V / .IIIIIIII'IIIII‘IIIII ()3 ‘ a This is to certify that the thesis entitled BIOCOMPOSITES FROM ENGINEERED NATURAL FIBERS AND UNSATURATED POLYESTER RESIN FOR HOUSING PANEL APPLICATIONS presented by GEETA MEHTA has been accepted towards fulfillment of the requirements for the MASTER OF degree in Chemical Engineering SCIENCE Ma or Professoyr’s Sbnflure a0, 3009! Date MSU is an Affirmative Action/Equal Opportunity Institution ' -~‘-—--n-n—-o-.--o-o-a-oon-.---.-.-.-.-. LIBRARY ‘ Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 c:lCIRC/DateDue.p65op.15 BIOCOMPOSITES FROM ENGINEERED NATURAL FIBERS AND UNSATURATED POLYESTER RESIN FOR HOUSING PANEL APPLICATIONS By Geeta Mehta A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemical Engineering and Material Science 2004 ABSTRACT BIOCOMPOSITES FROM ENGINEERED NATURAL FIBERS AND UNSATURATED POLYESTER RESIN FOR HOUSING PANEL APPLICATIONS By Geeta Mehta The aim of this project is to make low cost ‘green materials' (biocomposites) for use in various housing panel applications, for example, wall panels, roofs, doors, floors, etc, to compete with and substitute for glass-polyester panel systems used presently. The starting materials for these products are natural fibers as reinforcements, and unsaturated polyester resin as polymer matrix. The project consists of three interconnected parts: modification and engineering of natural fibers to make them suitable as reinforcements in polymer composites, alteration of the polymer composite matrix to make it more ‘green’ while maintaining its properties and performance, and the development of a sheet molding compounding panel process (SMC) as a new way for continuous production of biocomposites. The research on ‘designed’ fiber reinforcement sought to synergistically combine up to four different fibers, which have been effectively surface treated, physically or chemically, to improve the fiber matrix adhesion in the resulting bio- composite. Research was also being conducted to modify the polyester polymer matrix through the addition of modified vegetable oils while maintaining properties and processability. The functionalization of vegetable oils is done in a manner that it is compatible with polyester resin. We have also been successful in developing a SMC process for the continuous manufacturing of bio-composites using chopped natural fibers. C0pyn'ght by GEETA MEHTA 2004 DEDICATION And Life, a little bald and gray, Languid, fastidious, and bland, Waits, hat and gloves in hand, Punctilious of tie and suit (Somewhat impatient of delay) On the doorstep of the Absolute. T. S. Eliot January 1910 This work is dedicated to my father, mother, sisters, and niece, who have given me the wings to fly, strength to persist, inspiration to move forward, faith to be brave in desperate situations, and love to conquer it all. A special dedication to my best fi‘iend, philosopher and guide, who has always been there for me through all thick and thin. I thank all of you, I’m so lucky to have you all. Words can not express the warmth of love that I feel all around me because of you all. ACKNOWLEDGEMENTS I would like to thank Dr. Lawrence T. Drzal, my research advisor, for giving me the opportunity to work on the wonderful area of bio-composites. He has provided me with much opportunity to travel, network, and present my own work in all settings, showing his dedication to my academic and professional growth. My experience of working with him has made me realize exactly how to handle my future graduate students or employees. He was instrumental in writing this dissertation, and I am thankful for his guidance. I am indebted to Dr. Manjusri Misra, my supervisor for all her guidance, encouragement, and support throughout my graduate study. She acted as the driving force behind this research. She provided the opportunity for each student to do his/her own research, yet when asked, she was always willing to provide her knowledge and expertise to assist in understanding the problem. Her professional and personable attitude has made my stay at Michigan State a very rewarding and memorable experience. I have enjoyed her mentorship and fiiendship, and have learned about life as well as research from her. She always has an answer for experimental challenges. Her dedication to the work of the students and postdocs of the CMSC goes above and beyond the call of duty and her expertise has proven priceless. I would also like to thank Dr. Amar K. Mohanty, who not only wrote the original NSF- PATH proposal, but also guided my research. He is one of the most hard working and motivating individuals I have ever come across. When graduate life was bleak, it was his tough love therapy that got me through the dark patches. I have always cherished his honesty and forthrightness, and his wonderful habit of helping people, especially students. I am thankful for this thoughtfulness and his dedication towards my project. Special thanks and recognition also go to Dr. Rigoberto Burgueno, who has always been extremely generous with his time, whether I needed technical or professional advice. I am grateful to him for serving on my committee, for helpful discussions, and for reviewing my research program. He has been a great counselor and I wish him all the luck in the world for his future. Also, I am grateful to acknowledge all of past and present CMSC group members who provided not only scientific, but also moral support, and most of all friendship, and camaraderie throughout my study and research. Some of the special people I would like to acknowledge particularly are, Arief, LaKeya, Mario, Jean Pierre, Praveen, Masud, Hwan Man, Tao, Dana, Wanjun, and, Prasad who made coming into the lab fun and interesting and were always ready to lend a helping hand. My utmost appreciation and thanks are given to my best friend, Pinks, for all the love and support throughout my graduate career. My best friend has been my rock to whom I could always turn for comfort, understanding, and motivation. He made me see the silver lining in the clouds when all I could see was an abysmal black hole. My family has always stood by me, through all the ups and downs, and has been very patient and encouraging in everything I have ventured to do. Ma and Papa sacrificed their lives for us kids. And I wouldn’t be myself without my lovely sisters Gudia and Pooja. My niece J iya has filled our lives with joy. I could not have made it this far without them. vi I am thankful to Mr. Michael Rich and Mr. Kelby Thayer for opening my eyes to US work culture. Kelby also helped me with the SMC line. Bob ‘Jurek and Brian Rook have always been accommodating with all my queries. Jean Rooney and Karen Lillis, who keep things running in the CMSC, whether this calls for ordering materials or remembering somebody on a special day. The work they do is invaluable, and without them, students, staff and faculty alike would be lost. Special thanks are given to many faculty and staff members in the Department of Chemical Engineering and Materials Science, and the Graudate School, Michigan State University for their assistance during my graduate study. Finally, I would like to acknowledge the financial support provided by Dr. Drzal through grants from the NSF -PATH (National Science F oundation-Partnership for Advanced Technologies in Housing, Award number 0122108). I am also thankful to Kemlite Inc. and Flax Craft Inc., for their participation 1 am exceedingly appreciative of the Department of Chemical Engineering and Materials Science of Michigan State University for providing me with graduate assistantship, research assistantships and fellowship. vii TABLE OF CONTENTS LIST OF TABLES ................................................................................... xi LIST OF FIGURES ................................................................................ xiii CHAPTER ONE 1 INTRODUCTION ......................................................................... 1 References ................................................................................... 11 CHAPTER TWO 2 LITERATURE REVIEW ................................................................. 13 2.1 Natural Fibers ..................................................................... 13 2.2 Unsaturated polyester resin ..................................................... 34 2.3 Bioresins .......................................................................... 70 2.4 Processing of composites ....................................................... 77 2.5 Weathering ........................................................................ 89 References ................................................................................... 95 CHAPTER THREE 3 EXPERIMENTAL METHODS ......................................................... 110 3.1 Materials .......................................................................... 1 10 3.2 Surface Modifications of natural fibers ....................................... 113 3.3 Bioresin ........................................................................... 118 3.4 Fabrication of composites ...................................................... 122 3.5 Testing ............................................................................. 151 3.5.1 Natural fiber ............................................................. 151 3.5.2 Bioplastics and biocomposites ....................................... 158 CHAPTER FOUR 4 RESULTS ON ENGINEERED FIBERS .............................................. 173 References ................................................................................. 231 CHAPTER FIVE 5 RESULTS ON MATRIX MODIFICATION ......................................... 233 5.1 Curing Kinetics ................................................................... 233 5.2 Bioplastics from Bioresins ...................................................... 253 References ................................................................................. 294 viii CHAPTER SIX 6 RESULTS ON NEW PROCESSING .................................................. 296 Biocomposites from SMC ............................................................... 296 References ................................................................................. 324 CHAPTER SEVEN 7 RESULTS ON DURABILITY ......................................................... 325 7.1 Moisture Absorption ............................................................ 325 7.2 QUV Accelerated Weathering ................................................. 330 References ................................................................................. 340 CHAPTER EIGHT 8.0 CONCLUSIONS AND FUTURE WORK ............................................ 341 8.1 CONCLUSIONS ................................................................ 341 8.2 FUTURE WORK ................................................................ 347 ix LIST OF TABLES Table 2.1.1 Typical properties of glass fiber and natural fibers ............................... 18 Table 2.2.] Cost ofvarious polyester resins. .35 Table 2.2.2 Examples of monomers used for UPE system cross linking .................... 39 Table 2.2.3 Examples of inhibitors used for UPE system ..................................... 42 Table 3.1.1 Reinforcements for composites made in this project ........................... 111 Table 3.1.2 Materials and suppliers ............................................................ 112 Table 3.2.1 Pre-treatment and post treatment for surface treatments of the natural fibers ................................................................................................ 1 17 Table 3.3.1 Bioresins from grafting of vegetable oils ........................................ 119 Table 3.3.2 Bioplastics obtained from grafled oils ............................................ 120 Table 3.4.1 Samples made using silicone molds .............................................. 124 Table 3.4.2 Samples made using compression molding ..................................... 127 Table 3.4.3 Samples made using SMC line .................................................. 142 Table 4.1.1 Thermogravimetric results for surface treated hemp fibers ................... 175 Table 4.1.2 Elemental composition of surface treated hemp fibers (from XPS analysis) ............................................................................................. 17 8 Table 4.1.3 Elemental ratios of surface treated hemp fibers (fi'om XPS analysis) ....... 178 Table 4.2.1 Thermogravimetric results for surface treated pure hemp fibers. . . . . . ....196 Table 4.2.2 Elemental composition of surface treated pure hemp fibers (from XPS analysis) ............................................................................................. 199 Table 4.2.3 Elemental ratios of surface treated pure hemp fibers (from XPS analysis) ............................................................................................. 199 Table 4.3.1 Thermogravimetric results for surface treated kenaf fibers . .......217 Table 4.3.2 Elemental composition of surface treated hemp fibers (from XPS analysis) ......................................................................................................... 219 Table 4.3.3 Elemental ratios of surface treated hemp fibers (from XPS analysis) . ..219 Table 5.1.1 Comparison of maximum conversion at different temperatures from FTIR experiments .......................................................................................... 252 Table 6.1 Maximum degradation temperatures for natural fibers .......................... 301 Table 6.2 Atomic concentrations on the surfaces of fibers used for SMC line ........... 301 Table 6.3 Ratio of atomic concentrations of fibers ........................................... 302 Table 7.1.1 Maximum degradation temperature of the bioplastics ........................ 330 xi LIST OF FIGURES Figure 1.1 Schematic representation of the research plan for this project .................... 6 Figure 2.1.1 Examples of natural fibers .......................................................... 14 Figure 2.1.2 Classification of biofibers .......................................................... 16 Figure 2.1.3 Comparison of cost and density between glass fibers and biofibers. . . . . .....16 Figure 1.1.4 Chemical structure of cellulose molecules: Poly-B(1,4)-D-Glucose. . . ..18 Figure 2.1.5 Comparison of moduli of glass fibers and biofibers ............................ 20 Figure 2.1.6 Comparison of natural fiber (NF) and glass fiber (GF)—unsaturated polyester (UPE) composites ................................................................................... 23 Figure 2.1.7 North American major end markets of natural fiber composites, year 2000 ................................................................................................... 24 Figure 2.2.1 Matrix pattern of polymer composites ............................................ 35 Figure 2.2.2 A representative chemical structure of unsaturated polyester ................. 36 Figure 2.2.3 Preparation of polyester resin ...................................................... 38 Figure 2.2.4 Chemical structure of methyl ethyl ketone peroxide (MEKP) ................ 40 Figure 2.2.5 Chemical structure of benzoyl peroxides (BPO) ............................... 40 Figure 2.2.6 Reaction mechanism of polyester resin with cobalt activator . .....41 Figure 2.2.7 Crosslinking mechanism of UPE resins .......................................... 43 Figure 2.2.8 Curing reaction rate of UPE ........................................................ 45 Figure 2.2.9 A depiction of a simple IPN material ............................................. 69 Figure 2.2.10 Soybean plant ....................................................................... 72 Figure 2.2.10 Flax plant and flower .............................................................. 73 Figure 2.2.1] Photograph of ESO-clay nanocomposite showing high flexibility. . . . . ....77 Figure 3.1.1 Hemp mat with filler (Bastrnat 115) obtained from F laxCrafi Inc .......... 112 Figure 3.2.1 Vacuum oven for drying biofibers ............................................... 114 xii Figure 3.4.1 Conventional oven for curing silicone molds ................................. 124 Figure 3.4.2 Carver compression Press ......................................................... 126 Figure 3.4.3 Picture of bioplastic samples made using compression molding ........................................................................................................ 129 Figure 3.4.4 Scheme of BCSMCP process .................................................... 132 Figure 3.4.5 Scheme of vibratory feeder and screw feeder ................................ 134 Figure 3.4.6 Mixing of resin components in the Ross mixer .............................. 141 Figure 3.4.7 BCSMCP process for natural fiber-polyester resin composites ............. 143 Figure 3.4.8 Feeding natural fibers though a vibratory feeder .............................. 145 Figure 3.4.9 An E-Glass-CaCO3-UPE composite ............................................. 147 Figure 3.4.10 Composite samples made using SMC line .................................... 148 Figure 3.4.11 Composite samples made using SMC line .................................... 149 Figure 3.4.12 Composite samples made using SMC line 150 Figure 3.5.1 Hi-Res TGA 2950 (TA instruments) in CMSC lab ............................ 152 Figure 3.5.2 Perkin Elmer spectrum 2000 FT-IR in CMSC lab . .......153 Figure 3.5.3 PHI 5400 ESCA (XPS) system in CMSC lab .................................. 155 Figure 3.5.4 TA 2920 Modulated DSC (TA instruments) in CMSC lab. . . . . . . ...157 Figure 3.5.5 Sputter coater(1efi) and ESEM (right) Model no. 2020 at CMSC lab. . ...158 Figure 3.5.6 TA DMA 2890 in CMSC lab 161 Figure 3.5.7 UTS in flexural testing mode at CMSC lab16l Figure 3.5.8 Impact testing machine (left) and notching machine (right) at CMSC lab .................................................................................................... 162 Figure 3.5.9 United Corporation “SFM-20” Test System ................................... 163 Figure 3.5.10 TMA 2980 (TA instruments) in CMSC lab .................................... 164 Figure 3.5.11 (a) Working of a contact mode AFM. (b) MultiMode Scanning Probe Microscope (SPM) from Digital Instrument ................................................... 165 xiii Figure 3.5.12 Humidity chamber in CMSC lab ............................................... 167 Figure 3.5.13 QUV weatherometer in Dr. Kamdem’s lab ................................... 169 Figure 3.5.14 Open panel of QUV weatherometer ........................................... 169 Figure 3.5.15 CIELAB system for color measurement ...................................... 163 Figure 3.5.16 Surface profilometer for roughness measurement ........................... 170 Figure 3.5.17 Weathered samples, on left, a hempmat-UPE biocomposite, on right, a BBSG—GFC-UPE biocomposite ................................................................. 171 Figure 4.1.1 Schematic representation of the compression molding process ............. 173 Figure 4.1.2 TGA of surface treated hemp fibers ............................................ 174 Figure 4.1.3 Derivative thermo-gravimetric analysis (DTGA) of surface treated hemp fibers ................................................................................................. 175 Figure 4.1.4 DSC of surface treated hemp fibers .............................................. 176 Figure 4.1.5 Proposed reaction for surface treatment of hemp fibers. . . . . . . . . .....179 Figure 4.1.6 Proposed reaction for curing of surface treated hemp fibers and UPE. . ...180 Figure 4.1.7 Proposed reaction for curing of surface treated hemp fibers and UPE. . l 81 Figure 4.1.8 Optimization of fiber volume fraction by evaluation of mechanical properties ............................................................................................ 182 Figure 4.1.9 Comparison of tensile properties of surface treated composites. . . . . . ..184 Figure 4.1.10 Comparison of flexural properties of surface treated composites... . . . . ...186 Figure 4.1.11 Impact strength of surface treated composites ................................ 188 Figure 4.1.12 Typical storage modulus curves of surface treated composites ........... 189 Figure 4.1.13 Typical tan delta curves of surface treated composites ..................... 190 Figure 4.1.14 Storage modulus of surface treated composites at 40 OC .................... 191 Figure 4.1.15 ESEM micrographs of surface treated hemp mat fibers ..................... 192 Figure 4.1.16 Schematic representation of the fiber matrix interface and various factors affecting the fiber-matrix adhesion .............................................................. 193 Figure 4.2.1 TGA of surface treated pure hemp fibers ........................................ 195 xiv Figure 4.2.2 Derivative therrno-gravimetric analysis (DTGA) of surface treated pure hemp fibers ........................................................................................... 196 Figure 4.2.3 DSC of surface treated pure hemp fibers ....................................... 197 Figure 4.2.4 Optimization of styrene content by evaluation of flexural properties ............................................................................................ 200 Figure 4.2.5 Optimization of styrene content by evaluation of mechanical properties ............................................................................................. 201 Figure 4.2.6 Optimization of styrene content by evaluation of mechanical properties ............................................................................................. 202 Figure 4.2.7 Comparison of tensile properties of surface treated pure hemp mat based composites ........................................................................................... 203 Figure 4.2.8 Comparison of flexural properties of surface treated composites ........... 205 Figure 4.2.9 Impact strength of surface treated composites ................................. 207 Figure 4.2.10 Typical storage modulus curves of surface treated composites. . . . . . . . ....208 Figure 4.2.11 Typical tan delta curves of surface treated composites ...................... 210 Figure 4.2.12 Storage modulus of surface treated composites at 40 0C .................... 211 Figure 4.2.13 ESEM micrographs of surface treated pure hemp mat fibers. . . . . . . . ........212 Figure 4.2.14 ESEM micrographs of tensile fractured surfaces of biocomposites from pure hemp mat fibers ............................................................................... 214 Figure 4.3.1 TGA of surface treated kenaf fibers ............................................... 215 Figure 4.3.2 Derivative therrno-gravimetric analysis (DTGA) of surface treated kenaf fibers ................................................................................................... 216 Figure 4.3.3 DSC ofsurface treated kenaffibers................................................218 Figure 4.3.4 Comparison of flexural properties of surface treated kenaf based composites ............................................................................................ 220 Figure 4.3.6 Specific Flexural Properties of Kenaf-UPE composites ...................... 222 Figure 4.3.5 Notched Izod Impact strengths of kenaf-UPE composites ................... 224 Figure 4.3.6 Typical storage modulus curves of surface treated kenaf composites. . ....225 Figure 4.3.7 Typical tan delta curves of surface treated kenaf composites. ................. 226 XV Figure 4.3.8 Storage modulus of Kenaf-UPE composites at 40 OC.. ............................. 227 Figure 4.8.9 SEM micrographs of surface treated kenaf fibers. ................................... 229 Figure 4.3.10 SEM micrographs of surface treated kenaf fibers based composites ....... 230 Figure 5.1.1 Temperature sweep for curing UPE from 25 0C to 160 0C. . ...............235 Figure 5.1.2 Temperature sweep for curing UPE from 35 0C to 300 0C..... ..........2.35 Figure 5.1.3 Isothermal DSC cure ofUPE at 50, 60,70, 80 0c. . 236 Figure 5.1.4 Isothermal DSC cure ofUPE at 90, 100, 110, 120 0C237 Figure 5.1.5 Isothermal DSC cure ofUPE at 130, 140, 150, 160 0C. 237 Figure 5.1.6 Residual heat flow (temperature sweep) DSC UPE curing .................. 238 Figure 5.1.7 Comparison of isothermal heat of reaction at different temperatures. . .....239 Figure 5.1.8 Comparison of maximum conversions at different temperatures. . . . . ........239 Figure 5.1.9 FT IR spectra of pure UPE. . . .. ................................................................... 242 Figure 5.1.10 FTIR spectra of pure styrene .................................................................... 243 Figure 5.1.11 FT IR spectra of UPE system at 150 0C at t=0 minutes and F 250 minutes ...................................................................................... 244 Figure 5.1.12 Fractional conversion at 50 0C from in situ F TIR study .................... 245 Figure 5.1.13 Fractional conversion at 60 0C from in situ F TIR study .................... 245 Figure 5.1.14 Fractional conversion at 70 0C from in situ FTIR study .................... 246 Figure 5.1.15 Fractional conversion at 80 0C from in situ FTIR study .................... 247 Figure 5.1.16 Fractional conversion at 90 0C from in situ FTIR study .................... 247 Figure 5.1.17 Fractional conversion at 100 0C from in situ FT IR study ................... 248 Figure 5.1.18 Fractional conversion at 110 0C from in situ FTIR study. . . . . . . . . . . . . . ......248 Figure 5.1.19 Fractional conversion at 120 0C from in situ FTIR study ................... 249 Figure 5.1.20 Fractional conversion at 130 0C from in situ FTIR study. . . . . . . .............250 Figure 5.1.21 Fractional conversion at 140 0C fiom in situ FTIR study ................... 250 xvi Figure 5.1.22 Fractional conversion at 150 0C from in situ F TIR study .................... 251 Figure 5.1.23 Fractional conversion at 160 0C from in situ FTIR study .................... 252 Figure 5.2.1 FT IR spectra of pure acrylonitrile ................................................ 255 Figure 5.2.2 FT IR spectra of pure castor oil ................................................... 255 Figure 5.2.3 FT IR spectra of pure soybean oil ............................................... 256 Figure 5.2.4 FTIR spectra of pure methyl ester of soybean oil ............................ 256 Figure 5.2.5 FTIR spectra of grafted castor oil ............................................... 257 Figure 5.2.6 FTIR spectra of grafted soybean oil ............................................ 257 Figure 5.2.7 FTIR spectra of grafted methyl ester of soybean oil ......................... 258 Figure 5.2.8 Scheme showing probable reaction between vegetable oil and reactive monomer, resulting in modified vegetable oil ................................................ 259 Figure 5.2.9 Impact strength of the bioplastics ............................................... 260 Figure 5.2.10 Flexural properties of the bioplastics .......................................... 261 Figure 5.2.11 Impact strength of biocomposites containing bioresins .................... 262 Figure 5.2.12 Flexural properties of biocomposites containing bioresins ................ 262 Figure 5.2.13 ESEM micrographs of impact fractures surface of bioplastic containing MSO-PBMA-UPE, a) Magnification 2000X, Scale 25 um, b) 250 X, .................... 263 Figure 5.2.14 ESEM micrographs of impact fractures surface of bioplastic containing MSO-UPE, a) Magnification 300 X, Scale 150 um b) Magnification 2000X, Scale 25 um ................................................................................................... 264 Figure 5.2.15 a) Impact fiactured surface of neat polyester resin, scale 20 pm .........265 Figure 5.2.15 b) Impact fiactured surface of bioplastic (MSO-UPE-PBMA), scale 20 um ......................................................................................................... 265 Figure 5.2.15 c) Impact Fractured surface of bioplastic (MSO-UPE-PBMA), scale 2 pm . . ...................................................................................................... 265 Figure 5.2.16 AFM picture of neat unsaturated polyester resin (deflection image, scan size: 5 pm X 5 pm) ............................................................................... 266 Figure 5.2.17 AFM picture of bioplastic (MSO-UPE-PBMA) (deflection image, scan size: 60 um X 60 um) ............................................................................ 267 xvii Figure 5.2.18 AFM picture of bioplastic (MSO-UPE-PBMA) (deflection image, 9 mm X 9 pm) ................................................................................................ 268 Figure 5.2.19 TGA of pure soybean oil phosphate ester polyol (SOPEP) ................ 271 Figure 5.2.20 DSC of pure soybean oil phosphate ester polyol (SOPEP) ................ 272 Figure 5.2.21 FTIR spectra of pure maleic anhydride (MA) ............................... 273 Figure 5.2.22 FTIR spectra of pure soybean oil ............................................. 274 Figure 5.2.23 FTIR spectra of pure SOPEP ................................................... 274 Figure 5.2.24 FTIR spectra of modified SOPEP ............................................. 274 Figure 5.2.25 Proposed reaction mechanism for. grafting of SOPEP with MA .......... 276 Figure 5.2.26 Storage modulus of bioplastics made with GFTSOPEPI ................... 278 Figure 5.2.27 Storage modulus of bioplastics made with GFTSOPEP2 .................. 278 Figure 5.2.28 Storage modulus of bioplastics made with GFTSOPEP3 .................. 279 Figure 5.2.29 Tan delta of bioplastics made with GFTSOPEPI ........................... 279 Figure 5.2.30 Tan delta of bioplastics made with GFTSOPEP2 ........................... 280 Figure 5.2.31 Tan delta of bioplastics made with GFTSOPEP3 ........................... 280 Figure 5.2.32 Glass transition temperature of bioplastics made with GFTSOPEPl ....281 Figure 5.2.33 Glass transition temperature of bioplastics made with GFTSOPEP2 ....282 Figure 5.2.34 Glass transition temperature of bioplastics made with GFTSOPEP3 ....282 Figure 5.2.35 Crosslinking density of bioplastics with GFTSOPEPl ..................... 284 Figure 5.2.36 Crosslink density of bioplastics made with GFTSOPEP2 .................. 284 Figure 5.2.37 Crosslink density of bioplastics made with GFTSOPEP3 .................. 285 Figure 5.2.38 Impact strength of bioplastics made with GFTSOPEPI ................... 285 Figure 5.3.39 Impact strength of bioplastics made with GFTSOPEP2 ................... 286 Figure 5.2.40 Impact Strength of bioplastics made with GFTSOPEP3 ................... 286 Figure 5.2.41 ESEM micrographs of impact fractures surface of bioplastic containing a) 10%GFTSOPEP1-UPE, Magnification 3000 X, Scale bar 15 um xviii b) 20% GFTSOPEPl-UPE, Magnification 300 X, Scale bar 150 um c) 30% GFTSOPEPl-UPE, Magnification 800 X, Scale bar 50 um (I) 40% GFTSOPEPl-UPE, Magnification 285 X, Scale bar 150 um ...................... 288 Figure 5.2.42 ESEM micrographs of impact fractures surface of bioplastic containing a) 10%GFTSOPEP3—UPE, Magnification 400 X, Scale bar 100 um b) 20% GFTSOPEP3-UPE, Magnification 400 X, Scale bar 100 um c) 30% GFTSOPEP3-UPE, Magnification 600 X, Scale bar 200 pm (1) 40% GFTSOPEP3-UPE, Magnification 300 X, Scale bar 150 um ...................... 289 Figure 5.2.42 AFM picture of 20% GFTSOPEP3-UPE (deflection image, 50 um X 50 pm) .................................................................................................. 290 Figure 5.2.44 AFM picture of 20% GFTSOPEP3-UPE (deflection image, 50 um X 50 um) .................................................................................................. 291 Figure 5.2.45 AFM picture of 20% GFTSOPEP3-UPE (Z scale: 1000 nm, scan size, 16 um X 16 um) ...................................................................................... 291 Figure 5.2.46 AFM picture of 20% GFTSOPEP3-UPE (deflection image, 16 um X 16 um) .................................................................................................. 292 Figure 5.2.47 Picture of 20% GFTSOPEP3-UPE on TV screen ........................... 292 Figure 6.1 Flexural properties of glass composites .......................................... 297 Figure 6.2 Storage modulus of glass composites at 40 0C .................................. 298 Figure 6.3 DMA plot for glass composites ................................................... 298 Figure 6.4 DSC plots of natural fibers used for SMC line .................................. 300 Figure 6.5 TGA plots of natural fibers used for SMC line ................................. 300 Figure 6.6 Tensile properties of biocomposites .............................................. 304 Figure 6.7 Flexural properties of biocomposites ............................................. 306 Figure 6.8 Impact properties of biocomposites ............................................... 308 Figure 6.9 Storage modulus of biocomposites at 40 OC ..................................... 309 Figure 6.10 Tensile properties of biocomposites ............................................. 312 Figure 6.11 Flexural properties of biocomposites ........................................... 314 Figure 6.12 Impact properties of biocomposites ............................................. 316 xix Figure 6.13 Storage modulus of biocomposites at 40 OC .................................... 317 Figure 6.14 ESEM micrographs of SMC biocomposites, at magnification of 100 X and scale bar of 450 um ............................................................................... 318 Figure7.1.l Moisture absorption of bioresins ................................................ 325 Figure7.1.2 Moisture absorption of biocomposites of hemp mat .......................... 326 Figure7.1.3 Moisture absorption of SMC manufactured biocomposites .................. 326 Figure 7.1.4 TGA of bioplastics ................................................................ 329 Figure 7.1.5 DSC of bioplastics ................................................................ 329 Figure 7.2.1 Change in color parameter ‘L’ over time for biocomposites ................ 330 Figure 7.2.2 Change in color parameter ‘a’ over time for biocomposites ................ 331 Figure 7.2.3 Change in color parameter ‘b’ over time for biocomposites ................ 331 Figure 7.2.4 Change in color parameter ‘E’ over time for biocomposites ................ 332 Figure 7.2.5 Change in weight over time for biocomposites ............................... 333 Figure 7.2.6 Change in roughness parameter Ra over time for biocomposites ........... 335 Figure 7.2.7 Change in roughness parameter Rz over time for biocomposites ........... 335 Figure 7.2.8 Change in roughness parameter Rmax over time for biocomposites .........336 Figure 7.2.9 Change in storage modulus over time for biocomposites .................... 337 Figure 7.2.10 Change in T8 over time for biocomposites ................................... 338 Images in this thesis are presented in color XX CHAPTER ONE: INTRODUCTION 1. INTRODUCTION Fiber-reinforced plastic composite materials were first produced by combining cellulose fibers with phenolic resin in 1908. Since that time, composite materials have advanced to applications ranging from packaging materials to aerospace components and structure, largely based on the utilization of petroleum based constituents, i.e., glass, carbon or ararnid fibers reinforced with epoxy, unsaturated polyester resins, polyurethanes, or phenolics. Because of growing environmental consciousness, traditional composite structures are now becoming the subject of legislative initiatives governing their manufacture, use and removal. The most difficult aspect of the problem is the removal at the end of lifetime, as the components are closely interconnected, relatively stable, and therefore difficult to separate and recycle [1]. There is a new and growing requirement for composite materials which are recyclable and/or degradable. Natural Fibers By combining natural reinforcing fibers with biopolymer matrices, new fiber reinforced materials called ‘biocomposites’ are being developed [2-10]. These biocomposites consist of biofiber as the reinforcing element and usually a biodegradable polymer as the matrix material. Since both components are biodegradable, the composite is also expected to be biodegradable. Biopolymers are generally biodegradable, but they do not possess the necessary thermal and mechanical properties useful for engineering applications. On the other hand, common engineering plastics are obtained fiom synthetic polymers, but are non-biodegradable. Although much research has been carried out on biofiber reinforced synthetic polymers, they are not completely biodegradable. However, biofibers derived from annually renewable resources, used as reinforcing fibers in both thermoplastic and thermoset matrix composites, provide environmental benefits with respect to ultimate disposability and raw material utilization [1, 5, 7]. In biocomposites, the biofibers serve as reinforcement by enhancing the strength and stiffness of the resulting composite structures. The conventional fibers like glass, carbon, aramid, etc., are produced with very specific properties having little variability, whereas the properties of natural fibers vary considerably. Biofiber properties depend on whether the fibers are taken from plant stem or leaf, the quality of the plants as a function of agricultural practices, the age of the plant, the preconditioning of fibers, and the processing methods adopted for extraction of fibers [5, 6]. Properties such as density, ultimate tensile strength, initial modulus, and toughness are related to the internal structure and chemical composition of fibers [5]. However, pristine natural fibers have the mechanical properties that would allow them to be used as glass fiber replacements in certain composite applications. In addition to constituent selection, to obtain a good reinforcing effect for plastics composites, it is necessary to increase the adhesion between fibers and resins by using surface treatments [3, 9]. Surface chemical modifications of natural fibers, including dewaxing, alkali treatment, cyanoethylation, vinyl grafting, and treatment with various coupling agents, have been shown to improve fiber matrix adhesion in biocomposites [9]. Thermoset composites are more desirable than therrnoplastics because of their superior mechanical properties. Petroleum based matrix resins in thermoset biocomposites are in general, non-biodegradable. Compared to thermoplastics, thermoset polymers have a lower viscosity, complex formulations, better damage tolerance, easier fiber impregnation, longer processing cycles, and higher fabrication costs. Thermoplastics are also recyclable, easy to handle, and have unlimited shelf life. However, they are prone to creep, while thermoset are not. Unsaturated polyester resin Among various thermoset resins, unsaturated polyesters (UPEs) are the most widely used matrix materials in polymeric composites [11, 12]. They are processed over a wide temperature range involving hand lay-up at low temperatures, resin-transfer molding at medium temperatures, and sheet-molding compound (SMC) compression molding, bulk- molding compound (BMC) injection molding, and pultrusion at high temperatures [19]. UPE resins have many applications in automotive, aircraft, electrical, and appliance components [1 1-20]. Unsaturated polyester resins (UPE) refer to a large class of polymers containing reactive double bonds in the polymeric chain. The reaction of a UPE resin is a free-radical chain- growth crosslinking copolyrnerization between the styrene monomer (25-45 wt %) and the UPE molecules. A source of free radicals, usually organic peroxides, is needed to initiate the reaction. The initiator is disassociated with help of a promoter. Optimum performance can be achieved by controlling the type and concentration of peroxide and promoter. Typical initiator concentrations run from 1 to 3% by weight based on clear resin and, for accelerators, (mostly cobalt) from 0.25 to 4% based on 1% metal content solution [1 1-14]. Under or overdosing can result in deterioration of cure and mechanical properties. Project Motivation The ambitious goals set by the US. government for the creation of a biobased economy as an alternative to the existing petroleum-based products are challenging the industry, academia, and agriculture [21]. This project seeks to replace conventional glass fiber- polyester composites by a novel low-cost natural/biofiber composite for housing panel applications. Through use of biofibers, our goal is to make composites, which can outperform current housing panels, while maintaining competitive economic structure. Objective The overall objective of this project is to develop environmentally friendly biocomposites as alternatives to the present glass based composites. Although polyester resin is a petroleum based product and not eco-friendly, we have chosen it as the matrix resin as it is required to produce a material which can be integrated easily into the current housing material industries. In the future, replacement or modification of the petroleum based resin by a suitable biobased resin is a logical and desirable direction for this research. In order to achieve the above objective, several interrelated components are required: i) development of a Biocomposite Sheet Molding Compound Panel (BCSMCP) high volume manufacturing process; ii) fundamental studies for determining the relationship between structures and properties of the materials;, and iii) utilization of ‘engineered’ natural fibers having good fiber-matrix adhesion, a balance of mechanical properties and processability. Design of Fibers ‘Engineered’ biofibers are defined as the suitable blends of surface treated bast, leaf and link fibers, which provide an optimum balance in mechanical properties of the final composite. The blending of differently modified fibers is based on the principle that the individual fiber properties vary with their location in the plant and the level of adhesion will determine their contribution to the composite stiffness and toughness. Previous research has shown that fibers originating fiom the leaf have very high toughness while those originating from bast have very high modulus [1]. Thus, by combining different fibers in optimum weight ratios, it is possible to have very tough and stiff biofibers. The bast fibers used in this research are kenaf, hemp, jute, flax, while the leaf fibers are sisal, henequen, and pineapple leaf fiber. An optimum fiber matrix interface bond is critical for performance of composite materials. Surface chemical modifications of biofibers have been successful in improving the fiber-matrix adhesion of biocomposites. Surface treatments also improve compatibility between fiber and matrix, dimensional stability and resistance to biodeterioration. The use of low cost, water based surface treatment and sizing for biofibers, for example, low concentrations of alkali treatment, and silane treatment is mandatory if the balance between performance and economic viability is to be maintained. The reinforcing effect of the fiber is also dependent on other parameters, such as, fiber diameter, fiber length, chemical constituents of the fiber, fiber orientation in the composite, number of individual fibers in a fiber bundle, and the fiber volume fraction used in the composite. All of these variables and their effect on the final biocomposite properties will be factored into the experimental plan. Project: Biocomposites from Engineered Natural Fibers for Housing Panel Applications Innovative biofiber Formulation of Novel Processing treatment & ‘Engineered polyester resin (Biocomposite sheet biofibers’ molding compound .___I— panel processing) Novel l-step water BCSMCP Optimization based sizing of all curing constituents Fabrication of l biocomposite l . Water ‘ panels usrng emulsion flame Curing BCSMCP process retardant studies I Evaluation of Surface characterization physico-mechanical of treated biofibers propertles and optimization of Blending of BCSMCP different biofibers 1. . Design of ‘Engineered’ biofiber Research output: Integration of education with R&D, possible future commercialization Figure 1.1: Schematic representation of the research plan for this project [22] Curing of polymer matrix Temperature and time are the most important parameters in a cure cycle. Curing of thermosets requires an intimate knowledge of the chemical kinetics of the polymerization and crosslinking reactions. The parameters that must be determined in a cure cycle are the number of stages in the cure, the rate of temperature increase, the hold temperature at each stage, the pressure at which the cure takes place, and the time allotted for the cure cycle [23]. Once the kinetics are understood and the actual chemistry behind the curing is established, the cure cycle parameters can be chosen based on the desired polymer properties. The physical nature of chemical crosslinking is quantified and represented by crosslink density and degree of cure. The crosslink density is a quantitative measure of the number of crosslinks that exist in a given volume in the therrnosetting polymer. The degree of cure represents the chemical conversion of the curing reaction [24]. The crosslinking density is determined from Dynamic Mechanical Analysis (DMA), and degree of cure is evaluated from Diffrential Scanning Calorimerty (DSC), and Fourier Transform Infra Red Spectroscopy (FTIR). In addition, the kinetics of curing reaction, including, rate of reaction, order of reaction, etc. can be estimated from DSC and FTIR experiments. It should be noted that the optimum curing conditions will have to re-established after the addition of any fillers, fibers, additives to the polymer matrix. Matrix modification Unsaturated polyester resins (UPE) are brittle, undergoing a 7-10% volume shrinkage after curing, and are not very resistant to alkali exposure. Traditionally, these properties are improved by blending with various additives; for example, fracture properties of a cured resin improve after blending with reactive liquid rubber and the shrinkage of UPE is prevented by introduction of a polar low-shrinkage thermoplastic. Bioresins, or thermoset resins derived from vegetable oils like, soybean, castor, corn, peanut, cottonseed, etc, also increase toughness and reduce volume shrinkage, with the added advantage of lower cost, abundant availability, and lower environmental impact [25]. Adding bioresins to the matrix will reduce the amount of fossil fuel based products in the final composite. It is planned to eventually phase out polyester resin from the formulation of biocomposites, to produce a completely ‘green’ product. Processing Using thermoset resins, biocomposites can be fabricated by compression molding, resin transfer molding and hand lay-up. One goal of this project is to develop a continuous process for making biocomposites, which will be similar to the existing Sheet Molding Compound (SMC) process, and thus could be adopted in the industries without any major change in infrastructure. This new high volume processing technique will be named ‘biocomposite stampable sheet molding compound panel’ (BCSMCP) manufacturing process. SMC is a continuous sheet containing chopped fibers and mineral fillers embedded in a highly viscous thermoset resin [26]. In the commercial SMC process, continuous glass fibers rovings are fed to a chopper, cut to 6 mm in length, and then distributed onto a carrier film, forming a uniform layer of chopped glass fibers. Since natural fibers cannot be obtained in a continuous from, and making a continuous yarn or roving with these fibers would be difficult and expensive, chopped natural fibers can only be utilized if a new process is developed. Achieving a uniform and continuous dispersion of a controlled amount of fibers with little variability on the SMC line will be an important goal of this work. Testing Mechanical properties of the composites need to be determined in order to insure that the resulting biocomposite has structural parity with the conventional composite. After the composite has been fabricated and conditioned, it is subjected to various tests, in accordance with ASTM standards to evaluate its properties. Properties of interest are tensile strength and modulus, bending strength and modulus of elasticity, impact strength, storage modulus, loss modulus, tangent delta, coefficient of thermal expansion, compressive strength, creep, shear stress and stiffness. The experimental properties are compared with theoretical values obtained from micro-mechanical models such as the Halpin-Tsai, Halpin and Pagano and Piggott models [27]. Morphology and fiber distribution are ascertained by electron microscopy on tensile fractures samples of composites. The durability of biocomposites is tested by moisture absorption test, accelerated weathering test, and fire test. The untreated and surface treated biofibers and hybrid biofibers are characterized by DSC, Thermal Gravimetric Analysis (TGA), Fourier Transform Infra-Red Spectroscopy (FT IR), Environmental Scanning Electron Microscopy (ESEM), and X-Ray Photoelectron Spectroscopy (XPS) to evaluate degree of crystallinity, maximum degradation temperature, chemical reactions as a result of treatment, topology, and morphology, and surface atomic concentrations, respectively. Summary The desire for renewable materials from sustainable sources is increasing for a variety of applications. Polymer matrix composites reinforced with natural plant fibers are one such example. Combining natural fibers with an unsaturated polyester resin matrix, novel low- cost biocomposites with desired properties can be made. Such biocomposites can provide many beneficial additions to the American Advanced Housing program. Fundamental and applied research into the materials, their surface treatments and fabrication processes is necessary to transfer this technology to industry. The objective of this research is to investigate the properties and processing of biofiber reinforced Unsaturated Polyester Resin (UPE) composites. The biggest advantage of panels made from this type of biocomposite is their low cost, combined with their ecological and technological advantages. If composite panels with acceptable properties can be developed, along with a viable manufacturing method for their continuous processing, housing panels for the future can be produced. 10 References l. Biodegradable Polymers in North America & Europe (P081), Mar Tech, USA, July (1998), 21 2. A. K. Mohanty, M. Misra, G. Hinrichsen, Macromolecular Materials and Engineering, 276/277, (2000), l 3. A. K. Bledzki, J. Gassan , Progress in Polymer Science, 24, (1999), 221-274 4. A. K. Mohanty, M. Misra, L. T. Drzal, Journal of Polymers and the Environment, 10 ( 1/2) (2002) 19 5. J. N. McGovern, in: Encyclopedia of Polymer Science and Engineering, Vol.7, H. F. Mark, N. M. Bikales, C. G. Overberger, G. Menges, J. I. Kroschwitz, Eds, John Wiley & Sons; New York, (1987), p.16. 6. L. M. Lewin, E. M. Pearce, Handbook of Fiber Science and Technology, Volume IV, Fiber Chemistry, Marcel Dekker, New York (1985) 7. AK. Rana, B.C. Mitra, A.N. Banerjee, Journal of Applied Polymer Science, 71 (1999), 531 i 8. B. van Voorn, H. H. G. Smit, R. J. Sinke, B. de Klerk, Composites: Part A, 32, (2001),]271-1279 9. A. K. Mohanty, M. Misra, L. T. Drzal, Composite Interfaces, 8, 5 (2001), 313-343 10. D. Ray, B. K. Sarkar, A. K. Rana, N. R. Bose, Composites: Part A, 32, (2001),1 19- 127 11. R. Bums, Polyester Molding Compound; Marcel Dekker: New York, (1982) 12. V. Bellenger, B. Mortaigne, J. Verdu, Journal of Applied Polymer Science, 44, (1992), 653 13. H.V Boeing, Unsaturated polyesters: structure and properties, Elsevier Pub. Co., (1964) 14. M. Malik, V. Chaudhary, I. K. Varma, Journal of Macromolecular Science, Part C - Reviews in Macromolecular Chemistry and Physics, C40, 2&3, (2000), 139-165 15. C. H. Bamford, C. F. H. Tipper, Free-radical polymerization, Elsevier Pub. Co., (1976) 16. Y.J. Huang, J .8. Leu, Polymer, 34, 2, (1993), 295-304 17. H. Yang, L. J. Lee, Journal of Applied Polymer Science, 84, (2002), 211—227 11 18. D. G. Hepworth, D.M. Bruce, J. F. V. Vincent, G. Jeronimidis, Journal of Material Science, 35, (2000), 293-298 19. R. Crawford, Journal of Plastics Engineering, 2nd ed.; Pergamon: New York, (1987) 20. Y. S.Yang, L. Lee, Journal of Polymer Process Engineering, 5, (1987—1988), 327 21. C. Eckert (Kline& Co.), “Opportunities of Natural Fibers in Plastic composites”, 3 rd Annual Ag Fiber Technology Showcase, October 4-6, 2000, Memphis, TN, USA 22. L. T. Drzal, A. K. Mohanty, “Biocomposites from Engineered Natural Fibers for Housing Panel Applications”, NSF-PATH proposal (2001) (Award number 0122108) 23. G. Giindiz, “Unsaturated polyester resin overview”, in Polymeric Materials Encyclopedia, 11, (J .C. Salamone Ed.) CRC Press, Boca Raton, FL, (1996), 8469-8476 24. H. Yang, L. J. Lee, Journal of Applied Polymer Science, 84, (2002), 211—227 25. G. Mehta, A. K. Mohanty, M. Misra, L. T. Drzal, Green Chemistry, 6(5), (2004), 254-258 26. US 3615979, C. J. Davis; R. P Wood; E. R. Miller, “Glass fiber-reinforced sheet molding compound” (1971) 27. J. C. Halpin, and J. L. Kardos, “The Halpin-Tsai equations: A review”, Journal of Polymer Engineering and Science, 52, (1976), 344-352 12 CHAPTER TWO: LITERATURE REVIEW 2.0 Literature Review In this chapter, some of the basic definitions and concepts related to natural fiber composites will be introduced, followed by summaries of past research related to natural fiber modification, polyester resin, bioresins, composite fabrication techniques, and applications of natural fiber composites. 2.1 Natural Fibers The idea of using cellulose fibers as reinforcement in composite materials is not a new or recent one. Ever since the beginning of human civilization, natural fibers like grass and straw were used to reinforce mud bricks. However, with the advent of high performance man made materials, the use of natural fibers diminished. Until recently, the natural fibers were mainly used in the production of rope, string, clothing, carpets and other decorative products. Interest in the use of natural fibers has grown during the last decade due to their various advantages. Increasing environmental consciousness and demands of legislative authorities are leading to the scrutiny of manufacture, use, and removal of traditional composite structures, usually made of glass, carbon or aramid fibers reinforced with epoxy, unsaturated polyester resins, polyurethanes, or phenolics. The disadvantages of such composite materials is the use of organic materials during their manufacture, high processing temperature and energy use, and disposal and recycling at the end of their 13 lifetime. The composite constituents are closely interconnected, relatively stable, and ©Drzal. Mohanty, Misra, MSU 2001 Figure 2.1.1: Examples of natural fibers [1] therefore difficult to separate and recycle [1]. Sustainability, industrial ecology, coo-efficiency, and green chemistry are guiding the development of the next generation of materials, products, and processes. Biodegradable plastics and biobased polymer products based on annually renewable agricultural and biomass feedstock can form the basis for a portfolio of sustainable, coo-efficient products that can compete and capture markets currently dominated by products based exclusively on petroleum feedstock [2]. To be sustainable, a biobased product derived from renewable resources has to be recyclable, triggered biodegradable, commercially viable and environmentally acceptable. The Technology Road Map for Plant/Crop based Renewable Resources 2020, sponsored by the US. Department of Energy (DOE), has targeted to achieve 10 % of basic chemical building blocks arising from plant derived renewable sources by 2020, with development concepts in place by then to achieve a firrther increase to 50 % by 2050. The US. agricultural, forestry, life sciences, and chemical communities have developed a strategic vision [3] for using crops, trees, and agricultural 2100 residues to manufacture industrial products, and have identified major barriers [4] to its implementation. 2.1.2 Structure of biofibers Biofibers are generally lignocellulosic consisting of helically wound cellulose microfibrils in an amorphous matrix of lignin and hemicellulose [1, 2, 5]. A single fiber of all plant based natural fibers consists of several cells. These‘cells are formed out of crystalline microfibrils based on cellulose, which are connected to a complete layer, by amorphouse lignin and hemicellulose. Multiple of such cellulose-lignin/hemicellulose layers stick together to a multilayer composite, the cell wall [9]. These cell walls differ in their composition and in the orientation of the cellulose microfibrils. These fibers consist of several fibrils that run along the length of the fiber. The potential fibers are separated from the original plant in several ways like retting, scrapping, decorticating and pulping. Cellulose is the main component of almost all natural fibers. The elementary unit of a cellulose macromolecule is anhydro-D-glucose, which contain three hydroxyl (OH) groups (Figure 2.4). These hydroxyl groups form hydrogen bonds inside the macromolecules itself (intramolecular) and between other cellulose molecules (intermolecular). Therefore all natural fibers are hydrophilic in nature. 15 H OH O OH OH H H CHZOH 0 g "' I] Figure 1.1.2: Chemical structure of cellulose molecules: Poly-[3(1,4)-D-Glucose 2.1.3 Advantages of natural fibers Natural fibers offer various advantages over other kinds of reinforcements. Biofibers come fi'om annually renewable resources, and are biodegradable. They are much less expensive and less dense as compared to man made fibers like E-glass fibers and carbon fibers [1]. Environmental gains can be made through use of renewable biofibers instead of synthetic fibers. Insulation and sound absorption properties of natural fibers are much better than those of fiberglass. 180 ‘r‘ 3 150 - i n cowl" r 2.5 A O Densrty (gm/cm3) ME r 2 g 39 a? is D Figure 2.1.3: Comparison of cost and density between glass fibers and biofibers Natural fibers provide a net energy savings over man-made fibers. Biofibers are produced by solar energy while production of synthetic fiber needs large amounts of petroleum based energy. The processing temperature to make glass fiber exceeds 1200°C [1, 7, 8]. It 16 takes 6,500 BTUs of energy to produce one pound of kenaf (not including the energy to produce fertilizer, collect and process the fibers) while it takes almost four times that much energy (23,500 BTUs) to produce one pound of glass fiber [7]. 2.1.4 Properties of Natural Fibers Physical properties of natural fibers are strongly influenced by their chemical structure such as cellulose content, degree of polymerization, orientation and crystallinity, which are affected by the plant genetic makeup, conditions during growth of plants as well as extraction methods used. As a result, there is an enormous variability in fiber properties depending upon which part of the plant the fibers came from, the quality of the plant and its location [1, 5]. The properties of these fibers are very difficult to measure, because a considerable number of fibers need to be tested to obtain statistically significant mean values. The mechanical properties of biofibers depend greatly on the scale and structure. Smaller structures usually lead to more regular composition and less defects and hence a better mechanical properties. However, with the existing technologies the smallest fiber obtained is in a form of single fiber or microfibril. Research is going on the extraction of nano scale cellulose whiskers from biofibers [10]. 2.1.5 Comparison between biofibers and other fibers The properties of biofibers are comparable to those of some artificial fibers (Table 2.1). Although biofibers have moduli comparable to glass fibers, they are lower in strengths and toughness. They also have a higher hydrophilicity because of the presence of OH 17 groups on their backbones. Thermal degradation in biofibers begins approximately around 200 0C, therefore they must be processed below this temperature. The specific tensile modulus (modulus of fiber divided by its density) and specific modulus of elasticity of biofibers are higher than those of glass fibers (Figure 2.5). This is because of the low densities of biofibers. Table 2.1.1: Typical properties of glass fiber and natural fibers [1 1] Fibers Properties E-Glass Flax Hemp Jute Ramie Sisal Density (g/cm3) 2.55 1.40 1.48 1.46 1.50 1.33 Tensile 800 _ Strength 2400 550 - 900 400 - 800 500 600 - 700 1500 (MPa) E'MOdulus 73 60 - 80 7o 10 - 30 44 38 (Gpa) Specific Modulus 29 26 - 46 47 7 - 21 29 29 (E/density) Elonga‘io“ at 3 1.2 - 1.6 1.6 1.8 2 24. Failure (%) Moisture Absorptiori%)" 7 8 12 12-17 11 El Elastic Modulus (GPa) 0 Specific Modulus (GPa/grn/cm3) 80 p—a LII O L O T O\ O l J; O Elastic Modulus (GPa) m S o o N o Specific Modulus Glass fiber Hemp th Figure 2.1.4: Comparison of moduli of glass fibers and biofibers 18 2.1.6 Types of natural fibers There is a wide variety of natural fibers. These include wood fibers, and a variety of agro- based fibers such as stems, stalks, bast, leaves and seed hairs. These fibers are abundantly available throughout the world and they come from renewable resources [1]. Fibers are also obtained fi'om recycled agro fiber-based products such as paper, waste wood, and point source agricultural residues such as rice hulls from a rice processing plant [2]. Depending on their origin, natural fibers may be grouped into: grasses and straw fibers, non wood fibers, wood fibers, and cellulose nano-whiskers. l . Straws, grasses and reeds These fibers come from the stems of monocotyledonous plants such as bamboo and sugar cane, big blue stem grass, switch grass, corn [1, 2, 5]. Most grasses produce biofibers that can be used to reinforce plastics. Higher cellulose content leads to higher mechanical properties in the biofibers, so does a lower microfibrillar angle. The mechanical properties of fibers decrease as the diameter increases. 2. Leaf fibers Leaf fibers are fibers that run lengthwise through the leaves of most monocotyledonous plants such as henequen, banana, pineapple, sisal, screw pine, and palm [1, 2, 5]. These fibers are also referred 'to as ‘hard fibers’. Leaf fibers generally have a higher microfibrillar angle, leading to lower values of tensile modulus compared to bast fibers. 3. Bast fibers These fibers (bundles) come from the inner bark (phloem or bast) of the stems of dicotyledonous plants. Common examples are jute, flax, hemp, mesta, and kenaf [1, 2, 3]. l9 Bast fibers generally have a lower microfibrillar angle, and higher cellulose content leading to higher values of tensile modulus compared to bast fibers. 4. Seed and fruit hairs These are fibers that come from seed-hairs and flosses, which are primarily represented by cotton, coir, kapok, and oil palm [1, 2, 5]. 5. Wood fibers These fibers come from the xylem of angiosperrn (hardwood) and gymnosperm (softwood) trees. Examples include maple, yellow poplar and spruce [6]. REINFORCING BIOFIBERS j Straw/ Grass Non-wood ‘ Cellulose Wood Biofibers Biofibers nano-whiskers Biofibers , I I , Examples: Bast I I - Sources: Wood Com/ wheatl - Leaf Seed/Fruit and non-wood rice straw/ big fibers through bluestem grass/ explosion switch grass, etc. Examples: Examples: Kenaf, jute, Cotton, coir, S etc oft and flax, hemp, etc. ' hard woods Examples: Sisal, henequen, pineapple leaf, etc. Figure 2.1.5: Classification of biofibers [2] 2.1.6 Biocomposites 20 Biocomposites, in general are materials made naturally or produced synthetically that include some type of natural material in their structure. In our research, biocomposites are also known as natural fiber composites. Biocomposites are formed through the combination of natural cellulose fibers with other resources such as biopolymers or resins or binders based on renewable raw materials or synthetic polymers. The objective is to combine two or more materials in such a way that a synergism between the components results in a new material that is much better than the individual components. Some of the plant fibers with suitable properties for making biocomposites are: industrial hemp, kenaf, henequen, jute, flax, sisal, banana, kapok, etc. [1]. The most commonly used polymer matrices include thermoset polymers such as polyesters, epoxies and phenolics, and therrnoplastics like polyethylene (PE), polystyrene (PS), and polypropylene (PP). Since the polymer matrix is soft, flexible and light weight in comparison to fibers, their combination provides a high strength-to-weight ratio for the resulting composite [1, 2, 11]. The properties of composites also depend on those of the individual components and on their interfacial compatibility. The interface between the fiber and the matrix gives the composite its structural integrity. The interface consists of the bond between fiber and matrix and the immediate region adjacent to this bond. At least three types of bonding are thought to exist at the interface: chemical, electrical, and mechanical. A composite with weak fiber matrix interface will not be able to transfer the load from the matrix to the reinforcing fiber and usually leads to poor composite strength [12]. The stress transfer at the interface between two different phases is determined by the degree of adhesion. Strong adhesion at the interface is needed for effective transfer of stress and load 21 distribution throughout the composite. In the case of short-fiber composites, various factors influencing the properties are: (i) the fiber dispersion, (ii) the orientation and geometry (aspect ratio) of the fibers within the composites, (iii) the fiber volume fraction, and (iv) the quality of the interface between the reinforcing fiber and polymeric matrix phase. 2.1.7 Applications of biocomposites The applications for which biocomposites have been studied include interior and exterior paneling of automobiles, interior paneling in rail vehicles, fumiture, and replacement of wood products in housing applications. Daimler Chrysler has used biocomposites of green flax and hemp fiber mats with polyester for under body panels of the EvoBus (a touring coach) [13]. They found replacing the under body panel, previously a glass fiber composite, with biocomposites required 83 percent less energy to manufacture, and the resulting part was 40 percent cheaper. In addition, the same tools and machine used to manufacture the part with glass-fiber composites were used for manufacturing with biocomposites. Biocomposite materials have also been used for interior paneling applications in automobiles, including door paneling and rear shelf paneling [14]. The interior paneling of rail vehicles in Germany has also been manufactured using biocomposites, mainly for weight savings over glass-fiber composites [15]. In this application the standards for fire protection are very high in Germany. Thus, the biocomposites were treated with flame retardants to attain the high levels of fire-protection required by the standards. 22 I IUP 30 M%GF EIUP 35wt°/oM= 100%«1 ' 80%" 38 kJ/mm 60%“ 40%” 20 kJ/mmz' 20%“ 0%“ Flexural Flexural inpact Density Strength Modulus Strength Figure 2.1.6: Comparison of natural fiber (NF) and glass fiber (GF)—unsaturated polyester (UPE) composites [13] Phenix Biocomposites of Mankato, Minnesota have developed biocomposite materials with decorative surfaces for furniture, table tops, wall panels, and other home and office finished surface applications [16]. These applications show that biocomposites can be manufactured and used in products where attractive surface finishing is required. Jute and coir based biocomposites have been developed in India as substitutes for plywood and medium density fiber boards for low-cost housing needs [17]. The engine and transmission covers of Mercedes-Benz transit buses now contain biocomposites of polyester resin and natural fibers [18]. Other products under development include the use of sisal—based biocomposites as panels and roofing sheets, which could also be used as false ceilings and partition boards. The current market uses of bio-based composites in North America are shown in Figure 2.7 [19]. 23 Others (7%) . Automotive (8%) Industrial! Consumer (10%) Building Products (75%) Figure 2.1.7: North American major end markets of natural fiber composites, year 2000 [19] 2.1.8 Modification of Natural Fiber A major disadvantage of biofibers is their highly polar nature, which makes them incompatible with non-polar polymers. This incompatibility usually leads to poor dispersion of the fibers in the matrix material, poor interfacial strength and lower mechanical performance. In addition, the poor resistance to moisture absorption makes the use of natural fibers less attractive for exterior applications or applications where they are exposed to a moist environment. In order to improve compatibility with polymer matrices and to minimize water absorption, the natural fibers need to be modified. This situation calls for the development of strategies for the modification of the cellulose fiber surface, thereby gaining control over the fiber-polyrner interface. Fiber surface can be modified by either physical or chemical treatments. Physical treatments include fibrillation, plasma treatment, and corona treatment [20]. Physical treatments change the chemical, structural and surface properties of the fiber surface and thereby influence the mechanical bonding with the matrix polymer. Chemical treatments introduce chemical bonds between the fiber and matrix, and include dewaxing, alkali treatment, cyanoethylation, vinyl grafting, and treatment with various coupling agents (silane, isocyanate, titanate, maleic anhydride etc.) [21]. Pretreatment of 24 fibers by encapsulated coating with coupling agents also provides better dispersion by reducing the fiber-fiber interaction with the formation of coating on the fiber surface. 2.1.9 Past Research on biofibers and biocomposites 2.1.9.1 Studies on biofibers Han [22] reviewed the characteristics of the non wood fibers, and suggested their use in pulping. The important properties of the fibers are fiber length, lignin content, and cellulose content. Rowell [23] reviewed the various applications possible for the agro- based fiber reinforced composites, ranging from geotextiles to filers to sorbents to structural and nonstructural composites, to packaging and molded products. Peijs [24] reviewed the developments in the field of biocomposites and their applications. From wood fiber composites, to natural fiber reinforced composites to all green biocomposites, biocomposites have come a long way in the recent years. These fibers have found their own special place in the automotive, sports and transportation sectors, using conventional processing techniques including sheet molding compound (SMC), bulk molding compound (BMC), laminating and resin infusion (RTM and VARTM), etc [11 ]. Riedel and Nickel [25] advocated the use of natural fibers reinforced with biopoylmers as construction materials in their 1999 article. They discovered a new application of biocomposites, in covering structural elements in automobiles, railway, furniture, and leisure industry. 2.1.9.2 Studies on surface treatments 25 Mishra et al. [26] surface treated pineapple and sisal fiber (alkali treatment, cyanoethylation, and acetylation) to make composites with polyester resin. They also made hybrid composites reinforced with a combination of glass fibers and sisal, as well as glass and pineapple fibers. The water absorption tendency of the biocomposites decreased after surface treatments, and also after glass fibers were introduced in the system. Mwaikarnbo and Ansell [27] studied the effect of mercerization and acetylation on the properties of hemp, sisal, jute and kapok by XRD, DSC, FT-IR, and SEM. After the chemical treatments, the surface of the fibers became rough and clean. From XRD they observed that at low alkali concentrations, there was a slight increase in the crystallinity of the fibers, but at higher alkali concentrations, the fiber crystallinity index decreases. Thermal analysis pointed at the optimization of acetyl groups at elevated temperatures. They concluded that the structure of the fibers was altered by alkalization and acetylation. Bisanda [28] studied the effect of alkali treatment on the adhesion characteristics of sisal fiber reinforced epoxy composites. He observed that the alkali treatment of sisal fiber improves the wetting ability of the fiber with the resin, reduces the voids, and improves strength and water resistance. Rout et al. [29] studied the graft copolymerization of acrylonitrile onto coir fibers using CuSo4 and NaIOa, as well as inorganic salts and organic solvents as the initiators. Morphological studies showed the evidence of grafting taking place on the fibers and even penetrating onto the fiber matrix. Grafting resulted in higher maximum stress at break, and increased hydrophobicity of the coir fibers. 26 Stamboulis et al. [30] studied the effect of environmental conditions on mechanical and physical properties of flax fibers. They treated the flax fibers with Duralin treatment, which consists of heating the fibers in steam or water at 160 0C for 30 min in an autoclave, followed by drying at 250 0C for 2 hours. It was found that Duralin treated fibers absorbed less moisture than untreated green flax fibers. Treated fibers were smooth and well separated, and have uniform higher strength. Zeta potential measurements confirmed higher hydrophilicity of green flax, and hydrophobicity of Duralin treated fibers. Silva et al. [31] performed mechanical and thermal characterization of alkali treated native Brazilian coir fiber. Mechanical and morphological results pointed at higher ultimate tensile strength and modulus of coir fibers after treatment. Thermal stability of fibers increased up to 48 hours of mercerization. Gassan and Bledzki [32] determined the relationship between the structure and mechanical properties of alkali treated jute fibers. Due to alkali treatment, there is shrinkage of the jute fibers which influences fiber structure, crystallinity ratio, degree of polymerization and Herrnans factor, as well as mechanical properties. Bismarck et al. [33] studied the effect of fiber surface treatments on the thermal and electrokinetic properties of coir and sisal fibers. The fibers were dewaxed, alkali treated and grafted with methylmethacrylate (MMA). It was seen from SEM micrographs that coir fibers were larger in diameter than sisal. After dewaxing and alkali treatment, the fibers became rougher, while, after MMA grafting they became very smooth. The maximum degradation temperature of the fibers increased after alkali treatment and MMA grafting. All treatments led to an increase in the water absorbed by the fibers. 27 Higher accessibility of the surface functional groups resulted in lower zeta potential, improving the interaction between the fibers and the matrix. Mohanty et al. [34] alkali treated and exploded the corn stalks with ammonia and carbon dioxide and studied their morphology and thermal properties of the fibers. As a result of these treatments, some fibrils got separated from the fiber bundle, and the thermal stability increased. Samal and Ray [35] studied FTIR spectra of chemically modified pineapple leaf fibers. The treatments done on the fibers were: alkali, dinitrophenylation, benzoylation, and benzoylation-acetylation. Pavithran et al. [36] evaluated the impact performance of unidirectional sisal-polyester composites. Owing to the optimum microfibrillar angle of sisal fibers, they exhibit a very impact amongst lignocellulosic materials. Energy absorption in the sisal biocomposites occurred by fiber failure and fiber pull cuts. The same authors also examined the impact performance of pineapple, banana, coir and sisal fibers in an unsaturated polyester resin matrix [37]. Oksman et al. [38] evaluated the mechanical and morphological properties of unidirectional sisal-epoxy composites made using resin transfer molding (RTM). The sisal fibers were non-uniformly distributed in the polymer matrix, and the adhesion between fibers and matrix was weak. Belcher et al. [39] studied the fiber matrix adhesion of the silane treated aligned henequen fiber reinforced epoxy composites. Coupling with y-glycidoxypropy ltrimethoxy silane (GPS) led to compatibilization between henequen fibers and epoxy matrix, thereby improving the thermal and mechanical properties of the biocomposites. 28 Belcher et al. [40] also studied the effects of epoxy modifications and surface modifications on the aligned henequen fiber reinforced epoxy composites. They introduced various amounts of epoxidized soybean oil and epoxidized linseed oil into the matrix, and also modified the fibers by using alkali treatment, plasma treatments, silane treatment and ultraviolet treatment. Varma et al. [41] made hybrid composites with glass fibers and jute fibers in a polyester resin matrix. The jute fibers were modified with y-aminopropyl trimethoxy silane, isopropyl triisostearoyl titanate, and tolylene diisocyanate (TDI). Although titanate treatment resulted in the properties of hybrid composites, Silane and TDI treated jute fibers were not compatible with the polyester resin. Mishra et al. [42] studied the tensile, flexural, impact and hardness properties of the unidirectional untreated and bleached jute-epoxy composites. Bleaching the fibers caused delignification which led to improvement in toughness, hardness, interlaminar shear strength and flexural properties of the composite. Rout et al. [43] studied the influence of surface modifications of coir fibers on the coir- polyester resin composites. The fibers were either alkali treated, bleached or grafted with acrylonitrile. The authors observed that adhesion between the coir fibers and the matrix improved after all surface treatments. The water absorption tendencies of the composites also decreased after the treatments. Zimmerman and Losure [44] reported the use of non-woven kenaf fiber mats as reinforcements for epoxy matrix. Although they found voids in the samples led to lower properties, they recommended pursuing this research. 29 Rong et al. [45] studied the interfacial interaction in sisal-epoxy composites and its effect on impact properties. The fibers were subjected to alkali treatment, acetylation, cyanoethylation, y-aminopropyl triethoxy silane, heat treatment, and mixed treatments. The modifications led to increased surface energy of sisal fibers after treatments, and chemical bonding between the fiber and the matrix. The microfailure mechanism of the composites was a function of interfacial adhesion and the fiber length continuity. For chopped fiber reinforced composites, the interfacial strength should be tailored to enhance the energy dissipation through debonding and pull out of the fiber bundles. Mishra et al. [46] analyzed the effectiveness of various types and degrees of surface modifications of sisal fibers in improving mechanical properties of sisal—polyester composites. 5% alkali treatment, 10% acrylonitrile grafting and cyanoethylation at 60 0C were the optimum modification conditions with respect to mechanical properties. Sydenstricker et al. [47] performed thermal, morphological and pull out analysis on siasl- polyester composites. The Brazilian sisal fibers were treated with alkali and N-isopropyl- acrylamide. The fibers treated with acrylamide solution performed better in tensile properties, moisture content and pull out lengths as compared to untreated and alkali treated fiber based composites. Hill and Khalil [48] studied the effect of treatments on mechanical properties of composites made with polyester resin and coir or oil palm as reinforcements. The fibers were treated with acetylation, y-methacrylopropyltrimethoxy silane, and nepenty1(diallyl)oxytri(dioctyl)pyrophosphate titanate. Khalil et al. [49] observed the effect of various anhydride treatments on the mechanical properties and water absorption tendency of oil palm empty fruit bunches in a polyester 30 matrix. The fibers were reacted with acetic, succinic, or propionic anhydride without any catalysts at 100 0C fbr 1 hour. Modifications resulted in hydrophobic fibers and improved fiber-matrix bonding. Acetylated fibers had better properties compared to propionylated or succinylated fibers. Mohanty et al. [21] reviewed the surface modifications of natural fibers and the properties of resulting biocomposites. They highlighted recent studies and developments in the area of improvement of the fiber-matrix adhesion for thermoplastic as well as thermoset matrices reinforced with natural fibers. Mohanty et al. [50] also reported surface treatments of natural fibers in a polyester resin matrix which improved the interface. The modifications included: alkali treatment, acrylonitrile grafting, methyl methacrylate grafting, and cyanoethylation. Ray et al. [51, 52] studied the effect of alkali treatments on the flexural and tensile properties of jute-vinyl ester composites. The improvements in mechanical properties were highest for a composite made with 4 hours alkali treated jute fibers with 35% fiber weight. Ray et al. [53] analyzed the dynamic mechanical properties of alkali treated jute-vinyl ester composites. For alkali treated fiber based composites, the rate of all of storage modulus with temperature was inversely proportional to the defect concentration in the composites. Gassan and Bledzki [54] evaluated the effect of cyclic moisture absorption-desorption on the mechanical properties of silane treated jute-epoxy composites. The absorption- desorption cycle led to debonding of the resin from the fibers, and introduced craks in the 31 matrix. It also changed the fracture mechanism of the composite, but did not alter the tensile strength. Depworth et al. [55] produced high volume fraction of hemp and flax fiber composites using low viscosity epoxy and phenolic resins. The fibers used were untreated, retted, mechanically decorticated or soaked in urea. Rout et al. [56] studied the effect of fiber surface treatments on the mechanical properties of coir-polyester composites. The surfaces of coir fibers were changed by use of dewaxing, alkali treatment, and graft copolymerization with methyl methacrylate onto alkali treated fibers. Joseph et al. [57] studied the effect of hybridization in jute/cotton fabric reinforced polyester resin composites. Giacomini et al. [58] manufactured composites containing curaua fiber and unsaturated polyester resin. Pothan et al. [59] studied the vicsoelastic properties of banana fiber reinforced polyester composites. Oksman et al. [60] analyzed the morphology and mechanical properties of sisal-epoxy composites made using resin transfer molding (RTM). 2.1.9.3 Studies on biofiber reinforced thermoplastics As mentioned before, one of the major drawbacks of using cellulose fibers as reinforcement is because of their poor dispersion characteristics in many thermoplastic melts, such as polypropylene and polystyrene, due to their hydrophilic nature. Several methods have been suggested and described in the literature to overcome this problem. Among them are fiber surface modification, use of dispersing agents such as stearic acid, and fiber pre-treatrnents such as acetylation. Fiber dispersion can also be improved with 32 increased shear force and mixing time [61]. A careful selection of initial fiber lengths, processing aids, processing techniques as well as processing conditions then is necessary in order to produce high performance composites. Raj and Kokta [62] investigated the influence of using various dispersing aids (stearic acid and mineral oil) and a coupling agent (maleated ethylene) in cellulose fiber reinforced polypropylene composites. Good fiber dispersion is generally the ultimate objective of any mixing process [61]. Various mixers are available to mix short fibers in thermoplastics such as extruders, plasticorder, injection molding machines and kneaders. Different mixing techniques, however, do not produce composites with the same degree of fiber dispersion. Therrnokinetic mixers have also been used to mix cellulose fibers with thermoplastics to effectively disperse the cellulose fibers within thermoplastic matrices [63, 64]. Pereira et al. [65] investigated the effect of several processing techniques on the properties of polypropylene composites reinforced with short sisal fibers. The best processing method involved a twin-screw extruder. Childress and Selke [66] investigated the effectiveness of several additives in enhancing mechanical properties of wood fiber/high-density polyethylene composites. The additives used were ionomer-modified polyethylene (ION), maleic anhydride modified polypropylene (MAPP), and two low molecular weight polypropylenes [67]. The mechanical properties of the composites studied increased with increasing additive concentration. The most effective additive was MAPP, followed with ION. Thomas et al. [68] reinforced polystyrene with benzoylated sisal fibers. The results revealed better compatibility between treated cellulose fibers and the polystyrene matrix, 33 and this resulted in enhanced tensile properties of the resulting composite. These improvements were attributed to the similarity between the phenyl-structure present in both benzoylated sisal fibers and polystyrene, which makes them thermodynamically compatible with each other. Ali et al. [69] studied the effects of processing conditions on the viscoelastic and mechanical properties of biodegradable composites made with extrusion of sisal fibers and mater Bi-Y and mater Bi-Z. Tensile and creep behavior of composites was affected by type of polymer matrix and processing conditions such as: speed of mixing, time of mixing and temperature. Alvarez et al. [70] evaluated the mechanical properties and water uptake of the compression molded composites made of alkali treated sisal fibers and MaterBi-Y, which is a biodegradable polymer. They found that the fiber treatment produced an increase in the equilibrium moisture content and a decrease in the diffusion coefficient. 2.2 Unsaturated polyester resin The majority of resins used in the composite industry are thermosets (see Figure 2.2.1) [71]. About 65 % of all composites produced currently for various applications, use glass fiber and polyester or vinyl ester resins. Unsaturated polyester resins (UPE) are widely used in the composite industry because of their relatively low price, low density, ease of handling, thermal and dimensional stability, good chemical and weather resistance, and excellent mechanical, chemical and electrical properties. Furthermore, compared to other thermosets, unsaturated polyester resins over can be pigmented, and can be easily filled and fiber reinforced in a liquid form. The use of reinforced thermoset composites by 34 automakers has nearly doubled in the last decade, and is expected to increase 47 percent during the next five years through 2004 [50, 72]. l 5% other thermosets 1 5% . 70% Reinforced . . Reinforced thermoplastics unsaturated polyester Figure 2.2.1: Matrix pattern of polymer composites [71] The polyester resins, because of their versatility and low cost, are widely used throughout the world. Polyester resins are classified as: (l) ortho resins, (2) isoresins, (3) bisphenol- A-fumarates, (4) chlorendics, and (5) vinyl ester [73]. Ortho resins, known as general- Table 2.2.1: Cost of various olyester resins Cost (Cents/lb) Polyester Resins Ortho resin 60-68 Iso-resin 73-83 Bisphenol-A 123-153 Vinyl ester 147-161 purpose polyester resin, are based on phthalic anhydride, maleic anhydride and glycols (Figure 2.2.2). Ortho-resin is the least expensive among all the polyester resins (Table 2.2.1). The solutions of unsaturated polyesters and styrene vinyl monomers (reactive diluent) are known as UPE resins. 2.2.1 Making UPEs: 35 The first unsaturated polyester resins of similar type as used today were synthesized in the 1930s. Polyesters are subdivided into three classes: aliphatic, aromatic, crosslinked. The third category is thermosetting polymers. Polyester resins are manufactured by a step growth polymerization reaction of unsaturated acids (or anhydrides), saturated aromatic acids, and difunctional alcohols glycols [73]. HO/ Figure 2.2.2: A representative chemical structure of unsaturated polyester The basic chemistry of linear unsaturated polyesters is rather simple. A mixture of unsaturated and saturated dicarboxylic acids is reacted with diols in a melt polycondensation. Monofunctional alcohols and acids are also used in some formulations to tailor the properties. The most traditional composition is maleic anhydride, o-phthalic anhydride and 1.2-propanediol, which are cheap raw materials. Other common raw materials are filmaric acid, isophthalic anhydride, terephthalic acid, adipic acid, ethylene glycol, diethylene glycol, dipropylene glycol, neopentyl glycol and bisphenol A [74]. The properties of the final product can be varied almost endlessly by changing the composition of the unsaturated polyester using these raw materials. Generally aromatic groups improve the hardness and the stiffness while aliphatic chain components increase the flexibility. The unsaturated polyester has typically a molecular mass between 1000 and 5000. The molecular mass is regulated by the diol/dicarboxylic acid ratio [75]. Usually the diol is in 36 excess, as the used diols are liquids, while the dicarboxylic acids and anhydrides are solids. An excess of solid reactants can cause a problem in the form of sublimation of the reactants during polycondensation. A high molecular mass will give a higher hardness, tensile and flexural strength of the final cured material. If the molecular mass is too low, the mechanical properties of the cured resin will be poor. A too high molecular mass increases the viscosity of the resin solution, which will cause problems with the processing of the resin. Air entrapment in the laminate, poor wetting of the reinforcement, long mould filling times and processing times are typical practical problems due to the resin viscosity. Almost all commercial production of unsaturated polyesters is done by the melt polycondensation of unsaturated and saturated acids or anhydrides with glycols. No solvents are used, and the formed water is continuously removed, in order to force the esterification reaction towards completion. The condensation temperature is typically between 170 and 230 °C. At the end of the condensation, vacuum is often applied in order to remove remaining water from the viscous melt. The total reaction time can be from 8 h up to 25 h, and the reaction is followed by acid number titrations and viscosity measurements [76]. Azeotropic polycondensation in the presence of organic solvents such as xylene or toluene can also be used. The reaction takes place at lower temperatures and it is possible to avoid losses of volatile reactants. The drawbacks are longer reaction times and environmental problems with solvent removing and recycling. The most common glycol used is propylene glycol. It is low cost and has good balance of properties. A combination of bisphenol A and propylene glycol provides good chemical resistance and high HDT to the oligomer. The most common anhydride used is 37 maleic anhydride, it provides cure sites. Fumaric acid is the best unsaturated acid used. Amongst saturated acids/anhydrides, phthalic anhydride is low cost and hard, isophthalic acid improves strength and chemical resistance, and adipic acid imparts flexibility and toughness. The UPE provides polymer properties, including modulus, toughness, glass transition temperature, and durability to the resin formulation [77]. O O .l O + O + CH3—‘CH2—CH2 -—-> OH OH O o Phthalic anhydride Maleic anhydride Propylene glycol it 1? it it -— C—O—(EH-CHz—O—C—CH=CH—C—O—CH—CH2—0—— — . _ + H20 _ n Unsaturated polyester resrn Figure 2.2.3: Preparation of polyester resin [74] 2.2.2 Diluent (Monomer) The polyester resins are usually diluted by adding a low molecular weight comonomer to adjust the viscosity of the mixture. Reactive diluent or monomer is added to the oligomer in the weight range of 0-60 %, typically 35-45 %. It helps to control the viscosity of the UPE, acts as a crosslinker, and improves wetting behavior [78]. Because of its lower cost, the most common monomer used for UPE is styrene, but other monomers can also be used. The examples of monomers used for UPE system cross linking are shown in table 2.2.2 38 Table 2.2.2: Examples of monomers used for UPE system cross linking Mainly used for translucent parts. Always used in combination Methylmethacrylate with styrene (50:50). Excellent weathering. Used for some special applications where a good control of the :égtzrenethyl reactivity and good control of shrinkage is needed. Unpleasant odor High reactivity. Provide high chemical resistance. Lower impact Divinyl benzene . res1stance. Low volatility. Used in molded parts with good electrical Diallyl phthalate p erformances. Used in special applications like stratification of PS foam. ““34 acetate Low hydrolysis resistance. 2.2.3 Curing of UPE The curing of polyester resin occurs by free-radical chain-growth crosslinking. In addition to the monomer and oligomer, curing agents need to be added for the reaction to proceed. Curing agents include initiator, promoter, accelerator and inhibitor. Initiator To initiate the reaction a source of free radicals is needed. Organic peroxides are used as the source of fi'ee radicals [74]. They refer to a family of molecules containing at least two oxygen atoms, single-bonded together. The general structure is: R—‘-———0—0—R2 R‘—o—o—R—o—3 o—R2 Rl—o—o-H The 0-0 bond can be easily broken to generate very reactive species called free radicals. The free radicals can be generated either by the action of heat (homolitic scission), or at room temperature in presence of an activator like metallic salts compounds or an amine. The first step of initiation is the decomposition of the peroxide. The second step is chain propagation reaction via styryl radical formation capable of reacting with a double bond. 39 CH3\ H CH3\ /C H3 2 H2 H2 OOH— (It—OOH OOH—C— O—O— T— OOH C H3 CH3 CH3 Monomer Dimer Figure 2.2.4: Chemical structure of methyl ethyl ketone peroxide (MEKP) The main families of organic peroxides used for curing unsaturated polyester are: ketone peroxides, diacylperoxides, peresters, hydroperoxides, and perketals. Ketone peroxides are generally used at room or mid temperature in combination with Cobalt salts. Diacylperoxides like Benzoyl peroxide can be used at room temperature, activated by tertiary aromatic amines, or without promoter at higher temperature (100°C). Peresters and perketals are used at higher temperatures and are preferred in injection and compression molding processes. Hydroperoxides are used at higher temperatures in injection and compression molding processes [79]. Ketone peroxides are the most commonly used for curing UPE systems. They have 8-10 % active oxygen. For room temperature curing as in the case of hand-lay-up structures, methyl ethyl ketone peroxide (MEKP) is used; for moderate temperature curing benzoyl peroxide is used. For hot press di-t-butyl peroxide or t-butyl perbenzoate is used. A mixture of initiators is used when a large temperature increase is expected. Q—l‘l-o—o—il Figure 2.2.5: Chemical structure of benzoyl peroxides (BPO) 40 The selection of the peroxide determines the kinetics of reaction and is also an important parameter for the "pot life". Final part quality is also linked to the peroxide used (aspect, curing efficiency). The selection of the right peroxide system involves several parameters including: part produced, associated process, processing temperature, timing of the process, nature of the pure resin used, and formulation. Optimum performance can be achieved by controlling the type and concentration of peroxide and promoter [80]. Promoters and accelerators Promoters and accelerators are used to help initiate cure at room temperature by accelerate the decomposition of peroxides. Commom accelerators include metallic salts, amines, or mercaptants. Cobalt naphthenate (CoNap) and cobalt octanoate (CoOc) are the most widely used accelerators. Cobalt naphthenate (CoNap) is usually added to systems containing MEKP, in weight range of 0-0.3%. Dimethyl aniline (DMA) is usually added in the weight range of 0- 0.3% to systems containing BPO and MEKP. Gel time for a given resin depends on initiator level, promoter level, second promoter level, and temperature. ROOH+Co3+——-> Roo‘ + H+ + Co2+ lroorr+Co2+ ———> my + on" + (303+ Figure 2.2.6: Reaction mechanism of polyester resin with cobalt activator Inhibitors In order to prevent premature curing and to extend pot-life, inhibitors are also introduced into the curing system, typically in concentrations less than 100 ppm. Inhibitors are free radical scavengers like Hydroquinone derivatives, terbutylcatechol, cresol derivatives. Some examples of inhibitors used for UPE system are shown in Table 2.2.3 41 Table 2.2.3: Examples of inhibitors used for UPE system Hydroquinone Excellent for stabilization during storage No color issues Can be used in a wide range of temperatures. Very efficient Tertlobutylhydroqumone with systems initiated by Dibenzoylperoxide Provides very good stability to the resin until 60°C. Used D'tertmbutylhyquumone when short gelling time is needed (cold and hot process) Can be used in a wide range of temperature. Very efficient Toluhydroquinone with systems initiated by Dibenzoylperoxide. Provides excellent thermal resistance Excellent performances. Provides yellow color to the Benzoquinone system . Active until 40°C. Rapidly absorbed by the resin. Tertrobutylcatechol Excellent adjustment of pot-life C rosslinking Polymerization may occur by any of three mechanisms: free radical, cationic, anionic. Chain polymerization is characterized by the presence of a few active sites which react and propagate through a sea of monomers. It may be formed via a two step process, where a polymer with unsaturations is first formed (a thermoplastic formed via step polymerization), and then these unsaturated sites are reacted with a crosslinking agent in a second step to produce the final structure. The reaction of a UPE resin is a free-radical chain-grth crosslinking copolymerization between the styrene monomer and the UPE molecules. The curing reaction involves a sequence of steps: initiation, inhibition and/or retardation, propagation, and termination. Polyester molecules are the crosslinkers while styrene serves as the agent to link adjacent polyester molecules. The polymer chains grow and become crosslinked in three possible reactive processes: styrene—polyester co- polymerization, styrene homopolyrnerization, and polyester homopolymerization [81]. 42 The curing behavior is complex due to the interaction between the resin chenristry and the variation of the physical properties. There are four stages in the crosslinking of UPE: Gelation: incipient formation of an infinite molecular network Vitrification: glass transition temperature of the forming polymer rises above the temperature of cure Full cure: highest attainable degree of cure Devitrification: degradation Before the flee—radical polymerization proceeds, the resin is a viscous liquid. During the curing reaction, the chain length of the resin molecules grows through the crosslinking reaction of functional groups and the resin becomes more viscous. As the curing advances further, the reacting system forms a highly crosslinked network, which results in a rapid increase in the resin viscosity [73]. Gelation corresponds to the formation of the first insoluble fraction of the polymer with an infinite molecular weight. The onset of gelation is of critical importance to processing thermoset systems. The general crosslinking mechanism is shown in Figure 2.2.7. _ Liquid Solid: 3D network ‘8 Styrene monomer : Double bond ~ Polyester chain Figure 2.2.7: Crosslinking mechanism of UPE resins 43 The reactivity, the viscosity and the final properties of UP resins can be adjusted by changing either the nature of the UPE chain, chemical composition (aliphatic, aromatic, number of double bonds), resin molecular weight, nature and concentration of the crosslinking monomer (containing at least one double bond or ratio of the styrene/polyester content. [76-81 ]. Autoacceleration Autoacceleration effect can occur when the increasing viscosity limits the rate of termination because of diffusional limitations. The free radicals can't terminate with as easily as they can find other monomers to propagate with. Rate decrease occurs because of further diffusional limitations caused by vitrification []. Effect of Additives on Cure Surfaces, sizings, and surface treatments on fibers and fillers can significantly alter not only the final properties of a composite, but also resin cure kinetics. Polyester resins are generally associated with fillers (silica, CaCO3, MgO, clay), and fibers (glass fibers, glass fibers mats or fabrics) acting mainly as a binder. 2. 2. 3.1 Curing performance measurement Lots of methods are used for determination of curing kinetics. As an example, the curing kinetic and performance in a cold cure system can be measured by the viscosity increase over time and/or the temperature increase over time. The crosslinking reaction between the unsaturated polyester and styrene is exothermic. The exothermic heat generated is proportional to the level of unsaturation of the chain and the amount of styrene. The heat 44 generated can auto accelerate the hardening reaction but some processes cannot sustain high temperatures. In molding processes, high temperatures can lead to excessive shrinkage, warpage and cracking for high thicknesses. Figure 2.2.8 below shows the principle of one possible measurement and the typical curves obtained. Propagation/ Autoacceleration [I Diffusion limited [I Reaction rate Induction time B Time Figure 2.2.8: Curing reaction rate of UPE 2.2.4 Studies on UPE Considerable work has been reported on the synthesis, characterization, curing behavior of UP resins [82-155]. 2.2.4.1 Studies on curing of UPE Kama] and Sourour [82] modeled of the DSC cure reaction with empirical equation dx/dt=kx"'(1-x)" by assuming the total reaction order m+n=2 and a zero initial cure rate [82]. 45 Salla and Martin [83] observed the dynamic, isothermal and residual heats of curing UPE at different scanning rates, temperatures and setting times by use of DSC. At room temperature, UPE curing follows an autocatalytic reaction system. It was difficult to find extent of reaction for such a system because at lower temperatures, the instrument was not sensitive enough, and at higher temperatures, some heat was lost even before the instrument could detect it. Residual heat can help to determine the ultimate extent of curing of any system. Salla et al. [84] used DSC to study the therrno-dependence and thermodynamics of the curing of UPE with different catalyst systems. For curing the UPE, they used BPO (0.03- 3.0 % wt), DMA, MEKP (1 wt "/o) and CoOc (0.2-10 wt %). In dynamic DSC experiments, more than one exothermic peak was seen. In the absence of a promoter, only one peak was observed. When the amount of initiator is varied, the exothermic peak gets shifted. Without post curing, all the formulations are only partially cured. Paci et al. [85] used pulsed nuclear magnetic resonance (NMR) at 20 MHz to monitor the middle to end stage of curing of UPE. The gelling and network formation process was followed by measuring spin lattice relaxation time of the system. A semi-empirical kinetic equation was used to evaluate degree of cure as a function of time, kinetic constants, and activation energy. Rarnis and Salla [86] investigated the theoretical and experimental curing reaction of UPE system with different polyester/styrene ratios by means of DSC and gel permeation chromatography (GPC). The theoretical heat of reaction corresponding to complete conversion was extrapolated from DSC curves. Experimentally, complete conversion was not seen in nay of the resin systems. From the molecular weight distribution obtained 46 from GPC, it was seen that part of styrene and polyester, as well as low molecular weight oligomer s of styrene and branch styrene on polyester molecules, do not form the polyester network. Rouison et al. [87] used an autocatalytic model to study the curing kinetics of a polyester resin containing various promoters and an inhibitor. They used an isothermal DSC method to obtain curing rate, rate constants, enthalpy of reaction, and order of equation. The experimental plots fitted well with the model and were temperature dependent. Nzihou et al. [88] utilized DSC for studying the polymerization kinetics of a thermoset resin under isothermal and dynamic conditions. They proposed a phenomenological kinetic model, which took into account diffusion effects, for curing reaction. They found that under isothermal conditions conversion was less than 100 % because of presence of unreacted monomer in the system. Dynamic studies in DSC revealed higher conversion at higher temperatures. Vilas et al. [89] studied the curing kinetics of UPE using DSC and thermal scanning rheometry (TSR). When the concentration of initiator (MEKP), or the cure temperature is increased, the gel time decreases, and the reaction rate increases. Huang and Leu [90] studied the effects of temperature, initiator and promoter on low temperature curing (30-50 0C) of UPE. In the early stages of reaction, the copolymerization is azeotropic. During diffusion controlled propagation stage, the conversion of styrene is greater than that of vinylene groups. The cross link lengths of styrene decrease with increasing temperature. At moderate reaction rates, the intermicrogel crosslinking reaction is more dominant than intramicrogel crosslinking 47 .a' “J .lrl ‘1' , u. reaction. At later stage of reaction, intramicrogel crosslinking is more favorable than interrnicrogel crosslinking. Yang and Lee [91] quantified the effects of resin type, temperature, curing agents, on the reaction kinetics of UPE and vinyl ester cured at low temperatures by DSC and rheometry. When MEKP is used as the initiator, inhibitor benzoquinone provides a longer induction period and higher final conversion for UPE compared to vinyl ester resin. Pentanedione is a good retarder for vinyl ester but acts as an accelerator for UPE. They also developed a gel time model to quantify the effects of temperature, inhibitor, retarder, on both polyester and vinyl ester resin. UPE is more sensitive to temperature changes than vinyl ester resin. Yang and Lee [92] studied the reaction kinetics of three UPE resins having different C=C bonds per molecule, by using a differential scanning calorimeter (DSC) Fourier transform infrared (FT IR), and electron spin resonance (ESR) spectroscopy. They also developed a diffusion-controlled model to simulate the reaction rates and conversion profiles of polyester vinylene and styrene vinyl groups, as well as the total reaction rate and conversion. The results from DSC and F TIR match well. The reaction rate increases with increase in the degree of C=C unsaturation per molecule. The diffusion-limitation effect is more significant for the polyester resin with a higher degree of unsaturation, leaving more unreacted C=C bonds trapped inside the matrix after vitrification, resulting in lower final conversions of polyester and styrene C=C bonds The glass transition temperature measured by the DSC is a linear function of the final conversion within the temperature range studied, and it is used to monitor the change of the final conversion. 48 Elli t’i ' ..4 cl... ,4 I ‘5) According to ESR measurements, for the polyester resin with a higher degree of unsaturation, the polymer formed is more compacted due to more intramolecular reaction, and more radicals are trapped in the cured resin without termination. The mechanistic kinetic model was capable of reasonably predicting the reaction rate and conversion profiles of polyester vinylene and styrene vinyl groups, as well as the total reaction rate and conversion at different temperatures and at nonisothennal conditions. Chen and Yu [93] studied the microgelation phenomenon of UPE curing by static and dynamic light scattering as well as DSC. The formation of microgels due to intramolecular crosslinking reaction inside the UPE coils, was observed by studying the variation between UPE particle size and the depolarization ratio. There is a decrease in the compatibility between partially cured UPE and styrene monomer as the intermolecular crosslinking reactions amongst the microgel particles proceeds further. Kosar and Gomzi [94] evaluated the thermal effects of curing reaction of UPE by isothermal and dynamic DSC experiments. They also developed a numerical model which took into account the heat transferred by conduction through the resin, as well as the kinetics of heat, generated by cure reaction. There was good agreement between the predicted model and the experimental data. Yun et al. [95] examined the curing kinetics of unsaturated polyester resin system exhibiting apparent induction periods by modeling free radical initiation and propagation processes. After making a master curve of the reaction exotherms, a single activation energy of the reacting system was determined for both induction time (ti) and the maximum reaction-rate time (tm). Using the power-law equation for the initiation efficiency function, two elementary model equations were proposed for the radical and 49 monomer reaction rates. The developed model equations predicted isothermal as well as dynamic heating curing reactions very well with only five model parameters without any modification in different thermal conditions. Ziaee and Palmese [96] evaluated the effects of cure temperature on proprieties of vinyl ester resin system using NMR, FTIR, AFM and tensile properties. Post curing the samples after isothermal curing showed the same conversion of styrene and vinyl ester bonds regardless of the initial cure temperature. The glass transition temperature was observed between 118-122 0C, irrespective of curing temperatures. The fracture toughness of samples cured at lower temperature (30 0C) was three times higher than those cured at higher temperature (90 0C). AF M revealed nodular morphology which could be microgels. Brill and Palrnese [97] used FTIR spectroscopy to investigate vinyl ester-styrene bulk copolymerization curing kinetics. The depletion of vinyl ester and styrene double bonds was monitored separately. Styrene monomer continues to react even after vinyl ester has completely reacted. Overall conversion increases with increase in isothermal cure temperature. Near IR and mid IR spectroscopic techniques were used for cure characterization of UPE by Grunden and Sung [98]. In addition to NIR analysis, they calculated the extent of reaction of styrene and vinylene C=C, in comparison to extent of reaction values from conventional mid—IR. The differences between the conversion of styrene and vinylene C=C is due to reduced diffusion of styrene into tightly crosslinked microgels. Lin and Hsu [99] studied the cure reactions of aromatic dicyanate, diepoxide, diamaine system using in situ FTIR and DSC. They proposed six reaction paths for dicyanate 50 /diepoxide system, four reaction paths for diepoxide /diamine system, and four reaction paths for dicyanate Idiepoxide /diarnine system. 2.2.4.2 Studies on properties of UPE Shan et al. [100] studied the effect of vinyl ester-styrene network structure on the thermal and mechanical properties by using DSC and DMA. The crosslink density of the vinyl ester resin was changes by two methods: by changing molecular weight of oligomer and by changing amount of styrene added to the curing system. Vinyl ester /styrene systems have secondary relaxations, which are indistinguishable from each other, though the physical network and the composition of network chains are different. Tg of the systems increases with decrease in cross link density. Han and Lem [101] studied the curing kinetics and the rheological properties of unsaturated polyester resins. They found that as cure progressed, the steady shear viscosity increased very rapidly with cure time at all shear rates investigated. The normal forces showed negative values at low shear rates and positive values at high shear rates. The negative normal forces may be due to material shrinkage during cure, and positive normal forces due to deformation of large molecules, formed by crosslinking reactions during cure. They explained the increase of viscosity with the degree of cure, at various values of isothermal curing temperature by combining the rheological and DSC measurements. Han and Lem [102] investigated the effect of particulates on both the rheological properties and the curing kinetics of unsaturated resin. In the formulation, calcium 51 carbonate and clay were used as inorganic particulates and high-density polyethylene powder as organic particulates. They authors found that as the particulate content increased, the resin formulation gave rise to shear-thinning behavior and the rate of cure increased. The observed shear-thinning behavior is attributable to breakage of the particle agglomerates as the shear rate is increased, behavior which is typical of concentrated suspensions. The CaC03 particles also helped to control shrinkage during cure when the material was subjected to steady shear deformation. The gel time was shorter for mixtures of resin and particulates than for the neat resin alone. A UPE system with particulates showed negative normal stress effects at rest, which originated from polymerization shrinkage. However, when the fluid is subjected to steady shearing deformation, the normal stress response was found to vary greatly among the various particulates investigated. Han and Lem [103] explored the effect of low-profile thermoplastic additives on rheology and curing kinetics of unsaturated polyester resin. The authors used UPE with two different types of LPAs: poly(vinyl acetate) (PVAc) and poly(methyl methacrylate) (PMMA). They observed that during cure the UPE/PMMA system exhibits shear thinning behavior even before the cure time reaches the critical value whereas the UPE/PVAc system does not. Both PVAc and PMMA helped to reduce the shrinkage of the resin during cure. The rate of cure and the final degree of cure are decreased when the amount of low-profile additive is increased. The shrinkage control becomes effective only when the shear rate is greater than a certain critical value. The shrinkage of the resin during cure, judged by the normal stress response is decreased by the addition of thermoplastic additive. The shrinkage control is found to be more effective when the 52 mixture of UPE/LPA is subjected to intensive shearing deformation than when the material is at rest. An optimum amount of low-profile additive must be used in order to minimize the sacrifice of the rate of cure and yet maximize the shrinkage control. Serre et al. [104] examined the morphology of the outermost layer of a compression molded samples made of UPE, glass fibers, calcium carbonate, low profile additives and release agents by using atomic force microscopy (AFM), dispersive x-ray spectrometry (EDX), scanning electron microscopy SEM, and XPS. The positions of fibers, filler, and mold release, were observed. Serre et al. [105] used atomic force microscopy (AF M) in the tapping mode to investigate the morphology of cured four blends based on miscible UPE/LPA (low profile additive)/ST (styrene) systems. AFM identified particles, nanoparticles and nanogels (with sizes varying from 20 to 57 nm), besides aggregates and microgels previously observed by scanning electron microscopy (SEM). All samples were miscible with each other, and indicated that miscibility was an important parameter in the morphology of the whole network. Higher miscibility resulted in more numerous nanogels, and smaller microgels. They also developed a relation between the microgels sizes and the void volumes. Marieta et al. [106] utilized AFM for studying the morphology of thermoset resins toughened by incorporation of core shell particles and thermoplastics. They used AFM in tapping and contact modes to look at the topographical, force and phase images of these systems. 2.2.4.3 Studies on flammability of UPE 53 Mazrouh [107] analyzed the fire retardant characteristics of polyvinyl chloride (PVC) and antimony trioxide (Sb203) in a UPE-fiberglass formulation. Addition of PVC (l-6%) and Sb203 (4-9%) enhanced the fire retardancy of the composites, with very little effect on the mechanical properties of the composites. Hangzhou JLS Flame Retardants Chemical Co. [108] developed a flame retardants for thermosets or for intumescent coatings. These retardants are based on ammonium polyphosphate, which is a non halogen, environmentally friendly product. When combined with the standard flame retardant filler, aluminum trihydrate (ATH), it lowers the levels of ATH needed to reach the required fire retardation regulations. M. Chaote [109] described some of the chemical approaches for halogen free flame retardants in circuit boards industry. He recommended the use of phosphate esters, melamine cyanurate, and phosphorous compounds as fire retardants. However, in these compounds, nitrogen and phosphorus have an intrinsic affinity towards moisture. They are also greatly susceptible to oxidative attack, and are readily thermally degraded. Another approach is to reduce the amount of resinous fuel, and increase the amount of hydrated filler content (aluminum trihydrate). Avtec Industries has developed a polymeric additive, TSWBTM, for fire retardancy and smoke suppressing, which is flee from bromine, antimony, and magnesium. It is designed and engineered to be incorporated within composite laminate structures (UPE based) to provide fire protection and thermal insulation properties and smoke suppression [110]. 2.2.4.4 Studies on toughening of UPE 54 Rosa and Felisberti [111] modified the UPE with poly(organosiloxanes) to improve its impact resistance. Glycidyl methacrylate, 1,3-aminopropyltriethoxy silane(ATPS), and l,l,3,3-tetramethyl-l,3—diehtoxysiloxane were used as modifiers for UPE resin. From dynamic mechanical analysis, it was observed that APTS was incorporated into the resin by reaction of its amino group with GMA. In systems with lower weight fraction of additives (7.23%), there was improvement in impact strength, however due to phase separation; there was no improvement in systems with higher amounts of additives (14.5 %). Rosa et al. also studied the adhesion between siloxane modified UPE and glass fiber by acoustic emissions of composites [112]. Rajulu, et al. [113] blended epoxy and unsaturated polyester resin in chloroform and studied the viscosity, ultrasonic velocity, and refractive indices of the blends. Gawdzik et al. [114] evaluated the effect of concentrations of toluene diisocyanate (TDI) on the thermo-mechanical properties of unsaturated polyester resins. 1-3 % by weight of TDI was added to the UPE, and the samples were subjected to FTIR, thermo-gravimetric analysis, Charpy impact test, and rheometry. TDI should not be added to UPE resins in amounts exceeding 1 wt %. At higher concentrations if TDI, there is a huge increase in the viscosity of the materials, and thixotropy is observed in samples containing 3 % TDI. Min et al. [115] used polyurethane (PU) was used as a modifier for UPE to improve the toughness of the resin. They studied the effect of the polyol molecular weight as a polyurethane (PU) soft segment and the PU contents on the toughness of PU-modified UPE resins. The PU was prepared by reaction between polymeric methyl diisocyanate 55 (MDI), a difunctional poly(ether polyol), a trifunctional poly(ether polyol) and dibutyltin dilaurate as the catalyst. The maximum toughness was observed in the case of 2 wt % of PU in the curing system, due to the network formation of UPE with PU. Above 2 wt %, the toughness value deteriorated because of the presence of unreacted polyol in the system. In the case of difunctional polyol, polyol cannot participate in the network reaction effectively as the molecular weight increases, so the toughness decreased because the chain mobility decreased. But in the case of trifunctional polyol, that touglmess was not affected considerably by the molecular weight. Cherian and Thacil [116] performed the toughening studies of UPE by introducing maleated elastomers using physical or chemical methods. The elastomers are either dissolved in styrene and then blended with UPE, or elastomers are modified by grafting with maleic anhydride. Cherian and Thacil [117] prepared blends of UPE with functional rubbers, and studied their mechanical properties. As the compatibility of the UPE with unmodified elastomers is poor, and a suitably distributed rubber phase of optimum size is not formed, they used elastomers bearing ftmctional groups. The elastomers used were epoxidized natural rubber, hydroxyl terminated natural rubber, hydroxyl terminated polybutadiene, and maleated nitrile rubber. Toughness and tensile properties were maximum at 2.5 wt % of rubber in the system (studied range: 0-5 wt % rubber). Maleated nitrile rubber had the best performance compared to all other elastomers. Thouless et al. [118] discussed the mechanics of toughening of brittle polymers. There are tow principles for toughening of brittle polymers. The first is by manipulating the cohesive processes acting across a crack tip, to increase the intrinsic toughness of the 56 materials. The second is by manipulation of the microstructure of the polymer, to make it easier to trigger non-linear dissipative processes around the crack tip. Huang et al. [119] reviewed the developments in toughening mechanisms of thermoset polymers. A toughened thermoset usually contains elastic or thermoplastic domains dispersed in discrete for throughout the matrix resin to increase the resistance to crack growth initiation. The phenomenon of viscoelasticity, shear yielding, and dilatational deformation involving cavitations govern the fracture behavior of thermosets. McGarry and Subramaniam [120] prepared a toughened cross linked molecular network system with unsaturated polyester, styrene, diglycidyl ether of bisphenol A, and a reactive liquid runner, namely, amino terminated butadiene-acrylonitrile. The network is a two phase system,iwhich has one phase containing no rubber, and the second phase with uniform distribution of rubber. 2.2.4.5 Studies on modification of UPE Synthetic modifications of unsaturated polyester resins have been done in order to achieve lower styrene emission, better styrene solubility and lower processing viscosity. Two concepts were used; introduction of liquid crystalline segments into the unsaturated polyester and end-capping the unsaturated polyester with poly(ethylene glycol)s of various molar mass[121] . Hegemann [122] developed a comonomer free low viscosity unsaturated polyester resin for electrical industry. Two C=C bonds were found in a system by combining standard UPE resin (has a C=C from maleic/firmaric unsaturation), and the adduct of Cyclopentadiene with a functional group, which allows for a condensation reaction. These 57 unsaturations were incorporated in one resin, without copolymerization during the polycondensation of the polyester. Crosslinking was studied by using DSC. Douglas and Pritchard [123] studied the effect of substituents in the styrene ring, on the water absorption characteristic and thermal stability of UPE. They used styrene, 4-mehtyl styrene, 4-ehtyl styrene, 4-n-butyl styrene, 4-isoproyl styrene, tertiary butyl styrene, 4- chlorostyrene, and 3, 4,-dichlorostyrene to crosslink a maleic/phthalic anhydride based unsaturated polyester. Loss of water resistance was seen in systems with chloro substituents and most systems with alkyl groups. System with tertiary butyl group leads to a decrease in water absorption. Thermal stability is decreased by chloro substituents and increased by n-alkyl groups. Guo and Zheng [124] studied the miscibility and crystallization of thermosetting polymer blends of unsaturated polyester resin and poly(s-caprolactone) (PCL). Before curing, only one glass transition temperature (Tg) is present in (PCL) UPE blends, showing miscibility. But after crosslinking, the blend was partially miscible. FTIR studies revealed that intermolecular hydrogen bonding interaction between the components is an important driving force to the miscibility of PCL/UPE blends and the partial miscibility of crosslinked PCL/UPE blends. The spherulitic morphology of the blends was remarkably affected by crosslinking. Guo and Zheng [125] also studied the crystallization kinetics of thermosetting polymer blends of poly(e-caprolactone) and unsaturated polyester resin. The overall crystallization rate of PCL decreases with the addition of amorphous component, UPE. The kinetic rate constant decreases rapidly for both the PCL/UPE blends and the crosslinked PCL/UPE 58 blends with decreasing PCL concentration. However, neither of these papers presented a study on effect of PCL addition on volume shrinkage of the unsaturated polyester resin. 2.2.4.6 Studies on low profile additives for UPE In low temperature molding processes, control of resin shrinkage and residual monomer is an important concern. The presence of low profile additives (LPAs) can reduce the shrinkage of unsaturated polyester (UPE)/styrene resins under proper processing conditions but may increase the residual styrene content. It is believed that the reaction- induced phase separation and the polymerization shrinkage in both the LPA-rich and UP- rich phases result in the formation of microvoids, which partially compensates the resin shrinkage. The relative reaction rate in the two phases plays an important role in shrinkage control. Han and Lem [126] measured the bulk rheological properties of polyester resin formulation containing UPE/fillers and UPE/fillers/LPAs using cone-and-plate rheometer. The fillers used were clay, calcium carbonate, and milled glass fiber. The low-profile additives, poly(vinyl acetate) (PVAc) and poly(methyl methacrylate) (PMMA), were dissolved in styrene. The bulk viscosities of all blends of polyester resin and PVAc lie between those of the individual components, whereas the bulk viscosities of some blends of polyester resin and PMMA go through a minimum and a maximum, depending on the composition of the mixture. On adding both the filler and low-profile additive together in polyester resin, the rheological behavior became quite complex, indicating presence of interactions between the filler and the low-profile additive. 59 Han and Lem [127] investigation was made of the rheological behavior of unsaturated polyester resin during thickening in the presence of filler or low-profile additive alone and, also, in the presence of both filler and low-profile additive. They used CaCO3 and clay as fillers and poly(methyl methacrylate) (PMMA) and poly(vinyl acetate) (PVAc) as the low-profile additives. The viscosity behavior of the resin/filler/LPA system was similar to that of the resin/ LPA. From the rheological standpoint, the different types of low-profile additive had little influence on the thickening behavior of the respective filled systems. Ruffler et al. [128] characterized the fissures occurring during curing of blends of UPE with PVAc using fractal analysis. The LPA effect arises by fissuring with fractal geometry. The fractal dimensions of fissures depend on the PVAc concentration and the cure temperature. Huang and Su [129] studied the effects of two LPAs, PVAc and PMMA on the curing kinetics during the cure of UPE resins by use of DSC and FTIR. There is a kinetically controlled plateau (small shoulder) in the initial portion of the DSC rate profile depending on the molar ratio of styrene to polyester C=C bonds, in the sample containing LPA. Adding LPA enhanced the relative conversion of polyester C=C bonds. to styrene throughout the reaction, as shown by F TIR results. A microgel-based kinetic model was developed to explain the effects of LPAs on reaction kinetics (intramicrogel and intermicrogel crosslinking reactions), relative conversion of styrene and polyester C=C bonds, and the final conversion. The LPAs induced a globule microstructure during the curing, which enhanced the final conversion of polyester, but would result in an increase 60 or a decrease of the final conversion of styrene, depending on the initial molar ratio of styrene to polyester C=C bonds. Huang and Su [130] investigated the effects of two LPAs, PVAc and PMMA, on the morphological changes during the cure of UPE resins by using DSC, SEM, optical microsc0py, and low-angle laser light-scattering (LALLS). Under SEM, the microvoids and microcracks which are responsible for the volume shrinkage control were observed at the later stage of reaction. The morphological changes during curing varied considerably, depending on the types of LPA and the initial molar ratios of styrene to polyester C=C bonds. Adding LPA in the neat UP resin reaction system enhanced the clear identification of microgel particles, because a layer of LPA could cover the surface of microgel particles when the microgel particle phase separates from the matrix of unreacted resin containing LPA. The segregating effect of LPA led to reduced merging of microgel particles and helped to retain the identity of the individual microgel particle. Sun and Yu [131] used gel permeation chromatography (GPC) and DSC to study the effect of LPAs, PVAc and PMMA, on the curing reaction of UPE resins. According to DSC results the curing reaction rate decreased as the concentration of LPA increased. The concentration of LPA and the compatibility of LPA with UPE resins had a strong influence on the polyester microgel formation and the curing behavior. There was evidence of the shrinkage of UPE microgels at the early stage of curing reaction witnessed from GPC results. The formation of microgel particles has been found to be the key feature of the UPE-styrene copolymerization. For UPE resins mixed with LPA, the curing reaction rate of UPE with styrene decreased as the concentration of LPA increased. Increase of the LPA concentration facilitates the UPEs to form coils in the styrene 61 monomer and to undergo an intramolecular crosslinking reaction, which causes a delay of gelation. Huang and Liang [132] compared the volume shrinkage abatement caused by four LPAs, including PVAc, PMMA, thermoplastic polyurethane (PU), and polystyrene (PS), in polyester resin. The effectiveness of volume shrinkage agent decreased in the following manner: PVAc > PMMA > PS. The shrinkage effect with PU depended on its concentration. In case of relatively polar LPAs like PVAc and PMMA, the fractional volume change after curing is controlled by microvoid formation. For nonpolar LPAs like PS, it is controlled by intrinsic polymerization shrinkage. For polar LPAs which can react with UPE, both factors play a role. Siato et al. [133] studied the rheological and morphological changes in polyester resin after incorporation of methyl diphenyl isocyanate (MDI) and magnesium oxide (MgO) as thickeners into a UPE-LPA system. The MDI based system exhibited rapid viscosity build up and followed by stable viscosity for long storage times, and the shrinkage control by this system of poor due to lack of phase separation. For MgO based systems, the viscosity increased slowly and it changed during the long storage time. There was phase separation during curing which led to a better shrinkage control in MgO based system. Li and Lee [134] found out the shrinkage mechanism of UPE/LPA systems at low temperature cure using rheometry, kinetics, morphology and dilatometry study. The competition of the shrinkage induced by resin polymerization and the expansion induced by microvoid formation, strongly determine the shrinkage behavior of the UPE resin. At 62 low temperature cure, the systems with higher molecular weight and lower LPA content are more effective shrinkage controllers. Huang and Jiang [135] explored the effects of chemical structure and composition of UPE on the miscibility of UPE/styrene/LPA system having PU, PVAc, PMMA as LPAs. The dipole moments of UPE and LPAs were calculated, and it was found that the increasing order of polarity in terms of dipole moment per unit volrune was PMMA3$ con 03 _ oofl cog comm comm 00 H m 00m m comm : _ _ L _ _ _ _ _ 0 lllll 1» 11,1; 1 om l m m . ow m. m u .. 8 a % . ow oofi 243 0 (It 1 —t=0min —t=250rnin r“ 25~ \ "I T h Ur M M 1 n Transmittance (%) DJ LII \1 15 r 5 j I I T I I I\ 1050 1000 950 900 850 800 750 700 650 Wave nunber (cm!) Figure 5.1.1]: FTIR spectra of UPE system at 150 0C at t=0 minutes and t= 250 minutes. Figure 5.1.11 shows the FTIR specha of UPE system at 150 0C at t=0 minutes and t= 250 minutes. At t=0 minutes, when the reaction has not yet started, the peaks of UPE system at 980 and 910 cm'1 look very distinct. However, after t=250 minutes, when the reaction has been completed, the peaks at 980 and 910 cm'1 are no longer visible. The FTIR specha of UPE system at other temperatures, at the beginning and end of the reaction look similar to Figure 5.1.11, and thus have not been shown. Figures 5.1.12 through 5.1.23 show the fractional conversion of C=C bonds of UPE as well as styrene at 50 °C, 60 °C, 70 °C, 80 °C, 90 °C, 100 °C, 110 °C, 120 °C, 130 °C, 140 0C, 150 0C and 160 0C respectively. The peak heights were normalized and calculated by using the PerkinElmer Spechum 2000 software. The values of fractional conversion were calculated using equations 7 and 8 on peak heights of all FTIR specha taken for four hours. 244 Fractional Conversions at 50 0C 0-8 ‘ — UPE —— Styrene ,5 0.6 ~ (0 8 > :1 5 0.4 ~ 0.2 - O I T I I 0 50 100 150 200 250 Time (min) Figure 5.1.12: Fractional conversion at 50 0C from in situ FTIR study Fractional Conversions at 60 0C I v/ -— UPE -— Styrene E .2 8 > :1 O U 0 I l I I 0 50 100 150 200 250 Time (min) Figure 5.1.13: Fractional conversion at 60 0C fiom in situ FTIR study In Figure 5.1.12, fractional conversions of double bonds of UPE and styrene at different times are shown for an isothermal experiment at 50 0C. The conversion curves were very 245 smooth, and the initial slopes of both curves were very gradual. At the beginning of the reaction, the conversion of UPE was higher than that of styrene, but after 50 minutes, the conversion of double bonds of styrene was more than those of UPE. In Figure 5.1.13, fractional conversions of double bonds of UPE and styrene at different times are shown for an isothermal experiment at 60 0C. Again the curves were smooth, and the initial slope was gradual. As seen in figure 5.1.12, initially the conversion of UPE double bonds were higher, however after 50 minutes, the conversion of double bonds of styrene was more than those of UPE. l 0.8 — —-UPE —-Styrene t: 30.6 - 0 E 70 C E °0.4 i U 0.2 r O a fi I I I 0 50 100 150 200 250 300 Time (nrinutes) Figure 5.1.14: Fractional conversion at 70 0C from in situ FTIR study Figure 5.1.14 shows fractional conversions of double bonds of UPE and styrene at 70 0C. It can be seen that the initial slope became less gradual and more steep, compared to Figures 5.1.12 and 5.1.13. Also, the initial conversion of UPE double bonds was higher, however after about 20 minutes; the conversion of double bonds of styrene was more than those of UPE. 246 0-8 ‘ l —UPE -—Styrene 0 g 0.6 — 80 C 5 g 04 - O o 0.2 - 0 I I fir I I —5 45 95 145 195 245 Time (nu'nutes) Figure 5.1.15: Fractional conversion at 80 0C from in situ FTIR study 1 08 _, "— UPE — Styrene 0 90 C .5 0.6 1 g 0.4 e U 0.2 J O I I I I I 0 50 100 150 200 250 300 Time (nrinutes) Figure 5.1.16: Fractional conversion at 90 0C horn in situ FTIR study Figure 5.1.15 shows fractional conversions of double bonds of UPE and styrene at 80 OC. The initial slope of both curves became very steep, compared to Figures 5.1.12, 5.1.13 and 5.1.14. The initial conversion of UPE and styrene double bonds was almost the same. The maximum conversion of styrene was higher than that of UPE, as already discussed by many authors [15-18]. 247 The fractional conversion of UPE and styrene double bonds at 90 0C is shown in Figure 5.1.16. The initial slope of both curves became very steep, compared to Figures 5.1.12, 5.1.13 and 5.1.14. The initial conversion of UPE and styrene double bonds was almost the same. 9 oo 1 I — UPE —— Styrene 100 °C 9 ON 1 Conversion .9 h 1 .C’ N 1 O 1 1 1 r 1 0 50 100 150 200 250 300 Time (nrinutes) Figure 5.1.17: Fractional conversion at 100 0C from in situ FTIR study 1.2 Fractional conversions at 110 0C — UPE — Styrene 1 T / 0.8 - t: .2 t": 0.6 — > t: O U 0.4 — 0.2 ~ 0 I I I I I I -10 40 90 140 190 240 290 340 Time (min) Figure 5.1.18: Fractional conversion at 110 0C from in situ FTIR study 248 Figure 5.1.17 shows fractional conversions of double bonds of UPE and styrene at 100 0C. The initial slope of both curves was very steep, similar to Figures 5.1.15, and 5.1.16. The initial conversion of UPE and styrene double bonds was almost the same. The maximum conversion of styrene was higher than that of UPE. The fractional conversion of UPE and styrene double bonds at 110 0C is shown in Figure 5.1.18. The initial slope of both curves, again, became very steep. The initial conversion of UPE and styrene double bonds was almost the same. 1 - Fractional conversions at 120 0C —UPE — Styrene 0.8 — 8 g 0.6 e > t: O U 0.4 - 0.2 4 O 1 1 1 1 1 1 -10 4O 90 140 190 240 290 Time (nrin) Figure 5.1.19: Fractional conversion at 120 0C from in situ FTIR study Figure 5.1.19 shows fractional conversions of double bonds of UPE and styrene at 120 0C. The initial slope of both curves was very steep, similar to Figures 5.1.16, 5.1.17, and 5.1.18. The initial conversion of UPE and styrene double bonds was almost the same. The maximum conversion of styrene was higher than that of UPE. 249 1.2 Fractional Conversions at 130 0C — UPE — Styrene l; ,- p—l l Conversion .6 .o .o A as oo 1 L l .o N 1 0 I I I I I - 5 45 95 45 195 245 Time (rrrin) 1 Figure 5.1.20: Fractional conversion at 130 0C from in situ FTIR study 1.2 Fractional conversions at 140 0C -— UPE — Styrene W 1_. .0 00 1 Conversion o ex 1 0.4 ~ 0.2 - O I I T I l I - 10 40 90 140 190 240 290 Time (nrin) Figure 5.1.21: Fractional conversion at 140 0C from in situ FT IR study In Figure 5.1.20, fractional conversions of double bonds of UPE and styrene at 130 0C are shown. The conversion of double bonds of styrene was more than those of UPE. At this 250 temperature, the conversion rates were much higher than at 50 or 60 OC. The curves reached their maxima very quickly, in a matter of minutes. Figure 5.1.21 shows fractional conversions of double bonds of UPE and styrene at 140 0C. The initial slope of both curves was, again, very steep. The initial conversion of UPE and styrene double bonds was almost the same. The maximum conversion of styrene was higher than that of UPE. 1.2 Fractional conversions at 150 0C — UPE — Styrene 1 I 08 A K7 = .2 g g 0.6 — Q U 0.4 i 0.2 r O I I I I r -10 40 90 140 190 240 Time (min) Figure 5.1.22: Fractional conversion at 150 0C from in situ FTIR study 251 —s N Fractional Conversion of UPE at 160 oC __ UPE —— Styrene 1 -—4 __._/ 0.8 J :2 .2 a“: > O 6 ‘ = G U 0.4 ~ 0.2 ~ 0 I I I I -5 45 95 145 195 245 Time (nin) Figure 5.1.23: Fractional conversion at 160 0C from in situ FTIR study In Figure 5.1.22, fractional conversions of double bonds of UPE and styrene at different times are shown for an isothermal experiment at 150 0C. The initial slopes of the curves are again, very sharp. The conversion of double bonds of styrene was more than those of UPE. This temperature was 90 0C higher than the one shown in Figure 5.1.12; therefore the conversion rates are much higher. The curve reaches its maxima very quickly. In Figure 5.1.23, fractional conversions of double bonds of UPE and styrene at different temperatures are shown for an isothermal experiment at 160 0C. The conversion of double bonds of styrene was more than those of UPE. Here again, the temperature was 110 0C higher than the one shown in Figure 5.1.12; therefore the conversion rates were much higher. The maximum of the curves was reached in a few minutes. Table 5.1.1: Comparison of maximum conversion at different temperatures from FTIR experiments 252 Temperature Time a max 0. max (C) (minL UPE STY 50 220 0.87 0.89 60 220 0.88 0.90 70 26] 0.89 0.91 80 240 0.90 0.92 90 253 0.90 0.93 100 271 0.90 0.93 110 280 0.91 0.94 120 300 0.91 0.94 130 240 0.91 0.95 140 285 0.91 0.97 150 240 0.92 0.97 160 242 0.92 0.99 In Table 5.1.1, a comparison is made of the maximum fractional conversions of double bonds of UPE and styrene at different temperatures measured at the end of experiment. As expected, conversions are directly depending on temperature, and they increase with increment in temperature. The conversion of double bonds of styrene is more than those of UPE. This has been already reported in the literature [1-19]. 5.2 Bioresins and bioplastics The results of matrix modification experiments by introducing a derivatized vegetable oil in the UPE curing system are described in this section. The first part deals with the results from bioplastic obtained by grafting of acrylonitrile on vegetable oil like soybean, castor, and MSO. The second part discusses the results of bioplastics obtained after grafiing soybean polyol with maleic anhydride. 5. 2. I Bioplastics made by grafting of vegetable oils with acrylonitrile 253 The bioresins formed were characterized by FTIR, TGA, DSC, and the bioplastics developed from the bioresins were analyzed by mechanical, thermal and morphological tests. FTIR The bioresin obtained after reacting acrylonitrile with natural oils, were characterized by FTIR to validate the proof of reaction. FTIR scans of pure acrylonitrile, pure vegetable oils, and, modified vegetable oils were taken and compared. Figure 5.2.1 through 5.1.7 show the above mentioned FTIR spectra of acrylonitrile, pure natural oils, as well as grafted oils. In Figure 5.2.1, the spectrum of pure acrylonitrile is seen. The main peaks to be observed are: CEN stretch from 2260-2200 cm'l, C=C stretch of vinyldene from 1665- 1620 cm'l, and C-C-CEN bend from 580-530 cm". On comparing all of the spectra, it was seen that while the peak at 2200 cm'1 was not found in pure castor oil, pure soybean oil and pure MSO, it could be easily spotted on grafted castor oil, grafted soybean oil and grafted MSO. Also, the peaks at 1620 cm'1 and 530 cm'1 increased in height in grafted MSO, and grafted castor oil, compared to pure MSO and pure castor oil (Figure 5.2.2, 5.2.4, 5.2.5, and 5.2.7). 254 160 — AN Neat 120 4 g? 0 g 80 4 g E '8 40 ~ a h [- 0 I I I I l 4000 3400 2800 2200 1600 1000 400 Wavenunber (cm") Figure 5.2.1: FTIR spectra of pure acrylonitrile 0.8 Castor oil a 0.7 ~ 3, 0 g 0.6 ~ a f a 0.5 7 3E 0.4 ~ 0.3 I I I T I I I l 3900 3400 2900 2400 1900 1400 900 400 Wave nunirer (cm") Figure 5.2.2: FTIR spectra of pure castor oil 255 0.65 SoybeanOil 0.55 ~ E 8 :1 0.45 - a .n In 8 2 0.35 — 0.25 I I r I I I I 3900 3400 2900 2400 1900 1400 900 400 Wave nunirer (cm!) Figure 5.2.3: FTIR spectra of pure soybean oil 40 — MSO neat DJ 0 1 20* Transmittance (%) p-d O l 0 I I I I r 4000 3400 2800 2200 _1 1600 1000 400 Wavenunber(cm ) Figure 5.2.4: F TIR spectra of pure methyl ester of soybean oil 256 20 , —ANCASBP2 Transmittance (%) M -10 f I I T If 4000 3400 2800 2200 1600 1000 400 Wavenunber (cm") Figure 5.2.5: FTIR spectra of grafted castor oil 1.1 - SOY-BPO-AN g 0.9 ~ 0 1 U :1 I 3 f 0.7 ~ 8 a < 0.5 r 0.3 I I I I I I I 3900 3400 2900 2400 1900 1400 900 400 Wave nunber (cm'I) Figure 5.2.6: FTIR spectra of grafted soybean oil 257 OO O -—ANMSOBP3 §60 — 40 ~ :3 E m g 20 ~ I- 0 ‘I I W I I 4000 3400 2800 2200 1600 1000 400 Wavenumbe r(cm'l) Figure 5.2.7: F TIR spectra of grafted methyl ester of soybean oil Reaction scheme Figure 5.2.8 shows the probable reaction scheme of the grafting process of vegetable oil with acrylonitrile. This reaction mechanism has not been confirmed yet, but it might be the way acrylonitrile grafted to the unsaturated backbone of natural oils. 258 —— WEW I Vegetable Oil ' 1 Catalyst (Free radical) Temperature + CH2 =CH—GEN (Acrylonitrile) H2C H i—CN H2O HC—CN Acrylated Oil Figure 5.2.8: Scheme showing probable reaction between vegetable oil and reactive monomer, resulting in modified vegetable oil Notched Izod impact strength Figure 5.2.9 shows the notched Izod impact strength of the bioplastics. Impact energy is the energy absorbed into the specimen during the impact event divided by its cross- sectional area. For the bioplastics, the impact strength increased after addition of biobased oil into the matrix (Figure 5.2.9). The impact strength of plastic containing 259 MSO was 25 % higher than that of neat UPE resin, and impact strength of plastic containing EML is 96 % higher than neat UPE resin. 60 4 {I [:1 BS (MPa) A MOE(GPa) I _ LII O 1 A o 1 W O L N O L Bending Strength (MPa) . 8 Di 0 r A Figure 5.2.9: Impact strength of the bioplastics Legend: A= UPE control, B: MSO-UPE, C= GFT 2C-UPE, D= GFT ZB-UPE, E= GFT 2A-UPE, F= GFT 4A-UPE, G= GFT 4B-UPE, H= GFT 5A-UPE, I= GFT SB-UPE, J= MSO-PBMA-UPE, K= BML-UPE Flexural properties Figure 5.2.10 shows the bending strength and the elastic modulus of the bioplastics. The trend in flexural properties was opposite to that of impact strength. All bioplastics had lower bending strength and elastic modulus as compared to neat UPE resin (Figure 5.2.10). The bending strength of plastic containing modified MSO was 20% lower than strength of neat resin, and its modulus was 30% lower than that of neat resin. In the case of biocomposites made with blends of oils and resin, the impact strength increased, but the flexural properties decreased. 260 p—A A H N l (,I Impact Strength (J/m) Figure 5.2.10: F lexural properties of the bioplastics Legend: A: UPE control, B= MSO-UPE, C= GFT 2C-UPE, D= GFT 2B—UPE, E= GFT 2A-UPE, F= GFT 4A-UPE, G= GFT 4B-UPE, H= GFT 5A-UPE, I= GFT SB-UPE, J= MSO-PBMA-UPE, K= EML—UPE, L=AESO-UPE Biocomposites made with bioresins Biocomposites were also made by impregnating hemp matl and kenaf fibers with blends of natural oils and UPE. Impact strength of biocomposites The impact strength of biocomposite containing hemp fibers in a matrix of bioresin, was 96 % more than that of biocomposite made with hemp fibers and UPE resin (Figure 5.2.11). Similarly, the kenaf based biocomposites gained about 95% strength compared to a biocomposite containing kenaf fibers in UPE matrix. Flexural properties of biocomposites However, the bending strength of the hempmatl biocomposite with bioresin was 20% lower than that of biocomposite with neat resin and the modulus of elasticity for this 261 biocomposite was 7 % higher than that of biocomposite with neat UPE resin (Figure 5.2.12). In case of kenaf composite too, the bending strength decreased (20%) but the elastic modulus was higher (9%) on introduction of bioresin in the polymer matrix. 40 I 10~ ‘ , L ll 0 I l i l '1 A C B DJ 0 1 Impact Strength (J/m) N O D E F Figure 5.2.11: Impact strength of biocomposites containing bioresins Legend: A= UPE control, B: EML-UPE, C= Untreated hempmatl-UPE, D= Untreated hempmatl-EML-UPE, E= Untreated kenaf-UPE, F= Untreated kenaf-EML-UPE 160 8 ElBS (MPa) OMOE(GPa) I Q 2 . I e 5312M ~ 6 E a E“ 80 ~ — 4 m g I “5 U) m vs 5 E” 40 — , — 2 '5 E a i a: I 0 I I I I I 0 A B C D E F Figure 5.2.12: Flexural properties of biocomposites containing bioresins Legend: A: UPE control, B= EML-UPE, C= Untreated hempmatl-UPE, D= Untreated hempmatl -EML—UPE, E= Untreated kenaf-UPE, F = Untreated kenaf-EML-UPE 262 Dynamic mechanical analysis of bioplastics Following the trend of flexural properties, the storage modulus for bioplastics decreased in the same manner as the neat resin. As seen with bending strength and modulus of elasticity, there was a 20—30% reduction in the storage modulus of the bioplastic (figure not shown). This result was expected because vegetable oils are intrinsically low modulus materials because of their molecular structure, and when they are added to the polyester matrix, they lower the stiffness of the resulting material. The glass transition temperature of neat UPE resin, obtained from a tan delta plot was 95 0C (figure not shown). For all bioplastics, the T g decreased by 10—12 0C as compared to neat UPE resin. The glass transition temperatures of plastics based on vegetable oils have been reported from 240 to 80 0C in the literature [21-23]. Morphology of bioplastics / ’II If. ~' \ . i I ‘ 7 ~. . Figure 5.2.13: ESEM micrographs of impact fractures surface of bioplastic containing MSO-PBMA-UPE, a) Magnification 2000X, Scale 25 um, b) 250 X, Scale 200 um 263 ESEM micrographs of the impact fractured surfaces of the bioplastics show phase Figure 5.2.14: ESEM micrographs of impact fractures surface ofbioplastic containing MSO-UPE, a) Magnification 300 X, Scale 150 um b) Magnification 2000X, Scale 25 um separation (Figures 5.2.13 and 5.2.14). These pictures showed discrete microstructures of the dispersed phase which are scattered throughout the entire surface. SEM images (Figure 5.2.15) of the impact fi'actured surfaces of plastic samples showed contrast between the neat UPE resin and bioresin (MSO-PBMA-UPE). Phase separation was observed in the bioplastics samples. From ESEM and SEM images, it was observed that addition of small amount of liquid rubber, PBMA, improved the distribution of the second phase in the bioplastics, by making it more uniformly spread out over the entire surface. 264 265 Flatten Figure 5.2.16: AFM picture of neat unsaturated polyester resin (deflection image, scan size: 5 umX 5 pm) AFM images (Figures 5.2.16—18) also validated the existence of two phases in bioplastic samples, and a single phase in neat polyester resin. As can be seen in Fig. 5.2.17, there was uniformity in the distribution of hemispherical shaped craters of the second phase. According to all ESEM, SEM and AFM images, the second phase seemed to be present in craters or holes, which contain crosslinked molecules of the MESO, UPE and polystyrene. They appeared in different sizes, but had the same content. In Figure 5.2.17, the circular parts contain the second phase, which is the methyl ester of soybean oil. This picture specially looked at the smaller circular particles, and represented different sized domains distributed uniformly across the entire matrix. Figure 5.2.18 shows the inside view of one such crater of the second phase. It was different in structure compared to the neighboring UPE resin. 266 Flatten Figure 5.2.17: AFM picture of bioplastic (MSO-UPE-PBMA) (deflection image, scan size: 60 um X 60 um) A lamellar structure was observed inside the circular crater, validating the difference in mechanical properties of the UPE and MESO. The granular looking objects in Figure 5.2.16 (neat UPE), were probably micro gels formed during curing. It can be inferred that the lamellar structured second phase particles, absorbed the impact energy, and thus improved the impact strength of brittle neat polyester resin. 267 Flatten Figure 5.2.18: AFM picture of bioplastic (MSO-UPE—PBMA) (deflection image, 9 urn X 9 um) Most unmodified thermosets are brittle materials at ambient temperature. Consequently, most polymers have to be impact modified in order to satisfy end-use requirements for rigid applications. In most cases, the problem can be solved by incorporating rubber domains into the polymer matrix. However, the nature and the size of the elastomeric phase have to be adapted to each polymer matrix [24]. In most cases, it is also desirable to use the minimum amount of rubber modifier, in order not to affect other physical properties such as modulus. The most important characteristics of an impact modifier are: rubber glass transition temperature, rubber domain particle size in the matrix, quality of dispersion, and adhesion to the polymer matrix. Historically, several technical approaches for toughening have been developed, which can be divided into three categories all based on the incorporation of an elastomeric damping phase: elastomer introduction during 268 polymerization, dispersion of a thermoplastic elastomer phase during compounding (uncrosslinked), and incorporation of elastomeric core-shell particles. Traditionally, elastomeric additives have been used for toughening in the free radical cross-linking of unsaturated polyester resin [25, 26]. But polyester blends with vegetable based oils can also achieve the same purpose. The blends of oils and polyester resin form semi immiscible systems. Here, the oil phase provided the rubbery structures after curing. The oil phase was acting as an impact modifier; hence, it absorbed impact energy and delayed catastrophic failure. The main toughening mechanisms which have been identified for thermoplastics as well as thermosets are namely: crazing of the polymer matrix, shear yielding of this matrix, and cavitation of the rubber phase. Depending on the polymer system, either a single mechanism or a combination of different mechanisms will be activated [27]. The oil phase of the system consisted of small discrete rubbery particles, with an average size of 5—15 mm. They were randomly distributed in the glassy brittle polyester resin matrix. As reported in the literature about rubber toughened thermosets, these oil phase particles relieved the constraints in the matrix through principal mechanisms of cavitation and forming shear bands [28]. This phase separation phenomenon is similar to that in the low profile mechanism of unsaturated polyester resins. The miscibility and interfacial properties of additive and resin blends play a major role in the toughening process. Improvements in the fracture impact of thermoset polyester resin can be obtained by dispersing elastomer particles with diameters from 0.5 to 5 mm in the blends. The smaller particles produce more microcracks in the matrix and enhance the fracture—impact energy. The additive is immiscible with the resin and gets phase 269 separated from the matrix during curing. Unsaturated polyester resins systems are very reactive materials, which are initially in a liquid state, and become solid after curing. The morphology of the polyester resin systems is decided by thermodynamics and polymerization kinetics. Phase separation brings about a change in polydispersity. Usually, the size of additive particles in the matrix is 1—40 microns. A way to reduce this size is to decrease interfacial tension between two phases. Adding a small percentage of coupling agent like maleic anhydride solves this problem. In this system, poly- (butadiene—maleic anhydride) brought about uniformity in the dispersion, as well as uniformity in sizes of the second phase. Therefore, smaller microcracks were formed in the matrix leading to an increase in impact strength. These microcracks were produced during the large volumetric shrinkage which takes place during polymerization. Microvoids also occurred around the fibers in the composite. Microvoid formation is an important phenomenon for systems consisting of blends of bioresin and UPE. It compensates for resin shrinkage, but induces intrinsic brittleness. Increased additive content in the UP resin should produce higher impact strength, but if the microvoid content is also enhanced, impact strength is reduced. But in this system, the reverse effect was observed. Even after addition of 20% bioresin into the polyester matrix, the impact strength was improved, with a little sacrifice of modulus of elasticity. Well-dispersed particles in the resin matrix induced homogeneous distribution of internal stresses due to network formation. Thus, low energy impact fractures in the unsaturated polyester resin composites were eliminated by the use of this additive in the system. The addition of the blends of derivatized oils and UPE in the matrix of the natural fibre biocomposites lead to an increase in the impact strength of the composites. 270 Thus, using bioresins as polymer matrices provides a two fold benefit over thermoset resins in biocomposites. Their presence can improve the impact strength of the resulting bioplastic, as well as produce a material with higher biobased content. 5.2.2 Bioplastics made by reaction of vegetable oils with maleic anhydride Bioresins were also made by chemical modification of the vegetable oil (SOPEP) with a functional reactive monomer (MA). The OH content of this polyol is 122 mg of KOH. The pure polyol was characterized using mass spectroscopy (MS), nuclear magnetic resonance (NMR), DSC, TGA and FTIR. The pure maleic anhydride and modified polyol were also scanned under FTIR. The bioplastics developed from the bioresins were analyzed by mechanical, thermal and morphological tests. These results are discussed in this section. TGA 1.6 1 J _ t 1.2 2 0.8 ~ ——Werght% é °\e —Derivative Weight q_ 0 8 a E O 6 - ' ’33 .2» B ‘9 __ o 3 0.4 a 0'4 .3 fl \ e 0.2 ~ * 0 5 0 I I I I I -O.4 0 1 00 200 300 400 500 600 Temperature (0 C) Figure 5.2.19: TGA of pure soybean oil phosphate ester polyol (SOPEP) 271 Figure 5.2.19 shows the curves from therrno gravimetric analysis of pure SOPEP. As can be seen from the graph, pure polyol is stable until 200 oC. The maximum decomposition temperature of pure SOPEP is 400 0C. DSC Figure 5.2.20 shows the curves from differential scanning calorimetry of pure SOPEP. The DSC plot shows no crystallization peaks. -I I 3 .2 In. *5 -5 -‘ 0 I -7 -. " '9 I I I I I I I -70 -20 30 80 130 180 230 280 Tenperature (0 C) Figure 5.2.20: DSC of pure soybean oil phosphate ester polyol (SOPEP) FTIR The modified polyol obtained after reacting maleic anhydride with soybean oil polyol, were characterized by FTIR to validate the proof of reaction. FTIR scans of pure maleic anhydride, pure soybean oil, pure soybean polyol, and, modified polyol were taken and compared. Figure 5.2.21 through 5.2.24 show the FT IR spectra of MA, pure soybean oil, 272 pure SOPEP, and grafted SOPEP respectively. In Figure 5.2.21, the spectrum of pure MA is seen. The main peaks to be observed are: C=O symmetric stretch from 1765-1725 cm’1 (1772), C=C stretch from 1680-1630 cm‘1 (1631), CH out of plane deformation from 730-665 cm", C-O-C anti-symmetric stretch from 1280-1070 cm“, coo— at 1567 cm" and OH at 3600 cm". In pure SOPEP (Figure 5.2.23), a high peak of OH is seen at 3467 cm'l which is not seen in pure soybean oil (Figure 5.2.22). In modified SOPEP (Figure 5.2.24), the OH peak at 3460 cm'1 is small compared to that in pure SOPEP. Other significant peaks seen in modified SOPEP are: C=O at 1725 cm", C=C at 1643 cm’l, and COO- at 1547 cm’l. Absorbance N I» in w in l 1 l N L p—o VI 1_ 1 I I I I I I I7 3900 3400 2900 2400 1900 1400 900 400 Wavenurrber Figure 5.2.2]: FTIR spectra of pure maleic anhydride (MA) 273 0.59 - 2 0.49 I a .9 h 8 I: <1 0.39 ~ 0.29 r I I I I I I 3900 3400 2900 2400 1900 1400 900 400 Wavenunirer Figure 5.2.22: FTIR spectra of pure soybean oil 1.05 ~ 3 I §085 ~ I- 8 2 0.65 - 0.45 I I I I I I T 3900 3400 2900 2400 l 900 1400 900 400 Wavenumber Figure 5.2.23: FTIR spectra of pure SOPEP 4.5 3.5 r 8 E N £ a 2 5 r .e < 1.5 ~ 0.5 I I I T I T l 3900 3400 2900 2400 l 900 1400 900 400 Wavenunber Figure 5.2.24: FTIR spectra of modified SOPEP 274 Reaction mechanism Figure 5.2.25 shows the probable reaction scheme of the grafting reaction of soybean oil phosphate ester polyol with maleic anhydride. This reaction mechanism has not been confirmed yet, but it might be the way maleic anhydride attacks the OH groups of SOPEP, and gets grafted. DMA of bioplastics Dynamic mechanical analysis was done on bioplastics with modified polyols GFTSOPEPI, GFTSOPEP2, and GFTSOPEP3. Storage Modulus, tan delta and cross- linking density of the bioplastics were found out. Figures 5.2.26, 5.2.27, and 5.2.28 show the typical storage modulus curves of bioplastics made using GFTSOPEPl, GFTSOPEP2, and GFTSOPEP3. Typical tan delta curves of bioplastics made using GFTSOPEPl, GFTSOPEP2, and GFTSOPEP3 are shown in Figures 5.2.29, 5.2.30, and 5.2.31. 275 I OH OH o—-< OH OH + OH OH O OH OH Soybean Polyol 0” 0“ Maleic Anhydride N,N-Dimethylbenzylaminel T~= 80 0 C OOH OH COOH OH O O I_\/\/\I/H\/Y\/\/l\/ m 0 0 COOH O HO Modified Soybean Polyol Figure 5.2.25: Proposed reaction mechanism for grafting of SOPEP with MA The bioplastics samples with 30%, 40% and 50% GFTYSOPEP 2 had physical problems and couldn’t be tested. The bioplastic samples with 30%, 40% and 50% GFTSOPEP 3 were post-cured. For finding out the T8 of the bioplastic samples with GFTSOPEP2 and GFTSOPSP3, the DMA was carried out from -20 “c to 160 °C. In Figure 5.2.26, the storage modulus of bioplastics samples containing 40% and 50% of GFTSOPEP], dramatically decreases as compared to samples with 10%, 20% and 30% GFTSOPEPI. The storage modulus of samples with containing 40% and 50% of 276 GFTSOPEPI is very low as compared to neat UPE, while the bioplastics containing 10%, 20%, and 30% GFTSOPEPI (8.5-11.5 GPa at room temperature) have very higher storage modulus compared to UPE control (2.3 GPa at room temperature). Similarly, in the case of GFTSOPEP3 bioplastics, the 10%, 20%, and 30% GFTSOPEP3 samples have very higher storage modulus compared to UPE control (Figure 5.2.27). And, for bioplastics containing GFTSOPEP2, the samples with 10% and 20% GFTSOPEP2 samples have very high storage modulus compared to UPE control (Figure 5.2.28). Since flexural and tensile moduli follow the same trend as storage modulus, it can be argued that the bioplastics containing 10, 20, and 30% of GFTSOPEP], 2, and 3 have very high elastic moduli and tensile moduli compared to UPE control. The glass transition temperatures (T8) of the bioplastics samples were evaluated from the maxima of the tan delta plots (Figures 5.2.29, 5.2.30, and 5.2.31). 277 — UPE Control -— 10% GFTSOPEP] -—- 20% GFTSOPEPI 30% GFTSOPEP] — 40% GFTSOPEPI —- 50% GFTSOPEP] 12 ' Storage Modulus (GPa) .b 4 TA; _I._ O 20 70 Temperature(0C) 120 Figure 5.2.26: Storage modulus of bioplastics made with GFTSOPEP] l I—l Storage Modulus (GPa) 2 0 A N 0 — UPE control — 10% GFTSOPEPZ —— 20% GFTSOPEPZ — 30% GFTSOPEPZ I 20 70 0 Temperature (°C) 12 Figure 5.2.27: Storage modulus of bioplastics made with GFTSOPEP2 278 -— UPE Control — 10% GFTSOPEP3 — 20% GFTSOPEP3 — 30% GFTSOPEP3 — 40% GFTOSPEP3 — 50% GFTSOPEP3 l6 y—e N Storage Modulus (GPa) 0° 0 -1 0 40 90 1 40 Temperature (11 C) Figure 5.2.28: Storage modulus of bioplastics made with GFTSOPEP3 — UPE Control — 10% GFTSOPEP] -— 20% GFTSOPEPI ”023.0% GFTSOPEP] -- 40% GFTSOPEPI — 50% GFTSOPEP] Tan Delta 20 70 Temperature (°C) 120 Figure 5.2.29: Tan delta of bioplastics made with GFTSOPEP] 279 — UPE control — 10% GFTSOPEP2 —— 20% GFTSOPEP2 — 30% GFTSOPEP2 0.5 0.4 Tan Delta .0 U) .0 N 0.1 20 70 Temperature (°C) 120 Figure 5.2.30: Tan delta of bioplastics made with GFTSOPEP2 -—- UPE Control — 10% GFTSOPEP3 — 20% GFTSOPEP3 30% GFTSOPEP3 — 40% GFTSOPEP3 — 50% GFTSOPEP3 0.6 0.4 5 '5 a :I‘ a I- 0.2 0 — 10 40 140 Temperature (0090 Figure 5.2.31: Tan delta of bioplastics made with GFTSOPEP3 280 Figures 5.2.32, 5.2.33, and 5.2.34 show the T3 of the bioplastic samples made from all the grafted polyols. From Figure 5.2.32 it was seen that the glass transition temperature increases after addition of grafted soy polyol for 10, 20 and 30% GFTSOPEPI, but for samples with 40% and 50% GFTSOPEP] was not determined from this experiment. The highest T8 was observed for samples with 20% GFTSOPEP]. In the case of bioplastics with GFTSOPEP2 too, the sample with 20% GFTSOPEP2 registered the highest Tg (Figure 5.2.33). The highest Tg in case of bioplastics containing GFTSOPEP3 was demonstrated by the sample with 20% GFTSOPEP3 (Figure 5.2.34). The T8 of samples with 40% and 50% GFTSOPEP was below the T3 of neat UPE. The low T3 of the samples with 40 and 50% grafted polyol hits at higher flexibility and higher mobility of these samples, which was also physically evidenced. 108 92r Glass transition temperature () C) oo 00 I A B C D Figure 5.2.32: Glass transition temperature of bioplastics made with GFTSOPEPI Legend: A=UPE control, B=10% GFTSOPEPl-UPE, C=20% GFTSOPEPl-UPE, D=30°/o GFTSOPEPl-UPE 281 h—I N O 100 ~ _I_ —17 , 00 O 1 Glass transition temperature () C) Ch o 40 r p' . 20‘ ‘ i ,. a ‘ £4 F "fit“ 5;: "’.°'. I ' ‘ mu 0 l I A B C D Figure 5.2.33: Glass transition temperature of bioplastics made with GFTSOPEP2 Legend: A=UPE control, B= 10% GFTSOPEP2-UPE, C: 20% GFTSOPEP2-UPE, D= 30% GFTSOPEP2-UPE 120 .9 0 ’E *1 5 + E —-:r:— a 80‘ E Q) [- fl .2 '2 40— ‘ l a “,I I» ”t I 2 I; ”a I “I 0 1' ii- Tu 0 I 4 I I I" I I A B C D E F Figure 5.2.34: Glass transition temperature of bioplastics made with GFTSOPEP3 Legend: A=UPE control, B=10% GFTSOPEP3—UPE, C= 20% GFTSOPEP3—UPE, D: 30% GFTSOPEP3-UPE, E= 40% GFTSOPEP3—UPE, F= 50% GFTSOPEP3-UPE Figures 5.2.35, 5.2.36, 5.2.37 show the cross link density of the bioplastic samples. An important property from a characterization standpoint of crosslinked networks is the number of crosslinked sites and the molecular weight or length of chain between 282 sites. The crosslink density of bioplastics samples were measured in the rubbery region, 40 OC above the Tg of the material, by the following formula: DC: E, /(3*R*T) —(10) where, uc=crosslink density, R=universal gas constant, T=Tg+40 0C, and BI: storage modulus at temperature T. The change in glass transition temperature is directly proportional to crosslink density. As compared to neat polyester resin, 10, 20 and 30% GFTSOPEP] samples show a large increment in the cross linking density. With 40 and 50% GFTSOPEP], cross link density couldn’t be found out because their Tg was not in the range on the temperature profile of the DMA test (room temperature to 1600C @ 4 0C/min). For finding out Tg of these samples, the DMA will have to carried out from a much lower temperature, like -20°C to 1600C. Similarly, in the case of GFTSOPEP2 bioplastics, the 10%, 20%, and 30% GFTSOPEP2 samples have very higher cross linking density compared to UPE control (Figure 5.2.36). And, for bioplastics containing GFTSOPEP3, the samples with 10%, 20%, and 30% GFTSOPEP3 samples have very high cross linking density compared to UPE control (Figure 5.2.37). These results are in coherence with the storage modulus results. The high values of cross link density of the bioplastic samples demonstrate a tightly wound rigid network of the polymer formed by blending bioresins with UPE. 283 0.016 0.012 - 0.008 1 Crosslink densrty(mol/nf) .o 8 A 0- A B C D Figure 5.2.35: Crosslinking density of bioplastics with GFTSOPEPI Legend: A=UPE control, B=10% GF'I‘SOPEPl-UPE, C= 20% GFTSOPEPl-UPE, D= 30% GFTSOPEPl-UPE 0.012 “E =I e0.008 ~ 5 E E 1: 30.004 - E 8 l- U 0 - I I A B C D Figure 5.2.36: Crosslink density of bioplastics made with GFTSOPEP2 Legend: A=UPE control, B=10% GFTSOPEP2-UPE, C: 20% GFTSOPEP2-UPE, D= 30% GFTSOPEP2-UPE 284 0.016 F3 0 id N 1 20162261 *3‘ O 'o if 1 Cross link density (mol/rrf) 2 O 00 an . .7" EDI": Ti . a Y 33".. 0 _ A B C D E F Figure 5.2.37: Crosslink density of bioplastics made with GFTSOPEP3 Legend: A=UPE control, B=10% GFTSOPEP3-UPE, C= 20% GFTSOPEP3-UPE, D= 30% GFTSOPEP3-UPE, E= 40% GFTSOPEP3-UPE, F= 50% GFTSOPEP3-UPE Impact strength of bioplastics The notched Izod impact strength of the bioplastic samples made from all the grafted polyols are shown in Figures 5.2.38, 5.2.39, and 5.2.40. 20 H II! 1 Impact Strength (J/m) , 8 S—I ii -* 0 I I I I I A B C D E F Figure 5.2.38: Impact strength of bioplastics made with GFTSOPEPl Legend: A=UPE control, B=10% GFTSOPEPl-UPE, C: 20% GFTSOPEPl-UPE, D= 30% GFTSOPEPl-UPE, E: 40% GFTSOPEPl-UPE, F: 50% GFTSOPEPl-UPE 285 12 J -- Impact Strength (J/m) A B C Figure 5.3.39: Impact strength of bioplastics made with GFTSOPEP2 Legend: A=UPE control, B=10% GFTSOPEP2-UPE, C= 20% GFTSOPEP2-UPE 20‘ " 516‘ E B 312“ ‘I' E o 3: m 1 *5 8~ «I a. E 4_ O I A B C D E F Figure 5.2.40: Impact Strength of bioplastics made with GFTSOPEP3 Legend: A=UPE control, B=10% GFTSOPEP3-UPE, C= 20% GFTSOPEP3-UPE, D= 30% GFTSOPEP3-UPE, E= 40% GFTSOPEP3-UPE, F= 50% GFTSOPEP3-UPE It is well known that impact strength and storage modulus (DMA) track trends opposite to each other. It was earlier seen that storage modulus of 10%, 20% and 30% 286 GFTSOPEPI bioplastic samples was much higher than UPE control, while that of 40% and 50% GFTSOPEPI samples was much lower than the storage modulus of neat resin (Figure 5.2.26). From Figure 5.2.37 it was observed that impact strengths of 10%, 20%, and 30% GFTSOPEP] is much lower than neat UPE , while that of 40% and 50% GFTSOPEPl is higher than that of neat UPE. This trend was also seen in impact strength of bioplastics samples of GFTSOPEP2 and GFTSOPEP3. Thus, the flexible samples containing 40% and 50% grafted polyols were much higher in toughness, and low in stiffness. Morphology of bioplastics samples 287 ‘ 3.0“- ”A; . - A a I Figure 5.2.41: ESEM micrographs of impact fractures surface of bioplastic containing a) 10%GFTSOPEPl-UPE, Magnification 3000 X, Scale bar 15 um 1)) 20% GFTSOPEPl-UPE, Magnification 300 X, Scale bar 150 um c) 30% GFTSOPEPl-UPE, Magnification 800 X, Scale bar 50 um d) 40% GFTSOPEPl-UPE, Magnification 285 X, Scale bar 150 um Figures 5.2.41 and 5.2.42 show the ESEM micrographs of impact fractures surfaces of bioplastics made from grafted polyols. In all of the bioplastics examined under ESEM, phase separation was observed between grafted soy polyol rich phase UPE rich phase. But the surface features changed with different compositions. 288 and Figure 5.2.42: ESEM micrographs of impact fractures surface of bioplastic containing a) 10%GFTSOPEP3-UPE, Magnification 400 X, Scale bar 100 um b) 20% GFTSOPEP3-UPE, Magnification 400 X, Scale bar 100 um c) 30% GFTSOPEP3-UPE, Magnification 600 X, Scale bar 200 um (I) 40% GFTSOPEP3-UPE, Magnification 300 X, Scale bar 150 um In all of the bioplastics, circular or spherical shapes are observed on the surface. In bioplastics with 10 and 20% modified polyol, there were only circular trenches, where the dispersed phase might be present. The diameter of these circular objects varied greatly, but these were present all over the entire fractured surface. In 289 bioplastics with 30 % modified polyol, circular as well as spherical objects were seen on the surface. In bioplastics with 40 and 50% modified polyol, there are more spherical objects and lesser circular objects. The size of the spherical looking balls also varied over the entire surface, but the distribution was very uniform. Figures 5.2.43 through 5.2.46 show the AFM images of a sample containing 20% GFTSOPEP3 in a UPE matrix. The images for the GFTSOPEP3 samples show phase separation. The height features are indicators of the second phase in these pictures. The picture focuses on a relatively empty portion of the matrix, where small particles of second phase, which is modified polyol, can be seen, strewn about in the entire area. The morphology of this sample can be clearly observed from TV images of this sample shown below in Figures 5.2.47. Flatten 0 7.5.0 50.0 Figure 5.2.43: AFM picture of 20% GFTSOPEP3-UPE (deflection image, 50 um X 50 Inn) 290 Flatten 50.0 25.0 O 0 25.0 50.0 Figure 5.2.44: AFM picture of 20% GFTSOPEP3-UPE (deflection image, 50 um X 50 Ian) Flatten o 5.0 10 . o 15 . 0 Figure 5.2.45: AFM picture of 20% GFTSOPEP3-UPE (Z scale: 1000 nm, scan size, 16 mm X 16 pm) 291 Flatten 0 5.0 10.0 15.0 Figure 5.2.46: AFM picture of 20% GFTSOPEP3-UPE (deflection image, 16 um X 16 Inn) Figure 5.2.47: Picture of 20% GFTSOPEP3—UPE on TV screen Figure 5.2.44 focused on an area devoid of circular geometries, and looks at second phase particles distributed on the UPE matrix. Figure 5.2.45 showed the top view of a spherical object on the matrix. This object looks like a mercury drop on a solid surface, but is composed of both modified polyol and UPE resin. The concave shape of the spherical object can be easily seen. This is a height image, while Figure 5.2.46 292 is the deflection image of the same object. The composition of the spherical object was clearly observed here. Thus phase separated bioplastic samples resulted in changes in mechanical, thermal and morphological characteristics of the polymers formed. 293 References 1. M. R. Kamal, S. Sourour, Polymer Engineering and Science, 13(1), (1973), 59—64. 2. J. M. Salla, J. L. Martin, Thermochimica Acta, 126, (1988), 339-354. 3. J. M. Salla, X. Ramis, J. L. Martin, A. Cadenato, Thermochimica Acta, 134, (1988), 261-167. 4. M. Paci, E. D. Vecchio, F. Campana, Polymer Bulletin, 15, (1986), 21-27. 5. X. Ramis, J. M. Salla, Journal of Applied Polymer Science, 45, (1992), 227-236. 6. D. Rouison, M. Sain, M. Couturier, Journal of Applied Polymer Science, 89, (2003), 2553-2561. 7. A. Nzihou, P. Sharrock, A. Ricard, Chemical Engineering Journal, 72, (1999), 53- 61. 8. J. L. Vilas, J. M. Laza, M. T. Gary, M. Rodriguez, L. M. Leon, Journal of Applied Polymer Science, 79,(2001), 447-457. 9. Y-J. Huang, J-S. Leu, Polymer, 34(2), (1993), 295-304. 10.H. Yang, L. J. Lee, Journal of Applied Polymer Science, 79, (2001), 1230-1242. 11.H. Yang, L. J. Lee, Journal of Applied Polymer Science, 84, (2002), 211-227. 12.J. S. Chen, T. L. Yu, Journal of Applied Polymer Science, 69, (1998), 871-878. 13.V. Kosar, Z. Gomzi, Chemical and Biochemical Engineering Quarterly, 15 (3), (2001), 101—108. l4.Y.-M. Yun, S.-J. Lee, K.-J. Lee, Y.—K. Lee, J.-D. Nam, Journal of Polymer Science: Part B: Polymer Physics, 35, (1997), 2447—2456. 15.8. Ziaee, G. R. Palmese, Journal of Applied Polymer Science, 37, (1999), 725-744. 16R. 11’. Brill, G. R. Palmese, Journal of Applied Polymer Science, 76, (2000), 1572- 1582. 17B. L. Grunden, C. S. P Sung, Macromolecules, 36, (2003), 3166-3173. 18.R-H. Lin, J-H. Hsu, Polymer lntemational, 50, (2001), 1073-1081. 19.L. Shan, C. G. Robertson, K. N. E. Verghese, E. Burts, J. S. Riffle, T. C. Ward, K .L. Reifsnider, Journal of Applied Polymer Science, 80,(2001), 917-927. 294 20.1nfrared and Raman Spectroscopy, Chapter 7, Group Frequencies: Infrared and Raman, 159-182. Part B: Series Volume: 1 E. G. Brame; J. G. Grasselli, Print Published1977 21. F. Li and R. C. Larock, J. Polym. Sci., Part B: Polym. Phys., 2001, 39, 60—77. 22.13. Can, S. Kusefoglu and R. P. Wool, J. Appl. Polym. Sci., 2001, 81, 69. 23.1. Frichinger and S. Dirlikov, in Toughened Plastics I Science and Engineering, ed. C. K. Riew and A. J. Kinloch, 1993, (Advances in Chemistry, No. 233), American Chemical Society, Washington, DC, pp. 451—489. 24.B. Cherian and E. T. Thachil, Progr. Rubber Plastics T echnol., 2001, 17(4), 205— 224 25. L. Suspene, Y. S. Yang and J. P. Pascault, in Toughened Plastics I Science and Engineering, ed. C. K. Riew and A. J. Kinloch, 1993, (Advances in Chemistry, No. 233), American Chemical Society, Washington, DC, pp. 163—188. 26. F. J. McGarry and R. Subramanium, in Toughened Plastics [1 Novel Approaches in Science and Engineering, ed. C. K. Riew and A. J. Kinloch, 1996, (Advances in Chemistry Series, No. 252), American Chemical Society, Washington, DC, pp. 133— 149. 27.C. B. Bucknall, Toughened Plastics, Applied Science Pub. Ltd., London, 1977. 28. A. F. Yee and R. A. Pearson, J. Mater. Sci., 1986, 21, 2462. 295 CHAPTER SIX: RESULTS ON NEW PROCESSING 6.0 RESULTS ON NEW PROCESSING The results obtained from testing biocomposites manufactured with a new SMC technique are described in this chapter. 6.] Biocomposites from SMC The new SMC processing for natural fibers can be done on normal commercial SMC equipment with a minor change on the way fibers are fed to the line. In industrial SMC set-ups, the glass fibers rovings are fed to a chopper, and chopped fibers fall onto the carrier film [1-11]. But the natural fibers are not supplied as continuous rovings or yarns, and must be fed to the film in a chopped form. Therefore, in the new set —up, the chopped fibers fall from a calibrated vibratory feeder onto the carrier film, and get drenched in resin flowing from two resin pots, and advance to the compression rollers, and finished product is obtained at the end of the line. The prepreg from SMC line is refrigerated for certain time to reach desirable gelation. The gelled product is then compression molded in variety of molds to get the desired shape. For glass composites: The glass fiber polyester resin composites from SMC line were tested for mechanical and thermal properties including, bending strength, modulus of elasticity, storage modulus and tan delta. The flexural properties of glass composites are shown in Figure 6.1. There was consistency in the data for samples B and C, which represented glass composites made using SMC processing. This was a verification of the fact that SMC process had been optimized for Glass-UPE-CaCO3 system. The bending strength and modulus of elasticity 296 for samples B and C are almost same at 125 MPa, and 12.5 GPa respectively. The bending strength of glass composites was 32 % more than that of UPE control, while their modulus of elasticity was 370 % more than neat polyester resin. a BS (Mpa) o MOE(GPa) 160 20 a 33 °- T T — 16 Q E. 120 - a '5 '5 g I I 12 '5 g 80 ~ II-“I’ (D _ 8 n— O .3 g .2 40 ‘ _ 4 3 iii 5 g o I I 0 A B C Figure 6.1: F lexural properties of glass composites Legend: A= UPE control, B= SMC Composite 1, C: SMC Composite 2 The storage modulus of glass fiber—UPE composites at 40 0C is shown in Figure 6.2. Again, both B and C have almost the same values of storage modulus. As compared to UPE control, glass composites had very high storage modulus, (1600 % more than that of neat polyester) which reflected high stiffness of glass fibers. Figure 6.3 shows a typical dynamical mechanical analysis plot for glass fiber composites. As is common in all thermoset systems, the storage modulus decreased with increasing temperature. The glass transition temperature of this composite was 110 0C. As compared to storage modulus, loss modulus was very low over the entire range of test temperature. 297 60 I“,- w o 1 «h o 1 Storage Modulus (GPa) N (a) O O ...L 0 L1— A B C Samples Figure 6.2: Storage modulus of glass composites at 40 0C Legend: A= UPE control, B= SMC composite l, C= SMC composite 2 — Storage Modulus —- Loss Modulus — Tan Delta 60 0.3 A 50 ~ (I 6‘5 ;' 40 ‘I r 0.2 2 a § 30 - a a E a. 20 ~ — 0.1 E m 10 - 0 e 0 1 00 Temperature (°C) Figure 6.3: DMA plot for glass composites 1 The data for glass fiber composites confirmed that consistent materials had indeed been produced, meaning thereby, that the process parameters also had been optimized. Thus, the next step was development of SMC processed natural fiber composites. 298 For natural fibers: The thermal properties of natural fibers used for making biocomposites are shown in Figures 6.4, 6.5 and Table 6.1. Figure 6.4 shows the plots from DSC for five fiber samples. All the fibers showed same transition from -60 0C to 300 0C. A trough was observed in all of the five samples. This ranged from 74 0C for untreated BBSG, 94 0C for untreated green flax core, 106 0C for untreated hemp, 83 0C for silane treated BBSG, to 99 0C for silane treated green flax core. This was because of evaporation of water from natural fibers. This phenomenon is very common for natural fibers. From this data, it was interpreted that after silane treatment of fibers, the water evaporation occurs at a higher temperature as compared to that for untreated fibers. Figure 6.5 shows TGA plots for the same five fiber samples which were analyzed in DSC. Here too, all fibers showed same kind of transition. The maximum degradation temperatures were found out from the TGA plots and are listed in Table 6.1. The maximum degradation temperature for BBSG and green flax core increased after silane treatment. The percentage weight at 600 0C ranged from 18 % for untreated BBSG to 22.5 % for silane treated BBSG and silane treated green flax core. At higher temperatures, big blue stem grass was more stable than flax, as can be seen from the weight % curves. 299 Haet Flow (mW) L do -12 -70 30 130 230 Temperature (°C) Figure 6.4: DSC plots of natural fibers used for SMC line Legend: A= Untreated big blue stem grass, B= Untreated green flax core, C=Untreated hemp, D= Silane treated big blue stem grass, E= Silane treated green flax core —A —B -—-C ——D E Weight (%) O) O 40 r 20 r 0 Ai——— WiiI If I I F I I A 0 1 00 200 300 400 500 600 Temperature (°C) Figure 6.5: TGA plots of natural fibers used for SMC line 300 Legend: A= Untreated big blue stem grass, B= Untreated green flax core, C=Untreated hemp, D= Silane treated big blue stem grass, E= Silane treated green flax core Table 6.1: Maximum degradation temperatures for natural fibers Samples Td,max (C) A 305.47 B 305.79 C 307.96 D 326.54 E 307.21 Legend: A: Untreated big blue stem grass, B= Untreated green flax core, C=Untreated hemp, D= Silane treated big blue stem grass, E= Silane treated green flax core Table 6.2 and 6.3 show the results from XPS of natural fiber samples. These results also depict changes occurring in the fibers after surface treatment. According to Table 6.2, untreated BBSG and green flax core, contain no silicon, while the presence of silicon was found in silane treated BBSG and green flax core. As compared to untreated BBSG and green flax core, silane treated BBSG and green flax core had a decreased concentration level of carbon, and an increased concentration of oxygen. Table 6.3 shows the ratio of atomic concentrations of C/O, C/Si and C/N in all four samples. The C/O ratio decreased while moving from untreated BBSG to silane treated BBSG, and from untreated flax core to silane treated flax core. Between silane treated BBSG and silane treated flax core, C/Si ratio was higher for silane treated flax core. Table 6.2: Atomic concentrations on the surfaces of fibers used for SMC line C1s[.314] le[.499] Ols[.733] Si2p[.368] A 85.77 2.55 11.68 0 B 84.55 1.46 13.99 0 C 84.26 1.24 12.3 2.2 D 81.66 1.51 15.57 1.25 Legend: A= Untreated big blue stem grass, B= Untreated green flax core, C=Silane treated big blue stem grass, D= Silane treated green flax core 301 Table 6.3: Ratio of atomic concentrations of fibers C/Si C/O C/N 7.34332 33.6353 6.0436 57.911 38.3 6.85041 67.9516 65.328 5.2447 54.0795 Legend: A: Untreated big blue stem grass, B= Untreated green flax core, C=Silane treated big blue stem grass, D= Silane treated green flax core U003> Due to the large number of composites, the mechanical and thermal properties of these composites have been divided into two groups, and each group is individually discussed. The groups are: composites containing 20% calcium carbonate, and the composites containing no calcium carbonate at all. The properties of the group with CaCO3 are first discussed. The tensile strengths and moduli of SMC produced biocomposites containing calcium carbonate, are shown in Figure 6.6. The bars represent tensile strength and the points denote tensile modulus. The tensile strength of silane treated big blue stem biocomposite is 29% more than strength of untreated big blue stem (BBSG) biocomposite. While, the tensile strength of silane treated big blue stem & green flax core biocomposite, is 25% more than strength of untreated big blue stem & green flax core biocomposite. And, the tensile strength of untreated treated big blue stem biocomposite is 12% more than strength of untreated big blue stern & flax biocomposite. The strength of untreated jute-hemp (25 wt %) hybrid biocomposite is 70% higher than that of untreated BBSG biocomposite. The strength of untreated green flax core biocomposite is 40% lower than that of untreated BBSG biocomposite. The strength of untreated jute-hemp (25 wt %) hybrid biocomposite is 20% higher than that of untreated jute-hemp (20 wt %) hybrid biocomposite. The strength 302 of chopped E-glass composite is 20% higher than that of hybrid of E-glass mat-Hemp composite. The strength of chopped E-glass composite is 170% higher than that of untreated BBSG biocomposite. The strength of E-glass mat (30wt%)-Bioresin (20wt %)- is 40% higher than that of chopped E-glass composite. In case of tensile modulus, silane treated big blue stem biocomposite, has a modulus value 11.5% more than strength of untreated big blue stem biocomposite. The tensile modulus of silane treated big blue stem & green flax core biocomposite, is 12% more than that of untreated big blue stem & green flax core-UPE-CaCO3. And, the tensile modulus of untreated treated big blue stem biocomposite is 22% more than modulus of untreated big blue stem & green flax core biocomposite. The tensile modulus of untreated jute-hemp (25 wt %) hybrid biocomposite is 65 % higher than that of untreated BBSG biocomposite. The modulus of untreated green flax core biocomposite is 27 % lower than that of untreated BBSG biocomposite. The modulus of untreated jute-hemp (25 wt %) hybrid biocomposite is 28 % higher than that of untreated jute-hemp (20 wt %) hybrid biocomposite. The modulus of chopped E-glass composite is 18 % higher than that of hybrid of E-glass mat-Hemp composite. The modulus of chopped E-glass composite'is 100% higher than that of untreated BBSG biocomposite. The modulus of E-glass mat (30wt%)-Bioresin (20wt %)-is 60% lower than that of chopped E-glass composite. The highest tensile strength was of the samples containing E-glass mat (30wt%)-Bioresin (20wt %). The chopped E-glass composite and E-glass mat-hemp hybrid biocomposite had second and third highest tensile strengths, respectively. The highest tensile modulus was of the samples containing chopped E-glass. The E-glass mat-hemp hybrid biocomposite and untreated jute-hemp (25 wt %) biocomposite had second and third 303 highest tensile moduli, respectively. The low values of tensile strengths and moduli of composites containing big blue stem grass and grass flax core were because of short length of these fibers. 51 TS (MPa) 0 TM (GPa) 70 14 O) O 1 U" o 1 Tensile Strength(MPa) Tensile Modulus(GPa) s A G Figure 6.6: Tensile properties of biocomposites Legend: A=UPE Control, B= Untreated BBSG-Caco3-UPE, C= Silane treated BBSG- CaCO3-UPE, D= Untreated flax&BBSG—CaCO3-UPE, E=Silane treated Flax &BBSG- CaCO3-UPE, F= Jute-Hemp(25%wt)-CaCO3-UPE (SMC), G= Untreated core flax- CaCO3-UPE (SMC), H= Jute-hemp(20 wt %)- CaCO3-UPE (SMC), I= E-glassmat- Hemp(20 wt%)— CaCO3-UPE (SMC), J= Chopped glass(20 wt%)-CaCO3-UPE (SMC), K= E-glass mat(30wt%)-Bioresin(20wt %)-UPE The bending strengths and moduli of elasticity of SMC produced biocomposites containing calcium carbonate, are shown in Figure 6.7. The bars represent bending strength and the points denote modulus of elasticity. The bending strength of silane treated big blue stem-UPE-CaCO3, is 15% more than strength of untreated big blue stem-UPE-CaCOg. While, the bending strength of silane treated big blue stem & flax -UPE-CaCO3, is 10% more than strength of untreated big blue stem & flax-UPE-CaCO3. And, the bending strength of untreated treated big blue 304 stem-UPE-CaCO3, is 6% more than strength of untreated big blue stem & flax-UPE- CaCO3. The strength of untreated jute-hemp (25 wt %) hybrid biocomposite is 165% higher than that of untreated BBSG biocomposite. The strength of untreated green flax core biocomposite is 7% lower than that of untreated BBSG biocomposite. The strength of untreated jute-hemp (25 wt %) hybrid biocomposite is 40% higher than that of untreated jute-hemp (20 wt %) hybrid biocomposite. The strength of chopped E-glass composite is 14% lower than that of hybrid of E-glass mat-Hemp composite. The strength of chopped E-glass composite is 215% higher than that of untreated BBSG biocomposite. The strength of E-glass mat (30wt%)-Bioresin (20wt %)-is 40% lower than that of chopped E-glass composite composite. In case of modulus of elasticity, silane treated big blue stem-UPE-CaCO3, has a modulus value 21% more than strength of untreated big blue stem-UPE-CaCO3. The modulus of elasticity of silane treated big blue stem & flax -UPE-CaCO3, is 17% more than that of untreated big blue stem & flax-UPE-CaCO3. And, the modulus of elasticity of untreated treated big blue stem-UPE-CaCO3, is 5% more than modulus of untreated big blue stem & flax-UPE-CaCO3. The modulus of elasticity of untreated jute-hemp (25 wt %) hybrid biocomposite is 102 % higher than that of untreated BBSG biocomposite. The modulus of elasticity of untreated green flax core biocomposite is 17 % lower than that of untreated BBSG biocomposite. The modulus of elasticity of untreated jute-hemp (25 wt %) hybrid biocomposite is 25 % higher than that of untreated jute-hemp (20 wt %) hybrid biocomposite. The modulus of elasticity of chopped E-glass composite is 5 % higher than that of hybrid of E-glass mat-Hemp composite. The modulus of elasticity of chopped E- glass composite is 126% higher than that of untreated BBSG biocomposite. The modulus 305 of elasticity of E-glass mat (30wt%)-Bioresin (20wt %)-is 55% lower than that of chopped E-glass composite composite. The highest bending strength was of the samples containing E-glass mat-hemp hybrid. The chopped E-glass composite and untreated jute-hemp (25 wt %) hybrid biocomposite had second and third highest bending strengths, respectively. The highest modulus of elasticity was of the samples containing chopped E-glass. The E-glass mat-hemp hybrid biocomposite and untreated jute-hemp (25 wt %) hybrid biocomposite had second and third highest moduli of elasticity, respectively. The bending strengths and moduli of elasticity followed the same trend as tensile strengths and moduli. a BS (MPa) 0 MOE(GPa) 160 12 140 -« L 10 120 ~ 100 - {7 i 8 80— 60- i i i 88 20- 0 -. . A B C D E Bending Strength(MPa) Modulus of Elasticity (Gpa) ’///////////////.Z ///////////////fl/’IZ \ '11 Q I c. K Figure 6.7: Flexural properties of biocomposites Legend: A=UPE Control, B= Untreated BBSG-Caco3-UPE, C= Silane treated BBSG- CaCO3-UPE, D= Untreated flax&BBSG-CaCO3-UPE, E=Silane treated Flax &BBSG- CaCO3-UPE, F= Jute-Hemp(25%wt)-CaCO3-UPE (SMC), G= Untreated core flax- CaCO3-UPE (SMC), H= Jute-hemp(20 wt %)- CaCO3-UPE (SMC), I= E-glassmat- Hemp (20 wt%)— CaCO3-UPE (SMC), J= Chopped glass(20 wt%)-CaCO3-UPE (SMC), K= E-glass mat(30wt%)-Bioresin(20wt %)-UPE Figure 6.8 shows the impact strength of composites of SMC produced biocomposites containing calcium carbonate. The impact strength of silane treated big blue stem-UPE- 306 CaCO3, is 65% less than strength of untreated big blue stem-UPE-CaCO3. While, the impact strength of silane treated big blue stem & flax -UPE-CaCO3, is 20% less than strength of untreated big blue stem & flax-UPE-CaCOB. And, the impact strength of untreated treated big blue stem-UPE-CaC03, is 66% less than strength of untreated big blue stem & flax-UPE-CaCO3. The strength of untreated jute-hemp (25 wt %) hybrid biocomposite is 86% higher than that of untreated BBSG biocomposite. The impact strength of untreated green flax core biocomposite is 20% lower than that of untreated BBSG biocomposite. The impact strength of untreated jute-hemp (25 wt %) hybrid biocomposite is 20% higher than that of untreated jute-hemp (20 wt %) hybrid biocomposite. The impact strength of chopped E-glass composite is 83% higher than that of hybrid of E-glass mat-Hemp composite. The impact strength of chopped E-glass composite is 1330% higher than that of untreated BBSG biocomposite. The impact strength of E-glass mat (30wt%)—Bioresin (20wt %) is 7% higher than that of chopped E- glass composite composite. The impact strengths of composites followed a pattern completely opposite to that of bending and tensile strengths. This is a common behavior for fiber reinforced plastics. The highest impact strength was of the samples containing of E-glass mat (30wt%)- Bioresin (20wt %). The chopped E-glass composite, and E-glass mat-Hemp hybrid composite had second and third highest impact strength, respectively. The impact strengths of composites containing big blue stem grass and grass flax core were very small, because, these fibers very small in length. In particular, the length of BBSG fibers was about 4 mm, and that of green flax core was 1mm. 307 200 - Notched Izod Impact Strength ’5‘ 51504 5 O) 5 a ,"3 100 - 0 a a. E 50 - 0 atammmgagl, A B C D E F G H Figure 6.8: Impact properties of biocomposites Legend: A=UPE Control, B= Untreated BBSG-Caco3-UPE, C= Silane treated BBSG- CaCO3-UPE, D= Untreated flax&BBSG-CaCO3-UPE, E=Silane treated Flax &BBSG- CaCO3-UPE, F= Jute-Hemp(25%wt)-CaCO3-UPE (SMC), G= Untreated core flax- CaCO3-UPE (SMC), H= Jute-hemp (20 wt %)- CaCO3-UPE (SMC), I= E-glassmat- Hemp(20 wt%)- CaCO3-UPE (SMC), J= ChOpped glass(20 wt%)-CaCO3-UPE (SMC), K= E-glass mat(30wt%)-Bioresin(20wt %)-UPE The storage modulus of composites of SMC produced biocomposites containing calcium carbonate, are shown in Figure 6.9. The storage modulus of silane treated big blue stem- UPE-CaCO3, at 40 0C, is 17.5% more than that of untreated big blue stem-UPB-CaCO3. The modulus of silane treated big blue stern & flax -UPE-CaCO3 is 13 % more than modulus of untreated big blue stem & flax-UPE-CaCO3. The modulus of untreated 308 treated big blue stem -UPE-CaCO3 is 26 % more than modulus of untreated big blue stem & flax-UPE-CaCO3. The modulus of silane treated big blue stem -UPE-CaCO3 is 32 % more than modulus of silane treated big blue stem & flax-UPE-CaCO3. The modulus of untreated jute-hemp (25 wt %) hybrid biocomposite is 45 % higher than that of untreated BBSG biocomposite. The modulus of chopped E-glass composite is 105% higher than that of untreated BBSG biocomposite. The modulus of E-glass mat (30wt%)-Bioresin (20wt %) is 55 % lower than that of chopped E-glass composite. The data for storage modulus followed the same trend as tensile modulus and modulus of elasticity. Storage Modulus at 40 °C 1 2000 10000 ~ 8000 a 6000 _ 4000 ~ 2000 — I 0 _ :'.:":.l A B C D E F G H Storage Modulus (MPa) Figure 6.9: Storage modulus of biocomposites at 40 0C Legend: A=UPE Control, B= Untreated BBSG-CaCO3-UPE, C= Silane treated BBSG- CaCO3-UPE, D= Untreated flax&BBSG-CaCO3-UPE, E=Silane treated Flax &BBSG- CaCO3-UPE, F= Jute-Hemp(25%wt)-CaCO3-UPE (SMC), G= Untreated core flax- CaCO3-UPE (SMC), H= Jute-hemp(20 wt %)- CaCO3-UPE (SMC), I= E-glassmat- Hemp(20 wt%)- CaCO3-UPE (SMC), J= Chopped glass(20 wt%)-CaCO3-UPE (SMC), K= E-glass mat(30wt%)-Bioresin(20wt %)-UPE 309 The tensile strengths and moduli of SMC produced biocomposites containing no calcium carbonate, are shown in Figure 6.10. The bars represent tensile strength and the points denote tensile modulus. The tensile strength of untreated hemp (25 vol %)-UPE biocomposite is 145 % more than strength of untreated henequen (25 vol %)-UPE biocomposite. The tensile strength of untreated kenaf (25vol %)-UPE biocomposite, is 138 % more than strength of untreated henequen-UPE biocomposite. And, the tensile strength of untreated hemp-henequen - UPE biocomposite is 107 % more than strength of untreated henequen (25 vol %)-UPE biocomposite. The strength of untreated kenaf-henequen-UPE hybrid biocomposite is 13 % lower than that of untreated kenaf-UPE biocomposite. The strength of untreated hemp- henequen (30 wt %)- bioresin (20 wt%) hybrid biocomposite is 10% lower than that of untreated hemp biocomposite. The strength of untreated henequen biocomposite is 38 % lower than that of neat polyester resin. The strength of untreated hemp biocomposite is 51 % higher than that of neat polyester resin. The strength of untreated kenaf biocomposite is 47 % higher than that of neat polyester resin. The strength of untreated hemp-henequen hybrid biocomposite is 27 % higher than that of neat polyester resin. The strength of untreated kenaf-henequen hybrid biocomposite is 27% higher than that of neat polyester resin. The strength of untreated hemp-henequen (30 wt %)-bioresin (20 wt %) hybrid biocomposite is 35 % higher than that of neat polyester resin. In case of tensile modulus, the modulus of untreated hemp (25 vol %)-UPE biocomposite is 104 % more than modulus of untreated henequen (25 vol %)-UPE biocomposite. The tensile modulus of untreated kenaf (25vol %)-UPE biocomposite, is 125 % more than modulus of untreated henequen-UPE biocomposite. And, the tensile modulus of untreated 310 hemp-henequen -UPE biocomposite is 120 % more than modulus of untreated henequen (25 vol %)-UPE biocomposite. The modulus of untreated kenaf-henequen-UPE hybrid biocomposite is 24 % lower than that of untreated kenaf-UPE biocomposite. The modulus of untreated hemp-henequen (30 wt %)-bioresin (20 wt %) hybrid biocomposite is 14 % lower than that of untreated hemp biocomposite. The modulus of untreated henequen biocomposite is 195 % higher than that of neat polyester resin. The modulus of untreated hemp biocomposite is 500 % higher than that of neat polyester resin. The modulus of untreated kenaf biocomposite is 560 % higher than that of neat polyester resin. The strength of untreated hemp-henequen hybrid biocomposite is 550 % higher than that of neat polyester resin. The modulus of untreated kenaf-henequen hybrid biocomposite is 402 % higher than that of neat polyester resin. The modulus of untreated hemp-henequen (30 wt %)-bioresin (20 wt %) hybrid biocomposite is 415% higher than that of neat polyester resin. The highest tensile strength was of the samples containing untreated hemp fibers. The untreated kenaf biocomposite and untreated hemp-henequen hybrid biocomposite had second and third highest tensile strengths, respectively. The highest tensile modulus was of the samples containing untreated kenaf fibers. The untreated hemp-henequen hybrid biocomposite and untreated hemp biocomposite had second and third highest tensile moduli, respectively. The low values of tensile strengths and moduli of composites containing big blue stem grass and grass flax core were because of short length of these fibers. 311 a TS (MPa) 9 TM (GPa) 80 14 A 12... E E .C U) +0 3 g 8 a £40 103 CD 6 E 2 .2 '17) 4 "a :20 C Q 0 1- 2 1- 0 0 B C D E F G Figure 6.10: Tensile properties of biocomposites Legend: A=UPE Control, B= Untreated henequen (25 vol %)-UPE, C= Untreated hemp (25 vol %)-UPE, D=Untreated kenaf (25vol %)-UPB, E= Untreated hemp-henequen (25 vol %)-UPE, F= Untreated kenaf- henequen (25 vol %), G= Untreated hemp- henequen(30 wt %)- Bioresin (20 wt %)-UPE The bending strengths and moduli of elasticity of SMC produced biocomposites containing no calcium carbonate are shown in Figure 6.11. The bars represent bending strength and the points denote modulus of elasticity. The bending strength of untreated hemp (25 vol %)-UPE biocomposite is 28 % more than strength of untreated henequen (25 vol %)-UPE biocomposite. The bending strength of untreated kenaf (25vol %)-UPE biocomposite, is 12 % more than strength of untreated henequen-UPE biocomposite. And, the bending strength of untreated hemp-henequen - UPE biocomposite is 7 % less than strength of untreated henequen (25 vol %)-UPE biocomposite. The bending strength of untreated kenaf-henequen-UPE hybrid biocomposite is 11 % lower than that of untreated kenaf-UPE biocomposite. The bending strength of untreated hemp-henequen (30 wt %)-bioresin (20 wt %) hybrid biocomposite 312 is 20 % lower than that of untreated hemp biocomposite. The bending strength of untreated henequen biocomposite is 18 % lower than that of neat polyester resin. The bending strength of untreated hemp biocomposite is 4 % higher than that of neat polyester resin. The bending strength of untreated kenaf biocomposite is 8 % lower than that of neat polyester resin. The bending strength of untreated hemp-henequen hybrid biocomposite is 23 % lower than that of neat polyester resin. The bending strength of untreated kenaf-henequen hybrid biocomposite is 27% lower than that of neat polyester resin. The bending strength of untreated hemp-henequen (30 wt %)-bioresin (20 wt %) hybrid biocomposite is 17 % lower than that of neat polyester resin. In case of modulus of elasticity, the modulus of untreated hemp (25 vol %)-UPE biocomposite is 66 % more than modulus of untreated henequen (25 vol %)-UPE biocomposite. The modulus of elasticity of untreated kenaf (25vol %)-UPE biocomposite, is 70 % more than modulus of untreated henequen-UPE biocomposite. And, the modulus of elasticity of untreated hemp-henequen -UPE biocomposite is 16 % more than modulus of untreated henequen (25 vol %)-UPE biocomposite. The modulus of untreated kenaf- henequen-UPE hybrid biocomposite is 38 % higher than that of untreated kenaf-UPE biocomposite. The modulus of untreated hemp-henequen (30 wt %)- bioresin (20 wt%) hybrid biocomposite is 24 % lower than that of untreated hemp biocomposite. The modulus of untreated henequen biocomposite is 94 % higher than that of neat polyester resin. The modulus of untreated hemp biocomposite is 222 % higher than that of neat polyester resin. The modulus of untreated kenaf biocomposite is 230 % higher than that of neat polyester resin. The strength of untreated hemp-henequen hybrid biocomposite is 125 % higher than that of neat polyester resin. The modulus of untreated kenaf-henequen 313 hybrid biocomposite is 170 % higher than that of neat polyester resin. The modulus of untreated hemp-henequen (30 wt %)-bioresin (20 wt %) hybrid biocomposite is 145 % higher than that of neat polyester resin. The highest bending strength was of the samples containing unheated hemp fibers. The untreated kenaf biocomposite and untread hemp-hennequen hybrid biocomposite had second and third highest bending strengths, respectively. The highest modulus of elasticity was of the samples containing untreated kenaf fibers. The untreated hemp biocomposite and untreated kenaf-henquen hybrid biocomposite had second and third highest moduli of elasticity, respectively. The bending strengths and moduli of elasticity followed the same trend as tensile strengths and moduli. 521 as (MPa) 9 MOE (GPa) 120 13 A T - 11 ‘3‘ g 100 r I g ‘2' T a: :-: r 9 V 5 80 — to»: T 3? O) :e:e:e: I- 0 5 5:4: 7 '77, e 50 “ It??? 2 ”e. r5 w b I I. c 40 - 9:40: o ‘6 f 1360301 F 3 m c 3"“ 2 8 20 C e _ 1 1:3 0 0 I I I I I I -1 5 A B C D E F G Figure 6.11: Flexural properties of biocomposites Legend: A=UPE Control, B= Untreated henequen (25 vol %)—UPE, C= Untreated hemp (25 vol %)-UPE, D=Untreated kenaf (25vol %)-UPE, E= Untreated hemp-henequen (25 vol %)-UPE, F= Untreated kenaf- henequen (25 vol %), G= Untreated hemp- henequen(30 wt %)- Bioresin (20 wt %)-UPE 314 Figure 6.12 shows the impact strength of composites of SMC produced biocomposites containing no calcium carbonate. The impact strength of untreated hemp (25 vol %)-UPE biocomposite is 50 % less than strength of untreated henequen (25 vol %)-UPE biocomposite. The impact strength of untreated kenaf (25vol %)-UPE biocomposite, is 60 % less than strength of untreated henequen-UPE biocomposite. And, the impact strength of untreated hemp-henequen -UPE biocomposite is 30 % less than strength of untreated henequen (25 vol %)-UPE biocomposite. The impact strength of untreated kenaf- henequen—UPE hybrid biocomposite is 9 % higher than that of untreated kenaf-UPE biocomposite. The impact strength of untreated hemp-henequen (30 wt %)-bioresin (20 wt %) hybrid biocomposite is 16 % lower than that of untreated hemp biocomposite. The impact strength of untreated henequen biocomposite is 430 % higher than that of neat polyester resin. The impact strength of untreated hemp biocomposite is 170 % higher than that of neat polyester resin. The impact strength of untreated kenaf biocomposite is 115 % higher than that of neat polyester resin. The impact strength of untreated hemp- henequen hybrid biocomposite is 270 % higher than that of neat polyester resin. The impact strength of untreated kenaf-henequen hybrid biocomposite is 135 % higher than that of neat polyester resin. The impact strength of untreated hemp-henequen (30 wt %)- bioresin (20 wt %) hybrid biocomposite is 82 % higher than that of neat polyester resin. The impact strengths of composites followed a pattern completely opposite to that of bending and tensile strengths. This is a connnon behavior for fiber reinforced plastics. The highest impact strength was of the samples containing untreated henequen fibers. This result is not surprising because, it is known that leaf fibers have high toughness and low stiffiress, while, bast fibers have low toughness and high stiffness. A hybrid bio- 315 composite of 25 wt % untreated hemp and 10 wt % untreated henequen had second J I \ ‘, " a fi‘} D v t» » ccc c b»), H ccc l)». c» h)». ccc ‘ V >>») a»): ccc c c _ >5») rs»: ccc c c ’c’c’c’ ’c’c c. H r»:» I a)»: ccc ccc >>>: : c or»: ff)" ‘. 9 T 033‘: “ ccc ccc ccc c c 0 ‘j t»); a). v v a»): b» ) ccc cccc I c ccc c c b)»; to» a 3)): b ) a ccc cccc Iccc ccc ccc c ( l:)) v): >») ’DDI b»): a r a ccc cccc Iccc ccc ccc c 4 tr») 5)» r)» a»): >)>: ) ) ccc cccc Iccc ccc ccc cccc __‘ l»); a») a») a)»: >9») )) ccc cccc Iccc ccc ccc cccc >D)’ DD) .9: r»): bra» r») — c c ccc cccc [ccc ccc ccc cccc o :1 >9)» at» a)» a»): p»): c». ccc ccc cccc hccc ccc ccc cc 4 c)»: >3») or) 9»: a)» baa» r») ccc ccc cccc iccc ccc ccc cccc a)»: l”’ a») )b) a») >))) )) 0 ccc 11; ccc ccc ccc ccc ccc Figure 6.12: Impact properties of biocomposites Legend: A=UPE Control, B= Untreated henequen (25 vol %)-UPE, C= Untreated hemp (25 vol %)-UPE, D=Untreated kenaf (25vol %)-UPE, E= Untreated hemp-henequen (25 vol %)-UPE, F= Untreated kenaf- henequen (25 vol %), G= Untreated hemp- henequen(30 wt %)- Bioresin (20 wt %)-UPE The storage modulus of composites of SMC produced biocomposites containing no calcium carbonate, are shown in Figure 6.13. The storage modulus of untreated hemp (25 vol %)-UPE biocomposite is 22 % more than storage modulus of untreated henequen (25 vol %)-UPE biocomposite. The storage modulus of untreated kenaf (25vol %)-UPE biocomposite, is 30 % more than storage modulus of untreated henequen-UPE biocomposite. The storage modulus of untreated henequen biocomposite is 50 % more than that of neat polyester resin. The storage modulus of untreated hemp biocomposite is 83 % higher than that of neat polyester resin. 316 Storage Modulus at 40 °C sawswwawswawxc. . .«wflfifififlfifiifl "v“vvvvvvvvvvw..vvvvvvvRVX .3... V V»«vwvv«V3»(«vvvvvvvvvvwx cvsvvvvv.«V«xv»@«vvvvvvvvwx . VNVNVNVNVNVNVNVNVNVNVNVNVNVNVNVNVNVNM 0000000000000000000000000000 .aaaauacxxaawxxxxxxxxx “~vaVvvvvvvvvwvwvcxx«em .H .\«««\«««««««««««««««»«««««««««««««V« «w~w~w~w~wswawswswstwswswswwswswswxwxwawt .\./A/A/A/«VNVA/A/A/AlA/A/A/«VA/A/«VA/A/A/A4. ccccc O 000000000000 000000000000 O 0 o c o c a oooooooo The data for storage modulus followed the same trend as tensile modulus and modulus of 7000 The storage modulus of untreated kenaf biocomposite is 93 % higher than that of neat polyester resin. elasticity. LL _ .i q _ _ q 0 0 0 0 0 0 0 0 0 0 O 0 0 0 0 0 0 0 0 6 5 4 3 2 1 was: «228.2 322m Untreated henequen (25vol%)-UPE, C Untreated kenaf (25%vol)-UPE Storage modulus of biocomposites at 40 0C Unsaturated polyester resin, B SMC line. They are: untreated BBSG-UPE-CaCO3, silane treated BBSG- UPE-CaCOg, Figure 6.14 shows the ESEM micrographs of four biocomposite samples made using untreated BBSG-green flax-UPE-CaCO3, silane treated BBSG-green flax-UPE-CaCO3. All the micrographs are at the magnification of 100 X and scale bar of 450 pm. The Untreated hemp (25%vol)-UPE, D Figure 6.13 Legend: A ESEM pictures were taken from the tensile fractured surfaces of these composites. In all the pictures, fiber pull out was observed. The length of the fiber pulled out became 317 shorter after the chemical treatment, as is seen in b) and d). c) Untreated BBSG-GFC-UPE-CaCO3 d) Silane treated BBSG-GFC-UPE—CaCO3 Figure 6.14: ESEM micrographs of SMC biocomposites, at magnification of 100 X and scale bar of 450 um CONCLUSIONS Biocomposites have been successfully made using natural fibers, unsaturated polyester resin, and bioresin by sheet molding compound panel processing. These biocomposites were made in the same SMC equipment, which is used to fabricate glass-polyester 318 composites. As a comparison, we also fabricated glass-polyester composites on this equipment. The biocomposites were made on this SMC line after a few minor adjustments. Instead of using the traditional fiber feeding system, we used a screw feeder and a vibratory feeder to supply natural fibers to the set-up. Consistent and repeatable results were obtained showing that this process is consistent and can be used for fabrication of bio-composites. We have also found that glass-UPE composites have almost same specific strength and modulus as natural fiber-UPE composites. But, with optimization of the entire BCSMC process, use of engineered natural fibers, and inclusion of desirable additives, we seek to achieve best mechanical, thermal and physical properties as comparable as to glass based SMC, and thus replace/substitute glass-UPE composites with natural fiber biocomposites. We aim to use biocomposites sheet molding compound panel processing (BCSMC) for fabrication of biocomposites composed of natural fibers and unsaturated polyester resin. Chopped natural fibers like, hemp, kenaf, pineapple leaf fiber, glass fibers, hybrid fibers, will be used to reinforce unsaturated polyester resin and bioresins in high speed sheet molding compound panel processing. This process would result in continuous and high volume manufacture of biocomposites. As a result, the industrial scale production of biocomposites would be possible. This will lead to accessibility of environmental goods for multiple uses in automotives, buildings as well as in fiirniture industries. The newly developed process focuses on large-scale production of biocomposites containing thermoset resins and natural fibers. The equipment used for this processing is the common industrial SMC line. Currently, natural fibers reinforced thermoset composites are not prepared using SMC [12-14]. Our processing aims to use SMC 319 process to fabricate these natural fiber-thermoset composites. In commercial SMC set-up, continuous glass fibers rovings are fed to a chopper, which cuts them to a 6 mm size, and they fall on the carrier film, forming a uniform layer of chopped glass fibers. Since, natural fibers cannot be obtained in a continuous from, and making a continuous yarn or roving with these fibers would be a difficult and expensive, we make use of chopped natural fibers in this new process. Biocomposites have been successfully made using natural fibers, unsaturated polyester resin, and bioresin by sheet molding compound panel processing. These biocomposites were made in the same SMC equipment, which is used to fabricate glass-polyester composites. As a comparison, we also fabricated glass-polyester composites on this equipment. The biocomposites were made on this SMC line after a few minor adjustments. Instead of using the traditional fiber feeding system, we used a screw feeder and a vibratory feeder to supply natural fibers to the set-up. Consistent and repeatable results were obtained showing that this process is consistent and can be used for fabrication of biocomposites. We have also found that glass-UPE composites have almost same specific strength and modulus as natural fiber-UPE composites. But, with optimization of the entire BCSMC process, use of engineered natural fibers, and inclusion of desirable additives, we seek to achieve best mechanical, thermal and physical properties as comparable as to glass based SMC, and thus replace/substitute glass-UPE composites with natural fiber biocomposites. A novel high volume processing technique named ‘biocomposite stampable sheet molding compound panel’ (BCSMCP) manufacturing process was developed so as to mimic the continuous sheet molding compound (SMC) as is currently used in making 320 glass fiber-polyester resin composites. Natural fiber-unsaturated polyester resin biocomposites were fabricated using the biocomposite stampable sheet molding compound panel’ (BCSMCP) manufacturing process. The natural fibers used for making the biocomposites using this process were: big blue stern grass, green flax core, hemp, henequen, kenaf, coir, flax and jute. For biocomposite fabrication using SMC line, 20% calcium carbonate by weight was added to the matrix as a filler. The natural fiber content was 20% by weight in biocomposites with big blue stem grass, silane treated big blue stem grass, silane treated big blue stem grass and green flax core, and, hemp and jute. Unsaturated polyester resin content in the formulations was 60% by weight. The highest fiber content for biocomposites, achievable using SMC process with UPE and CaCO3 in the matrix was 25 "/0 by weight, achieved in the case of a hybrid of untreated jute and hemp. Hybrid biocomposites were made by combining, hemp with jute mats, hemp with kenaf, hemp with henequen, big blue stern grass with green flax core, silane treated big blue stern grass with silane treated green flax core, hemp with sisal. A hybrid composite was made by combining E-glass mats with hemp. To reduce the amount of UPE in the matrix, 20% by weight of bioresin was added to a hybrid biocomposite with hemp and henequen. The bioresin was a soybean oil phosphate ester polyol modified with maleic anhydride. The mechanical properties of hybrid biocomposites with BBSG and green flax core increased after chemical treatment with 1% methacryloxypropyltrimethoxy silane (y-MPS). For the SMC samples containing calcium carbonate, the highest tensile strength was of the samples containing E—glass mat (30wt%)-Bioresin (20wt %). For the SMC samples containing calcium carbonate, the chopped E-glass composite and E-glass mat-hemp hybrid biocomposite had second and 321 third highest tensile strengths, respectively. For the SMC samples containing calcium carbonate, the highest tensile modulus was of the samples containing chopped E-glass. For the SMC samples containing calcium carbonate, the E-glass mat-hemp hybrid biocomposite and untreated jute-hemp (25 wt %) biocomposite had second and third highest tensile moduli, respectively. For the SMC samples containing calcium carbonate, the highest bending strength was of the samples containing untreated hemp fibers. For the SMC samples containing calcium carbonate, the untreated kenaf biocomposite and untreated hemp-henequen hybrid biocomposite had second and third highest bending strengths, respectively. For the SMC samples containing calcium carbonate, the highest modulus of elasticity was of the samples containing untreated kenaf fibers. For the SMC samples containing calcium carbonate, the untreated hemp biocomposite and untreated kenaf-henequen hybrid biocomposite had second and third highest moduli of elasticity, respectively. For the SMC samples containing calcium carbonate, the highest impact strength was of the samples containing of E-glass mat (30wt%)-Bioresin (20wt %). For the SMC samples containing calcium carbonate, the chopped E-glass composite, and E- glass mat-Hemp hybrid composite had second and third highest impact strength, respectively. For the SMC samples containing no calcium carbonate, the highest tensile strength was of the samples containing untreated hemp fibers. For the SMC samples containing no calcium carbonate, the untreated kenaf biocomposite and untreated hemp- henequen hybrid biocomposite had second and third highest tensile strengths, respectively. For the SMC samples containing no calcium carbonate, the highest tensile modulus was of the samples containing untreated kenaf fibers. For the SMC samples containing no calcium carbonate, the untreated hemp-henequen hybrid biocomposite and 322 untreated hemp biocomposite had second and third highest tensile moduli, respectively. For the SMC samples containing no calcium carbonate, the highest bending strength was of the samples containing untreated hemp fibers. For the SMC samples containing no calcium carbonate the untreated kenaf biocomposite and untreated hemp-henequen hybrid biocomposite had second and third highest bending strengths, respectively. For the SMC samples containing no calcium carbonate, the highest modulus of elasticity was of the samples containing untreated kenaf fibers. For the SMC samples containing no calcium carbonate, the untreated hemp biocomposite and untreated kenaf-henquen hybrid biocomposite had second and third highest moduli of elasticity, respectively. For the SMC samples containing no calcium carbonate, hybrid biocomposite of 25 wt % untreated hemp and 10 wt % untreated henequen had second highest impact strength. For the SMC samples containing no calcium carbonate, untreated henequen biocomposite had the highest impact strength. 6.2 Biobeams and bioplates from VARTM The work on bioplates and biobeams was done in collaboration with Mario Quagliata, and Dr. Rigoberto Burguefio of the Department of Civil and Environmental Engineering. More details on this part of the work can be obtained from Mario’s M.S. thesis, “Development and Characterization of Biocomposite Cellular Beams and Plates for Load-Bearing Components”, Masters Thesis, Michigan State University, East Lansing, MI; 2003. A brief description of this work is presented in the literature review. References 323 1. C. J. Davis, R. P. Wood, E. R., Miller, “Glass fiber-reinforced sheet molding compound”, US 3615979, (1971). 2. F. C. Peterson, L. P. Theard, “Thermosetting sheet molding compounds”, US 3713927, (1973). 3. A. N. Piacente, US 3835212, “Resinous sheet like products”, (1974). 4. F. C. Peterson, L. P. Theard, “Sheet molding compounds”, DE 2357000, (1974). 5. T. Kishino, “Unsaturated polyester compositions for moldings”, JP 49076988, (1974). 6. K. Horiuchi, T. Kamiya, Y. Ogawa, Y. Nakamura, M. Sugimori, “Unsaturated polyester compositions”, JP 49107086, (1974). 7. K. Takaishi, K. Masamoto, K. Tagawa, “Molding of unsaturated polyesters”, JP 50005463, (1975). 8. M. Sekiguchi, K. Nomaguchi, H. Tanaka, “Molding resin compositions”, JP 50008882, (1975) 9. Y. Takayama, Y. Ichimura, T. Aoyagi, “Sheet molding compositions”, JP 50014792, (1975) 10. M. Sato, T. Watanabe, W. Koga, “Sheet molding compositions”, JP 50036588, (1975) 11. R. Kondo, K. Nakagawa, M. Fukuda, I. Kishi, T. Ohtsuki, “Fire retardant thermosetting resin composition”, US 3931095, (1976). Kondo, Renichi; Nakagawa, Koji; Fukuda, Makoto; Kishi, Ikuji; Ohtsuki, Tateki. 12. R. P. Wool, J. Lu, S. N. Khot, “Sheet molding compound resins from plant oils”, US 2003088007, (2003). 13.B. van Voorn, H. H. G. Smit, R.J. Sinke, B.de Klerk, Composites: Part A, 32 (2001),1271-1279. 14. D. N. Goswami, P. C. Jha, K. Mahato, K. K. Kumar, Popular Plastics & Packaging, 48(3), (2003), 68-71. 324 CHAPTER SEVEN: RESULTS ON DURABILITY 7.0 Results on weathering, moisture absorption, and flammability The biocomposites were exposed to moisture absorption tests, accelerated weathering and flammability tests to assess their durability. The results of these tests are illustrated in the following sections. 7.1 Moisture Absorption The samples of bioplastics and biocomposites were subjected to controlled moisture absorption analysis in a humidity chamber. The percent weight gained by the samples was plotted against square root of time. The test was continued until the curve plateued off, and the equilibrium moisture content was achieved. Figures 7.1.1, 7.1.2, and 7.1.3 show the moisture absorption characteristics of the bioplastic samples made from grafted soybean polyol, compression molded hempmatl samples, and biocomposites manufactured using SMC, respectively. — 10% B'nresin --- 20% Bioresin — 30% Bioresin 0.1 g 0.06 . E 0.02 . a CD fi-ODZ ~ 0 B -0.06 - -0.1 T r l I O 10 20 30 0.5 40 50 Square root of time (hr) Figure7.1.l: Moisture absorption of bioresins 325 0.9 —— Untreated hempmat-UPE — AN treated hemprmt-UPE 0.7 ~ $5 8 0.5 ~ .5 a U E 0.3 ~ .29 Q) 3 0.1 7 “0.1 I I r Ti 0 10 4O 50 20 , 30 05 Square Root of Time (hrs) Figure7.1.2: Moisture absorption of biocomposites of hemp mat 0.7 — Silane treated BBSG-CaCO3-UPE —— Sihne treated BBSG & Green Flax Core-CaCOS-UPE 0.5 * Weight Gained (%) O U) 0.1 ~/ -0.1 r r r I 0 10 4O 50 20 3O 0 5 Square Root of Time (his) ' Figure7.1.3: Moisture absorption of SMC manufactured biocomposites From Figure 7.1.1, it was seen that the bioplastics samples made using 10%, 20%, and 30% GFTSOPEP3, lost weight in the beginning of the experiment until 100 hours, 326 and then started gaining weight. The equilibrium moisture content for these samples was about 0.06 % moisture. This shows very high hydrophobicity of the samples made using bioresins, because under the same conditions, UPE control would have the equilibrium moisture content of about 1 %. However, the initial weight loss of these bioplastics samples was not understood. Possible explanations for this cause were either leaching out of bioresin from the plastic samples, or presence of any unreacted bioresin in the plastic. To investigate these claims, thermal analysis of the samples was done after their initial exposure to humidity chamber. Figure 7.1.2 shows the moisture absorption characteristics of biocomposite samples made with untreated and acrylonitrile treated hempmatlina UPE matrix. The samples started gaining weight right fiom the beginning of the experiment, and continued doing so until the equilibrium was reached. This tendency was due to hydrophilic nature of the fibers due to the presence of many OH groups in the cellulosic backbone of biofibers. Water molecules get hydrogen bonded to the hydroxyl groups within the fiber cell wall of the biofibers. The equilibrium moisture content for the samples with untreated hemp matl was 0.7 %, while that for the one with acrylonitrile treated hempmatl was about 0.3 % moisture. Thus, it was observed that in addition to increment in mechanical and thermal properties, acrylonitrile treatment leads to less moisture absorption, giving rise to a more stable composite. In Figure 7.2.3 the plot of weight gain % versus time is seen for silane treated BBSG- CaCO3-UPE and silane treated BBSG & green flax core-CaCO3-UPE, both of which were processed by BCSMCP. It was seen that these biocomposites too started gaining weight right from the beginning of the experiment, and continued doing so until the 327 equilibrium was reached. This weight gain has been observed by many authors, and was due to hydrophilic nature of the fibers due to the presence of many OH groups in the cellulosic backbone of biofibers [1-9]. The equilibrium moisture content for the samples with silane treated BBSG was 0.28 %, while that for the one with silane treated BBSG & green flax core was about 0.55 % moisture. Here, the equilibrium moisture content of silane treated BBSG & green flax core was more than that of silane treated BBSG. Green flax core, which constitutes 50 % fiber and 50 % core, is more hydrophilic than grass because of the presence of powdered core. This led to higher weight increase in the hybrid biocomposite containing grass and green flax. Thermal analysis of bioplastics The bioplastics which lost weight in the initial period of moisture absorption test were thermally investigated by TGA and DSC. These samples were at room temperature before the test. Figure 7.1.4 and 7.1.5 show the plot from TGA analysis and DSC analysis of the bioplastics. Table 7.1.1 shows the maximum degradation temperature of the bioplastics examined by TGA. It was observed that on increasing the amount of bioresin in the bioplastic, the maximum degradation temperature decreased. For a bioplastic with 10% GFTSOPEP3, the maximum degradation temperature was 346 0C; for a bioplastic with'20% GFTSOPEP3, it was 341 OC; and for a bioplastic with 30% GFTSOPEP3 it was 333 0C. The initial decomposition temperature of the bioplastics was measured from Figure 7.1.4. The initial decomposition temperature clearly became lower with an increase of bioresin concentration. The reduction of the initial decomposition temperature is indicative of the existence of unreacted 328 constituents. Generally, thermoset polymers having higher cross-link density show higher maximum decomposition temperature. 120 T ___,_L_ 1.20 30% Biores'n 30% Bioresil ! 100 , "\11) 20% Bioresin T 1-00 20% B'nres'n - l A 11‘10 ' ' 1 U T lO/oBroresn + 0.80 3 .\ so a . L :7 ' 10% Bioresin j I 5 § i i 0.60 g In a: 60 I , 1 i. :5 I I 0.40 .5 .39 | a g 40 i N " l 020 .E, \ i . a 20 4 —* "‘f K \ i I 0.00 O + I I I I . I I .020 O 100 200 300 400 500 600 700 Temperature (0 C) Figure 7.1.4: TGA of bioplastics 2- _ : 2 - f T 1 § 10%Bioresin s 1 .. B a 0? 20%Bioresin 1 ’5 -1 “I ‘i - l I -2 ~ 30%B'nres'm -3- wt __..,_,,_ l l I -4 if I fl I i -70 30 130 230 330 Temperature (0 C) Figure 7.1.5: DSC of bioplastics 329 Figure 7.1.5 shows the best flow curves from -60 to 300 OC, obtained from DSC. There were no exothermic peaks in the DSC curves of any of the bioplastic samples. This indicated the absence of any unreacted bioresin in the bioplastic samples. Table 7.1.1: Maximum degradation temperature of the bioplastics Maximum decom osition Sample temperature ( C) A 346.15 B 341.23 C 332.66 Legend: A= 10% Bioresin, B=20% Bioresin, and C=30% Bioresin 7.2 Accelerated weathering The accelerated weathering of the biocomposite samples was done to study the effect of harsh weather elements on the properties of biocomposites. The exposed samples were evaluated by color test, surface roughness, weight change, and dynamical mechanical analysis. The results from these analyses are presented below. Color change +A +8 +C —x—D 0 x r I l T I I 15 30 45 6O 75 90 Time (Days) Figure 7.2.1: Change in color parameter ‘L’ over time for biocomposites 330 Legend: A=Untreated hempmatl-UPE, B=Acrylonitrile treated hempmatl-UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE +A +B +C —x—D 0 15 30 45 60 75 90 Time (Days) Figure 7.2.2: Change in color parameter ‘a’ over time for biocomposites Legend: A=Untreated hempmatl -UPE, B=Acrylonitrile treated hempmatl -UPE, C=Silane treated BBSG-CaCO3—UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE +A +8 +C —x—D 15 10* 0 15 3O 45 60 75 90 Time (Days) Figure 7.2.3: Change in color parameter ‘b’ over time for biocomposites Legend: A=Untreated hempmatl-UPE, B=Acrylonitrile treated hempmatl -UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE 331 —I—A +8 +C —x—D O x ‘1‘ I I T I l 15 3O 45 . 60 75 90 Time (Days) Figure 7.2.4: Change in color parameter ‘E’ over time for biocomposites Legend: A=Untreated hempmatl-UPE, B=Acrylonitrile treated hempmatl-UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCOB-UPE Figures 7.2.1 to 7.2.4 show the change in various color parameters of the biocomposite samples with exposure to accelerated weathering. In the CIELAB system, there are three parameters for color, L*, a*, b*. The L* axis represents the lightness, whereas, a* and b* axes are the chromaticity coordinates. While +a* is for the red, -a* is for green, +b* for yellow, -b* for blue, and L* varies from 100 (white) to zero (black). The changes in the values of L, a, b were used to calculate dE*, which is the cumulative color change. The values of dL increased from 0 at beginning of the experiment to 50 for samples exposed for 85 days. This was because the samples were changing color from greenish brown to white after being exposed to UV, rain, condensation and humidity. The combination of water, oxygen and UV irradiation promotes the change of color of these biocomposite samples. The change in L was less pronounced for samples containing BBSG and green flax core which were processed in SMC. 332 The values of da decreased from 0 for all samples. The values of db changed from 0 to positive in case of hemp matl composites, but changed from 0 to negative in case of BBSG and GFC biocomposites. The overall color change of the biocomposites, dE, increased from 0 to 50 for all biocomposites. Here again, the dE values of biocomposites with hemp matl were higher than those of samples containing BBSG and green flax core which were processed in SMC. The color change of untreated hemp matl based biocomposites was slightly higher than that of the acrylonitrile treated hemp matl based biocomposite. Weight change Figures 7.2.5 shows the plot of percentage weight gain with exposure time for biocomposite samples. All of the samples were gaining weight in the beginning of the weathering test, until about 30 days into the test. The weight gain was due to hydrophilicity of the biofibers. Weight Gain (%) -5 T I I I 0 20 40 80 100 60 Time (days) Figure 7.2.5: Change in weight over time for biocomposites 333 Legend: A=Untreated hempmatl-UPE, B=Acrylonitrile treated hempmatl -UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE Afier 30 days of exposure, the biocomposite samples started losing weight and continued doing so until the end of the exposure time. This weight loss was due to the biodegradation of the biocomposites, which began afier 30 days of exposure to weathering. However this weight loss was less than 2% for samples with untreated hempmatl and silane treated BBSG. The weight loss was less than 5% for biocomposite sample with silane treated BBSG and GFC, while, it was about 1.25% for the sample with acrylonitrile treated hemp matl. Therefore, it was established that acrylonitrile treatment of the hemp fibers makes them more stable compared to untreated hemp based ones. This could be due to improved adhesion between fibers and the matrix, and better interfacial bonding in a composite made with surface treated fibers. The values of weight loss was highest for the biocomposite containing silane treated BBSG and GF C was again because of presence of core particles which do not have the characteristics as the fibers. Surface roughness Figures 7.2.5, 7.2.6, and 7.2.7 show the comparison of surface roughness parameters of the biocomposite samples as a function of exposure time. In general, the rouglmess of the composite samples increased with exposure time. Of all surface roughness parameters, the values of Ra were smallest, followed by those of R2, and finally those of Rm. The roughness parameters of the samples with untreated hemp matl were the highest, followed by samples with silane treated BBSG and GFC, followed by sample 334 with acrylonitrile treated hemp matl. The biocomposite samples with silane treated BBSG had the lowest values of surface roughness parameters. 9A IB AC oD M 12~ ‘ A A10~ ‘1 o E 6~ ‘ A o I 41‘ . ’ I I I . 2’ . O O OI I l I I 0 20 40 6O 80 100 Tirm (days) Figure 7.2.6: Change in roughness parameter R, over time for biocomposites Legend: A=Untreated hempmatl -UPE, B=Acrylonitrile treated hempmatl-UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE « 0A IB AC CD 1% m— A A E t; 60 F . E A ‘ . N 40 ‘ I m . I .4 C 20 z : . . O l I I I I 20 40 60 80 100 Time (days) Figure 7.2.7: Change in roughness parameter Rz over time for biocomposites 335 Legend: A=Untreated hempmatl-UPE, B=Acrylonitrile treated hempmatl -UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE 0A IB AC 0D 1% 100 — A ‘ E g 80 ~ ‘ A o E 60 + I a 40 . 3 a o 0 O 20 I I 9 O T I I I T O 20 40 60 80 100 Time (days) Figure 7.2.8: Change in roughness parameter Rmax over time for biocomposites Legend: A=Untreated hempmatl -UPE, B=Acrylonitrile treated hempmatl —UPE, C=Silane treated BBSG-CaCOB-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE The samples get washed over by water, irradiated by UV, and react with ambient oxygen while in the accelerated weathering chamber. These conditions lead to breaking of the linkages between the fibers and the matrix, and the biocomposite surface stats crumbling. On increasing the exposure time of weathering, this spreads though the thickness of the composite. The reactions involved in the weather degradation are oxidation, reduction, dehydration, hydrolysis, swelling, shrinking, freezing, and cracking. The cell wall polymers responsible for the moisture sorption of biofibers are: hemicellulose, accessible cellulose, non-crystalline cellulose, lignin, and crystalline cellulose. The cell wall polymers accountable for the ultraviolet degradation are: lignin, hemicellulose, accessible cellulose non-crystalline cellulose, and crystalline 336 cellulose. While, the cell wall polymers responsible for the thermal degradation properties of biofibers are: hemicellulose, cellulose, and lignin. The strength of the biofibers is controlled by crystalline cellulose, matrix (non-crystalline cellulose + hemicellulose + lignin), and lignin. The biological degradation if cell wall of biofibers is influenced by hemicellulose, accessible cellulose, and non—crystalline cellulose. Dynamic mechanical analysis Figures 7.2.9 and 7.2.10 show the storage modules at 40 0C and the T8 of the biocomposites as a function of the exposure time of weathering, respectively, analyzed by DMA. With the physical and chemical changes occurring in the samples in the course of artificial weathering, it was no surprise to see that the values of storage modules at 40 0C and T3 of the biocomposites decreased as the weathering time increased. However, this decrease in modulus was reasonably small, as was seen by other authors earlier [10-12]. AA OB IC OD 10 E g 8 . . . O . 3 A A A 2 4 . 0 O Q Q 8’» E s 2‘ m 0 I I I I 0 20 40 60 80 100 Time(Days) Figure 7 .2.9: Change in storage modulus over time for biocomposites 337 Legend: A=Untreated hempmatl-UPE, B=Acrylonitrile treated hempmatl -UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE AA 0B IC OD 97 I :8. 95 I ‘ . an I . E- 0 ‘ A 0 I 93 « . I O 91 T I I I O 20 40 . 6O 80 100 Tune (Days) Figure 7.2.10: Change in Tg over time for biocomposites Legend: A=Untreated hempmatl-UPE, B=Acrylonitrile treated hempmatl-UPE, C=Silane treated BBSG-CaCO3-UPE, D= Silane treated BBSG & Green Flax Core- CaCO3-UPE The highest storage modulus was of the sample containing acrylonitrile treated hemp matlbased biocomposite, followed by untreated hempmatl , silane treated BBSG, and silane treated BBSG & green flax core. The same trend was observed for glass transition temperatures of biocomposites. The overall decrease in Tg for all samples was less than 3 0C for the entire exposure time. The overall decrease in modulus of surface treated hemp fibers based biocomposites was 8.6% over the entire weathering exposure time, 10% for silane treated BBSG-UPE, 10.3 % for untreated hempmatl- UPE, and 11.5% for silane treated BBSG and GFC based biocomposite. The change in storage modulus was related to change in weight of these samples. The tendency of biofibers to absorb moisture causes off-gassing (void formation) during compounding. This results in a molded article with a microstructure having 338 variable porosity and resembling that of high-density foam. The pores formed will act as stress concentration points which then lead to an early failure of the composite during loading. Another major drawback of using biofibers as reinforcing agent is the high moisture absorption of the fibers due to hydrogen bonding of water molecules to the hydroxyl groups within the fiber cell wall. This leads to a moisture build-up in the fiber cell wall (fiber swelling) and also reduction in the fiber-matrix interface bonding. This is responsible for changes in the dimensions of biofiber-based composites, particularly in the thickness and the linear expansion due to reversible and irreversible swelling of the composites. As a consequence, the fiber-matrix adhesion is weak and the dimensional stability of biofiber-based composites particularly for outdoor applications will be greatly affected. Nature builds lignocellulosic resources from carbon dioxide and water, and it has all the tools to recycle them back to the starting chemicals. Possible ways of degradation include biological, thermal, aqueous, photochemical, chemical, and mechanical means of degradation. In order to produce cellulose fiber-based composites with a long service life, the degradation processes caused by nature need to be retarded. One way of preventing or slowing down the natural degradation process is by modifying the cell wall chemistry of the material which is responsible for many of its properties. This can be accomplished by chemical modification of the fibers. 339 References 1. J. Gassan, A. K. Bledzki, Journal of Applied Polymer Science, 82, (2001), 1417— 1422. 2. J. Gassan, Composites: Part A, 33, (2002), 369-374. 3. S. Mishra, A. K. Mohanty, L. T. Drzal, M. Misra, S. Parija, S. K. Nayak, S. S. Tripathy, Composites Science and technology 63, (2003), 1377-1385. 4. J. Rout, M. Misra, S. S. Tripathy, S. K. Nayak, A. K. Mohanty, Journal of Applied Polymer Science, 84,(2002), 75-82. 5. A. Bismarck, A. K. Mohanty, I. Aranberri-Askargorta, S. Czapla, M. Misra, G. Hinrichsen, J. Springer, Green Chemistry, 3, (2001), 100-107. 6. J. Rout, M. Misra, S. S. Tripathy, S. K. Nayak, A. K. Mohanty, Composites Science and Technology, 61, (2001), 1303-1310. 7. T. H. D. Sydenstricker, S. Mochnaz, S. C. Amico, Polymer Testing, 22, (2003), 375-380. 8. H. P. S. A. Khalil, H. Ismail, M. N. Ahmad, A. Ariffin, K. Hassan, Polymer International, 50, (2001), 395-402. 9. J. Gassan, A. K. Bledzki, Polymer Composites, 20 (4), (1999), 604-611. 10. L. M. Matuana, D. P. Kamdem, and J. Zhang, “Photoageing and Stabilization of Rigid PVC/Wood-Fiber Composites,” Journal of Applied Polymer Science, 80 (11), (2001), 1943-1950. 11. L. M. Matuana and D. P. Kamdem, 5'” Pacific Rim Bio-based Composites Proceedings, 10-14 December 2000, 644-651 12. N. Stark, L. M. Matuana, Proceedings of Society of Plastics Engineers Annual Technical Conference ANT EC 2002, 2209-2213 340 CHAPTER EIGHT: CONCLUSIONS AND FUTURE WORK 8.] CONCLUSIONS Environmental and economic consciousness around the world had led to a revolution in interest in the use of renewable and sustainable materials for a variety of applications. Biocomposites, which comprise of biodegradable or synthetic polymers reinforced with natural fibers, are one example of this new class of sustainable products. The objective of this research was to develop novel low-cost biocomposite panels with desired properties by incorporation of natural fibers in an unsaturated polyester resin matrix so that these biocomposites could then be used in structural and building applications. Three interconnected parts of this project were completed in order to achieve the research objective and included: i) design of engineered natural fibers, ii) modification of the polymer matrix, and iii) development of a new process for continuous production of biocomposites. In addition, the durability of these biobased materials was investigated as well. Conclusions: 1. Engineered Biofibers The objective of this part of the work was to design the fibers responsible for the reinforcement of the composite. Engineered biofibers with optimized fiber—matrix adhesion are required for good mechanical, thermal and physical properties. The optimum fiber volume fraction was found to be 30% based on a comparison of the mechanical properties of composites with three different volume fractions of fibers. The surface chemical modifications of natural fibers like alkali treatment, vinyl grafting and treatment with various coupling agents, were some means we employed to improve 341 fiber matrix adhesion of the resulting biocomposites. Surface modification also resulted in enhancement of the aspect ratio, improved the wettability of the fibers, and formed a strong interface between polar natural fiber and non polar matrix. An acrylonitrile treatment for the hempmatl fibers produced a 80% higher increase in tensile strength over untreated hempmatl fiber based composites, a 430% increase in tensile modulus over the UPE control, and a 53% increase in impact strength over the untreated hempmatl-UPE composite. The gap between performance of glass based composites and biocomposites was bridged by fabricating a hybrid composite comprising of glass and biofibers. Hybrid biofibers were also custom made to give maximum stiffness and toughness by blending different weight fractions of different untreated or surface treated biofibers (bast, leaf, seed, hit, or grasses). In terms of specific modulus and strength, glass composites and biocomposites were in the same range. Hybridization of the composite through combining E-glass mats with hempmatl led to a 23% decrease in tensile strength, a 25% decrease in bending strength, and a 26% reduction in storage modulus at 40 OC compared to a composite of 100% E-glass fiber mat in the UPE matrix. The specific elastic modulus of a biocomposite containing acrylonitrile treated hempmatl fibers was 70% greater than that of glass-UPE composite. Hybrid biofibers can be designed to meet the desired requirements of a particular application. From the various experimental results, it can be concluded that the fiber-matrix adhesion for a natural fiber reinforced composite system can be improved by the introduction of chemical bonding at the interface. The effect of increasing the surface area and hence the interfacial friction was not as significant compared to the effect of increasing chemical 342 bonds across the fiber-matrix interface. It can be inferred that the interfacial adhesion is directly related to the amount of reactive fiinctional groups, in this particular case the hydroxyl content on the biofiber surface. When the possibility of chemical bonding is removed from the fiber matrix interface, the interfacial adhesion no longer relates to the interfacial chemistry and becomes a pure mechanical phenomenon. It is also important to note that the composite properties can also depend on various other factors such as void content and processing conditions. Good interfacial adhesion between the natural fiber and matrix polymer can only be significant when other major factors are Optimized. In summary, biocomposites were successfully made using various natural fibers with different surface treatments. An increase in mechanical and thermal properties is seen for all surface treated fiber based biocomposites most probably due to the formation of covalent bonds across the fiber-matrix interface. The UPE and alkali treatments are low cost treatments, which increase the performances of the resulting biocomposites. The specific modulus of this type of biocomposite is comparable to that of conventional glass fiber-UPE composites. Therefore, these biocomposites have the potential to replace glass- based composites. Conclusions: 11. Matrix Modification The second major portion of this project was directed at modification of the petroleum- based polymer matrix by incorporation of substantial concentrations of natural oil based resins to increase its biocontent and to improve the matrix properties. Polymerizable bioresins were developed from natural oils by chemical modification (grafting). 343 Bioresins were prepared from vegetable based oils by two methods: i) grafting vegetable oils with acrylonitrile, and ii) grafting maleic anhydride onto the natural polyols. Utilization of bioresins as polymer matrices provides a two fold benefit over thermoset resins in biocomposites. Their presence can improve the toughness of brittle unsaturated polyester resin as well as produce a material with higher biobased content. The bioresins produced can be ‘tuned’ to produce a range of bioplastics which varied from high in stiffness to high in toughness by changing the amount of bioresin introduced into the UPE matrix. The kinetic parameters of the UPE as well as UPE-bioresin curing reaction were investigated by kinetic studies. The fractional conversion of styrene double bonds was higher than those of UPE. The reaction rate of curing increases with an increase in the temperature. 100% conversion was not observed in any of the curing systems because of the entrapment of monomers in crosslinked segments of the matrix however, fractional conversion of styrene double bonds was 99% at a temperature of 160 °C, and for UPE double bonds was 92% at a temperature of 160 0C. In all bioplastics, phase separation was observed between the bioresin component and the UPE component. The impact strength of the bioplastics and biocomposites was increased by 100% by acrylonitrile grafting to the natural oils. At the same time, the bending strength reduced by 20% (by adding 20% bioresin). Alternatively, maleic anhydride was grafted onto soybean polyols. The bioplastic samples containing 40% and 50% grafted polyols had higher toughness, and lower stiffness, while bioplastic samples containing 10%, 20%, and 30% grafted polyols had a very high modulus (from 2.8 GPa for UPE control to 10 GPa for 10%GFTSOPEP1 at 40 OC) and very low impact strength. 344 Conclusions: 111. New Processing The objective of developing a new continuous biocomposites sheet molding compound panel (BCSMCP) processing was to create a process for making biocomposites that could be scaled up to commercialization levels and at the same time take advantage of the unique properties of these biocomposites. The BCSMCP method is a variation of the traditional SMC processing method and was invented to enable high volume industrial scale continuous production of biocomposites using any type of natural fiber, UPE and even bioresins. After the SMC process was optimized for natural fiber-polyester resin composites, several composites made with various untreated and surface treated chopped hybrid fibers, biofibers, calcium carbonate, UPE, and bioresin were successfully produced. As part of this research, a new fiber feeder was designed to handle biofibers and produce a uniform distribution of chopped biofibers onto the SMC line. A hybrid composite of E- glassmat-hemp was made in the BCSMCP which had a tensile strength only 20% lower than a chopped E-glass composite, a tensile modulus only 8% lower than, and a bending strength only 4% lower, the tensile modulus only 5 % lower than chopped E-glass composite, also made in the BCSMCP. The chopped E-glass composite had the tensile strength 170% higher than untreated BBSG biocomposite, and had a storage modulus 105% higher than untreated BBSG biocomposite at 40 0C (untreated BBSG biocomposite had lowest properties of all SMC composites). 345 The new BCSMCP process can easily be integrated to any existing SMC infrastructure supporting the commercialization of biocomposites by SMC. Conclusions: Durability The objective of this part of the project was to evaluate performance and longevity of bioplastic and biocomposite samples after exposure to the weather elements. The durability of bioplastics as well as biocomposites were evaluated by moisture absorption tests, accelerated weathering and flammability tests. Moisture absorption tests indicated that the equilibrium moisture content was 0.06 % for these polyester bioplastics, and was less than 0.7 % for biocomposites. The environmental exposed biocomposites changed color from brown to white as the exposure time increased. The weight of the samples first increased (up to 1% in 30 days), and then decreased (up to 5% in 90 days). Surface roughness of the samples increased with exposure time, and warpage was also observed. The maximum loss in storage modulus after moisture equilibration and subsequent drying was about 12% for a sample containing silane treated BBSG-green flax core hybrid. Surface treated fiber based composites performed better in terms of these characteristics compared to untreated fiber reinforce composites. These tests suggest that biocomposites can be used for interior applications for prolonged periods of time without excessive degradation due to environmental conditions. General Conclusions 346 Biocomposites were made using various types of natural fibers with different surface treatments by use of different processing techniques. Hybrid composites were made by combining different varieties of biofibers in different weight fractions, and also by combining biofibers and E-glass fibers. Biocomposites could be produced within 30-80% of the mechanical and thermal properties of glass composites in most cases. Parity in properties could be achieved by hybridization of the biofiber and glass fibers based composites. The impact strength and modulus could be improved by adding different types of bioresins into the UPE matrix resulting in a phase separated material. Incorporating certain bioresins in the matrix increased the biomass content of composites up to 60%. Continuous production of biocomposites was demonstrated by utilization of the SMC process. Engineered fibers were designed to compete with glass fiber based SMC composites. The mechanical properties of biocomposites do not change significantly on accelerated weathering. Through the integration of all parts of this project, we were able to develop new cost effective biocomposites for future applications in structural and building industries. Our ultimate goal of replacing existing glass fiber-polyester composite panels currently used in the housing industry by biocomposites has been demonstrated. Biocomposites have the potential to replace glass fiber-UPE composites. 8.2 FUTURE WORK 347 We have been successful in making structural biocomposites for future applications, however, some more additional research needs to be completed before this technology is placed in widespread use. The adhesion between fibers and the matrix must be improved by surface treatment. Alternative surface treatments such as acetylation, bleaching, UV/plasma, and microwave should be investigated and optimized for each type of natural fiber in the future. Bioresins having different chemistries can be developed from natural oils. These bioresins then must be polymerized by blending with synthetic resins or by themselves to increase the amount of ‘green’ material in a thermoset composite. Further experiments need to be aimed at studying in detail the behavior of the second phase, the actual mechanism of toughening and control of the dispersed phase in the bioresin-UPE systems. The fracture studies must be performed in further detail to find the critical parameters like fiber-matrix adhesion and fracture dynamics. Alternative strategies to improve the toughness of brittle UPE resins by adding rubbery particulates also should be investigated. The curing times of the UPE or bioresin-UPE systems also must be optimized to get maximum conversion in shortest amounts of time. There should be some solutions to the volume shrinkage problems of the UPE and bioresin-UPE systems. The amount of renewable and biodegradable materials in the biocomposites and bioplastics must be increased to at least about 60% of the entire weight of the product. One option to consider would be the incorporation of a small volume fiaction of economic nanoparticles into the biofiber-bioresin-UPE system, to make a bio-nano-composite, with the added advantages of increased stiffness, strength, and heat resistance; and decreased moisture absorption, flammability, and permeability. 348 An effort must be made to study in detail the various parameters that affect the properties of composites in hybrid fibers systems. A universal model needs to be developed which allows for inclusion of natural fiber properties to estimate the theoretical values of the biocomposite properties. A cost analysis for the biocomposite sheet molding compound panel processing should be performed, and compared with conventional SMC technique processing glass-UPE composites. Finally, collaboration with our industrial partners, F laxCraft Inc. and Kemlite Chemical Company Inc. needs to be enhanced to promote and commercialize this technology in housing panel applications. 349 EEEEEEEEEEEEEEEEEEEEEEEEEEE IIIIIIIIIIIIIIIIIIIII ”J_-_-_-_-_-_-_-_-_-_-j_-,