slab, or the FRP panel, Of the bridge system. Particularly for CMB bridges, shear force resistance is required at the membrane/slab interface under flexural actions. A study on possible forms Of shear connectors is thus needed to allow implementation for the developed CMB bridge systems. Finally, post Optimality studies on the response Of Optimal CMS bridges indicated that the distribution of in-plane stresses depends on the form and spacing Of diaphragms, which affect the efficiency of load transfer to the FRP membrane. A future study is thus required to investigate geometries and Optimal placement Of diaphragms for CMS bridges. 167 .2... L. ._ ..f ..; 1.1«N..... r 13.1.. .1