Mechanics and Design of Flexible Composite Fish Armor by MASSACHUSETTS INSTIt9TE OF TECHNOLOGY Ashley Browning JUN 28 2012 B.S., University of Texas at Austin (2008) LIBRARIES Submitted to the Department of Mechanical Engineering in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering ARCHNVES at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2012 o Massachusetts Institute of Technology 2012. All rights reserved. A uthor ........................... ,........... Department oMechanical Eriering May_2 , 2012 Certified by .......................... Ford Professor of Mary C. 4 oyce IE e Certified by ........................ Christine Ortiz Professor of Material Science and Engineering ThesiSpervp A ccepted by .......................... David Hardt Chairman, Department Committee on Graduate Students Mechanics and Design of Flexible Composite Fish Armor by Ashley Browning Submitted to the Department of Mechanical Engineering on May 25, 2012, in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering Abstract Inspired by the overlapping scales found on teleost fish, a new composite architecture explores the mechanics of materials to accommodate both flexibility and protection. These biological structures consist of overlapping mineralized plates embedded in a compliant tissue to form a natural flexible armor which protects underlying soft tissue and vital organs. Here, the functional performance of such armors is investigated, in which the composition, spatial arrangement, and morphometry of the scales provide locally tailored functionality. Fabricated macroscale prototypes and finite element based micromechanical models are employed to measure mechanical response to blunt and penetrating indentation loading. Deformation mechanisms of scale bending, scale rotation, tissue shear, and tissue constraint were found to govern the ability of the composite to protect the underlying substrate. These deformation mechanisms, the resistance to deformation, and the resulting energy absorption can all be tailored by structural parameters including architectural arrangement (angle of the scales, degree of scale overlap), composition (volume fraction of the scales), morphometry (aspect ration of the scales), and material properties (tissue modulus and scale modulus). In addition, this network of armor serves to distribute the load of a predatory attack over a large area to mitigate stress concentrations. Mechanical characterization of such layered, segmented structures is fundamental to developing design principles for engineered protective systems and composites. Thesis Supervisor: Mary C. Boyce Title: Ford Professor of Mechanical Engineering Thesis Supervisor: Christine Ortiz Title: Professor of Material Science and Engineering 3 4 Acknowledgments I would like to thank my advisors, Professor Mary C. Boyce and Professor Christine Ortiz for their endless support, wisdom, and enthusiasm. They have helped me grow and develop into the confident woman I am today. They are both excellent role models and have taught me the importance of professional success in the balance of life. Thanks to the all of the Boyce and Ortiz group members for your help and advice. Special thanks to Meredith Silberstein, Juha Song, and Shawn Chester for showing me the ropes and to Narges Kaynia, Ting Ting Chen, Swati Varshney, and Matt Connors for being an absolute pleasure to work with. I'm grateful for the technical assistance of Dr. Ara Nazarian at Beth Israel Deaconess Medical Center (BIDMC) Orthopedics Biomechanics Laboratory and Dr. Alan Schwartzman at the MIT Nanolab. I'd also like to acknowledge Mark Belanger at the MIT Edgerton Shop, Bill Buckley at the Laboratory for Manufacturing and Productivity (LMP) Shop, and Ken Stone at the MIT Hobby Shop for doing things faster, safer, and more easily than I thought was possible since before I was born. None of this would have been possible without the support of my friends, family, and fiance. Thanks to Adam and Jason for visiting me in this arctic tundra and Jaclyn for being the sister I never had. Thanks Mom and Dad for the weekly reminders to graduate, in case I somehow forgot. Thanks James for his patience and support of my career, in spite of the challenges it created. Thanks JetBlue flight 1264 for getting him here every other week and Beyonce for reminding him to put a ring on it. I'm so grateful that Joe and Jenna Petrzelka have been my rock and source of coffee since day one and that Tiffy and Mark spice up my ramen noodle diet with molecular gastronomy. And thank you, litter, for coming into my life and making everything so much more enjoyable. This research was supported by the U.S. Army Research Office through the MIT Institute for Soldier Nanotechnologies under contract W911NF-07-D-0004. 5 6 Contents 15 1 Introduction 2 1.1 Bioinspired flexible protective systems . . . . . . . . . . . . . . . . . 15 1.2 Fish scales as a structural biological material . . . . . . . . . . . . . . 17 1.3 O verview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 2.1 2.2 3 19 Background Geometry of elasmoid fish scale assembly . . . . . . . . . . . . . . . 2.1.1 Fish integument overlap and arrangement . . . . . . . . . . . 19 2.1.2 Ex-vivo characterization with pCT . . . . . . . . . . . . . . . 22 2.1.3 Model formulation . . . . . . . . . . . . . . . . . . . . . . . . 24 Fish scale and tissue materials properties . . . . . . . . . . . . . . . . 26 29 Mechanical protection under blunt loading 3.1 3.2 Materials and methods . . . . . . . . . . . . . . . . . . . . . . . . . . 29 3.1.1 Prototype fabrication and mechanical testing . . . . . . . . . . 29 3.1.2 Finite element micromechanical model . . . . . . . . . . . . . 33 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 Infinitely long (periodic) structures . . . . . . . . . . . . . . . . . . . 48 . . . . . . . . 48 Experimental and simulated finite-sized structures 3.3 3.4 19 3.3.1 Deformation mechanisms in periodic structures 3.3.2 Energy absorption . . . . . . . . . . . . . . . . . . . . . . . . 60 3.3.3 Effect of scale morphometry . . . . . . . . . . . . . . . . . . . 64 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 7 4 Mechanical protection under predatory tooth attack 69 4.1 Finite element micromechanical model . . . . . . . . . . . . . . . . . 70 4.2 Indentation simulations results . . . . . . . . . . . . . . . . . . . . . . 71 4.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 5 Bending and flexibility 6 79 5.1 Materials and methods . . . . . . . . . . . . . . . . . . . . . . . . . . 79 5.2 R esults . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 Conclusions 87 6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 6.2 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 6.3 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 A Materials characterization of the scales of Arapaima gigas 91 A. 1 Elemental analysis with EDX and BSEM . . . . . . . . . . . . . . . . 91 A.2 Mineralization analysis with TGA . . . . . . . . . . . . . . . . . . . . 94 A.3 Nanoindentation 94 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . B DIC and load synchronization 99 C Periodic boundary conditions 103 C.1 General mathematical formulation for periodicity in two directions C.2 Case I: Fixed corner node with periodicity in two directions . . . . . 103 105 C.3 Case II: Fixed corner node, roller surface, with periodicity in one direction 109 C.4 Abaqus input file . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 C.5 MATLAB code to generate Abaqus input file . . . . . . . . . . . . . . 115 8 List of Figures . . . . 16 . . . . . . . . . . . . 20 2-2 Scale overlap for the common rudd . . . . . . . . . . . . . . . . . . . 21 2-3 Scale overlap for the family of Cyprinidae fish . . . . . . . . . . . . . 21 . . . . . . . . . . . . . . 23 2-5 pCT slices of salmon integument across body length . . . . . . . . . . 24 2-6 Overview of fish assembly and geometric variables . . . . . . . . . . . 25 2-7 Map of scale volume fraction 0, Ra = 50) . . . . . . . . . . . . . 26 2-8 Materials properties of fish scales and tissue . . . . . . . . . . . . . . 27 3-1 3-D printed prototypes . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3-2 Compression testing of silicone rubber . . . . . . . . . . . . . . . . . 31 3-3 Tensile testing of ABS dogbones . . . . . . . . . . . . . . . . . . . . . 32 3-4 Selected experimental geometries for testing . . . . . . . . . . . . . . 33 3-5 Experimental setup on Zwick . . . . . . . . . . . . . . . . . . . . . . 34 3-6 Undeformed finite-sized geometries . . . . . . . . . . . . . . . . . . . 35 3-7 Undeformed geometries for periodic simulations . . . . . . . . . . . . 36 3-8 Definitions of energy and stiffness. . . . . . . . . . . . . . . . . . . . . 37 3-9 Definitions of deformation mechanisms. . . . . . . . . . . . . . . . . . 38 1-1 Fish armor and its human analogues. 2-1 Histology of fish integument in different species 2-4 Sample prep for pCT of salmon integument #(Kd, 3-10 Experimental deformation under blunt loading for select geometries. 40 3-11 Experimental deformation under blunt loading for all geometries. 41 3-12 Simulations of select finite-sized geometries under blunt loading. . 3-13 Experimental cracking and delimitation . . . . . . . . . . . . . . . . . 9 44 45 3-14 Stiffness and back deflection for experiments . . . . . . . . . . 46 3-15 Stiffness and energy absorption for finite-sized simulations . . 47 3-16 Spring model of layered composite. . . . . . . . . . . . . . . . 48 3-17 Periodic simulations for select geometries under blunt loading 49 3-18 Loading curves for all periodic simulations under blunt loading 50 3-19 Deformation mechanisms, stiffness, and energy absorption 52 3-20 Simulated contours of u2 for blunt loading . . . . . . . . . . . . . . . 53 3-21 Simulated contours of 622 for blunt loading . . . . . . . . . . . . . 54 3-22 Simulated contours of 612 for blunt loading . . . . . . . . . . . . . 55 3-23 Simulated contours of ol for blunt loading . . . . . . . . . . . . . . . 56 3-24 Simulated contours of o 22 for blunt loading . . . . . . . . . . . . . . . 57 3-25 All deformation mechanisms for blunt loading . . . . . . . . . . . . 58 3-26 Key deformation mechanisms for blunt loading . . . . . . . . . . . . 59 3-27 Deformation resistance vs. energy absorption for blunt loading . . . . 62 3-28 Maps of energy absorption for blunt loading . . . . . . . . . . . . . . 63 3-29 Effect of scale morphometry . . . . . . . . . . . . . . . . . . . . . . . 65 3-30 Deformation mechanisms for blunt loading and changing Ra 66 . . . . . 4-1 Predatory attack by a tooth ................. 4-2 Indentation loading conditions . . . . . . . . . . . . . . . . . . . 71 4-3 Contours of indentation of fish scale assembly . . . . . . . . . . 73 4-4 Indentation force and back deflection for R/t, = 0.1 . . . . . . . 74 4-5 Indentation loading curves for varying indenters (R/t = 0.1, 10) 75 4-6 Deformation resistance vs. energy absorption for indentation . . 76 5-1 Experimental setup for bending over die. . . . . . . . . . . . . . 80 5-2 Experimental deformation under bending for all geometries. 81 6-1 Summary of deformation mechanisms under blunt loading. 6-2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 A-1 Overview and SEM of Arapaima gigas scales . . . . . . . . . . . 92 10 ...... 69 . 88 A-2 Materials characterization of Arapaima gigas scales by EDX and BSEM 93 A-3 Mineralization of Arapaima gigas scales using TGA . . . . . . . . . . 95 . . . . . . . . 96 . . . . . . . . . . . . . . . 97 A-4 Location for nanoindentation of Arapaima gigas scales A-5 Nanoindentation of Arapaima gigas scales C-1 Representative volume element for periodic boundary conditions. . . . 104 C-2 Node sets for periodic boundary conditions. . . . . . . . . . . . . . . 105 C-3 Periodic boundary conditions used in model. . . . . . . . . . . . . . . 109 11 12 List of Tables 2.1 Control parameters for pCT scan . . . . . . . . . . . . . . . . . . . . 22 2.2 Description of parameters and variables used in study . . . . . . . . . 28 3.1 Volume fractions studied where Ra = 50. . . . . . . . 30 3.2 Material properties used in compression simulations . . . . . . . . . . 37 3.3 Normalized areal density PA for 0, Ra = 50) where p, ~ pt. . . . 60 3.4 Normalized areal density PA for #(Kd, #(Kd, 0, Ra = 50) where p8 = 2 .6pt. 61 3.5 Aspect ratio study parameters. . . . . . . . . . . . . . . . . . . . . . 65 4.1 Material properties and geometry used in indentation simulations. . . 70 A. 1 Thermal profile for thermogravimetric analysis . . . . . . . . . . . . . 94 # for geometries 13 14 Chapter 1 Introduction 1.1 Bioinspired flexible protective systems Natural protective systems may hold the key to the design of new structural materials that are flexible and lightweight while still affording mechanical protection. These biological materials have the potential to inspire rapid advances in the design and engineering of new structures for a variety of applications, from microsystems to body armor to vehicle and building architecture. In order to realize these designs, a full understanding of the structure-function relationships and underlying mechanics of such systems is essential. Biological exoskeletons have evolved over many millions of years to maximize survivability by incorporating various structural design principles (Ortiz and Boyce, 2008). In the case of ancient fish, dermal armor became thinner, lighter, and subdivided into small plates to accommodate threats without compromising mobility (Colbert and Morales, 2001; Arciszewski and Cornell, 2006). Today, most species of bony fish (e.g. zebrafish Danio rerio, Arapaima gigas) possess a network of flexible yet stiff and strong scales that provide a lightweight defense against predators (Fig. 1-1). These scales are angled, overlapped, and embedded in the tissue to provide spatially homogeneous protective coverage and flexibility that can be tailored across the body. Unlike the articulating exterior scales on creatures such as snakes, lizards, pangolins, or butterfly wings, the teleost fish scales are mineralized and deeply embedded within the dermis of the integument (Whitear, 1986). 15 Embedding scales fully within the tissue creates a unique robust composite system in which the interplay between the mineralized scales and the compliant tissue plays a nontrivial role in tailoring and controlling deformation mechanisms. A careful balance of mobility and protection is an important goal in animal evolution, just as it is in the engineering of human armor (Arciszewski and Cornell, 2006), and hence imbricated fish scales are an appealing biomimetic system. (a) (b) (c) Figure 1-1: Fish armor and its human analogues. (a) Teleost fish Atlantic salmon (Salmo salar) is covered in scales. A cross section (below) shows that scales are overlapping and embedded in a tissue. (b) Backplate of an early 16th century brigandine. Brigandines were used as body armor and composed of small iron plates riveted to a canvas or linen fabric backing. (c) Dragonskin is modern segmented armor developed by Pinnacle Armor, consisting of ceramic discs encased in an aramid textile cover. Images adapted from Ostman, Elisabeth, Iduns kokbok, 1911, Last accessed on May 1,2012. http: //commons. wikimedia. org/wiki/Salmosalar; Evans (1994); Mitsuhiro Arita, Exoroche on Arms and Armor, Last accessed on May 1, 2012. http: //wiki.ffxiclopedia.org/wiki/Exoroche-onArmsandArmor; Daderot, Brigandine backplate, probably Italy, early 16th century - Higgins Armory Museum, Last accessed on May 1, 2012. http://commons.wikimedia.org/wiki/Category: PiecesofarmorintheHigginsArmoryMuseum; Ling Li, April 27, 2012; Pinnacle Armor, Dragon Skin, Last accessed on May 1, 2012. http: //www .pinnaclearmor. com/body-armor/dragon-skin/; Wikipedia, Dragon Skin, Last accessed on May 1, 2012. http://en.wikipedia.org/wiki/DragonSkin; Neal and Bain (2004). Modern elasmoid scales evolved from the ganoid and rhomboid scales of primitive fish and are now nearly ubiquitous in present day. Compared to their ancient counterparts, elasmoid scales are thinner, lighter, and more flexible, having eliminated bone and reduced hypermineralized components (Sire, 1990; Sire et al., 2009). Fish scales 16 vary in both morphology (Burdak, 1979; Ibafiez et al., 2009) and mechanical properties (Garrano et al., 2012) across the surface of the body to accommodate inherent body curvature and performance requirements, e.g., increased biomechanical flexibility near the tail for locomotion, and on the main body for increased protection. Such structure-function relationships have evolutionary implications for swimming mode (Breder, 1926; Lindsey, 1979), body stiffness and flexural response (Long et al., 1996; Hebrank and Hebrank, 1986; Hebrank, 1980), hydrodynamics (Whitear, 1986), and defensive armor (Burdak, 1979; Vermeij, 1983) in which species have adapted their integument in response to environmental stimuli to maximize survivability. 1.2 Fish scales as a structural biological material The elasmoid fish scale has recently been investigated as a viable structural biological material. Material structure and properties vary by species, but elasmoid scales have been reported to be mineralized 16-59%, slightly anisotropic, and have hydrationdependent material properties (Whitear, 1986; Torres et al., 2008; Zhu et al., 2012; Garrano et al., 2012). They are composed of a bony mineralized outer layer and a compliant basal collagen layer (Sire, 1990; Sire et al., 2009). Experimental measurements of mechanical properties have been limited to the mechanics of an individual scale resisting penetration by a sharp indenter or uniaxial tensile testing. Testing of scales in a hydrated state found the tensile strength to range 22-65 MPa and toughness ~1 MPa (Meyers et al., 2011; Zhu et al., 2012; Ikoma et al., 2003; Torres et al., 2008; Garrano et al., 2012). The multilayered fish scale localizes penetration by promoting circumferential (Bruet et al., 2008; Song et al., 2011) or cross-like (Zhu et al., 2012) fracture patterns in the outer layer. On a larger length scale, work by Vernerey and Barthelat (2010) introduced a two-dimensional mathematical model of overlapping fish scales in accommodating larger length scale bending and radius of curvature as a function of scale geometry and material properties. They find a strainstiffening behavior of the assembly with increasing scale overlap provides the ability to distribute deformation over a large area during bending. However, the mechan17 ics of embedded overlapping scales in response to blunt and predatory penetration loading remains unexplored. The ability of fish scales to function as a dermal armor and provide protection via minimizing back deflection and compressive stresses in underlying tissues remains to be explored. In addition, a composite model is needed to fully understand the interplay of the mineralized scales with a hyperelastic tissue as a means of tailoring deformation mechanisms. 1.3 Overview Here, the architectural structure of the overlapping scales embedded in a soft tissue is investigated as a key determinant in mechanical protection to advance understanding of the mechanical design principles of flexible protective biological exoskeletons. The composite system developed herein considers fish scales fully embedded in a tissue matrix, more closely emulating physiological conditions. The connections amongst the composite structure, the deformation mechanisms, and the macroscopic response to a variety of physiological loading conditions are explored. In particular, the roles of tissue shear and scale bending and rotation as a function of scale geometry are identified. To accomplish these objectives, a finite element based micromechanical model was developed in which protective response was determined as a function of scale aspect ratio, scale orientation angle, spatial overlap, and volume fraction of scales. Macroscale synthetic models were constructed using a combination of three-dimensional printing and molding methods; mechanical behavior under blunt compressive loading was measured using mechanical testing where digital image correlation (DIC) was used to measure the strain field in the composite material. These experimental and numerical models yield detailed insights into the roles and the tradeoffs of the composite structure providing constraint, shear, and bending mechanisms to impart protection and flexibility. 18 Chapter 2 Background 2.1 2.1.1 Geometry of elasmoid fish scale assembly Fish integument overlap and arrangement To simplify the complex geometry of a fish, this study focuses only on the integument (i.e. the epidermis, dermis, and embedded scales). A superficial examination of the surface of a teleost fish reveals articulating scales; however, histological sections (Whitear, 1986; Park and Lee, 1988; Hawkes, 1974; Sire, 1990; United States Fish and Wildlife Service, 2010) further reveal that the scales are fully surrounded by soft tissue comprised by the dermis and epidermis and also embedded deeply into the soft tissue at an angle (Fig. 2-1). The primary features of the structure of mineralized scales embedded in a soft tissue are captured in a micromechanical model consisting of (1) an upper layer of relatively high modulus (E,) scale of aspect ratio L8 /t, oriented at angle 0 and embedded in a low modulus (Et) tissue, and (2) a soft (low modulus Et) tissue substrate support representing the underlying collagenous layer. The scales are initially oriented at angle 0 relative to a compliant tissue support. Scales overlap each other such that a fraction of the scale is embedded beneath an adjacent scale. The length d is the exposed length of scale, while the remaining length (L, - d) is embedded beneath another scale. This degree of imbrication, or overlap, Kd - d/L, is a measure for the spatial overlap defined by Burdak (1979), 19 (a)%% _ _ -__ME 400 (b) V 100 pm (d) (d) Figure 2-1: Histology of integument in different species of teleost fish. Scales are shown to be embedded in a dermal tissue at varying angles 0 and overlap Kd. Adapted figure showing section of integument of (a) trout (United States Fish and Wildlife Service, 2010), (b) mangrove killifish (Park and Lee, 1988), (c) unknown (Guinan), (d) cichlid (Sire, 1990), (e) coho salmon (Hawkes, 1974), and (f) unknown (Smith, 2012). who measured Kd for 16 species of the Cyprinidae family of fish and found that values range from 0.24 < Kd 1 depending on species and body location (Fig. 2-2, 2-3). Fish with highly overlapping, heavily armored skins such as the tench (Tinca tinca), have Kd ~ 0, whereas lightly scaled fish have Kd ~ 1, as is the case for undulating eel-like fish. The number of scales per unit length is given by Ki-. The overlap along the circumference was determined to be constant for all species studied by Burdak (1979). Typical values for the scale thickness in teleost fish are on the order of t, ~ 0.1 20 (a) (b) (c) 1(d Figure 2-2: Scale overlap Kd along the length of the common rudd (Scardinius erythrophthalmus). Image adapted from Burdak (1979). Kd along length of fish alburnus alburnus alburnus rutilus rutilus rutilus blicca bjoerkna bjoerkna leuciscus cephalus cephalus scardinius erythrophthalmus chondrostoma nasus variabile carassius auratus gibelio carassius carassius cyprinus carpio carpio leuciscus idus idus tinca tinca gobio gobio gobio misgurnus fossifis cobitis taenia taenia nemachilus barbatulus barbatulus 1 0.8 cc 0.6 a) 0 0.4 0.2 increasing fish size 0 3 Kd = d/L region of fish Figure 2-3: Scale overlap Kd along the length body for the family of Cyprinidae fish. Image adapted from Burdak (1979). 21 mm, but reported values range from 44 pum to 1 mm, (Whitenack et al., 2010; Meyers et al., 2011). Typical scale lengths, LS, range from 5 to 15 mm, but have been reported up to 100 mm long (Francis, 1990; Garrano et al., 2012; Zhu et al., 2012; Meyers et al., 2011; Torres et al., 2008; Jawad and Al-Jufaili, 2007). Despite the large range of observed scale lengths and thicknesses, the geometry of the fish scale preserves an aspect ratio Ra = L8 /t, between 25 to 100 based on available literature and measurements. The effect of various geometric parameters will be studied (as discussed later) but the scale will be held at constant aspect ratio of either Ra = 50 or 100, which were taken as intermediate and maximum values, respectively. 2.1.2 Ex-vivo characterization with CT The aspect ratio of scales has been observed to vary not only from species to species, but also across the length of an individual fish. Micro-computed tomography (iCT) was employed to obtain full three-dimensional characterization of the integument of a fresh Atlantic salmon (Salmo salar) obtained from a local fishery (New Deal Fish Market, Cambridge, MA). Samples were excised from the main body and caudal peduncle (tail) near the lateral line and immediately scanned to reduce the effects of dehydration (Fig. 2-4). Scans were performed with a pCT40 (Scanco Medical AG, Switzerland) using the parameters in Table 2.1 and analyzed following previously reported methods (Song et al., 2010; Connors et al., 2012). Mimics (Materialise, Belgium) and Adobe Photoshop were used to process the DICOM slices and threshold the partially mineralized scales. The ttCT slices reveal the aspect ratio, overlap, and angle of the scales of the salmon vary as a function of location along the body shown in Fig. 2-5. Table 2.1: Control parameters for puCT scan E (energy) I (current) FOV (tube 0) voxel (resolution) 45 kVp 177 pA 36.9 mm (largest) 18 Pm (high) 22 (a) Figure 2-4: Sample prep for pCT of fish integument. (a,b) Scales of Atlantic salmon (Salmo salar) are - 10 mm long and transparent. (c) After underlying muscle is removed, resulting tissue and scale composite is flexible. (d,e) Tissue is cut into ~ 25 x 25 mm sections and (f) scanned in puCT tube between layers of foam. Salmon image adapted from Ostman, Elisabeth, Iduns kokbok, 1911, Last accessed on May 1, 2012. http: //commons. wikimedia.org/wiki/Salmosalar. 23 2 mm Figure 2-5: (a) Teleost fish Salmo salar,the Atlantic salmon, covered with (b) integument of overlapping stiff scales. (c) Two-dimensional pCT slices of structure (false coloring), showing scales embedded in a dermal tissue at various geometric configurations. Salmon image adapted from Ostman, Elisabeth, Iduns kokbok, 1911, Last accessed on May 1, 2012. http://commons.wikimedia.org/wiki/Salmo.salar. 2.1.3 Model formulation Relevant geometric parameters and structures are generalized in a model geometry shown in Fig. 2-6. This non-dimensional formulation allows for length-independent analysis of overlapping composite assemblies across a range of length scales. All parameters and variables in model are shown in Table 2.2. In most teleost fish species, the upper epidermis follows the contours of the scale pockets (Whitear, 1986), and hence, the irregularity of the surface topology increases as a function of Kd and 0. A variable of scale volume fraction, #, can be calculated in terms of the thickness of the dermis between scales td, and scale thickness to, as # = ts/(ts + td). This fraction # represents the scale-to-tissue ratio in the upper, scaled layer and does not include the tissue support, as this bottom layer serves only to capture physiologically relevant back deflection in the underlying integument. This bottom tissue layer is fixed at height Htissue = Ls/2. Here, the scale volume fraction is reported as #(Kd, 0, Ra) as shown in Fig. 2-7, whereas previous models, neglecting a composite system, only consider a linear # ~ 1/Kd. This study will vary geometric parameters as governed by the following relationship: tan2 0 +1 KdRatan0 24 k I 9 CliJ Figure 2-6: Overview of fish assembly assembly and variables. This non-dimensional formulation allows for length-independent analysis of overlapping composite assemblies across a range of length scales. An approximate biologically relevant range of #(Kd, 0, Ra) is shown in Fig. 2- 7 based on available histological sections and puCT scans (Whitear, 1986; Park and Lee,.1988; Hawkes, 1974; Sire, 1990; United States Fish and Wildlife Service, 2010; Guinan). To compare relative weights of architectures given tissue density pt and scale density ps, the approximate areal density PA (mass per unit area) of the composite prototypes is given by PA - ptL2 + L, sin[p# -+pt(1 - @)] fixed in study function of architecture The areal density of the structure PA is normalized by the areal density of the bottom layer alone (fixed), highlighting the additional weight of the scale layer. PA = 1 + 2 sin0[(ps/pt)4 + 1 - $)] Assuming similarity of the scale and tissue densities (p8 ~ pt), this weight is primarily 25 Volume Fraction <p 40 __ _ _00__ _ __ _ __ _ .0e_ 30 OC 20 10 0 0.5 00 1- Kd (scale overlap) 0 0.2 0.4 0.6 0.8 1 Figure 2-7: Map of scale volume fraction q(Kd, 0, Ra = 50) indicating approximate biological range based off histological sections and measurements in Figs. 2-1 and 2-5. Corresponding geometry from model shown on right. a function of scale angle: PA - PA(O), as discussed later in Section 3.3.2. 2.2 Fish scale and tissue materials properties Fish scales are composed of a cross-ply collagen layer and an outer mineralized hydroxyapatite layer as detailed in Appendix A (Sire, 1990; Sire et al., 2009) and, for the purpose of this work, can be approximated as an elastic material. Tensile testing of fish scales reports an effective scale elastic modulus is ~ 1 GPa but varies among species and along the length of the body of the fish (Fig. 2-8). Testing of scales in a hydrated state (more closely resembling in vivo conditions) yields a reduced elastic modulus 0.1-0.8 GPa and tensile strength to range 22-65 MPa (Torres et al., 2008; Zhu et al., 2012; Garrano et al., 2012; Hebrank and Hebrank, 1986; Meyers et al., 2011; Ikoma et al., 2003). The elastic properties of fish scales are in the range of common engineering polymers (e.g., polycarbonate, acrylonitrile butadiene styrene), as shown in Fig. 2-8. The density of fish scales (ps) is approximated using results obtained from thermogravimetric analysis (see Appendix A.2) and the relative densities of hydroxyapatite (mineralized) and collagen (organic). The density of fish scales 26 ranges 1.92-2.43 g/cc based on available literature and measurements (Ikoma et al., 2003; Torres et al., 2008). Little work has been done to characterize the tissue in the integument of fish. Work by Hebrank and Hebrank (1986) reported tensile testing of fish tissue samples (both with and without scales) to determine an approximate modulus. For comparison, these results are shown along with other biological tissues in Fig. 2-8. Many biological tissues exhibit hyperelastic, nearly incompressible materials properties similar to rubber. In general, the tissue is orders of magnitude more compliant than the embedded scales. For reference, the density of fish tissue (pt) is approximated as an intermediate value between human adipose tissue (p = 0.92 g/cc) and human muscle (p = 1.06 g/cc) density (Heymsfield et al., 2002). As an upper bound, the scales are approximately 2.6 x as dense as the surrounding tissue. (a) scale elastic modulus, E 3D printed ABS extruded ABS polymer fish scale1 (wet or 0 1 GlPa 2GPa 0 ) (b) tissue shear modulus, G 10 kPa 100 kPa 1 MPa 10OMPa Hebrank, 1986; Meyers et al., 2011; Ikoma et al., 2003). 27 100OMPa 1GPa Table 2.2: Description of parameters and variables used in study Parameter Description Ls ts scale length scale thickness dermal tissue thickness epidermal tissue thickness exposed scale length td te d Htissue Hscale total height scale overlap scale angle scale volume fraction scale modulus tissue modulus scale density tissue density normalized areal density PA = t,(1/ te - Property geometric geometric geometric geometric geometric 1) ts Htissue Ls/2 Hcae ~ L, sin 0 Htotal = Hcae + Htissue Kd = d/L scale layer height Kd Es Et Ps Pt td tissue support height Htotai 0 Formula # PA 28 tan =1 + 2 sin 9[-Pt# 10- #)] geometric geometric geometric spatial spatial composition material material material material material & composition Chapter 3 Mechanical protection under blunt loading 3.1 3.1.1 Materials and methods Prototype fabrication and mechanical testing Macroscale synthetic prototypes were fabricated and tested to characterize a row of scales over a range of geometric architectures. A computer aided drawing package (SolidWorks, Dassault Systemes) was used to create models of the scales in a variety of geometric configurations. Scale aspect ratio was kept constant at Ra - 50 and scale volume fraction <5 was varied from 0 to 1 by parametrically varying overlap Kd (0.1, 0.3, 0.5, 0.7, 0.9) and scale angle 0 (2.50, 50, 10', 200, 40') for a total of 21 arrangements (Table 3.1). These models were exported as stereolithography (.STL) files for three-dimensional printing using fused deposition modeling (Dimension BST 1200es, Stratasys). Parts were built with thermopolymer acrylonitrile butadiene styrene (ABS) with a layer thickness of 0.010 inches and a breakaway support. The resulting parts were placed in an acrylic frame and used as a mold for a translucent silicone rubber (Mold Max 10T, Smooth-On). After the rubber cured, the ABS scales were permanently embedded in the silicone and the frame was removed, creating the prototype parts shown in 29 Table 3.1: Volume fractions # # 0.1 2.50 50 - bZ 100 - d 200 400 0.62 0.41 for geometries studied where Ra = 50. overlap Kd 0.3 0.5 0.7 0.77 0.39 0.21 0.14 0.92 0.46 0.23 0.12 0.08 0.66 0.33 0.17 0.09 0.06 0.9 0.51 0.26 0.13 0.07 0.05 (d) Figure 3-1: (a) Fused deposition modeling (FDM) 3-D printing was used to obtain macroscopic prototypes for a variety of architectures. ABS parts were placed in an acrylic frame, creating a mold. (b) Silicone rubber was cured in the mold, embedding the rigid ABS scales in the prototypes, which easily broke away from the PLA support material. (c-d) Resulting fabricated samples with a U.S. quarter for reference. Fig. 3-1. The final macroscale models had scale geometry LS = 50 mm, t' = 1 mm, were 100 x 40 x -50 mm in size, and contained 2-6 repeating scale units depending 30 on angle 0 and overlap Kd. Material properties of the silicone rubber and ABS were measured Eo,t = 390 kPa and E, = 880 MPa using compressive and tensile testing, respectively (Figs. 3-2, 3-3). These properties are in the range of biological materials for wet elasmoid scales (Lin et al., 2011; Torres et al., 2008; Garrano et al., 2012) and fish tissue (Hebrank and Hebrank, 1986), as shown in Fig. 2-8. The ABS to silicone rubber modulus ratio is -2000. H =25mm W =40 mm L = 100 mm A = L*WF 0.2 expAFE FF uniaxial compression CU ~0 6 Eo =0.39 MPa H0 plane strain W L 0 0 e=6/H o=F/A 0.1 e from DIC 0.2 Figure 3-2: Elastic modulus of the silicone rubber was evaluated using uniaxial compression testing of rectangular samples. A rigid frame was used (see Fig. 3-5) to impose experimental plane strain conditions and results were verified with 2-D model in Abaqus. Plane strain compression testing of the synthetic prototypes was performed (Zwick Z2.5kN, Zwick/Roel) at a loading rate of 0.1 mm/s to a maximum force of 400 N. Each sample was tested three times and repositioned between each test to estimate measurement errors. Error bars in the experimental data correspond to the standard deviation of these results. Macroscopic engineering stress o- = F2 /A is calculated from the measured vertical reaction force, F2 , and the initial total sample cross-sectional (surface) area, A. To provide plane strain conditions for finite thickness geometry, a frame was constructed with thick clear acrylic (PMMA) sheets, constraining the sample in the 3-direction, which was precisely adjusted to account for slightly varying sample thicknesses (Fig. 3-5). Chalk powder and Teflon sheets were used to reduce friction on the sides and top/bottom of the rubber, respectively. A speckle pattern was applied on the samples with an airbrush and India ink for use with optical ex31 (a) V (b) I (d) (e) Zwick Data 80 a-e from DIC 10 8 60 6 40 0- 4 20 0 0 2 0 0.1 0.2 0.3 0.4 gage elongation (mm) 0 ' 0.005 0.01 8 0.015 0.02 Figure 3-3: Elastic modulus of the 3-D printed ABS was evaluated using tensile testing of laser cut dogbone samples. Data from the digital extensometer is synchronized to mechanical testing data to obtain engineering stress-strain curves as detailed in Appendix B. tensometry via digital image correlation (DIC). Images of the testing were taken at a rate of 1 fps (VicSnap, Correlated Solutions) and DIC (Vic-2D 2009, Correlated Solutions) was used to track global deformation and obtain local contours of logarithmic (Hencky) strain. Engineering strain E is calculated as the total vertical displacement from DIC, 6, normalized by the initial total height, Htotai. Shear strain is reported as engineering shear strain. Data from the digital extensometer was then synchronized to mechanical testing data to obtain engineering stress-strain curves as detailed in Appendix B. 32 (a) (b) Volume Fraction < 40 30 (c) C 20 10 0 0 0.5 Kd (scale overlap) 0 0.5 (dl 1 Figure 3-4: (a) Map of scale volume fraction # and selected geometries for Ra = 50. Macroscale 3-D printed and molded synthetic fish scale inspired assembly of ABS (yellow) and silicone rubber (translucent): (b) Kd = 0.3, 0 = 200, # = 0.21, (c) Kd = 0.3, 0 = 5, # = 0.77, (d) Kd = 0.9, 0 = 50, #S 0.26. Selected geometries have similar scale overlap (b, c), angle (c, d), and volume fraction (b, d) to demonstrate the functional effect of each structural parameter. 3.1.2 Finite element micromechanical model A two-dimensional plane strain composite scale-tissue micromechanical model was simulated using finite element analysis (FEA). The composite structure was discretized and modeled with quadratic continuum elements in ABAQUS/Standard (library types CPE6H and CPE8RH) and a fine mesh with elements of length - t,/4 was needed to capture the deformation within the scales. Tissue and scale elements are modeled as perfectly bonded with properties found in Table 3.2. A blunt loading event was conducted in the form of a distributed compression of the scale assembly with a rigid plate. Non-linear, large deformation theory with frictionless contact was assumed. Simulations which consider the same finite length geometries and the same loading conditions as the experimental samples, were conducted and allow for direct comparison of model with experimental results. These geometries correspond to the archi33 Figure 3-5: Experimental setup on Zwick with plane strain acrylic and aluminum frame, Teflon and chalk powder for frictionless boundaries, and pre-loading (rubber bands) to secure the sample against the base. The load from the platens is transferred via a polycarbonate plate the width of the sample. Plastic shims (here, yellow and brown) were used to precisely adjust the width of the aluminum frame to account for varying sample thicknesses. tectures #(Kd, 0, Ra = 50) found in Table 3.1 as and shown in Fig. 3-6. To simulate structures with many scales, such as those found on the entire length of a fish, periodic boundary conditions were also imposed on the lateral edges of the scale assembly while allowing overall expansion in the 1-direction, as detailed in Appendix C. These geometries are the same architectures #(Kd, 6, Ra = 50) found in Table 3.1 and have varying length proportional to LS cos 0 as shown in Fig. 3-7. Note that the representative volume element (RVE) used is 3x as long as necessary, but used to better visualize the deformations in the structures. Further simulations were carried out changing the aspect ratio of the scales, Ra = L/t,, from 50 to 100 by increasing the length of the scales L, while holding scale volume fraction results in overlap Kd half of that shown in Table 3.1. 34 # and angle 0 fixed. This Kd = 0.3 Kd = 0.1 K, - 0.7 Kd = 0.5 Kd = 0.9 0 LO $ = 0.92,1PA= 1.07 (1.22) 0.66, A 1.07 (1.18) $P= 0.51, p5A = 1.08 (1.16) a) 0 LO 11 $0.46, j5= 1. 15 (1.30) 4 = 0.33, AA= 1.16 (1.27) $= 0.23, PA= 1.33 (1.48) $P 0.17, PA= 1 .3 3 (1.44) P = 0.13, 5A= 1. 7(1.76) 1.65 (1.91) $P =0.12,1A= 1.66 (1.82) $ =0.09, p= 1.67 (1.78) $=0.07, PA =1.67 (1.76) 2.25 (2.56) (P ='0.08, PA = 2.26 (2.45)' CD $ =0.77, PA= 1.14 (1.39) 0 $= 0.39 PA= 1. 3 1 (1.56)2 I0.26, 5= 1.16 (1.25) 0 I 11 S= 0.62, = 1.58 (2.37) $P 0.21, PA CID =O.41, p = 2.15 (0.14, $= 0.05, PA= 2 .2 7 (2.38) Figure 3-6: Undeformed finite-sized geometries #(Kd, 0, Ra = 50) used to simulate experiments. Areal density PA is given for experimental and silicone rubber and ABS and biological tissue and scales in parentheses. .. .. .. .... .... I.......... Kd = 0.3 Kd = 0.1 Kd = 0.5 K, = 0.7 Kd =0.9 0 LO Ci II CD $= 0.21A 1.07 (1."22) $= 0.66 TA= 1.07 (1.18) 4)=0.51, I, 1.08 (1.16) 4) =046 1.15 (1.30) 4) 0.33, 1& 1.16 (1 27) P = 0.26,15A 1.16 (1.25) 0 LO $ = 0.77, 15A 14 (1.39) 11 0 0 4= 0.39, CD A= 1. 3 1 (1.56) t= 0.23, 5A1.33 (1.48) 4) = 0.17, jY = 1.33 (1.44) 0 0 C\J CID $ I=0.6 , PTA= .58 (0.37) PA $. =021, 91 (1 *165 4)=0.12, FA*= 1.66 (.82) *.07,5A= *.7(.6 0 0D 041, pTA .1 (3.1 2) $0K.08,15A=2.26 (2 5) Figure 3-7: Undeformed geometries #(Kd, 0, R, = 50) with periodic boundary conditions used to simulate very long structures. Areal density PA is given for experimental and silicone rubber and ABS and biological tissue and scales in parentheses. These structures have varying length proportional to L, cos 0. Note that the representative volume element (RVE) used is 3 x as long as necessary, but used to better visualize the deformations in the structures. Table 3.2: Material parameters experimentally determined from synthetic prototype and employed in simulations. Scale Tissue Material Behavior Modulus 3-D printed ABS Elastic Es = 880 MPa Silicone rubber Neo-Hookean Eo,t = 0.39 MPa Poisson's ratio Density v, = 0.4 Ps = 0.80 g/cc vt - 0.49 pt = 1.07 g/cc Macroscopic engineering stress and strain were calculated in the same manner as done experimentally (stress o = F2 /A and strain c = 6 /Hota). Local true Cauchy stress and logarithmic Hencky strain contours were evaluated at different load levels to identify the deformation mechanisms which accommodate the loading for the different composite geometries. An effective secant stiffness Elo-=.o5 MPa = a/E was determined (Fig. 3-8) and normalized by tissue elastic modulus Et, highlighting the stiffness amplification effect of the scales. Strain energy density U o0.05 MPa = o Ed was determined and normalized by the energy absorbed by the tissue at a same macroscopic stress, Ut. E 0* Figure 3-8: Definitions of energy E and stiffness U taken at a*. For this work, both E and U were evaluated at a* = 0.05 MPa. Angular rotation dO of individual scales in response to loading was calculated 37 with respect to initial angle: dO = 9' - 9 (Fig. 3-9). For small bending within an individual scale, the local deformed curvature is described by K = a2y'/Ox'2 in a rotated coordinate frame (1', 2') aligned with original scale angle 9. The average bending curvature along a scale of deformed length s is given by ks = fLS Kds and found by extracting the displacements ui and u 2 of nodes along the scale in the mesh. (a) LC 0 0 0 0 0 0 0 0 0 11 (b) 1 K(S) 21 Figure 3-9: Definitions of deformation mechanisms under blunt loading with periodic boundary conditions. (a) Undeformed configuration and boundary conditions for Kd = 0.5, 9 = 5', # = 0.46, PA = 1.33 and (b) deformed configuration exhibiting local scale curvature K(s), rotation dO = 0' - 9, and shear strain E12 in tissue. 3.2 Experimental and simulated finite-sized structures The geometric arrangement of the scale and tissue structure governs the response to blunt loading, resulting in a substantial dependence of back deflection (6 tissue, refer to Fig. 2-6) and stiffness (E) on geometric arrangement of the assembly. Macroscopic loading curves for the finite-sized geometries are shown in Fig. 3-10a for three selected geometries (all with scale aspect ratio Ra = 50): (i) Kd = 0.3, 9 = 200, # = 0.21 (ii) Kd = 0.3, 9 = 50, # = 0.77 38 (iii) Kd= 0.9, 0 = 50, # = 0.26 These particular geometries are highlighted because they demonstrate the functional effect of each structural parameter individually by sharing similar scale overlap Kd (i, ii), scale angle 0 (ii, iii), or scale volume fraction # (i, iii). The experimen- tal and simulated loading curves for the same finite-length geometries show that all structures studied have an initial nonlinear response due to contact conditions from the irregular loading surface. It was also observed that the bilayer structure of the finite-sized composite allowed the bottom tissue layer of the samples to laterally expand significantly (seen in Fig. 3-12). This edge effect has a strong influence on several aspects of the results of the simulations, as discussed later in Section 3.3.1 where the influence of finite-sized specimens are compared to infinitely long periodic structures. When evaluated at a fixed macroscopic strain, the effect of scale geometry on the maximum back deflection in the underlying tissue was found to be modest. As shown in Fig. 3-10b, increasing Kd from 0.9 to 0.3 (a -3x increase in #) offered a <10% decrease in back deflection tissue at any fixed strain. However, the effective stiffness of the composites is strongly dependent on the scale arrangement, and hence, at a fixed macroscopic stress, the same ~3x increase in # offers a 30-40% decrease in back deflection (Fig. 3-10c). Given that predatory attacks are often force-limited, these results show how geometric scale configurations and compositions observed in nature can protect underlying tissue by limiting tissue straining as expected. By tailoring the structural parameters, species are able to control the local protective function of their armors. 39 (a) 0.1 0.08 (c) 0.12 * Kd=0.3, 0=5*, III * 0.08 $=0.77, PA114 E- 0.06 0 (b) 0.12 6=20*, P=0.21, P,= 1 .6 5 Kd=0.3, 0.08 K,=0.9, 0=5*, 0.04 P=026, PA= 0.04 0.02 0 0.00 0 0.05 0.1 0 0.05 U2 (mm) 0.00 0 0 -0.15 0.05 a (MPa) 0.1 -12 -22 -2.5 (d) 0. 1 0 -0.15 -0. 15 1 K-=0.3, 0=20*, $)=0.21, PA= .65 (e) (f) K =0.9 9=50 <=0_26 5 =1.16 Figure 3-10: Effect of blunt loading on ABS and silicone rubber prototypes under plane strain compression for three selected geometries with Ra = 50. (a) Corresponding experimental and simulation compression stress-strain curves, and back deflection as a function of (b) strain and (c) macroscopic stress. Error bars represent standard deviation of n = 3 experiments and solid lines represent simulations of finite-sized geometry. (d-i) Selected experimental contours of displacement and strain are shown at macroscopic stress o = 0.05 MPa and indicate deformation within the composite structures. 40 undeformed s=0.05 6=0. 10 e12 at e=0.10 10.15 -0.15m Figure 3-11: Undeformed geometry and experimental deformation under blunt loading at E= 0.05, 0.1 and contours of shear strain evaluated at c = 0.1. 41 e1 undeformed e=O.10 s=0.05 -0.15 at e=O.10 10.15 U I I (continued) Undeformed geometry and experimental deformation under blunt loading at e = 0.05, 0.1 and contours of shear strain evaluated at c = 0.1. 42 e1 at e=0.10 undeformed e=0. 10 &=0.05 -0.15 iUKWJO.15 (continued) Undeformed geometry and experimental deformation under blunt loading at E= 0.05, 0.1 and contours of shear strain evaluated at E= 0.1. 43 u2(mm) e22 0 -2.5 12 0 -0.15 -0.15 0.15 (a) (b) Figure 3-12: Simulated contours of displacement and strain are shown at macroscopic stress o = 0.05 MPa corresponding to experiments of blunt loading of physical models in Fig. 3-10. (a) Kd = 0.3, 0 = 200, corresponding to # = 0.21, PA =1.65 (b) Kd = 0.3, 0 = 50, # = 0.77, PA = 1.14 and (c) Kd = 0.9, 0 = 5', 4 = 0.26, PA = 1.16. Experimental contours of displacement and strain agree well (Figs. 3-10, 3-12) with simulated results of the same geometry, demonstrating the quantitative capabilities of the model and providing insights into the deformation mechanisms. Depending on the geometric arrangements and volume fractions of the scales, macroscopic loading is accommodated by either (1) global tissue shear between scales and scale rotation, (2) local tissue shear at scale tips and scale bending, or (3) constraint of the tissue and overall stiffening. These deformation mechanisms are illustrated using the three selected geometries in Figs. 3-10, 3-12. Tissue shear and scale rotation. As shown in the contours of Fig. 3-10d for Kd = 0.3, 0 = 20', the anisotropy from the angled scales couples the compressive loading to a shear strain 612 within the tissue. For similar architectures with high scale overlap (low Kd) and initial angle 0, this region of shear strain is broad and evenly distributed over the length of the scales. This mechanism of shear between scales promotes uniform scale rotation, as shown in Figs. 3-10d, 3-12a. Here, the scales deform minimally as they rotate relative to the tissue; deformation is accommodated through tissue shearing. This effect was largely observed for structures with 0 > 200. Experimental results capture this effect well with the exception of geometries with 44 # ~ 1, in which the imaging pattern could not resolve the very narrow tissue layer between scales (Fig. 3-10e). Scale bending. Upon blunt loading, it can be seen in Fig. 3-10f that the scales in the arrangement with Kd = 0.9, 0 = 50 have noticeably deformed from their original configuration (Fig. 3-4). Here, the scales locally bend to accommodate the loading, resulting in highly localized shear strain in the regions of scale overlap. Such low overlap results in undesirable localized regions of strain in the tissue substrate, shown in the contours of 622 and 612. Due to the small feature size and non-linear deformation of the scales during bending, digital image correlation (DIC) was not able to fully capture the motion of the scales. It can also be seen for other architectures with Kd > 0.5 that the scales with low overlap underwent significant local bending, often constrained by three or more points of inflection. At high load (F > 500 N), the scales in same samples were observed to crack and subsequently delaminate from the rubber tissue (Fig. 3-13). delamination delamination (a) (b) Figure 3-13: Cracking and delimitation observed experimentally under high loads (F > 500 N) for (a) Kd = 0.7, 0 = 400, # = 0.06, (b) Kd = 0.7, 0 = 20', # = 0.09. Tissue constraint. The scale layers imposed a significant constraint on the underlying tissue, constraining lateral expansion. This constraint is apparent in the deformed images of Fig. 3-12, which show the constrained bulging of the underlying 45 tissue in the finite strained samples. This constraint effect will be shown to provide significant stiffening and hence dramatic reduction in back deflection of the overall composites when evaluating the periodic structure. (b) 0.10 (a) 4 experiment FE (periodic) 0.08 experiment FE (periodic) O 0.06 0.04 io U) O O I 1 I0.02 01$' 1 0.5 scale fraction $ 00 0 0.5 scale fraction $p 1 Figure 3-14: Experimental and finite length simulation results for 21 total geometries with Ra = 50. (a) Stiffness and (b) back deflection in response to blunt loading are evaluated at a = 0.05 MPa. Loading curves from experiments on physical prototypes agree well (Fig. 3-10a) with simulations of the same geometry and display the behavior of the composite structure. A secant stiffness was extracted from loading curves at a fixed macroscopic stress of a = 0.05 MPa (corresponding to e = 0.02 - 0.17) and was found to range from E = 1 to 3 across the geometries studied (Fig. 3-14a). This secant stiffness captures the effective "springs in series"-like stiffness of the tissue support layer and the scaled layer (Fig. 3-16). Assemblies with a low volume fraction of scales had E ~ Et; indirectly, these low volume fraction scale assemblies do not limit back deflection. Increasing # from 0 to 0.25 promotes a nearly linear increase in E from 1 to 2 and further increasing # gives a plateau of E = 3 at show the effect of increasing # on limiting back deflection and suggest a plateau of influence at #= # = 0.5. These results 0.5. Effective stiffness for all geometries is shown in Fig. 3-15. While effective stiffness of the composite is one indicator of the protective response of the system, it does not explicitly describe the effects experienced in the underly46 Kd = 0.1 Kd = 0.3 Kd = 0.5 K, = 0.7 K = 0.9 0 LO C\i 4)= 0.9 2 , pA = 1.07 (1.22) 11 CD E=2.82 U=0.49 -E/p=2.64-Ulp=0.4a $=O.66, =1.07 (1.18) E=2.74 U=0.65 E./p=5BUp.=1L60. $=0.51, p =1.08 (1.16) E=2.46 U=0.70 E4p=22B___U4L=065 = 0.46, E=2.49 F /p-9&i. $ = 0.33, Fp= 1.16 (1.27) U=0.94 E=2.05 U/p=081 $=0.26, A= 1.16 (1.25) E=2.08 U=O.91 $=0.17, Y =1.33 (1.44) E=1.66 U=1.11 U/p=0.83 E/p=1.25 $ =0.13, p= 1.34 (1.42) U=1.33 U/p=0.99 $ 1067 76) U-. U/ =Q-57 0 4)= 0.77, E=2.90 CD F/p=2.54 =1.14 (1.39) U=0.47 Up=~041 i = 1.15 (1.30) U=0.72 UI/p-6AF/p=177 C a) $ =0.39, A= 1.31 (1.56) E-2.52 U=0.65 E/p=1.92 U/p=0.50 $=0.23, PA= 1.33 (1.48) E=2.10 U=0.78 E/p=1.58 U/p=0.59 1.65 191) p0.21,> U=0.78 E/p=123 Ulo-p.47 $=0.12 1 .82) E=1.'61 = 0 82 FE/p~l 011 PQ65 U/p=0.7 F/p=1.8 E=1.48 E/p=1.11 0 CJ I E./p16 E=2.4WA Elp=1.79 IU=3P $ =U=0.,59 U/p= .38 L E=2.02 E=7 E/p=0.73 1.67 1.78) U=.50 U/p=0.50 .07 E=1.64 E/P=0.64 0 II CD $ = 0.41, A = 2.15(3.12) E=2.05 E/p=0.95 U=0.84 $= 0.14, E=1.64 JU/p=3 E/ = 2.25 (2.56) U=1.08 p0.8 $ = 0.08, jA= 2.26 (2.45) E=1.39 U=1.21 /p=06 9I /p=n. $ = 0.06, p = 2.27 (2.40) E=0.97 U=0.67 E/p03U/p=0.3i 4)= 0.05, E=1.08 E/p=0.47 A = 2.27 (2.38) U=1.60 U/p=0.71 Figure 3-15: Resistance to deformation (E, E/PA) and the resulting energy absorption (U, U/PA) for each geometry under blunt loading for the finite-sized structure simulated for experimental geometries. Each structure is characterized by #(Kd, 0, Ra = 50) with areal density PA given for experimental and silicone rubber and ABS and biological tissue and scales in parentheses. ing tissue. To function as an effective dermal armor, fish scales must minimize the deflections, strains and/or stresses beneath the scale layer to protect the body. Experimental results show that back deflection in the underlying tissue, 6 tissue, decreases as scale volume fraction increases for 4 < 0.5 and then remains at 6 tissue ~ 1 mm (ts) for #$> 0.5 (Fig. 3-14b). F F &scale Utissue 6total - kscale (variable) kscaie kti ektotal = E ktissue (constant) 6scale 6total-6tissue ktotai = (1/kscaie+1/ktissueY Figure 3-16: "Springs in series" model of tissue support layer and the scaled layer is captured by effective secant stiffness E. 3.3 3.3.1 Infinitely long (periodic) structures Deformation mechanisms in periodic structures Since the actual structures consist of many repeating scales along the length of the fish body, simulations with periodic boundary conditions which capture an infinitely long structure provide a more physiologically relevant assessment of the mechanical response than the simulations of the finite-sized specimens. Recall the finite-sized specimens exhibited significant lateral expansion (bulging) of the bottom soft tissue substrate (Fig. 3-12) due to the unconstrained lateral edges. The relaxed boundary conditions in these finite-sized specimens causes the structures to exhibit a more compliant response. In an infinitely periodic structure, the lateral strain of both layers 48 must be identical based on compatibility. The implications of this constraint have a dramatic impact on tissue deformation for geometries with stiff upper scale-tissue layers, in which the expansion of the lower layer is constrained by the presence of and compatibility with the scale layer. To eliminate these boundary effects, a periodic structure was simulated to more accurately model a larger number of repeating scales along the length, such as those found along the length of a fish body. 0.1 0.08 'C 0.06 C. o 0.04 (a) Kd=0.3, 0=20-,P=0.21, A=1.65 Kd= 0.3, 6=5', $p=0.77, PA11 Kd=0.9, 6=5*, $=0.26, yA= 1 . 1 6 0.02 0' 0 -2.5 U2 0.05 e22 i 0 -0.15 0.1 e E,2 1 0 -0.15 in o, (MPa) 0.15 -2.5 a22 (MPa) 2.5 -0.1 0.i (b) (c) (d) Figure 3-17: Periodic simulations for Ra = 50 yield insights into deformation mechanisms of periodic scale assembly under blunt loading. (a) Loading curves demonstrate stiffening constraint for structures with high #. Contours are evaluated at o- = 0.05 MPa. (b) Large initial scale 0 promotes rotation of scales via uniform tissue shearing (612). (c) High scale volume fraction directly results in a stiff composite, which reduces back deflection in the underlying tissue u 2. (d) Geometries with little scale overlap (high Kd) exhibit excessive scale bending (ou) and localized high stress (o 22 ) and strain in the underlying tissue. The resulting loading curves and contours for the periodic structure are shown in 49 Figs. 3-17 through 3-24, which demonstrate the large difference the constraint from periodicity has on the results for geometries with high #. Additionally, all simulated loading curves can be seen on a single plot in Fig. 3-18, where it can be seen that the non-flat loading surface of several geometries result in a softer overall response than the tissue (rubber) alone. Deformation mechanisms (Jk, dO), resistance to deformation (E, E/PA), and the resulting energy absorption (U, U/PA) can be shown for each geometry in Fig. 3-19. tissue scales 0.1 irregular loading surface 0.05"void oid" 0-- 0 0.05 0.1 0.15 Figure 3-18: Loading curves for all periodic simulations under blunt loading. The non-flat loading surface of several geometries with scale volume fraction # ~ 0 (e.g. bottom right structure) result in a softer overall response than the tissue (rubber) alone due to "voids". Future work should normalized the response of each structure to a similar structure compose of tissue (rubber) only to highlight the stiffening response of the scales. Imposing periodic boundary conditions has little effect for small volume fraction (# < 0.2), but as # increases, the effective stiffness of the periodic scale structure increases by an order of magnitude over the finite length structures (Fig. 3-26a), which had permitted lateral deformation (bulging) of the underlying tissue. The effective stiffness of the periodic structure follows a nearly linear relationship with scale fraction # (Fig. 3-26a) and shows the direct effect that increasing scale armor has on limiting back deflection at a given macroscopic stress level (Fig. 3-26b). This relationship between back deflection and scale volume fraction demonstrates that the 50 volume fraction of scales governs the constraint of the tissue layer and in turn, has a significant influence on protection. Therefore, effective stiffness and back deflection can be adjusted by tailoring the scale fraction #(Kd, One aspect of the constraint effect of increased 0). # on effective stiffness is demon- strated in the geometry of Fig. 3-17c, which has the same overlap as Fig. 3-17b but a significantly higher scale volume fraction, which was achieved by reducing the angle of the scales from 200 to 5', resulting in a stiffer loading curve and far less penetration in the underlying tissue (U 2 , 6 tissue) at a given force. In this case, by reconfiguring the angle of the same number of scales per unit length (Kd-1), the system can achieve a 4x increase in penetration resistance. These results show how the geometric arrangement of the scales and the scale volume fraction, together with their influence on constraining the deformation of the underlying tissue, serve to provide protection. Furthermore, such scale-imposed tissue constraint could serve as a design principle for these composite structures to locally tailor stiffness. The underlying deformation mechanisms for the structures with periodic boundary conditions were also identified. All mechanisms (scale bending, scale rotation, and tissue shear) are shown as a function of architecture (Kd, 0) and composition (#) in Fig. 3-25. Similar to results from finite-length geometry, scale bending, scale rotation, and tissue shear were found to be the primary modes of deformation. The mechanisms of scale bending and rotation are dependent on initial overlap (Kd) and angle (0) (Fig. 3-26). Architectures with low overlap (high Kd - 1) are dominated by scale bending, while those with high initial angle 0 are dominated by tissue shearing and corresponding scale rotation. This behavior is demonstrated by the high values of contours of bending stress (oll) found within the minimally-overlapping scales in Fig. 3-17d for a macroscopic stress of or = 0.05 MPa. Scales in such configurations at this load have an effective curvature of , > 0.25LS 1, or equivalently, an effective bending radius of R < 4L,. Conversely, the large initial angle 0 in Fig. 3-17b results 51 0 LO) $ = 0.92, PA=1.07 (1.22) U=0.17 E=9.78 K=0.08 dG=0.02* E/p=9.17 U/p=O.1 CD Kd = 0.9 K, = 0.7 Kd = 0.5 Kd = 0.3 Kd = 0.1 $ =0.66, PA = 1 07 (1 .18) K=0.17 E=6.60 U=0.28 d8=0.45* E/p=6.15 U/p=0.26 0 LO 4P= 0.46, QA= 1.15(13) U=0.36 K=0.11 E=4.71 U=0 19 E=8.18 E/p=7.17 U/p- .17 d8=0.25* E/p=4.08 U/p=0.31 (P= 0-77, PA 11 CD K=0.02 d6=0.0* 1.14 (1.3 ) 4= 0.33, pA = 1.16 (1.27) K=0.20 E=3.36 U=0.54 d=0.840 E/p=2.90 U/p=0.46 4 = 0.51, PA = 1.08 (1.16) U=0.45 E=4.28 K=0.19 0 dO=-0.03 E/p=3.97 U/p=0.42 =.26, P = 1.16 (1.25) E=2.51 U=0.81 dO=0.52* E/p=2.16 U/p=0.70 K=0.26 0 0 39 $ = 0. , PA= 1.31 (1.56) U=0.47 E=3.77 K=0.05 0 dB=0.35 E/p=2.87 U/p=0.36 a) = 0.62, o K=0.0 od8=-0.5 I=045 CD o t U/ $p= 0.137, p, = 1 .33 (1.44) K=0.26 E=2.01 d8=1.62* E/p=1.51 U=0.89 U/p=1.67 1 42 $= 0.13, p= 1.34 ( . ) U=1.37 E=1.36 K=0.31 d=1.87* E/p=1.02 U/p=1.02 1.58 (2.37 3.66 =2.31 $ = 0.07, pA= 1.67 (1. 7 8 ) K=0.05 E=1.09 U=1 .58 $ =10.21, p= 1.6 5 (1 91) U0.78 E=2.21 U/ =0.2 K=0.1 4 U/ =0.47 d8=1.2)* E/p=1.3 $=0.41, K=0.1 dO=3. CD $= 0.23, P= 1.33 (1.48) U=0.65 E=2.68 K=0.16 0 8=0.90 E/p=2.03 U/p=0.49 dO=2.450 E/p=0.65 U/p=0.95 .15 (3.12, =2.10 p=0.97 8 $=0. 4,p =2. 25( 2 K=0.20 __=4.110 E=1.65 E/p=0.74 56 ) U=-1.04 U/ =0.47 0.08,P =2.26(2.4) U=1. 5 K.3=20 E=1.36 d644.98* E/p=0.60 U/p=0 55 Figure 3-19: Deformation mechanisms ('-, dO), resistance to deformation (E, E/PA), and the resulting energy absorption (U, U/PA) for each geometry. Each structure is characterized by <(Kd,0, Ra = 50) with areal density PA given for experimental and silicone rubber and ABS and biological tissue and scales in parentheses. K= = 0.5 Kd = 0.1 LO C\J -2.5 U2 Kd = 0.9 0 0n I0 LO a) 0 0 CD 'I a) Q1 0 a) 3 Kd= 0. Kd= 0.1 0.1 U Kd= 0.5 Kd=0. 9 Kd= 0. 7 -=2.5' - - -0=5* cd a . 20 . - 0U 0 0.05 0.1 '0 0.05 0.1 0.1 0.05 =200 8=400 0.1 Figure 3-20: Loading curves and contours of u2 for blunt loading of architectures with periodic boundary conditions for Ra = 50 evaluated at o = 0.05 MPa. Kd = 0.9 Kd = 0.5 Kd = 0.1 - 22 0 -0.15 CD LO OD 0 0) 0 CDi Kd= 0. 3 Kd= 0.1 as IIa) Kd= 0. 7 Kd= 0.5 Kd=0.9 E=2.5* - - - 0=50 0=10* a. 20 -- 0 II 0 0.05 C 0.1 0.05 PE 0.1 ~0 0.05 C 0.1 0=20* 0=400 0.1 Figure 3-21: Loading curves and contours of E22 for blunt loading of architectures with periodic boundary conditions for Ra = 50 evaluated at o = 0.05 MPa. - - K 0= 0.5 Kd = 0.1 0' c'O II Kd = 0.9 812 -0.15 0.15 0 LO a) 0 C0 C14 11 CD C-" V-I 0 0 Kd= 0.1 Kd= 0.3 Kd= 0. 7 Kd= 0.5 Kd= 0.9 0.1 Figure 3-22: Loading curves and contours of e2 for blunt loading of architectures with periodic boundary conditions for Ra = 50 evaluated at o = 0.05 MPa. .. ...... .. .... .... .... ...... .. .... . ............ .... - K= = 0.5 Kd = 0.1 Kd = 0.9 0 LO al -2.5 (MPa) i 2.5 CD 0 LO CD 0 C\D 0 .0I a) Kd= 0.1 0.1 Kd= 0. 7 Kd= 0.3 -0.1 Kd= 0. 9 9=2.5 - - - 8=5* -0=10' - 20 .05 0=20 9=40* 0 0.05 C 0.1 0 0 0.05 C 0.1 0 0.05 C 0.1 0.05 C 0.1 0.05 r- 0.1 Figure 3-23: Loading curves and contours of oa for blunt loading of architectures with periodic boundary conditions for Ra = 50 evaluated at -= 0.05 MPa. Kd = 0.1 K LO CD -0.1 022 ino = 0.5 Kd = 0.9 (MPa) 0.1 0 a) 0 C). I CD Q1 0 CL Kd= 0.1 Kd= 0. 3 Kd= 0.7 Kd= 0.5 0.1 ~0 0.05 0.1 0.05 Kd= 0. 9 0.1 Figure 3-24: Loading curves and contours of o 22 for blunt loading of architectures with periodic boundary conditions for Ra = 50 evaluated at -= 0.05 MPa. ................. in a shear straining of the tissue, which in turn, leads to the rotation of the (E12) scales. Structures with both high Kd and 0 (corresponding to #< 0.1) exhibit both significant scale bending and scale rotation, but such configurations are typically not observed in nature. 0 K 6 0.01 vs overlap Kd: E a C SHEARING ROTATION BENDING 0 0 00 7-. 0.5 o = 2.50 0.6 M * 4 0 S2 .005 S21 1 '4 6=5 0.4 -' o= 10' o 40' A) CM 0 W 00 1 1 Kd (scale overlap) Kd (scale overlap) Kd (scale overlap) 0.5 0 vs volume fraction $s:s E S6 .2 4 0 0.01 E .005 ?0.4 0 A~ 0.2 okiP;, C W 0. 0 0.5 scale fraction (P 0 scale fraction $p 4 0.5 scale fraction $p 06 0.01 vs angle60: E = 2.5* 0 5 0 = 10 0.6 Kd = 0.1 -c 0.6 E E ad .005 - 0 04 CM Ne 0 0 40 20 0 (scale angle, *) 0 40 20 0 (scale angle, *) C Ka = 0.3 0.4 9 0.2 Kd = 0.5 0- Kd =0.9 K = 0. 0 (scale angle, *) Figure 3-25: All deformation mechanisms for uniform compression evaluated at a macroscopic stress of a = 0.05 MPa for 21 total geometries with Ra = 50. Scale bending P, is primarily a function of scale overlap Kd, scale rotation dO is primarily a function of scale angle 0, and average inter-tissue shear strain (Ce) is only a mechanism to promote scale rotation. 58 (a) 10 (b) 0.10 experiment FE (finite length) 8 FE (periodic) O 0.081 0 1W an a) 0 NC 4 0 0)( O9 0 0.04 II 0.02 F 1 0.5 0 01 10 0.5 1 scale fraction $p scale fraction $ (c) 0.01 0.008 0 0 2 ' 00 0 0 0 0.06 6 U, experiment FE (finite length) FE (periodic) (d) 6 e=40* 0 E 0.006 C .4 0 ~0.004 *0 2 2.50 0.002 f 01 0 0.5 scale overlap Kd 1 0 40 20 0 (scale angle, 0) Figure 3-26: Deformation mechanisms for uniform compression evaluated at a macroscopic stress of o- = 0.05 MPa for 21 total geometries with Ra = 50. (a) For low scale volume fractions (0 < 0.25) in experiments and finite length simulations, effective loading stiffness E of the composite increases linearly with scale fraction then plateaus for # > 0.5. Similarly, back deflection in the underlying tissue, tissue, decreases linearly for # < 0.5. Periodic simulations suggest that this plateau for large # is a boundary effect. For periodic simulations, (c) microscopic scale bending is primarily a function of overlap Kd, where as (d) scale rotation is dependent on initial scale 0. 59 3.3.2 Energy absorption To determine the ideal armor architecture under blunt loading, the deformation resistance of the composite must be evaluated against competing mechanical functionalities. Assuming similarity of the scale and tissue densities (PS ~ Pt) in this study for the experimental samples of silicone rubber and ABS, this weight is primarily a function of scale angle: PA - PA(0), as shown in Table 3.3. Thus to minimize the weight of the armor while stiff maximizing its protection against blunt loading via increased stiffness, the composite should have both a minimum scale angle 0 and high scale density #. Given that the stiffness of structures (E) and their energy absorption (U) are inversely related when evaluated at a fixed load, it is problematic to select a structure that combines both high stiffness and high energy absorption. Thus, if both properties of high E and U are desired, an intermediate value of # must be selected to balance these two criteria, as shown in Fig. 3-27b and Fig. 3-28a. Furthermore, it can be seen in Fig. 3-27a that the stiffness (in blue) exponentially decreases with increasing weight, while the energy absorption (in red) nearly linearly increases with weight. Hence, it is advantageous to select a structure where PA < 1.5 (equivalently, O < 100, after which increasing weight has diminishing returns in terms of stiffness and energy absorption. Table 3.3: Normalized areal density PA for #(Kd, 0, Ra = 50) where p, ~ pt. overlap Kd 0.7 0.5 0.3 0.9 PA 0.1 2.50 - - 1.06 1.07 1.08 50 - 1.14 1.15 1.16 1.16 0 100 - Cd 200 400 1.58 2.17 1.31 1.65 2.24 1.33 1.66 2.26 1.33 1.67 2.27 1.34 1.67 2.27 < Alternatively, if the density of the mineralized scales is greater than that of the 60 tissue, as often the case in biological systems, the resulting weight is a function of both scale volume fraction <5 and angle 0. For physiologically relevant biological materials of tissue and fish scale, an approximate density ratio is p, = 2 .6pt where PA is shown in Table 3.4. Energy absorbed and energy absorbed per unit weight for this density ratio is shown in Fig. 3-27b and Fig. 3-28b. As the density of the scales increases, E the intersection of the and U curves is increasing on the lower end of the weight spectrum, suggesting that added weight doesn't necessarily imply added protection. Table 3.4: Normalized areal density PA for < b- <;(Kd, 0, Ra = 50) where ps = 2.6pt. overlap Kd 0.7 0.5 0.3 0.9 PA 0.1 2.50 50 - 1.39 1.22 1.30 1.18 1.27 1.16 1.25 10 200 - 2.37 1.56 1.91 1.48 1.82 1.44 1.78 1.42 1.76 400 3.12 2.56 2.45 2.40 2.38 61 2.50 50 100 20* 40* (a) 2 10 0.9 0 0 E3 'j AK, y0 0.7 0.5 Kd 0 0 E3 0.1 0.3 D1 -5 'ui o0O 0.3 0.1 E3 0 0 0 E30 0.7 0.9 00 E3 009 01 0.5 1.5 2 2 -0 PA 2.50 50 (b) 2 100 03 0.9 0 0 0 0 0 0.1 0 1 0.1 0 0.3 0 Kd 0 0 I3 1 F 0.5 400 10 Kd 0.7 20* 1.5 5 ILo 0 0 0.5 0.7 0.9 3 2.5 2 0.3 S'o 3.5 3 j5A )0 2.50 50 100 20* 40* (c) 2 0 i O1 AAK Kd 0.9 00 0.1 0 0.7 iD 1 0.5 1W 0.5 0.3 3 0 0.7 E30E 0.1 0.9 0 1 1 2 0 3 4 j5A 5 60 6 7 Figure 3-27: Deformation resistance and energy absorption vs. loading for 0.3 #(Kd, 0, Ra = weight for blunt 50), evaluated at a macroscopic stress of o = 0.05 MPa. Stiffness E and energy absorption U for (a) the experiments and simulations where Ps ~ pt, (b) biological densities of tissue and scales where p, = 2 .6pt, and (c) for an hypothetical case where ps = 10pt. 62 (b) Stiffness E (a) Energy Absorption U 40 35 30 25 a) o 0.2 0.4 0.6 Kd(scale overlap) (c) 40 0.2 0.4 Kd(scale (d) 40 E/o. 35 35 30 30 25 25 * 20 20 0.6 overlap) O/S 15 10 5 0.3 0.1 (e) 0.5 0.7 Kd(scale overlap) 0.1 0.9 (f) ff/PA 40 0.3 0.5 0.7 Kd(scale overlap) 0.9 U/PA, 40 35 35 30 30 25 25 C 20 20 15 15 10 10 5 5 0.2 0.4 0.6 Kd(scale overlap) 0.2 0.4 0.6 Kd(scale overlap) Figure 3-28: Deformation resistance and energy absorption for blunt loading for #(Kd, 0, Ra = 50), evaluated at a macroscopic stress of o- = 0.05 MPa. Corresponding map of scale density # is shown in Fig. 2-7. (a) Stiffness E and (b) energy absorption U as a function of scale architecture (scale overlap Kd and scale angle 0. E/IPA and U/pA are normalized by areal density to reflect the stiffness and energy absorption as a function of weight for (c-d) the experiments and simulations where p, ~ pt and (e-f) simulations of biological densities of tissue and scales where ps = 2.6pt. 63 3.3.3 Effect of scale morphometry Increasing the aspect ratio of scales Ra = L/t, from 50 to 100 was performed in three ways: (a) fixing fixing #, Kd #, 0 and varying Kd, (b) fixing Kd, 0 and varying #, and (c) and varying 0 as shown in Table 3.5. The resulting geometries, stress- strain curves, and contours of o22 are shown in Fig. 3-29. To compare the same composition and contact geometry of varying architectures, scale volume fraction # and angle 0 were held constant and evaluated for all geometries as a function of #(Kd, 0, Ra = 100). This combination of parameters is reflected in Fig. 3-29b, Fig. 3-29a shows the original structure from the previous discussion. Here for the same volume fraction of scales # and a given fixed 0, the increase of aspect ratio reveals a tradeoff between dominant deformation mechanisms. Increasing the aspect ratio via reducing Kd increases the effective stiffness E of the assemblies, resulting in a reduced back deflection in the underlying tissue (Fig. 3-26). To examine the mechanics that govern this response, scale bending dO and rotation i were evaluated for the infinitely-periodic case. Bending within the scales was reduced for higher aspect ratio simulations, which is explained by the corresponding decrease in scale overlap. However, the trend in Fig. 3-26c suggests that scale with the same overlap Kd and increased aspect ratio would bend much more than those with the original aspect ratio, showing the large influence of Ra on the local deflections and bending curvature of the scales. Despite the same fraction of tissue in the composite, shearing of the tissue was notably reduced. Hence, while scale bending increased, scale rotation was found to significantly decrease. This is noteworthy because previous results suggest scale angle 0 (here, fixed) governs tissue shear and scale rotation. This transition from tissue shear plus scale rotation to scale bending with increasing aspect ratio served to stiffen the composite, yet resulted in elevated levels of local bending stress within the scales. These higher bending stresses will lead to early failure of the scales via yielding or fracture. 64 Table 3.5: Aspect ratio study parameters. (a) (b) (c) (d) Ra Kd 0 $ PA 50 100 100 100 0.3 0.15 0.3 0.3 20 20 20 10 0.21 0.21 0.10 0.19 1.65 2.30 1.67 1.66 0.1 0.08 - 0 06 . 0.04 0.02 00 0.05 0.1 8 (b)- (a) (c) (d) undeformed Figure 3-29: Effect of scale morphometry. Increasing scale aspect ratio Ra = Ls/t, from 50 to 100 as a function of varying Kd, 0, #. Stress-strain curves and contours of -22 are shown for these combinations of geometric parameters corresponding to Table 3.5. Architecture in (a) is used in Section 3.3.1 and (b) in Section 3.3.3. Alternatively, the aspect ratio of the scales can be increased by fixing the scale volume fraction # and the scale overlap Kd and varying the scale angle 0. Changing 65 (a) (b)0.10 15 V FE (Ra= 100, periodic) O FE (R ,= 50, periodic) 10 1 I1W 0.08 o FE (Ra= 50, finite len.) I experiment . 0.06 to CO 0) V C e 0.04 V V"- 5 0.02 (c) 0.01 0 0 1 0.5 scale fraction p 0.5 1 scale fraction $) (d) 8 = 2.5* 50 100 200 40* 0.008 - :r E 0.006 C a) Ra = 100 00 0.004 0 M R, = 50 0.002 0 00 Kd 0.5 (scale overlap) 1 20 30 0 (scale angle, 0) Figure 3-30: Deformation mechanisms for uniform compression evaluated at a macroscopic stress of o- = 0.05 MPa for 21 total geometries with Ra = 50 and 100. (a) For low scale volume fractions (# < 0.25) in experiments and finite length simulations, effective loading stiffness E of the composite increases linearly with scale fraction then plateaus for # > 0.5. Similarly, back deflection in the underlying tissue, tissue, decreases linearly for # < 0.5. Periodic simulations suggest that this plateau for large # is a boundary effect. For periodic simulations, (c) microscopic scale bending is primarily a function of overlap Kd, where as (d) scale rotation is dependent on initial scale 0. Transition from rotation to bending is evident for large Ra (dashed line). the fixed parameters results in a completely different structural arrangement, as shown in Fig. 3-29d. This structure with fixed #, Kd, Ra results in both more bending and rotation than the original structure, yet is less stiff. These results suggest that aspect ratio alone can not govern the deformation mechanisms of the structure and that 66 the response is a function of composition (#), structural arrangement (Kd, 0) and morphometry (Ra). In general, the morphometry of the scales helps to govern the dominant deformation mechanism of the assembly, but is dependent on the spatial arrangement of the scales. Typically, thin, high aspect ratio scales result in deformations primarily dominated by bending, whereas thick (low aspect ratio) scales are dominated by tissue shear plus scale rotation. This tradeoff can be tailored by adjusting individual scale aspect ratios in addition to other properties, e.g. elastic modulus, internal structure, and porosity of the scales. 3.4 Conclusions Scale architecture (Kd, 0), composition (#), and morphometry (Ra) can be used to locally tailor composite stiffness and back deflection in the underlying tissue in order to achieve protection under blunt loading events. Overlapping scales distribute stresses across a large volume of material and provide penetration resistance at a reduced weight compared to a continuous armor layer. The flexibility of the scales and tissue permits the scales to rotate and bend under applied loading. Assemblies of these scaled structures are formed without joints or hinges, forming a simple, easily adaptable design. The degree of flexibility or protection is easily tailored and adaptable by simply changing morphometry and the overlap distribution of the scales. Inclusion of a compliant tissue is key to understanding the true behavior of the composite, which is dominated by tissue constraint, tissue shearing (and resulting scale rotation), and scale bending. The composition primarily governs the effec- tive response of the composite, which is achieved by a combination of deformation mechanisms controlled by the microstructural arrangement. Volume fraction # gov- erns constraint, effective stiffness, and back deflection; scale overlap Kd governs scale 67 bending, and initial scale angle 0 governs tissue shearing and scale rotation. The influence of these mechanisms can be tailored by adjusting the scale morphometry: for scales with a large aspect ratio, the deformation transitions will be dominated by scale bending. Conversely, scales with small aspect ratio will rotate under blunt loading, accommodated by tissue shearing. 68 Chapter 4 Mechanical protection under predatory tooth attack While blunt loading captures the response of the assembly to large indenters (R > KdLS), threats are often localized to a small area. This is the case for predatory biting with a sharp tooth, with common fish predators shown in Fig. 4-1. These biting attacks are idealized as force-limited penetration with indenter of radius R. (a) (b) (C) Figure 4-1: Predatory biting attacks by (a) birchir Polypterus senegalus (b) nothern pike Esox lucius biting salmon and (c) river otter. Images adapted from Song et al. (2011); Alaska Department of Fish and Game, http://www.adfg.alaska.gov/ index. cfm?adfg=wildlifenews.view_ article&articlesid=50&issueid=13; Richard Mousel, http://www.critterzone. com/animal-pictures-nature/ mammal-otter-river.htm last accessed on May 1, 2012. 69 4.1 Finite element micromechanical model A two-dimensional plane strain composite scale-tissue micromechanical model was simulated using finite element analysis (FEA) similar to blunt loading simulations in Section 3.1.2. Quadratic continuum elements were used in ABAQUS/Standard (library types CPE6H and CPE8RH) and a fine mesh with elements of length ~ t,/4 was needed to capture the deformation within the scales. Scales and tissue were modeled with isotropic linear elastic properties from experimental characterization as shown in Table 4.1. Tissue and scale elements are modeled as perfectly bonded. Non-linear, large deformation theory with frictionless contact was assumed. Here, predatory attack via tooth indentation was simulated with rigid indenters of varying radii (R/ts = 0.1, 1, 10) and a fixed half cone angle (0/2) of ~ 30 . Table 4.1: Material parameters experimentally determined from synthetic prototype and employed in indentation simulations. Scale Tissue Behavior Modulus Elastic E, = 880 MPa Neo-Hookean Eot = 0.39 MPa Poisson's ratio Density Geometry v, = 0.4 Ps = 0.80 g/cc L, = 50 mm, t, = 1 mm vt 0.49 pt = 1.07 g/cc Htissue= L,/2 = 25 mm - The thin epidermal tissue directly beneath the indentor was removed to increase computational efficiency, as this thin outer layer of tissue would be pierced in a physiological setting. In addition, these indentations were performed on a very large sample (length > L cos0) to neglect edge effects (Fig. 4-2). All indenters were restricted to motion in the 2-direction and reaction forces F 1 , F2 , and displacement 6 were computed. Lateral reaction force F1 was negligible compared to normal force F2 for all indentations, so force reported is F2 . Local true Cauchy stress and logarithmic Hencky strain contours were evaluated 70 Figure 4-2: Indentation loading conditions, showing very large sample (length > L cos 0) to neglect edge effects, removed tissue at loading surface to increase computational efficiency, and a constrained (fixed) bottom edge. at different load levels to identify the deformation mechanisms which accommodate the loading for the different composite geometries. Indentation force is reported F2 per unit width (plane strain) and indentation depth is normalized by sample height, 6 /Htotai. Deflections in the underlying tissue 6 tissue were reported at depth t, below the scales. A normalized indentation force per depth (stiffness) is given by F2H/6tisse and evaluated at F = F2 = 5 N/mm. Strain energy density per unit thickness UlF=5N/mm 4.2 OF ydF was also determined. Indentation simulations results Idealizing predatory biting as a sharp tooth of radius R, the corresponding contact area (- 2R) and the resulting deformation-induced scale-tissue interactions govern the number of scales that are involved in the deformation. For physiologically relevant indentation (R < KdL,), the indenter only comes into direct contact with one scale. Underlapped and adjacent scales are also deformed, as governed by Kd. Under low penetration loads, the scales are not locally indented, but rather undergo significant bending with the compliant tissue between the scales shearing and the tissue beneath the scales also shearing (Fig. 4-3). For the case of a conical rigid indenter tip at low 71 loads (F = 5 N), there is a single point of contact and the resulting loading curves and underlying tissue deformation were found to be nearly identical for the range of R/t = 0.1, 1, 10 simulated. Note that this force of F = 5 N is only relevant for the materials properties and scale size shown in Table 4.1. The similarity of both the loading curves and the back deflection corresponding to Kd = 0.3, 0 = 20', # common overlap of Kd = = 0.21, and Kd = 0.3, 0 = 5', # = 0.77 indicate that the 0.3 controls the mechanical response of the composite. In contrast, the response for Kd = 0.9, 0 = 5', # = 0.26 is much more compliant and permits more back deflection tissue in the underlying tissue because of the minimally overlapping scales of Kd = 0.9. Localized indentations of R/t, = 0.1 are primarily accommodated through the mechanism of scale bending, which is isolated to only a few scales. Evaluating across all architectures #(Kd, 0, Ra = 50), the overlap of the scales (Kd) governs the ability of the composite to resist penetration (Fig. 44). Normalized indention force (local stiffness) increases with increasing scale overlap (decreasing Kd). Similarly, back deflection in the underlying tissue was found to decrease with overlap such that architectures with highly overlapping scales (Kd - 0) are very stiff and had minimal back deflection. Due to the localized, nonuniform deformation of the assembly under finite indentation, scale rotation is no longer a dominant deformation mechanism. In these layered composite structures, the overlap, angle, and volume fraction work together to give interplay between scale bending and shearing of the tissue to involve more material. These geometries are mechanisms to absorb more energy by involving greater amount of material laterally as well as to deeper depths. Larger indenters (R > KdL, and simulated for R/t, = 10 as shown in Fig. 4-5) under large penetration loads permit the scale to bend around the indenter. Above F = 10 N, these larger radius indenters produce a stiffer behavior as the contact area between scale and tooth increases. For indenters whose radii span multiple 72 (a)8 (b) 0.08 - 0 -0.08 -1 Kd=0.3, 0=20*, $)=0.21, PA=1 .65 Kd= 0.3, 0=5, $=0.77, PA 1 1 4 Kd=0.9, 0=50, $=0.26, PA 1.16 -0.16 0 0.15 .0 6/H u2 (mm) -2.5 0 -0.2 0 40 tissue path (mm) 822 uj0.2 E12 -0.2 i o 0.2 -2.5 (MPa) ui2.5 022 -0.1 (MPa) inu0.1 (d) (e) (f) Figure 4-3: Indentation on fish scale assembly for radius R/t, - 0.1. (a) Loading curves for three geometries show that indentation force is nearly identical for architectures with similar overlap, despite very different 6, $. (b) Corresponding underlying back deflection and (d-f) contours of stress, strain taken at F = 5 N. (b)0.20 (a) 80 00 0 70 0 0 0 0 0.12 z 00 g0 40 00 0 0 0 0 2 50 M U- 30 0 00 0 0.16 EO E 60 0 0 0.081 0 OD 00 0 0 0.04' 0 0.5 q)(scale fraction) 0 (c) 80 (d)0.20 70 0.16 E 60 z V7 7 1050 7 7 v' v V 0.12 *vV7 7 ~2 5 1 0.5 $ (scale fraction) 1 V7 V7 0.08 40 30 7 0 (scale angle, 0.04 L 0 4 20 0 40 20 0 (scale angle, 0) 0) (f) 0.20 (e) 80 70 0.16 LI El E 60 E _r S0.12 z LI 50 U- 0.08[ El 40 30 0 Kd 0.5 (scale overlap) 0.04' 0 1 0.5 Kd 1 (scale overlap) Figure 4-4: Indentation force and back deflection evaluated at F = 5 N across all scale assembly architectures #(Kd, 0, Ra = 50) for a sharp indenter with radius R/t, = 0.1. (a-d) While little relationship existed between scale angle 6 and volume fraction 4, (e-f) there is a direct trend between overlap Kd and the ability of the composite to resist penetration. scales, the load is distributed across the scale assembly which reduces deflections, scale deformation, and stress concentrations, thereby protecting the fish. This explains 74 why predators benefit from having sharper teeth: if one isolated scale sees too much bending, the scale can be defeated. By localizing the indentation deformation to a finite region, predators with a given biting force are able to penetrate the underlying tissue. While this elastic model does not incorporate failure mechanisms such as fracture, delamination, or plasticity, the stress distributions within the scales and tissue give insight to failure modes and locations resulting from excessive deformation. High shear strain 12 of the tissue promotes failure by delamination of the tissue-scale interface, which permits scales to dislodge and damage the structure of the composite. Alternatively, regions of high bending stress o-l of the scale promote yield and fracture of the scale. 30 radius: R/t = 0.1 .indenter 25 K= 0.3, = 20*, $=0.21 Kd 0.3, 0 = 5-, $=0.77 Kd= 0.9, 0 = 5*, P=0.26 20 E Z_ 15 u.- 10 indenter radius: R/t = 10 ----- Kd =0.3,E=5', $=0.77 ----- Kd=0.9,6=5*, p=0.26 5 0 0 0.1 0.2 0.3 0.4 0.5 6/H Figure 4-5: Indentation loading curves for varying indenter geometry (R/t = 0.1, 10) show little effect of R on mechanical response due to point-wise interaction for small loading. At larger loads, the scale deforms around the indenter radius and the reaction force increases. Contours shown of &. To determine the ideal armor architecture, the penetration resistance of the com- posite must be evaluated against competing mechanical functionalities. While segmented scale armor is inherently more flexible than a solid plate of the same thickness, 75 (a) 0. 4 0 j Kd 0. 0.3 5 00 0 .3 0 -4 0 200 100 2.50 50 0 40* 80 0 0 o 0.1 60 _E 0.3 0.9 0. 7 0. E 0 03 0 .2 0.1 50 130 0. 0.1 o 0 0 5 0.2 1 Kd 70 E0 0 Q 1.6 0. 1.4 5% 8B 1.2 1.8 2.2 2 0.5 U- 0.7 0.9 40 30 2 .4 TA 2.5050 100 (b) 0. 400 200 20*0 40* O0 0 - -70 0.315-00 -Kd 3 0.1 0. 0. 9 0. 0. Kd E 5 3 D -- 0. 1 0.3 - 60 0 03 E $ 0.3 LL E 0.2 5 0 1 SO - 50 0 0 .2 0.1 5b 1 . 29 1.5 0 40 3 30 3.5 O. 2.5 2 0.5 M.. 0.7 ,_ 0.9 5A Figure 4-6: Deformation resistance and energy absorption vs. weight for indentation for <(Kd, 0, Ra = 50) and R/t = 0.1, evaluated at indentation load of F=5 N/mm. Stiffness FH/6 and energy absorption U for (a) the experiments and simulations where p, - pt and (b) biological densities of tissue and scales where p, = 2 .6pt. ongoing and future work remains to characterize the flexibility of the these systems. The relative weight of the composites is given by the areal density PA of each architecture, which is primarily a function PA(L sin 0) given the similar density p of the scale and tissue. It can be shown in Fig. 4-4 that geometries with high overlap (e.g. Kd = 0.1) resist penetration best. However, the geometric constraints of the structure (as shown in Fig. 2-7) restrict that high overlap is correlated with high scale angle 0, thus increasing the weight PA of the structure. Thus to minimize the weight of the armor while still maximizing its protection against indentation, the composite should 76 have both a minimum scale angle 0 and high overlap Kd. As these are conflicting requirements, a median value such as Kd ~ 0.5 is a compromise (as found in nature), though other factors such as flexibility should be considered. This tradeoff is shown in Fig. 4-6. 4.3 Conclusions Finite indentation of overlapping scaled structures models physiologically relevant predatory biting attacks. Simulations reveal that as the contact area of the loading surface is reduced to a finite radius penetration indentation (R = oc - R ~ 0), there is a transition from tissue constraint and scale rotation (under blunt loading) to a deformation dominated by localized scale bending, which is accommodated by shearing of the tissue interlayers and the underlying tissue substrate and governed by the overlap of the scales. Architectures with high overlap (low Kd) resist penetration by distributing loading over a large area. Under penetration, back deflection in the underlying tissue is nearly proportional to the overlap of the scales (6 tsase oc Kd). Deformation of a scale during localized bending could risk failure by plastic deformation or fracture, exposing the underlying tissue and defeating the armor. Here, highly overlapping scales are beneficial as they provide multiple layers of defense; in order to penetrate the tissue, the indenter must deform and defeat Kd-1 scales. Alternatively, scale rotation deforms the compliant tissue which could delaminate or tear gracefully, leaving the stiff scales intact. 77 78 Chapter 5 Bending and flexibility A segmented armor system, such as the layered, overlapping scale and tissue structure found on fish is advantageous due to its unique combination of mobility and protection. While segmented scale armor is inherently more flexible than a solid plate of the same thickness, the flexibility of the these systems must be characterized. This work presents an experiential introduction to evaluating the flexibility of overlapping, layered composite structures under microscopic bending. 5.1 Materials and methods To roughly evaluate the response of the composites under bending, the samples were manually deformed over a steel die, as shown in Fig. 5-1. The die is of radius R - 80 mm and was chosen because it was the largest die readily available. Smaller radius dies were hypothesized to easily break the samples under a small applied load. The tested samples correspond to those tested under blunt compression loading in Chapter 3 for <(Kd, 0, Ra = 50) as shown in Table 3.1. They are fabricated of 3-D printed ABS and silicone rubber and measure 100 mm by 40 mm by -50 mm tall, depending on scale angle 0. 79 5.2 Results The resulting deformations are shown in Fig. 5-2. Delimitation was the primary mode of failure for these samples under bending. Structures with high scale angle 0 delaminated most easily. The scales showed little sign of mechanical deformation (scale bending) and instead rotated with the underlying tissue substrate to accommodate the curvature. These results of tissue delamination correspond to results observed experimentally with the true biological fish integument: under extreme curvature (e.g. folding the integument upon itself) the scales "popped out" of the composite. In contrast, the same samples were found to fracture and crack under blunt and indentation loading. Hence, to function as a flexible armor, the composite must resist multiple failure mechanisms. Future testing should be performed on longer samples of length of at least 3-4 L, to neglect boundary effects. Testing should be done using a fixture and using instrumented equipment to record the load applied evenly to the samples. In addition, digital image correlation (DIC) should be used to track the deformation of the samples. Finally, these samples should be in good condition; several of the samples were damaged from previous compression testing. Figure 5-1: Experimental setup for bending over die of radius R = 80 mm. 80 LO, Ci CD LO CD Figure 5-2: Experimental bending of 3-D printed ABS and silicone rubber structures over a die of radius R = 80 mm for 0 = 2.5'. 81 LO a) 'LO CD II (continued~I) f3DpitdASadslcn edn Eprmna 0m frdu overa d~~~~~~~i 08 o ' ubrsrcue 0r t6 6 edn (cniue)Eprietl ovradi 0-1' frdisR -80m I83 o o - ritdAB n ilcn rbe trcue 0 iI N 0 Ci a *o I Ci CD silicone rubber structures (continued) Experimental bending of 3-D printed ABS and over a die of radius R = 80 mm for 0 = 20'. 84 0 a) 0 0 LO a) (cnine) ovradi xermntlbedngo frdisR 0-4' =80m 3DprnedAS o 85 n slcoerubr tucue 86 Chapter 6 Conclusions 6.1 Conclusions Fish scale armor provides a model biomimetic composite for the design of structures to provide global penetration resistance and flexibility. Biological systems have a limited library of materials available, often consisting of rigid mineralized components and compliant organic tissues. The morphology and architecture of the material organization is one unique way to provide a spectrum of structural functionalities. Overlapping scale units distribute stresses across a large volume of material and provide penetration resistance at a reduced weight (and subsequent cost of mineralization) compared to a continuous armor layer. Scale architecture (Kd, 0), composition (#), and morphometry (Ra) can be used to locally tailor composite stiffness and back deflection in the underlying tissue in order to achieve biomechanical protection. Under blunt loading, the composition primarily governs the effective response of the composite, which is achieved by a combination of deformation mechanisms controlled by the microstructural arrangement. Inclusion of a compliant tissue is key to understanding the true behavior which is dominated by tissue constraint, tissue shearing (and resulting scale rotation), and scale bending. Scale volume fraction 87 # governs constraint, effective stiffness, and back deflection; scale overlap Kd governs scale bending, and initial scale angle 0 governs tissue shearing and scale rotation. The influence of these mechanisms can be tailored by adjusting the scale morphometry: for scales with a large aspect ratio, the deformation transitions will be dominated by scale bending. These results are summarized in Fig. 6-1. high mineralization cost increasing <b, tissue constraint decreasing overlap (increasing Kd) increasing indentation deflection CD CO, .C scale bending E12 high back deflection scale rotation, tissue shear Figure 6-1: Summary of deformation mechanisms under blunt loading. As the contact area of the loading surface is reduced to a finite radius penetration indentation, there is a transition from tissue constraint and scale rotation to a deformation dominated by localized scale bending, which is accommodated by shearing of the tissue interlayers and the underlying tissue substrate and governed by the overlap of the scales. Architectures with high overlap (low Kd) resist penetration by distributing loading over a large area. Deformation of a scale during localized bending could risk failure by plastic deformation or fracture, exposing the underlying tissue and defeating the armor. Here, highly overlapping scales are beneficial as they provide multiple layers of defense; in order to penetrate the tissue, the indenter must deform and defeat K - scales. Alternatively, scale rotation deforms the compliant 88 tissue which could delaminate or tear gracefully, leaving the stiff scales intact. Such mechanical characterization could have useful applications in developing principles for protective systems (Neal and Bain, 2004), composite textiles (Hatjasalo and Rinko, 2006), and even electronics (Kim et al., 2012). 6.2 Contributions This thesis presents three main contributions that have been achieved: 1. A parametric model is developed to describe the fish scale assembly of the integument. Here, the scale architecture (Kd, 0), composition (#), and morphometry (Ra) are systematically varied to determine the effect of each variable. 2. Micromechanical models are used to simulate these various geometries under blunt and indentation loading to describe the effective response and protective features of the structures. Deformation mechanisms are characterized, revealing transitions across scale and indenter geometries. 3. Physical prototypes are fabricated and tested, showcasing the strengths and limitations of experimental techniques such as 3-D printing and digital image correlation. 6.3 Future work Simulations should explore the role of these layered structures as designs to involve more material energy absorption (Fig. 6-2a). Macroscopic bending of the scale-tissue structure should be investigated (Fig. 6-2b). It is important to understand the tradeoff between protection and mobility in flexible, segmented armor design. This could be performed analytically, simulated using a finite element micromechanical model, 89 and compared to preliminary bending experiments performed in Chapter 5. In addition, advanced micromechanical models should incorporate viscoelasticity, plasticity, failure (delimitation shown in Fig. 6-2c). Experimental work should focus on validating indentation simulations in Chapter 4. These indentation experiments need to be performed on samples with a larger length than those used for blunt loading to neglect boundary effects. Simulations predict that a sample length of at least 3-4 L, is needed for the stress fields to approach zero given similar loading. Indenters of varying radii have already been fabricated (Fig. 6-2d). (a) (b) B'(x,y) = (R-h)sine, (R-H)(1-cosO)+h = UR=LKB' B L (d) Figure 6-2: Future work. (a) Bilayered composite structures that involve greater material in energy absorption (contours of e12 shown). (b) Macroscopic bending of composite. (c) Delamination and failure observed experimentally in early prototype under bending needs to be characterized. (d) Experimental indentation of samples could be achieved with fabricated plane strain indenters. 90 Appendix A Materials characterization of the scales of Arapaima gigas The primary objective of this research is to study the mechanical behavior of the elasmoid scales of a variety of fish. Giant fish measuring over two meters in length and weighing in at over 100 kilograms like the Amazonian fish Arapaima gigas are fascinating biological systems. These fish possess large scales of varying degrees of mineralization that serve as a dermal armor to natural predators. These scales are elasmoid and consist of an elasmodine basal plate of a twisted plywood arrangement of collagen fibers capped with different hypermineralized tissues (Meunier and Brito, 2004; Sire et al., 2009; Torres et al., 2008). Scales of A. gigas have a partially mineralized base and well mineralized limiting layer of hydroxyapatite. These layers can be easily seen with optical and scanning electron microscopy (SEM) as shown in Fig. A-i. A.1 Elemental analysis with EDX and BSEM Energy-dispersive X-ray spectroscopy (EDX) and backscattered scanning electron microscopy (BSEM) were performed of the fish scales of Arapaima gigas to determine 91 dorsal (a) anterior (b) posterior ventral (c) anterior/ rostral field, embedded in dermis posterior/caudal field, covered in epidermis (exposed) Figure A-1: (a,b) Arapaima gigas are large fish that possess equally large scales (- 50 mm long and ~1 mm thick) with human hand shown as reference with dry scales. (c) Off-white portion of scale in embedded in dermis, while dark portion is covered in the epidermis (exposed). Scanning electron microscopy (SEM) images of embedded (d,f) and exposed (e,g) section of scale showing external mineralized layer and basal collagen layer. Scales in (d,e) are sectioned with scissors, while those in (f,g) are mounted in epoxy and polished. 92 the material composition. A section of the scale was embedded in a room-temperature curing epoxy resin and cross-sectioned in the transverse and longitudinal directions To reduce using a diamond-impregnated annular wafering saw at 800-900 r.p.m. surface roughness, samples were polished on a polishing wheel with 15 Pm, 6 pm, and 1 pm silica nanoparticles on a soft pad, and again with 50 nm silica nanoparticles on a cloth pad in distilled water. Surface roughness was verified using atomic force microscopy (AFM). Sample was sputter coated with gold for conductivity. Samples were scanned in a JEOL JSM-6700F at 15 kV and 10.0 pA using a 26 mm sample holder. EDX was used for spatial imaging as shown in Fig. A-2a. Spectral analysis with BSEM (Fig. A-2b) confirms composition of an internal layer of collagen and and outer layer of hydroxyapatite. Silicon was also present, likely from polishing contamination. phosphorus oxygen (a) (b) Z IP 0 1 Z:2 gradient even distribution Ca internal 2 4 3 5 keV - collagen 6 7 8 9 10 Figure A-2: Materials characterization of Arapaima gigas scales by (a) energydispersive X-ray spectroscopy (EDX) and (b) backscattered scanning electron microscopy (BSEM), confirming an internal collagen layer and an outer mineralized hydroxyapatite layer. Agrees with similar results published in Meyers et al. (2011). 93 A.2 Mineralization analysis with TGA Thermogravimetric analysis (TGA) was used to assess the mineralization of the scales of Arapaima gigas using a TGA Q50. It was found that a mortar and pestal created a paste that was not suitable for TGA so scissors were used to create a very fine powder of dried fish scales. A 10 mg powdered sample was placed in a platinum pan and into the instrument. After the thermal profile in Table A.1, air was used to cool the device. Table A.1: Thermal profile for thermogravimetric analysis Equilibrate temperature Initial isothermal time 30 0 C 5 minutes Ramp rate 10-20 'C/min Maximum temperature Final isothermal time 800 0 C 5 minutes After water evaporated from samples, organic and inorganic composition were determined by computing the mass that burns off during thermal ramp and the remaining mass, respectively (Fig. A-3). Density of hydroxyapatite (p = 3.17 x 103 kg m3 ) and collagen (p = 1.33 x 103 kg M3 ) were used to determine volume fraction of mineralized and organic components and compared against previously published values (Ikoma et al., 2003; Torres et al., 2008). The results conclude that the scales of Arapaima gigas are A.3 - 24% mineralized by volume. Nanoindentation Nanoindentation was employed to evaluate the spatially-specific stiffness, hardness, and deformation mechanisms of the different scales through the thickness of the two different material layers. Samples were embedded and polished as in Section A.1. Nanoindentation of the scale samples was performed on a Hysitron Triboindenter 94 (collagen) -0.3 -0.2 0.2 % inorg 0.150 40 -0.1 40 0.05 Temperature ("C) experimental Ikoma et al, 2003 Torres et al 2008 inorganic organic 24 62 S 2 25 20 43 5 54 Figure A-3: Mineralization of Arapaimagigas scales using TGA. Density of hydroxyapatite and collagen were used to determine volume fraction of mineralized and organic components. Results for volume % characterization consistent with other elasmoid fish scale mineralization data (Ikoma et al., 2003; Torres et al., 2008). using a Berkovich (triagonal pyramid) diamond probe tip following work by Bruet et al. (2008). Due to the nonideal tip geometry of the indenter and frame compliance, the system was calibrated prior to each set of experiments using a standard fused quartz sample. This calibrated indentation also provides the probe geometry (tip end radius and truncate height) for use in data analysis. The piezoelectric transducer was allowed to equilibrate for 660 seconds prior to each indent. The loading profile consisted of constant loading rate of 50 pN/s until the maximum set peak load is reached, hold period of 3 s at this maximum set peak load, a constant unloading rate of -50 pN/s until a zero force load is reached. A maximum set load of 500 MN was used as an initial estimate based on successful penetration depth. Indentions was performed along the length of the sample approximately every 2 pm to observe the different material layers. All indentation load-displacement data collected from the experiments were ex95 ported for processing in MATLAB. Material parameters including elastic modulus and hardness were determined from the indentation unloading curve based on the OliverPharr method (Oliver and Pharr, 1992) and preliminary experiments are shown in Fig. A-5. The reduced elastic modulus is calculated as Eo-p = '/ 2 S Amax where S is the slope of the initial region of the unloading indentation-depth curve, and Amax is the projected area of contact at the maximum contact depth. This projected area is derived using the geometry of the probe tip along with contact depth from the instrumented indentation curves. The hardness is calculated as H0 P -- Pmax Amax where Pmax = 500 pN is the maximum load of the indentation curve. The hardness obtained represents the load-bearing capacity of the material. ; ocation o; Lnanoindentation Figure A-4: Location for nanoindentation of Arapaima gigas scales. croscopy of scale embedded in epoxy, ground and polished. 96 Optical mi- 60 C!, 20 . 0- w -...... ............ V+ 0 z -. - - 40 .--- .-.. 12. ...-. .. . .. . .... .... .. .. .. . . . . . . . . . .*. . .. 3 -V-w- #W+* ..... -------L_ 100 0 200 300 :L 500 400 600 700 Distance (t m) 3r 200 400 600 Indentation Depth (nm) 800 a- 2k 6 I .... -. -.-.. ... 0T 0 ............ 1.0 0.5 -............. *:'**.~ -~ . I 100 ***4*~* 4. * * I *. 200 **~ 44~4* Y--Y 70 I 300 *~ 400 . 500 600 700 Distance (t m) IK Figure A-5: Preliminary data for Oliver-Pharr stiffness and hardness from nanoindentation of external region of Arapaima gigas scale where Pmax = 500 pN. This work is preliminary and needs refinement to determine appropriate indentation loading profile and indentation spacing to accurately capture the laminate collagen layers. While tensile testing reports elastic modulus an order-of-magnitude lower around E ~ 2 GPa (Torres et al., 2008), similar results are published in Meyers et al. (2011). ......... . .......... .. 98 Appendix B DIC and load synchronization The following programs are meant to extract and analyze data from a mechanical testing machine (e.g., Zwick). The program importZwickCycle .m I reads the . tra file and produces a N x 1 cell array of data (e.g., "Test time" ;"Cycle number"; "Standard travel"; "Standard force") or any output specified by the Zwick program file. These vectors will all be the same length N (and typically quite large due to the data acquisition rate of the Zwick). The displacement or strain data from digital image correlation (DIC) extensometry is imported in a similar manner. Due to the finite frame rate of the camera, there will be far less data points from DIC than from the Zwick (vector of length M < N). These data sets are conditioned as needed (e.g. noise removed below a threshold, zeroed) such that the curves all start from the same time corresponding to the start of the test. Synchronize the strain data from DIC to load data from Zwick using interpZwCam. m. This program will linearly interpolate the camera strain over the Zwick data such that the strain vector is also of length N. This will permit you to plot displacement (or strain) vs the force (or stress) to extract true loading data and calculate elastic moduli without concern about frame compliance. 'Code written for MATLAB R2009b. 99 function [data] = importZwickCycle(filename) % written by ashley browning, 2011 X reads force, displacement, etc data from a zwick.tra file X---------------------------------------------------------------fid = fopen(filename,'r'); % read line from file tline = fgetl(fid); ~= -1 while tline if or(tline(2:5) == 'Test', tline(2:5) == 'Stan') % "Test time";"Cycle number";"Standard travel";"Standard tline = fgetl(fid); % skip dimensions data = textscan(fid,'% f u%fu % fu%f','delimiter',';'); else tline = fgetl(fid); end end fclose(fid); end 100 force" = interpZwCam(timeCam, [strainInt] function strainCam, timeZwick) . ---------------------------------------------------------------- % written by ashley browning, 2011 to interpolate Y from DIC to the Load data from Zwick %/ % the strain data timeCam is time from extensometer (seconds) timeCam = (frame number) / (frame rate in fps) % strainCam is from extensometer (unitless) %/ /0 % timeZwick timeZwick is real time from zwick can be read from the .tra fiLe in seconds, with the Loading rate and displacement or computed ideally, timeCam, strainCam, timeZwick have all been zeroed and thresholded appropriately so they coincide with true start shortest time (all common time) so that you interpolate over the same if timeCam(end) >= timeZwick(end) % truncate extensometer time if use I = timeCam <= time interval timeZwick(end); timeCam = timeCam(I); strainCam = strainCam(I); else % truncate zwick time display('youucutuoffutheuextensometerutoousoonu:(') display('timeZwick(end)utimeCam(end)') [timeZwick(end) timeCam(end)] end % interpolate over finer time abscissa strainInt = interpl(timeCam, strainCam, timeZwick); if(length(strainInt) ~= length(timeZwick)) error('fixulengthsuofuinterpolateduvectors') end % fix weird interpolation dstrain = strainInt(end-1) error at end - strainInt(end-2); strainInt(end) = strainInt(end-1)+dstrain; end 101 102 Appendix C Periodic boundary conditions This methodology was developed by Danielsson, Parks, and Boyce Danielsson et al. (2002) and based off Nuo Sheng's M.S. thesis with the input of Narges Kaynia. C.1 General mathematical formulation for periodicity in two directions Deformation Often when analyzing complex structures, it is convenient to only model a single repeating unit, or representative volume element (RVE). Using RVEs is computationally efficient and neglects artificial boundary effects from a larger, finite structure. In order to simulate the RVE in a finite element package such as Abaqus, periodic boundary conditions are needed. To develop equations to implement in Abaqus, consider the 2-D solid with periodic boundaries in the 1 and 2-directions undergoing deformation in Fig. C.1. 103 TL 2 TR YY2 Ia XX2 Xxi t-1 YY1 BR BL 4 Li - Figure C-1: Representative volume element for periodic boundary conditions. BL= bottom left node, BR= bottom right node, TL= top left node, TR= top right node, XX1= left node set, XX2 = right node set, YY1 = bottom node set, YY2 = top node set. For a given 2-D solid under plane strain, let the deformation gradient tensor F be Ox F ax Fn F 12 F F 22 21 Let the displacement tensor H be F 11 - 1 H=F-1= F F 12 F 22 21 -1 Hn H H H 21 12 2 2 Periodicity The following general relationships exist for displacement u, given symmetry in the 1- and 2-direction for a 2-D, element shown in Fig. C.1. Here, we relate both the vertical displacement (ui) and the horizontal displacement (U2) on the two sets of paired edges (top and bottom, left and right). These relationships are described by the following equations: UXX2 u1v2 UYY2 _ - xXi1 HnL1 12L2 YY1 H U XX2 xX2 UYY2 U2 104 _ - = Xx11=i UYY1 U2 H21L1 H2 2 L2 (C.1) Depending on boundary conditions and specific periodic constraints used, some or all corner nodes may or may not be included in the edge sets. Assume here that all corner nodes are excluded from the edge sets. To fully define periodicity on the RVE, these corners must also be related. All relationships between the four corners (top left, bottom left, top right, bottom right) are described by the following equations: uf" =HnL1 -uL TL u = uBR HL _ UBL = UTR - L = H21L1 H21L1 (C.2) uT TR BR - 2 uUTL 1 C.2 u BL 1 BR =H22L2 - _ TL - H12 2 BL H2 Case I: Fixed corner node with periodicity in two directions To implement the most general case of the framework developed in Section C.1, let's consider the geometry shown in Fig. C.2. Here we've fixed the bottom left (BL) node and redefined our edge sets. TL 2 TR L2 t-1 1BI I1 R Figure C-2: Node sets for periodic boundary conditions. When defining our edge sets, we have to be sure to not over-constrain any nodes. Each degree of freedom (ui, U2 ) must only be declared once. In addition, Abaqus requires that the equations are in the simplest possible form. To do this, we must 105 ensure that the equations have been reduced as much as possible and that we don't repeat variables. We've made the following modifications to our edge sets: 1. XX1 is the left side, including BL node, not including TL node 2. XX2 is the right side, including BR node, not including TR node 3. YY1 is the bottom side, including BL node, not including BR node 4. YY2 is the top side, including TL node, not including TR node Now only the TR node must need corner periodic equations (C.2), as all the other corners are a part of the edge sets and defined by (C.1): Boundary conditions To prevent pure translation and rotation, a single node must be fixed. Here, we choose BL to be the fixed node, but this is arbitrary. The following boundary conditions must be specified: UBL U1 UBL 2 - - o To respecify the sets to include/exclude corners, we must go back and simplify our corner periodic equations (C.2): 0 BR - Ui = -[L TR HnL1 =HL uBR - e= H 21 L 1 T2 L uTR U2 =H21L 1 ui - UBR UTL -K= UTR 12L2 - 0 BR 0 H12L2 U2L -e=H22L2 Rearranging (and subsequently eliminating any equations with variables uTL), uBR or this reduces to UJR = Hn L1 + H 12 L 1 u2' 106 = H 21 L 1 + H 2 2L 2 (C.3) Together, equations (C.1) and (C.3) define our periodic constraints. Double checking with the way we defined our edges in Fig. C.2, we can verify that each degree of freedom is only declared once. Loading An external load is applied on the body by prescribing the displacement gradient H = F - 1. Let H be defined be creating two arbitrary virtual dummy nodes, VI and V2, such that the displacements u1 , U2 of these virtual nodes correspond to the components of H. H= Hu H12 U 71 H 21 H22 U U2 Hence, to apply uniaxial strain of A F 22 uV2 1 = U2= - U V2 2 ' 2 1.1 in the 2-direction, prescribe H 22 = 0.1. Alternatively, to apply simple shear y = 0.1, prescribe H 12 = 0.1. Implementation Refer to Section C.4 for implementation in an Abaqus input file. Virtual nodes V1, V2 are unattached to the meshed part. If included at the *Part level, their number must be larger than the maximum number of nodes in the mesh. If included at the *Assembly level, they are not related to the part at all and their number is arbitrary (can be 1, 2, etc.). H= Hu1 H21 H 12 H22 u 1 = UV2 1 99971 99981 1 U V1 U2 99972 99982 They can be defined at any location, though placing them at the origin simplifies 107 interpreting their displacements: *Node, nset=FakeDefNodes 9997, 0., 0. 9997, 0., 0. The previous relationships can be simplified as the following, which can then be programmed into Abaqus using *Equation (for L1 =10 and L1=20): xi - HuL 1 = 0 Yx1 - YY1 1 - H12L2 urx2 _ x2 _ YY2 U1Y2 U Y 2 _ Y1 - U R = TR U2 NOTE: - H =H H21L1=0 XX1, 1, 1, XX2, XX1, 2, 1, XX2, 1, -1, 9997, 2, -1, 9997, 2, 1, 10. 10. 0 YY1, 1, 1, YY2, 1, -1, 9998, 1, 20. H22L= 0 YY1, 2, 1, YY2, 2, -1, 9998, 2, 1 L1 21 L, = _ + H 1 2 Li TR, 1, 1, 9997, 1, -10. + H22L2 TR, 2, 1, 9997, 2, 9998, -10. 20. 1, -10. 9998, 2, -20. The order of the variables in these equations has a significant effect on how the stiffness matrix is solved. When defining the node sets, ensure that they remain in the correct order by declaring unsorted in the node definition. In addition, assure that the x, y coordinates of these node sets match up identically. For complex geometries, this can be done by verifying in MATLAB, as shown in Section C.5. The boundary conditions for true and dummy node sets can be implemented using *Boundary: U BL U F11 = 1.1 0 uBL = 0 H11 = 0.1 108 BL, 1, 2, 9997, 0. 1, 1, 0.1 C.3 Case II: Fixed corner node, roller surface, with periodicity in one direction This is the periodic boundary condition used in the model in Chapter 3.3.1. Consider the 2-D solid with periodic boundaries in the 1-direction undergoing deformation in Fig. C.3. This would be used for modeling compression or tension of a very long structure with finite height. 1. XX1 is the left side, not including BL node 2. XX2 is the right side, not including BR node 3. YY1 is the bottom side, including both BL and BR nodes TL TR YY2 2 L2 Xxi XX2 1BL I ;M R Li Figure C-3: Periodic boundary conditions used in model. BL= bottom left node, BR= bottom right node, TL= top left node, TR= top right node, XX1= left node set, XX2 = right node set, YY1 = bottom node set, YY2 = top node set. Boundary conditions To prevent pure translation and rotation, a single node must be fixed. Here, we choose BL to be the fixed node. The following boundary conditions must be specified: u1 BL B 2uL= 109 0 To allow frictionless free expansion in the 1-direction and to prevent rotation, impose a roller condition for bottom surface YY1: ? 4 Y1 =o The periodic equations are as follows. Note the corner equation is simplified based on our fixed node boundary condition: XX2 XX2 _ Xxi = HuLi Xxi H21Li uB= H1 1 L 1 Loading For this problem, we will apply a uniform pressure in the 2-direction for plane strain elements. We want to specify no shear and ensure that the elements are free to expand in the 1-direction. Thus, we need to set the shear deformation F2 1 = 0. Apply traction, displacement, contact, or pressure conditions as normal in Abaqus CAE, e.g. we can just specify the pressure using *Dsload. You can also impose a finite indentation, but note that the indentation would be repeating along the length due to the periodic boundary conditions. Alternatively, you can impose a uniform compression using the deformation gradient F by declaring a displacement on the F 22 (99982) node. Implementation Refer to Section C. 5 for implementation by defining all boundary conditions in MATLAB. Now the previous relationships can be simplified as the following, which can then 110 be programmed into Abaqus using *Equation (for L 1=10): XX2* - XX1* - 9997 1 L1 = 0 xx1, 1, 1, XX2, 1, -1, 9997, XX2* - XX1* - 9997 2 L1 xx1, 2, 1, XX2, 2, -1, 9997, 2, = 0 BR 1 - 9997 1 L1 = 0 BR, 1, 1, 9997, 1, 10. 10. 1, -10. The boundary conditions for true and dummy node sets can be implemented using *Boundary: BLo BL, 1 1= H21 C.4 0 =0 1, 1, 0. YY1, 2, 2, 9997, 2, 2, 0. 0. Abaqus input file The following Abaqus input file implements the above formulation of periodicity in both x and y by use of virtual "dummy" nodes. In addition, this input file demonstrates the ability to impose a deformation by declaring the deformation gradient F. By selectively uncommenting the ** lines and changing the components of F, the load can be changed from uniaxial compression/tension, biaxial compression/tension, or simple shear. 111 *Heading ** Job name: Mesh2D Model name: Model-1 ** Generated by: Abaqus/CAE Version 6.8-2 *Preprint, echo=NO, model=NO, history=NO, ** ** PARTS ** *Part, name=Part-1 *End Part ** ** ASSEMBLY ** *Assembly, name=Assembly ** *Instance, name=Part-1-1, part=Part-1 *Node, nset=Nall 1, 1., 0. 2, 0., 0. 3, 1., 1. 4, 0., 1. 5, 0.5, 0. 6, 1., 0.5 7, 0.5, 1. 8, 0., 0.5 9, 0.5, 0.5 *Element, type=CPE4H, Elset=ElAll 1, 5, 9, 8, 2 2, 8, 9, 7, 4 3, 1, 6, 9, 5 4, 3, 7, 9, 6 *Solid Section, elset=ElAll, material=NH *End Instance ** *Nset, 2, *Nset, 1, *Nset, 4, *Nset, 3, nset=BLNode , instance=Part-1-1 nset=BRNode, instance=Part-1-1 nset=TLNode, instance=Part-1-1 nset=TRNode, instance=Part-1-1 ** ** PERIODIC BOUNDARY CONDITIONS: ** *Node, nset=FakeDefNodes 9997, 0., 0. 112 contact=NO 9998, 0., 0. ** XX1 is the left side, including BLnode , not including TLnode *Nset, nset= XX1, instance=Part-1-1, unsorted 2,8, ** XX2 is the right side, including BRnode , not including TRnode *Nset, nset= XX2, instance=Part-1-1, unsorted 1,6, ** YY1 is the bottom side, including BLnode , not including BRnode *Nset, nset= YY1, instance=Part-1-1, unsorted 2,5, ** YY2 is the top side, including TLnode , not including TRnode *Nset, nset= YY2, instance=Part-1-1, unsorted 4,7, *Equation 3 XX2, 1, XXi, 1, 9997, 1, 3 XX2, 2, XX1, 2, 9997, 2, 3 YY2, 1, YY1, 1, 9998, 1, 3 YY2, 2, YY1, 2, 9998, 2, 3 TRNode, 9997, 1, 9998, 1, 3 TRNode, 9997, 2, 9998, 2, 1 -1 -1.0 1 -1 -1.0 1 -1 -1.0 1 -1 -1.0 1, 1 -1.0 -1.0 2, 1 -1.0 -1.0 ** *End Assembly ** *Material, name=NH *Hyperelastic, neo hooke 50.,0.0004027 *Step, name=Strain, nlgeom=YES *Static 0.01, 1., le-05, 1. 113 *Boundary BLNode, 1, 2, 0.0 ** APPLY LOAD TO DEFORMATION ** unixial **9997, **9998, **9998, ** ** compression 2, 2, 0.0 1, 1, 0.0 2, 2, -0.1 biaxial **9997, **9997, **9998, **9998, compression 1, 2, 1, 2, simple 9997, 9997, 9998, 9998, GRADIENT 1, 2, 1, 2, 1, 2, 1, 2, -0.1 0.0 0.0 -0.1 shear 1, 2, 1, 2, 0.0 0.0 0.1 0.0 ** *NODE PRINT, NSET=FakeDefNodes u,tf *Output, field, variable=PRESELECT *node output, nset=FakeDefNodes u *End Step 114 C.5 MATLAB code to generate Abaqus input file The following MATLAB programs take an input file from Abaque CAE, invoke periodic boundary conditions in x (as in Section C.1), and rewrites the original input file to a new file. 115 maininput m written by ashley browning 2010 to make a bunch of input files periodic in x by the program 'nodesortalignrewrite ', location, part name location, new file file which inputs original 0/ %0 when CAE adds a part to an assembly , it calls 'part' -- > 'part-i' can be added in loop below '-1' or any other postscripts this 'input2red ', 'input3red ', if you have 'inputired', simi larly, add 'red' in the loop below you can just clear all; clc; % name of input file, INPUTNAME={'input1', ie: inputl.inp 'input2','input3'}; X name of part in associated input file PARTNAME={'partl', 'part2','part3'}; for i=1:length(INPUTNAME) display(INPUTNAME(i)) nodesortalignrewrite (... char(strcat('I:/AbaqusTemp/',INPUTNAME(i),'.inp')),... char(strcat('E:/Ash/Abaqus/pbc/',INPUTNAME(i),'pbc.inp')),... char(strcat(INPUTNAME{i},'-1'))) end 116 function [] = nodesortalign_2Drewrite3(fname,fnnewpart) % Ashley Browning modified from code by Matthew J. Rosario % do NOT specify any boundary conditions on the part in CAE Z all of the BCs are written in this program. % only specify loading BCs (indentation, pressure, etc) % example: (original file loc, new file loc, part name) % nodesortalign_2Drewrite3('I:\AbaqusTemp\frommodel.inp', 'E:\Ash\Abaqus\pbc\input.inp', 'part -1') % Open file for reading fid1=fopen(fname,'r'); fid =fopen(fnnew,'wt'); if (fid1 % % input output < 0) error('couldunotuopenufileus',fname); end % Find the end of header head = 0; key = '*Node'; while 1 readin = fgetl(fidl); len = length(readin); if len~=5 head = head + 1; fprintf(fid,readin); Zcopy this to new file fprintf(fid,'\n'); Zcopy this to new file elseif and(readin(len)==key(len),len==5) head = head + 1; break elseif head>20 error('headeroidentificationuerror') end end % Store nodes and their positions data = textscan(fidl,'%du%fuf','delimiter',','); node=data{1 ,1}; xpos=data{1,2}; ypos=data{1 ,3}; le-4; rtol = 1/tol; % Round all data to 4 decimal places xpos = round(xpos*rtol)/rtol; ypos = round(ypos*rtol)/rtol; tol = 117 % Find maximum dimensions of prism xmax=max(xpos);xmin=min(xpos); ymax=max(ypos);ymin=min(ypos); % The following procedures sort through the node sets % to seIect nodes % Define the face nodes that are found on faces, edges, and corners XX1=[]; for k=1:length(xpos) if xpos(k)==xmin XX1=[XX1; double(node(k)) xpos(k) ypos(k)]; end end XX1=sortrows(XX1,3); XX2=[]; for k=1 :length(xpos) Yif xpos (k)==xmax if abs(xpos(k)-xmax)<=2*tol XX2=[XX2; double(node(k)) xpos(k) ypos(k)]; end end XX2=sortrows(XX2,3); YY1=[]; for k=1:length(ypos) if ypos(k)==ymin YY1=[YY1; node (k) xpos(k) end end YY1=sortrows(YY1,2); ypos(k)]; YY2=[]; for k=1:length(ypos) if ypos(k)==ymax YY2=[YY2; node (k) xpos(k) ypos(k)]; end end YY2=sortrows(YY2,2); Y make sure the XX1 and XX2 (Left % have identicaL ypos and rewrite 118 and right if not sides) % display this on the screen so you can determine aliasing LengthLeft=length(XX1) LengthRight=length(XX2) maxDiff = max(abs( ypos(XX1(:,1))-ypos(XX2(:,1)))) % error before alignment: (XX1, XX2, difference): [ypos(XX1(:,1)), ypos(XX2(:,1)), ypos(XX1(:,1))-ypos(XX2(:,1))]; % adjust the ypos of XX2 to match that of XX1 for k=1:length(XX2) ypos(XX2(k,1))=ypos(XX1(k,1)); end Y error after alignment: (XX1, XX2, difference): [ypos(XX1(:,1)), ypos(XX2(:,1)), ypos(XX1(:,1))-ypos(XX2(:,1))]; L1=xpos(YY1(length(YY1),1))-xpos(YY1(1,1)); L2=ypos(XX1(length(XX1),1))-ypos(XX1(1,1)); cornernodes=[XX1(length(XX1),1) XX2(length(XX2),1); XX1(1,1) XX2(1,1)1 % Write sorted node numbers to file % fid = fopen(fnnew,'wt'); Z Write new aligned node assigments to the same file fprintf (fid,'*Node\n'); for k=1:length(node) fprintf(fid,'\td,\tf,\tf\n',node (k),xpos(k),ypos(k)); end % find the key = body while end of instance '*EnduInstance'; 0; 1 readin = fgetl(fid1); len = length(readin); if len~=13 body = body + 1; %copy this fprintf(fid,'\n'); Xcopy this elseif and(readin(len)== key (len ) ,len body = body + 1; fprintf(fid,readin); Xcopy this fprintf(fid,'\n'); %copy this break end fprintf(fid,readin); to new file to new file ==1 3) to to end fprintf(fid,'*Node,unset=FakeDefNodesu\n'); 119 new file new file fprintf(fid,'9997,uO.,uO.u\n9998,uO.,uO.u\n'); fprintf(fid,'*Nsetanset=BLNode,uinstance=Xsu\nd,u\n',partXX1(1)); fprintf(fid,'*Nsetunset=BRNodeuinstance=%su\nd,u\n',partXX2(1)); fprintf(fid,'*Nset,unset=XX1,uinstance=%s,uunsortedu\n',part); for k=2:length(XX1) Y starts at k=2 to fprintf(fid,'XAd,\n',XX1(k)); omit BLnode end fprintf(fid,'*Nset,anset=XX2,uinstance=%s,uunsortedu\n',part); for k=2:length(XX2) Z starts at k=2 to fprintf(fid,'X/d,\n',XX2(k)); end omit BRnode fprintf(fid,'*Nset,anset=YY1,uinstance=%s,uunsortedu\n',part); for k=1:length(YY1) fprintf(fid,'Xd,\n',YY1(k)); end fprintf(fid,,'*Equationu\n'); fprintf(fid,'3u\nXX1,u1,u1a\nXX2,u1,u-1u\n9997,u1,u%du\n', fprintf(fid,'3u\nXX1,u2,ulu\nXX2,u2,u-1u\n9997,u2,udu\n', fprintf(fid,'2a\nBRNode,u1,ula\n9997,u1,u-Xdu\n', Li); key = while Li); Li); '*Boundary'; 1 readin = fgetl(fidl); len = length(readin); if len~=9 fprintf(fid,readin); fprintf(fid,'\n'); %copy %copy this this to new file to new file elseif and(readin(len)~=key(len),len==9) fprintf(fid,readin); %copy this to new file fprintf(fid,'\n'); Xcopy this to new file elseif and(readin(len)==key(len),len==9) break end end % No Deformation BCs fprintf(fid,'*Boundaryu\n'); fprintf(fid,'BLNode,u1,u1u\n'); fprintf(fid,'YY1,u2,u2u\n'); fprintf(fid,'9997,u2,u2u\n'); fprintf(fid,'*Boundaryu\n'); % Copy other boundary conditions while 1 readin = fgetl(fidl); len = length(readin); 120 and the rest of the input file if len>1 fprintf(fid,readin); fprintf(fid, '\n'); %copy 7.copy else break end end fclose(fidl); fclose(fid); disp('complete') end 121 this this to to new file new file 122 Bibliography Arciszewski, T., Cornell, J., 2006. Bio-inspiration: Learning creative design principia, in: Smith, I. (Ed.), Intelligent Computing in Engineering and Architecture, Springer, Berlin, Germany. pp. 32 - 53. Breder, C.M., 1926. The locomotion of fishes. Zoologica 4, 159-297. Bruet, B.J.F., Song, J., Boyce, M.C., Ortiz, C., 2008. Materials design principles of ancient fish armour. Nature Materials 7, 748-756. Burdak, V.D., 1979. Morphologie fonctionnelle du t6gument 6cailleux des poissons. Kiev: La Pense Scientifique (in Russian). French translation, Cybium 10, 1986, 1-147. Colbert, E.H., Morales, M., 2001. Colbert's Evolution of the Vertebrates: A History of the Backboned Animals Through Time. Wiley-Liss. Connors, M.J., Ehrlich, H., Hog, M., Godeffroy, C., Araya, S., Kallai, I., Gazit, D., Boyce, M.C., Ortiz, C., 2012. Three-dimensional structure of the shell plate assembly of the chiton Tonicella marmorea and its biomechanical consequences. Journal of Structural Biology 177, 314-328. Danielsson, M., Parks, D.M., Boyce, M.C., 2002. Three-dimensional micromechanical modeling of voided polymeric materials. Journal of the Mechanics and Physics of Solids 50, 351379. Evans, H.E., 1994. Anatomy of Tropical Fishes, in: Gratzek, J.B., Matthews, J.R. (Eds.), Aquariology: The Science of Fish Health Management. Tetra Press, pp. 1-8. Francis, R.I.C.C., 1990. Back-calculation of fish length: a critical review. Journal of Fish Biology 36, 883-902. Garrano, A.M.C., Rosa, G.L., Zhang, D., Niu, L.N., Tay, F., Majd, H., Arola, D., 2012. On the mechanical behavior of scales from Cyprinus carpio. Journal of the Mechanical Behavior of Biomedical Materials 7, 17 - 29. Guinan, M.J., . Anatomy 100: Integument, University of California at Davis Lecture. http://guppiesonli.wordpress.com/guppy-articles/ skin-structure-of-guppy/. Last accessed May 1, 2012. 123 Hatjasalo, L., Rinko, K., 2006. Elastic composite structure. US Patent 7153789. Hawkes, J.W., 1974. Structure of fish skin 1. general organization. Research 149, 147-158. Cell Tissue Hebrank, M.R., 1980. Mechanical properties and locomotor functions of eel skin. The Biological Bulletin 158, 58. Hebrank, M.R., Hebrank, J.H., 1986. The mechanics of fish skin: lack of an "external tendon" role in two teleosts. The Biological Bulletin 171, 236. Heymsfield, S., Gallagher, D., Kotler, D., Wang, Z., Allison, D., S., H., 2002. Bodysize dependence of resting energy expenditure can be attributed to nonenergetic homogeneity of fat-free mass. American Journal of Physiology - Endocrinology and Metabolism 282, E132-8. Ibafiez, A.L., Cowx, I.G., O'Higgins, P., 2009. Variation in elasmoid fish scale patterns is informative with regard to taxon and swimming mode. Zoological Journal of the Linnean Society 155, 834-844. Ikoma, T., Kobayashi, H., Tanaka, J., Walsh, D., Mann, S., 2003. Microstructure, mechanical, and biomimetic properties of fish scales from Pagrus major. Journal of Structural Biology 142, 327-333. Jawad, L.A., Al-Jufaili, S.M., 2007. Scale morphology of greater lizardfish Saurida tumbil (Bloch, 1795)(Pisces: Synodontidae). Journal of Fish Biology 70, 1185-1212. Kim, S., Su, Y., Mihi, A., Lee, S., Liu, Z., Bhandakkar, T.K., Wu, J., Geddes III, J.B., Johnson, H.T., Zhang, Y., Park, J.K., Braun, P.V., Huang, Y., Rogers, J.A., 2012. Imbricate scales as a design construct for microsystem technologies. Small 8, 901-906. Lin, Y.S., Wei, C.T., Olevsky, E.A., Meyers, M.A., 2011. Mechanical properties and the laminate structure of Arapaima gigas scales. Journal of the Mechanical Behavior of Biomedical Materials 4, 1145-1156. Lindsey, C.C., 1979. Form, Function, and Locomotory Habits in Fish. Fish Physiology 7, 1-100. Long, J.H., Hale, M.E., McHenry, M.J., Westneat, M.W., 1996. Functions of fish skin: Flexural stiffness and steady swimming of longnose gar Lepisosteus osseus. Journal of Experimental Biology 199, 2139-2151. Meunier, F., Brito, P., 2004. Histology and morphology of the scales in some extinct and extant teleosts. Cybium 28, 225-235. Meyers, M., Lin, Y.S., Olevsky, E.A., Chen, P.Y., 2011. Arapaima vs. Piranha. Advanced Biomaterials , 1-13. 124 Battle in the Amazon: Neal, M.L., Bain, A.D., 2004. Method and apparatus for defeating ballistic projectiles. US Patent 6745661. Oliver, W.C., Pharr, G.M., 1992. Improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. Journal of Materials Research 7, 1564-1583. Ortiz, C., Boyce, M.C., 2008. Bioinspired structural materials. Science 319, 10531054. Park, E.H., Lee, S.H., 1988. Scale growth and squamation chronology for the laboratory-reared hermaphroditic fish Rivulus marmoratus (Cyprinodontidae). Japanese Journal of Ichthyology 34, 476-482. Sire, J.Y., 1990. From ganoid to elasmoid scales in the actinopterygian fishes. Netherlands Journal of Zoology 40, 75-92. Sire, J.Y., Donoghue, P.C.J., Vickaryous, M.K., 2009. Origin and evolution of the integumentary skeleton in non-tetrapod vertebrates. Journal of Anatomy 214, 409440. Smith, F., 2012. Integument, in: Anatomy 3421, Dalhousie University, Department of Anatomy and Neurobiology. pp. 1-35. Song, J., Ortiz, C., Boyce, M.C., 2011. Threat-protection mechanics of an armored fish. Journal of the Mechanical Behavior of Biomedical Materials 4, 699-712. Song, J., Reichert, S., Kallai, I., Gazit, D., Wund, M., Boyce, M.C., Ortiz, C., 2010. Quantitative microstructural studies of the armor of the marine threespine stickleback (Gasterosteus aculeatus). Journal of Structural Biology 171, 318-331. Torres, F.G., Troncoso, O.P., Nakamatsu, J., Grande, C.J., Gomez, C.M., 2008. Characterization of the nanocomposite laminate structure occurring in fish scales from Arapaima gigas. Materials Science and Engineering C: Materials for Biological Applications 28, 1276-1283. United States Fish and Wildlife Service, 2010. Normal trout histology. http: // training.fws.gov/ec/fish/histol.html. Last accessed May 1, 2012. Vermeij, G.J., 1983. Shell-breaking predation through time. Biotic interactions in recent and fossil benthic communities. Plenum, New York , 649-669. Vernerey, F.J., Barthelat, F., 2010. On the mechanics of fishscale structures. International Journal of Solids and Structures 47, 2268-2275. Whitear, M., 1986. The skin of fishes including cyclostomes: epidermis, in: BereiterHahn, J., Matoltsy, G.A., Richards, S.K. (Eds.), Biology of the Integument, Springer, Berlin, Germany. pp. 8 - 38. 125 Whitenack, L.B., Simkins Jr., D.C., Motta, P.J., 2010. Biology meets engineering: The structural mechanics of fossil and extant shark teeth. Journal of Morphology 272, 169-179. Zhu, D., Ortega, C.F., Motamedi, R., Szewciw, L., Vernerey, F., Barthelat, F., 2012. Structure and mechanical performance of a "modern" fish scale. Advanced Engineering Materials 14, B185 - B194. 126