A Split Hopkinson Pressure Bar Apparatus for High Strain Rate Testing of Interpenetrating Phase Composites (IPC): Measurements and Modeling by Chandru Periasamy A thesis submitted to the Graduate Faculty of Auburn University in partial fulfillment of the requirements for the Degree of Master of Science Auburn, Alabama May 14, 2010 Keywords: structural foams, dynamic failure, interpenetrating phase composites, syntactic foam, split Hopkinson pressure bar, finite element analysis Approved by Hareesh V Tippur, Chair, Professor of Mechanical Engineering Jeffrey C Suhling, Quina Distinguished Professor of Mechanical Engineering James S Davidson, Associate Professor of Civil Engineering Abstract Novel materials with enhanced mechanical performance and multifunctionality are of interest to automotive, aerospace and marine industries alike. Designing materials with multiple thermo-mechanical attributes while satisfying lightweight constraint is rather challenging. This challenge becomes even greater when materials are to withstand elevated rates of loading. In this thesis, (i) a 3D Interpenetrating Phase Composite (IPC) made of Syntactic Foam (SF) in an open-cell aluminum scaffold is proposed for high- strain rate applications and (ii) a split-Hopkinson pressure bar apparatus is developed for characterizing low-impedance materials such as the SF and IPC materials at high-strain rates. The SF used for making IPC is prepared by dispersing hollow glass microballoons into an epoxy matrix. The resulting SF is a two phase composite on a microscopic scale and a lightweight homogeneous, isotropic material on a macro scale. The infusion of SF in its uncured state into the aluminum scaffold results in an IPC with a continuous 3D interpenetrating network throughout the material volume. A split Hopkinson pressure bar has been developed and calibrated for carrying out high-strain rate stress-strain response measurements on low impedance materials. Dynamic compression characteristics, including strength and energy absorption features of IPC made of four different volume fractions of microballoons in SF are measured using this apparatus and compared. Measurements on pure SF counterparts are also carried out for evaluation relative to IPC. The results show that in general, the IPC foams ii iii outperform the SF samples in terms of compressive strength and energy absorption per unit volume. The underlying deformation processes are evaluated optically using real- time high-speed photography as well as post-mortem analysis of deformed samples using scanning electron microscopy. The former includes the development of a novel grid- based method for measuring surface displacements and strains. An idealized elasto-plastic unit cell finite element (FE) model based on explicit dynamics is proposed for studying IPC foams. The 3D geometry of the aluminum component in the cubic unit cell is modeled as a tetrakaidecahedron, a 14 sided polyhedron, also called the Kelvin cell. The SF constituent of the IPC is modeled as an occupant of the rest of the unit cell. The computational model incorporates infinite elements to represent the far-field regions surrounding the unit cell. The infinite elements help simulate stress wave dynamics of a sample undergoing deformation in a split Hopkinson pressure bar. The computations are successfully compared with measurements for all the four cases. The contour plots of von-Mises stress, equivalent plastic strain and displacements in the loading direction are analyzed at different strain levels. In addition, the kinetic and strain energy histories absorbed by SF and aluminum ligaments in the IPC are quantified. Acknowledgments I thank my advisor, Dr. Hareesh V. Tippur, for his financial support and technical guidance without which my work would not have been a success. I consider the time I spent in the ?Failure characterization and optical techniques? laboratory at Auburn University, USA under Dr. Tippur?s supervision to be of immense value in advancing my technical education. I thank Dr. Jeffrey Suhling and Dr. James Davidson for their willingness to be on my MS committee and for reviewing this thesis. I thank the U.S. Army Research Office (ARO) for funding this research through grants W911NF-08-1- 0285 and W911NF-06-1-0126 (DURIP). Also, this work was made possible in part by a grant of high performance computing resources and technical support from the Alabama Supercomputer Authority. It would be unfair on my part not to acknowledge and thank Dr. Robert Jackson, as without his support and help my induction into Auburn University would not have been smooth. It was the opportunity to work on his team that introduced me to technical research and helped me develop my modeling skills. Dr. Dongyeon Lee, Kailash Jajam, Vinod Kushvaha, Allen Craven and Robert Bedsole deserve mention here as my good colleagues who made my experience inside and outside the laboratory stimulating, pleasant and eventful. I thank Rahul Jhaver for sharing his experimental knowledge and results. My parents, brother and grandmother have been my greatest believers and supporters in all of my aspirations. I dedicate this thesis to them. iv Table of Contents Abstract ............................................................................................................................ ii Acknowledgements ......................................................................................................... iv List of Tables ................................................................................................................ viii List of Figures ................................................................................................................. ix 1. Introduction ...................................................................................................................1 1.1 Motivation ..............................................................................................................1 1.2 The syntactic foam / aluminum foam based IPC ...................................................3 1.3 Objectives ...............................................................................................................5 1.4 Organization of the thesis .......................................................................................6 2. Literature review ...........................................................................................................8 2.1 Split Hopkinson Pressure Bar (SHPB) ...................................................................8 2.2 Previous works on IPC .........................................................................................10 3. Material preparation and sample dimensions .............................................................13 3.1 Material preparation .............................................................................................13 3.2 Sample dimensions ...............................................................................................16 4. Experimental procedure ..............................................................................................18 4.1 Experimental setup ...............................................................................................19 4.2 Location of strain gages .......................................................................................22 4.3 Measurement of strain ..........................................................................................23 v 4.4 Stress, strain and strain rate ..................................................................................24 5. Evaluation of the setup ................................................................................................27 5.1 Dynamic equilibrium and constant strain rate ......................................................27 5.2 Effect of pulse shapers and striker length ............................................................31 5.3 SHPB calibration ..................................................................................................33 6. Experimental results ....................................................................................................34 6.1 Quasi-static compressive stress-strain response ...................................................34 6.1.1 Syntactic foam ...........................................................................................34 6.1.2 IPC .............................................................................................................37 6.2 Dynamic compressive stress-strain response .......................................................39 6.2.1 Syntactic foam ...........................................................................................40 6.2.2 IPC .............................................................................................................43 6.3 Energy absorption .................................................................................................46 6.3.1 Quasi-static ................................................................................................46 6.3.2 Dynamic .....................................................................................................48 6.3.3 Dynamic energy absorption history ...........................................................50 6.4 Failure progression ...............................................................................................51 7. Finite element modeling and analysis of IPC foam ....................................................56 7.1 Geometric modeling .............................................................................................56 7.2 Computational model ...........................................................................................58 7.2.1 Material properties .....................................................................................58 7.2.2 Yielding and hardening ..............................................................................59 7.2.3 Explicit integration scheme in ABAQUS ..................................................63 vi vii 7.3 Discretization ........................................................................................................64 7.3.1 Infinite elements ? Background .................................................................67 7.4 Boundary conditions and loading .........................................................................68 7.5 Other analysis parameters ....................................................................................70 7.6 Results from finite element analysis ....................................................................70 7.6.1 Stress-strain response .................................................................................70 7.6.2 Comparison between experimental and finite element results ..................71 7.6.3 Analysis of contour plots ...........................................................................73 7.6.4 Energy absorption ......................................................................................78 8. Summary .....................................................................................................................81 References .....................................................................................................................85 Appendix A ...................................................................................................................92 List of Tables Table 6.1 Quasi-static properties of syntactic foam and IPC foam samples ..................39 Table 6.2 Dynamic properties of syntactic foam and IPC foam samples .......................42 Table 7.1 Elastic properties of SF and aluminum used in the finite element model ......59 Table 7.2 Inelastic properties of Al 6101 used in the finite element model ...................59 viii List of Figures Figure 1.1 Cross section view of a human bone ...............................................................2 Figure 1.2 Syntactic foam used in submersible vehicles in marine environments ...........4 Figure 1.3 (a) Open-cell metal foams made of different materials, (b) Closed-cell aluminum alloy foam ......................................................................................4 Figure 1.4 Schematic of a 3D interpenetrating network ...................................................5 Figure 2.1 Hopkinson?s setup ...........................................................................................9 Figure 3.1 (a) Rubber mold with uncured SF (b) Al foam inserted into mold containing uncured SF (c) Cast and machined SF and IPC samples ...........14 Figure 3.2 (a) Cross-section of IPC foam with 30% volume fraction of SF, (b) SEM micrograph of IPC.........................................................................................15 Figure 3.3 (a) Representation of pores and cells, (b) Sample dimensions ......................16 Figure 4.1 Photographs of the actual SHPB setup used for dynamic compression experiments ...................................................................................................18 Figure 4.2 Split Hopkinson pressure bar setup ...............................................................20 Figure 4.3 Schematic representation of the locations of incident, reflected and transmitted waves in a SHPB setup ..............................................................23 Figure 4.4 Schematic illustration of the SHPB ...............................................................24 ix Figure 5.1 Representative SHPB strain histories for (a) Syntactic foam sample with 30% microballoon volume fraction, (b) IPC foam sample with 30% microballoon volume fraction .......................................................................28 Figure 5.2 Demonstration of constant strain-rate for (a) SF-30 and (b) IPC-10 and dynamic stress equilibrium for (c) SF-30 and (d) IPC-10 ...........................30 Figure 5.3 Effect of pulse shapers and repeatability on SF-30 samples .........................31 Figure 5.4 Effect of striker length on SF-30 samples .....................................................32 Figure 5.5 True stress-strain response for cast acrylic at a strain rate of ~ 1250 / sec ...33 Figure 6.1 Effect of microballoon volume fraction on quasi-static compression response of syntactic foam (SF) samples ......................................................35 Figure 6.2 SEM images of 30% V f syntactic foam at (a) 14% strain and (b) 58% strain showing crushing of microballoons at various stages of static deformation ...................................................................................................36 Figure 6.3 Effect of microballoon volume fraction on quasi-static compression response of IPC foam samples ......................................................................37 Figure 6.4 SEM image of statically deformed IPC foam at 58% strain .........................38 Figure 6.5 Effect of microballoon volume fraction on dynamic stress-strain response of SF samples ................................................................................................41 Figure 6.6 (a) Side view of a dynamically deformed syntactic foam sample revealing a network of shear bands, (b) SEM image of cross section of dynamically deformed syntactic foam ...............................................................................43 Figure 6.7 Effect of microballoon volume fraction on dynamic stress-strain response of IPC foam samples .....................................................................................44 x Figure 6.8 (a) Side view of a dynamically deformed IPC foam sample revealing multiple shear bands and (b) SEM image of cross section of dynamically deformed IPC foam .......................................................................................45 Figure 6.9 Energy absorbed (a) per unit volume and (b) per unit mass by SF and IPC foam samples under quasi-static loading up to 22% true strain ....................47 Figure 6.10 Energy absorbed (a) per unit volume and (b) per unit mass by SF and IPC foam under dynamic loading up to 22% true strain .............................49 Figure 6.11 Strain energy absorption history of (a) SF and (b) IPC with increasing strain in SHPB tests ......................................................................................51 Figure 6.12 Schematic of the experimental setup used for high-speed photography to study real-time failure progression in SF and IPC samples in the SHPB set up ............................................................................................................52 Figure 6.13 Stages of dynamic deformation of SF-30 ....................................................53 Figure 6.14 Stages of dynamic deformation of IPC-30 ..................................................55 Figure 7.1 Schematics representing the building blocks of an idealized Kelvin cell based IPC ......................................................................................................57 Figure 7.2 Check for appropriateness of discretization ..................................................65 Figure 7.3 Illustration showing (a) the infinite element layer in the FE model and stress wave behavior in the (b) absence and (c) presence of infinite element layer .................................................................................................66 Figure 7.4 Load and boundary conditions ......................................................................69 Figure 7.5 True stress?true strain responses of different IPC obtained from finite element analysis ...........................................................................................71 xi xii Figure 7.6 Comparison of experimental and FEA results of (a) IPC-10, (b) IPC-20, (c) IPC-30 and (d) IPC-40.............................................................................72 Figure 7.7 Mises stress contour plots from dynamic FE model for IPC-30 ...................74 Figure 7.8 Equivalent plastic strain contour plots from dynamic FE model for IPC-30 ...........................................................................................................75 Figure 7.9 x-displacement contour plots from dynamic FE model for IPC-30 ..............77 Figure 7.10 History of (a) kinetic energy and (b) strain energy absorption by aluminum and SF in unit-cell IPC-30 ..........................................................78 Figure 7.11 History of percentage of (a) kinetic energy and (b) strain energy absorption by SF and aluminum in IPC-30 .................................................79 Figure A1 Quasi-static and dynamic compressive responses of neat epoxy (?Epo- Thin?) used in SF and IPC foams .................................................................92 1 INTRODUCTION 1.1 Motivation The continued human quest for making lives more comfortable has increased demands on engineering and technology. Lately, researchers have been looking towards nature for inspiration to meet some of these demands such as those involving materials and structures. The field of science and engineering which focuses on studying and researching bio-inspired materials and structures is called biomimetics [1]. Naturally occurring structural materials have undergone millions of years of evolutionary processes to adapt to the changing needs of their environment and functionality. So, it is only wise to adopt some of these examples in the process of developing new structural materials. Recent advances in commercial and defense automotives, as well as marine construction, require innovative lightweight materials that could serve in the design of stiff and tough structures capable of absorbing impact energy. This concurrence of material properties cannot be easily formulated into a single material system. The microstructure of some naturally occurring materials such as bone (Fig 1.1) and nacre have a 3D interpenetrating structure [2], often made up of two or more biological phases. Thinking along these lines, and considering the combined requirement of structural integrity and energy absorption, an Interpenetrating Phase Composite (IPC) [3] made of two phases ? one, brittle Syntactic Foam (SF) and the other, ductile aluminum (Al) foam 1? ? is being proposed in this thesis work. In general, IPCs are multiphase materials in which each constituent forms a continuous 3D network within the material volume. Thus, each phase in its stand alone state can be imagined to have an open-cellular microstructure. Figure 1.1: Cross section view of a human bone (From www.health.state.ny.us/diseases/conditions/osteoporosis/needknow.htm) Even though IPCs are heterogeneous on a micro/meso scale, the macro scale response is often isotropic. Each phase offers its own unique property to the IPC as a whole. For example, one phase might offer good toughness and thermal conductivity, while the other might enhance stiffness and dielectric properties. This co-existence of desirable properties without significant directional dependency or distinct weak planes (as opposed to conventional layered composites) makes an IPC attractive for structural applications. That is, each phase of an IPC contributes its unique property to the overall structural response synergistically. The above mentioned possibilities invite an indepth study on the IPC. 2? ? 1.2 The syntactic foam / aluminum foam based IPC In this work, the IPC is made of two phases ? a polymer based Syntactic Foam (SF) and an open-cell aluminum (Al) foam. SF is a low density composite comprised of epoxy (or, any polymer) and hollow glass (or, any ceramic) microballoons. The microballoons are microscopic hollow spheres, usually a few micrometers to several hundred microns in diameter, and wall thickness ranging from fraction of a micron to a few microns. When these two materials are mechanically blended and cured, they form a two phase composite on a micro scale. Considering the size of the microballoons and their uniform distribution within the epoxy matrix, the resulting composite can be regarded as an isotropic material on a macro scale. The microballoons are hollow (with a partial pressure) which makes the syntactic foam lighter as well as buoyant. This unique combination of low density and buoyancy makes SF an ideal choice for marine applications (Fig. 1.2). Metal foams, on the other hand, are broadly classified into open- cell (Fig. 1.3(a)) and closed-cell (Fig. 1.3(b)) foams. They are typically used for energy absorption and heat dissipation applications. The closed-cell foams in particular are used as layers in sandwiched sheets for energy absorption. The open-cell foams, on the other hand, are also used for other applications [2] such as cryogenics, filteration, missile baffles, etc. Due to their relatively high cost, metal foams are only used in areas of advanced technology, such as aerospace structures. 3? ? Figure 1.2: Syntactic foam used in submersible vehicles in marine environments (From www.naval-technology.com) 4? ? between the two phases is occupied by one of the two phases.) It can be seen that each (a) (b) Figure 1.3: (a) Open-cell metal foams made of different materials (from www.ergaerospace.com); (b) Closed-cell aluminum alloy foam (from Reference [4]) The IPC used in this work is made by infusing the SF into the open-cell aluminum foam. The term ?Interpenetrating? signifies that each phase of the composite is continuous and forms a network of its own. Figure 1.4 shows a schematic of an interpenetrating network. The region shown in red (gray) represents one phase of the network, and the region shown in white is the second phase. (In reality, the empty space phase is three dimensionally continuous and also forms an interpenetrating network. A material of this type can prove to be advantageous in many applications as it can offer the combined properties of the two different materials synergistically. For example, in the SF / aluminum foam IPC, the aluminum foam might add to the stiffness of the whole structure, whereas the aluminum foam by itself does not have appreciable stiffness. Figure 1.4: Schematic of a 3D interpenetrating network 1.3 Objectives wing are the primary objectives of this work: ? Dev up for characterizing the h different volume fractions (V f ) of microballoons in dynamic compression properties, viz., stress-strain response, compressive strength and energy absorption of SF and IPC foams using the SHPB. The follo elop a split Hopkinson pressure bar (SHPB) set compression behavior of low impedance materials and calibrate it for accuracy and repeatability of measurements. ? Prepare SF and IPC samples wit SF. ? Test 5? ? ? Comparatively study the dynamic compression properties of SF and IPC and the effect of microballoon volume fraction V f in the SF on the overall performance of erparts. erstand the deformation processes in SF and IPC materials. This thesis has been structured in such a way that the experimental aspects are putational work is detailed later on. The next chapter present IPC. ? Evaluate the high-strain rate performance of SF and IPC relative to the quasi-static count ? Perform high-speed photography and microstructural investigation of the deformed samples to und ? Develop a finite element model of the IPC and simulate dynamic compression response using the measured mechanical responses of its constituents. ? Validate the computational model by comparing the computed high-strain rate IPC responses with the measurements. 1.4 Organization of the thesis presented first and then the com s a literature review on (a) the Split Hopkinson Pressure Bar (SHPB) setup used for high-strain rate experimental characterization of materials and (b) the previous works on IPC. Chapter 3 describes material preparation methods for both SF and IPC foams. The experimental procedure and the evaluation of the SHPB apparatus, respectively, are presented in Chapters 4 and 5. The results of the experiments are detailed and analyzed in Chapter 6. For the purpose of comparison of the strain rate effects, the appropriated quasi-static results from a previous work [2] are also presented intermittently throughout this chapter. Chapter 7 presents the finite element procedure, including details on 6? ? 7? ? geometric modeling, relevant material properties, model discretization and application of boundary conditions. A comparison of the experimental and computational results is also presented in this chapter. Finally, Chapter 8 summarizes the results of this work. 2 LITERATURE REVIEW 2.1 Split Hopkinson Pressure Bar (SHPB) A split Hopkinson pressure bar is a device used to study the mechanical behavior of a material at high-strain rates. The principle behind the SHPB was first introduced by Bertram Hopkinson [5] in 1914. Hopkinson used the principle to measure the pressure developed in a bar when a blast occurs at one end of it. In his set up (Fig. 2.1), a bar was suspended with one end attached to a detonation mechanism and the other end in magnetic contact with a time piece. When a blast occurs at one end, a pressure pulse travels through the suspended bar and impacts the time-piece. At the free end of the time piece, the pressure pulse reflects back as a tensile pulse, and when it reaches the time piece ? bar interface, the time piece separates from the bar and flies into a momentum trapping mechanism. The energy recorded by the momentum trap corresponds to the energy of twice the wave length of the pressure pulse in the time piece. Subsequently, Hopkinson?s principle was used by Robertson [6] in 1921 and by Landon and Quinney [7] in 1923. The technique was then modified in 1948 by Davies [8] who used condensers for measuring displacements in the pressure bar. In 1949, Kolsky [9] modified the setup further by adding a second pressure bar and placed the specimen between the two bars. He used a cylindrical condenser microphone to measure the amplitude of the pressure pulse produced by firing a detonator at the free end of the 8? ? incident bar. At the free end of the second bar, he used a parallel plate condenser microphone to measure displacement. He used the signals from the two condensers to measure the stress-strain response of the specimen. Most recently, researchers have replaced the detonator with a striker impacting the incident bar (at a known velocity) and the condensers with strain gages pasted on the pressure bars. Momentum Trap Gun Cotton Cylinder Bar Time Piece Recorder Suspension Rails Figure 2.1: Hopkinson?s Setup (Reproduced from reference [5]) The invention of the split Hopkinson bar has helped researchers to dynamically test a variety of solids including metals, polymers, ceramics and composites [10]. Modifications to the conventional Kolsky bar have been made by many researchers to 9? ? test brittle materials [11, 12], low-impedance or soft materials [13, 14, 15] and structural foams [16]. To test low-impedance materials at high strain rates, Chen et al. [13] used a hollow high-strength aluminum alloy for the bar material in place of the conventionally used solid circular cross sectioned steel bars. It has been reported in the same work that the reduced cross sectional area of the hollow aluminum bars increased the amplitude of the transmitted strain signal in the bars by an order of magnitude. Song and Chen [16] obtained the dynamic compressive response of epoxy syntactic foam using a SHPB setup in which they used an annealed copper pulse shaper at the striker-incident bar interface. In their work, it has been demonstrated that the specimens were under dynamic stress equilibrium during loading due to the effect of the pulse shaper. Recently, Subhash and Liu [17] tested an epoxy-based polymeric foam for its crushability under rigid lateral confinement. In their work, they used magnesium bars to test samples under low stress, and steel bars for samples under high stress. Thus, it is evident that material properties (density in particular) play a significant role in the stress wave history in the bars. Likewise, Bouix, et al. [18] used nylon bars in their SHPB setup to dynamically characterize the compression properties of polypropylene foams. It is to be noted that nylon which is visco-elastic, was the choice material for testing a low impedance rubbery material. 2.2 Previous works on IPC The primary material system under investigation in this work is an IPC, which is a two phase composite. The following is a review of some of the works on different IPC systems available in the literature. Among the existing works on IPCs, Prielipp et al. [19] 10? ? studied the mechanical properties of aluminum/alumina interpenetrating composite and measured fracture strength and fracture toughness of the composite as a function of ligament diameter and volume fraction of the metal reinforcement. The metal reinforced interpenetrating composites consistently had higher fracture strength. Breslin et al. [20] characterized an aluminum/alumina IPC obtained using the liquid phase displacement reaction method. The resulting IPC showed enhanced density, thermal conductivity and CTE characteristics, without compromising stiffness or fracture toughness. Travitzky et al. [21] showed that the residual stresses developed in silicon during solidification of molten Si within an Al 2 O 3 matrix results in the greater strength and fracture toughness of the resulting Al 2 O 3 /Si interpenetrating system. Wegner and Gibson [22] developed finite element models to predict the elastic, strength and thermal expansion properties of two phase IPCs. They attributed the enhancement in the thermo-mechanical properties to the contiguity of the phase with the most desirable property. In a later work [23], the same authors studied the mechanical behavior of resin-impregnated porous stainless steel. The yield strength, ultimate strength, elongation at failure and the elastic modulus were all found to increase with an increasing volume fraction of steel. Skirl et al. [24] examined the thermal expansion behavior of alumina/aluminum IPC. A pressure infiltration technique was used to introduce aluminum into slip cast and then sintered alumina. The tensile and compressive residual stresses in alumina and aluminum phases were found to enhance the overall thermal coefficient of expansion. An increase in failure strain with increased metal content was also reported. Mayer and Papakyriacou [25] studied the fatigue behavior of graphite/aluminum IPC. The lightweight metals, such as aluminum, were infiltrated into polycrystalline graphite to improve the fracture toughness of 11? ? 12? ? polycrystalline graphite. A 30% increase in the cyclic strength and a 10% increase in the endurance limit were also reported. Tilbrook et al. [26] measured the effective mechanical properties of alumina-epoxy interpenetrating composites and reported strong dependence of properties on the composition and the processing of the material itself. Etter et al. [27] examined the flexural strength and fracture toughness of graphite/aluminum IPC at room temperature and at 300 o C. A 200% improvement in both these characteristics for IPC over un-infiltrated material at room temperature was reported. Additionally, no significant drop in properties was seen at elevated temperatures. Jhaver and Tippur [28] demonstrated the feasibility of a lightweight IPC foam by infiltrating a syntactic foam (SF) [29-31] into an open-cell aluminum network and examining quasi-static uniaxial compression response experimentally and numerically. They showed that the compressive strengths and energy absorption capacities of IPC samples were greater than those of the corresponding syntactic foam samples. As noted in the above reports, the enhancements in the effective mechanical properties of IPC are dictated by the contiguity of the phase with the most desirable properties. An IPC foam formed by a continuous interpenetrating 3D network of syntactic and aluminum foams has the potential for improving structural integrity with good energy absorption characteristics, thus making it a worthwhile material system for further study. This work builds on the feasibility study of the IPC foam first reported in Ref. [28]. 3 MATERIAL PREPARATION AND SAMPLE DIMENSIONS This chapter details the preparation of the IPC foam material and the choice of dimensions of the samples. In the first section, the sequential procedure adopted for the preparation of the SF samples is described, followed by a description of the preparation of the IPC samples. Later, a microscopic view of an IPC sample cross section is discussed with regard to the bonding between SF and aluminum and the randomness of the microballoon distribution. Lastly, a discussion on the sample dimensions and the reason for choosing those dimensions will be presented. 3.1 Material preparation The constituents used for preparing the SF foam were a low viscosity epoxy (Epo- Thin TM from Beuhler, Inc. USA, mass density of the resin ~1100 kg/m 3 ) and hollow glass microballoons (K-1 TM microballoons from 3M Corp., bulk density 125 kg/m 3 ) of average diameter of ~60 ?m and wall thickness of ~0.6. Syntactic foam samples were prepared by mixing the desired quantity of microballoons into epoxy resin. The uncured syntactic foam was then transferred into a silicone rubber mold (Fig. 3.1 (a)) after vacuuming (at approximately -75 kPa gage pressure) to remove any trapped air bubbles in the mixture. The mixture was then allowed to cure for at least seven days before being removed from the mold and machined to size as shown in Fig. 3.1 (c). 13? ? Al Foam Uncured SF Rubber mold (a) (b) Cured SF Cured IPC Machined SF sample Machined IPC sample Machined IPC sample (c) Figure 3.1: (a) Rubber mold with uncured SF, (b) Aluminum foam inserted into mold containing uncured SF, (c) Cast and machined SF and IPC samples The metal foam used in the IPC was a commercially available open-cell Duocel ? aluminum (Al 6101-T6) preform obtained from ERG Aerospace Corp., with a pore density of 40 pores per inch (~8% relative density). The metal foam was cleaned with 14? ? acetone and then coated with silane, ?-aminopropyltrimethoxysilane (H 2 NC 2 H 4 NHC 3 H 6 Si(OCH 3 ) 3 ), to enhance the bond strength between aluminum ligaments and syntactic foam. To prepare the IPC, the degassed, uncured SF was prepared the same way as before. Then, the silane coated aluminum foam was slowly inserted into the mold (as shown in Fig. 3.1 (b)) containing uncured syntactic foam so that the syntactic foam filled the open-pores of the aluminum network. After curing for at least seven days, the unfinished sample was removed from the mold for machining (Fig. 3.1 (c)). Figure 3.2: (a) Cross-section of IPC foam with 30% volume fraction of SF, (b) SEM micrograph of IPC; (From Ref. [28]) A photograph of the cross-section of the cured IPC foam is shown in Fig. 3.2(a). The shiny metallic ligaments of the aluminum preform interspersed within the syntactic foam can be readily observed. Figure 3.2(b) shows a micrograph of the cross section of the IPC with 30% microballoon volume fraction in SF. The interface between aluminum ligaments and syntactic foam is well defined without voids or cracks suggesting a good bond between the two phases. The microballoons also appear to be uniformly distributed within the epoxy matrix. 15? ? 3.2 Sample dimensions The cured SF and IPC samples were removed from the mold and rested for about a week before machining. The sample dimensions were determined based on a previous work by Gibson [32]. Gibson has analyzed size effects on foam test specimens, and proposed a specimen thickness (length) to cell size ratio (t/d) of 8 or above, so that the specimen better represents the bulk. By size effects, Gibson means the effect on the measured material properties when the macroscopic sample dimensions are of the order of the pore size. The aluminum foam used in this thesis work has a pore density of 40 pores per inch with 22 ? 26 cells per inch according to the manufacturer. Therefore, it can be assumed that there are approximately 24 cells per inch on an average (Figure 3.3(a) shows the relative sizes of pores and cells). Thus, the individual cell size is approximately 1.06 mm (0.042 in). A t/d ratio of 9 was chosen for the SHPB tests. Accordingly, the samples used in the dynamic testing had a length of 9.5 mm (0.375 in) and a diameter of 12.7 mm (0.5 in) as shown in Fig. 3.3(b). 12.7 mm 9.5 mm Figure 3.3: (a) Representation of pores and cells (From www.ergaerospace.com); (b) Sample dimensions (a) (b) 16? ? 17? ? From the dimensions shown in Fig. 3.3(b), the thickness (t) to diameter (D) ratio (or, slenderness ratio S 0 ) is 0.75. This S 0 was chosen with reference to the work published by Malinowski and Klepczko [33] in which they suggested an optimum value for use in SHPB tests for different material categories based on the axial inertia and interfacial friction of the sample material. According to their work, samples made of low flow stress materials such as gold and uranium should have 0.1 ? S 0 ? 0.5. For samples made of moderate density metals like aluminum and copper they suggest an optimum S 0 range of 0.5 to 1.0, and for hard materials like high strength steels, an optimum S 0 range of 1.0 to 1.5. The IPC used in this work consists of aluminum and SF, which are made of moderately dense and brittle, but not hard materials. Logically, the integrated effect of combining the two materials in the IPC suggests a S 0 in the 0.5 to 1.0 range for the SHPB test samples. 4 EXPERIMENTAL PROCEDURE The dynamic experiments were performed using a Split Hopkinson Pressure Bar [9] (SHPB) apparatus (Fig. 4.1) developed for this study. This chapter provides a detailed description of the experimental setup, the working principle and the theoretical basis for obtaining the dynamic stress-strain response of SF and IPC samples. (a) Pressure cylinder Pressure gage Air inlet I-beam Bronze bearing (b) Gun barrel Solenoid valve Striker Incident bar 18? ? (c) Oscilloscope Signal conditioning unit Location of strain gage Figure 4.1: Photographs of the actual SHPB setup used for dynamic compression experiments 4.1 Experimental setup Figures 4.1 and 4.2 show photographs and a schematic of the SHPB setup. It consists of an air pressure cylinder with a barrel, a striker (25.4 mm diameter and 406 mm long), a pulse shaper, an incident bar (25.4 mm diameter and 2438 mm long), a transmitter bar (25.4 mm diameter and 2438 mm long), a stopper, two strain gages (one on each bar), a signal conditioning unit (Ectron TM Wheatstone bridge with signal conditioning amplifiers) and a data acquisition system (LeCroy TM high-speed digital oscilloscope). The cylinder, barrel and the bars were all mounted on two steel I-beams supported by adjustable legs. The adjustable legs were used for the leveling and aligning the bars. A solenoid valve was connected the cylinder and the barrel was used to launch the striker. The solenoid valve was encased within a ?-metal sheet to attenuate the magnetic flux field that would otherwise generate electrical noise in the output signals and trigger the Incident bar Transmitter bar 19? ? data acquisition circuit prematurely. A series of through holes were drilled into the end of the gun barrel so that when the striker was launched, the air ahead of the striker escaped through the holes laterally and the air resistance on the striker was minimized. To reduce friction, the incident and transmitter bars were supported on a series of bronze bearings mounted on the I-beams. Both the striker and the two bars were made of Al-7075 of the same diameter in order to eliminate the impedance mismatch between them. The 7075 aluminum alloy was chosen for its strain rate independent characteristics. The sample was held snugly in between the incident and the transmitter bars prior to loading. The sample ends were coated with a thin layer of grease to position the sample and minimize friction at the interfaces. A 6.4 mm diameter disc made of a stack of four sheets of adhesively backed copper tapes each 0.1 mm thick was used at the front end of the incident bar as the pulse shaper. The interface between the pulse shaper and the incident bar was also coated with grease for tackiness. The use of pulse shaper helped to achieve dynamic stress equilibrium (to be described later) in the sample for the loading duration. The pulse shaper, the incident bar, the sample and the transmitter bar were all pushed against each other so that there were no gaps at any of the interfaces. Transmitter bar Strain gage Sample Strain gage Incident bar Striker Figure 4.2: Split Hopkinson pressure bar setup Data acquisition system Signal Conditioner Stopper Cylinder Pulse shaper 20? ? The strain histories in the bars were captured by two strain gages fixed one on each bar. The strain gages (CEA-13-062UW-350) had a grid resistance of 350? and a gage length of 1.6 mm (0.062 in) and were purchased from Vishay ? Micro- Measurements ? . One practical problem while conducting experiments was that the lead wires from the strain gages were breaking off at the terminals on the gages due to the intense stress waves. To overcome this, one long piece of masking tape was used to first wind the lead wires together. Then, the wire pair was attached to the bar using the remaining free end of the tape which was wrapped around the bar. In addition to the gages on the bars, two external dummy gages of the same specifications were mounted on an unstressed bar placed in the same atmospheric conditions as the incident and the transmitter bars. The leads from the strain gages on the bars and the external dummy gages were connected to the signal conditioning unit. The output from the signal conditioning unit was connected to the digital oscilloscope. The strain gage signals were filtered using a built in (?Eres? or Enhanced resolution filter) filter available in the oscilloscope. A resolution enhancement of 1.5 bits was chosen. The filtered data was then used to calculate the actual strain in both the bars as a function of time. The striker was pushed all the way to the back of the gun barrel, and the pulse shaper was fixed on the incident bar. The sample was then held between the bars and they were all pushed against the stopper at the end of the transmitter bar. Next, the air cylinder was pressurized and the striker was launched towards the incident bar by opening the solenoid valve using an electrical switch. When the striker impacted the incident bar, a compressive stress wave (incident wave) originated at the front end of the bar and traveled towards the sample. A part of the stress wave (transmitted wave) traveled into 21? ? the sample and subsequently into the transmitter bar. Simultaneously, a portion of the stress wave was reflected (reflected wave) at the incident bar-sample interface and traveled back into the incident bar as a tensile stress wave. All three pulses were captured by the two strain gages mounted on the incident and the transmitter bars, and recorded as voltage histories in the data acquisition system after being processed by a signal conditioning unit 1 . The oscilloscope acquired the strain signals as voltage histories at a rate of 2.5 x 10 6 samples per second. These voltage histories were then used to calculate the strain histories of the bars. 4.2 Location of Strain Gages The location of the strain gages on the incident bar was chosen based on the elastic wave speed (C 0 ) in the bar, the time period of the wave and the total length of the bar (L bar ) to make sure that there was no overlapping of the incident and the reflected stress wave signals. The bar wave speed nullnull null null null null in the Al 7075 bars is 5123 m/s [16], the striker length (L striker ) was 406.4 mm (16 in) and the length of the bars (L bar ) were 2438.4 mm (96 in). From the above data, the time period (T) [34] of the stress wave pulse was calculated as ker 0 2 stri L C = 160 ?s. Then, the time taken by the wave to traverse the incident bar once was calculated as 0 bar L C = 475 ?s. ???????????????????????????????????????????????????????????? 1 ?To ensure that the circuitry and output voltage values from the signal conditioning unit were correct, a strain gage of the same specifications was mounted on a cantilever beam and the strains for various loads were measured. These measured strains were compared against theoretically calculated strains based on the one-dimensional beam equations, and the error factor was determined. This error factor was multiplied to the output voltages corresponding to the gages on the incident and the transmitter bars to ensure accuracy. 22? ? Figure 4.3: Schematic representation of the locations of incident, reflected and transmitted waves in a SHPB setup From the above calculations, it is evident that the time period of the wave corresponds to twice the length of the striker. So, the strain gage on the incident bar was located at approximately 1200 mm from the incident bar?specimen interface which is more than twice the length of the striker. The schematic representation of the above discussion is shown in Fig. 4.3 using the so-called x-t diagram. Accordingly, the incident and the reflected wave signals will not overlap during a typical test. 4.3 Measurement of strain The strain gages on the bars were separately connected to Wheatstone bridge circuits in the signal conditioning unit. A half-bridge circuit was used in this study. The external dummy gage of the half bridge circuit was mounted on a rod of the same material as the bars so that the temperature effects on the working gage would be nullified. The rod with the dummy gage was placed in the vicinity of the bars so as to Specimen Strain gage Transmitter bar Strain gage Incident bar Striker Distance (x) Incident wave T Reflected wave i Transmitted wave m (t) e 23? ? maintain the same atmospheric conditions. The strain in the bars is recorded as voltages, V 0 using the oscilloscope. If S g is the gage factor of the strain gage, ? is the strain in the bar and R g is the resistance of the strain gage, then change in resistance ?R g in the strain gage is given by, ?null null null null null null null null (4.1) voltage for an initially balanced bridge V 0 when the gain or amplification is G, null null . The output is given by null null null null null ?null null null null (4.2) where V e is the exc n voltage fo From the last two equations, the s null null null null null itatio r the bridge. train can be expressed as, nullnull nullnull null . (4.3) .4 Stress, strain and strain rate 4 The stress ()t? , strain rate ()t?& and strain ()t? histories of the sample were calcula e sp ba Figure 4.4: Schematic illustration of the SHPB ted using th lit Hopkinson r equations. The details of the derivation [34] of the equations are provided as follows. Transmitter bar (A 0 , ? 0 ) Incident bar (A 0 , ? 0 )Striker 1 2 v 1 v 2 24? ? The strain rate or the rate of change of linear dimension in the specimen is given by nullnull nullnull null nullnull null null null nullnullnullnullnull null nullnullnull null . (4.4) As seen in Fig. 4.4, null null and null null are interface velocities at 1 and 2 and L is the specimen = (4.5) where C 0 and are the wave speeds n the bars and the specimen respec null null (4.6) From Eqs. (4.4), (4.5) length. At the interfaces, at t 0, null null null null null null null , null null null null null null null C i , tively, and the subscripts I, R (and T) denote incident, reflected and transmitted, respectively. At t > 0, null null is reduced because of the reflected wave. That is, null null null nullnull null null null null null . and (4.6), the strain-rate expression can be re-written as, nullnullnullnullnull null null 0 nullnull null nullnull null nullnull null 0 null null null . The strain in the specimen can be obtained by integrating the strain rate: null nullnull null nullnullnull null null null nullnullnullnull null null nullnullnullnullnullnull null null . nullnullnullnull null null null null (4.7) By assuming stress equilibrium in the specime l times from 0 to t, t stres null null nullnullnullnull null null nullnullnull 2null n at al he s in the sample can be written as the average of the stresses on the specimen at the two interfaces as, nullnull where, null null null nullnull null nullnull null null , null null nullnullnull null null null null null null null . null null nullnullnull null null 25? ? 26? ? Here, A 0 and A are cross sectional areas of the bars and specimen, respectively, and E 0 is the elastic modulus of the bars. Thus, nullnull null null null null nullnull nullnull null nullnullnull null null null nullnullnullnull null null nullnullnullnull . For dynamic equilibrium, null null nullnullnull nullnull null nullnullnull d null null nullnull null null null null . Hence, we get, null an nullnullnullnull null null null null null null null nullnullnull , (4.8) nullnullnullnullnull null null nullnull null null null null nullnullnullnull null null nullnull null null , (4.9) null null null null The last three equatio SHPB and were used to calculate the stress, strain and strain rate in the specimen as a function of time, and subsequently, to obtain stress-strain curves. Once the engineering stresses (? )? and strains (? ) were obtained, constant volume approximation was invoked to obtain the corresponding true stress (? ) and true strain (? ) values using: nullnull ) null nullnull. (4.10) ns (4.8), (4.9) and (4.10) are the governing equations of the eng eng true true null nullnullnullnull nullnull null1nullnull nullnullnull null , (4.11) null nullnullnullnull nullnull nullnullnull null1 nullnull nullnullnull null . (4.12 5 EVALUATION OF THE SETUP The SHPB setup developed in the laboratory was tested for its ability to produce a constant strain rate, dynamic stress equilibrium and experimental repeatability. Strikers of different lengths, ranging from 8 inches to 16 inches, were used to check the appropriate striker lengths for ensuring that the samples would undergo an appreciable strain/deformation within one stress pulse. In this chapter, the above issues are presented in addition to the calibration tests of the setup. 5.1 Dynamic equilibrium and constant strain rate The standard SHPB equations for strain rate, strain and stress in the specimen assume that the specimen is under dynamic stress equilibrium and experiencing a constant strain rate during deformation. The dynamic stress equilibrium essentially means that the instantaneous forces at the front and back ends of the specimen are equal. That is, the sum of the strains or output voltages corresponding to the incident and reflected stress waves should equal the strain or output voltage corresponding to the transmitted wave. In real situations, however, this does not occur unless the rate of loading is reduced to allow sufficient time for the specimen to reach dynamic equilibrium. This is often achieved by the use of so-called pulse shapers. A pulse shaper essentially moderates the initial impact allowing sufficient time for the specimen to attain stress equilibrium before undergoing 27? ? large deformation. So, different types of pulse shapers were used in this work at the striker / incident bar interface to attain the best possible dynamic equilibrium in the sample. Time (x 10 -6 ) sec 0 200 400 600 800 1000 St r a i n ( x 1 0 -6 ) -2000 -1500 -1000 -500 0 500 1000 1500 2000 Incident wave Reflected wave Transmitted wave Strain rate ~ 1500 /sec ? (a) Time (x 10 -6 ) sec 0 200 400 600 800 1000 Str a in (x 1 0 -6 ) -2000 -1500 -1000 -500 0 500 1000 1500 2000 Incident wave Reflected wave Transmitted wave (b) Figure 5.1: Representative SHPB strain histories for (a) Syntactic foam sample with 30% microballoon volume fraction, (b) IPC foam sample with 30% microballoon volume fraction 28? ? The pulse shapers made of a stack of copper tapes that are more compliant than aluminum resulted in reducing the rate of loading and smoothing the instantaneous input strain wave profile as seen in Fig. 5.1. The representative strain histories for SF-30 and IPC-30 samples obtained by using the above mentioned pulse shaper that resulted in good dynamic stress equilibrium are shown in Fig. 5.1. In Figs. 5.1(a) and (b), the solid line corresponds to the strain history in the incident bar, and the dotted line corresponds to the one in the transmitter bar. As can be seen from the strain histories, the initial compressive wave ramps up for about 20 microseconds to attain the maximum amplitude instead of an instantaneous jump. Subsequently, the amplitude remains constant for about 125 ?s before starting to ramp down. The tensile reflected wave from the incident bar / specimen interface reaches the strain gage on the incident bar at approximately 250 ?s after the end of the incident wave. The transmitted wave is seen as a compressive signal in Fig. 5.1, and it lasts for almost 125 ?s before its amplitude starts to ramp down. The condition that the sample should be under a constant strain-rate during the deformation was verified by plotting the sample strain-rate histories (Figs. 5.2(a) and (b)). The equations [34] for SHPB reveal that the strain rate in the specimen is proportional to the reflected wave induced strain in the incident bar. The nearly constant strain region in the reflected wave (Fig. 5.1) suggests that the specimen was under a constant strain rate for about 125 ?s during the dynamic loading event. The strain-rate histories shown in Figs. 5.2 (a) and (b) for the representative SF-30 and IPC-10 samples confirm this. As was the case for the incident (as well as reflected) strain waves, the strain-rate histories also follow the ramp-up and ramp-down patterns during the initial 29? ? and final stages of the pulses, respectively. Also, it is evident that the strain-rates remain nearly constant for about 125 ?s for both SF and IPC cases. Time (microseconds) 0 50 100 150 200 250 S t rain rat e ( s -1 ) 0 500 1000 1500 2000 2500 3000 ? Time (microseconds) 0 50 100 150 200 250 S t rai n ra t e (s -1 ) 0 500 1000 1500 2000 2500 3000 (a) (b) Time (microseconds) 0 50 100 150 200 250 E rror % -100 -80 -60 -40 -20 0 20 40 60 80 100 ?? Time (microseconds) 0 50 100 150 200 250 Er ro r % -100 -80 -60 -40 -20 0 20 40 60 80 100 ? (c) (d) Figure 5.2: Demonstration of constant strain-rate for (a) SF-30 and (b) IPC-10, and dynamic stress equilibrium for (c) SF-30 and (d) IPC-10 The dynamic stress equilibrium condition was verified by plotting the percentage difference between the incident wave and the (reflected + transmitted) wave with respect to the incident wave. From Figs. 5.2(c) and (d), it can be seen that the percentage difference is about 9% or less for most part of the loading pulse duration and that the dynamic equilibrium condition is satisfied quite well given the highly transient nature of 30? ? the dynamic event. Among the various shapers that were tried (Fig. 5.3), a stack of four 0.25 inch diameter copper tapes resulted in good equilibrium and was used in all experiments unless stated otherwise. 5.2 Effect of pulse shapers and striker length As noted earlier, different types of pulse shapers were used at the striker / incident bar interface to attain the best possible dynamic equilibrium in the sample. The stress- strain curves corresponding to the different pulse shapers are shown in Fig. 5.3. From the figure, it can be seen that there is no significant difference in the stress-strain curve, even when different pulse shapers were used. The plots also confirm the repeatability of the experimental setup. Nevertheless, the pulse shaper that gave the best dynamic equilibrium of the test sample was used for all other experiments. Strain 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 St r e s s ( M Pa ) 0 20 40 60 80 100 No pulse shaper 0.04" thick, 1/8" dia Cu 0.04" thick, 1/4" dia annealed Cu 4 stacks, 1/4" dia Cu tapes 3 stacks, 1/4" dia Cu tapes Figure 5.3: Effect of pulse shapers and repeatability on SF-30 samples (strain-rate ~ 1500/s) 31? ? As mentioned at the beginning of this chapter, strikers of different lengths were used to test samples to select the most appropriate and practical striker length. From the discussion in Section 4.2 it is clear that the loading pulse duration is directly proportional to the striker?s length when the striker and the bars are made of the same material. It should also be noted that the longer the loading pulse, the larger the sample deformation within the first pulse. This is evident from Fig. 5.4. The line-dot-dot patterned line corresponds to an 8 in. long striker. The maximum strain attained by the sample by using this striker was ~12%. Longer strikers of lengths 12 in. and 16 in. yielded higher strains of ~16% and ~20%, respectively. In Fig. 5.4, the line-line patterned line corresponds to the 12 in. striker and the line-dot-line patterned line corresponds to the 16 in. striker. Strain 0.00 0.05 0.10 0.15 0.20 0.25 S t ress (M P a) 0 20 40 60 80 100 203 mm (8") striker 305 mm (12") striker 406 mm (16") striker ? Figure 5.4: Effect of striker length on SF-30 samples (strain-rate ~ 1500/s) 32? ? 5.3 SHPB calibration Since the primary motivation for this study is to study low impedance materials, the SHPB apparatus was also calibrated relative to the published results [35] in the literature for a widely studied low impedance material. The material tested was a commercially produced cast acrylic rod stock of 12.7 mm dia. Cylindrical samples of length 6.35 mm were machined from the stock and tested under conditions similar to those used for SF and IPC samples. The resulting true stress-strain diagram for a strain rate of approximately 1250 /sec is shown in Fig. 5.5. The results are compared with those reported in Ref. [35] for specimens made from cast acrylic sheet stock of identical specimen dimensions. True strain 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 True st ress (M P a ) 0 50 100 150 200 250 300 350 Current work Ref. 35 Figure 5.5: True stress-strain response for cast acrylic at a strain rate of ~ 1250 / sec The overall similarity in the stress-strain response with the reported ones is clearly evident. The minor deviations are attributed to differences in the specimen preparation, stock material (sheet vs. rod) and strain rates (~ 1500 / sec in Ref. [35]). 33? ? 6 EXPERIMENTAL RESULTS This chapter presents the dynamic compression responses of SF and IPC obtained from the experiments. To appreciate the strain rate effects on these two material systems, quasi-static [2, 28] stress-strain responses of SF and IPC previously reported by Jhaver and Tippur [28] are presented first. Later, the dynamic stress-strain responses obtained from the current work are presented along with a comparative analysis of quasi-static and dynamic responses. Finally, a summary of the energy absorption characteristics of SF and IPC under quasi-static conditions [2, 28], which are also later compared against the corresponding results under dynamic conditions, is also presented. 6.1 Quasi-static compressive stress-strain response 6.1.1 Syntactic foam The static compressive true stress vs. true strain responses for four types of syntactic foams (SF) (10, 20, 30 and 40% V f ) are shown in Fig. 6.1. These responses show three distinct stages of deformation typical of structural foams. An initial linear elastic region is followed by a distinct drop in stress with increasing strain after yielding. (This softening response is attributed to the one seen for neat epoxy sample tested under identical static loading conditions, and is presented in Appendix A.) With increasing strain, the softening response is followed by a region of nearly constant stress. In this 34? ? plateau region, microballoons within the SF fail progressively. Further increase in load causes densification (signified by an increasing stress) as microballoons collapse completely. As seen in Fig. 6.1, the stress-strain responses of syntactic foam samples of all four volume fractions follow a similar trend. It is also evident that the compressive strength decreases with increasing volume fraction of microballoons in the SF. The 10% V f samples (SF-10) have a compressive strength of 63.2 MPa, whereas the SF-40 samples have 34.4 MPa, or a 45% decrease. (The elastic modulus values also show a decreasing trend with increasing microballoon volume fraction. The SF-10 samples have an elastic modulus of ~1600 MPa and the SF-40 samples, ~1260 MPa, which is a 21% decrease.) The true strains corresponding to the initial maximum stress are 4.5%, 4.0%, 3.3% and 3.1% for the SF-10, SF-20, SF-30 and SF-40 samples, respectively. True strain 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 True st ress (M Pa) 0 20 40 60 80 100 SF-10 SF-20 SF-30 SF-40 Figure 6.1: Effect of microballoon volume fraction on quasi-static compression response of syntactic foam (SF) samples [2] 35? ? The plateau stress values of SF samples in the order of increasing volume fractions are approximately 40 MPa, 33.5 MPa, 26 MPa and 20.6 MPa. This trend of decreasing plateau stress with increasing microballoon volume fraction is also consistent with those predicted by Kim and Plubrai [36]. Furthermore, as the volume fraction of microballoons increase in SF, the densification strain also increases monotonically, from ~50% true strain for 10% V f to ~75% true strain for 40% V f . The deformed samples were sectioned and examined in a scanning electron microscope. Figure 6.2 shows micrographs of deformed SF-30 samples corresponding to 14% (Fig. 6.2(a)) and 58% (Fig. 6.2(b)) strains. Figure 6.2: SEM images of 30% V f syntactic foam at (a) 14% strain and (b) 58% strain showing the crushing of microballoons at various stages of static deformation. (The loading was in the vertical direction) [2] The onset of microballoon failure is evident in Fig. 6.2(a). Figure 6.2(b) shows the microballoons to be almost completely crushed and flattened, which explains the densification or the increase in stress with increase in strain after ~50% strain. Furthermore, compression of the matrix and crushing of microballoons is uniform in the 36? ? entire field of view. The direction of load in the micrographs is along the vertical axis, but a skew in the direction of the fractured microballoons is seen in Fig. 6.2(a) suggesting a tendency for shear failure. 6.1.2 IPC The quasi-static results for four different IPC samples with different volume fraction of microballoons in SF (10%, 20%, 30% and 40%) are shown in Fig. 6.3. The IPC foams respond in a way similar to SF samples (Fig. 6.1). There is an initial linear elastic region, followed by yielding and softening signifying the onset of nonlinearity. The softening is then followed by regions of plateau stress and densification. True strain 0.00.20.40.60.81.01.21.4 True stress (M Pa) 0 20 40 60 80 100 IPC-10 IPC-20 IPC-30 IPC-40 Figure 6.3: Effect of microballoon volume fraction on quasi-static compression response of IPC foam samples [2] 37? ? In the case of IPC, however, the onset of nonlinearity is also affected by progressive bending of the aluminum ligaments, which in turn causes crushing of microballoons in between the ligaments (Fig. 6.4). The plateau region for IPC foam is relatively narrow when compared to the one for the corresponding syntactic foam. That is, densification starts at a relatively lower strain in IPC. The elastic modulus for 10%, 20%, 30% and 40% volume fraction IPC foam samples are 2354 MPa, 2123 MPa, 1852 MPa and 1702 MPa, respectively. The compressive strengths of the IPC samples in increasing order of volume fraction are 81.0 MPa, 63.1 MPa, 51.5 MPa and 44.0 MPa. The plateau stresses for the samples show a similar trend as the compressive strengths. That is, samples with a lower V f of microballoons have higher plateau stresses. The true plateau stress values for the 10%, 20%, 30% and 40% V f samples are approximately 59 MPa, 50 MPa, 40 MPa and 31 MPa, respectively, which is a ~10 MPa increase in plateau stress value for a 10% decrease in V f of microballoons. Figure 6.4: SEM image of statically deformed IPC foam at 58% strain. (The loading was along the vertical direction) [2] 38? ? After yielding, the onset of densification also follows a uniform trend, wherein lower V f samples have lower densification strain. The approximate true strain values which mark the onset of densification are in the 45%-55% range. The quasi-static compression properties are consolidated in Table 6.1. Designation Microballoon volume fraction (%) Density (kg/m 3 ) Compressive strength (MPa) Elastic modulus (MPa) SF-10 10 995 63 ? 1.6 1600 ? 25 SF-20 20 870 52 ? 2.2 1530 ? 35 SF-30 30 796 42 ? 1.4 1448 ? 28 SF-40 40 696 34 ? 1.8 1261 ? 42 IPC-10 10 1169 81 ? 1.6 2354 ? 26 IPC-20 20 1044 63 ? 2.3 2123 ? 32 IPC-30 30 986 52 ? 3.6 1852 ? 27 IPC-40 40 862 44 ? 1.9 1702 ? 26 Table 6.1: Quasi-static properties of syntactic foam and IPC foam samples 6.2 Dynamic compressive stress-strain response SF and IPC samples with 10%, 20%, 30% and 40% microballoon volume fractions were tested at a strain rate of ~1500 per second using the split Hopkinson pressure bar described in the previous chapter. The maximum strain attained by a sample depends on the time period of the incident stress wave. This in turn depends on the length 39? ? of the striker used [34]. Several striker lengths, 203 mm (8 in.) ? 406 mm (16 in.), were considered. In view of a relatively large expected strain to failure, the longest feasible striker length was considered. The incident stress pulse generated by the 406 mm long striker (impact velocity ~15 m/s) had a total time period of ~200 ?s with a constant strain lasting over 100 ?s as shown in the representative strain histories in Fig. 5.1 for SF and IPC samples with 30% volume fraction of microballoons. The maximum engineering strain experienced by the sample for this pulse was a little over 25%. So, conservatively, all analyses were based on a maximum specimen strain value of 25% and the stress-strain data once the stress pulse started dropping off was not considered in the analyses. 6.2.1 Syntactic foam The dynamic true stress - true strain responses of SF with four different volume fractions of microballoons obtained using a 406 mm striker at a strain rate of ~1500 per second are shown in Fig. 6.5. The response of syntactic foam has two distinct regions. An initial linear elastic response is followed by a monotonically decreasing stress region with increasing strain. As in the quasi-static cases, the compressive strengths of the samples decrease with increasing microballoon volume fraction. The compressive strengths of the 10%, 20%, 30% and 40% volume fraction samples were approximately 104 MPa, 80 MPa, 62 MPa and 50 MPa, respectively. The relative decrease in the compressive strengths for every 10% increase in the microballoon volume fraction is ~20%. Among the notable differences, dynamic loading clearly suppresses the distinct softening seen in quasi-static cases after the initial peak stress in all volume fractions, as reported in Ref. [28]. Instead, a more gradual softening is evident in Fig. 6.5. The tendency for the SF 40? ? with lower volume fraction of microballoons (10% and 20%) to soften after attaining the maximum stress is somewhat more distinct than for those with higher volume fractions of microballoons (30% and 40%). These can be further attributed to the differences in neat epoxy response under static and dynamic loading conditions as described in Appendix A. After yielding, the stresses for SF with lower V f of microballoons remain consistently higher than that for SF with higher V f . The difference in stress values after yielding between specimens with different volume fractions of microballoons is approximately constant at all strains (within the observation window). The elastic responses under static and dynamic conditions also show differences. The measured elastic modulus for the four cases of syntactic foams is shown in Table 6.2. True Strain 0.00 0.05 0.10 0.15 0.20 0.25 Tr ue Stres s ( M Pa) 0 20 40 60 80 100 120 140 SF-10 SF-20 SF-30 SF-40 Figure 6.5: Effect of microballoon volume fraction on dynamic stress- strain response of SF samples (Strain rate ~1500/sec) 41? ? Designation Microballoon volume fraction (%) Density (kg/m 3 ) Compressive strength (MPa) Elastic modulus (MPa) SF-10 10 995 104 ? 4 3460 SF-20 20 870 80 ? 3 3100 SF-30 30 796 62 ? 5 2700 SF-40 40 696 50 ? 3 2400 IPC-10 10 1169 120 ? 4 - IPC-20 20 1044 100 ? 5 - IPC-30 30 986 80 ? 6 - IPC-40 40 862 60 ? 3 - Table 6.2: Dynamic properties of syntactic foam and IPC foam samples It should be noted that the dynamic elastic modulus values shown in Table 6.2 were obtained using ultrasonic pulse-echo measurements [37]. (This method was preferred over direct measurement from the stress-strain response due to weak stress equilibrium in the specimen during the initial stages (up to about 30 ?s) of stress wave propagation in the specimen. The elastic moduli of the IPC are not reported here as the pulse-echo transducer could not be used effectively to determine wave speeds in IPC due to wave dispersion problems). The results indicate a higher elastic modulus under dynamic conditions when compared to the respective quasi-static [28] counterparts. A backlit photograph of a deformed SF sample with 30% V f of microballoons is shown in 42? ? Fig. 6.6(a). The sample in this case was loaded along the vertical direction. A network of shear bands crisscrosses the entire sample and the bands are oriented at approximately ?45 o to the loading direction. Figure 6.6: (a) Side view of a dynamically deformed syntactic foam sample revealing a network of shear bands, (b) SEM image of cross section of dynamically deformed syntactic foam; Careful observation reveals crisscrossing shear failure planes. (The loading was along the vertical direction) A micrograph of a 30% V f specimen is shown in Fig. 6.6(b). Again, the direction of loading was along the vertical direction of the image. The cracks that appear in the image are skewed at an angle of approximately 45 o to the direction of application of the load suggesting failure due to shear localization. In regions away from the crack, the microballoon footprints are essentially circular suggesting very little deformation away from the shear bands. 6.2.2 IPC The IPC samples of four microballoon volume fractions viz. 10% (IPC-10), 20% (IPC-20), 30% (IPC-30) and 40% (IPC-40) were tested and the results are shown in Fig. 43? ? 6.7. The dynamic compression response of IPC foams followed trends similar to that of the corresponding syntactic foams. The response shows a linear region in the beginning, followed by a modest nonlinear response until a maximum stress is reached. Subsequently a monotonic reduction of stress with increasing strain is seen in the observation window up to 22% true strain. As in the case of SF samples, the yield strengths of IPC samples decrease with increasing volume fraction of microballoons. True strain 0.00 0.05 0.10 0.15 0.20 0.25 True st ress (M P a ) 0 20 40 60 80 100 120 140 IPC-10 IPC-20 IPC-30 IPC-40 Figure 6.7: Effect of microballoon volume fraction on dynamic stress- strain response of IPC foam samples (Strain rate ~1500/sec) In the order of increasing microballoon V f , the maximum stress values attained are approximately 120 MPa, 100 MPa, 80 MPa and 60 MPa. The percentage decrease in the compressive strength for the IPC-20 with respect to that of IPC-10 is 17%, and that of IPC-30 with respect to that of IPC-20 is 20%. The IPC-40 has a 25% decrease in 44? ? compressive strength with respect to the IPC-30. Figure 6.8 shows a photograph and a micrograph of a deformed 40% V f IPC. As in the SF samples, the photograph of the deformed sample in Fig. 6.8(a) clearly shows a well-defined network of shear bands throughout the specimen (the shear bands are slightly more pronounced and father apart relative to the SF counterparts due to the presence of the aluminum network.). The SEM image in Fig. 6.8(b) shows underlying microscopic failure mechanisms in the IPC. Two types of failures can be identified from this image. One is the formation of cracks in the syntactic foam (similar to the one seen for SF in Fig. 6.6(b)). The crack in Fig. 6.8(b) is again inclined at ?45? to the direction of the load along the vertical direction of the image, which suggests that the failure occurred due to shear localization. However, shear bands here are interrupted by the metallic ligaments of the IPC. The other type of failure in IPC, which is absent in pure SF samples, is the debonding of the interfaces between the syntactic foam and the aluminum ligaments. Figure 6.8: (a) Side view of a dynamically deformed IPC foam sample revealing multiple shear bands and (b) SEM image of the cross section of dynamically deformed IPC foam. (The loading was along the vertical direction) 45? ? 6.3 Energy absorption An aspect of foams that is attractive for structural applications is energy absorption. A comparison of energy absorbed per unit volume and per unit mass by the SF and the corresponding IPC samples under dynamic loading conditions was made. The energy absorbed per unit volume (U) was evaluated by calculating the area under the true stress ? true strain curve as, 7 L ? ? : ? ; @? 4 (6.1) where ()? ? is the stress as a function of strain. The energy absorbed was evaluated up to a strain value of 22% as U 0.22 . The energy absorption per unit mass was also evaluated by dividing the energy absorbed per unit volume by the density of the sample. Again, to appreciate the effect of strain rate on energy absorption, the quasi-static results on energy absorption [2, 28], are presented before the dynamic counterparts. 6.3.1 Quasi-static The absorbed energy (U 0.22 ) evaluated from true stress-true strain graphs [2, 28] is plotted as histograms for all the cases in Fig. 6.9. As was the case for the compressive strength and the elastic modulus, the absorbed energies also show a decreasing trend with increasing volume fraction of microballoons. The energy absorbed per unit volume (Fig. 6.9 (a)) by the SF samples of 10%, 20%, 30% and 40% microballoon V f are 10.4 MJ/m 3 , 8.5 MJ/m 3 , 6.8 MJ/m 3 and 5.4 MJ/m 3 , respectively. The relative decrease in U 0.22 for every 10% increase in microballoon V f is 18-20%. The energy absorbed by the IPC samples is higher than that by the corresponding syntactic foam samples by 34%-37%. The actual values are approximately 14.0 MJ/m 3 , 11.5 MJ/m 3 , 9.2 MJ/m 3 and 7.4 MJ/m 3 46? ? for 10%, 20%, 30% and 40% cases, respectively. The percentage decreases in U 0.22 for the IPC-20 / IPC-10 pair, IPC-30 / IPC-20 pair and the IPC-40 / IPC-30 pair are approximately 20% in each case. Microballoon volume fraction 10% 20% 30% 40% E ner gy abso r ptio n p er unit v olume (M J/ m 3 ) 0 5 10 15 20 25 30 SF IPC (a) ? Microballoon volume fraction 10% 20% 30% 40% E ner gy ab sor p ti on pe r u n it mas s ( kJ/kg) 0 5 10 15 20 25 SF IPC (b) Figure 6.9: Energy absorbed (a) per unit volume and (b) per unit mass by SF and IPC foam samples under quasi-static loading up to 22% true strain 47? ? The energy absorption per unit mass are plotted in Fig. 6.9(b) and they also show a similar trend. The IPC samples have absorbed 12.0 kJ/kg, 11.0 kJ/kg, 9.4 kJ/kg and 8.6 kJ/kg which are higher when compared to the syntactic foams of corresponding volume fraction of microballoons, respectively. The percentage increase in the energy absorption per unit mass in the IPC foams with respect to the syntactic foams is approximately 10%-14%. 6.3.2 Dynamic The data on energy absorption, U 0.22 , under dynamic loading conditions is shown in Fig. 6.10. The absorbed energies of SF and IPC foam samples under dynamic loading conditions are clearly higher than those observed under quasi-static loading conditions (Fig. 6.9). Energy absorbed per unit volume by SF-10, SF-20, SF- 30 and SF-40 samples are 21.9 MJ/ m 3 , 19.6 MJ/m 3 , 16.4 MJ/m 3 and 12.1 MJ/m 3 , respectively and that by IPC-10, IPC-20, IPC-30 and IPC-40 samples are 24.5 MJ/m 3 , 22.0 MJ/m 3 , 18.3 MJ/m 3 and 12.7 MJ/m 3 , respectively. The percentage reduction in U 0.22 per unit volume for the SF-20 / SF-10 pair, SF-30 / SF-20 pair and the SF-40 / SF-30 pair are approximately 11%, 16% and 26% respectively. The percentage reduction in U 0.22 per unit volume for the IPC-20 / IPC-10, IPC-30 / IPC-20 pair and the IPC-40 / IPC-30 pair are approximately 10%, 17% and 30%, respectively. The trend of the percentage reduction in the energy absorption per unit volume suggests that the rate of reduction in U 0.22 would be greater than the rate of increase in the microballoon volume fraction for dynamic loading. 48? ? Microballoon volume fraction 10% 20% 30% 40% E n e r g y abso r ption pe r unit v o l u m e (M J/m 3 ) 0 5 10 15 20 25 30 SF IPC ? Microballoon volume fraction 10% 20% 30% 40% E n ergy absor p tion per unit mass (kJ/kg) 0 5 10 15 20 25 SF IPC (a) (b) Figure 6.10: Energy absorbed (a) per unit volume and (b) per unit mass by SF and IPC foam under dynamic loading up to 22% true strain However, in the case of energy absorbed per unit mass, the syntactic foam appears to perform better than the corresponding IPC foam under dynamic loading conditions. The energies absorbed per unit mass by the SF-10, SF-20, SF-30 and SF- 49? ? 40 samples were 22.0 kJ/kg, 22.5 kJ/kg, 20.6 kJ/kg and 17.4 kJ/kg, respectively and that absorbed by IPC-10, IPC-20, IPC-30 and IPC-40 samples were 20.9 kJ/kg, 21.0 kJ/kg, 18.5 kJ/kg and 14.8 kJ/kg, respectively. Again, these are nearly twice the respective quasi-static values. When compared with the SF samples, the IPC samples had lower energy absorption per unit mass by 5%, 6.7%, 10.2% and 14.5%, respectively. (It should be noted that the differences between 10% and 20% SF and between the SF and IPC of the two volume fractions are relatively small with overlapping error bars. Hence, no claim regarding the trend between these two volume fractions is made.) Accordingly, if energy absorption at a slight increase (~10%) in weight is acceptable for an application, IPC still outperforms SF even under dynamic conditions. Another interesting observation is that the increasing trend of the energy absorption per unit mass with decreasing microballoon volume fraction appears to be reversed at a microballoon volume fraction of approximately 10%. Therefore, the 10% microballoon volume fraction could be the threshold level in terms of absorbed energy per unit mass. 6.3.3 Dynamic energy absorption history Plots of cumulative absorbed energy at various strain values were also plotted to evaluate the history of energy absorption during dynamic compression. The variations in energy absorption per unit volume corresponding to the dynamically loaded SF and IPC samples are shown in Fig. 6.11. The cumulative energies absorbed per unit volume for both the SF and IPC samples follow a nearly linear trend beyond yielding. Also, as seen 50? ? from the graphs, the energies absorbed by lower V f samples are consistently higher than that absorbed by higher V f samples throughout the loading history. Engineering strain 0.00 0.05 0.10 0.15 0.20 0.25 S t rai n energy (MJ/ m 3 ) 0 5 10 15 20 25 SF-10 SF-20 SF-30 SF-40 Engineering strain 0.00 0.05 0.10 0.15 0.20 0.25 S t rai n en ergy (M J/m 3 ) 0 5 10 15 20 25 IPC-10 IPC-20 IPC-30 IPC-40 (a) (b) ? Figure 6.11: Strain energy absorption history of (a) SF and (b) IPC with increasing strain in SHPB tests 6.4 Failure progression Another interesting outcome under dynamic loading conditions is strain recovery in SF and IPC foam samples. The final lengths of the deformed samples were measured after experiments on both SF and IPC. Interestingly, for SF samples, the final measured lengths were more than that predicted by the SHPB equations after imposing a 25% engineering strain. This suggests that the SF samples had partially recovered (sprung- back) after the loading pulse had ended. Although the source/s of spring-back in SF is/are unclear at this point and would be interesting to study in the future, this phenomenon was negligible in the case of IPC samples. Whatever caused the spring-back to occur in the SF, is probably overcome in the IPC by the aluminum ligaments. Once aluminum ligaments undergo plastic deformation, they prevent the SF from springing-back. This 51? ? aspect of SF relative to IPC needs to be considered in an application if spring-back is a design factor. This issue was further examined optically by tracking the progression of deformation under static and dynamic loading conditions for both SF and IPC foams. To understand the failure of SF and IPC foams including the above noted spring-back behavior in SF, real-time photography of progression of deformation during loading was carried out. The schematic of the experimental setup is shown in Fig. 6.12. This was also particularly important under dynamic loading conditions to assess progressive (failure originating at the impact end and propagating along the length of the bar) vs. steady-state deformation as stress waves propagate and deform the sample. Grating lines on flattened sample surface Incident bar Transmitter bar Lamp Lamp Hi speed camera Figure 6.12: Schematic of the experimental setup used for high-speed photography to study real-time failure progression in SF and IPC samples in the SHPB set up 52? ? Two narrow (3 mm wide) diametrically opposite flat surfaces were machined on the foam samples and were printed with a line grating (pitch = 0.4 mm). The grating lines were first drafted using Solid Edge TM software and then printed on a decal paper. The printed grating lines on the decal paper were then transferred to the specimen surface. A few selected photographs of the specimen with gratings recorded using digital high-speed photography (186,000 frames/sec) are shown in Fig. 6.13. Figure 6.13: Stages of dynamic deformation of SF-30. Images (a)-(c) correspond to he first loading pulse and (d)-(f) correspond to later time instants showing presence ot f ? sp ring-back. The pitch of the coarse grating lines is 0.4 mm. The sample was loaded in the horizontal direction; t = 0 corresponds to the time when deformation begins. ? 53? ? The image in Fig. 6.13(a) corresponds to the undeformed sample just before the onset of the loading pulse in SHPB. The subsequent images show the deformed samples at different time instants (t). It is evident from the deformed length (l) that there was continuous decrease in the sample length up to 135 ?s. This time instant is also when the loading pulse starts to decay due to the finite striker length. That is, t = 135 ?s marks the end of the constant strain rate region of the loading pulse. The approximate strain at 135 ?s calculated directly from the image is ~23% which is close to the obtained engineering strain of 22% using SHPB equations. Around 195 ?s the sample starts to spring-back. At the 850 th ?s the sample has regained about 0.7 mm in length which is about 9.5% of the deformed length at the end of the loading pulse at 135 ?s. Figures 6.13(d), (e) and (f) correspond to time instants after the end of the first incident stress wave, and before the second incident stress wave. From the images shown in Fig. 6.13, due to the coarseness of the gratings used, shear bands seen on the specimen surface in Fig. 6.6(a) are not readily evident although a hint of these bands can be visualized in Fig. 6.13(c) and beyond. Much higher gratings densities and magnification are needed to achieve the necessary spatial and temporal resolutions in this regard. To contrast the response of SF with that of IPC, a similar optical observation using high-speed photography was also carried out for an IPC sample. A selection of corresponding real-time images is shown in Fig. 6.14. Noticeably, the quality of the images in these experiments is not as good as the ones in Fig. 6.13 due to the presence of shiny aluminum ligaments in the IPC which perturb the overall light field making it difficult to image the specimen. Nevertheless, qualitative observations regarding the nature of deformation progression can be made. The spring-back 54? ? phenomenon seen in SF is nearly absent in IPC samples under stress wave loading as evident from Figs. 6.14 (d), (e) and (f). Figure 6.14: Stages of dynamic deformation of IPC-30. Images (a)-(c) correspond to the first loading pulse and (d)-(f) correspond to later time instants showing absence of spring-back. The pitch of the coarse grating lines mm. The sample was loaded in the horizontal direction; t=0 corresp to the time when deformation be is 0.4 onds gins. 55? ? 7 FINITE ELEMENT MODELING AND ANALYSIS OF IPC FOAM In view of the geometric complexity of modeling the IPC foam computationally, a unit-cell analysis was undertaken. A single cell of the IPC foam was used to predict the dynamic compressive response. The geometry was modeled using solid modeling software Solid Edge and imported into the finite element analysis software ABAQUS/Explicit. In ABAQUS, the dynamic/explicit algorithm was chosen to solve the temporal deformation process numerically. The following details the methodology adopted for geometric modeling and finite element analysis. 7.1 Geometric modeling In 2005, Gong et. al., [38] studied the compressive response of unfilled open-cell foams. In their work, they reported that the number of sides of a cell range from 9 to 17, the average being 13.7 (see, insert in Fig. 3.1(b)). Accordingly, they chose a 14 sided polyhedron (tetrakaidecahedron) [39] called a Kelvin cell to represent a single cell of the open-cell foam in their analysis. In this research, the same Kelvin cell was adopted as an idealized representation of a unit-cell of the aluminum open-cell foam. The polyhedron consists of 6 squares and 8 hexagons, with all the edges having the same length l, as shown in Fig. 7.1. The height, h of the Kelvin cell is nullnull2?2null [38]. The length of the ligaments in the actual aluminum foam was ~1.5 mm and the same dimension was 56? ? chosen for the ligaments in the idealized model. The resulting wire mesh of the Kelvin cell model had a height, h ~ 4.24 mm. Further, the Kelvin cell was modeled as a solid with triangular cross section ligaments based on the actual cross sectional shape of the aluminum ligaments, increasing the total height of the model to ~ 4.69 mm. The cross section of the ligaments was chosen to be an equilateral triangle with ~ 0.78 mm sides for reasons to be discussed shortly. In reality, the cross section along the actual ligaments was not a constant, but considering the complexity of geometric modeling, the ligaments were modeled as having a constant cross sectional dimension. 4.69 mm 1.5 mm 3D array of Kelvin cells Unit cell IPC Ligament Side view of unit cell IPC Figure 7.1: Schematics representing the building blocks of an idealized Kelvin cell based IPC 57? ? Though there are some non conformances of the model geometry to the actual geometry, the aluminum volume fraction in the model matches the volume fraction range (8 -10% as claimed by the manufacturer of the aluminum foam) of the aluminum in the actual test sample. Thus, the resulting IPC model had an overall dimension of ~ 4.69 mm x 4.69 mm x 4.69 mm. The region not occupied by aluminum was ?filled? with the SF material. In Fig. 7.1, the periodic arrangement of the Kelvin cells in 3D space resulted in a regular, well organized idealized open-cell foam whereas in actual aluminum foam, the cells are not periodic. Nevertheless, the randomness of the actual aluminum foam cells were assumed to be uniform on the scale of the test sample dimensions. 7.2 Computational model 7.2.1 Material properties The material definitions for the SF region were assigned the compressive material properties of the respective microballoon volume fraction SF obtained from the SHPB experiments (Chapter 6). The elastic moduli were obtained by computing the slopes of the stress-strain curves obtained from the SHPB experiments. The Poisson?s ratios and the density of the SF are shown in Table 7.1. The aluminum region was assigned a bilinear stress strain property corresponding to Al 6101 based on Alcoa Inc. datasheet 1 . The elastic regime for aluminum was assigned a Young?s modulus of 69 GPa and a Poisson?s ratio of 0.35. The plastic regime was assigned an initial yield stress of 172 MPa at a plastic strain of 0, and a stress of 200 MPa at a plastic strain of 0.148, for Al 6101 as shown in Table 7.2. ???????????????????????????????????????????????????????????? 1 ?The metallic phase of the IPC was assumed to be strain rate independent. 58? ? Material Elastic Modulus (MPa) Poisson ratio Density (kg/m 3 ) Al 6101 69000 0.35 2700 SF-10 5122 0.34 995 SF-20 4529 0.34 870 SF-30 4459 0.34 796 SF-40 3835 0.34 696 Table 7.1: Elastic properties of SF and aluminum used in the finite element model Plastic strain Stress (MPa) 0 172 0.148 200 Table 7.2: Inelastic properties of Al 6101 used in the finite element model 7.2.2 Yielding and hardening The von-Mises yield criterion was used to define material yielding. Further, yielding was assumed to be isotropic. This criterion uses the equivalent uniaxial plastic strain and the corresponding uniaxial stress input to define the yield surface [40, 41]. The material property input for the plastic regime was supplied in the form of stress as a tabular function [40] of plastic strain. In ABAQUS, the stress at any given state within the maximum plastic strain defined will be interpolated [40] based on the supplied tabular 59? ? data. Beyond the maximum plastic strain defined, the stress remains constant and equal to the stress value in the previous step in the data [40]. The hardening and softening behaviors are also assumed to be isotropic. This means that the yield surface is the same [40] in all directions based on increasing or decreasing stress as plastic straining occurs. In the current material system, aluminum could be undergoing strain hardening, while the SF could be softening. The yield function [41] for isotropic hardening is expressed as, nullnullnullnull null null null nullnull nullnull ,nullnull nullnullnull nullnull nullnull nullnull nullnull nullnullnullnullnullnull where ? 0 is the equivalent uniaxial stress, ? pl is the equivalent plastic strain and ? is temperature. Reference [41] details the algorithm that ABAQUS uses to solve the nonlinear material model based on the von-Mises yield criterion. Some of the details are as follows: The total strain can be partitioned into elastic and plastic strains as, (7.1) where ? is the total strain and ? el and ? pl are the elastic and plastic strain components, respectively. For a given stress state, the strain states could be different depending on the loading path. Due to this, increments of plastic strain (nullnull ) need to be computed based on the loading path before obtaining the total strain for a particular state of stress [42]. The Prandtl-Reuss plasticity relations state that the increments of plastic strain depend only on the instantaneous deviatoric stresses [42]. Hence, the deviatoric stress matrix, S is obtained using (7.2) where ? is the stress matrix, I is the identity matrix and p is the volumetric stress given by 60? ? nullnullnull null null nullnullnullnullnullnullnullnull. nullnullnullnullnull nullnullnull null2nullnull nullnull nullnull nullnull nullnull nullnullnullnull nullnull null ; null null null null (7.3) Further, p and S are related to strains under elastic conditions as and null , (7.4) where, K and G denote the bulk modulus and shear modulus, respectively, ? vol is the elastic volumetric strain, and null is the deviatoric elastic strain. It is to be noted that bold symbols ?, S and I represent (3x3) matrices. Now, writing the incremental plastic strain as a product of its scalar value and its direction, , null null null null null null:null null null (7.5) where, e pl is the deviatoric plastic strain, null is the scalar deviatoric plastic strain, n is the direction vector and q is the estimated equivalent uniaxial stress. Equation (7.5) is called the integrated flow rule. null nullnull pl nullnull nullnull nullnull step. It is understood that nullnull corresponding to a particular nullnull nullnull is obtained from the material definition input as stated at the beginning of this section. Therefore, the condition to be verified is, nullnullnullnullnullnullnull nullnull null. (7.6) To check this condition, the expression for q would first be obtained as follows: The above equations are then solved for the state of the material at the end of every increment. The state of the material refers to the estimated equivalent uniaxial stress (q) as a function of deviatoric equivalent plastic strain (e ). The state of the material will be verified by checking if the estimated stress (q) equals the scalar equivalent stress ( ) corresponding to the scalar deviatoric plastic strain ( ) of that load 61? ? The change in the deviatoric plastic strain (?e pl ) between the current and the previous increment is computed as ?null nullnull null?nullnull nullnull null nullnull2nullnullnull nullnull | null null?nullnull?nullnull nullnull nullnull ?null null1null nullnull null . (7.7) Using Eqs. (7.1) and (7.7), Eq. (7.2) is modified as (7.8) where, is the change in the total strain. In the above equation, the term in the parentheses is the estimated elastic strain value at the end of the current iteration. The subscript t represents the previous time step. Now, substituting for n in the above equation from Eq. (7.5), ?nullnull nullnull nullnullnull2nullnullnull ; nullnull null nullnull nullnull | null null?nullnull nullnull3null?nullnull nullnull null3nullnull? ; null? null null null null . (7.9) The inner product of the above equation with itself gives nullnull:nullnull . (7.10) As q must now satisfy Eq. (7.6), the above equation is modified as 3nullnullnull? null?nullnull nullnull nullnullnullnullnull0 The above equation is then solved using Newton?s method to get, null nullnull null nullnullnullnull?null?nullnull nullnull nullnullnullnull nullnullnullnull . (7.11) ; null null nullnullnull nullnullnull nullnull , (7.12) where, null nullnull is the error in the plastic strain increment in case Eq. (7.6) is not satisfied. H is the strain hardening function, which is nothing but the relationship between the respective stresses and plastic strains given as the material definition input (in this work, the input is given in a tabular format). Now, the new ?nullnull or the new plastic strain increment value for the next iteration for the same load step is obtained as ?nullnull null?nullnull nullnull . (7.13) nullnull nullnull nullnull nullnull 62? ? The iterati zero. A , the deviatoric and volumetric stresses corresp .2.3 Explicit integration scheme in ABAQUS motion , the explicit central difference operator satisfie nullnull nullnull nullnullnull nullnull null null (7.14) where nullnull nullnullnull represents th em ?null on is then carried out until ?nullnull nullnull converges, or until null nullnull approaches t this stage, the solution for that increment is fully defined. The converged change in scalar equivalent plastic strain (?nullnull nullnull ) is used in Eq. (7.9) to obtain the deviatoric stress (S) that corresponds to the equilibrium state for that particular increment. The volumetric stress (p) is also now computed using Eq. (7.3). In essence, for each increment of load onding to the equilibrium states are computed based on the converged value of change in the scalar equivalent plastic strain at every integration point in the model. 7 The explicit central difference integration rule is used to solve the equations of governing a dynamic event in ABAQUS. The implementation of this rule in ABAQUS [40] is summarized in the following: At the beginning of every increment, t s the dynamic equilibrium equations; the result of which is the acceleration at that time step, nullnull nullnullnull null nullnullnull e acceleration with null as the degree of freedom, the subscript i is the incr ent in an explicit dynamic step, null is the mass matrix, null is the applied load vector and null is the internal force vector. The velocity for the time nullnull null is computed using the acceleration at time t. Then, using this computed velocity and the displacement at time t, the displacement at time nullnull?null is computed as, [40]: nullnull nullnullnull null null null nullnullnull nullnullnull null null null null ?null nullnullnullnullnull null?null nullnullnull null 63? ? nullnull nullnullnull (7.15) and null nullnullnullnullnull nullnull nullnullnull null?null nullnullnullnullnull nullnull nullnullnull null null null . (7.16) At time t = 0, the initial v fied otherwise, and the mid-increments nullnull nullnull null null alues of velocity and acceleration are set to zero, unless speci null and nullnull nullnull null null null are given as nullnull nullnull null null null nullnullnull nullnullnull null ?null nullnullnull null nullnull nullnullnull and nullnull null (7.17) nullnull null null null nullnullnull nullnullnull null ?null nullnullnull nullnull nullnullnull , (7.18) respectively. The displacements will then be used to obtain the state of the material as described above. Unlike the implicit scheme, iterations are not used here to achieve convergence. Instead, the stability of the solution is governed by the appropriate time step ch is dictated by the time required by the stress wave to travel the smallest elemen US and discretized using four noded linear tetrahedron (C3D4) [40] elements. To check for the appropriateness of the odel meshed with elements having 0.2 mm and 0.1 mm edge length were simulated and the stress vs. strain results corresponding to the two mesh sizes were size (?t), whi t and the maximum wave speed in the model defined [40]. The time step size should be chosen to be sufficiently small, typically a third or lower than the time taken by the stress wave to travel the smallest element in the mesh. 7.3 Discretization The unit cell model of the IPC was used to predict the overall response of the test sample. The geometric model was imported into ABAQ mesh refinement, a m 64? ? compared. The comparison (Fig. 8.2) shows that the results nearly overlap on each other. So, for all further analyses, the model was meshed with elements having an approximate element edge of length 0.2 mm. True strain 0.00 0.05 0.10 0.15 0.20 0.25 0.30 True st ress (M P ) a 40 60 80 100 IPC-30 SHPB 0 20 IPC-30 FEA 0.2 IPC-30 FEA 0.1 Figure 7.2: Check for appropriateness of discretization In order that the model better represents the sample, the unit cell is imagined to be located deep within a large material volume. The Solid Edge model of the unit cell has exposed faces, and does not represent a cell within the material volume. This is because, in a dynamically ire length of the sample before reflecting at the end of the sample, whereas in the model the stress waves would loaded sample, the stress waves travel through the ent get reflected at the boundaries of the cell itself. To overcome this shortcoming, a skin / layer of SF and aluminum material was modeled adjacent to the corresponding material at the boundary of the cell. This skin was meshed using the so-called infinite elements [43]. 65? ? Syntactic foam Infinite element skin Aluminum (a) (c) Figure 7.3: Illustration showing (a) the infinite element layer in the FE model and stress wave behavior in the (b) absence and (c) prese ment layer Reflected wave Incident wave Free surface (solid line) (broken line) (b) T Incident wave Infinite elem layer ent ransmitted wave nce of infinite ele 66? ? 7.3.1 Infinite elements ? Background small c ly the most important work on infinite elements was carried out by Bettess [43] in 1977. He suggested Lagrange polynomial type shape functions that include an exponential decay ts should satisfy two main requirements, one that the null nullnull null null nullnull null nullnull The infinite elements are used in situations where the region of analysis is very ompared to the region surrounding it [40]. One of the earliest and probab term. Bettess states that the infinite elemen shape functions should be realistic and the other that it should lead to integrations over the finite element domain. Accordingly, Bettess proposed shape functions for one dimensional domain presented as, null null null? null null nullnull null null nullnull null null nullnullnull i varying from 1 to n-1 and j ? i. Among all the points in the domain, the n point is infinitely distant and all other points have finite x-coordinates. As reported by Bettess, in the above expression, L is an arbitrary distance and serves as the measure of severity of the exponential decay. il regions and plane frames. Two types of infinite elements were develop nullnullnull nullnullnull where, n is the number of points in the domain, x is the coordinate and N is defined for i th Since its inception, infinite elements have been developed and used by several researchers. In 1987, Haggblad and Nordgren [44] used infinite elements to model the far field regions to study a nonlinear soil-structure interaction effect. Viladkar et al. [45] also used infinite elements coupled with finite elements to idealize and study interactions between the far field so ed and used by Park et al. [46] to analyze the hydrodynamic forces on offshore structures. One type was used to model radiation in the fluid at infinity and the other type 67? ? used to model the fictitious bottom boundary of the deep water. Later in 1993, Zhao and Valliappan [47] developed a 3D dynamic infinite element that satisfies four different requirements all of which are very useful in representing practical problems much more efficiently. Two of those requirements used in the FE model of the current work are, interfacial displacement compatibility between finite and infinite elements and displacement continuity at the common boundary between infinite elements possessing different material properties. In ABAQUS, as infinite elements are associated with the far field regions, their behavior is assumed to be linear and the material response isotropic [40]. The particular infinite element used in this analysis is a 3D, 8 node, one-way infinite brick (CIN3D8) [40]. The infinite elements (CIN3D8) transmit any stress wave incident on them to a hypothetically infinite distance in one direction and thus prevent any reflection of stress waves side. On the side opposite to that of the support, and on the surface of the unit cell which at the boundary of the unit cell. This phenomenon made possible by the implementation of a layer of infinite elements in the finite element model enforces the condition that the unit cell is located within a larger volume of IPC material. Those infinite elements adjacent to aluminum were assigned the properties of aluminum, and those that are adjacent to SF were assigned the SF properties as seen in Fig. 7.4. Even though all the sides of the IPC have infinite element layers, for clarity purposes, the infinite element layer on only one side is shown in Fig. 7.4. 7.4 Boundary conditions and loading The unit cell is supported on the free surface of the infinite element layer on one 68? ? marks the boundary between the unit cell IPC and the infinite element layer, a velocity pulse of magnitude 10250 mm/s for a time period of 250 microseconds is applied. This velocity corresponds to the approximate particle velocity in the incident bar at the cle velocity (V PI ) in the incident bar is given by, null nullnull nullnull Figure 7.4: Load and boundary conditions incident bar / sample interface. The parti nullnullnull nullnull nullnullnull null 5.123null 10 null nullnull null? null0.002 null 10250nullnull null? , where C bar is the incident bar wave speed and null nullnullnull (~ 0.002) is the maximum plateau strain in the incident bar caused by the stress wave. The time period of 250 microseconds was chosen as it is the approximate duration for which the actual test samples were loaded in the SHPB experiments. x y z Infinite element layer with aluminum properties lay rtie Infinite element er with SF prope s Uniform velocity pulse v(t) = 10250 mm/s 69? ? 7.5 Other analysis parameters The analysis was set up to simulate the dynamic problem for 250 microseconds, same as the loading pulse duration. The load steps were automatically selected by the ABAQUS software based on the fastest wave speed and the smallest element size in the model [40]. The load step period is chosen to be less than the time taken by the stress wave to traverse the smallest element in the model. As a verification procedure, the software checks if the formation speed of the material is less than one-third the wave speed in the material at each load step. As rge deformations were expected during the deformation, geometri ded in the analysis procedure. analysis 7.6.1 de la cal nonlinearity effects were also inclu 7.6 Results from finite element Stress-strain response The stress and strain histories of all the elements (aluminum and SF) on the surface of the cell on the side directly opposite to the loaded surface were obtained. From the data obtained, the average stress and strain histories were obtained, which were used to plot the dynamic stress-strain curves. The same procedure was followed for all the numerical IPC models corresponding to 10%, 20%, 30% and 40% microballoon volume fractions in SF, and the results are shown in Fig. 7.5. The trends are similar to those observed from the experimental results discussed in Ch. 6, Fig. 6.7. 70? ? True strain 0.00 0.05 0.10 0.15 0.20 0.25 True st 0 20 40 60 80 140 100 120 ress (M P a ) IPC-10 FEA IPC-20 FEA IPC-30 FEA IPC-40 FEA Figure 7.5: True stress ? true strain responses of different IPC obtained from finite element analysis 7.6.2 Comparison between experimental and finite element results The graphs in Fig. 7.6 show the dynamic stress-strain responses of all volume fraction IPC models obtained from the finite element analysis. In general, a good agreement between the measured results and finite element results is observed. The differences between the experimental results and the simulations are attributed to the following ? In the actual specimens micro cracking occurs whereas the numerical model does not issues: allow any debonding to occur at material interfaces. 71? ? ? The material property of aluminum used in the model corresponds to the quasi-static compressive response. This assumption was used considering the fact that aluminum is r in rate independent. lections back into the sample at elatively stra ? The infinite elements transmit all the stress waves that are incident on them. In the real situation however, there will be stress wave ref the sample / transmitter bar interface within the duration of 250 microseconds. True strain 0. ? True strain 000 0.05 0.10 0.15 0.20 0.25 True s t ress (M P a ) 0 20 100 140 0 20 100 140 40 60 80 120 IPC-10 EXP IPC-10 FEA .00 0.05 0.10 0.15 0.20 0.25 Tru e stre ss (M Pa) 40 60 80 120 IPC-20 EXP IPC-20 FEA ? ? True strain 0.00 0.05 0.10 0.15 0.20 0.25 T r ue st ress (M P a) 0 20 40 60 80 100 120 140 IPC-30 EXP IPC-30 FEA ? True strain 0.00 0.05 0.10 0.15 0.20 0.25 T r ue str e ss (M P a ) 0 20 40 60 80 100 120 140 IPC-40 EXP IPC-40 FEA Figure 7.6: Comparison of experimental and FEA results of (a) IPC-10, (b) IPC-20, (c) IPC-30 and (d) IPC-40 (b) (c) (d) (a) 72? ? 7.6.3 alysis of contour plots From the finite element analysis, the von-Mises stresses, the equivalent plastic strains, displacements along the loading direction, the kinetic energy and the total strain energy of the whole model for up to 22% true strain were obtained. To appreciate the spatial and temporal distribution of these quantities in a dynamically deforming IPC, the FE results (contour plots) corresponding to the IPC-30 model is presented. In these contour p he IPC. The right hand side image section made at exactly half the length of the unit cell along the z-direction. From the stress contour plots (Fig. t minum ligaments experience the maximum stress, approx An lots, the left hand side images show contours on the outer surface of t s show contours on the x-y plane of the 7.7), i can be seen that the alu imately 200 MPa, equal to the yield stress of aluminum. The maximum stress attained by the SF however is approximately 70 MPa, which is close to the yield stress of SF-30 obtained from the experiments. From the figures, it is also evident that the stress concentrations occur in the vicinity of the ligament joints. This is expected considering the geometry and the material discontinuity present in the model. 73? ? Figure 7.7: Mises stress contour plots from dynamic FE model for IPC-30 Left side images: whole IPC, Right side images: sectioned IPC (Loading is along the x-direction) (b) 12% strain (c) 20% strain (a) 5% strain A A B C B C A B C A B C B C C A A B 74? ? strain contour plots from ic FE model for IPC-30 side images: whole IPC, Rig ectioned IPC (Loading is along the x-direction) Figure 7.8: Equivalent plastic ft dynam sLe ht side images: (a) 5% strain (b) 12% strain (c) 20% strain A B C A B A B C A B A B C C B A C C 75? ? The contour plots of the equivalent plastic strain ar the contour plots, it is obvious that the SF regions have experienced larger pla c strains compared to that in aluminum. A much more interesting observation from the figures is that the locations of localized maximum plastic strain are not too close to the ligament joints. This suggests that very close to the ligament joints, the stiffer aluminum regions ic strain have restrained the SF from ng as much as the SF regions at are further away. The contours also help in appreciating the percentage of strain energy absorbed by the aluminum and the SF regions. Unlike the stress contours and equivalent plastic strain contours, the disp ent contours are rather continuous (Fig. 7.9). This is expected as debonding between the aluminum and SF regions are not allowed to occur in the model. That is, the material interfaces are modeled as perfectly bonded. The contours on the cross-section view reveal that up the length of the unit cell fr end the displacem at th ce s reversed. Th d gradually decreas e shown in Fig. 7.8. From sti with low plast deformi th lacem to half om the loading eent nter is more than that at the periphery. Beyond the mid-section of the cell, this trend i at is, the displacement is a maximum at the periphery an es towards the center symmetrically. Near the mid-section, the displacement is nearly constant along the cross section. 76? ? .9: x-displacement contour plo odel for IPC- Left side images: whole IPC, Right side images: sectioned IPC (Loading is along the x-direction) Figure 7 ts from dynamic FE m 30 (a) 5% strain (c) 20% strain A A B B A (b) 12% strain C C C C C C B B B B A A A 77? ? 7.6.4 Energy absorption It was shown earlier (see Section 6.2.2) that the compressive strengths and the stresses beyond yield point in IPC are higher than their corresponding pure syntactic foam. In other words, the area under the stress-strain curve (or the energy absorbed per unit volume) for a dynamically deforming IPC is greater than that for a SF. The introduction of aluminum into the SF has helped enhance this property among others such as the compressive strength. At this stage it was thought to be a worthwhile effort to analyze the proportion of various energies absorbed by the aluminum and SF phases in the IPC. The kinetic energy (KE) trends and the total strain energy (SE) trends are of particular interest. Previously, the equivalent plastic strain contours gave a hint regarding the proportion of strain energy absorbed by the two phases of the composite foam. To quantify the proportions shared between the SF and the aluminum phases, the KE history and the SE history for a duration of 125 microseconds, which approximately corresponds to the 22% strain, are plotted below. Engineering strain 0.00 0.05 0.10 0.15 0.20 0.25 0.30 Kin e tic e nergy (mJ) 0 1 3 5 (a) 2.0 (b) 2 4 SF-30 AL IPC-30 Engineering strain 0.00 0.05 0.10 0.15 0.20 0.25 0.30 S tra in ene rgy (J) 0.0 1.2 0.4 0.8 1.6 SF-30 AL IPC-30 Figure 7.10: History of (a) kinetic energy and (b) strain energy absorption by aluminum and SF in unit-cell IPC-30 78? ? Figures 7.10(a) and (b) show the amount of KE and SE absorption by SF and aluminum per unit cell of IPC. To quantify the proportion of these two energies shared by SF and aluminum, the percentage of KE and SE absorbed by SF and aluminum per cell is plotted in Figs., 7.11(a) and (b), respectively. From the Fig. 7.11(a), it can clearly be seen that the KE absorbed by the aluminum region accounts for more than 20% of the total KE absorbed by the IPC throughout the loading history. It should be realized that the volume fraction of aluminum in the model IPC is ~8 o, 8% of the IPC material absorbs at least 20% of the total KE. %. S Engineering strain 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0 20 40 60 80 100 20 40 60 80 100 Kine t i c en er gy % SF-30 AL Engineering strain S t r ain e nergy % SF-30 AL 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0 E >> KE) is to be noted here. Although, the strain energy absorbed per unit cell (Fig. 7.11 (b)) by the aluminum region is considerably lower with (a) (b) Figure 7.11: History of percentage of (a) kinetic energy and (b) strain energy absorption by SF and aluminum in IPC-30 From Fig. 7.11(b), it is seen that about 11% of the total SE is absorbed by aluminum during most of the loading history. Again, comparing the proportion of the SE absorbed by aluminum (~11%) to that of the volume fraction of aluminum in the model IPC (~8%), there is an encouraging, though not very prominent trend. The difference in KE and SE magnitudes (S 79? ? 80? ? res to that by the SF, the fact that aluminum prises only ~ 8% of the total volume of the IPC cell makes the energy absorbed by the aluminum phase more significant. This suggests that if the aluminum phase is replaced with a more compliant material, the energy absorption by the resulting IPC might be higher than that observed in the aluminum foam / SF IPC. It should also be noted that using a more compliant material might prove to be detrimental to the overall structural integrity of the IPC. Nevertheless, the proposed FE m characteristics of di pect com odel could be effectively used to predict the energy absorption fferent material combinations. 8 SUMMARY A split Hopkinson pressure bar was developed for characterizing the compression response of low impedance materials. The apparatus was successfully calibrated against compression data in the literature for acrylics at strain rates of 1250 /sec. The repeatability of stress-strain responses was demonstrated for different stress pulse durations as well as modest variations in pulse shapers. Using this SHPB apparatus, the compression characteristics of Interpenetrating Phase Composite (IPC) foams were studied. The IPC foams were prepared by infusing uncured epoxy-based syntactic foam (SF) into open-cell aluminum preforms/scaffolds. Curing of SF resulted in an IPC structural foam with a 3D interconnectivity. The dynamic compression responses of IPC foams made of SF with 10% - 40% V f of hollow glass microballoons were studied at strain rates of ~1500/sec. The responses of IPC foams were also evaluated relative to their pure SF counterparts. The failure progression in both types of samples was recorded using high-speed photography to examine the real-time deformation and failure processes. The micrographs were also used to explain the underlying failure mechanisms. A unit-cell based finite element model of an IPC was developed using a Kelvin cell representation of aluminum ligaments of the open-cell scaffold. The dynamic compression response was simulated using ABAQUS structural analysis software. The 81? ? 3D geometries for the SF and aluminum regions of the unit cell were modeled using Solid Edge. Elasto-plastic experimental measurements were used as the material property inputs for the SF region in the finite element IPC model. Infinite elements were used to model the external surface boundary of the unit-cell to idealize its presence within a large volume of the material. The major conclusions of this work are as follows: ? The V f of microballoons in SF plays a dominant role in the overall response of the SF and IPC foams. A monotonic increase in elastic modulus, yield stress, and plateau stress are evident as V f of microballoons decreases. The IPC foams consistently have higher values for each of these characteristics relative to the corresponding SF. ? Unlike static results, under dynamic conditions SF and IPC samples show a stress- strain response that has only two dominant regimes ? an initial proportional loading zone up to a maximum stress, and a subsequent monotonically softening zone. The foams with lower V f of microballoons tend to soften more rapidly than the ones with higher V f . The elastic modulus and maximum stress increase with decreasing V f of microballoons under high-strain rate loading. The values for IPC are again higher than those for SF under dynamic conditions. The dynamic maximum stress values are also higher for both SF and IPC relative to the quasi-static ones. ? The dynamic energy absorption values (U 0.22 ) up to 22% strain are nearly twice the corresponding static values for both IPC and SF samples. The energy absorbed per unit volume by the IPC under dynamic conditions is about ~10% higher than the SF counterparts. The energy absorbed per unit mass, however, favors SF under dynamic conditions. This is unlike the observation under static conditions. 82? ? ? The energy absorbed per unit mass for both the SF and IPC foams under dynamic loading follow a linear trend with respect to the applied strain. ? The failure of SF and IPC under dynamic conditions is dominated by the formation of an extensive network of shear bands in SF. Also, under dynamic conditions the microballoons are not as uniformly crushed as in static cases. The IPC samples show debonding of aluminum ligaments from the surrounding SF as well. ? Unlike static results, the high-speed optical recordings of the deformation process reveal significant spring-back in SF whereas it is negligible in case of IPC. Also, the failure process is rather uniform in both SF and IPC and no evidence of spatially progressive failure (crushing) is seen at the loading rates of 1500/sec used in this study. ? A unit-cell model of the IPC using a space-filling polyhedron ? tetrakaidecahedron or a Kelvin cell was successfully developed using ABAQUS. Infinite elements were also used to model the external surfaces of the unit-cell. ? The stress-strain responses of the model IPC?s with 10% - 40% V f of microballoons in SF were compared to those from experiments. A good agreement between the simulated and the experimental results was observed. ? Contour plots of von-Mises stress, equivalent plastic strain and displacement in the loading direction were analyzed. It was observed that maximum stress concentrations occurred at the vicinity of the aluminum ligament joints and the SF region experienced higher plastic strains than the aluminum region. The proportions of kinetic and strain energies absorbed by aluminum and SF regions were also analyzed. 83? ? 84? ? ? Citing the conformance of the results of the unit-cell model to those of the experimental results, the proposed method of finite element analysis could be used effectively to predict the energy absorption characteristics of different material combinations. REFERENCES 1. Sarikaya M, 'An introduction to biomimetics - A structural view point', Microscopy research and technique, 27(5), pp 360-375, 1994. 2. Jhaver R, 'Compression response and modeling of interpenetrating phase composites and foam-filled honeycombs', MS thesis, Auburn University, 2009. 3. Clark D. R., ?Interpenetrating phase composites?, Journal of the American Ceramic Society, 75(4), pp 739-759, 1992. 4. Nammi SK, Myler P, Edwards G, 'Finite element analysis of closed-cell aluminum foam under quasi-static loading', Materials and design, 31(2), pp 712-722, 2010. 5. Hopkinson B, ?A Method of Measuring the Pressure Produced in the Detonation of High Explosives or by the Impact of Bullets?, Philosophical Transactions of the Royal Society of London. Series A, Containing Papers of a Mathematical of Physical Character, 213, 437-456, 1914. 6. Robertson R, ?Some properties of explosives?, Journal of the Chemical Society, Transactions, 119, 1-29, 1921. 85? ? 7. Landon JW and Quinney H, ?Experiments with the Hopkinson pressure bar?, Philosophical Transactions of the Royal Society of London. Series A, containing Papers of a Mathematical or Physical Character, 103, 622-643, 1923. 8. Davies RM, ?A critical study of the Hopkinson pressure bar?, Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences, 240, 375-457, 1948. 9. Kolsky H, ?An investigation of the mechanical properties of materials at very high rates of strain?, Proceedings of the Physical Society. Section B, 62, 676-700, 1949. 10. Ramesh KT, Narasimhan S, 'Finite deformations and the dynamic measurement of radial strains in compression Kolsky bar experiments', International journal of solids and structures, 33(25), pp 3723-3738, 1996. 11. Frew DJ, Forrestal MJ, Chen W, 'A split Hopkinson pressure bar technique to determine compressive stress-strain data for rock materials', Experimental mechanics, 41(1), pp 40-46, 2001. 12. Frew DJ, Forrestal MJ, Chen W, 'Pulse shaping techniques for testing brittle materials with a split Hopkinson pressure bar', Experimental mechanics, 42(1), pp 93-106, 2002. 13. Chen W, Zhang B, Forrestal MJ, 'A split Hopkinson bar technique for low-impedance materials', Experimental mechanics, 39(2), pp 81-85, 1999. 86? ? 14. Chen W, Lu F, Zhou B, 'A quartz-crystal-embedded split Hopkinson pressure bar for soft materials', Experimental mechanics, 40(1), pp 1-6, 2000. 15. Chen W, Lu F, Frew DJ, et al., 'Dynamic compression testing of soft materials', Journal of applied mechanics - Transaction of the ASME, 69(3), pp 214-223, 2002. 16. Song B, Chen W, 'Dynamic compressive response and failure behavior of an epoxy syntactic foam', Journal of composite materials, 38(11), pp 915-936, 2004. 17. Subhash G, Liu QL, 'Quasistatic and dynamic crushability of polymeric foams in rigid confinement', International journal of impact engineering, 36(10-11), pp 1303- 1311, 2009. 18. Bouix R, Viot P, Lataillade JL, 'Polypropylene foam behavior under dynamic loadings: Strain rate, density and microstructure effects', International journal of impact engineering, 36(2), pp 329-342, 2009. 19. Prielipp H, Knechtel M, Claussen N, ?Strength and Fracture toughness of aluminum alumina composites with interpenetrating networks?, Materials Science and Engineering A, 197(1), pp 19-30, 1995. 20. Breslin MC, Ringnalda J, Xu L, ?Processing, microstructure, and properties of co- continuous alumina-aluminum composites?, Materials Science and Engineering A, 195(1-2), pp 113-119, 1995. 87? ? 21. Travitzky NA, Gutmanas EY, Claussen N, ?Mechanical properties of Al2O3/Si composites fabricated by pressureless infiltration technique?, Materials letters, 33(1- 2), pp 47-50, 1997. 22. Wegner LD, Gibson LJ, ?The mechanical behavior of interpenetrating phase composites ? I: modeling?, International Journal of Mechanical Sciences, 42(5), pp 925-942, 2000. 23. Wegner LD, Gibson LJ, ?The mechanical behavior of interpenetrating phase composites ? I: resin-impregnated porous stainless steel?, International Journal of Mechanical Sciences, 43(4), pp 1061-1072, 2001. 24. Skirl S, Hoffman M, Bowman K, ?Thermal expansion behavior and macrostrain of Al2O3/Al composites with interpenetrating networks?, Acta Materalia, 46 (7), pp2493-2499, 1998. 25. Mayer H, Papakyriacou M, ? Fatigue behavior of graphite and interpenetrating graphite-aluminum composite up to 10(9) load?, Carbon, 44 (9), pp 1801-1807, 2006. 26. Tilbrook MT, Moon R, Hoffman M, ?On the mechanical properties of alumina-epoxy composites with an interpenetrating network structure?, Materials Science and Engineering A, 393(1-2), pp 170-178, 2005. 27. Etter T, Kuebler J, Frey T, ?Strength and fracture toughness of interpenetrating graphite/aluminium composites produced by the indirect squeeze casting process?, Materials Science and Engineering A, 386(1-2), pp 61-67, 2004. 88? ? 28. Jhaver R and Tippur HV, ?Processing, compression response and finite element modeling of syntactic foam based interpenetrating phase composite (IPC)?, Materials Science and Engineering A, pp 507-517, 2009. 29. Hiel C, Dittman D and Ishai O, ?Composite sandwich construction with syntactic foam core ? A practical assessment of post impact damage and residual strength?, Composites, 24(5), pp 447-450, 1993. 30. Shutov FA, ?Syntactic polymer foams?, Advances in Polymer Science, 73-4, pp 63- 123, 1986. 31. Gupta N, ?A functionally graded syntactic foam material for high energy absorption under compression?, Materials Letters, 61(4-5), pp 979-982, 2007. 32. Gibson LJ, ?Mechanical Behavior of Metallic Foams?, Annual review of materials science, 30, pp 191-227, 2000. 33. Malinowski JZ, Klepaczko JR, 'A unified analytic and numerical approach to specimen behavior in the split-Hopkinson pressure bar?, International journal of mechanical sciences, 28(6), pp 381-391, 1986. 34. Meyers M, ?Dynamic Behavior of Materials?, John Wiley & Sons, Inc., pp 305-307, 1994. 35. Mulliken AD and Boyce MC, ?Mechanics of the rate-dependent elastic-plastic deformation of glassy polymers from low to high strain rates?, International Journal of Solids and Structures, 43, pp 1331-1356, 2006. 89? ? 36. Kim HS, Plubrai P, ?Manufacturing and failure mechanisms of syntactic foam under compression?, Composites Part A: Applied Science and Manufacturing, 35(9), pp 1009-1015, 2004. 37. Butcher RJ, Rousseau CE, Tippur HV, ?A Functionally graded particulate composite: Preparation, measurements and failure analysis?, Acta Materialia, 47(1), pp 259-268, 1998. 38. Gong L, Kyriakides S, Jang WY, ?Compressive response of open-cell foams. Part I: Morphology and elastic properties?, International Journal of Solids and Structures, 42, pp 1355-1379, 2005. 39. Thompson W (Lord Kelvin), 'On the division of space with minimum partitional area', Philosophical Mag., 24, pp 503-514, 1887. 40. Abaqus analysis user's manual v 6.8. 41. Abaqus theory manual v 6.8. 42. Mendelson S, ?Plasticity: Theory and application?, Robert E. Krieger publishing company, pp 98-121, 1983. ISBN 0-89874-582-9. 43. Bettess P, 'Infinite elements', International Journal for numerical methods in engineering, 11, pp 53-64, 1977. 44. Haggblad B, Nordgren G, 'Modeling non linear soil-structure interaction using interface elements, elastic-plastic soil elements and absorbing infinite elements?, Computers & Structures, 26, pp 307-324, 1987. 90? ? 91? ? 45. Viladkar MN, Godbole PN, Noorzaei J, 'Soil-structure interaction in plane frames using coupled finite-infinite elements', Computers & Structures, 39, pp 535-546, 1991. 46. Park WS, Yun CB, Pyun CK, 'Infinite elements for evaluation of hydrodynamic- forces on offshore structures', Computers & Structures, 40, pp 837-847, 1991. 47. Zhao CB, Valliappan S, 'A dynamic infinite element for 3-dimensional infinite- domain wave problems?, International journal for numerical methods in engineering, 36, pp 2567-2580, 1993. APPENDIX A The epoxy resin used for preparing SF and IPC foam samples was characterized under both static and dynamic loading conditions. For completeness, the corresponding data is reported in Fig. A1. The quasi-static response of neat epoxy shows three distinct regions: initial linear elastic region, softening region following yielding, and a plateau stress region. A yield stress of approximately 80?2 MPa is followed by a plateau stress of ~ 48 MPa up to 22% true strain. At higher strains (not shown), there is a densification of stress. (The sample continues to densify up to 70% strain). True strain 0.00 0.05 0.10 0.15 0.20 0.25 True st ress (M Pa) 0 25 50 75 100 125 150 175 200 Dynamic Static Strain rate ~ 1200 / sec Strain rate ~ 0.001 / sec ? Figure A1: Quasi-static and dynamic compressive responses of neat epoxy (?Epo-Thin?) used in SF and IPC foams 92? ? 93? ? The dynamic response, on the other hand, has only two distinct regions. Following an initial linear response with a higher elastic modulus than the quasi-static counterpart, a maximum yield stress of ~185?5 MPa is attained by the neat epoxy sample. Subsequently, the sample shows a monotonic softening response until the stress pulse ends.