Illlfllllll lull]llllllllllllllllllll 1293 00840 1303 lIBRARY Michigan Suite:4 University This is to certify that the - “ thesis entitled A / . ”‘6 Bmvcfltng fiffi [MK Vl lg W7 <3 t HWWW CL PM vs ‘lo {19%;} W/NLV I'm-53 presented by ‘J'TAWID KH/flmmfl has been accepted towards fulfillment - of the requirements for a ,, n a . ,J' ’ ”7‘06 lLE'r 0%: SClC'I’lCC‘ degree in —1-/Cgpf‘~’+l7!h7i 0:; ‘ Che“ worked Elli/H7901" l g Major professor Datefil ”/7’Y? 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution - W ~ ‘_ . ‘——r MSU RETURNING MATERIALS: Place in book drop to LIBRAMES remove this checkout from -_—. your record. FINES will be charged if book is returned after the date stamped below. PLEASE NOTE: In all cases this material has been filmed in the best possible way from the available copy. Problems encountered with this document have been identified here with a check mark 4 . cull 99N9999’N .s P d d o 12. 13. 14. 15. 16. Glossy photographs or pages / Colored illustrations. paper or print Photographs with dark background Illustrations are poor copy . Pages with black marks. not original copy Print shows through as there is text on both sides of page __ Indistinct. broken or small print on several pages Print exceeds margin requirements __ Tightly bound copy with print lost in spine__ Computer printout pages with indistinct print Page(s) lacking when material received. and not available from school or author. Page(s) seem to be missing in numbering only as text follows. Two pages numbered . Text follows. Curling and wrinkled pages Dissertation contains pages with print at a slant. filmed as received Other U-M-I THE BONDING MECHANISM OF ARAMID FIBERS TO EPOXY MATRICES By Javad Kalantar A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemical Engineering Winter 1988 ABSTRACT THE BONDING MECHANISM OF ARAMID FIBERS TO EPOXY MATRICES By J AVAD KALANTAR The interface between aramid fibers and epoxy matrices lacks the level of adhesion attained by other reinforcing fibers. For aramid composites, the interfacial —~~-~ properties attained to date are acceptable, but less than optimal for some -— applications. By developing a basic understanding of the interfacial interactions between aramid fibers and epoxy matrices, fundamental approaches for improving the adhesion can be identified. W ”Three types of interfacial interactions have been examined: (mechanical) ’W Wthat include thermal strain and Poisson’s ratio differences between fiber and matrix, (chemical interactions that include covalent bonding and fiber-matrix wetting, and physrcochemical weak boundary layers. Each of these interactions has been evaluated by manipulating the interface and the curing conditions. Both aramid and carbon fibers have been examined in order to access the interfacial interaction by comparing the behavior of these two fibers. Our results indicate that the adhesion of aramid fibers to epoxy matrices lacks the mechanical and chemical interactions present in carbon-epoxy adhesion. Aramid fibers exhibit an interfacial shear strength as much as four times lower than the expected theoretical value. Direct observation of the aramid-epoxy interface, by transmission electron microscopy, shows fibrillar separations within the fiber surface. This type of interfacial failure suggests that Mamifippxyhadhesion could be due ejectflrfaqsiailw W W W W W TO MOHAMMAD J AHAN iii Acknowledgement This work was accomplished at the Composite Materials and Structures Center (CMSC), at Michigan State University. I want to thank Dr. L. T. Drzal for the opportunity to be exposed to the project and his support, assistance, and patience during my work. I wish to gratefully acknowledge E. I. du Pont de Nemours & Co. for its financial and research support. Finally, I would like to express my appreciation to the Staff of Center for Electron Optics for their help and informative discussions. iv Table of Contents List of Tables .................................................................................................... vii List of Figures .................................................................................................. viii Nomenclature .................................................................................................... ix Introduction ....................................................................................................... 1 Background ........................................................................................................ 4 Composite Interface ........................................................................................... 4 Polyaramid Fibers .............................................................................................. 5 Mechanical Properties of Aramid Fibers .......................................................... 12 Carbon Fibers ..................................................................................................... 15 Epoxy Matrices .................................................................................................. 16 Aramid-Epoxy Adhesion ................................................................................... 21 Aramid-Epoxy Interface .................................................................................... 23 Microstructural Approach ................................................................................. 23 1. Mechanical Interlocking ............................................................... 24 2. Adsorption Interactions ................................................................ 24 3. Electrostatic Attractions ................................................................ 25 4. Polymer Interdiffusion .................................................................. 25 Macrostructural Approach ................................................................................ 26 1. Mechanical Interactions ................................................................ 26 Surface Topography ............................................................... 26 Thermal Stresses .................................................................... 28 Poisson’s Ratio Difference .................................................... 29 2. Wetting .......................................................................................... 30 3. Weak Boundary Layers ................................................................ 30 Theory .................................................................................................................... 33 Experimental ..................................................................................................... 37 Results and Discussion ................................................................................ 42 Thermal Stresses ................................................................................................ 42 Chemical Bonding ............................................................................................. 46 Poisson Contraction ........................................................................................... 48 Fiber Wetting ..................................................................................................... 50 Three dimensional stress model ........................................................................ 50 Electron Microscopy .......................................................................................... 60 Conclusions and Recommendations ................................................... 69 Appendix A: Three-Dimensional Stress Model ....................................... 72 Appendix B: Critical Length distributions ............................................... 76 Appendix C: Thermal Expansion Data ..................................................... 101 Appendix D: Microtoming Technique ........................................................ 108 References ........................................................................................................... 113 vi List of Tables Table 1 Material properties of Kevalr 49 and As4 fibers. ......................................... 13 vii List of Figures \i Figure l The sample conditions and the interactions present at each condition. ............. Figure 2 Structure of polyaramid PP'I‘A monomer. ................................................... Figure 3 Dobb and Johnson model for the aramid pleated structure. ............................. Figure 4 Morgan’s model for the aramid fibrillar morphology. ................................... Figure 5 Model of carbon fiber ribbon morphology. .................................................. Figure 6 Structure of DGEBA epoxy. ..................................................................... Figure 7 Structure of DETA, MPDA, and DETDA curing agents. ............................... Figure 8 SEM micrographs of fracture surfaces of unidirectional AS -4/epoxy and Kevlar 49/epoxy composites. .................................................................... Figure 9 Tensile jig for the critical length technique and a sample mold. ...................... Figure 10 Plot of % thermal shrinkage of the epoxy matrices. ..................................... Figure 11 Plot of experimental interfacial shear strength of untreated carbon and aramid fibers with epoxy matrices cured at different temperatures. ................. Figure 12 Plot of experimental interfacial shear strength of untreated and gold coated carbon and aramid samples made with 75°C and 125°C cured epoxy. ............. Figure 13 Plot of interfacial shear strength of untreated, gold coated, and silicone coated carbon and aramid samples made with the room cured epoxy. ............. Figure 14 Plot of interfacial shear strength of untreated, gold coated, and silicone coated aramid with epoxy matrices cured at different temperatures. ................ Figure 15 Plot of epoxy elastic modulus cured at different temperatures. ....................... Figure 16 Plot of theoretical and experimental critical lengths for aramid and carbon fibers with epoxy matrices cured at different temperatures. ................. Figure 17 Plot of theoretical and experimental average interfacial shear stress over the critical lengths for aramid and carbon samples. ..................................... Figure 18 Optical micrographs of fiber fragment at their critical lengths with bright-field and cross-polarized light. ......................................................... Figure 19 Plot of theoretical average radial stresses for aramid and carbon fibers with epoxy matrices cured at different temperatures. .................................... Figure 20 TEM micrographs of Kevlar 49 and (50/50) epoxy matrices cured at 175°C (lognitudinal cut). ......................................................................... Figure 21 TEM micrographs of Kevlar 49 and room-cured epoxy (radial cut). ................ Figure 22 SEM micrographs of a single Kevlar 49 fiber. ............................................ Figure 23 TEM micrographs of a surface damaged Kevlar 49 in a epoxy matrix cured at 175°C (radial cut). ..................................................................... Figure 24 TEM micrographs of a surface damaged Kevlar 49 in an epoxy matrix crued at 175°C (radial cut). The section is stained with 0504. .................... Figure 25 TEM micrographs of a shear damaged Kevlar 49 in an epoxy matrix cured at 75°C (radial cut). ........................................................................ Figure 26 TEM micrographs of a shear damaged Kevlar 49 in an epoxy matrix cured at 75°C (radial cut). ....................................................................... Figure 27 Diagram of stress distributions around a fiber fragment. ............................... viii 22 39 43 45 47 49 51 52 54 55 57 59 62 63 65 67 68 75 Nomenclature Elastic constant for fiber longitudinal stress, MPa Elastic constant for fiber radial stress, MPa Elastic constant for fiber radial displacement First elastic constant for matrix radial displacement Second elastic constant for matrix radial displacement Axial fiber elastic modulus, MPa Radial fiber elastic modulus, MPa Matrix elastic modulus, MPa Radial fiber elastic shear modulus, MPa Matrix elastic shear modulus, MPa Plane strain bulk modulus, MPa Critical length, um Radial coordinate, urn Fiber radius, tun Curing temperature, °C Glass transition temperature, °C Longitudinal fiber displacement Longitudinal matrix displacement Radial fiber displacement Radial matrix displacement Longitudinal coordinate, um Dimensionless longitudinal coordinate Weibull shape parameter Matrix coefficient of thermal expansion, ppm/°C Axial fiber coefficient of thermal expansion, ppm/°C Radial fiber coefficient of thermal expansion, ppm/°C Weibull scale parameter, um Temperature difference between ambient and oven condition, °C Far field axial strain Fiber fracture strain Matrix thermal strain Longitudinal fiber thermal strain Radial fiber thermal strain Matrix Poisson’s ratio Axial fiber Poisson’s ratio Radial fiber Poisson’s ratio, assume to be equal to vuf Fiber tensile strength, MPa Longitudinal fiber stress, MPa Radial fiber stress, MPa Longitudinal matrix stress, MPa Radial matrix stress, MPa Average interfacial radial stress, MPa Interfacial shear strength, MPa Fiber shear stress, MPa Matrix shear stress, MPa Average interfacial shear stress, MPa Bulk constant INTRODUCTION Aramid fibers have a unique combination of stiffness, high strength, and low ~ density which rivals the properties of inorganic reinforcing fibers such as glass and —— carbon. Advanced composites made of aramid fibers have excellent axial properties - as compared to inorganic fibers, but their off-axis properties are less than optimum for some applications. The off-axis properties of fiber-reinforced composites are generally controlled by the level of fiber-matrix adhesion. In this study, attempts are made to understand the adhesion interactions of aramid fibers with epoxy resins. These adhesion interactions are : mechanical strains, chemical interactions, and effects of physicochemical weak boundary layers. Mechanical strains are caused by thermal shrinkage and Poisson’s ratio differences between fiber and matrix. Chemical interactions include wetting and covalent bonding. The wetting of the fiber with liquid epoxy insures intimate molecular contact between them which is a prerequiste for covalent chemical bonding. Weak boundary layers can reduce the efficency of load transfer at the interface and significantly affect other interfacial interactions. Application of controlled weak boundary layers is used to manipulate fiber-matrix interactions. The approach of this study is to separate and qualitatively analyze the adhesion interactions by modifying the curing conditions and the fiber surface. Figure 1 illustrates the experimental plan. Mechanical Chemical Sample Interactions Interactions Conditions Thermal Poisson Covalent Wetting Stress Contraction Bonding Untreated Fibers & + + + + Elevated Temp. Cured Untreated Fibers & + + + Room Temp. Cured Gold Coated Fibers & + + + Elevated Temp. Cured Silicone Coated Fibers & + + Elevated Temp. Cured Gold Coated Fibers & + + Room Temp. Cured Silicone Coated Fibers & + Room Temp. Cured Figure 1 - The experimental sample conditions and the interactions present at each condition. As shown in Figure 1, various combinations of epoxy curing temperatures and fiber surface modifications have allowed distinguishing different interaction mechanisms. Gold and silicone coating of the fibers are the two types of surface treatments examined. Both treatments introduce physicochemical weak boundary layers which effectively eliminated all the covalent bonds that might be present at the fiber-matrix interface, but also modified the thermodynamics of the fiber surface. Gold coating has produced an inert, yet wettable fiber surface. Silicone coating has produced an inert, but non-wettable fiber surface. Carbon fibers have physical and chemical properties that are distinctly different from aramid fibers. Comparison of these two types of fibers in samples made with the same epoxy has enabled the effects of different interactions to be distinguished. A three-dimensional linear elastic stress model is used to compare the experimental data with their theoretical values. The model has been developed by Whitney et al. [1]. For a single fiber fragment, stress distributions and critical lengths are predicted by the model. A direct observation of the aramid-epoxy interface is carried out by microtoming the single-fiber samples and examining them by transmission electron microscopy. BACKGROUND Composite Interphase The mechanical behavior of composite materials reflects the interactions r— ” between their various constituents. When a load is applied to a fiber reinforced W /’ —.__.—— composite, the load is transferred between matrix and fiber through their interphase. ~- A sgggg interphase promotes greater involvement of the fibers, thus contributing to “r the. composite strength. The fiber-matrix interphase also determines the failure mode of the composite. At high levels of adhesion, the failure would start with matrix cracks, but at lower adhesion levels, failure occurs along the fiber-matrix interphase. For continuous fibgiéinforced composites, the fiber-matrix interphase is symmetric ~— along the fiber longitudinal axsis, and is commonly referred to as the fiber-matrix ~ "interface". Composite behavior is significantly affected by the condition of its interface [2]. In particular, off-axis properties such as interlaminar shear and transverse strength can be improved by increasing interfacial bonding. Improved interfacial -— adhesion also enhances the environmental stability of polymeric composites by —- reducing the formation of weak boundary layers. However, for some applications -- such as fracture toughness, a low level of fiber-matrix adhesion is desirable. In . general, the optimum condition for interfacial strength depends on the particular -— application and its expected loads. In elementary treatments of the composite tensile properties, the effects of the interface are usually ignored [3]. In practice, the interfacial properties have moderate /' to critical influences on many mechanical or thermal properties of the composite. /‘ Studies by Drzal et al. [4] and Owen [5] on carbon fibers with different surface prOperties have demonstrated the significance of interfacial properties on fiber- dominated composite properties. Peters et al. [6] have shown that the mechanical properties of the composite are affected more by the interfacial condition of the fiber-matrix than by the degree of the cure of the matrix. Such observations suggest that the curing cycles of the resin should be optimized with respect to the desired fiber-matrix adhesion rather than optimum matrix mechanical properties. At low levels of fiber-matrix adhesion, fracture toughness and impact resistance of the composites usually increase. A report by Chang et a1. [7] on carbon-epoxy composites with controlled interfaces has shown an inverse relation between interlaminar shear strength and impact resistance of the composite. Similar works by Mai et al. [8], on fracture toughness of Kevlar-epoxy composites has / demonstrated a 200% to 300% greater fracture toughness for Estapol-7008 coated / fiber composites than uncoated fiber composites. The above studies has shown that a ’ very high level of fiber-matrix adhesion can be detrimental to fracture toughness and ’ impact resistance of the composites. Polyararrrid Fibers At molecular levels, the strength of organic polymers is related to the rupture / of their carbon-carbon bonds. In theory, the material strength can be calculated from / the carbon-carbon bond dissociation energy (~ 83 kcal/mole) and the packing of the polymers [9]. However, for most solid materials, the measured strength of the bulk is several orders of magnitude smaller than the theoretical values. The main reason is the existence of flaws or defects in the structure of the material. Misalignment in the orientation of the polymer chains, broken chain ends, and slippage of the chains can lead to stress concentrations on a few bonds which cause chain rupture and catastrophic failures. To reach high material strengths, certain highly ordered *“ polymer morphologies are required. Polymer chain packing, orientation, and / extension significantly affect the material strength. The distribution of flaws and cracks which are detrimental to the strength and must also be minimized. During the past two decades, considerable progress has been made in the production of high performance synthetic fibers [10,11]. These fibers have high degrees of crystallinity and their ultimate properties approach their theoretical maximums. The most successful high performance organic fibers have been prepared 1 from wholly aromatic polymers [12]. These fibers have high modulus, high strength, and are not brittle. Preston [13] has reviewed the development of aromatic polymer fibers. To date, the most successful high-performance organic fibers have been , polyaramid fibers. EJ. du Pont is the major manufacturer of one type of aramid fiber ’ which is marketed under the trade name Kevlar®. Three types of Kevlar fibers are available for specific applications : (1) High modulus Kevlar 49 for composite ' reinforcement, (2) intermediate modulus Kevlar 29 for ropes and fabrics, and (3) tire " cord Kevlar. Since its introduction in 1971, Kevlar has become the major reinforcing fiber for applications where toughness and impact resistance is required [14]. Kevlar fibers consist of extended chains of highly oriented rod-like molecules /" [15,16] formed into fibers with a nominal diameter of 12 um. The aramid monomer i\s_pa_ra;Phenylene Terephthalamide (PPTA) and its chemical structure is shown in Figure 2. The polymer chains are oriented in the fiber longitudinal direction and are hydrogen bonded to each other. The structure of Kevlar fibers is not well documented, but some conclusions about their morphology can be made. 880:9: «Chm 223338 Co 222:5 - N 2ng 0/ ©e ‘lz . o aim ah\o€h Dobb et al. [17], using electron diffraction and electron microscope dark-field image studies, have reported that the structure of Kevlar 49 fiber consists of sheets of polymer chains radially arranged and held together by hydrogen bonding (Figure 3). These sheets are regularly pleated along the axial direction of the fiber, with a pleat angle of about 170°. Over small transitional sections between the pleated sheets, the PPTA polymers are parallel to the axial plane; this feature eliminates the possibility of rotational molecular orientation. Dobb has observed two main types of 500 nm and 250 nm periodicities in the fibers, but near the edges of the fiber he has reported evidence of marked changes in the spacing. Morgan et al. [18] have studied the relation between Kevlar fibers failure process and its structure. While not disputing the pleat morphology of the fiber, they have suggested that the primary structural factor affecting the deformation and failure of the fiber is the concentration and distribution of the supermolecular chain ends within the fiber. Based on the fiber fabrication procedure, structure of the PPTA crystals, microscopic deformation, and fracture topography studies, they have proposed a model of chain-end distribution. In this model shown in Figure 4, chain- end distributions are random in the fiber exterior, but progressively more aligned and clustered in the interior. Morgan’s model suggests a skin-core morphology for the fiber with random chain distribution at the skin and periodic weak planes at the core. The periodicity of the weak planes is about 200 nm, which is the suggested average length of a PPTA rrricromolecule rod. Morgan further suggests that PPTA macromolecules are clustered into cylindrical crystals with 60 nm diameter and 200 run length. A large percentage of macromolecules transverse the weak plane, keeping the continuity of the crystals in the axial fiber direction. It can be inferred from , Morgan’s model that crack propagation can readily occur parallel to the rods and , across the weak planes, leading to fiber fibrillations. Rm 3 - Dobb and Johnson model for the aramid fiber structure. The diagram shows a system of radially pleated polymer planes. A small vertical section is located between each pleat. 10 ] 100 to 1000 nm skin Path of a break Figure 4 - An exaggerated model of aramid fiber morphology proposed by Morgan. The PPTA chains are ramdomly distributed in the fiber exterior and progressively more clustered at the interior. Such a sturcture results in skin-core difference in the fiber. 11 Morgan’s suggested skin-core morphology and chain-end distribution model has been implied by other workers. Skin-core morphology of the fibers has been suggested in studies by Chatzi et al. [19]. Using' the Dignam-Roth theory, their photoacoustic FI‘IR spectroscopic studies have determined a difference in the chain orientation between the exterior and the interior of the Kevlar 49 fibers. Brown et al. [20], using electron paramagnetic resonance studies, have determined that the concentration of the stress-induced free radicals is more than estimated concentration for the fracture surface. They have suggested that the excess free radicals are produced by polymer chain scission at weak planes within the fiber. A fiber fabrication process for the aramid fibers has been reported by Morgan [21]. Aramid fibers are produced by the condensation polymerization of terephthaloyl chloride and p-phenylene diamine [22]. The PPTA is polymerized using a stoichiometric ratio of the reactants. The polymer solution is washed with NaOH to neutralize the HCl formed during polymerization. The solution is then extruded into hot walled cylinders, whereupon the solvent is removed and the shear forces cause the PPTA liquid crystals to orient in the direction of the shear. The resulting yarns are washed and the subsequent stretching and drawing Wtr'eatrnents. - amuse stifmsssarrd “$9.31!: In another report by Morgan et al. [23], the chemical impurities in Kevlar 49 fibers have been investigated. They have determined that there are ~ 0.7% of Na2804 impurities within the fiber, which are the result of sulfuric acid neutralization step. Similar impurity concentrations have been reported by Penn et al. [24]. Morgan has suggested that NaZSO4 residues in the interfibrillar regions are paths for moisture diffusion, which during fiber fabrication can generate microvoids in the fiber. Ashbee et al. [25] have proposed a chemical volume expansion model to describe the hydration expansion which can result in fiber fracture. Based on "salt- weathering" mechanisms in geology, Whalley et al. [26] have confirmed the 12 possibility of a NaZSO4 hydration fracture mechanism. Small angle X-ray scattering studies by Lee et al. [27] have suggested that failure of Kevlar 49 fibers is due to increases in the volume fraction of microvoids and their enlargement along the fiber axis direction. Mechanical Properties of Aramid Fibers Chiao et al. have documented the common properties of aramid fibers and their composites [28]. Table 1 lists the mechanical properties of Kevlar 49 relevant to this study. Kompaniets et al. [29] have examined the statistical aspects of aramid fiber tensile strength. They have reported that for both monofilament and yarns, the tensile strength decreases with an increase in gage length, but the tensile strength of the unidirectional composites was found to be unaffected by the gage length. Their tensile strength data were also independent of the deformation rate (1 to 20 rum/min). Many investigators have attempted to model the strength and modulus behavior of the Kevlar fibers. Knoff [30] has proposed a modified weakest-link model to describe the tensile strength of the fibers as a function of test length. His experimental results have suggested that for aramid fibers below 1 cm gage length, the tensile strength only slightly increases with decreasing length, but above 1 cm the tensile strength is strongly dependent on the gage length. Their model can be considered to describe the aramid fibers as a series of approximately 1 cm uniform strength links. 13 Table 1 - Material properties of Kevlar 49 and AS—4 fibers Property Ref. Kevlar 49 AS—4 E1f (GPa) [83] 119 231 E2f (GPa) [1] 6.9 21 v12f [1] 0.35 0.25 G2f (GPa) [1] 2.6 8.3 (11f ppm/°C [791,[1] -5.72 -2 (12f ppm/°C [791,[1] 65 8.5 ouf (GPa) [83] 3.31 5.86 euf (%) [83] 2.5 1.4 l4 Compressive buckling of aramid fibers has been modeled by DeTeresa et al. [31]. They have suggested that the compressive strength of the fibers should be / proportional to either the shear modulus or the shear strength of the fibers. Another / -' report by DeTeresa et al. [32], describes a new technique to determine the compressive and torsional behavior of the Kevlar 49 fibers. A torsional pendulum has been developed and ratios of (5:1) tensile-to-compressive strength, (17:1) tensile- to-shear strength, and (70:1) tensile-to-shear moduli have been reported. White et al. [33] have proposed an interesting mechanical model to describe compressive buckling of the aramid fiber. Their model involves the classical mechanics spring, dashpot, and rigid rods elements. This model can describe both the compressive buckling and the stress-strain characteristic under tensile loading after the buckling. Fatigue, creep, and failure behavior of aramid fibers have been studied by several workers. Wagner et al. [34] have investigated the creep-rupture statistics of the Kevlar 49 fibers. They have observed that for a given stress level, fibers with 7% lower tensile strength show an order of magnitude lower rupture lifetime. Wagner et al. have suggested a power law dependence for the creep behavior, with different regions of power-law exponents. Lafitte et al. [35] have reported similar multi-region power law creep behavior for Kevlar 29 fibers. In another fatigue study by Lafitte et al. [36] the tensile strength variability of Kevlar 29 fibers has been attributed to the distribution of defects in the fibers. Cook [37] has suggested a kinetic model for Kevlar 49 creep-failure behavior that allows prediction of its rupture lifetime. Several techniques for determining the failure mode of aramid fibers has been reported. Fracto—emission studies by Dickinson et al. [38] have indicated a correlation between total emission of electron and positive ions with the extend of fibrillation in a Kevlar fiber. Acoustic emission (AB) techniques on Kevlar 49 fibers by Hamstad et al. [39] have been less fruitful. The load at which the failure of a 15 single fiber occurred was not directly correlatable with the AE event peak amplitude. An interesting model for longitudinal splitting from surface defects has been set forth by Wagner [40] which allows the prediction of onset and growth of the cracks from surface flaws. Carbon Fibers Carbon fibers are the most widely used high-performance reinforcing fibers. / Carbon fibers have distinctly different properties from aramid fibers. This study has / concentrated on the adhesion properties of the aramid fibers, but comparing aramid and carbon fibers is helpful in highlighting the interfacial interactions. The morphology and the surface properties of carbon fibers will be described briefly. Additional details and more extended discussions on carbon fibers and composites are provided in a review article by Riggs et al. [41] and the reference book by Donnet and Bansal [42]. In a carbon fiber, the carbon atoms form hexagonal rings which are extended by covalent bonds to form a plane of carbon rings called "basal planes". These planes stack and form layers which are held together by weak van der Waal forces; however, there is little alignment between equivalent carbon atoms in adjacent planes. This crystallographic structure of carbon is referred to as "turbostratic graphite". In the plane of the rings, the tensile moduli is very high (910 GPa), but normal to the rings, the moduli is considerably reduced (30 Gpa). The graphite crystals form ribbon-like structures along their basal planes. The general orientation of the ribbons is along the fiber axis. There is a large number of microvoids existing between the ribbons. As the number of microvoids decreases and the ribbons are more aligned, the tensile modulus and strength of the fiber increases. Diefendorf and Tokarsky [43] have shown that the orientation of the ribbons varies from the surface to the center; with more alignment closest to the surface. This morphology shown in 16 Figure 5, results in a fiber with more load-carrying capacity on the surface than in its COI'C. Carbon fibers are usually surface treated in order to enhance their adhesion properties. Comparison of composite properties of surface treated A54 and untreated AU -4 carbon fibers by Drzal et al. [4] has demonstrated the improvement of interface sensitive properties. Most surface treatments tend to modify the interphase by creating fiber surface porosity and increasing the fiber-matrix contact areas. These treatments also improve the reactivity of the surface by forrnin g reactive functional groups, mostly carbonyl and carboxylate. Drzal et al. [44] have demonstrated that another effect of surface treatment is the removal of weak structural layers from the fiber surface which can be a source of interfacial failure. It is usually difficult to assess the relative importance of each mechanism on the improvement of the carbon fiber adhesion. Mechanical properties of carbon fibers vary significantly among different brands due to different precursors and production techniques available. Table 1 includes some of the mechanical properties of the AS-4 carbon fibers utilized in this study. More detailed discussions of the mechanical properties of carbon fibers are available elsewhere [41,42]. A discussion of statistical aspects of carbon fiber strength distributions has been presented by Phani [45]. Epoxy Matrices Epoxy resins are the most common resins used in high-performance composites. Epoxies are therrnoset resins with a wide range of viscosities and can be reacted with a variety of curing agents to obtain a spectrum of mechanical properties. The chemical reaction between epoxy and the curing agent is referred to as "curing". The reaction forms a three-dimensional crosslinked solid structure. The curing mechanisms and kinetics of epoxy systems have been discussed extensively in A carbon ribbon Figure 5 - Arr exaggerated model of carbon fiber ribbon structure. The amplitude and wavelength of the ribbons vary from the core to the surface with increasing alignment near the surface. 18 the literature [46,47]. The mechanical properties of various epoxy systems have been documented by Lee and Neville [48]. ‘(hrring reactions of epoxies do not release any volatiles or excess by-products, thus, they usually have low volume shrinkage. Epoxy resins exhibit a high level of adhesion among polymers and cured epoxy resins have excellent chemical resistance and provide good electrical insulation. The epoxy system used in the present study is Diglycidyl Ether of Bisphenol A (DGEBA) shown in Figure 6. This epoxy has been cured with three types of primary amine curing agents: DiEthyleneTriAmine (DETA); M—PhenyleneDiAmine (MPDA); and DiEthleolueneDiAmine (DETDA) shown in Figure 7. These three curing agents all contain two similar primary amines at each end. The chemistry of the epoxy-amine crosslinking is identical in all three systems, but due to steric hindrances, each system has different kinetics. Cure kinetics of the DEGBA epoxy with other aromatic diamines has been investigated by Moroni et al. [49]. In their study, using thermal analysis data, they have distinguished between different kinetic mechanisms. A report by Wiggins [50] compares the curing behavior of DGEBA and DETDA curing agent with other aromatic curing agents. \ Mechanical properties of DGEBA systems cured with amine curing agents have been investigated by several workers. For the DGEBA/MPDA system, Gupta et al. [51] have reported the dependence of the mechanical properties on temperature, composition, and the relaxation process. Their results have suggested that, in the glassy state, the tensile modulus is dependent on the intermolecular packing, but in the rubbery state, the crosslinking density is the important factor. The effect of crosslinking density on the glass transition of several amine-cured epoxies has also been examined by Bellenger er al. [52]. Another report by Gupta et al. [53] have examined the free volume and its effect on moisture transport in DGEBA/MPDA system. They have determined that below the glass transition, there is no covalent 19 are lmo lame- / \ o .583 < film< vmuwoonb U or O (\l O I") o¢ 0m Aouenbatg Appendix B 87 ratifirlfnuuré 48.3 .4 003 00: 00_m_ 00_0_ 0mm 000 00¢ 00m 0 L h h _ _ b _ L 0 E rim u L r So u o. ooo.m u e . L oom a as n .25.. so go 4 .40 :4 go o. B Aouanbard Appendix B 88 ooo¢ A53 .4 000m, 000m 000E o . h — . _ E L _ . O J T L IL I L L L L Wm rd l E L x J a L . .m L L - m L . 0 L to_ ,A t L Rom u o. L L oooo u e . mam H bmmm H .93.. ago 4442 o .40 :4 §§ ago D .9 Appendix B 89 o. _ . firm? uh. ... m... . in... a 1.3.2.... E 83 .4 oo¢_ oomo GEM? ofiF— COOP on « ¢o44.u.o>< :3 u 5.: E §§§ flu ¢— Aouenbetg L . .. A53 .4 .85 . _ . oo_o_ . 8.2 . . ooxoo L TN o to 1.. No4 a 4.44.4 u .03. to ill/go .. .40 3 so 2.... UR ill! 0 Aouenberg Appendix B 91 ooom 2.. . .22 o 483 .4 ofion oo_oo oo_om oo_o¢ oo_om OOON L _ . . h . _ _ . _ p b . L P O w L l L L -N L 10 man.“ H 003. H .034. é. 2% 3 Egg gash UN. Aouenbetg 463 .4 ooof .oo_¢4. . .oomf . .83. .oo_: oooo m . . . o -N .2 o wo .2. .2: oo4 H 83 n .oéwfi 42:4 u o 4404 444444.45 I : Aouenborfl Appendix B 93 oonr A83 .4 _ .ooof . .oo_o_. .oo_¢_. _ .oomf _ .ooflo l rm to 24. to oo4 H oo¢4 u .3... 2o 6.84 n it 3.x 3480 Bee HU Aouenberg Appendix B 94 000m 050 .4 oown oowo oowo oow¢ oowo ooom — - b p _ _ . p p P h p L L b O L .L Tl F L 1N L L... L L L It? otm H ¢ooo n .92 8.84 .2. 5.4.0 o§ 4.3.28 2.844% NH Aouenberfl Appendix B 95 009. A83 .4 _ .ooLE. . .oo? . .83. _oo.:. . oooo. . .oooo mm H. 83 n .25.. woF Goo ...... 22.4.0 2.44 4.8825 .U . Aouenberd Appendix B 96 FL .L 2.4 H 83 n .3... Cop u it 2.x 838 28 NU fiouonbetg Appendix B 97 000m A83 .4 ooon oooo oooo ooo¢ ooon L _ L L L L L L lr — L L L h L L L O 1N L ..¢ to 2o ..o_ H L H oooo u 25.4.9 8.? u .25 9.x 4.328 384% Nu .. ¢— Aouanbarg Appendix B 98 483 .4 00¢P 002. 00_NF 00__. F 000* p b b h L L L L _ P 0 TN 1¢ 10 mm H 4.4m; H .93.. Adam... 500.4 .3 5.5 034 03.825 WU Aouenber‘g Appendix B 99 483 .4 ooo_ . .oo_t. .oo_o_ _ .822 28$o 2; 2N to 2... 4.2 H 54.4 n .23 ..o 283 68.. 4.. 2.2.4.0 3% 338 How U Aouenberg Appendix B 100 0000 A53 .4 ooo... oooo oooo ooo¢ ooom L r L L L _ _ P L b h L h L L L O . 2N . 2.. L . 2o L 83 H mooo u oz... 383 808 4.. 5.4.0 2.x 4.880 384% DU Aouenbard Appendix C Appendix C Thermal Expansion & Tg Data TMA plots of % thermal expansion of the 75°C, 125°C, and 175°C cured epoxies are presented. The plots also show the T8 of these epoxy systems. Plots of post-curing schedules for the room temperature and 75°C cured epoxy systems are also included. 101 Appendix C 102 A2305 mEE. mm #0 00 mm mm m? t1» 0% mm mm mm #N 0N 0L NF _ a _ _ th _ nu _ P _ _ _ L _ 0 Lu _ _ I 100— 1 (3°) 81n1913du19‘1‘ Appendix C 103 (3,) oanqndoduel f———-—) LL20 asah om m« m« o v 0 ON P L p P b L i P P L s > p L i p ii? P > > r L “00. a o«.o.. oooxczcw Locum / ov1 \IIIIIlulul. 0.0 83:3th .35 iv __ 0. 4 ‘U'I'l' W _ w _ _ r 8.. _ willful. T... _ _ _ ..._ . j — mlll'Illl ‘ 1 fl _ 1 our IIIIIIL 1m.o Y . ocean Lav o.om.on.om.om.ov an . nozoLLoL .aeou soon a a>ou m 1 co“ 1 xi cod-conxw oLueom 1m.o (hm0\«nn Ema m onuuua uot euauto Appendix C 104 3.505 m5? 0 Tfirj (3°) aanqeaedmal 8L 9:39.3th com om“ om“ ov« emu co“ on on ow em rFL ppPPPPPFF>L >PL>> Lererilerr C .m o.exezo.mmto . m. D. .A 1 T a«mmm.o r o.oo.mn m UOmm'OF j o.mo o L: w .u.nn o e: o . ooexea.mn«.u «Lma\«nm mun o.« m.« o.m 86091.13 00‘ SUOUtO (X) Appendix C Loo. unauoeuneok ov« om“ b p L p p p L a » 00“ p P t u.exea.mm«nu vmmn.o oooo.mmu Down“ I ah ooE\51«.hwlu oLau comma o L: VN Lom\omv