as k l..0,- I. r In], ‘ 'T'T’figsgiffiéwuxwég . _ *Z‘v-“(hwh ."s f.wf‘°\ :é" U'nH U01 («H l )4 '}\ I , l‘fi-t‘t..‘,, tit: 25;: ::\:'L:;.;,::%\:’?13;‘ : \ ‘ I “Q 152-: .og‘rsgg. §\( "'53. V..‘0;4:y ‘ , i" ‘1' £6}:\§, c £55k ' 23“- i .s; ’ 3].? . ‘ 2;) ‘ (Raf; ‘.) :J:\ "5' “Mums...” l I .(C'JI': 51$! I '0- 2:? J - ,'.".-lv7. ..'.".,I .1 '/(,"..”"'_:.:-. .u'. I... ’3’... .'__ . ' I" ”I "l:“' / l"|:"'“’ King. ’2' ”I. .' I' ”HM. 1'1'7‘, [I 21,315.! .'.'/,I.’ :Z’w’fié, 4, ("fit/.1 , {I}. lll'lllllllllllllllllllllll L 3 1293 00100 182 LIBRARY Michigan State University This is to certify that the thesis entitled EFFECTS OF MICROWAVE PROCESSING ON THE FIBER-MATRIX INTERPHASE IN COMPOSITES presented by RAJ K . AGRAWAL has been accepted towards fulfillment of the requirements for MASTER OF SCIENCE degree in CHEMICAL ENGINEERING WZ/Dmfl Major professor Date 02' /43 ’33" 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution MSU LIBRARIES ”3—. RETURNING MATERIALS: PIace in book drop to remove this checkout from your record. FINES will be charged if book is returned after the date stamped below. um :65 1'0 “399‘- rf‘fi. . er-W‘ A I" \ . ? 1&5 CV 5 i9gg 7M” ‘ {42:52) EFFECTS OF MICROWAVE PROCESSING ON THE FIBER-MATRIX INTERPHASE IN COMPOSITES BY Raj K. Agrawal A THESIS Submitted to Michigan State University in partial fulfillment of the requirements For the degree of MASTER OF SCIENCE Department of Chemical Engineering 1988 ABSTRACT EFFECTS OF MICROWAVE PROCESSING ON THE FIBER-MATRIX INTERPHASE IN COMPOSITES BY Raj K. Agrawal Microwave curing of composites offers the potential for the development of an extremely fast and versatile method of composite processing. Control of the energy input, duration and location provide an inherent flexibility not present in any other technique. A key factor in achieving this capability lies in creating an "acceptable" interface between fiber and matrix in the cured composite. Differences in dielectric properties of fibers, polymers, and other constituents of the composite could cause selective absorption of microwave energy as well as other identifiable molecular phenomena in the interphase regidn. These phenomena could be: creation of interfacial volatiles; desorption of specific reinforcement surface chemical groups and their reaction with the polymeric matrix; and the O variation in cure morphology in the interphase region. In order to assess the consequences of these possible factors, experiments have been done using a single mode (TE 2.45 GHz) cylindrical microwave cavity with single 111’ fiber composite specimens. After obtaining an optimum cure cycle with microwaves to match that achieved with a conventional thermal cure cycle as measured by tensile tests, dynamic mechanical analysis and differential scanning calorimetry, quantitative measurements of interfacial strength and physical properties have been carried out and compared with the results from thermal cured systems. Under the conditions studied for single fiber specimens, the fiber-matrix interfacial shear strength decreases in both glass-epoxy and aramid-epoxy cases as compared to thermally cured specimens. Graphite fiber-epoxy adhesion, on the other hand, increases significantly in these single fiber studies in microwave processed specimens as indicated by an increase in the interfacial shear strength. The failure mode changes from axial (thermal curing) to matrix failure. To my parents, for all their love, support and understanding throughout my career. iv ACKNOWLEDGEMENTS I would like to express my great appreciation to my advisor, Prof. L.T. Drzal who was generous both with his time and advice during the entire course of this investigation. This 'initiation’ into graduate study has been very enjoyable for me because of his constant encouragement. My thanks are also due to Prof. M.C. Hawley and Prof. J. Asmussen for letting me use their microwave experimental setup. I would like to thank Mr. M.J. Rich, Mr. J. Jow, Mr. S. Iyer and Mr. J. Kalantar for their timely help and suggestions throughout this study. The financial assistance of the Composite Materials and Structures Center and the Defense Advanced Research Project Agency (DARPA) under contract DAAG 46-85-K-0006, P00004 is gratefully acknowledged. TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES 1. INTRODUCTION 2. BACKGROUND AND LITERATURE REVIEW 2.1 Background 2.2 Literature Review 2.2.1 Microwave Effects in Polymer Composites 2.2.2 Interface Effects Under Microwave Radiation 3. EXPERIMENTAL 3.1 Approach 3.2 Material 3.2.1 Thermoset Resin System 3.2.2 Fibers 3.2.2.1 Glass Fiber 3.2.2.2 Kevlar Fiber 3.2.2.3 Carbon Fiber 3 3 Thermal Curing 3 4 Microwave Curing 3.5 Material Properties 3 6 Single Fiber Critical Length Test ESULTS AND DISCUSSION 1 Matrix Properties 2 E-glass Fiber 3 Kevlar Fiber . 4 Carbon Fiber 5 Interphase Temperature in Graphite—Epoxy Specimen 5. CONCLUSIONS AND RECOMMENDATIONS 6. APPENDICES 6.1 Appendix A. DSC and DMA scans 6.2 Appendix B. Critical length data and analysis 7. LIST OF REFERENCES vi PAGE vii viii 10 10 15 27 27 30 32 37 37 4O 41 43 44 50 51 55 55 59 62 65 69 77 81 85 91 Table Table Table Table Table Table Table 1. LIST OF TABLES Material properties of DER 331 cured with stoichiometric amount of mPDA. [26] Electrical properties of DGEBA cured with aromatic diamines. [36] Properties of A84, E-glass and Kevlar-49 fibers reported in literature. Mechanical properties of epoxy cured under thermal and microwave environments. Interfacial shear strength of E-glass fiber specimens cured under thermal and microwave environments. Interfacial shear strength of Kevlar 49 fiber specimens cured under thermal and microwave environments. Interfacial shear strength of graphite fiber specimens cured under thermal and microwave environments. vii PAGE 33 36 39 6O 61 64 67 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 10. 11. LIST OF FIGURES model of the epoxy - reinforcement interphase highlighting the possible components, size and imposed environments acting in this region. [16] Schematic Arrangement of shear debonding (top) and enlarged schematic of a resin droplet on a fiber under the shearing plates (bottom). [13] Typical load-displacement curve for button type test. [21] Diagram of fiber critical length experiments. [22] Single fiber compression test specimen to measure (a) shear strength (b) tensile strength of interface. [20] Tests on unidirectional laminae to measure (a) interlaminar shear strength and (b) transverse. [37] Axial normal stress and interfacial shear stress distribution in a broken fiber fragment. [26] Diagram of possible failure modes in the fiber critical length experiment (a) matrix cracking, (b) frictional stress transfer, (c) shear stress transfer. [27] Dielectric constant vs. temperature of DGEBA cured with mPDA. [36] Dissipation factor vs. temperature of DGEBA cured with mPDA. [36] On-line temperature and complex dielectric constant vs. microwave curing time for DER 332 and DDS at an average input power viii PAGE 18 20 20 21 24 24 29 31 34 35 38 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 13. 14(a) 14 (b) 15. 16. 17(a) 17(b) 17(c) 18. 19. 20. 21. 22. level of 5.72 watts at 2.458 GHz. [35] Schematic illustration of the potential surface chemical groups which have been found on the surface of carbon fibers. [41] Experimental microwave system circuit diagram. Cylindrical brass cavity for microwave curing. Cross-sectional view of the cylindrical cavity (the 6 = 0 plane passes through the excitation probe). [35] Adjustable microwave excitation probe. Schematic diagram of a single fiber interfacial shear strength specimen. Temperature-time profiles of DER 331 and mPDA specimens with different orientation inside the microwave cavity. Temperature-time profiles of DER 331 and mPDA specimens at microwave power level between 10.5-11 watts and varying duration. Temperature-time profiles of DER 331 and mPDA specimens at different microwave power levels. Micrographs of (a) E-glass, (b) Kevlar 49 and (c) AS 4 fiber fracture in thermally cured specimens. Degraded matrix around graphite fiber in microwave'environment. Micrographs of graphite fiber fracture in (a) thermally cured specimen and (b) microwave cured specimen. Diagram of the graphite fiber-epoxy interphase temperature measurement experiment. Temperature-time profiles in the graphite fiber-epoxy interphase region. ix 42 45 46 47 49 52 56 57 58 63 66 68 72 74 Figure Figure Figure Figure Figure Figure Figure Figure 24. 25. 26. 27. 28. 29. 30. DSC scan of uncured DER 331 and mPDA mixture. DSC scan of thermally cured DER 331 and mPDA specimen. DSC scan of partially cured DER 331 and mPDA specimen under microwave environment. DMA scan of thermally cured DER 331 and mPDA specimen. lc/d (glass fiber, thermal curing) lc/d (glass fiber, microwave curing) lc/d (carbon fiber, thermal curing) lC/d (carbon fiber, microwave curing) 81 82 83 84 85 86 87 88 INTRODUCTION The processing of high performance and thick composite materials requires the transfer of energy efficiently into the polymer mass. The introduction of high temperature matrix materials necessitates the use of very high temperature processing equipment for the production of these new generation composite materials if traditional thermal processing techniques are used. Electromagnetic heating of polymer systems is an alternative to conventional thermal processing. Radio frequency waves and microwaves provide the most suitable frequencies for polymer processing. Microwave curing of composites offers the potential for the development of an extremely fast and versatile method of composite processing. Control of energy input, duration and location provides an inherent flexibility not present in any other technique. Limitations in thermal processing due to poor heat conductivity of polymers and accelerated thermal degradation at elevated temperatures due to prolonged heating are of no relevance in microwave processing. Some of the advantages and flexibilities of 2 microwave heating are as follows: 1. Selective and controlled heating: Depending on the dielectric properties, microwave energy is absorbed by materials selectively. This makes it possible for energy to be transferred to the desired location with little coupling to it’s surroundings. Further, by manipulation of the dielectric properties of the material, the amount of energy absorbed can also be controlled. 2. Time sequencing of the process and pulsed heating: There is no thermal inertia in microwave heating. By controlling the duration and intensity of microwave radiation, time sequencing of a process can easily be done. Microwave power can be turned on and off as quickly as desired . 3. Rapid bulk heating and decreased thermal degradation: The microwave heating phenomenon is very different from thermal heating. In thermal heating, the material surroundings are heated first and then conduction of heat takes place from the outside to the inside of the material. In microwave heating, radiation penetrates through the bulk of the material and heat is produced inside the material. 3 The energy transfer is very efficient. The required processing temperature can be achieved very rapidly without long conduction times which also result in less thermal degradation for high temperature materials. The use of microwave heating as an industrial tool is becoming commonplace as a means of conserving energy and increasing productivity. Successful composite processing requires that not only the matrix but the polymer-fiber interphase be processed in an optimum manner. Composite properties are highly dependent on adhesion between fiber and matrix and the interphase region is responsible for the type and level of adhesion. One ‘of the unknown parameters necessary to effectively process uncured polymeric matrix reinforced with continuous fibers is the interaction of microwave radiation with the fiber-matrix interface/interphase. This study investigates the coupling of microwave energy with the fiber-matrix interphase and the resultant effects on fiber-matrix adhesion in composites. Coupling of microwave energy depends on the complex dielectric property of materials. Differences in dielectric properties of fibers, polymers and other constituents of composites could cause selective absorption of microwave energy as well as other identifiable molecular phenomena in the interphase region. These phenomena could be: generation [and release of volatiles from reinforcements or the polymer 4 matrix in the interphase region; alteration of the specific reinforcement surface chemical groups and their reactions with the polymeric matrix; and the introduction of variations in cure morphology in the interphase region. Moisture is present in both the polymeric matrix and on the surface or in the bulk of the reinforcement fiber. Although the quantities of moisture present in the bulk material may be low on an absolute scale, concentration of even one percent moisture at the fiber-matrix interface would be a disruptive force causing the nucleation and growth of voids which are a source for the reduction of composite properties. For graphite or glass reinforcements, the moisture is located on the fiber surface. For a polymeric reinforcement, bulk absorption of moisture is an additional source. Preferential heating of water molecules by microwave radiation would cause water vaporization. Interfacial free energies between fiber and matrix could dictate that the water molecules would preferentially concentrate and nucleate at the interface. The reinforcement fiber surface contains surface chemical groups either native to the surface or placed there through surface treatments. These groups are effective in the promotion of wetting of the fiber surface with the polymer and in chemical coupling of the polymer with the fiber surface in order to promote mechanical, thermal and 5 environmental stability. In the presence of microwave radiation, these surface groups (e.g. carboxylic, phenolic, carbonyl, lactone and hydroxyl) could desorb from the surface or could react with the matrix in a different manner. Either the reaction between the adjacent surface groups resulting in deactivation of the surface chemical species or enhanced reactivity of the surface groups with the matrix will affect the fiber-matrix adhesion. Absorption of microwave energy as heat is directly proportional to the loss factor of the material being processed and is given by [1]: 1 II

= —E2w€ 2 o where

is the rate of power absorption, E0 is the electric field intensity, w is the frequency of radiation and e" is the loss factor. Differences in dielectric properties between fiber and matrix may give rise to energy localization effects in the interphase region. Further the reflection of microwave radiation from the conducting carbon fiber surface in laminated composites could cause different unidentified phenomena in the interphase region. Sharp temperature gradients at the interface may cause the processing of the interface to be different from the bulk. All these could be the source for variations in the local morphology of polymers in the interphase region. 6 All the different possible phenomena occuring in the interphase region as discussed above could be a beneficial or detrimental development depending on the polymer chemistry. The creation of an excessively crosslinked region could increase the efficiency of stress transfer from fiber to matrix while reducing the composite fracture toughness. The promotion of fracture toughness with a reduction in mechanical properties could be a result if matrix crosslink density is reduced. Microwave heating can also be utilized for the processing of high temperature thermoplastics and thermoplastic matrix composites. In order to use microwave radiation in this area, interactions of microwave radiation with thermoplastic matrices need to be explored. Microwave radiation could alter the conformation of the matrix and could also cause morphological changes in the matrix and in the fiber-matrix interphase which would change the structure-property behaviour of the system. In the case of semi crystalline polymers, rate of nucleation, crystal growth rates, and degree of crystallinity could all be affected by the microwave heating mechanisms. Again, all these effects could be beneficial or detrimental and need to be investigated. BACKGROUND AND LITERATURE REVIEW 2.1 BACKGROUND A dielectric material may be defined as one in which it is possible to store electric energy by the application of an electric field and recover the energy when the field is removed. Dielectric materials are usually very poor heat conductors. Heating such substances throughout their volume is very difficult with conventional processes that apply heat to the surface only. Electromagnetic energy in the radio-frequency (RF) range, on the other hand, can act below the surface of a material and heat all parts of the volume simultaneously with substantially greater speed and uniformity of heating than conventional methods. This heating process is termed "dielectric heating". For dielectric heating, two ranges of radio-frequencies are used: a frequency somewhere in 1-200 MHz range, usually known as high-frequency or radio-frequency heating; frequency in 1-5000 GHz range, known as microwave heating. In high-frequency heating, the material to be heated is usually placed between two electrodes where as in microwave 8 heating, energy is applied by specially designed microwave applicators and wave guides. Dielectric heating at any frequency is the result of the interaction of electromagnetic energy with the atomic and molecular structure of materials. Many polymer systems contain various polar functional groups with permanent dipole moments. These dipole moments are randomly oriented due to the thermal brownian motion. The resultant polarization, which is defined as the net dipole moment per unit volume, is zero because of the random orientation of all the dipoles. The incident electromagnetic radiation creates an electric field which interacts with the dipole of the molecule. The molecule rotates, thereby aligning itself with the field and thus increasing the polarization. Polarization would be maximum if all the dipoles were in alignment, but the random molecular motion continuously knocks dipoles out of alignment, keeping the polarization below the maximum level. Molecules can’t rotate instantaneously into alignment with the electric field as they have mass spread over a certain volume and there are retarding forces exerted by the surrounding molecules. The response time of a system can be considered in terms of the decay of its polarization if an electric field is suddenly turned off. The response time of a system determines whether the dipole moments can keep up 9 with the oscillating electric field in an electromagnetic wave . At low frequencies, the time taken by the electric field to change direction is longer than the response time of the dipoles and polarization keeps in phase with the electric field. The field provides energy to make the molecules rotate into alignment. Some energy is transferred to the random motion each time a dipole is knocked out of the alignment and realigned. The transfer of energy is so small, however, that the temperature hardly rises. At high frequency, the electric field oscillates so rapidly that it changes direction faster than the response time of the dipoles. Since the dipoles do not rotate, no energy is absorbed and the system does not heat up. In RF and microwave range of frequencies, the time in which the field changes is about the same as the response time of the dipoles. Dipoles rotate because of the resulting torque they experience but the resulting polarization lags behind the changes in.the direction of the electric field. When the field is at its maximum strength, the polarization may still be low and keeps rising as the field weakens. The. lag indicates that the system absorbs energy from the field. Dielectric heating has been in use for quite some time in medical, food, textile, and paper industries but has 10 recently been applied in the field of composites. Measurements of dielectric properties have been used to monitor chemical reactions in organic materials for more than fifty years but the use of electromagnetic processing of composites is still under investigation. Work in this area has been greatly stimulated in the past decade due to the increasing importance of thermosets and thermoplastics as matrix resins in fiber reinforced composites. 2.2 LITERATURE REVIEW 2.2.1 MICROWAVE EFFECTS ON POLYMER COMPOSITE : Microwave processing of epoxy resins at 2.45 GHz has been studied in conventional microwave (multimode) ovens [2,3], in TE01 wave guides [4] and in a TE10 wave guides [7]. Karmazsin [7] has used continuous and pulsed microwave radiation at 2.45 GHz in TE wave guides to cure epoxy. 10 Springer [8,9] has tried to model the interactions of electromagnetic radiation with polymeric composite and electromagnetic processing of composite materials. A kinetic model of epoxy with DDS has also been proposed by Sheppard [11-13]. The following section will summarize and discuss the significant recent research work in this area. Wilson and Salerno [2] have used a conventional microwave oven at 2.45 GHz to study curing behaviour of ll epoxy with different curing agents at power levels of 245 and 700 watts. Time-temperature profiles of two systems: 100 parts of Epon 828 with 49 parts of T-403 (Jefferson Chemical Company) and 100 parts of Epon 828 with 20 parts of curing agent Z (Shell Chemical Co.) were investigated under the microwave environment. They have identified three different regions in the time-temperature profiles. In the lower temperature zone, the temperature slope increases with time due to the rapid microwave heating. The central portions of the curves are of approximately constant slope or slightly concave downward as most of the epoxy is reacted. In the third zone, the slope increases with time after the microwave source is turned off and this is due to the heat produced by exothermic reactions. They have also measured the dielectric properties of these two gelled, uncured epoxy systems at 23, 60 and 70 0C at frequencies ranging from 1.0 to 2.45 GHz. using a slotted coaxial transmission line. Their experiments indicate that the dielectric properties increase with increasing temperature and are not strong functions of frequency in the desired range. The rate of increase of dielectric properties with increasing temperature is faster in 828-z than 828-T403 .due to the characteristics of the curing agent. The proposed model for the absorption of microwave energy as heat is based on the assumption that the curing 12 epoxies exhibit a bulk loss tangent close to that of pure water, rather than the value measured for the gelled epoxy, and the dielectric constant is in the same range as that of gelled material. They have not considered the problem of microwave reflection by laminate structure or the dependence of dielectric properties on the extent of conversion. Strand [3] has studied the feasibility of microwave processing in the plastic molding industry. He has compared the temperature-time profiles of thermally cured epoxies, polyesters, and polyurethanes at different temperatures with the ones under microwave environment at various power levels. The cure time of epoxy by the 6 KW microwave energy is found to be 30 times less than that of thermal heating at 177 oC. The experiments indicate that fast cure and high efficiency of energy utilization can be obtained by the microwave curing compared to the thermal curing. Gourdenne, et al. [4] have described the thermal interactions between the microwaves and some organic materials to be polymerized and have measured the transfer of energy from the electromagnetic beam to the irradiated samples. They have used a TE wave guide operated at 2.45 01 GHz to study the temperature-time profile of a DGEBA type epoxy cured with DDM (diamino diphenyl methane) at different initial power inputs of 40, 60, 80, and 100 watts. They have also theoretically modeled two microwave heating profiles: 13 before and after polymerization. In the experimental temperature-time profiles the temperature increases regularly in the beginning, then more rapidly until a maximum from which it decreases slowly afterwards. The shape of the peak and the intensity and position in time of its maximum depend on the input power level. The peak is broadened and translated towards the high values of time, whereas the intensity of its maximum is lowered, with decreasing input power. The uncured and cured portions of the temperature-time profiles match well with the simulated curves . Gourdenne and Van [5] have used the above mentioned epoxy-resin curing-agent and wave-guide systems to study the microwave curing of glass fiber filled epoxy systems. Two types of experiments are reported: 1. microwave treatment at given epoxy/glass composition and at variable power levels, 2. microwave treatment at given power values and variable glass composition. The curves of temperature at different power inputs are found to be similar to the curves without glass fiber. Also the temperature profiles at fixed microwave power but at increasing fiber contents are found to be similar to those temperature profiles of epoxy without fiber at decreasing power inputs. They have concluded that the absorbed energy in epoxy is reduced by increasing fiber content the same as by lowering the input power to epoxy without fiber. 14 Lee and Springer [8] have developed a model of electromagnetic waves interactions with organic matrix composites in a wave guide to calculate the complex dielectric constants by measuring the reflectance and transmittance. They have also developed a thermochemical model [9] in order to predict the temperature distribution, degree of cure, resin viscosity, resin content and void sizes during microwave processing of continuous fiber reinforced organic matrix composites. The model is verified by testing Fiberite 82/91348 glass epoxy and Hercules AS/3501-6 graphite epoxy composites using a 2.45 GHz , 700 watt microwave oven (Litton model 1290). Karmazsin and Satre [6] have studied the continuous and pulsed microwave processing of epoxy system and also have compared its thermomechanical properties with that of conventional heat polymerization samples. They have used a 2.45 GHz rectangular wave guide along with a 75 watt continuous or an equivalent (150 watt with 50% cyclic ratio) pulsed microwave power.supply. The epoxy system studied is resin AY 103 with HY 991 as hardener and the conventional heat polymerization cycle used is 1 hr. at 373 K. They have reported that the thermomechanical behaviour of the samples depend on the frequency of the pulsed microwave energy. Tg values of the various microwave cured and conventional heat cured samples are reported to be of the same order. The 15 Young’s modulus of the samples polymerized under microwave environment of 75 watt for 600 seconds are reported to be 10% greater than the Young’s modulus of the sample cured thermally. The authors have not reported the various degrees of polymerization in different microwave and conventional heat polymerization processes. Sheppard and Senturia [11] have demonstrated a relationship between the decrease in the relaxed permittivity and the consumption of polar reactive groups during the curing of DGEBA (Epon 825) with DDS. Isothermal extent of conversion versus time results from DSC are fit to a kinetic model, which is then used to predict reactive group concentrations. Microdielectrometry is used to measure the dielectric constants at frequencies between 0.1 Hz to 10 MHz. They have proposed a combined kinetic/dielectric model which has shown good agreement between experiments and model values up to 70% conversion. They have also determined dipole moments for the reactive groups by fitting the data to an empirically modified Onsager relation for the permittivity. 2.2.2 IRIRRFAQE EEEEQIS URDER MICROWAVE RADIATIO : Composite properties are dependent on the fiber and matrix properties and the specific interactions at the fiber-matrix interface. It is well established that the 16 fiber-matrix interface gives fiber composites their structural integrity and strength. Load is transferred from matrix to fiber through the interface and the interphase region determines the type and level of adhesion between fiber and matrix. The interface in a fiber-matrix composite is common to both fiber and matrix. The interphase exists from some point in the fiber where the local properties begin to change from fiber bulk properties, through the actual interface into the matrix where the local properties again equal the bulk properties. Thus it has physical and mechanical properties which are neither those of the fiber nor those of the matrix. Within this region, various components of known and unknown effect on the interphase can be identified. The fiber may have morphological variations near the fiber surface which are not present in the bulk of the fiber. The surface area of the fiber can be much greater than its geometrical values because of pores and cracks present on the fiber surface. The atomic and molecular composition of the fiber surface can be quite different from the bulk of the fiber. Surface treatments can add or remove surface chemical groups giving rise to a chemically and structurally different region. Fiber coatings to improve the compatibility of the fiber and the matrix is a major source for causing variations in the interphase region. Both chemical and physical bonds exist at the interface and causes the structure of the matrix in the interphase region to be different from the bulk. Further unreacted matrix 17 components and impurities can diffuse to the interphase region altering the local structure. The various characteristics of this interphase region are best illustrated by a schematic model shown in fig. 1 [16]. Characterization of interface region can be done on the basis of single fiber tests and multifiber or composite tests [44]. Several methods has been used to obtain a measure of the stress state and the strength of the bond at the interface. These methods can be divided into two groups. One group deals with direct measurements and involves model studies with a single fiber in a matrix casting. The other group involves indirect measure of the bond strength at the interface and involves testing of actual composites. single fiber tests have the advantage that they are easy to perform and require small quantities of materials. Sample preparation is generally inexpensive. A disadvantage is that the single fiber test is possibly not indicative of the performance of an actual composite. Composite tests on the other hand will give a good indication of the expected composite performanceobut these tests are expensive, time consuming and require large quantities of materials. Two common single fiber tests are the fiber pull out test and the fiber critical length test. A third category of single fiber tests are compression tests. 18 thormot , mocMnIcat and chemical onvlronmonts - matrix morpho'oov - unreacted epochs - lmpurltlos - voids - wrtaco chuntstry - topography - tlbor morphology Figure 1. Schematic model of the epoxy - reinforcement interphase highlighting the possible components, size and imposed envrronments acting in this region. [16] 19 Fiber Pull Out: In a fiber pull out test, a fiber is embedded in a very thin film (normally about 0.5 mm) [17] or head [18] or a disc [19] of polymer. In the bead test, the bead is held between two shearing plates (fig. 2) and the fiber is pulled out. In the disc test, the force is applied on the disc (fig. 3) to cause debonding between fiber and matrix. The force required to remove the fiber is measured with a load cell. The bonding strength is calculated by dividing the measured load by the area of contact of the fiber with the polymer. The area of contact is normally measured in a scanning electron microscope. Fiber Critical Length Test: In the fiber critical length experiment, a single fiber is embedded in a polymeric matrix [20-28]. The specimen is then subjected to increasing strain. Tensile stress applied to the specimen is transmitted to the fiber through shear stresses at the fiber-matrix interface. When the shear stress exceeds the local tensile strength of the fiber, the fiber breaks into fragments inside the specimen (fig. 4). The fiber breakage continues with the increasing strain until the fiber fragments become so small that the matrix can no longer transfer stresses over the fragments to further fracture the fiber. The length of the broken fiber 20 Figure 2. Arrangement of shear debonding (top) and enlarged schematic of a resin droplet on a fiber under the shearing plates (bottom). [18] LOAD I FORCI /80N0 MK . 0N SIAIt FRKZUON DISK / amnauwmwum Figure 3. Typical load-displacement curve for button type test. [21] 21 Ic 2 to Figure 4. Diagram of fiber critical length experiments. [22] 22 fragment is referred to as the fiber critical length (1c). The fiber critical length is an indication of the ability of the polymeric matrix to transfer stress to the fiber . Determination of fiber fragment length to diameter ratio allows an interfacial shear stress r to be determined according to [28] where of is the fiber tensile strength at the fragment length 1C for the fiber diameter d. Due to random defects in the fiber, fiber fragments in an experiment will have a range of lengths. The lengths of the broken fragments will range from one half of the critical length to the critical length. Ohsawa, et at. [23] have used a simple average to calculate lc from the average fiber length la as follows Drzal, et a1. [24] have taken a statistical approach and have used a Weibull distribution to describe the fragment lengths. In order to calculate the interfacial shear stress in 23 the fiber critical length experiment, the tensile strength of the fiber at the critical length must be known. However, the strength of the fiber depends on the flaw distribution of the fiber [31,32]. The longer the length of the fiber, the more defects it will have, and the defects will lower its strength. Rich and Drzal [25] have measured the strength of carbon fibers at their critical lengths (zO.5 mm). Single Fiber Compression Test: Single fiber compression test methods are used to determine the shear strength and the tensile strength at the interface. Two typical compression specimens are illustrated in Fig. 5 [20]. The parallel sided specimen in Fig. 5(a) represents a block of resin containing a short fiber along the central axis. When this specimen is compressed along the axis parallel to the fiber, shear stresses are created at the ends of the fiber because of the different elastic properties of the fiber and the matrix. The shear strength rs of the interface bond can be determined as where "C is the applied compressive stress causing debonding at the ends of the fiber. The neck-down specimen in Fig. 5(b) is used to measure hi1 \ \ \ ¥------o-—-4 no Figure 5. (a) 2i: Single fiber compression test specimen to measure (a) shear strength (b) tensile strength of interface. [20] (I)) Figure 6. L— r L...— 1......1: Tests on unidirectional laminae to measure (a) interlaminar shear strength and (b) transverse. [37] 25 the the tensile strength of the interface. When this specimen is compressed the difference in the Poisson’s ratio of the fiber and the matrix results in a tensile stress in the center of the neck perpendicular to the fiber matrix interface which is given by: ac ("m-Vf) Ef 2 (1+vm) Ef + (1-uf-2uf ) Em where um and uf are the Poisson’s ratios of the matrix and the fiber respectively, 0C is the neck section compressive stress at which debonding occurs and E is Young’s modulus. Broutman [21] used these models to measure the interfacial bond strength. in boron fiber-epoxy composites. He found that the tensile bond strength was approximately 800 psi and the shear about 8000 psi. According to these findings, the bond shear strength is ten times greater than the tensile. Indirect Methods: An indirect approach to measure the interface bond strength is to test unidirectional laminae in such a way that failure occurs in a shear mode parallel to the fibers or in a tensile mode normal to the fibers. Two such tests, the interlaminar shear strength test and the transverse strength test, are illustrated in Fig. 6 [37]. The 26 interlaminar shear test is a three point bending test and requires careful selection of span to depth ratio S/a (Fig. 6a) to insure shear failure along xx’ rather than tensile or flexural failure at point 0. The transverse tensile test shown in Fig. 6b is simple but requires great care for reliable data. Both the interlaminar shear strength and the transverse tensile strength depend on the fiber volume fraction. Results obtained using the interlaminar shear strength test are reported by Goan and Prosen [33] and Daniels, et al. [34]. Daniels, et al. found that the fiber microstructure, as reflected by its modulus and the type of fiber surface treatment, affects interlaminar shear strength and therefore influences the interfacial bond strength. An indication of the interlaminar shear strength dependence on fiber modulus for composites with untreated fibers is illustrated by Goan and Prosen [33]. There is no single method universally accepted to characterize the interface. Depending on the properties of the materials and the situation, one can choose a particular method or combinationoof methods to evaluate the interface. In the microwave field, most of the experiments are carried out from the electrical point of view to determine the dielectric properties of polymeric systems. There is no archival literature on the interaction of microwave radiation with fiber-matrix interface or with the individual fibers itself. EXPERIMENTAL 3 . 1 APPROACH Adhesion between fiber and matrix was evaluated by single fiber critical length tests [22]. When a single fiber specimen is subjected to an increasing tensile load, the fiber starts fragmenting until the remaining fiber fragments are shorter than the crirical length. Determination of the fiber critical length to diameter ratio allows the interfacial shear strength to be determined according to: where is the fiber tensile strength at the fragment ”f length 1C for the fiber diameter d. Due to random defects in the fiber, fiber fragments have a range of lengths. It has been shown that the critical length data fit the two parameter Weibull distribution [24] 27 28 F(x) = 1-exp - - : x>0 where a and B are the two parameters of the distribution and can be evaluated by maximum likelihood methods (see attached computer code). Using these two parameters, a mean value of interfacial shear strength can be calculated as follows: a Var (r) = f F (1- —) - r2 (1- —) 419 a a Valuable information regarding the stress transfer between the fiber and the matrix can be obtained by examining the fractured ends of a fiber fragment under polarized light. The axial normal stress and the interfacial shear stress distribution in a broken fiber fragment is shown in Fig. 7 [26]. Failure modes at the fiber tip have been reviewed by Mullin and co-workers (27, 28). If the fiber-matrix adhesion is poor, failure will occur at the interface as shown in Fig. 8(b). If the matrix is brittle and the adhesion between fiber and matrix is high, matrix cracking will occur with fiber failure as shown in Fig. 8(a). If a ductile matrix is used and the fiber-matrix adhesion is high, the matrix will fail by shear as shown in 29 Hem. .ucmemmuu Honwu coxoun m c“ cowusnfluumfio mmouum Hmonm Hmflommumu:w 0cm mmouum Hmfiuoc Hmfixd .b ousaflm J mm .mm If. i w... m. m" two N 5.». H ~f\\\\\\||lllliy .J1\\\\|IIIIIIIVHHH 8.3182063 6335. 30 Fig. 8(c). Evaluation of the interface in microwave cured specimens as compared with thermally cured specimens requires that the bulk matrix properties be the same in both cases. In order to achieve this the following steps were followed: 1.Establish a database of mechanical properties and extent of cure for the matrix cured thermally under a standard cure cycle. 2.Develop a microwave cure cycle which would produce the same matrix properties as the baseline data. 3.Evaluate and compare the interfacial shear strength and failure mode of various fibers in thermally cured and microwave cured (under the above determined cure cycle) specimens. 3.2 MATERIALS The materials used in this study were chosen very carefully to facilitate the detection of different mechanisms of microwave interactions at the interphase. 31 A a iL 7 High energy radial crack normal to fiber [W— Interface unbonding due to high shgar stress at newly formed on S é). 166%? c EN - Low energy resolved shear 'stress induced tensile cracks in the matrix Figure 8. Diagram of possible failure modes in the fiber critical length experiment (a) matrix cracking, (b) frictional stress transfer, (c) shear stress transfer. [27] 32 3.2.1 THERMOSET RESIN SYSTEM: The matrix material used was an epoxy resin based on diglycidyl ether of Bisphenol—A made by Dow Chemicals (DER 331). This resin cured with the stoichiometric amount of metaphenylenediamine (mPDA) was shown to be very suitable for single fiber critical length tests [24]. The chemical structure of the resin is: 43..-...{.QEQS_C.._I._C.JI00:04....1m2... The curing agent has the following chemical structure: NH, l 0 Physical properties of DER 331 cured with the \ NH, stoichoimetric amount of mPDA are listed in table 1. The dielectric constant and the dissipation factor of thermally cured DGEBA with mPDA at various frequencies and temperatures are shown in fig. 9 and fig. 10 respectively whereas the values for other diamines are listed in table 2 [36]. A typical value of dielectric constant of cured epoxy resin is 3.8 and that of loss factor is between 0.01-0.08. Jow et al. [35] have reported the on-line dielectric 33 Table 1. Material properties of DER 331 cured with stoichiometric amount of mPDA. [26] PROPERTY DER 33] E], GPa (M51) 3.8 (0.55) £2, GPa (M51) 3.8 (0.55) 012 0.35 023, GPa (M51) 1.4 (0.20) a], 10'5/°c (10'5/°r) 68 (32) a2, 10’6/°c (10’5/°F) 68 (32) 5.50 10’ C05 / .1 5.00 V 104 y//// ,0. E 3.12“" ‘1 g/ 10. :3 4.50 \\ . -— —-———- . r<—— 10' U .2! / o L / ‘ _______________——- 10“ / 3.50 / L ___——-0-/ 3.00 1 ‘ 1 I I 1 1 20 40 60 80 100 120 140 160 Temperature. ’C Figure 9. Dielectric constant vs. temperature of DGEBA cured with mPDA. [36] 35 3 1o5 ‘6 .13 C .9 E .9 i3 0004 0002 0001 20 40 60 80 100 120 140 160 Temperature. ‘C Figure 10. Dissipation factor vs. temperature of DGEBA cured with mPDA. [36] 36 Table 2. Electrical properties of DGEBA cured with aromatic diamines. [36] Eutectic Property M I’DA M DA DA DPS blend Arc resistance, seconds ............. 98 97 ...... 100 Dielectric strength. volts/mil ........ 400 - 420 410 430 Volume resistivity, ohm-em ......... 1 x 10“ 0.18 x 10" 2 x 10" >2 X 10' Surface resistivity. ohms ........... >4 x 10'” >4 x 10" 4 x 10' >4 X 10' Dielectric constant at 25"C : At 60 cps ...................... 4.42 4.35 4.22 4.20 At 10’ cps ...................... 4.34 4.27 4.15 4.13 At 10' cps ...................... 3.80 3.72 3.94 3.58 Dielectric constant at 80“C : At 60 cps ...................... 4.74 4.61 4.71 4.47 At 10’ cps ...................... 4.70 4.58 4.68 4.43 At 10‘ cps ...................... 4.33 4.22 4.35 3.86 Dielectric constant at l00"C : At 60 cps ...................... 4.70 4.58 4.71 4.72 At 10' cps ...................... 4.68 4.56 4.62 4.66 At 10‘ cps ...................... 4.42 4.35 4.38 4.12 Dissipation factor at 25"C: At 60 cps ...................... 0.0068 0.0066 0.0044 0.0055 At 10' cps ...................... 0.0183 0.0172 0.0128 0.0133 At 10‘ cps ...................... 0.0346 0.0352 0.0268 0.0318 Dissipation factor at 80°C : At 60 cps ...................... 0.0038 0.0029 0.0084 0.0335 At 10' cps ...................... 0.0035 0.0025 0.0059 0.0523 At 10‘ cps ...................... 0.0353 0.0358 0.0273 0.0352 Dissipation factor. 100°C ° At 60 cps ...................... 0.0030 0.0021 0.0177 0.0407 At 10’ cps ...................... 0.0032 0.0022 0.0076 0.0627 At 10‘ cps ...................... 0.0268 0.0260 0.0214 0.0422 37 properties of the reacting DGEBA and DDS system under microwave environment. Some of their results are shown in fig. 11. 3.2.2 FIBERS: Based on dielectric properties and surface chemistry three different fibers were chosen for reinforcement. 3.2.2.1 Glass Fiber: Unsized E-glass fibers made by Owens-Corning Fiberglas Corp. (lot no. 456, BG-740) were used in this work. Physical properties of this fiber are listed in table 3. E-glass fiber has a surface rich in hydroxyl groups and absorbs water. The surface of the glass fiber contains randomly distributed groups of oxides [37]. Some of the oxides, such 2, Fe203, and A1203 are non-hygroscopic and adsorb water both as hydroxyl groups and as molecular water which as $10 is held to the hydroxyl groups by hydrogen bonding. Other oxides are hygroscopic and when water is adsorbed at the surface they become hydrated. Thus, glass picks up water very rapidly to form a well-bonded surface layer which may be many molecules thick. The presence of moisture on the fiber surface along with the fiber’s low dielectric dissipation factor could produce some interesting effects in the interphase region under microwave environment. Glass 5.51%.? 9' 1 ‘1 3131331910 . ‘0. .33 (.9) 1110151103 3131391310 38 “mm“ .nmo mm4.~ am noon; ms.m «o H0>0H um3oa wanna amaua>m ca um man one man mmo now was» mcfluao o>azouafis .m> ucaumcoa ofluuaaaofio xmamsoo can annuaummemu mafiaico .HH magmas AmmSEEV mE: .Ov .mm .0” .ON . N .0— .OF .0 .O . s s _L p p p — h p p s b s u p p _ s p . p p . h p b L s p p s — h p n p . . OAuN OLNI 2% r . dem 0.? .. 1 .% IO.OO_. i meagoumdrcap (go) elnialadwei 39 TABLE 3 Properties of A84, E-glass and Kevlar-49 Fibers Reported in Literature Property Units A84 E-glass Kevlar-49 Diameter pm 8 10 11.9 Density 103Kg/m3 1.80 2.54 1.45 Young's modulus GPa 235 76 125 Modulus (perpend- GPa 21 76 _ ieular to fiber axis) Tensile strength GPa 3.58 3.45 3.6 Elongation to frac. % 1.53 4.8 2.6 Coeff. of thermal 10‘6/C(10'6/r) -o.11(-0.5) (2.8) -2 expansion parallel parallel parallel 8.54(4) _ 59 radial radial radial Thermal conductivity me'C-1 105 1.04 0.04 (parallel to axis) Specific heat Cal/g C 0.22 0.197 _ Btu/1b F Poisson's ratio 0.25 0.22 0.33 40 fiber is non conductive and its dissipation factor ranges between 0.002-0.005 at 1 MHz. 3.2.2.2 Kevlar Fiber: An aramid fiber Kevlar-49 manufactured by DuPont (lot no. 123481, 7200 denier) was used as the second reinforcement. Kevlar-49 fiber comes supplied with a surface finish. The fiber was washed with absolute ethanol according to the manufacturer’s recommendations prior to use in order to remove the finish. The washing procedure involved three successive washes in absolute ethanol with six hours of soaking in between each wash and then drying in a vented oven for 2 hrs. 0 125 °C. Kevlar is a non conducting fiber. Its dielectric constant is 3.8 and the dissipation factor lies between 0.0002-0.001. The fiber composed of 500 to 700 nm fibrils [38], has a surface that is relatively inert because the surface atoms are covalently bonded into the polymeric structure. Kevlar being very hygroscopic, can absorb upto 3 wt% water [40]. Some of the physical properties of Kevlar-49 are listed in table 3. 41 3.2.2.3 Carbon Fiber: Unsized carbon fiber AS4-12K made by Hercules (lot no. 638-4B) was used as the third reinforcing fiber. Physical properties of this fiber are listed in table 3. Carbon fiber is a very good conductor of electricity and has a very high value of loss factor. Carbon fiber has a highly active surface and readily . absorbs gases which affect the surface properties [37]. A range of active functional groups like carboxylic, lactone, carbonyl, phenolic,and hydroxyl (fig. 12, [36]) are present on the surface. Additional species besides oxygen are also present on the fiber surface [41]. Nitrogen in the form of amine or cyano groups is almost always present on the low heat treatment temperature fiber surface. Traces of elements such as silicon and iron can also be present. Drzal [42] has identified sodium, trapped in the lower modulus fibers from the earlier polymer fiber spinning steps, as being present on the fiber surface and has shown that the sodium has the ability to diffuse to the fiber surface from the bulk of the fiber under moderate elevated temperature conditions. Thus, there are a large number of sites for chemical bonding and there is a large area of contact with the resin. 42 Figure 12. Schematic illustration of the potential surface chemical groups which have been found on the surface of carbon fibers. [41] 43 3.3 THERMAL CURING Single fiber specimens of epoxy were prepared with the help of a silicone RTV-664 eight cavity mold. Standard ASTM 63.5 mm (2.5") dogbone specimen cavities with a 3.175 mm (1/8") wide x 1.59 mm (1/16") deep x 25.4 mm (1") long gage section were molded into a 7.62 cm (3") x 20.32 cm (8") x 1.27 cm (1/2") thick silicone piece. 0.8 mm (1/32") deep sprue slots were molded at both the ends of each dogbone cavity for aligning the fiber axially in the dogbone. Single filaments of desired fiber were taken out of a fiber bundle and placed tightly in the sprues of a dogbone cavity with rubber cement such that the fiber runs across the cavity axially. Epoxy resin (DER 331) and the stoichiometric amount (14.5 phr) of mPDA were weighed in separate beakers. These beakers were then transferred to an oven and heated at 65 0C until the mPDA melted. Then the epoxy resin and the curing agent were mixed thoroughly and degassed in a vacuum oven along with the silicone mold at 65 oC and 29" Hg vacuum. The degassed resin and curing agent mixture was then poured in all the cavities of the mold with the help of a drip rod and to a level just above the height of the mold surface. The assembly was then transferred to an electronically temperature controlled air circulating oven which was 44 preprogrammed for a curing cycle of 2 hrs @ 75 °C followed by 2 hrs 0 125 0C. At the end of the cure cycle, the oven temperature was reduced to room temperature at the given rate. The cured specimens were then taken out of the silicone mold and the extra material was sanded off the surface of the specimens in order to get smooth and uniform specimens. The samples were then stored in a dessicator until ready for analysis. 3.4 MICROWAVE CURING Coupling of microwave energy to polymer specimens was achieved by interactions between a sustained electric field in a coupling applicator and the dipoles contained in the polymer system. The schematic of the experimental setup is shown in fig. 13. There are four basic components in this setup: (1) a microwave power source (2) transmission lines (3) microwave applicator (4) temperature measuring device A cylindrical brass cavity [43] of 7" inner diameter covered by two transverse brass shorting planes, constructed in M80 machine shop, was used as a microwave applicator (fig. 14 (a,b)). The cavity length could be adjusted by 45 .eauuafio uwaaufio Eoumam a>mzou0fis Hmucaewuomxm .mH shaman 8.02 830.. 88.. 588.5 28. .3880 8.8.8.2 38368.2 .225 032...; 8.950 38.825 3. I 8.980 38.825 83 .82.: 8.8.3.5. 8.9.5.2 8328...... 03308.2 .833... 882 .28.. 8.0.). 830d . 52.8% 2... one... 0....quth 2.83.3... Figure 14(a) Cylindrical brass cavity for microwave curing. (U '4' r4 a.) Li :1 U" "-1 in 47 Hun. .Amnoud coaumuaoxo any nusousu mommaa ocean o u m an». >ua>aa H80wupcfiaxo on» ma 30fl> Hac0fluaamimmouo An.va madman L .l J _ m 7 m _ i _ N/i F _ P F M _ 4 WK 11.1 _ _i ~11 7.4 i _ j_ _ n l i. r... e. .r _ _ _ F. .80.. u 48 moving the top plane of the cavity through gears while the bottom plane was .fixed but could easily be removed. The cavity wall and the bottom had several diagnostic holes. Microwave power was coupled into the cavity through an adjustable excitation probe. The coaxial probe had an outer conductor of 1.27 cm. diameter and an inner conductor of 0.442 cm. diameter and was located 3.81 cm. above the bottom plate (fig. 15). A magnetron based fixed frequency (2.45 GHz) generator (Opthos MPG-4) with adjustable duty cycles from 5% to 85%, and power output from 0 to 120 watts, was used as the microwave source. For stability considerations, the power source was operated at higher power levels and the desired power was channelled through a 10 db directional coupler. A circulator was used to protect the source from the reflected power. 50 ohm impedence coaxial cables were used to transmit the power from the source to the cavity. Two 20 db directional couplers were used to decouple the incident and reflected signals. Both incident and reflected signals were attenuated and measured directly by power meters. Silicone dogbone cavities filled with the epoxy mixture were placed in the center of the bottom of the cavity and in alignment with the excitation probe. Due to non-uniformity of the fields inside the cavity, only one dogbone specimen was cured at a time. A fluoroptic temperature measuring Figure 15. Adjustable microwave excitation probe. 50 system (Luxtron Model 750) equipped with four channel measurements was used to continuously monitor the temperature of the reacting mixture at different locations. The cavity was tuned initially to critically couple the microwave energy to the loaded material in the lowest order made TE by moving the the top plate of the cavity and 111’ the excitation probe. Afterwards the cavity was tuned continuously throughout the experiment as the dielectric properties of the reacting mixture inside the cavity were changing continuously. 3.5 MATERIAL PROPERTIES Specimens cured thermally or under the microwave environment were tested for thermal and mechanical properties. The Extent of conversion (a) and glass transition temperature (T9) were determined using Differential Scanning Calorimeter (DSC DuPont 9900). DSC scans of cured and uncured matrices were made at 5 oC/min under nitrogen purge using open pans. In a DSC plot of heat flow vs. temperature, the area underneath the curve of a partially cured material is the heat of reaction required for complete conversion. The extent of conversion can be calculated as follows: Heat of react. for part. cured Extent of conversion = 1 - Heat of react. for uncured matl. 51 Mechanical properties, including tensile strength and tensile modulus were determined using a servohydraulic tensile tester (Material Testing System or MTS 880). ASTM Dogbone specimens were loaded in the hydraulic grips using specially designed supplementary grips to protect the specimens. The load-elongation curves were recorded on a chart recorder at a strain rate of 0.02"/min (about 2%/min). In calculating the strain from crosshead displacement, the gage length of the sample was taken as the distance between the supplementary grips (25.8 mm.). Young’s modulus was calculated from the initial slope of the load-elongation curve whereas the tensile strength to failure was recorded from the machine itself. Dynamic mechanical properties such as flexural modulus (E’) and loss tangent (tan 6) of cured specimens were determined using a Dynamic Mechanical Analyser (DMA, DuPont 9900). Fixed frequency (1 Hz) DMA scans were made at 10 oC/min and 0.2 mm. amplitude. Glass transition data were also obtained from the DMA scans. I 3.6 SINGLE FIBER CRITICAL LENGTH TEST Single fiber dogbone specimens (fig. 16) of the desired fiber were prepared and cured thermally or with microwaves as discussed above. Once fabricated, these specimens were 52 $2R 5 / / 5 SECTION A-A Figure 16. Schematic diagram of a single fiber interfacial shear strength specimen. 53 then tested for fiber critical lengths. Prior to testing, each specimen was examined visually and under the microscope for voids, fiber breaks, additional fibers and other foreign materials, and other fabrication defects which would affect the testing and the results. The testing procedure involved the careful straining of the specimen in a hand operated tensile testing fixture [24], designed and built in the Composite Materials and Structures Center laboratory and Michigan State University machine shop. The fixture was capable of applying enough load to the specimen to cause fracture and incorporated a dial gage to give an indication of the amount of strain. The fixture was then mounted on the stage of a microscope (Olympus BH-2). The clarity of the image was improved by using two cover glasses along with the proper refractive index fluid, one on the top surface of the specimen and one on the bottom surface, to eliminate the effect of specimen roughness. The specimen was then subjected to an increasing tensile load and fiber breaks were monitored in-situ through the microscope. When the. specimen reached a state of constant number of fiber breaks irrespective of the increasing load, the length of each fiber segment was measured using a A calibrated filar eyepiece. Fiber diameter was also measured using the same filar eyepiece. The fiber fractures were then examined under transmitted polarized light and photomicrographs were taken using the attached camera assembly to the microscope. 54 55 RESULTS AND DISCUSSION 4.1 MATRIX PROPERTIES Baseline mechanical properties of the thermally cured specimens under the curing cycle of 2 hrs @ 75 °C followed by 2 hrs. @ 125 0C are listed in table 4. For microwave processing, specimens are cured at various power levels and durations and their temperature-time profiles are shown in fig. 17(A,B,C). In the beginning of the process, the temperature increases very rapidly as the dipoles are free to rotate and thus the coupling of microwave energy is very efficient. As the reaction progresses, the network starts to form and as a result, dipoles are no longer free to rotate. Power absorption decreases and temperature decreases slowly. Orientation of dogbone specimens either parallel or perpendicular to the excitation probe is crucial in microwave curing as the fields inside the cavity are not uniform. Experiments are performed by aligning the specimens 56 .mufi>aa o>asonowe on» mowmcfi casuaucofluo ucauouuflo spa; mcosfiaaam some can Hmm man no moaauoum ofifluiouzumuamaoa Aa.ha anamfim 2:3 m2: mm om m: 0.. m h p p p F b b — b P 1— » h b b p _ p p p QUE OH 4wj can muuas Hatm.oa cmazumn Ho>ma bosom m>mzouafls um mamafioamm «ems use anm mma mo mmafiuoua mefluiauauauomEma Anvha ausmam Cs... “.2... mm ON 2 o. m. o F» F _ p L h b F 17 H p L i» r P L k F L H p p b p — O .8st o. 4-3 2 mmHDZE or win ”3.32:2 I; «.14 mEbZE or his. was woe. is: . WON. we... .1101. 102 M33) Eiéj'fllVEZidl/Bl l .‘u- (. 58 .maa>oa mason o>a3ouofie ucmuouuwo um mcmefiaomm «one can Hnn man no maafiuoum afifiuiouapaumaeae onha ousmflm 423.02; om m. 0. m 0 in L h n h p b b [P — p P h b — b b hi 8 b O 05.3 0.0 To ._ mtg... 0.0 I \ 05.3 0.0 I ..\ mu mtg; 0.0. Ta \\ wt%. 0.: aim \ n 000 m m d . .10... m m m 1 “G 100. B n O u a: .. N was hit - . - h won. 59 parallel to the E field as this configuration resulted in a more efficient coupling of microwave energy (higher temperatures, fig. 17(a)) as opposed to the placement of specimens perpendicular to the E field. Fig. 17(b) shows the temperature-time profiles of specimens cured at a power level between 10.5-11 watts of varying duration. Temperature-time profiles of specimens cured at different power levels are shown in fig. 17(c) . In this graph the comparison of various profiles shows that the peak temperature of a curing cycle increases with .increasing power level. These specimens are then tested for mechanical properties including modulus and tensile strength. The results of mechanical properties are listed in table 4. By comparing mechanical properties at various power levels with the baseline data, it is found that a curing cycle of 9.5 watts for 15 minutes would produce the same mechanical properties as the baseline data. This curing cycle of 9.5 watts for 15 minutes is used to evaluate the interfacial shear strength. 4.2 E-GLASS FIBER Interfacial shear strength (ISS) data for E-glass fiber is tabulated in table 5 and shows a 15% decrease when cured with microwave radiation. ISS in the microwave cured specimen is found to be 8.29 Ksi as compared to 9.73 Ksi for 60 Table 4. Mechanical properties of epoxy thermal and microwave environments. cured under Thermal Curing 2 hrs @ 75 °C, 2 hrs @125 °C. M ierowave curing* 9.0 W for 15 min. 9.5 W for 15 min. 12 W for 15 min. 5.0 W for 15 min. 4.5 W for 25 min. *- 1(- Tensile strength E (Ksi) (Kfi) Mean (8. D.) Mean (8. D.) 12.27 (0.97) 231.63 (10.73) 12.77 (1.09) 238.24 (6.34) 12.72 (0.83) 234.61 (9.27) 12.68 (0.62) 13.79 (0.649) 13.20 (0.40) 219.27 (15.29) 234.48 (6.92) 230.00 (4.1 1) it» wnnzhrs@tzs°cpostcure **‘ Values are low as compared with the literature values due to the use of machine strain in evaluating Young’s modulus. These values are used only for comparison in this study. 61 Table 5. Interfacial shear strength of E-glass fiber specimens cured under thermal and microwave environments. Thermal curing Microwave curing Total Fragments 580 681 Weibull a 4.366 3.6276 Weibull 8 18.4285 22.749 Fiber tensile strength 299 Ksi 299 Ksi (1” gage length) Interfacial shear 9.73 Ksi 8.29 Ksi strength 62 thermally cured specimens. The fracture behaviour of E-glass in epoxy matrix is shown in the transmitted light micrograph (fig. 18(a)). Black spots show the point of fiber failure in the matrix. At the point of fiber breakage, the matrix also fails and a crack starts to grow in the matrix perpendicular to the fiber. This type of failure is termed as matrix failure and is the characteristic of good adhesion between fiber and matrix. No difference is observed in the failure mode of E-glass fibers whether under microwave curing or thermal curing. 4.3 KEVLAR FIBER Similar experiments are done with aramid fiber specimens and the results are tabulated in table 6. The bonding strength also decreases by 15% in the microwave cured specimens. ISS in thermally cured specimens is 3.23 Ksi which goes down to 2.74 Ksi in microwave cured specimens. The ISS value of aramid fiber is much lower than that of glass fiber showing poor adhesion between aramid fiber and epoxy. This is also confirmed by the transmitted light micrograph (fig. 18(b)) of aramid fiber fracture. An aramid fiber does not break at sharp points as in the case of glass fiber. Instead the fiber breaks by fibrillation and the whole dark region is a fiber break. The fracture extends longitudinally along the fiber surface showing poor adhesion. The failure mode in microwave cured specimens is 63 TRANSMITTED LIGHT MICROGRAPH OF E-GLASS FIBER FRACTURE ..*§ “5.: ’97' t i 24 ‘-~ .- : ,‘f r at (a . .1. ‘3 « -mu‘s‘fll'.‘ 71:.» w. . H. TRANSMITTED LIGHT MICROGRAPH OF KEVLAR-49 FIBER FRACTURE TRANSMI’I‘TED POLARIZED LIGHT MICROGRAPH 0F A84 FIBER FRACTURE Figure 18. Micrographs of (a) E—glass, (b) Kevlar 49 and (c) As 4 fiber fracture in thermally cured specimens. 64 Table 6. Interfacial shear strength of Kevlar 49 fiber specimens cured under thermal and microwave environments. Thermal curing Microwave curing Total Fragments 292 186 Critical length 0.904 mm. 1.0645 mm. Fiber tensile strength 487 Ksi 487 Ksi (1" gage length) Interfacial shear 3.23 Ksi 2.74 Ksi strength. 65 observed to be the same as in thermally cured specimens. 4.4 CARBON FIBER The carbon fiber (AS4)—epoxy interphase was observed to have been thermally degraded under the microwave curing conditions of 9.5 watts for 15 minutes (FIG. 19). Carbon fiber, being a good conductor absorbs the majority of power causing a very high temperature rise in the interphase region. This necessitated the development of a lower power curing cycle for carbon fiber specimens. The results of ISS are tabulated in table 7 for this reduced power cycle. The ISS in the microwave cured specimens is found to be 14.15 Ksi as compared to 8.3 Ksi in thermally cured specimen. This shows a significant increase of 70% in bonding strength between the carbon fiber and the matrix. The mode of failure also changes from axial (thermal curing) to matrix (microwave curing). Fig. 20(a) is a polarized light micrograph of a thermally cured specimen showing an axial mode of failure. Light around the fiber shows that debonding has occurred between the fiber and the matrix. Fig. 20(b) shows a carbon fiber failure in a microwave cured specimen under transmitted light. The crack grows in the matrix direction depicting a matrix failure mode. 66 .ucwficoufl>co a>msouowe :fl nonwu dawnnmuv pcsoua xfluumfi omomuooo .ma owsofim 0%,. m. ., WWW. a a... mu 5...... 08.358 0759:. 8.0:: can... 67 Table 7. Interfacial shear strength of graphite fiber specimens cured under thermal and microwave environments. Thermal curing Microwave curing Total Fragments 515 553 Weibull 0 3.5677 3.7891 Weibull e 60.00 34.953 Fiber tensile strength 795 Ksi 795 Ksi (0.3 mm. gage length) Interfacial shear 8.30 Ksi 14.15 Ksi strength 68 TRANSMITTED POLARIZED LIGHT MICROGRAPH OF A84 FIBER FRACTURE (THERMAL CURING) TRANSMITTED LIGHT MICROGRAPH OF AS4 FIBER FRACTURE (MICROWAVE CURING) Figure 20. Micrographs of graphite fiber fracture in (a) thermally cured specimen and (b) microwave cured specimen. 69 4.5 INTERPHASE TEMPERATURE IN GRAPHITE-EPOXY SPECIMENS The immediate question that remains unanswered is what makes the graphite-epoxy interphase so unique under microwave processing that it shows dramatic changes interfacial properties as opposed to the glass-epoxy and Kevlar-epoxy interphases. The conductive nature of the carbon fiber causes the concentration of microwave energy in the form of heat in the interphase region. Carbon fiber acts as a heat source causing some unidentified phenomena in the interphase region which need to be investigated. Various possible interphase parameters, which could be affected significantly under microwave environment affecting the interfacial shear strength and the failure mode are: (1) Material properties of the interphase region could possibly be different from the bulk properties under a microwave environment due to the high interphase temperature. (2) The garphite fiber surface chemistry could be altered under microwave environment resulting in an improved chemical bonding between the fiber and the matrix. (3) Sharp temperature gradient in the interphase region could affect the material properties of the graphite fiber. (4) In microwave curing, graphite fiber absorbs majority of the power as heat and the curing proceeds from the fiber surface out to the bulk matrix. This could result in a 70 stress free interphase with improved adhesion between the fiber and the matrix. The determination of the graphite-epoxy interphase temperature is essential in order to develop an understanding of the interphase. All the above mentioned parameters could very well be affected by the interphase temperature. Evaluation of material properties of the interphase requires the knowledge of this interphase temperature. A three step procedure was developed to address this point. 1. Determination of interphase temperature 2. Curing dogbone specimens under microwave environment at the interphase temperature 3. Evaluation of material properties The interphase region is usually 5-5000 0A thick and so it is very difficult to evaluate its local temperature directly and accurately. There are two indirect methods that could be attempted to.determine the interphase temperature a. The first approach is based on the variability of the resistance of the graphite fiber with temperature. Known electric current is passed through the conducting graphite fiber and the resistance of the fiber is monitored. From the resistance data, fiber temperature can be obtained from a 71 temperature-resistance curve. This method seems to be easy but has it's own problems. Reliable fiber temperature- resistance calibration curve is very difficult to obtain because of variations in the fiber diameter. There are other experimental difficulties with this method. Graphite fiber has to extend beyond the dogbone specimen in order to complete the electric circuit. The extra length of the graphite fiber will be exposed to the microwave in the cavity and will cause erroneous resistance measurements. b. The second method is based on directly measuring the temperatures at different locations in the vicinity of the fiber in the dogbone specimen using fluoroptic microprobes. The interphase temperature can then be evaluated analytically using inverse heat conduction methods. The details of the experimental procedure is given below. Four different holes of about 1 mm diameter were made in the gage section of the silicone dogbone mold (fig. 21). The silicone mold was then placed on the bottom plate of the cavity in such a way othat the plate holes were aligned with the silicone mold holes. Fluoroptic temperature probes were then inserted through the holes and fixed at the different locations and heights across the thickness of the dogbone. The distances of these probes from the mold surface were then measured using a digital caliper. A single graphite fiber was then laid in the mold and it’s distance was also 72 DO 0 O Och—<1: Figure 21. Diagram of the graphite interphase temperature experiment. Silicone Dam Carbon Fiber fiber-epoxy measurement 73 recorded. Heated (70 oC) epoxy resin (DER 331) was then poured in the mold with the help of a drip stick. In order to protect the fluoroptic probes, no curing agent was used with the epoxy, since with the curing agent the epoxy mixture would solidify and destroy the fluoroptic probes. The cavity was then tuned in the TE mode and the specimen 111 was heated under 4.5 watts (low power curing cycle for graphite fiber specimens) of microwave power. Temperatures of all the four probes were then recorded every second with the help of an automated data acquisition system. Temperature profiles at distances of 0.09 mm (T1), 0.34 mm (T2) and 0.59 mm (T3) from the fiber are shown in Fig. 22. The graph shows that the steady state temperature in the interphase region is attained in less than 150 seconds. A steady state heat balance around the fiber allows us to back calculate the fiber surface temperature from the above obtained steady state temperature data. The heat balance can be shown as: d dT —(r—) dr dr II C 74 0:80. 00081.89 xxoaol .188: 3:880 mm .90 Ooomv 60:; com com com 0mg own 0mm ow. @ oo—l .58 00. 0. E 1.. 0m AEE¢M10V N... 4111.6 5......“ T. AEE 00.0. C. Q10 .. .. 30¢ .. r .. iom .................... .- - .. 100 . 100. row. (3) almaledule]~ 75 dT r - = 0 dr T = C1 ln r + C2 Constants C and C are determined from the temperature 1 2 data using the Least Square method and the values are- C1 = -l3.81, C2 = 73.98 The fiber surface temperature is calculated to be— T5 = 150.23 °c (at r = 4 p) This simplistic approach is good for the initial approximation of the interphase temperature as it is based only on the steady state data. Fiber is assumed to be infinitesimally small as compared to the mold dimensions and thus the boundaries are ignored in the heat balance. In practice, the fluoroptic temperature probes were placed at significant distances away from the fiber to ignore the boundaries. Further, the heat absorbed by the fluoroptic probes is also ignored in the heat balance. DSC studies show that it takes about 23 minutes to cure the DER 331 and mPDA mixture completely at the above determined fiber surface temperature (150 0C) as compared with more than 4 hrs. when cured at the bulk temperature (75 0C). This result confirms our observation that in a graphite fiber specimen under the microwave environment, the major 76 difference is that the epoxy curing front moves from the fiber surface out to the bulk. In aramid and glass fiber systems the cure would proceed from the bulk to the fiber surface. The temperature of the fiber might also be expected to be lower than the surroundings. This phenomena could result in a stress free interphase for the carbon fiber which inturn might cause better adhesion between the carbon fiber and the epoxy matrix while conversely providing a ’poorer’ interface for the aramid and glass systems. CONCLUSIONS AND RECOMMENDATIONS CONCLUSIONS The conclusions derived as a result of this study are as follows: . Microwave curing of epoxy is much faster and flexible than any other conventional thermal heating techniques and provides if not superior, then equally good mechanical properties. . In this study for microwave processing at fixed power levels, the temperature increase is very high for an uncured epoxy system and decreases as the reaction progresses, finally becoming constant when the reaction is complete. . Under the curing conditions used in this investigation with single fiber specimens, the fiber-matrix interfacial shear strength decreases by 15% when the glass-epoxy specimens are cured with microwave energy as opposed to conventional 77 78 thermal curing, and in addition the fiber failure mode (matrix failure) does not change. The adhesion in Kevlar-epoxy specimen is lower than the adhesion in a glass-epoxy specimen. Under the curing conditions studied, the bonding strength decreases by 15% in microwave cured specimens as compared to that in thermal cured specimens and the fiber failure mode is not affected by the curing techniques. Microwave curing of specimens with carbon fibers requires extra caution because the conductive nature of the carbon fiber can cause the degradation of the interphase due to selective absorption of microwave energy resulting in localized heating of the fiber. A lower power microwave curing cycle with one carbon fiber is required to cure single fiber specimens compared to the power level for glass-epoxy and Kevlar-epoxy specimens. . Microwave curing increases the level of adhesion in single fiber carbon-epoxy specimens. Under the curing conditions used for carbon fiber specimens, the interfacial shear strength in microwave cured specimens increases by 70% as compared to that in thermally cured specimens. However, the fiber fracture made changes from interfacial failure in thermally cured specimens to matrix failure in microwave cured specimens. 79 . The interphase temperature in the single fiber carbon-epoxy specimens under 4.5 watts of microwave power is found to be 75 oC higher than the bulk temperature which results in a much faster curing reaction in the interphase region than that in the bulk matrix. The curing front moves from the fiber surface out to the bulk possibly resulting in a stress free interphase. RECOMMENDATIONS FOR FURTHER WORK . Further studies need to be done to understand the effects of microwave radiation on the surface chemistry of the carbon fiber and various other fibers. This will give a better insight in understanding the various phenomena occuring in the interphase region under microwave environment. X-ray photo-electron spectroscopy of the fiber surface and transmission electron microscopy of the interphase region will provide needed understanding of the molecular phenomena. . The effect of microwave radiation on carbon fiber mechanical properties needs to be investigated. Being a good conductor, a carbon fiber absorbs microwave power selectively and as a result, its temperature goes up quickly which could affect its mechanical properties. 80 . Multifiber and actual composite specimens need to be cured under microwave environment for complete understanding of the whole process. . Use of matrix additives to promote efficient coupling of microwave energy to the matrix and to the interphase would be of interest in low dielectric loss factor materials. Microwave susceptible fiber coatings to improve interfacial properties of aramid and glass fibers under microwave curing needs to be explored. APPENDIX A 81 .ouauxfis «one can Hmn mmo causes: no caam Omc .mm gunman comm acoaao no..> Hmeocmo goo. meauneousuh 0mm oom om« oo« om 0 fl . L . L L 1L . . 0«.OI . s o s . Tm°.°1 Doom.mh 1 O '4 0' (fi/H) "01$ 393H 0., —- 9-.-. 00mm.mfia r—c---- m—p-o-c- ellln. 00.....000'1'1' I- I," . om.o ZmGOEHHZ 2Hz\uo om Hucmeeou “Nun“ mm\NO\N« .munu cam 00 0mm 0» sz\Uo N ”cacao: 4Haaeumnu no snow Omo .vm madman comm ucoano Oo.«> uncocoo «De. OLDuuLmQEUP com cow own ov« ON“ 00« om T h P > b b b b b > h b b, s... 01 m l _ _. m 00.0.04. 100.01 _ . . I I m Iom.01 . H _ 0x300...01 . m oomv.«v« A111 cofiufimcmcu mmm.o 3 u. . 1 o Tmm o ”M M .m . a ./ . 5 _ UONO.K.mfl It Tom.01 1mv.01 I r 06.01 can uoasdeu .30". won 02 .ucoseoo vo.m. mm\m.\mo .ouou cam coomm cu c«s\ooo om “cacao: 33084 2.: .8883 U m D as 00.: .33 .o.umoi>xoaw ”cask ooeau nausea >xoao .ennsom 83 .HCOEOHMKICO 0>fl3OHUHE ROUGH. casflammm some can 000 mun cause Saaawuuaq mo seam own .mm madman oomm ucoaao do..> Hoeecoo .oo. oeaueeonso» com 0mm oom emu co.“ on o p r n s b r b r L 1? 0" . Ol- 0xuvv..s iv..01 0.00.00 im..01 10..01 100.01 100.01 T (DIM) Hora neaH 100.01 1 0.00.00. 100.01 ' 0.00.0 j No.0 «50: a com 915: M 3:00:00”. «m .m. nm\om\vo .300 can 0. com op 2Hz\ooo.m 8050: 1234mm? .34.... .8888 D m D as o«.vm 82m «c.0zlqu13i 82“. sum 1:: >xoaw omcau w> 4X0 r~———1 94180 U91 oom 0.01. 1111 seawaoqm dams can 000 mmo cause Saaasuocu no seam «to .om gunman Ago.mcaumemusmh om« 0m. 0?" ON« Om“ om ow on iLIIIIII.IIIr.|.OIIIL'.‘ L't‘l‘llrI'Lll L P11 ilFIl ‘ \ es om.or.atuvmunufifias< -101- 1'. I! 111.111! ’. I071...“ 'l'lll.1‘.‘ t‘.‘ O 0‘... lll‘n.0:1-t11ll'lll.l'-Illi.l 1.11.311 . .1..- .,‘ ill. NI « >Uzw30mmm Omme mmumfi hm\m«\m« ”mama cam xix >M\14 Uo oom Oh sz\me o" 4<24m04 34m HLoumcmuo /\ 4 FL 65 mm.“ x No.w x mm.ha «o.ml