. .31! in. I . .viza . .31.: v .9422, $3. “tit Evin...“ !. “5"”; Huang... . . n. I! .I. 3... h- ...x.‘ 3 $3. h.........h..... . ) a..z.......w.u. .m V .n...:.v.bv..§¥3nu......auui= . .. . .1 .9 arr. . ?.¥# Erma“... 2...... r. 35553.... .. . .. ... . n f a V “but! .351}... Air... , -2. . . 41.5". in..- was... L. , i... a...:....,.: . .3. 3‘5)... unto: .b)!.¢.Vl.-.I.»¥. .‘u .v I . ab”. w. .5 3.3.... .z ‘01. x .2 ii... .1. 1...! 3.... 3.. ‘35.... tilts: )1... $34 .dtt‘iv - i. :., Id.) $1.. .AO. w I) 5.. ‘1... i=3... ...5.(‘. 0.1.... .i. {thawifinr‘ run}?! .3. . a 3.... . :3 Eli! .‘Izzk. a:.(...a.zia a}... 3.551. 15:. a; .sztxx... .. {.15. . . va\..v...u..g:t.......: in»... :31! . Ink $1.?tntw i! 235:4 a. )1? 2:1 I in... I} 13.3.}... 1 s .334 tr! A 2. loo/V This is to certify that the dissertation entitled 1 Antimicrobial Whey Protein Isolate-Based Edible Casings presented by Arzu Cagri has been accepted towards fulfillment of the requirements for Ph.D degree in Food Science Major professor ; Date 5/3/02 MSU is an Affirmative Action/Equal Opportunity Institution 0- 12771 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 AUG 0 6 20M 060305 Pg“ 1 L0 6 1 ZUU/ MG 2 1 mi 6/01 c:/CIRC/DateDue.p65—p.15 ANTIMICROBIAL WHEY PROTEIN ISOLATE-BASED EDIBLE CASINGS By ARZU CAGRI A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements For the degree of DOCTOR OF PHILOSOPHY Department of Food Science and Human Nutrition 2002 ABSTRACT ANTIMICROBIAL WHEY PROTEIN ISOLATE-BASED EDIBLE CASINGS By ARZU CAGRI Methods were developed to optimally produce low pH (5.2) antimicrobial whey protein isolate (WPI) edible films containing sorbic acid (SA) or p-aminobenzoic acid (PABA). Solutions of lactic acid and acetic acid (1:0, 1:1, and 23:1) were used to acidify the film solution. Films containing either antimicrobial agent inhibited the growth of Listeria monocytogenes, Escherichia coli 0157:H7 and Salmonella Typhimurium DT104 on trypticase soy agar. Increased concentrations of PABA or SA in WPI film increased the percent elongation (%E) and water vapor permeability (WVP). While tensile strength (TS) of WPI films decreased with increasing levels of SA, TS of films was not affected addition of PABA. Antimicrobial properties of films containing PABA, SA, or PABA:SA were tested over 21 days of refrigerated storage on slices of summer sausage and bologna inoculated to contain 106 L. monocytogenes, E. coli 0157:H7 or S. Typhimurium DT104 CFU/ g. These films decreased populations of the test pathogens and retarded the growth of mesophilic aerobic bacteria, lactic acid bacteria and mold/yeast throughout this period of storage. Contact with the slices decreased TS of the films, but increased % E. In the last objective, heat-cured WPI films containing PABA were heat-sealed to form casings. A commercial-type hot dog batter was stuffed into WPI, collagen, or natural casings. After cooking, the hot dogs were surface-inoculated with Listeria monocytogenes, vacuum packaged, and examined for numbers of L. monocytogenes, mesophilic aerobic bacteria (MAB), lactic acid bacteria (LAB), and yeast/mold during 42 days of storage at 4°C. Listeria populations on hot dogs prepared with WPI-PABA casings remained relatively unchanged; however, numbers of Listeria on hot dogs prepared with WPI, collagen, and natural casings increased during 42 days of refrigerated storage. Populations of MAB, LAB, and mold on WPI-PABA casings were lower compared to other casings. Physical (e.g. purge loss, color), mechanical (shear force), chemical (e. g. moisture, fat, protein content, pH, TBA) and sensory properties of hot dogs prepared with different casings were also analyzed. Taste panelists ranked flavor, juiciness, and overall acceptability higher for hot dogs prepared with WPI- PABA casings compared to hot dogs prepared with WPI, collagen or natural casings. Hence, WPI casings containing PABA appear to be a promising means to prevent growth of Listeria on hot dogs and extend shelf-life. To my parents ACKNOWLEDGMENTS This dissertation has been made by the support and help from many people. Foremost among them is my advisor Dr. Elliot T. Ryser. I would like to express my sincerest and most heartfelt thanks to him. His thoughtful guidance and advice helped me grow educationally and professionally. I cannot thank him enough for the all the support he has provided me for the successful completion of this research project. I would like to express my sincere appreciation to all my committee members, Dr. James Pastka, Dr. Jack Giacin, Dr. Susan E. Selke, Dr. Wesley N. Osburn, Dr. Zeynep Ustunol, for their guidance and support. I am grateful to Dr. Zeynep Ustunol for all support and advice throughout my graduation program. I would also like to thank Dr Seong-Joo Kim for her advice in film formation technique. I owe a great deal of gratitude to Dr. Wesley N. Osburn and his research group, Dr. Jamie Willard, Christine Ebeling, Deanna Bloom and Jeff Sindelar for their great input and collaboration in this project. They helped me in hot dog processing and other analysis. They also allowed me to use of the lab equipment, time, and expertise in finishing this project. Without their help the study could not have been completed. I would like to thank Dr. Bruce Harte for allowing me to use the labs and equipment in the Department of Packaging. I also would like to thank teaching assistants in Dept. of Packaging, Dr. Rujida Leepipattanawit and Krittika Tanprasert for all their help in this project. I sincerely appreciate Dr. Janice Harte for her advice on setting up the sensory tests. I would like to thank Turkish Government, Center for Food and Pharmaceutical Packaging Research and USDA FSIS for their financial support for this research. New Zealand Milk Product, Inc. and Strahl and Pitch, Inc. are also acknowledged for providing whey proteins and candellila wax, respectively. I would like to acknowledge all the past and present colleagues in my laboratory for their continues encouragement and friendship. I would also like to thank Can Mavruk for all his support and help with data analysis. My deepest appreciation goes to my family and friends who supported me while accomplishing this work. vi TABLE OF CONTENTS LIST OF TABLES ......................................................................... xii LIST OF FIGURES ........................................................................ xv ABBREVIATIONS .......................................................................... xvii INTRODUCTION ........................................................................... 1 CHAPTER 1 LITERATURE REVIEW .................................................................... 6 1.1 . Edible Films .................................................................... 6 1.1.1 Definition and historical background ........................... 6 1.1.2 Formation of edible films ......................................... 7 1.1.2.1 Components of edible films .............................. 7 1.1.2.2 Film formation techniques .............................. 8 1.1.3 Protein-based edible films ......................................... 9 1.1.3.1 Whey protein .............................................. 9 1.1.3.1.] Composition and structure ................. 9 1.1.3.1.2 Film formation ............................... 10 1.1.3.2 Corn zein .................................................. 11 1.1.3.2.] Composition and structure ................. 11 1.1.3.2.2 Film formation ............................... 13 1.1.3.3 Casein ..................................................... 16 1.1.3.3.] Composition and structure ................. 16 1.1.3.3.2 Film formation .............................. 17 1.1.3.3.3 Film properties ............................... 18 1.1.3.4 Wheat gluten .............................................. 19 1.1.3.4.1 Composition and structure ................. 19 1.1.3.4.2 Film formation ............................... 19 1.1.3.4.3 Film properties ............................... 20 1.1.3.5 Soy protein ................................................ 22 1.1.3.5.] Composition and structure ................. 22 1.1.3.5.2 Film formation ............................... 22 1.1.3.6 Collagen ................................................... 25 1.1.3.6.] Composition and structure ................. 25 1.1.3.6.2 Film formation ............................... 25 1.1.3.7 Gelatin ..................................................... 27 1.1.3.7.1 Composition and structure ................. 27 1.1.3.7.2 Film formation ............................... 28 1.1.4 Polysaccharide-based edible films ............................... 28 vii 1.1.4.1 Cellulose film .............................................. 28 1.1.4.2 Alginate .................................................. 31 1.1.4.2.] Film formation ............................... 31 1.1.4.3 Pectin ...................................................... 33 1.1.4.4 Chitosan ................................................. 34 1.1.4.4.] Film formation ................................ 34 1.1.4.5 Starch films ............................................... 35 1.1.4.6 Dextrins ................................................... 37 1.1.5 Lipid films .......................................................... 37 1.1.5.1 Wax films .................................................. 38 1.1.6 Evaluating of the properties of edible films .................... 40 1.1.6.1 Barrier properties ........................................ 4O 1.1.6.1.1 Water vapor permeability .................. 41 1.1.6.1.2 Oxygen permeability ........................ 42 1.1.6.2 Mechanical properties ................................... 43 1.1.7 Factors affecting film properties ................................. 43 1.1.8 Application of edible films ....................................... 44 1.2 Antimicrobials .................................................................... 50 1.2.1 Benzoic acid ........................................................ 51 1.2.2 Sorbic acid .......................................................... 52 1.2.3 Propionic acid ...................................................... 60 1.2.4 Mode of action of benzoate, sorbate, and propionic acid. 62 1.2.5 Parabens ............................................................. 63 1.2.6 Fatty acids ........................................................... 65 1.2.7 Organic acids ....................................................... 66 1.2.7.1 Acetic acid ................................................ 66 1.2.7.2 Lactic acid ................................................. 69 1.2.7.3 Mode of action of organic acids ...................... 70 1.2.8 Bacteriocins ......................................................... 70 1.2.8.1 Nisin ........................................................ 71 1.2.8.2 Pediocin .................................................... 74 1.2.8.3 Mode of action of bacteriocins ........................ 76 1.2.9 Natural antimicrobials ............................................. 77 1.2.9.1 Lysozyme .................................................. 77 1.2.9.1.] Mode of action .............................. 78 1.2.9.2 Spices, Herbs, and Essential oils ...................... 79 1.2.9.3 Lactoferrin ................................................. 81 1.2.9.4 Liquid smoke ............................................. 83 1.2.10 Cure agents ......................................................... 84 1.2.9.1. NaCl ....................................................... 84 1.2.9.2. Nitrite ...................................................... 85 1.3. Antimicrobial Edible Films ................................................ 87 viii CHAPTER 2 ANTIMICROBIAL, MECHANICAL, AND MOISTURE BARRIER PROPERTIES OF LOW pH WHEY PROTEIN-BASED EDIBLE FILMS CONTAINING p-AMINOBENZOIC OR SORBIC ACIDS ............................................. 95 2.1 . Abstract ................................................................................... 96 2.2. Introduction ............................................................................... 96 2.3. Materials and methods .................................................................. 99 2.3.1. Film preparation .............................................................. 99 2.3.2. Bacterial strain ................................................................ 99 2.3.3. Diffusion-type assay ......................................................... 100 2.3.4. Film thickness ................................................................ 100 2.3.5. Mechanical properties ....................................................... 101 2.3.6. Water vapor permeability ................................................... 101 2.3.7. Statistical analysis ............................................................ 102 2.4. Results and discussion .................................................................. 102 2.4.1 . Antimicrobial properties .................................................... 102 2.4.2. Mechanical properties ....................................................... 108 2.4.3. Water vapor permeability ................................................... 1 12 2.5. Conclusion ................................................................................ 1 14 CHAPTER 3 INHIBITION OF LISTERIA MONOC YT OGENES, SALMONELLA TYPHIMURIUM DT104 AND ESCHERICHIA COLI 0157:H7 ON BOLOGNA AND SUMMER SAUSAGE SLICES USING WHEY PROTEIN ISOLATE BASED EDIBLE FILMS CONTAINING ANTIMICROBIALS ................................................... 1 15 3.1. Abstract .......................................................................... 116 3 .2. Introduction ...................................................................... 1 16 3.3. Materials and methods ......................................................... 119 3.3.1. Products .............................................................. 119 3.3.2. Film preparation ..................................................... 119 3.3.3. Bacterial strains ...................................................... 120 3.3.4. Product Inoculation and Storage ................................... 120 3.3.5. Microbiological analysis ............................................ 121 3 .3 .6. Mechanical properties ............................................... 122 3.3.7. Statistical analysis ................................................... 123 3.4. Results and discussion ......................................................... 123 3.4.1 . Antimicrobial properties ............................................ 123 3.4.2. Mechanical properties .............................................. 137 3.5. Conclusion ...................................................................... 140 CHAPTER 4 INHIBITION OF LISTERIA MONOC YT OGENES ON HOT DOGS USING ANTIMICROBIAL WHEY PROTEIN-BASED EDIBLE CASINGS ................ 141 4.1 . Abstract .......................................................................... 142 4.2. Introduction ...................................................................... 143 4.3. Materials and methods ......................................................... 144 4.3.1 . Experimental Design ................................................ 144 4.3.2. Casing preparation .................................................. 145 4.3.3. Culture preparation .................................................. 146 4.3.4. Hot dog processing .................................................. 149 4.3.5 Chemical analysis .................................................... 149 4.3.6. Rancidity measurement ............................................. 149 4.3.7. pH Test ................................................................ 150 4.3.8. Diffusion Test ........................................................ 150 4.3.9. Microbiological analysis ........................................... 152 4.3.10. Film thickness ...................................................... 152 4.3.1 1. Mechanical properties ............................................. 152 4.3.12. Purge ................................................................. 153 4.3.13. Shear test ............................................................ 153 4.3.14. Color measurement ................................................ 154 4.3.15. Sensory evaluation ................................................. 154 4.3.16. Statistical Analysis ................................................. 155 4.4. Results and Discussion ........................................................ 155 4.4.1 Chemical analysis ................................................... 155 4.4.2 Thiobarbutiric acid values .......................................... 155 4.4.3. pH ...................................................................... 157 4.4.4. Diffusion Coefficient ............................................... 157 4.4.5. Antimicrobial Analysis ........................................... 160 4.4.6. Mechanical properties ............................................. 169 4.4.7 Purge loss and Cook Yield ......................................... 170 4.4.8 Shear Force ........................................................... 170 4.4.9 Color .................................................................. 173 4.4.10. Sensory Analysis ................................................. 177 CONCLUSIONS ............................................................................. 181 APPENDIX I Moisture Analysis .......................................................... 184 APPENDIX II Fat Analysis ................................................................. 186 APPENDIX III Protein Analysis ............................................................ 188 APPENDIX IV TBA measurement ......................................................... 191 APPENDIX V Sensory Questionnaire ................................................... 194 APPENDIX VI Written consent form .................................................... 199 BIBLIOGRAPHY ............................................................................ 202 xi LIST OF TABLES Table 1.1 Water vapor permeability and oxygen permeability of edible films ........... 12 Table 1.2 Tensile strength and elongation of edible films .................................... 14 Table 1.3 Antimicrobials ......................................................................... 54 Table 2.1 Antimicrobial activities of whey protein based edible films containing p-aminobenzoic acid (PABA) against 4 strains of L. monocytogenes ........ 103 Table 2.2 Antimicrobial activities of whey protein based edible films containing sorbic acid (SA) against 4 strains of L. monocytogenes. ....................... 104 Table 2.3 Antimicrobial activities of whey protein based edible films containing p-aminobenzoic acid (PABA) against 3 strains of E. coli 0157:H7 .......... 106 Table 2.4 Antimicrobial activities of whey protein based edible films containing sorbic acid (SA) against 3 strains of E. coli 01572H7 ........................... 107 Table 2.5 Antimicrobial activities of whey protein based edible films containing p-aminobenzoic acid (PABA) against 5 strains of S. Typhimurium DT104.. 109 Table 2.6 Antimicrobial activities of whey protein based edible films containing sorbic acid (SA) against 5 strains of S. Typhimurium DT104 ................... 110 Table 2.5 Thickness, Tensile strength (TS), Percent elongation (%E), and Water Vapor Permeability (WVP) of whey protein isolate-based films containing sorbic acid (SA) and p-aminobenzoic acid (PABA) ............................. 113 Table 3.1 Population decrease (log CFU/ g) of L. monocytogenes, E. coli 01572H7, and S. Typhimurium DT104 with WPI films containing p-aminobenzoic acid (PABA), sorbic acid (SA), or SAzPABA on bologna (B) and fermented summer sausage (SS) slices after 21 days of refrigerated storage ............. 132 Table 3.2 Inhibition of mesophilic aerobic bacteria (MAB) and lactic acid (LAB) on bologna slices using whey protein isolate edible films containing p-aminobenzoic acid (PABA) or sorbic acid (SA) at 4°C ...................... 133 Table 3.3 Inhibition of mold and yeast on bologna slices using whey protein isolate edible films containing p-aminobenzoic acid (PABA) or sorbic acid (SA) at 4°C ................................................................................... 134 Table 3.4 Inhibition of mesophilic aerobic bacteria (MAB) and lactic acid (LAB) on fermented summer sausage slices using whey protein isolate edible films xii containing p-aminobenzoic acid (PABA) or sorbic acid (SA) at 4°C. . . . 1 35 Table 3.5 Inhibition of mold and yeast on fermented summer sausage slices using whey protein isolate edible films containing p-aminobenzoic acid (PABA) or sorbic acid (SA) at 4°C ............................................................ 136 Table 3.6 Thickness, percent elongation (% E) and tensile strength (TS) of whey protein isolate edible films containing p-aminobenzoic acid (PABA) or sorbic acid (SA) while in contact with bologna slices at 4°C ............... 138 Table 3.7 Thickness, percent elongation (% E) and tensile strength (TS) of whey protein isolate edible films containing p-aminobenzoic acid (PABA) or sorbic acid (SA) while in contact with fermented summer sausage slices at 4°C ........................................................................... 139 Table 4.1 Hot Dog Smokehouse schedule .................................................... 148 Table 4.2 Moisture, fat and protein content of hot dogs prepared with WPI-p- aminobenzoic acid (PABA), WPI, collagen and natural casings .............. 156 Table 4.3 TBA values and pH of hot dogs prepared with WPI-p-aminobenzoic acid (PABA), WPI, collagen, and natural casings during 42 days of refrigerated storage ................................................................... 159 Table 4.4 Analysis of Variance on the effect of easing type on microbiological, chemical, and physical characteristics of hot dogs during 42 days of refiigerated storage (F-values for independent variables and interactions)... 160 Table 4.5 Tensile strength (TS) and percent elongation (% E) of WPI-p- aminobenzoic acid (PABA), WPI, collagen, and natural casings before and after cooking and smoking of hot dogs ..................................... Table 4.6 Purge lossesof hot dogs prepared with WPI-p-aminobenzoic acid (PABA), WPI, collagen, and natural casings during 42 days of refrigerated storage ................................................................... 172 Table 4.7 Interior color values of hot dogs prepared with WPI-p-aminobenzoic acid (PABA), WPI, collagen and natural casings during 42 days of refiigerated storage .................................................................................. 174 Table 4.8 Exterior color values of hot dogs prepared with WPI-p-aminobenzoic acid (PABA), WPI, collagen and natural casings during 42 days of refrigerated storage ................................................................................. 178 Table 4.9 Sensory attributes and chemical content of hot dogs prepared with WPI- p- aminobenzoic acid (PABA), WPI, collagen, and natural casings after 35 xiii days of refrigerated storage .......................................................... 179 xiv Figure 3.1 Figure 3.2 Figure 3.3 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Figure 4.6 LIST OF FIGURES Inhibition of L. monocytogenes on bologna (A) and fermented sausage (B) slices with WPI edible films containing 0.75%, and 1.0% p-aminobenzoic acid (PABA); 0.75% and 1.0% sorbic acid (SA) and 0.5% SA: 0.5%PABA (1 :1). Control + film (WPI edible film without SA or PABA) and control (no film used) ......................... 122 Inhibition of E. coli 01572H7 on bologna (A) and fermented sausage (B) slices with WPI edible films containing 0.75%, and 1.0% p-aminobenzoic acid (PABA); 0.75% and 1.0% sorbic acid (SA) and 0.5% SA: 0.5%PABA (1 :1). Control + film (WPI edible film without SA or PABA) and control (no film used). . . ....123 Inhibition of S. Typhimurium DT104 on bologna (A) and fermented sausage (B) slices with WPI edible films containing 0.75%, and 1.0% p-aminobenzoic acid (PABA); 0.75% and 1.0% sorbic acid (SA) and 0.5% SA: 0.5%PABA (1:1). Control + film (WPI edible film without SA or PABA) and control (no film used) ................ 124 Preliminary study. Growth of L. monocytogenes on hot dogs prepared with WPI- 1% PABA, WPI- 1% SA, WPI- 0.5% SA: 0.5% PABA, WPI (antimicrobial-free), and Collagen casings ................................. 163 Growth of Listeria monocytogenes on hot dogs prepared with WPI- 1% PABA , WPI, Collagen, and Natural casings .................................... 166 Growth of mesophilic aerobic bacteria on hot dogs prepared with WPI- 1% PABA , WPI, Collagen, and Natural casings .................................... 167 Growth of lactic acid bacteria on hot dogs prepared with WPI- 1% PABA, WPI, Collagen, and Natural casings ............................................... 168 Growth of mold on hot dogs prepared with WPI- 1% PABA , WPI, Collagen, and Natural casings ...................................................... 169 Cook Yield (A) and Shear force (B) of hot dogs prepared with WPI- p-aminobenzoic acid (PABA), WPI, collagen, and natural casings....l73 XV ASTM CDC CMC Coll % E FDA Gly GRAS HPC HPMC LA LAB MAB MC OP PABA PEG PW SA SAS SDS Sol SPI SS TBA TS ABBREVIATIONS Acetic acid American Society for Testing and Materials Bologna slices Centers for Disease Control and Prevention Carboxymetylcellulose Collagen Percent elongation Food and Drug Administration Glycerol Generaly recognized as safe Hydroxypropyl cellulose Hydropropionate methylcellulose Lactic acid Lactic acid bacteria Mesophilic aerobic bacteria Methylcellulose Oxygen permeability p-aminobenzoic acid Polyethylene glycol Peptone water Relative humidity Sorbic acid Statistical Analysis System Sodium dodecyl sulfate Sorbate Soy protein isolate Summer sausage 2- Thiobarbutiric acid Tensile strength xvi TSA TSAYE TSB TS WG WPI Trypticase soy agar Trypticase soy agar + yeast extract Trypticase soy broth Trisodium phosphate Wheat gluten Whey protein isolate Water vapor permeability xvii INTRODUCTION Microbial susceptibility of the food surface is a major determinant of product quality and safety during refiigerated storage and distribution since most Class 1 product recalls in the US. result from post-processing contamination during subsequent handling and packaging. The increasing rate at which ready-to-eat meat products are becoming contaminated with Listeria monocytogenes after processing has raised new safety concerns. From April 1998 to December 2001 over 75 Listeria-related Class I recalls involving more than 100 million pounds of cooked ready-to-eat meats were issued (USDA-FSIS, 2001a). In 1998, more than 35 million pounds of hot dogs and luncheon meats contaminated with L. monocytogenes were voluntarily recalled by one Michigan manufacturer in response to a listeriosis outbreak that resulted in 101 cases in 22 states, including 21 fatalities (CDC, 1998). Another outbreak involving 29 cases in 10 states (including 7 fatalities) prompted the recall of approximately 14.5 million pounds of turkey and chicken delicatessen meat with the product again becoming contaminated with L. monocytogenes after processing (CDC, 2000). In 2000, almost 2,300 cases of food- borne listeriosis were reported in the United States at an estimated cost of $2.33 billion (~$ 1 million /case), making L. monocytogenes the second most costliest food-borne pathogen known to date after Salmonella (S 2.38 billion) (U SDA-FSIS 2001b). Two other pathogens have emerged as major public health concerns. Escherichia coli 0157:H7 has been highly publicized due to outbreaks of illness associated with ground beef (Ostroff et al., 1990, Bell et al., 1994, Mead et al., 1997, Tuttle et al., 1999) and dry-cured salami (Tilden et al., 1996). Unlike most food-borne pathogens, E. coli 0157:H7 shows high tolerance to acidic environments. Transfer of E. coli 0157:H7 from contaminated meat or utensils to other foods such as fresh fruits and vegetables is also a major concern (CDC, 1995). The annual cost of the 62,458 E. coli 0157:H7 infections reported in 2000 was estimated at $659.1 million (~$10,500/case) (U SDA-F SIS, 2001b). Salmonella is the second most frequently reported cause of foodbome illness in the United States after Campylobacter. An estimated 1.34 million cases of salmonellosis occur annually, resulting in 600 deaths. The annual economic cost of foodbome salmonellosis was estimated at 2.38 billion dollars (U SDA-FSIS 2001b). The incidence of salmonellosis appears to be rising both in the US. and other industrialized nations. S. enteritidis isolations from humans increased dramatically during the early to mid 1990’s particularly in the northeast United States (6-fold or more). Salmonella is most commonly found in meat, poultry, eggs and water, with kitchen surfaces also a common breeding ground for Salmonella. Salmonella Typhimurium DT104, a multiantibiotic resistant strain, is also emerging as a serious threat to public health (Glynn et al., 1998). Food-bome transmission of S. Typhimurium DT104 has been documented for several meat-related outbreaks; suspected vehicles included roast beef, ham, pork sausage, salami sticks, cooked meats, frozen sausage samples, and chicken legs (Davies et al., 1996; Anonymous, 1996). In England, 17% of 786 fresh or frozen sausage samples yielded Salmonella spp., including S. Typhimurium DT104 (Nichols and de Louvois, 1995). Sharma et al. (2001) also recovered S. Typhimurium DT104 at several points in pork production. These findings indicate that meat products may pose a serious health risk if not cooked and handled properly. Post-processing pasteurization is one means for minimizing post-processing contaminants on meat products. Using this strategy, packaged products are individually pasteurized by heat or other means (e. g. high pressure, UV) to inactivate surface bacteria. However, some of these treatments makes the products organolepticaly unacceptable to consumers (Farber and Peterkin, 1999). Alternatively, various antimicrobial dips and sprays have been applied to ready-to-eat foods to minimize microbial growth (Robach and Sofos, 1982; Cunningham, 1979). However, their effectiveness over time is limited because the preservatives diffuse into the food, allowing surface organisms to grow. Therefore, it is important to prevent or control migration of these preservatives. Guilbert and his research group have assessed the ability of edible films to control the diffusion of sorbic acid and potassium sorbate (Guilbert, 1988; Giannakopoulus and Guilbert, 1986; Guilbert et al., 1985). In another study, lactic acid-treated casein films containing sorbic acid (Guilbert, 1988) retained 30% of the sorbic acid at the surface after 30 days of storage at 95% RH, whereas all of the sorbic acid with no casein film application on control samples diffused after 24 hours of storage. They confirmed that the edible film matrix entrapped antimicrobials or other food additives and reduced diffusion dming storage. Research on edible films as packaging materials has increased because of two potential that these films may improve overall food quality, extend shelf life, and improve economic efficiency of packaging materials. In addition, environmental concerns regarding disposal of fluid whey and increased production of non-biodegradable packaging materials have prompted interest in the development of edible films from whey protein isolate and whey protein concentrate as well as wheat gluten, soy protein isolate, zein protein and others. Previously, coating foods with edible films provided only barrier and protective functions. However, incorporating various substances into packaging material can improve its fimctionality. In addition to acting as a barrier against mass diffusion (moisture, gases, and volatiles), edible films can also be used as carriers for a wide range of food additives such as antimicrobial agents, flavoring agents, antioxidants, and colorants. Incorporating antimicrobial compounds such as nisin, pediocin, benzoic acid, sorbic acid/propionic acid, or lysozyme into edible films or coatings is another means of enhancing the safety and shelf-life of products subject to surface contaminants such as ready-to-eat meats and cheeses. For example, soy protein and corn zein films containing pediocin or nisin and calcium alginate coatings containing organic acids were shown to inhibit the growth of pathogenic (e. g. L. monocytogenes, S. Typhimurium and E. coli 0157:H7) and spoilage bacteria (Lactobacillus plantarum) on meat surfaces (Sirugusa and Dickson, 1993; Dawson et al., 1997; Padget et al., 1998). In addition, coating with an antimicrobial whey protein solution containing propionic/sorbic acid prevented growth of L. monocytogenes on frankfiirters (McDade et al., 1999). However, non-uniformity of the coating on frankfuiters after dipping, draining and drying would likely produce a less effective antimicrobial barrier as compared to pre-cast films. Whey protein isolate (WPI) edible films containing antimicrobials would be uniform in thickness. Consequently, these films would be better suited to inhibit post- processing surface contaminants such as L. monocytogenes. The underlying hypothesis for this research was that low pH whey protein-based edible films (pH 5.2) containing sorbic acid (SA) or p-aminobenzoic acid (PABA) could be developed to inhibit L. monocytogens, E. coli 0157:H7 and S. Typhimurium DT104 both in a model system and on hot dogs when used as heat sealed tubular casings and extend product shelf-life. The specific objectives of this research were as follows: (1) (2) (3) (4) (5) Develop a series of low pH whey protein isolate (WPI) edible films (pH 5.2) containing p-aminobenzoic acid or sorbic acid and determine water vapor permeability and mechanical (tensile strength and elongation) properties. Test the antimicrobial properties of these films against L. monocytogenes, E. coli 0157:H7 and S. Typhimurium DT104 on a laboratory medium. Compare the barrier and mechanical properties of these films to commercial collagen and cellulose films and modify the mechanical and barrier properties if needed using heat curing. Assess the ability of these films to retain their desirable antimicrobial and mechanical properties while in direct contact with slices of bologna and summer sausage during 21 days of refrigerated storage. Develope WPI hot dog casings containing PABA, SA, and PABA:SA by heat-sealing the aforementioned films into a tubular shape in order to: a) Test the ability of WPI casings containing PABA, SA or PABA:SA to inhibit L. monocytogenes on surface-inoculated hot dogs. b) Determine the ability of the WPI casings to retain their desirable mechanical properties during hot dog manufacture. c) Assess the impact of different casings on storage stability of hot dogs based on numbers of total aerobic mesophilic bacteria, lactic acid bacteria, yeast and mold, 2-thiobarbutiric acid (TBA) values, purge loss, shear force, color, pH, chemical analysis and sensory evaluation. CHAPTER 1 LITERATURE REVIEW 1.1. EDIBLE FILM 1.1.1. Definition and historical background Edible films or coatings are defined as continuous matrices that can prepared from proteins, polysaccharides, and/or lipids. These films and coatings have the potential to increase food quality and reduce food-packaging requirements since they act as mass transfer barriers to moisture, oxygen, lipids and solutes and can also carry a wide range of food additives including vitamins, colorants, flavoring agents, preservatives and antimicrobial agents. Historically, yuba, the first freestanding edible film, was developed in Japan from soymilk during the 15th century and used for food preservation (Guilbert and Biquet, 1996). Edible coating for food products date back even further with waxes applied to oranges and lemons in China to retard water loss during the twelfth century (Hardenberg, 1967). During the sixteenth century, coating food products with fat (e. g. lard) was used to control moisture loss in foods in England (Labuza and Contrereas-Medellin, 1981). Hot melt paraffin waxes have been used to coat citrus fruits in the United States since the 19308, and camauba wax and oil-in-water emulsions have been used for coating fresh fruits and vegetables since the 19505 (Kaplan, 1986). Currently, edible films and coatings find use in a variety of applications, including casings for sausages, and chocolate coatings for nuts and fruits. 1.1.2. Formation of edible films 1.1.2.1. Components of edible films Edible films typically contain three major compounds; proteins, polysaccharides, and lipids. Proteins used in the edible film include wheat gluten, collagen, corn, soy, peanut, casein and, whey protein (Kester and F ennema, 1986). Polysaccharides incorporated in edible films have included alginate, dextrin, pectin, carrageenan and cellulose derivatives (Greener and F ennema, 1994). Suitable lipids include waxes, acylglycerols, and fatty acids (Greener and Fennema, 1993; Park et al., 1994; Debeaufort and Voiley, 1995). In addition, composite films containing both lipid and hydrocolloid components have been developed. Plasticizers can be added to film-forming solutions to enhance properties of the final film. Plasticizers will decrease brittleness and increase flexibility of the film, which is important for packaging applications. Common food-grade plasticizers include sorbitol, glycerol, mannitol, sucrose, and polyethylene glycol. Plasticizers used for protein-based edible films decrease protein interactions, and increase both polymer chain mobility and intermolecular spacing (Lieberman and Gilbert, 1973). Plasticizers must be small molecules with low molecular weights and high boiling points that are highly compatible with the polymers (Banker, 1966). The type and concentration of the plasticizer influences the properties of protein films (Gennadios et al., 1994a; Cuq et al., 1997; Gueguen et al., 1998). When large amounts of plasticizer are used, mechanical strength, barrier properties, and elasticity decrease (Gontard et al., 1993; Park et al., 1994; Cherian et al., 1995; Debeaufort et al., 1997; Galietta et al., 1998). Crosslinking agents are used to improve water resistance, cohesiveness, rigidity, mechanical strength and barrier properties (Guilbert, 1995; Marquie et al., 1995; Remunan-Lopez et al., 1997). Commonly used covalent crosslinking agents include glutaraldehyde, calcium chloride, tannic and lactic acids. Exposure to ultraviolet light will increase the cohesiveness the protein films through the formation of crosslinks (Brault et al., 1997). Alternatively, enzymatic crosslinking treatments with transglutarninases or peroxidases can be used to stabilize films. 1.1.2.2. Film Formation Techniques Many techniques have been developed for forming films. These include coacervation, thermal gelation, solvent removal and solidification of melt. In coacervation, two solutions of oppositely charged hydrocolloids are combined, causing interactions and precipitation of the polymer complex. Solvent removal is another common means of forming hydrocolloid edible films. In this process, a continuous structure is formed and stabilized by the interactions between molecules through the action of various chemical or physical treatments. Macromolecules in the film-forming solution are dispersed in a solvent medium such as water, ethanol or acetic acid that contains several additives (plasticizers, crosslinking agents, solutes). The film forming solution is then cast in a thin layer, dried, and detached from the surface. In preparing some types of protein films (whey protein, casein, soy protein, wheat gluten), the macromolecule solution is heated for protein gelation and coagulation which involves denaturation, gelification or precipitation, followed by rapid cooling of hydrocolloid solution. Intrarnolecular and intermolecular disulfide bonds in the protein complex are cleaved and reduced to sulfhydryl groups during protein denaturation (Okamota, 1978). When the film-forming solution is cast, disulfide bonds are reformed which link the polypeptide chains together to produce the film structure with hydrogen and hydrophobic bonds also contributing to film structure. Another common film-forming technique is melting followed by solidification. Solidification of the melt by cooling is commonly used to prepare lipid films. Wax films can also be formed by casting molten wax on a dried film of methylcellulose and then dissolving away the methylcellulose film (Donhowe and Fennema, 1993). 1.1.3. Protein-Based Edible Films The proteins that have been used most extensively for production of films and coatings include whey protein, casein, corn zein, soy protein, gluten, collagen, and gelatin. 1.1.3.1. Whey Protein Films 1.1.3.1.1. Composition and Structure Whey protein is the protein that remains in milk serum after acid/rennet coagulation of fluid milk (de Wit, 1989; Morr and Ha, 1993). Whey protein consists of five different proteins: B-lactoglobulin, a—lactalbumin, bovine serum albumin, proteose- peptones, and immunoglobulins (Kinsella and Whitehead, 1989). B-Lactoglobulin (B-Lg) which comprises 50-75% of whey protein, is a globular protein with a molecular mass of approximately 18,300 Da (Eigel, 1984). B-Lg consists of 162 amino acid residues and contains two disulfide groups, one free sulfhydryl group, and several hydrophobic groups located in the interior of the globular structure (Brunner, 1977). The thiol group is particularly important since it facilitates molecular thiol- disulfate interchange reactions which allow formation of intermolecular disulfate-bonded dimers and polymers upon heating (Kinsella, 1984). The structure of B-Lg is pH dependent, with this protein existing as a dimer in solutions above pH 5.2. Below pH 3.5 and above 7.5, the dimer dissociates to a monomer, and between pH 3.5 and 5.2 the dimer polymerizes to an octomer. B-Lg undergoes time and temperature dependent denaturation reactions at temperatures above 65°C, which result in a general molecular expansion, exposure of the internal SH group, and hydrophobic and s-NHZ groups (Brunner, 1977; Kinsella, 1984). Structurally, B-Lg has about 15% helix, 43% B-sheet, and 15-20% B-turns (Papiz et al., 1986). a-Lactoalbumin (on-La) which comprises about 19% of total whey protein (Dybing and Smith, 1991), is another globular protein that contains 123 amino acid residues and four disulfide bonds. It has a molecular weight of 14,000 Da. The secondary structure of this molecule is composed of 30% a- helix, 9% B-sheet, and 61% unordered structure (Alexandrescu et al., 1993). Conformational changes occur in a-La at pH 4 with the molecule losing Ca+2 which is tightly bound at higher pH. a-La is denatured at 65.2°C and pH 6.7 with 80 to 90% of this denaturation reversed afier cooling. This reversibility is lost if the native S-S bonds are broken, for example, by heat-induced thiol- disulfate interchange reactions between a-La and B-Lg (de Wit, 1989). 10 1.1.3.1.2. Film Formation Film production requires heat denaturation since the hydrophobic and SH groups are deep within the globular whey protein. Denaturation changes the three-dimensional protein structure and exposes internal SH and hydrophobic groups (Shimada and Chefiel, 1998), which promote formation of new intermolecular S-S and hydrophobic bonds during drying (McHugh and Krochta, 1994a). Film formation is favored in more alkaline film solutions since SH reactivity increases at pH > 8 (Kella and Kinsella, 1988; Baneijee and Chen, 1995). Glycerol, sorbitol and polyethylene glycol have been commonly used as plasticizers to produce whey protein films that are transparent, bland, and flexible, and have excellent oxygen, aroma and oil barrier characteristics. However these films are poor moisture barriers due to their hydrophilic character (Table 1.1). Incorporation of lipids, including acetylated monoglycerides, waxes, fatty alcohols and fatty acids, into WPI film decreases water vapor permeability (WVP) by increasing hydrophobicity (McHugh and Krochta 1994b; Sherwin et al., 1998; Kim and Ustunol, 2001). Addition of beeswax and fatty acids to WPI film improves moisture barrier properties more than fatty alcohols with the longer chains more effectively reducing WVP (McHugh and Krochta, 1994c). Sherwin et a1. (1998) reported that particle size in the emulsion film increased with increasing fatty acid chain length from C12 to C22. McHugh and Krochta (1994c) showed that incorporation of large lipid particles decreased WVP of emulsion WPI film due to interactions at the protein-lipid interface or the dispersed phase. WPI film properties are affected by the extent of intermolecular disulfide bonding and plasticizers competing for protein chain-to-chain hydrogen bonding. Plasticizer content and relative humidity (RH) both have an exponential effect on film permeability. ll Table 1.1. Water vapor permeability (WVP) and oxygen permeability (OP) of edible films. Film WVP OP References g.mm/m2.d.kPa cm3.p.m/m2.d.kPa Coll - <0.O4a Lieberman and Guilbert (1973) Zein 12-24 a 1 1.8 ° Gennadios et a1. (1993), Parris and Coffin (1997) WG 53 b 3.9-6.1d Gennadios et a1. (1993) SP1 72-154 a 1.6-4.5 ° Park and Chinnan ( 1990), Gennadios et al. (1993) WPI 62.0- 70.2 c 18.5 —76.1° Krochta (1994) Casein 45.2 d 1.8 a Chick and Ustunol (1998) Alginate 42.2 e - Parris et al. (1995) Chitosan 1.6 f 90.2 ° Caner et a1. (1998) MC 7.7 ‘ 187.4 ° Park and Chinnan (1994) HPMC 9.5 f 297.3 ° Park and Chinnan (1994) Pectin 41.4 f 57.0 c Wong et al. (1992) Starch 220 g 2.9 c Allen et al. (1963) Beeswax 0.05 f 1.3 c Kester and Fennema (1989) Candelilla 0.02 f 0.3 c Kester and Fennema (1989) Camauba 0.03 f , 0.2 c Kester and Fennema (1989) Microcrystalline 0.03 f 2.2 c Kester and Fennema (1989) Acetylated 20.04 — 53.7 f - Lovegren and F euge (1954) monoglycerides Coll: Collagen, HPMC: Hydropropionate methylcellulose, MC: Methylcellulose, WG: Wheat gluten, WPI: Whey protein isolate, SP1: Soy protein isolate WVP: Temp. (°C), RH (%) a= 25°C, 50/ 100% RH, b= 21°C, 85/0% RH, c= 25°C, 0/79% RH, d= 378°C, 100/90% RH, e=30°C, 0/ 100% RH, f=25°C, 0/ 100% RH, g=38°C, 30/ 100% RH OP: Temp. (°C), RH (%), a= 23°C, 0% RH, b = 25°C, 63% RH, c= 25°C, 0% RH, d=23-38°C, 0% RH, e= 23°C, 50% RH At film compositions where glycerol- and sorbitol—plasticized WPI films have equal mechanical properties, sorbitol-plasticized films have lower oxygen permeability (OP) (McHugh and Krochta, 1994d, e). Inhibition of intermolecular disulfide bond formation with sodium dodecyl sulfate (SDS) increases WPI film solubility, extendibility and flexibility, while having little effect on WVP. Inhibition of sulfliydryl-disulfide interchange with N-ethyl maleimide reduces WPI film solubility and elongation, with little effect on other film mechanical properties or WVP. Reduction of disulfide bonds with cysteine has no effect on WVP of WPI films. Tensile strength of WPI films is comparably higher than most other protein-based films; with the opposite true for elongation (Table 1.2). Heat curing is commonly used to cross-link synthetic polymers and improve mechanical and barrier properties (El-Hibri and Paul, 1985; Perkins, 1988). Miller et al. (1997) also showed that heat curing of glycerol-plasticized WPI films increases film strength, while decreasing film extendibility, flexibility and WVP. Exposing free sulfliydryl groups promotes intermolecular disulfide bond formation to produce insoluble films. 1.1.3.2. Corn zein 1.1.3.2.1. Composition and structure Zein, the prolamin (soluble in 70% ethanol) fraction of corn gluten, comprises approximately 70% of corn gluten. This protein is soluble in aqueous ethanol and insoluble in water due to higher levels of nonpolar amino acids (i.e. leucine, proline, and 13 alanine) (Shewry and Miflin, 1985). Zein is also rich in glutamine, the amide derivative of glutamic acid, which promotes protein association by hydrogen bonding (Reiners et al., 1973; Wall and Paulis, 1978). Corn zein can be fractionated into three distinct classes: a-zein, B-zein, and y-zein (Esen, 1987). a-Zein and B-zein constitute 75-85 and 10-15% of total zein and are comprised of polypeptides with molecular masses of 21,000-25,000 and 17,000-18,000, respectively. Finally, y-zein is made up of a proline-rich polypeptide of 27,000 and constitutes 5-10% of total zein. Zein is produced commercially by extracting corn gluten with 80-90% aqueous isopropyl alcohol containing 0.25% sodium hydroxide at 60-70°C (Reiner, 1973). The centrifugally clarified extract is chilled to precipitate the zein with additional extractions and precipitation used to increase purity. 1.1.3.2.2. Film Formation Preparation of zein films generally involves casting alcohol solutions on inert, flat surfaces. Formed films are peeled after the solvent has evaporated (Gennadios and Weller, 1990). Films also have been prepared from acetone solutions (Yarnada et al., 1995). Zein films are naturally brittle, and therefore, plasticizers such as oleic acid (Kanig and Goodman, 1962), glycerol (Mendoza, 1975; Aydt and Weller, 1988) and lactic acid (Wu, 1995) are needed to improve their flexibility. An alternative method to prepare zein films involves plasticization of zein by forming an emulsion with oleic acid followed by precipitation of the protein-lipid mixture to form a soft moldable resin (Lai et al., 1997). The plastic resin then can be stretched over rigid frames to obtain thin membranes that set into flexible films. 14 Films formed upon solvent evaporation contain hydrophobic, hydrogen and disulfide bonds and have been characterized as tough, glossy and scuff-resistant (Pomes, 1971). Cross-linking agents such as aldehydes will improve moisture barrier and tensile properties of zein films (Szyperski and Gibbons, 1963). Epoxy resins also increase tensile strength of zein films by crosslinking between epoxy groups and phenolic, and aliphatic hydroxyl protein groups (Howland, 1961; Howland and Reiners, 1962; Yamada etaL,1995) The oxygen permeability of zein films is quite low when compared to synthetic films such as low density polyethylene (LDPE) and polyester film; however, oxygen permeability of zein film is greater than that of collagen (Hanlon, 1992; Krochta, 1997). Higher plasticizer levels will increase the oxygen permeability of zein films. Films prepared from zein have lower or similar WVP compared to methylcellulose, ethyl cellulose, hydroxypropylmethyl cellulose, and hydroxypropyl cellulose under similar conditions (Table 1.1). Zein films produced using different solvents possess distinct properties. For example, Yamada et a1. (1995) found that zein films prepared by dissolution in 30% acetone film showed lower WVP than when dissolved in 20% ethanol. They also reported that film strength increased linearly as film thickness increased except for films prepared from the aqueous acetone solution with 1, 2- Epoxy-3-chlor0propane. At similar plasticizer levels, zein film appears to have tensile strength (TS) and percent elongation (% E) similar to collagen film (Table 1.2). The concentration of plasticizer greatly affects the mechanical and water vapor barrier properties of zein protein films (Aydt et al., 1991; Butler and Vergano,1994). Table 1.2. Tensile strength (TS) and percent elongation (% E) of edible films (23°C, 50% RH). Film TS (Mpa) % E References Coll:Sor:Gly (3.4:0.8:l) 8.1 25 Hood (1987) Coll:Sor:Gly(8.8:0.8:l) 9.1 38 Hood (1987) Corn ZeinzGly (120.5) 2.7—15.7 43-198 Gennadios et al. (1993), Aydt et al. (1991) WGzGly (2.5:1) 4.4 194.7 Gennadios et al.(1993) SPI:Gly (2:1) 4.3 78 Brandenburg et a1. (1993) WPI:Gly (5.7:1) 29.1 4.1 Krochta (1994) WPlzGly (2.3:1) 13.9 30.8 Krochta (1994) WPlzGly (1.521) 18.2 5.0 Krochta (1994) Casein: Gly (1 .4:1) 4.5 223 Chick and Ustunol (1998) AlginatezGly (2:0.7) 2.5 7.9 Parris et al. (1995) Pectin: Gly (3:0.5) 2.3 5.0 Parris et al. (1995) Chitosan 6.3 -— 31.8 14-70 Caner et al. (1998) MC 12.5 20 Debeaufort and Voilley (1997) Starch:ethylene-acrylic acid 23.9 260 Otey et al. (1977) (2:0.8) Pectin:Starch: Gly (1:1:0.5) 27-34 1.8-l3 Coffin and Fishman (1994) Coll: Collagen, Gly: Glycerin, MC: Methylcellulose, WG: protein isolate, SPI: Soy protein isolate, Sor: Sorbitol Wheat gluten, WPI: Whey When Park et al. (1994) evaluated the impact of two plasticizers, glycerin (0.36 ml/g protein) and polyethylene glycol (PEG) (0.39 ml/g protein), on the properties of zein film, % E of corn zein films containing glycerin or PEG was 3 and 94%, respectively. TS of zein films reportedly increased from 13.4 to 25.8 MPa during 20 days of storage; however, percent elongation decreased from 76 to 12%. WVP of these zein films also decreased from 0.59 to 0.41 ng.m/m2.s Pa as the ratio of glycerin to PEG decreased. 1.1.3.3. Casein 1.1.3.3.1. Composition and Structure Casein, the major protein in milk, represents 80% of total milk protein (Dalgleish, 1989). Casein consists of three principal components, a, B, and K-casein, and a minor component, y-casein, which together form colloidal micelles in milk containing large numbers of casein molecules that are stabilized by calcium-phosphate bridging (Kinsella, 1984). Caseins are phosphoproteins that precipitate at pH 4.6 and 20°C. The a-caseins have an approximate molecular weight of 23,500 and an isoelectric point of pH 5.1 (Xiong, 1997). They are more phosphorylated than other caseins, more calcium sensitive than B-casein and contain more charged residues and fever hydrophobic residues than B- casein (Dalgleish, 1989). B-casein, which comprises up 30-35% of the total casein content in milk, has a molecular weight of 24,000 and an isoelectric point of pH 5.3 (Xiong, 1997). The last 50 residues of the terminus are charged, whereas the rest of the molecule is highly hydrophobic (Dalgleish, 1989). B-casein shows temperature, concentration, and pH 17 dependent association-dissociation. At neutral pH and high temperatures, B-casein associates into threadlike polymers (Dalghleish, 1989). K-casein, which comprises approximately 15% of the total casein fraction, has a molecular weight of 19,000 and an isoelectric point of pH 3.7 to 4.2 (Xiong, 1997). Being calcium sensitive, K-casein associates with 011- and B-caseins in the presence of calcium to form thermodynamically stable micelles. Apolar residues in K-casein are concentrated at the N-terminus with the charged portion being near the C-terminus. The hydrophobicity of K-casein is between that of a- and B-caseins (Dalgleish, 1989). The negatively charged K-casein is cleaved by the enzyme rennin used to coagulate milk in cheese production. When K- casein is cleaved, the micelle structure is stabilized and the casein precipitates. Casein contains low levels of cysteine, and thus few disulfide cross-linkages which produces an open random structure. Caseinates function as good emulsifiers and foaming agents because of even hydrophilic and hydrophobic amino acid distribution. 1.1.3.3.2. Film Formation Casein films are formed from aqueous solutions of casein. No further treatment is necessary because of their random coil structure and propensity for extensive hydrogen bonding. Hydrophobic, ionic, and hydrogen bonding involve interactions in the film matrix with the surfactant nature of casein making it very suitable for producing emulsion films that are transparent, flavorless and flexible. l8 1.1.3.3.2. Film properties The inherently poor water barrier properties of caseinate films can be improved by treatment with lactic or tannic acid (Guilbert, 1986). Films can also be prepared from sodium or calcium caseinate without addition of plasticizer. Addition of plasticizers (glycerol or sorbitol) to casein film solutions at ratios of 0.6:1, 1:1, and 1.421 decreased tensile strength, but increased percent elongation, with glycerol being more effective than sorbitol (Chick and Ustunol, 1998). Calcium caseinate films have lower WVP compared to sodium caseinate films. However, when treated with calcium chloride sodium caseinate films have even lower WVP, likely because of more effective calcium cross- linking (Avena-Bustillos and Krochta, 1993). Incorporating various lipids, waxes, and fatty acids into the caseinate film solution also has been examined for reduction of water vapor permeability, with beeswax being more effective than paraffin or camauba (Krochta et al., 1990; Ho, 1992). Caseinate film formation at the isoelectric point of casein insolubilizes the film and also reduces water vapor permeability by about one-half compared to films at higher pH levels (Krochta et al., 1988). Overall, caseinate films have moisture barriers that are similar to wheat gluten and soy protein films and somewhat greater moisture barriers than corn zein films (Table 1.1). Additional modification techniques, including UV light and y-irradiation, also can be used to improve the properties of casein films. The formation of bityrosine with y- irradiation (16-64 kGy) or UV light reportedly improved tensile strength and WVP properties of calcium caseinate film (Brault et al., 1997; Ressounany et al., 1998; Mezhgheni et al., 1998). Furthermore, addition of CaClz increased the fluorescence signal 19 of bityrosine by shortening molecular distances between polypeptides, thus aiding the formation of bityrosine. 1.1.3.4. Wheat Gluten 1.1.3.4.]. Composition and Structure Wheat kernels contain 8-15% protein based on dry weight (Kasarda et al., 1976). Gliadin and glutenin are the main wheat endospenn storage proteins, comprising 85% of wheat flour protein, the remaining 15% consisting of various albumins, globulins and smaller peptide structures (Holme, 1966). While gliadin is soluble in 70% ethanol, glutenin is not. By hydration gliadin and glutenin form a colloidal complex known as gluten (Pomeranz, 1988). These properties are used to produce films from gluten protein. The amino acid composition of wheat gluten is characterized by high levels of glutamic acid and low levels of lysine and other basic amino acids (Kasarda et al., 1976; Lasztity, 1986). The amide group of glutamine promotes hydrogen bonding between gluten chains. The relatively high amounts of nonpolar amino acids, such as proline and leucine cause wheat gluten insolubility in water at neutral pH (Krull and Inglet, 1971). Both gliadin and glutenin contain disulfate bonds. Disulfate bonds link glutenin chains together to produce polymers of high molecular weight. The gliadin-glutenin complex involves both covalent and noncavolent bonding in the film matrix. 1.1.3.4.2. Film Formation Wheat gluten (WG) films are produced by deposition and drying of wheat gluten dispersions. Aqueous ethanol is the most common solvent used in film-forming solution. 20 Mechanical mixing and heating under acidic or alkaline conditions are used to form homogeneous gluten dispersions (Gontard et al., 1992; Gennadios et al., 1993a). Covalent disulfide bonds are also important in wheat gluten film formation. Disulfide bonds in gluten are reduced to sulthydryl groups in wheat gluten dispersed under alkaline conditions (Okamoto, 1978). During casting of film-forming solutions, disulfide bonds are reformed by reoxidation in air and sulfliydryl-disulfide interchange reactions. Reformed disulfide bonds link polypeptide chains, to produce the film structure. Hlynka, (1949), Beckwith et al. (1965), and Wall et a1. (1968) describe the reduction and reoxidation of disulfide bonds while Steward and Mauritzen (1966) and McDermott et al. (1969) describe sulthydryl-disulfide interchange reactions in wheat gluten. 1.1.3.4.3. Film Properties Mechanical and barrier properties of cast WG films have been reported by several investigators (Krull and Inglett, 1971; Aydt et al., 1991; Gontard et al., 1992; Gennadios et al., 1993b; Park et al., 1994; Rayas and Ng, 1997; Rayas et al., 1998). Extensive intermolecular interactions in WG result in quite brittle films, which require plasticizers. Plasticizer molecules mediate between polypeptide chains, which decreases the rigidity of the film structure (Wall and Beckwith, 1969). However, increasing film flexibility by raising the plasticizer content (glycerin) reduces film strength and WVP (Gontard et al., 1993) Oxygen permeability of wheat gluten is similar to commercial nylon (Rayas et al. 1998). Unlike WVP, at comparable plasticizer content and test conditions, the oxygen permeability of gluten films prepared from alkaline solutions is an order of magnitude 21 lower than oxygen permeability of zein films. The effects of subjecting wheat gluten films to several treatments, including lactic acid, calcium chloride have been investigated by Gennadios et al. (1993c). Soaking the film in calcium chloride and in buffer solution at the isoelectric point of wheat gluten reportedly increased tensile strength (47 and 9%, respectively) and decreased WVP (14 and 13%, respectively). Effects of replacing a portion of the wheat gluten with keratin, soy protein, com zein, or soy protein with cysteine on film properties was also studied (Gennadios et al. 1993a, b; Were et al., 1999). The addition of hydrolyzed keratin into a wheat-gluten based film solution reportedly lowered oxygen permeability by 83%, WVP by 23%, and TS by 35%, but increased % E by 32%. Replacing 30% of the wheat gluten with SP1 decreased oxygen permeability, % E, and WVP by 40, 9, and 16%, respectively, while increasing TS by 69%. Corn zein addition had no effect on OP, but decreased WVP and % E by 23 and 37%, while increasing TS by 58%. Adding cysteine (1%) to WG: SPI (4:1) increased the disulfate content of the film forming solution and thus increased TS, but had no affect on WVP and oxygen permeability (Were et al., 1999). Properties of wheat gluten films are affected by the pH of the film-forming solution (Gontard etal., 1992; Gennadios eta1., 1993a). Gontard et al. (1992) showed that wheat gluten films produced at acidic conditions had significantly lower WVP than wheat gluten films produced at alkaline conditions (Table 1.1). However, Gennadios et al. (1993a) reported that films prepared under alkaline conditions had greater TS than films prepared under acidic conditions with %E and WVP not affected by the pH difference. 22 1.1.3.5. Soy Protein 1.1.3.5.1. Composition and Structure Soybeans contain 38-44% protein, most of which is insoluble in water but soluble in dilute neutral salt solutions. Soy protein is a globular protein and is separable into four different fractions: ZS, 78, 118 and 158 (Wolf and Smith, 1961) with the 7S and 118 fractions comprising about 37% and 31% of the total protein, respectively (Wolf et al., 1962). Soy protein is rich in asparagine and glutamine residues. Both 7S and 118 fractions are tightly folded proteins, with the former containing two or three cystine groups and the later containing 20 intermolecular disulfate bonds (Gennadios et al., 1993c) Protein in the form of meal is one of the typical end products from industrial soybean processing. Protein meal can be further concentrated for the production of soy protein concentrates and soy protein isolates, which contain at least 70% and 90% protein on a dry basis, respectively. 1.1.3.5.2. Film Formation Edible films from soybeans have been traditionally produced in the Orient on the surface of heated soymilk. In these films, lipids and carbohydrates are incorporated with protein. Films obtained from soymilk are known as “yuba” in Japan, “tou-fu-pi” in China, “kong kook” in Korea, and “fu chock” in Malaysia (Snyder and Kwon, 1987). The mechanism for forming these films involves polymerization of heat- denatured protein followed by surface dehydration of the soymilk emulsion (Farnum et al. 1976). Soy protein films contain a protein matrix formed by heat-catalyzed protein— 23 protein interactions that result in disulfate, hydrogen and hydrophobic bonds. Fukushima and Van Buren (1970) showed that the mechanism of polymerization involves intermolecular disulfate and hydrophobic bonds. Film formation is favored in more alkaline soymilk solutions. Soymilk used in making these films is prepared from soybeans. In one soymilk manufacturing procedure, (Wu and Bates, 1972) soybeans that were presoaked in water for 1 hour at 65°C were by drained, ground with water, passed through a screw expeller and finally followed by a clarification to produce soymilk with a pH of about 6.7. Guilbert (1986, 1988) produced films by casting aqueous solutions containing 10% soy protein isolate and 5% glycerin as a plasticizer. Films were characterized as flexible, smooth and transparent. Production of films and coatings from soymilk and soy protein has been reviewed by Gennadios and Weller (1991). Heating and alkaline conditions favor soy protein polymerization by unfolding the protein secondary structure and exposing sulfhydryl and hydrophobic groups. Sian and Ishak (1990) and Gennadios et al. (1993) reported that film formation was possible only within pH 1-2 and pH 6-12. The optimum pH range for producing soymilk films is between 7.0 and 8.0, with darkening of the film occurring above pH 8.0 (Wu and Bates, 1972). Brandenberg et al. (1993) produced film by heating alkaline solutions of soy protein isolate (5%) and glycerin (3%) at 60°C for 10 min. These films had a smoother appearance and fewer insoluble particles, as evidenced by scanning electron microscopy. Heat treatment of SPI film-forming solutions at 85°C promoted intermolecular cross-linking and produced SP1 films that were smoother and more transparent, and possessed lower WVP and increased % E, compared to those produced from unheated solutions (Stuchell and Krochta, 1994). 24 The moisture barrier properties of SPI films are similar to wheat gluten films prepared from alkaline solutions but somewhat inferior to corn zein films (Table 1.1). SP1 films having similar TS to zein and alkaline wheat gluten films have lower E (Table 1.2). The oxygen permeability of SPI films has been measured, but only at 0% (Brandenburg et al., 1993, Gennadios et al., 1994). Like wheat gluten films, the oxygen permeability of SPI films appears to be an order of magnitude lower than the oxygen permeability of zein films at similar plasticizer levels (Table 1.2). Soy protein films are transparent and flexible when plasticizer is added, but have poor water barrier properties (Guilbert, 1986). When the plasticizer content was increased from 20 to 40% the % E of soy protein films increased from 1.5 to 106%; however, tensile strength decreased from 15.8 to 2.6 MPa (Cunningham et al., 2000). Ionizing radiation, such as y-irradiation, cross-links soy protein films, thus improving their functional properties. Exposing soy protein film to a y-irradiation dosage of 5-30 ng in the presence of poly(ethylene oxide) or poly(vinyl alcohol) increased tensile strength (Grorpade et al., 1995; Sabato et al., 2001). Gennadios et al.(1998) also reported that UV irradiation (103.7 J/mz) improved tensile strength of soy protein films from 3.7 to 6.1 MPa; however, percent elongation decreased from 124 to 85% with increasing UV irradiation. Modification of SPI film properties by heat curing at 80 and 95°C was also studied by Gennadios et al. (1994), with this treatment reportedly reducing WVP, % E, moisture content, and water solubility and increasing TS. Changes in film properties were attributed to heat-induced cross—linking and lower moisture content within the film, with the greatest effect observed at 95°C. 25 Cross-linking interactions of calcium chloride or calcium sulfate with SP1 can also improve tensile strength (Park et al., 2001). In the same study, glucono-delta-lactone (GDL) incorporation into soy protein increased elongation. In addition, both calcium salts and GDL-treated SPI films had lower WVP than the SP1 control film. 1.1.3.6. Collagen 1.1.3.6.1. Composition and Structure Collagen is a fibrous protein generally isolated from hides, tendon, cartilage, bone and connective tissue (Ballian and Bowes, 1977). This protein is arranged in fibrils which are packed together to form bundles of parallel fibers having cross-sectional diameters of 20-40 microns (Ballian and Bowes, 1977). Collagen is unique in that every third amino acid residue throughout most of the structure is glycine, while proline and hydroxyproline account for about one-fourth of the residues (Harrington and Von Hippel, 1961, Ramachanran and Rarnakrishnan, 1976). Collagen fibers swell with immersion in alkali, acid or neutral salt solutions. They are resistant to proteolytic enzymes but are easy targets for collagenase. When heated, collagen fibers shrink and are converted to gelatin (Harrington, 1966). 1.1.3.6.2. Film Formation Collagen films can be produced from animal hides using a dry or wet process. The dry process involves alkaline treatment of hide corium and acidification to pH 3.0 (Hood, 1987) followed by shredding of acid-swollen coriums to preserve maximum fiber structure, mixing of acid-swollen fibers to produce a high-solids dough (>12%), addition 26 of plasticizing and cross-linking agents, high pressure pumping and extrusion of dough to form tubular casings, drying, conditioning, neutralizing and/or providing additional cross- linking. The wet process consists of acid or alkaline dehairing of hides, deacidification of hide coriums, grinding mixing of ground collagenous material with acid to produce a swollen slurry (4-5% solids), slurry homogenization, extrusion into tubular casings, washing to remove salts, treatment with plasticizing and cross-linking agents and drying. The tensile strength of collagen is less than that of methyl cellulose, corn zein, WPI, and Chitosan-based films (Table 1.2). Several techniques have been developed to improve the properties of collagen films and casings. The strength of collagen casings can be improved by drying to a moisture content of 15 to 40% in the presence of a polyhydric alcohol, such as poylethylene glycol, or a salt, preferably sodium chloride (Kidney, 1970). Treatment with glyceraldehyde was found to promote cross-linking which increased the strength and thermal resistance of collagen casings (Jones and Whitmore, 1972). Alkyl diols also improved the mechanical properties of collagen casings by reducing internal hydrogen bonding while increasing intermolecular spacing (Lieberman and Gilbert, 1973; Boni, 1988). Furthermore, exposing edible tubular collagen casings to ultraviolet irradiation (180-420 nm) increased casing strength (Miller and Marder, 1998). Exposure to proteolytic enzymes such as papain, bromelain, ficin, trypsin, chymotrypsin, pepsin and fungal proteose can be used to partially stabilize collagen, which will improve uniformity in diameter and wall thickness of collagen casings (Fujii, 1967; Tsuzuki and Lieberman, 1972; Miller, 1983). 27 Water vapor pemeability of collagen films can also be altered. Lieberman and Gilbert (1973) found that formaldehyde cross-linking and chrome tanning significantly reduced the gas permeability of collagen films. Formaldehyde reacted with collagen by combining with free amino groups of basic amino acids, whereas chrome tanning promoted reactions between carboxyl groups of amino acids. 1.1.3.7. Gelatin Films 1.1.3.7.1. Composition and Structure Gelatin, a protein produced from partial hydrolysis of collagen, ranges in molecular weight from 3000 up to 200,000 depending on production methods. Gelatin lacks internal order with the polypeptides configured randomly in aqueous solutions (Xiong, 1997). The process for hydrolysis of collagen to gelatin includes thermal denaturation (40°C), which cleaves hydrogen bonds and electrostatic bonds, and hydrolytic breakdown of covalent bonds (Eastoe and Leach, 1977). Thermally reversible gels are produced by forming cross-links between amino and carboxyl components of amino acid residual side groups (Glicksman, 1982). Raw materials commercially used for gelatin manufacture include pork skin as well as bovine hide, skin, and bones. A pretreatment or curing step is applied before gelatin extraction of the raw material. Curing removes impurities and initiates collagen hydrolysis. Two different curing methods, acid or alkali, produce Types A and B gelatin, respectively. Type A gelatin with an isoelectric point between pH 7 and 9 is mostly made from pork skin, while Type B gelatin with an isoelectric point between pH 4.6 and 5.2 is made from bones and bovine hides (Xiong, 1997). 28 2.1.3.7.2. Film Formation Gelatin films are prepared by casting aqueous film-forming solutions of 20% gelatin and 0-10% glycerin (Guilbert, 1986, 1988). Such films are clear, flexible, strong and oxygen impermeable when the aqueous solution contains plasticizer such as glycerol or sorbitol. Gelatin gels dry to produce films with poor moisture barrier and tensile strength properties. Treatment with lactic or tannic acid improves water-barrier properties although these films are less flexible and less transparent (Guilbert, 1986). Tensile strength of gelatin films can be improved by drying 5% solutions at 20°C rather than 60°C, due to a higher degree of crystallization in films dried at the lower temperature (Bradbury and Martin 1952). In addition, gelatin also reportedly has a configuration that promotes interchain hydrogen bonds at temperatures less than 35°C (Robinson, 1953). 1.1.4. Polysaccharide-based edible films Several carbohydrate-based edible films were developed and tested. This review will address some common carbohydrate films including cellulose-, chitosan-, alginate-, starch-, pectin-, and dextrin—based edible films. 1.1.4.1. Cellulose Films Cellulose, composed of linear chains of B-D-glucopyranosyl units joined by glycosidic bonds (1 —+ 4), is the most abundant organic compound on earth because it is the principal component of plant cells. Native cellulose is insoluble in water because of the high level of intermolecular hydrogen bonding, but can be converted into cold water— 29 soluble gums by esterification. In this process, cellulose is reacted with aqueous caustic, then with methyl chloride, propylene oxide or sodium monochloroacetate to yield methylcellulose (MC), hydroxypropyl methylcellulose (HPMC) and sodium carboxymetylcellulose (CMC), respectively. MC, HPMC and hydroxypropyl cellulose (HPC) are available commercially in powder or granular form in varying molecular weight. Their relative hydrophilities increase in the order of HPC o>E :5 Ana: .3 a chasm . . . c . w . B _ . .58: .3 co Hon—m Ammac camzeEBw 238m? 653 088 0385 .3 8 352m .23: .Qzesehaoum mom—E. .mfifl. .833. -585 05 .3 a 8:23 .88: 8532.885 :80ch .25 5:5 BE: Ea €84.53 .3 8 meow .833 33 .N 58 683:8 33¢ .3262; 8583 0528 .38: :52 Ba x5 .m SE“ $3.8m .53 .EN .38 .mvoow 33m $83538 odv 533 2 .4 cashew mo 83E: 8?: x83 .83: 208 93 mowmeokfi .3335 5:5 86335 5.32 33 $355-5 38> .Eztsemsnfi... nmc 93 .32: ”corona Avg: .3 a cocgmav— .m. .28 M 53.920 £3 638%? 23.552: 0528 .83: .3 Ho 3:55. .umzoMSAoeeoE 4 68—230 5832 53G :3 £82 adv .533 0:4 «unencum— coaooba Efiaozaa< not: aonaouom mfinmuauueouum—Z 5:.”— Ezuaomaad‘ each be 0.52 In bans—om «Ma 285.3. 2:39.282: .3 2.3. 54 8.3.1.02 :5 8&5. .808: .3 s 35.0.... .330 .53: 052...: .6 roams... :5 53:50:02 .88: 90:. :5 :magik... 0.: :5. 3.8. :5 83520 33:823.... :8 83 .025. 0.5852. 2.3.8 20.2.. .83: 0.3.3! Q 80305300205 4 .5855 - =00 5.03:5: - 33>» - b3: .§a§0a50§00300wNN 9.005 :00935 .300 3 .333 33.55 035.8200 00.5 :5 :02 .300: a00§0§003h :00925 :5. 20:3 .5330M003U .53.. 0.58 5323.3 .888: 523025.00 5.833 03:05.0. 93 5550.. A50: .3 :0 ..&%.83:05.03=0a 82030 .838: .002 . :5 5.52 .888: ...W 30.38.383.03: 8.5.: a 8:3. .88: 80...... . 8.538%...» u........@. :5 5:0”. A50: 3.8.33. w. 55.55593 35:09:00 0.5.0.0 .85: .02... 33:85.30 0.83.38 .85: 25.0.0 .83: :8 .3 .83. 2.33: 00 030.2 5.35:: ::.0 8.0.5.5 5383\0300000303m. 0305 dab? 05 33.3005 .93: £050.: 8.305 3:53:88: 2:05 053500.85. :5 .3301 Ana: .53....033MOE .m .3200 .0230 :0 050.5505 $303.5 :5 03v. 032.63% 53.3338: >00 .002 95003030 Ammo: 03 .9 SC .3 :0 53.3.5005 .3053 65.3935 0.: 0: 035.0: 05200 E: 8.8:... s 9.2 0. 00:55:22.. q 20...: EB .55: .830... 8 ..N .2....» 8.3-88 82.8.... 9303:: 553:9}: 553:9}. :35: 32203: 3:50:22: 5...: :8... 0.. 0.5:. 5. 5.338 a... 0.30.2. 5.2—00v m—aschomfitnd‘ .m 033—. 55 flea—veg began 98 5E .82? can.» .350 333 aoflovfix Ea :an .8835 $2on $6353: :8on .28: .3 63.83.85 £83: a. .53 Jon—3% 9328 28 no :Smtoso .333 58 .m .mgmMSAucneE .cowE—oo .3an .83. - 533 wed 0503 c8259 can 555 d .uzuzozcum. 856 95m 23 :3an 2% mafia .358 dun—om .8323 £8.50 «8.335 .33.: .8033 589$ 23 826% team—8b EB 98 3mm 55020 05 833 awe—«5 8.605 Eot no: .3 an 8.59 .833 NOE“ .momoono EEO—among .130 gong .83: .8333 €2on 02558 .3 Ho 832:2 «an «SSSSQQSM. i3. .865935 28 Ana: .3 a 225 cages; $320 $29.06 :2.me Ass: .3 s 225 :8 m ..§ 353% :3 33 .2535 2823 0328 28 .83: .3 8 =om .umamMSAugoE d £3920 £0305 beam 833082 - 53>? Euv oboo< ‘88:: aouaum—aaA‘ .333 89:03qu mfiaaauuocuflz EE— ucfla9=na< .53— uc 0.82 In 5:335 53 256.3 8.238 23:225.: .3 as: 56 3&0ng 33:0 22 Asa“: 2830 €58 330a .8800: 33 EB 88 3 80:20 .20me sotunEwov 0528 083:8 .EoN 68.3% =2» :00 SA 833 2 293063 .3 Ho Home.» amzmmoiueacfi d .m.-uumu\§uuououmucm .383 38986“: .mmnmwouoku U ,Qmwzrxsax .0 .Eafizéca 5333.85 .383 28:88.93 .EEESR 83: .3 .3 .més 3583293 Ho woman; .33: 63% .w .328 35.25% 82603 .3 8 bananas; .muawMSAuczcE 4 3.23.00 82: BEEF—om o-m - Eoomvom 3:38:50 fin“. manucuofiuhm Sago 3.355 . i5. mauveuofimm ghee—o no 82210300 23 “QR“. mauuoueum: acute EB “Hm—omwwmmwmfimw rum“. 836.6933 832:8 .3 Ho 8382:. . in“. $383083 ram“. Boa $5on .233 .3 Eatmuugmzbob smash. $2603 3 .0358 Ho swig 5.3: £33385 9&9... «Emucm 3.03:8 H88 vegans—hum 28.5806 0328 .3th EB x2522 .8§M§OQ§E atmfiq .53 .53 .8850 $955 96v .583 - 5&2 =c53:nn< acuao=na< E52. 839.03% 339E: gamma-.3988: 8F.— ..ch a: 9.52 In bass—cm av:— ding—i r88 fizegaui .3 2.19 57 Emwfigum £33385 £3933 §88§3§B .&3M~§ 329$ .aaofimfmu accoschamuk .atfiis Eatmfiaamnbob £3835 azuueuoiuhm ..~§.®.?8ug£b&£ub 6&58985 $8 : 33$» .9: figo .3 6 25m #3: sfimfi .53 5 .130 8:2. Ana: :8 .m 388.83% .6580 Amoas .E Eroscnami 838m 2:0on 6 __._m§an§§a3 gag—:29? m: we 8:33 .333 .E 53.8230 ocfinEoE 05 mo a 523 .88: Jam £33»? £58.58 2: .3 Ho 059$ .333 ram guauocxumg how—u £562.38 .130 via .33: .umzmMSAuozoE EB 8333 520M .933 Q .hEEQcELmSPSmB mm :03 mm 0338 Stud 28 880 .m .533. 3:29am - - :9: mo cow—20:0 - 58? md :Eomoaoaq :83 .3 s 8533. A258 ._m 3883.2? 833: s 335 .253 35889.03 suave; €0.52 Ba .3 8 0950 .09 .mmtmwoiugos 0380»? tamanb 8.5020 0328 $0 $3: .3 Ho :32. SL8»: £3953 - can “32 Eon Kama - 53>» - Escommm A mg: :8225 azuuousémsw Sou .m 3~3h< 2:53:93 nectaumam< 82.20qu nfiflaauhcoaufiz EE— each .853 a: £55 In mam—325 53 285—5. @388 $329.55 .3 2...; 58 corn—bonnmonq $33 .3 wmzmwotonofi 335% 338:8 was 8 xohom .233 ._a scam “6:059?me now—8n: combs go atone”— .Ahwav . dam $8833: £858.: .130 SEED ram Satunuucogi 038m 8525 Aofi £333 3.33— ..n% smuugagfiam ram .33 3583 v5 .835 Amway gfiuuoofioéuux 58 .m 05 £53» .w=a> Ammo: 6:936 38835333 3260.:— moxm TOE—Ea 0328 £503 23 38>» . dam £33335 - 322 wfixooa csv .633 - BEE 33 02.020 can 829% 8 26 m=oo 33838 8m 32 owe a Ea .mg .oxSmSSg new $38 flog—a 5:528 60 A53: Eco 98 use» 5:823 858%? combs Ana—v beam 03509880 £583 £2605 £8: quote 0328 “2.2 6335mm 02339580 - 302 £50833 - 53$ - Caz cougb< ancmaauzma unczau=na< 82.9.0.3: nfiflauwucchfiz EE— ecoh :83: ac 255 In @2338 «M.— 255.5 APERVV n?329_83=< .MA 033—. 59 Gandhi et al., 1973; Kasrazadeh and Genigeorgis, 1995; Briozzo et a1, 1985) and to prevent aflatoxin and enterotoxin production (Lueck, 1980). Antimicrobial activity of sorbate is generally enhanced at low temperature. Investigators showed that L. monocytogenes, Y. enterocolitica, A. hydrophila, and Pseudomonas putida were more sensitive to potassium sorbate at refrigeration temperature than at higher temperature (El-Shenawy and Marth, 1988; Moir and Eyles, 1992). Increased microbial inhibition by sorbate in the presence of higher solute concentrations also has been reported in both laboratory media and foods. Sodium chloride reportedly acts synergistically with potassium sorbate to inhibit S. aureus (Robach and Stateler, 1980) and outgrth of C. sporagenes (PA 3679) spores (Robach and Statelen 1980). Razavi-Rohami and Griffiths (1999) also found that the inhibitory activity of sorbic acid increased with increasing NaCl concentration from 3 to 20% (w/v) and lower pH (3.5) for Candida, Sporothrix, Fusarium, Penicillium, Paecilomyces and Aspergillus spp. The inhibitory mechanism of 3% NaCl is not know but may be related to the effects of added solutes on membrane activities as suggested by Walter et a1. (1987). 1.2.3. Propionic Acids Propionic acid, a monocarboxylic acid (pKa = 4.87), is produced by Propionabacterium freudenreichi subsp. shermani. Swiss cheese contains up to 1% propionic acid from the growth of propionibacteria, which gives it a characteristic flavor and prevents mold growth (Davidson and Branen, 1993). Antimicrobial activity of the propionates is again pH dependent, with the undissociated form being 45 times more inhibitory than the dissociated form (Eklund, 60 1985). Therefore, it is only effective in low pH edible films such as collagen and chitosan. Propionic acid or its salts are commonly used food preservatives due to their wide spectrum of activity. While primarily active against molds, some yeasts and bacteria are also inhibited (Table 1.3) (Chung and Goepfert, 1970; Eklund, 1985; El-Shenawy and Marth, 1989; Cherrington et al., 1990; Cherrington et al., 1991). Bacillus cereus which causes rope formation in bread dough, can be inhibited by the addition of propionic acid at pH 5.6 to 6.0 (Woolford, 1975; O’Leary and Kralovec, 1986). Propionates can be added directly to bread dough because they have no effect on the activity of baker’s yeast. Amounts of propionate used in foods are generally less than 0.4% (Robach, 1980). Propionic acid was most effective among the organic acids at inhibiting aflatoxin production by Aspergillus flavus on betel nuts after 2 (62%) and 4 (85%) weeks (Raisuddin and Misra, 1991). Kwon et al. (1998) reported that when grown in tryptic soy broth (TSB) containing buffered propionic acid (EPA), the growth rate of S. Typhimurium gradually decreased as the level of EPA increased and the broth pH decreased. No growth was detected at BPA concentrations above 3% (v/v), with grth also markedly suppressed under anaerobic conditions and at pH 5.0 compared to 7.0. Propionic acid and the propionates (8-12%) were also effective in controlling the growth of mold on the surface of cheese and butter (Deane and Downs, 1951). When used at pH 4.5 at a level of 0.2%, propionic acid completely inhibited growth and aflatoxin formation (Ghosh and Haggblom, 1985). 61 1.2.4. Mode of action of benzoate, sorbate, and propionic acid Benzoic, sorbic and propionic acid act by inducing changes in the morphology and appearance of microbial cells. Incorporating these acids into specific cell structures may also inhibit specific biosynthetic pathways in the cell (Sofos et al., 1986). As one example, sorbate prevented amino acid (L-serine and L-histidine) uptake by S. Typhimurium at low pH (Tuncan and Martin, 1985). These acids also have been shown to reduce the intracellular pH (pHi) in E. coli cells (Salmond et al., 1984) and vesicles (Eklund, 1985), suggesting that this common preservative may act as a protonophore. Ronning and Frank (1987) reported that undissociated sorbic acid reduced the pHi in vegetative C. sporogenes PA 3679 cells, and decreased the protonmotive force which energizes cellular activities such as amino acid transport. They also proposed that the loss in energy and reduced uptake of essential amino acids induced a stringent-type response that inhibited but did not kill the cells. Past evidence has suggested that the inhibitory action of these antimicrobials on yeasts is due to reduction of pH. However, using a novel method to measure pHi in growing cells, little correlation was found between reduced growth rate on exposure to sorbic acid and reduction of pH. In fact, growth inhibition correlated with an increase in the intracellular ADP/ATP ratio due to increased ATP consumption by the cells. This was partly attributed to the activation of protective mechanisms, such as increased proton pumping by the membrane H+ —ATPase, which ensured that pH did not decline when the cells were exposed to sorbic acid (Bracey et al., 1998). Thus, the inhibitory action of sorbic acid was likely due to the induction of an energetically expensive protective 62 mechanism that compensated for any disruption of pHi homeostasis but resulted in less available energy for normal growth. Benzoic, sorbic and propionic acid also alter cell membrane function by producing pores (Freese and Levin, 1978) that interfere with uptake of substrates, electron transport and the proton-motive force involved in transport functions. In addition, sorbic acid can bind sulfhydryl groups on various enzymes, thus leading to disruption of vital processes involved in transport functions, cell metabolism and growth. Sorbate also reportedly inhibits microbial growth by combining with coenzyme A (Nose et al., 1982) with Harada et al. (1968) theorizing that sorbate may act competitively with acetate at the site of acetyl coenzyme A. 1.2.5. Parabens Esterification of the carboxyl group of benzoic acid produces parabens. Having the ability to remain undissociated at pH values up to 8.5, most parabens are active at pH 3.0 to 8.0. The methyl, propyl and heptyl parabens can be used as food preservatives in most countries, while the ethyl and buthyl esters are more restricted. Parabens can be used effectively in a wide range of foods (Table 1.3). Parabens with a longer alkyl chain posses more antimicrobial activity than those having a shorter alkyl chain (Aalto et al., 1953; Shibasaki, 1969). Parabens are more inhibitory on Gram-positive than Gram-negative bacteria due to their decreased in polarity. Methyl, ethyl, propyl and butyl parabens completely inhibit the growth of Gram- positive bacteria and Gram-negative bacteria at concentration levels of 40-2000 and 50- 4000 pg/ml, respectively (Table 1.3) (Aalto et al., 1953; Jurd et al., 1971; Lee, 1973; 63 Kato and Shibasaki, 1975; Robach and Pierson, 1977; Dymicky and Huhtanen, 1979; Eklund, 1980; Lueck, 1980; Eklund, 1981; Reddy and Pierson, 1982; Reddy et al., 1982; Payne et al., 1989; Juneja and Davidson, 1992; Moir and Eyles, 1992; Davidson, 1993). However, parabens are generally more active against molds and yeasts than bacteria. Using esters of p-hydroxybenzoic acid, concentrations of 32 to 1000 ug/ml are normally needed for complete inhibition of bacteria and fungi (Table 1.3) (Aalto et al., 1953; Kato and Shibasaki, 1975; Marwan and Nagel, 1986; Jermini and Schmidt-Lorenz, 1987; Juneja and Davidson, 1992; Thompson, 1991) The antimicrobial mechanism for paraben is very similar to that of phenols and related phenolic compounds, since parabens are phenolic derivatives. Vas (1953) and Judis (1963) both proposed that phenol physically damages the cytoplasmic membrane of microorganisms, which causes the release of cytoplasmic compounds. Furr and Russel (1972) detected similar leakage of intracellular RNA by Serratia marcescens in the presence of parabens, with the amount of leakage proportional to the alkyl chain length of the paraben. Additional studies have shown that the parabens inhibit nutrient uptake through the cytoplasmic membrane, inhibiting both membrane transport and the electron transport system (Freese et al., 1973; Eklund, 1980). According to Freese et al. (1973) serine uptake by cytoplasmic membrane of B. subtilis is inhibited by parabens with Eklund (1980) showing inhibition of alanine, serine, phenylalanine, and glucose uptake by vesicles in E. coli, B. subtilis, and P. auruginosa. 64 1.2.6. Free Fatty Acids and Their Esters Low concentrations of long-chain fatty acids are inhibitory to microorganisms, especially Gram-positive bacteria and yeasts (Kabara, 1978). Saturated fatty acids having chain lengths of C12 to C16 and Cm to C12 possess the most antimicrobially activity against bacteria and yeasts, respectively (Kabara, 1993). Decreasing effectiveness of longer chain fatty acids may be related to increased hydrophobicity and decreased solubility (Wang and Johnson, 1991). Fatty acids are also more active at low pH (<5.0). Fatty acid structure and function has been reviewed by Kabara (1982, 1993). The inhibitory effects of unsaturated fatty acids increase as the number of double bonds increases. For example, linoleic acid was far more inhibitory than oleic acid (Fuller and Moore, 1967). Fatty acids and monoglycerides are inhibitory to bacterial species (Table 1.3) (Noterrnans and Dufrenne, 1981; Baker et al., 1985; Knapp and Melly, 1986). According to Sprong et al. (1999), increased intake of bovine milk fat enhanced gastrointestinal killing of L. monocytogenes, but had little effect on infection with Salmonella enteritidis in rats. Free fatty acids C100, Cm) and Cm) and the monoglycerides of Cm), Cm), and Cm) likely play a pivotal role in this enhanced listeriacidal activity. In contrast, infection with Salmonella was not affected by milk fat consumption. Monoesters of glycerols and the esters of sucrose are more antimicrobial than their corresponding free acids. Monolaurin (lauricin), the most effective of the glycerol monoesters, is inhibitory to various Gram-positive bacteria and some fungi at 5-100 ppm (Andrews and Grodner, 1997). Monolaurin is most effective at pH 5.0-8.0. Antimicrobial activity of monolaurin has been demonstrated in various foods including mechanically 65 deboned chicken meat, minced fish, chicken sausage (Baker et al., 1982), soy sauce (Kato, 1981), meat slurries (Noterrnans and Dufrenne, 1981), cottage cheese, pork homogenate (Robach et al., 1981), skim milk (Wang and Johnson, 1992), and crawfish tail (Oh and Marshal, 1994). However, use of monolaurin as a food preservative is limited due to off-flavors and the loss of activity from the interaction with lipophilic proteins, fat globules and starch. Fatty acids and poylglycerides are added to edible films and coatings to decrease WVP properties. Long-chain alcohols (e.g. stearyl alcohol) and fatty acids (e.g. stearic, palmitic) are commonly used as additives in edible coatings due to their high melting point and hydrophobic chracteristic (Hagenmaier and Shaw, 1990). Vojdani and Tores (1990) developed composite films with MC and fatty acids of different chain lengths to decrease migration of preservatives like potassium sorbate from the surface of foods such as cheese. The cellular membrane has been suggested as a primary target for antimicrobial activity (Freese et al., 1973) with fatty acids impairing cell permeability and the transport of nutrients (Greenway and Dyke, 1979). 1.2.6. Organic Acids 1.2.6.1. Acetic Acid Acetic acid (CH3COOH), the primary component of vinegar, is produced by Acetobacter species. Acetic acid and its salts are commonly used in variety of different foods (Table 1.3). Acetic acid, like other organic acids, can be used to acidify edible films prepared from chitosan, alginate, collagen and WPI. Addition of acetic acid also increases the activity of other antimicrobial agents such as sorbic acid and benzoic acid that can be incorporated into edible films. 66 Application of acetic acid and its salts in the meat industry has met with variable success. Adding acetic acid to chiller water (pH 2.5) in poultry processing reportedly extended the shelf-life of poultry (Mountney and O’Malley, 1965). Processing poultry in scald tank water containing 0.1% acetic acid decreased the heat resistance of Salmonella newport, S. Typhimurium, and Campylobacter jejuni (Okrend et al., 1986). However, in another study, incorporating 0.5% acetic acid in scald water had no effect on Salmonella, total aerobic bacteria, or members of the family Enterobactericea on poultry carcasses (Lillard et al. 1987). Acetic acid solution dips (1 to 3%) can reduce the number of pathogenic and spoilage microorganisms on beef and lamb tissue (Bell et al., 1986; Anderson et al., 1988). Growth of Listeria innocua, S. Typhimurium, E. coli 0157:H7, Clostridium sporogenes, aerobic bacteria, lactic acid bacteria, and Pseudomonads was suppressed or eliminated in ground beef prepared from carcasses which were washed with 2% lactic acid, 2% acetic acid, and 12% trisodium phosphate solution as compared to the untreated control (Dorsa et al., 1998). The combination of 100 mM NaOH (pH 10.5) and 76.7 mM acetic acid (pH 5.4) applied sequentially at 55°C for 5 min was very effective at eliminating L. monocytogenes from biofilms (~ 4.5- to 5.0 logs CPU/cm2 decrease) (Arizcun et al., 1998) Acetic acid is variably effective as an antimicrobial when used as a spray on meat carcasses. Washing with water (35°C) followed by a 2% acetic acid spray (55°C) treatment was more effective at reducing the numbers of E. coli 0157:H7 and S. Typhimurium than trimming or washing alone (Hardin et al., 1995). Similarly, a 5% acetic acid spray decreased populations of E. coli 0157:H7 on beef tissue by 1 log/cm2 as compared to the untreated control (Delzari et al., 1998). When used alone or in 67 combination with a pulsed electric current, another acetic acid treatment reportedly decreased Salmonella populations 1 log on carcasses (Tinney et al., 1997). Acetic acid caused greater inactivation of L. monocytogenes than lactic and citric acid with inhibition increasing at lower incubation temperatures. Lactic acid concentrations > 0.3% inhibited L. monocytogenes at 13 and 35°C (Ahamad and Marth, 1989, 1990). Addition of 0.1% acetic to TPB minimized the inhibitory effect of 100 and 200 CFU/ml SML during the first 32 h of incubation. Staphylococcus aureus was inhibited to a lesser extent when cultured in TSB supplemented with SML alone compared to SML with organic acids (Monk et al., 1996). A 2% acetic acid wash reportedly killed L. monocytogenes in a model system of fresh meat fluids within 24 h while the pathogen increased 1.0 to 2.0 logs using a nonacid water wash (Samelis et al., 2001). Sodium diacetate is also inhibitory to L. monocytogenes, E. coli, Pseudomonas fluorescens, Salmonella enteritidis, and Shewanella putrefaciens (Shelef and Addala, 1994). Sodium acetate at 1% reportedly increased the shelf life of catfish fillets by 6 days when stored at 4°C compared to the control (Kim et al., 1995). Sodium acetate will also inhibit the rope-forming bacteria and various molds at pH 3.5 to 4.5 in baked goods (Table 1.3) (Glabe and Maryanski, 1981) and prevented mold growth in cheese spreads (Doores, 1993). Vinegar vapor (4-6% acetic acid) has been used to inactivate conidia of fungal pathogens on various fruits. Vapors of several common Vinegars containing 4.2% to 6.0% acetic acid effectively prevented conidia of brown rot (Monilinia fiucticola), gray mold (Botrytis cinerea), and blue mold (Penicillium expansum) from germinating and causing decay of stone fruit, strawberries, and apples, respectively. Vapor from 1.0 ml of red 68 wine vinegar (6.0% acetic acid) eliminated decay on apricot caused by M. fructicola. Similarly, vapors from 1.0 ml of white vinegar (5.0% acetic acid) reduced decay on strawberries by B. cinerea from 50% to 1.4%. Eight different Vinegars, ranging from 4.2% to 6.0% acetic acid, of which 0.5 ml of each vinegar was heat-vaporized, reduced decay by P. expansum to 1.0 % or less on apples (Sholberg et al., 2000). 1.2.6.2. Lactic Acid Lactic acid (CH3CHOHCOOH), produced naturally by lactic acid bacteria during food fermentation, is primarily used for improving and controlling the quality and microbial stability of foods. Lactic acid sprays (1 to 3% solutions) often have been used to sanitize meat surfaces as discussed in several reviews (Smulders, 1986; Cherrigton et al., 1991; Dickson and Anderson, 1992) with psychrotrophic Gram-negative bacteria generally being more sensitive than Gram-positive organisms to this treatment. Lactates are generally used as humectants and flavor enhancers in meat and poultry products (Duxbury, 1988). Levels of 2% are used in hot dogs, frankfurters and similar products. Evans et a1 (1991) reported that injecting a 4% lactic acid solution into beef roast before cooking increased the cooking yield (15%) and lowered the aerobic plate count during refrigerated storage. The antimicrobial activity of lactic acid depends on the food application and the target microorganism. Lactic acid is more effective than malic, citric, propionic or acetic acid in inhibiting growth of Bacillus coagulans in tomato juice (Rice and Pederson, 1954). Lactic acid is also inhibitory to spore forming bacteria as well as S. aureus and Y. enterocolitica (Minor and Marth, 1970; Woolford, 1975; Bracket, 1987). 69 Like other organic acids, lactic acid can be used in edible film formulations such as chitosan and collagen for acidification purposes. As mentioned earlier, lactic acid also can be used modify both the tensile strength and antimicrobial properties of collagen casings. 1.2.6.3. Mode of action of organic acids Organic acids inhibit oxygen uptake and resultant ATP production in whole cells of B. subtilus (Sheu and Freese, 1972; Freese et al., 1973). However, they do not inhibit NADH oxidation by isolated membranes. Glycerol phosphate- or NADH-energized uptake of serine transport in membrane vesicles yields similar results. Hence, they appear to inhibit growth by uncoupling substrate transport and oxidative phosphorylation from the electron transport system, which in turn inhibits uptake of metabolites by the cell. Sheu et al. (1975), later determined that acetic acid interfered with PME, with these organic acids also acting on cellular enzymes to reduce intracellular pH (Huang et al., 1986). 1.2.7. Bacteriocins Bacteriocins are antimicrobial substances that have a peptide or protein component essential for their activity. Although most bacteriocins have a narrow spectrum of inhibition and only inhibit closely related species, nisin and pediocin exhibit a broader spectrum of activity and have consequently some attention as antimicrobial additives to edible film. 70 1.2.7.1. Nisin Nisin, the first bacteriocin to be used in the food industry, was recognized as a safe biological food preservative by a joint F AO/WHO commission on food additives in 1968 (FAO/WHO 1969). Twenty years later, nisin was accepted by the United States Food and Drug Administration (FDA, 1988). Nisin has been shown to be very effective in cheese products, including processed cheeses and cold-pack cheese spreads (Delves- Broughton, 1990). The legal precedent for use of nisin in US. foods was set with pasteurized cheese spreads (900 IU/mg) (Food and Drug Administration, FDA, 1988). Nisin, a protein of 34 amino acids produced by Lactococcus lactis subsp. lactis, (Jung, 1991) possesses amphiphilic characteristics with clusters of hydrophobic and hydrophilic residues at the N and C-terminus, respectively. Gross and Morell (1967) identified didehydroalanyllysine and isoleucine at the C- and N-terminus of nisin, respectively (Gross and Morel], 1970). The thioether cross-linkages and highly reactive double bonds are likely responsible for important functional properties of nisin, including acid tolerance, therrnostability and bactericidal activity. One of the mostly investigated bacteriocins for antimicrobial edible film studies, nisin can be incorporated into film solution or applied directly to the film surface after casting. Various nisin-containing protein-based films (e.g. whey protein, corn zein-, wheat protein-, and soy protein) have been assessed for antimicrobial activity against Gram-positive bacteria such as L. monocytogenes and lactic acid bacteria. Since nisin is more active in hydrophilic environments, WPI films which contain higher numbers of hydrophilic residues than zein or wheat protein films reportedly produce larger inhibition zones against L. monocytogenes. 71 The solubility, stability and biological activity of nisin are highly pH dependent. Biological activity and stability decrease sharply at higher pH values (Barranova et al., 1976; Liu and Hansen, 1990). Nisin is insoluble at neutral and alkaline conditions (Hirsch, 1951; Hurst, 1981; Liu and Hansen, 1990) and more stable to heat at low rather than high pH (Tramer, 1964). Stability of nisin to heat and storage depends on pH, chemical composition of the solution and the temperature. Nisin inhibits the majority of Gram-positive bacteria (Table 1.3) (Mattick and Hirsch, 1947; Ogden and Tubb, 1985). The effectiveness of nisin against various bacteria in meat and meat products has been investigated. Nisin can reportedly inhibit the growth of Bacillus licheniformis (Bell and DeLack, 1986), CIostridium sporogenes (Rayman et al., 1981) and lactic acid bacteria in cured and fermented meat products (Collins- Thompson et al., 1985). When combined with chelating agents, several reports suggest that nisin may be effective against Gram-negative bacteria in food (Shelef et al., 1995; Mahmoud and El- Baradei, 1998; Wells et al., 1998; Boziaris and Adams, 1999). While the outer membrane of Gram-negative bacteria prevents penetration of nisin into the cytoplasmic membrane (Stevens et al., 1991; Schved et al. 1994), chelating agents such as EDTA as well as sublethal heating, low pH, or freezing disrupt the permeability barrier thus increasing the sensitivity of Gram-negative organisms such as Salmonella enterica, S. Typhimurium, E. coli, and Shigella to nisin (Stevens et al., 1991; Kalchayanand et al., 1992; Ganzle et al., 1999). EDTA (0.3mM) and nisin (50ug), in combination with heating (0.3 min) at 37°C, reduced the numbers of Erwinia carotovora, E. chrysanthemi, Pseudomonas fluorescens, and P. viridiflava in trypticase soy broth by 2 logs, and at 49°C by 3 logs compared to the 72 unheated control at 25°C. Nisin activity against some Gram-negative bacteria can be enhanced by trisodium phosphate (TSP). Following a 10 minute exposure to sublethal concentrations of TSP (0.5 to 5 mM), Cameiro de Melo et al. (1998) reported that cell suspensions of Campylobacter jejuni, Escherichia coli, Pseudomonas fluorescens and Salmonella enteritidis showed greatly enhanced susceptibility to nisin (luM). Under optimal conditions at 37°C, viable counts decreased up to 6 logs after 30 minutes. Pulsed- electric field treatment, which induced sublethal injury, also enhanced the bactericidal action of nisin against Bacillus cereus and Escherichia coli (Pol et al., 2000; Terebiznik etaL,2000) Activity of nisin is strongly influenced by various environmental factors including pH, temperature and NaCl. Benkerroum and Sandini (1988) showed that L. monocytogenes was more sensitive to nisin in tryptose soy broth at pH < 5.94 than at pH 7.0. Increased antimicrobial activity of nisin against various pathogens at low pH was also reported by Harris et al.(1991) and Tatini (1992) with low pH increasing the sensitivity of E. coli and S. enterica to nisin (Ganzle et al., 1999). Tolerance of vegetative B. cereus cells to nisin increased as the pH of the broth increased from 5.53 to 6.01 and 6.57. Nisin activity is also directly related to incubation temperature. According to Tatini (1992), the minimum concentration of nisin needed to inhibit grth of L. monocytogenes was two to three times higher at 35 than at 4°C. Similarly, 5 and 50 ug/ml were needed to inhibit outgrowth of B. cereus spores at 8 and 15°C, respectively. Harris et a1 and Ganzle et al. (1999) also reported that addition of 2 to 3% sodium chloride enhanced the activity of nisin against L. monocytogenes, S. Typhimurium, and E. coli (< 10 pg /ml) in laboratory media. 73 1.2.7.2. Pediocin Pediocin is produced by Pediococcus acidilactici. Among the pediocins isolated from different strains, only pediocin PA1 (Pediococcus acidilactici PAC 1.0) and pediocin AcH (P. acidilactici LB42-923) have been well characterized. Pediocin AcH, having a molecular mass of about 2700 Da (Bhunia et al., 1987), contains 62 amino acids and two disulfide bonds. The pediocins are another commonly studied group of bacteriocins for edible film use, due to their wide spectrum of antimicrobial activity and their effectiveness over a wide range of pH and temperature. Antimicrobial activity of pediocin is retained at 100°C, reduced at 121°C and most evident at pH 4-7 with substantial losses at pH <3 or > 9. Pediocin remains active following treatment with lipase, phospholipase C, lysozyme, DNase or RNase, but its activity is destroyed by protease, papain and a-chymotrypsin (Gonzalez and Konka, 1987). While pediocin activity was unaffected during 6 months of frozen storage, over 50% of the activity was lost after 12 weeks at ambient temperature (Ray, 1996). Pediocin is inhibitory to a broad range of bacteria (Table 1.3). Inhibition of L. monocytogenes with pediocin AcH in various foods has been well documented. L. monocytogenes numbers decreased by 0.74 log within 2 h in beef wieners that were inoculated with a pediocin-producing strain of P. acidilactici (Yousef et al., 1999). Numbers of L. monocytogenes increased initially and then markedly decreased with pediocin production by P. acidilactici with pediocin activity also detected in the wiener exudate during the last stage of P. acidilactici growth. Degnan et al. (1992) investigated 74 survival of L. monocytogenes in temperature-abused, vacuum-packed wieners which contained pediocin- and non-pediocin-producing P. acidilactici. Listeria populations increased 3.2 logs after 8 days of storage at 25°C in the absence of any Pediococcus strain, remained unchanged in the presence of the non-pediocin-producing strain, and decreased by 2.7 logs using the pediocin-producing strain. Inhibition of L. monocytogenes by pediocin-producing starter cultures was reported during fermentation of dry and semi-dry sausage (Berry et al., 1991). Using a pediocin-producing strain of Pedioccocus and a nonbacteriocinogenic starter, L. monocytogenes populations decreased ~2 logs and <1 log respectively, during fermentation of semi-dry sausage. Antilisterial activity from a pediocin-producing strain also was demonstrated during manufacture of turkey summer sausage (Luchansky et al., 1992) with pediocin and non-pediocin—producing strains decreasing populations of L. monocytogenes by 3.4 and 0.9 log CFU/g, respectively, after 12 h of fermentation. Several investigators observed potential benefits of adding pediocin to dairy products. Pucci et al. (1988) showed that incorporating a crude extract of pediocin PAl into cottage cheese, cheese sauce and half-and-half decreased the numbers of L. monocytogenes in all inoculated samples after the first day of refrigerated storage. The effectiveness of pediocin AcH in controlling L. monocytogenes in dry milk, ice cream, and cottage cheese was also studied by Motlagh et al. (1992), who found that pediocin decreased numbers of L. monocytogenes by ~l log in sterile ground beef, sausage mix and cottage cheese after 1 h of storage at 4°C. Pediocin (1350 AU/ml) also reduced Listeria populations 2 to 4 logs in dry milk after 1 day of storage at 4°C. 75 In addition to meat and dairy products, other foods may benefit from biopreservation with pediocin and pediocin-producing strains. Choi and Beuchat (1994) added a bacteriocin extract from P. acidilactici M to kimchi during fermentation. This treatment immediately reduced numbers of L. monocytogenes in the inoculated product and inhibited growth of the organism during 16 days of fermentation. 1.2.7.3. Mode of Action of Bacteriocins Bacteriocins disrupt membrane activity by pore formation and may have additional effects on electron transfer chain components. Binding of bacteriocin to the cell membrane results in the formation of peptide pre-aggregates that induce conformational changes and pore formation (Sahl, 1991; Benz et al., 1991; F reund et al., 1991) which in turn decreases the membrane potential and proton gradient across the membrane. For example, Bruno et a1 (1992) showed that the addition of 2.5 pg of nisin/ml completely dissipated both components of the proton motive force in L. monocytogenes. Okereke and Montville (1992) also found that nisin inhibited the growth C. sporogenes PA 3679 by dissipating the proton motive force and possibly inhibiting ATPase activity, thereby exhausting intracellular ATP. Like other bacteriocins, pediocin’s bactericidal mode of action against Gram- positive bacteria involves adsorption to lipoteichoic acid receptors on the cell surface followed by penetration through the wall and contact with the cytoplasmic membrane (Gonzales and Kunka, 1987; Bhunia et al., 1991). Gram-negative bacteria do not possess lipoteichoic acid and are therefore unable to adsorb pediocin. Pediocin produces a 76 conformational change in the cell wall of Gram-positive bacteria which interferes with proper barrier function (Jack et al., 1995) Sublethally stressed Gram-negative bacteria can become sensitive to bacteriocins even though the molecule is not adsorbed by the cell. In such cells, the cell-surface barrier function is impaired, thus allowing bacteriocins to pass through the impaired wall, come in contact with the cytoplasmic membrane, and destabilize its function leading to cell death. Pediocin is also active against some Bacillus and Clostridium species (Kalchayanand, 1990), preventing outgrowth of germinated spores. 1.2.8. Natural Antimicrobials 1.2.8.1. Lysozyme Lysozyme, another popular choice for the production of antimicrobial films, is an enzyme comprised of 129 amino acids, crosslinked by four disulfide bonds. Dried egg white, the commercial source for lysozyme, contains about 3.5% lysozyme. However, lysozyme is also present in mammalian milk, tears and other secretions, insects and fish. It is heat stable (100°C) at pH <5.3, but is inactivated at lower temperatures when the pH is increased (Smolelis and Hartsell, 1952; Matsuka et al., 1966). This remarkable stability is attributed to the four disulfide bonds present in this small protein. Plasticizers used in edible film such as glycerol and sorbitol help stabilize lysozyme against heat through hydrophobic interactions that reduce the complete transfer of hydrophobic groups from an aqueous to a non-polar environment (Yashitake and Shininichiro, 1977; Back et al., 1979). Therefore, lysozyme is highly suited for heat processed films such as heat-pressed corn zein-based films (Dawson et al., 1997; Padget et al., 1998). 77 Lysozyme is most active against Gram-positive bacteria, most likely because peptidoglycan of the cell wall is more exposed. Penetration through the outer membrane can be accomplished by using either chelating agents or osmotic shock. The outer membrane of Gram-negative organisms contains divalent cations that stabilize lipopolysaccharide within the membrane. Chelating agents remove cations and thus increase cell permeability of Gram-negative bacteria to lysozyme (Hancock, 1984; Samuelson et al., 1985; Hughey et al., 1989; Payne et al., 1994; Razavi- Rohani and Griffits, 1996). Vedmina et al. (1979) tested the sensitivity of lysozyme against 476 strains of Gram-negative bacteria and found high resistance to lysozyme in Vibrio cholera El Tor and Pseudomonas. Organisms exhibiting varying sensitivity include Aeromonas and enteropathogenic Escherichia coli. 1.2.8.1.1. Mode of Action Lysozyme catalyzes the hydrolysis of B-1,4 glycosidic bonds between N- acetylmuramic acid and N-acetylglucosamine of peptidoglycan in bacterial cell walls, which leads to cell wall degradation and lysis in hypotonic solutions. Depending on the enzyme source, chitin and certain esters are also susceptible to lysozyme. 1.2.8.2. Spices, Herbs, and Essential Oils Essential oils are responsible for the odor, aroma, and flavor of spices and herbs. These compounds can be added to edible film to modify flavor, aroma and odor as well as antimicrobial properties. Films containing these ethanol-soluble compounds will also 78 show activity against both Gram-negative and Gram-positive bacteria, with Gram- positive organisms being more sensitive (Beuchat and Golden, 1989). Ting and Deibel (1992) reported that addition of cloves, oregano, sage, rosemary and nutmeg to TSB inhibited growth of L. monocytogenes at 24°C. According to Bahk et al.(1981), addition of 5% cinnamon also inhibited growth of L. monocytogenes in TB at 4°C; however, 0.5% ginger, garlic, onion and mustard as well as ginseng, saponin and mulberry extract were noninhibitory. Activity of garlic powder and cloves was also investigated against S. Typhimurium and S. aureus by Teerapom (1995). Results showed that 12% garlic powder and 0.4% clove oil led to a > 5-log reduction of S. Typhimurium, whereas 0.4% clove oil and a much higher concentration of garlic powder (58%) were needed for similar inhibition to S. aureus. Antimicrobial activities have been known among various plant essential oils for centuries. However, their use as food additives is limited by their strong flavor. These extracts contain mostly phenolic compounds such as abietane diterpenes (Moujir et a1, 1993), camosol and ursolic acid (Collins and Charles, 1987) which are presumably responsible for their antimicrobial action. Antimicrobial activity and composition of the essential oils from Micromeria cristata subsp. phrygia was investigated against Gram- negative and Gram-positive bacteria by Tabanca et al. (2001). The essential oils showed inhibitory activity against E. coli, S. Typhimurium, S. aureus, P. aeruginosa, E. aerogenes, P. vulgarus, and C. albicans with the major oil compenent being bomeol (39%). 79 Ethanol extracts of rosemary (Oxy’less) (100 mg/ml) effectively killed L. monocytogenes, Leuconostoc mesenteroides, S. aureus, Streptococcus mutanss and Bacillus cereus, whereas no activity was observed with Gram-negative bacteria (e.g. E. coli, Salmonella enteriditis, Erwinia carotovora) or yeasts (e.g. Rhodotorula glut, Cryptococcus laurentii) (Del Campo et al., 2000). Elgavyar et a1. (2000) evaluated the antimicrobial activity of essential oils from parsley, anise, angelica, carrot, cardamom, coriander, dill weed, fennel, oregano and rosemary using a disc diffusion assay. Essential oils from cardoman, coriander, oregano, basil, celery and parsley inhibited growth of S. Typhimurium, Y. enterocolitica, E. coli, S. aureus, L. monocytogenes and Aspergillus niger. However, essential oils from dill weed, fennel, carrot, rosemary and anise were noninhibitory to L. monocytogenes on standard methods agar (SMA). Inhibition zones were observed for all essential oils except those from carrots against S. Typhimurium, E. coli and S. aureus on SMA. The effectiveness of some essential oils is enhanced at lower temperatures and lower pH values. Ting and Deibel (1992) showed that refrigeration temperatures increased the inhibitory effect of sage, but not that of cloves or oregano against L. monocytogenes. In addition, essential oils from thyme and oregano were ten-fold more effective against B. cereus at 8 compared to 30°C (U ltee et al., 1998). Surface-active agents in foods such as dairy cream and albumin also were shown to reduce the inhibitory effect of essential oils (Davidson, 1993; Tassou and Nychas, 1994; Juven et al., 1994; Del Campo et al., 2000). For example, Juven et al. (1994) showed that serum albumin (9 mg/ml) eliminated the antimicrobial activity of thymol against S. Typhimurium. 80 Essential oils alter the fatty acid composition and phospholipid contact of the cell membrane, which results in the release of cellular constituents that interfere with metabolism, disrupting both electron transport and nutrient uptake. These essential oils also have the ability to disrupt various enzyme systems, including those involved in cellular energy and structural synthesis 1.2.8.3. Lactoferrin Lactoferrin (lactotransferrin), an iron binding glycoprotein, is present in bovine milk and can bind two iron atoms per molecule (Ashton and Busta, 1968; Reiter, 1983). This protein was shown to effectively inhibit the growth of several bacteria including Bacillus subtilis, B. stearothermophilus, L. monocytogenes, E. coli, Microccoccus spp, and Klebsiella spp. (Oram and Reiter, 1968; Reiter, 1978; Payne et al., 1989; Payne et al., 1990; Korhonen, 1978). Payne et al (1990) showed that lacroferrin had a bacteristatic effect against L. monocytogenes. At 30 mg/ml Apo-lactoferrin (iron-free lactoferrin) reduced Listeria population 10 fold. At 2.5 mg/ml, lactoferrin had no activity against S. Typhimurium or P. fluorescens with only minimal activity against E. coli 0157:H7 and L. monocytogenes (Payne et al., 1994). Some Gram-negative bacteria may be lactoferrin resistant due to the presence of siderophores that aid in adaption to low-iron environments (Crichton and Chaloteux-Wauters, 1987). In addition, bacteria with low iron requirements such as lactic acid bacteria would not be adversely impacted by lactoferrin (Reiter and Orarn, 1986). Ashton and Busta (1968) reported that the inhibitory activity of lactoferrin was likely due to chelation of iron as well as calcium and magnesium ions. Inhibition of L. 81 monocytogenes by lactoferrin is directly related to iron availability in the medium, with L. monocytogenes surviving best in iron-rich media (Payne et al. 1989). However, Arnold et al. (1982) showed that lactoferrin inhibited many bacteria in an iron-rich environment. Lactoferrin caused the release of anionic polysaccharides from the outer membrane of E. coli by chelation of cations that stabilize lipopolysaccharides. Thus, lactoferrin may increase outer membrane permeability to hydrophobic compounds. Lactoferricin B, the active region of lactoferrin, was isolated by acid-pepsin hydrolysis from the N-terminal region of the molecule (Bellamy et al., 1992). Lactoferricin contains 25 amino acid residues. Bellamy et al. (1992) and Jones et al. (1994) determined that lactoferricin was inhibitory to bacteria at concentrations of 0.3 to 150 ug/ml (Table 1.3). Pseudomonas fluorescens, Enterococcus faecalis and Bifidobacterium biflidum strains were highly resistant to this peptide. These results confirm and expand on earlier inhibition studies with lactoferricin B (Tomita et al., 1992;Wakabayashi et al., 1992). According to Kumar et al. (1999), lactoferricin B (100 ug/g) also reduced E. coli 0157:H7 populations by 0.8 log CF U/g in ground beef. While the mode of action of lactoferricin has not been fully elucidated, it is thought to alter the permeability of the membrane because of its cationic feature (Conner, 1993; Jones et al., 1994). 1.2.8.3. Liquid Smoke Liquid smoke is a solution of natural wood smoke flavors prepared by burning a wood (e.g. hickory, maple) and capturing the natural smoke flavors in water. 82 Alternatively, liquid smoke can be derived from the destructive distillation of a wood - i.e the breakdown or cracking of the wood fibers into chemical constituents which are distilled out of the wood residue. Most liquid smokes are very acidic, although some partially neutralized liquid smokes are also available. While some commercially available smoke can be used at filll strength, others are normally diluted in water or another appropriate diluent. Commercial liquid smoke products used in processed meats, sausages and cheeses contain phenols and acetic acid, which are bactericidal at relatively low concentrations. Liquid smoke can inactivate common food-borne pathogens including E. coli, Salmonella, S. aureus, and L. monocytogenes. Several investigators examined the potential of commercial liquid smokes to inactivate L. monocytogenes in culture media and meat products. Wendorff (1989) found that liquid smoke compounds (0.5%) were listericidal in both phosphate buffer and processed meats. Faith et al. (1992) also reported that L. monocytogenes was inactivated by addition of 0.2 and 0.6% liquid smoke to wiener exudate with D-values of 36 and 4.5 h, respectively; however, numbers of Listeria in the untreated exudate increased to 8 logs after 3 days at 25°C. Among 11 individual phenols tested, only isoeugenol showed antilisterial activity in tryptose broth (TB) during incubation at 37°C. In the presence of isoeugenol (100 ppm), greater inhibition of the pathogen was observed in TB acidified with acetic acid to pH 5.8 compared to pH 7.0. In addition to meat products, antimicrobial activity of liquid smoke was also ‘ evaluated against molds on Cheddar cheese (Wendorf et al., 1993). Liquid smoke reportedly inhibited the growth of Aspergillus oryzae and increased the lag phase of Penicillium camemberti and P. roqueforti. Among 8 phenolic compounds tested, only 83 isoeugenol retarded the growth of all three molds. P. camemberti was slightly inhibited by m-cresol and p-cresol, while A. oryzae was inhibited by guaiacol, 4-methylguaiacol, m-cresol and p-cresol. Based on these findings, liquid smokes which possess antimicrobial, antioxidant, color, and flavor properties, have the potential to become attractive edible film additives. Thus a, incorporation of liquid smoke has only been studied for edible collagen casings. Liquid smoke was introduced into the acid-swollen collagen mass before extrusion as a casing or film (Miller, 1975). Since liquid smoke is generally very acidic (pH 2.5 or less), it is compatible with the gel system and, in fact, can replace a portion of the acid normally added to induce swelling. The resultant edible collagen casings with uniformly dispersed liquid smoke reportedly had increased tensile strength and improved film clarity. 1.2.9. Curing Agents 1.2.9.1. Sodium Chloride Sodium chloride (NaCl), recognized as a food preservative since ancient times, can be used alone or in combination with other preservation techniques such as pasteurization or fermentation. Bacterial food-borne pathogens are generally susceptible to NaCl, except S. aureus, which can grow at low water activities (0.83-0.86) (McLean et al., 1968). Another salt-tolerant pathogen is L. monocytogenes, which can grow at concentrations up to 10% NaCl and survive for long periods of time in saturated brine solutions. Yeasts and molds are also more tolerant to low water activity than bacteria. with xerotolerant fungi growing at a water activity value as low as 0.61 (Corry, 1987). 84 Sodium chloride inhibits microbial growth by its plasmolytic effect. The antimicrobial activity of sodium chloride is related to its ability to reduce water activity in the food. Microbial cells lose water when the water activity of the external environment is reduced, which results in growth inhibition or possible death (Sperber, 1983). In addition to this osmotic effect, NaCl limits oxygen solubility, alters pH, and is itself toxic for microbial cells (Banward, 1979). Incorporation of NaCl into protein-based films as an antimicrobial agent is of limited use since physical properties of protein films decrease with increasing ionic strength of the film solution. At high ionic strength, proteins aggregate to form turbid opaque gels possessing a high-water holding capacity (Doi and Kitabastake, 1997). 1.2.9.2. Nitrite Sodium nitrite (NaNOz) and potassium nitrite (KNOZ) are primarily used to inhibit C. botulinium growth and toxin production in cured meats. Nitrite inhibits bacterial sporeforrners by preventing outgrowth of the germinated spores (Cook and Pierson, 1983; Duncan and Foster, 1968). Nitrite effectiveness also depends on other environmental factors. For example, nitrite is more active at low pH and anaerobic conditions (Buchanan and Solberg, 1972; Woods et al., 1989). When used as reducing agents, ascorbate and isoascorbate enhance the antibotulininal action of nitrite (Tompkin et al., 1978; Roberts et al., 1991). Nitrite is also inhibitory to other bacteria including C. perfiingens, E. coli, Acromobacter, Enterobacter, Flavobacterium, Micrococcus, and Pseudemonas spp. at 200 ug/g (Gibson and Roberts, 1986b). Growth of L. monocytogenes was inhibited for 40 days at 5°C by treating smoked salmon with 200 85 ppm sodium nitrite (Pelroy et al., 1994). However, Gibson and Roberts (1986a, 1986b) found that enteropathogenic E. coli, Salmonella spp., fecal streptococci, Lactobacillus (Castellani and Niven, 1955) and Bacillus (Grever, 1974) were resistant to 400 ug/g nitrite when used with 6% salt. Incorporation of nitrite into edible films has not yet been studied, even though nitrite appears to be a suitable antimicrobial agent for antimicrobial edible film production. In this regard, application of the films containing nitrite on ready-to-eat-meat products may provide a possible solution for preventing growth of L. monocytogenes and spoilage bacteria that may contaminate such products after processing, with the potential benefit of also improving surface color. Nitrite works by inactivating various enzymes or enzyme systems. Using vegetative cells of C. sporogenes, Woods et al. (1981) showed that pyruvate-ferrodoxin oxidoreductose and ferrodoxin were susceptible to nitrite. Inhibition of these enzymes causes a reduction in intracellular ATP and excretion of pyruvate. The phosphoroclastic system of C. botulinium and C. pasteurianum is also inhibited by nitrite (Woods and Wood, 1982; Carpenter et al., 1987). Rowe et al. (1979) showed that the other antimicrobial mechanism for nitrite was inhibition of active transport, oxygen uptake, and oxidative phosphorylation by oxidizing ferrous iron from an electron carrier (Rowe et al., 1979, Yang, 1985) 86 1.3. ANTIMICROBIAL EDIBLE FILMS Various antimicrobial edible films have been developed to control the growth of spoilage and pathogenic microorganisms that may contaminate the surface of foods after processing. In most solid foods, contamination and microbial growth occurs on the food surface, which leads to a reduction in product shelf-life. Edible films containing various antimicrobials such as benzoic acid, sorbic acid, propionic acid, lactic acid, nisin, and lysozyme have been investigated to retard the growth of bacteria, yeasts and molds on different product surfaces. The primary advantage of antimicrobial edible films is that the antimicrobial agents in these films can be specifically targeted to post-processing contaminants on the food surface, with the diffusion rate of the antimicrobial into the product partially controlled by levels incorporated into the film. Guilbert and his research group (Giannakopoulos and Guilbert, 1986; Guilbert et al., 1985; Guilbert et al., 1988) evaluated diffusivity of sorbic acid from casein films in different model systems. Using a cell membrane separated by casein film, their data showed that low temperatures (10°C) decreased the diffusivity of sorbic acid; however, lower water activity had no effect. They theorized that at higher levels, increased networking within the gel acted to restrict the movement of sorbic acid. Vojdani and Torres (1989a, b, 1990) also evaluated the permeability of several polysaccharide-based films prepared both with and without various combinations of lipids and potassium sorbate. Using permeability cells methylcellulose-palmitic acid films appeared to be most promising, with permeability of the film to sorbic acid decreasing from 10'8 to 10'10 mg/sec.cm2 as pH increased from 3 to 7 and aw decreased from 0.8 to 0.65 (Rico-Pena and Torres, 1991). 87 Chitosan, like other polysaccharides, forms a strong film that can carry a high amount of antimicrobials. Chitosan also has been reported to be a good choice for antimicrobial films due to its good film-forming properties, ability to adsorb nutrients used by bacteria (Darmadji and Izumimoto, 1994), capacity to bind water and inhibit various bacterial enzyme systems (Young et al., 1982). However, neutralized chitosan alone had no effect on bacterial growth when it was applied to the surface of meat products. Antimicrobial chitosan films were subsequently prepared by dissolving chitosan in hydrochloric, formic, acetic, lactic and citric acid solutions (Begin and Calstren, 1999). Films made from hydrochloric, formic and acetic acid were hard and brittle, whereas those containing lactic or citric acid were soft and could be stretched. The same research group designed antimicrobial chitosan films containing acetic acid or propionic acid, with or without the addition of lauric acid or cinnemaldehyde, to improve refrigerated shelf-life of vacuum-packaged processed meats (Ouatarra et al., 2000a). They indicated that film application delayed or completely inhibited enteric bacteria, Lactobacillus sakei and Serratia liquefaciens, on meat products. Films containing propionic acid were more effective than films containing acetic acid for reducing growth of L. sakei with the opposite observed for S. liquefaciens. Diffusion of acetic acid from the film matrix was limited by addition of lauric acid, with 2-22% of acetic acid remaining in the chitosan after 168 h of storage at 4°C. However, propionic acid almost totally diffused from the film after 48 h of storage. Release of organic acids from chitosan film is dependent on many factors, including electrostatic interactions between the acid and polymer chains (Demarger-Andre and Domad, 1994), ionic osmosis and structural changes induced by presence of the acid (Narisawa et a1, 1996). Ouattara et al. (2000b) 88 also tested the impact of temperature (4 to 24°C) and pH (5.7 to 7.0) on diffusion of acetic and propionic acid from chitosan films immersed in water. Whereas diffusion was unaffected by pH, a decrease in temperature from 24 to 4°C decreased the diffusion coefficients for acetic and propionic acid from 2.59 x 10'12 mZ/sec to 1.19 x 10"2 mz/sec and 1.87 x 10'12 mz/sec to 0.91 x 10'12 mz/sec, respectively. The dependency of diffusion on temperature is explained by effects on solubility of the diffusing molecule, the nature of adhesive forces at interfaces, and molecular mobility (Vojdani and Torres, 1990; Myint et al., 1996). Ouattara et al. (2000b) stated that addition of lauric acid (1%) or essential oils (0.5%) (cinnemaldehyde or eugenol) decreased the diffirsion of propionic acid, since these additives increased film hydrophobicity and modified the pore construction/blind porosity of the film, thereby impairing water uptake and molecular transformation. Besides propionic acid, fatty acids, and essential oils, sorbate and benzoate also have been tested in methylcellulose and chitosan films. For example, Chen et al. (1996) developed antimicrobial methylcellulose, chitosan and methylcellulosezchitosan films (3:2) containing 2, 4, or 5% sodium benzoate or potassium sorbate. Methylcellulose films containing 2% sorbate or benzoate yielded clear inhibition zones for Rhotorula rubra and Penicillium notatum on potato dextrose agar. Chitosan films containing 2% sorbate or benzoate yielded no zones of inhibition, since the high affinity between chitosan and the preservatives prevented diffusion of the antimicrobials. Incorporation of both potassium sorbate and sodium benzoate into methylcellulose/chitosan films did not change the tensile strength or percent elongation. In a glycerol-water model system (aw=0.8), 40% and 50-60% of both antimicrobial agents were released from the films after 6 h at 4 and 89 25°C, respectively. Another research group from Taiwan evaluated the antimicrobial activity of methylcellulose coatings containing benzoic and palmitic or stearic acid against two osmophilic yeasts (Zygosaccharomyces rouxii and Zygosaccharomyces mellis) on Taiwanese-style fruit preserves made from plums (Chen et al. 1999). Coatings containing 50-100 ug/g benzoic acid inhibited Z. rouxii and Z mellis at room temperature, with sensory characteristics of preserves such as flavor, texture, appearance and overall acceptability not affected by the coating. Protein-based edible films are also very good carriers of food additives, including antimicrobial and flavor agents, due to their favorable encapsulated nature. Zein films have been used in conjunction with potassium sorbate to control surface microbial growth. The diffusion barrier properties of zein films were confirmed in microbial tests using a model food system and S. aureus as the challenge organism. A reduced preservative diffusion rate due to barrier properties of zein films was identified as the mechanism for product shelf-life enhancement (Torres et al., 1985; Torres and Karel, 1985). The diffusion of sorbic acid from various wheat gluten films into a model food was also measured and modeled. When Penicillium notatum was used as the test organism, simple gluten-based films had no fungicidal effect. However, the gluten/lipid- based films showed strong sorbic acid retention and marked fungicidal activity at 30 and 40°C, delaying P. notatum growth for more than 21 days. Similarly, antimicrobial soy and corn protein-based films were developed by Dawson and his lab group at Clemson University. When prepared to contain nisin and lysozyme, these films were inhibitory to Gram-positive bacteria both on solid and in liquid media (Orr et al., 1996a, b; Dawson et al., 1995; Dawson et al., 1997; Padget et al., 90 1995; Padget et al., 1998). Addition of EDTA to these films also led to inhibition of Gram-negative organisms (Orr et al., 1996a). Modifying the water permeability by incorporating short chain fatty acids (lauric acid) reduced the effectiveness of nisin on solid media, whereas films with lauric acid were as effective as nisin against Gram- positive bacteria in liquid media (Dawson et al., 1997). Padget et al. (1998) incorporated nisin and lysozyme into soy protein and corn zein films using the heat press and casting methods. Both antimicrobial films containing lysozyme (10 to 133 mg/g film) or nisin (0.1 to 6.0 mg/g film) inhibited Lactobacillus plantarum on MRS media. Orr et al. (1998) found that com zein films containing 150 mg of nisin reduced L. monocytogenes populations 1.3 to 2.2 logs in milk after 72 h at 4°C, with no inhibition observed with nisin free-films. Use of zein film coatings containing nisin (1000 IU/g) also reportedly reduced L. monocytogenes populations 1 to 3 logs on ready-to-eat chicken during 30 days of refi'igerated storage (J anes et al., 1999). Rodrigues and Han (2000) reported that when incorporated into WPI films, lysozyme and nisin effectively inhibited Brochotrix thermosphacta but not L. monocytogenes. Addition of EDTA increased the inhibitory effect of these films against E. coli and L. monocytogenes on trypticase soy agar. Other antimicrobial WPI films containing sorbic acid (SA) or p-aminobenzoic acid (PABA) were developed by Cagri et al. (2001). Both of these films reportedly inhibited the growth of L. monocytogenes, E. coli 0157:H7, and S. Typhimurium on TSAYE agar. Subsequently, these films were tested between beef bologna and summer sausage slices that were surface inoculated with the same pathogens at a level of 106 CFU/g (Cagri et al., 2002a). WPI films containing SA or PABA decreased Listeria, E. coli and S. Typhimurium populations 3.4-4.1, 3.1- 91 3.6, and 3.1-4.1 logs on bologna and sausage after 21 days of aerobic storage at 4°C, respectively. Growth of mesophilic aerobic bacteria (MAB), lactic acid bacteria (LAB) and mold/yeast on slices was also inhibited with WPI films containing SA or PABA compared to antimicrobial-free control films. In the same study, film tensile strength decreased while % elongation remained unchanged following 72 h of product contact. Subsequently, heat—sealed WPI casings containing SA, PABA, or SAzPABA (1:1) were developed for hot dog manufacture, with these casings compared to commercial collagen and natural casings (Cagri et al., 2002b). WPI casings containing PABA inhibited L. monocytogenes grth on hot dogs during 42 days of refrigerated storage; however, films with SA or SAzPABA were less effective. Sensory (texture, flavor, juiciness, overall acceptability), chemical (TBA, pH, moisture, fat, protein), physical (purge, color), and mechanical (shear force) characteristics of hot dogs with WPI casings containing PABA were comparable to hot dogs prepared with collagen and natural casings. Consequently, WPI casings containing PABA may eventually prove useful in minimizing risk of Listeria growth on hot dogs. Ko et al. (2001) also tested the antilisterial activity of nisin (200 — 8000 IU/g film) when incorporated into WPI, soy protein isolate, egg albumin, and wheat gluten films. All of these films inhibited Listeria, with greatest activity observed at low pH (2.0 or 3.0). WPI films containing nisin were most effective against L. monocytogenes due to their increased hydrophobicity, with their mechanical properties also remaining unchanged by the addition of nisin. Various edible antimicrobial films have also found use in vegetables to prevent the grth of spoilage and pathogenic bacteria. Zhuang et al. (1996) investigated use of 92 antimicrobial cellulose-based edible films containing citric acid, acetic acid, or sorbic acid (0.2 to 0.6%) on tomatoes inoculated with Salmonella montevideo. Although coating with a hydroxypropyl methylcellulose (HPMC) solution reduced Salmonella populations by 4.5 logs on the surface of tomatoes; a reduction of only 2.0 logs was achieved in core tissue. S. montevideo cells penetrating into the core tissue when tomatoes were dipped in the 30°C bacterial suspension were likely protected from inactivation during coating (Zhuang et al.,l995). Among the antimicrobials tested in HPMC films, only 0.4% sorbic acid enhanced the inactivation of S. montevideo (~1.0 log) on the surface of tomatoes. However, tomatoes coated in HPMC containing 0.4% sorbic acid appeared chalky, less firm and exhibited color changes that may limit possible commercial applications. Most recently, Cha et al. (2001) reported that nisin-containing K-carrageenan, MC and HPMC films prepared by either a heat press or casting method were inhibitory to Micrococcus luteus in agar well diffusion assay. Nisin reportedly diffused faster from MC than from K-carrageenan or HPMC films. Not surprisingly, the heat-pressed films had lower antimicrobial activity than the cast films. In conclusion, the use of edible films or coatings on various food products continues to expand. The numerous benefits of edible films as carriers of antimicrobial as well as flavors, antioxidants and color agents justifies further research in this field. Edible films containing antimicrobial agents have been shown to effectively inhibit both pathogenic and spoilage organisms on a wide variety of ready-to-eat food surfaces. These films have the ability to control the diffusion rate of the antimicrobial agent and also serve as a good barrier against oxygen and water vapor transmission. The antimicrobial edible films appeared to be a possible solution to reduce the incidence of pathogens 93 especially L. monocytogenes on food surfaces. However, only some of these antimicrobial edible films received commercial acceptance. Further research is needed for development of feasible application methods for the industry such as extrusion of film solution into tubular shaped film in various sizes for processed meat products. 94 CHAPTER 2. ANTIMICROBIAL, MECHANICAL, AND MOISTURE BARRIER PROPERTIES OF LOW PH WHEY PROTEIN- BASED EDIBLE FILMS CONTAINING P- AMINOBENZOIC OR SORBIC ACIDS Cagri, A., Ustunol, Z., Ryser, E.T. Journal of Food Science. 2001. 66(6) : 865-870 95 2.1. ABSTRACT Low pH (5.2) whey protein isolate—based edible films containing p-aminobenzoic acid (PABA) or sorbic acid (SA) were developed and assessed for inhibition of Listeria monocytogenes, Escherichia coli 0157:H7, and Salmonella Typhimurium DT104 in a disc diffusion assay. Water vapor permeability (WVP), tensile strength (TS), and percent elongation (% E) were also determined. Using 1.5% PABA and SA, average inhibition zone diameters were 21.8, 14.6, 13.9 and 26.7, 10.5, 9.7 mm for L. monocytogenes, E. coli 0157:H7 and S. Typhimurium DT104, respectively. Three strains of S. Typhimurium DT104 were resistant to 0.5% SA. Addition of PABA and SA increased %E, but decreased TS. WVP was not affected by 0.5 and 0.75% SA; however, PABA increased WVP. 2.2. INTRODUCTION Microbial stability of a food surface is a major determinant of product quality and safety during storage and distribution since most Class I product recalls in the US. result from post-processing contamination during subsequent handling and packaging. In December 1998, new food safety concerns were raised when consumption of hot dogs was traced to over 100 cases of listeriosis, including 21 fatalities in 22 states (CDC, 1999). A nationwide recall was subsequently issued ‘for 35 million pounds of contaminated product. This pathogen continues to threaten the processed meat industry. Sixty-three of 97 microbiologically related Class I recalls issued from January 1999 to October 2000 involved a total of more than 3.5 million pounds of cooked/ready-to-eat 96 meats contaminated with Listeria monocoytogenes (USDA-FSIS, 2000). Two additional foodbome pathogens, namely Escherichia coli 0157:H7 and Salmonella Typhimurium DT104 are also raising considerable public health concerns. E. coli 0157:H7 has been responsible for many widely publicized outbreaks involving ground beef (Bell et al., 1994), fermented meat products (Tilden et al., 1996) and fresh produce (Besser et al., 1993). In addition, over 30 million pounds of raw ground beef have been recalled since 1995 due to E. coli 0157:H7 contamination (USDA-FSIS, 2000). S. Typhimurium DT104, a multiantibiotic resistant strain, is also emerging as a serious foodbome pathogen of public health concern with 103 of 306 (34%) S. Typhimurium isolates serotyped at CDC (Centers for Disease Control and Prevention) resistant to arnpicilin, chloramphenical, streptomycin, sulfonomides and tetracyclines (Glynn et al. 1998). Incorporating antimicrobial compounds into edible films or coatings provides a novel means for enhancing the safety and shelf life of ready-to-eat foods. Dawson et al. (1997) and Padget et al. (1998) used nisin and pediocin in soy protein and corn zein films to inhibit Lactobacillus plantarum and E. coli on laboratory media. Antimicrobial edible films are receiving attention as a potential pathogen intervention strategy for various muscle foods. Sirugusa and Dickson (1993) demonstrated that calcium alginate coatings containing organic acids were marginally effective on beef carcasses, reducing levels of L. monocytogenes, S. Typhimurium and E. coli 0157:H7 by 1.80, 2.11, and 0.74 logs, respectively. According to Ming et al. (1997), pediocin-coated cellulosic casings inhibited L. monocytogenes on ham, turkey breast meat, and beef. In addition, McDade et al. (1999) reported that dipping frankfurters in an aqueous whey protein solution (pH 5.2) 97 containing propionic/sorbic acid prevented growth of L. monocytogenes on the product during the first 2 to 3 weeks of storage at 4 °C. Sorbic acid, p-aminobenzoic acid, lactic acid and acetic acid have a long history as GRAS food preservatives. The World Health Organization has set the acceptable daily intake for sorbic and p-aminobenzoic acid at 25 and 30 ppm, respectively (Kabara et al., 1991). When used in combination with lactic and/or acetic acid, sorbic acid can inhibit the growth of L. monocytogenes, S. Typhimurium, and E. coli 0157:H7 in many low acid foods including cold-pack cheese (Ryser and Marth, 1988), bologna (Wederquist et al., 1994), beaker sausage (Hu and Shelef, 1996) and apple cider (Zhao et al., 1993; Uljas and Ingham, 1999). p-Aminobenzoic acid reportedly exhibited greater inhibitory activity against L. monocytogenes, E. coli and Salmonella enteritidis than formic, propionic, acetic, lactic or citric acids (Richards et al., 1995). Use of whey protein antimicrobial-containing films as a casing for frankfurters appears to be a promising means of retarding surface microbial growth, thereby enhancing product safety and extending shelf life. However, if used as a sausage casing, the mechanical properties of such an edible film (that is tensile strength, percentage elongation and water vapor permeability) are of equal importance if the film is to function properly and provide adequate physical protection for the product during production and storage. Consequently, our objectives were to (1) develop an edible film (pH 5.2) from whey protein isolate (WPI) containing p-aminobenzoic acid (PABA) or sorbic acid (SA) that is inhibitory to L. monocytogenes, E. coli 0157:H7 and S. Typhimurium DT104 and (2) assess the film for water vapor permeability (WVP), tensile strength (TS) and percentage elongation (%E) at break. 98 2.3. MATERIALS AND METHODS 2.3.1. Film Preparation Whey protein isolate (WPI, Alacen 895) (New Zealand Milk Products, North America, Inc., Santa Rosa, Ca., USA) (5% w/v) and glycerol (Sigma Chemical Co., St. Louis, Mo.) (2% w/v) were dissolved in distilled water containing 0.04% CaClz (w/v) (Sigma). After mixing and adjusting the pH to 8.0 with 1.0 N NaOH, the solution was heated at 90 °C for 30 min in a shaking water bath (170 Marcel Drive water bath, Precision Scientific, Winchester, VA). Following the addition of candelilla wax (Stahl Pash, Inc., New York, NY) (0.4%, w/v) during the last 5 min of heating, the solution was homogenized for 2 min in a SD-45 homogenizer (Tekmar Co., Cincinnati, Ohio, USA), filtered through cheese cloth and cooled to 23 i 2 °C. After incorporating 0.5, 0.75, 1.0, or 1.5% (w/v) sorbic acid (SA) or p-aminobenzoic acid (PABA), the pH was adjusted to 5.2 using three different solutions of lactic acid and acetic acid at ratios of 1:0, 1:1, and 7:3 lactic acid (1.0 N): acetic acid (1.0 N). Following degassing by vacuum, the whey protein solution (40 ml / plate) was cast by pipetting the solution into sterile 17 cm- diameter Teflon plates. The solutions were dried for approximately 24 h at 23 i 2 °C / 50 i 5% RH, after which the films were peeled from the plates and stored at 23 i 2 °C / 50 i 5% RH until used. 2.3.2. Bacterial Strains Four strains of Listeria monocytogenes (CWD 95 and CWD 249 from silage, CWD 201 from raw milk, and CWD 1503 from ground turkey) and three strains of 99 Escherichia coli 0157:H7 (AR, AD 305, AD 317) were obtained from C. W. Donnelly (Dept. of Nutrition and Food Sciences, University of Vermont, Burlington, Vt., USA). Five strains of Salmonella Typhimurium DT104 (G01074, G11601, G10931, G10601, G10127) were obtained from B. Swaminathan (Centers for Disease Control and Prevention, Atlanta, Ga., USA). All strains were maintained at —70 °C in trypticase soy broth containing 10% (v/v) glycerol and subcultured twice in trypticase soy broth containing 0.6% (w/v) yeast extract (Difco Laboratories, Detroit, Mich., USA) at 35 °C / 18-24 h before use. 2.3.3. Diffusion-type Assay WPI films were aseptically cut into 16-mrn diameter discs using a sterile cork borer. The discs were then aseptically transferred to pour plates containing exactly 15 ml of either trypticase soy agar + 0.6% yeast extract (TSAYE) (pH ~6.5) or TSAYE acidified to pH 5.2 with 1.0 N lactic acid (Difco) which had been previously seeded with 0.1 ml of an 18-24 h culture of the test organism. After 24 h of incubation at 35°C, the diameter of the inhibition zone around the edible film disc was measured perpendicularly to the nearest millimeter. The end result was the average of two measurements. 2.3.4. Film Thickness A model M micrometer (Testing Machine Inc., Amityville, N.Y., USA) was used to determine film thickness. Measurements were taken at five different locations and the mean value was used in further calculations for moisture barrier and mechanical properties. 100 2.3.5. Mechanical Properties Films were cut into strips measuring 101.6 mm by 25.4 m using a Precision Sample Cutter (Thawing Albert Instrument Co., Philadelphia, Pa., USA). All films were conditioned for 48 hours at 23 i 2 °C / 50 i 5% RH before testing. Tensile strength (TS) and percent elongation at break (%E) were determined according to standard D-882-91 (ASTM, 1992). The test was run using the Instron Universal Testing Machine Model 2401 (Canton, Mass, USA) at 23 i 2 °C / 50 i 5% RH with a static load cell of 1 kN and a cross head speed of 50.8 cm/min. TS was calculated in MPa from the following equation: TS = load /(sample width x sample thickness) % Elongation at break was determined by the following equation: % E = (distance sample stretched / original length of sample) x 100 2.3.6. Water Vapor Permeability Standard Method E96 - 80 (ASTM 1992) was used in which the film was sealed on top of an aluminum test cup containing desiccant (calcium sulfate) and then placed in a chamber at 37 °C / 85% RH. The area of the cup mouth was 54 cm2 and the cup well depth was 1.1 cm. Cups were weighed at 2-h intervals during ~12 h of controlled storage. Square edible film samples having a surface area of 9 cm2 were placed in the chamber to examine moisture absorption. WVP was calculated from the water vapor transmission rate through film, the partial vapor pressure difference between the two sides of the film, and the thickness according to McHugh et al. (1993). 101 2.2.7. Statistical Analysis All experiments were replicated three times using a complete randomized design. Two- way analysis of variance (ANOVA) was performed using the SAS Statistical Analysis System (SAS Institute Inc., 1990). Means were compared using the Duncan Grouping test at p=0.05. 2.4. RESULTS AND DISCUSSION 2.4.1. Antimicrobial Properties Increasing the concentration of PABA and SA in the film discs increased the diameter of inhibition zones for L. monocytogenes (4 strains), E. coli 0157:H7 (3 strains), and S. Typhimurium (5 strains) on TSAYE (pH 5.2) (p<0.05). SA and PABA are weak acids and are most effective in the undissociated form (Luck, 1980) due to their increased ability to penetrate the cytoplasmic membrane of bacteria. (Chichester and Tanner,1972). At pH 5.2 and 6.5, 28.48 and 1.25 % and 26.18 and 1.11 % of SA (pKa = 4.75) and PABA (pKa = 4.8) is undissociated. Hence, no inhibition was observed in TSAYE adjusted to pH 6.5 (results not shown). Control films (pH 5.2) without antimicrobials were non-inhibitory. Therefore, the antimicrobial containing films developed in our study would be best suited for foods such as meats and cheeses that have pH values 5 5.2. All L. monocytogenes strains were inhibited using WPI film discs (pH 5.2) containing SA or PABA at levels of 0.5, 0.75, 1.0 or 1.5% with inhibition zones ranging from 12.0 to 32.0 and 4.0 to 27.0 mm, respectively (Table 2.1, 2.2). El-Shenawy and Marth (1988) also showed that L. monocytogenes was inhibited when 0.2 to 0.3% 102 Table 2.1. Antimicrobial activities of whey protein based edible films containing p- aminobenzoic acid against 4 strains of L. monocytogenes Diameter of Inhibition Zone (mm) PABA. LA:AA* CWD 95 CWD 249 CWD 201 CWD 1503 (%)(w/V) 0 1:0 0a 0a 0a 0a 7:3 0a 0a 0a 08 1:1 02‘ 0a 0a 0“ 0.50 1:0 13.3 i 2.6bcd 8.7 i 2.3bc 15.0 i 1.7cd 11.7 i 2.6cd 7:3 9.3 i 1.5 b“ 8.0 i 2.2 “ 8.3 i 3.7“f 17.0 i 3.2“f 1:1 8.0 i 0.8“” 4.3 i 1.2bc 5.3 i- 0.1“ 4.0 i 0.2“b 0.75 1:0 18.0 i 0.6“f 15.0 i 1.6“ 20.0 i 0.5““ 13.0 i 12°“ 7:3 13.0 i 0.4bed 12.0 i 0.9cd 17.0 8: 1.8““ 14.0 i 15°“ 1:1 14.0 i 32°“ 8.7 i 0.4bcd 12.3 i 1.1“ 9.0 i 3.5“ 1.00 1:0 21.0 i 0.4"f 17.0 i 5.3bc 22.3 i 0.5“f 19.3 i 2.3“f 7:3 18.3 i 2.9“" 19.3 i 4.9“ 16.0 i 0.6“ 20.3 i 3.5cf 1:1 18.7 :r. 2.3“f 13.0 i 5.7“ 18.3 i 2.1“f 17.3 i 3.8“f 1.50 1:0 24.7 i 4.6f 19.7 i 3.3“d 25.0 i 3.8f 22.3 i 4.1f 7:3 22.8 i 0.1cf 22.8 i 0.9“ 22.7 i 0.1“:f 16.7 i 3.2“f 1:1 20.3 i 2.3“f 18.1 i 1.16 27.0 .4: 0.7cf 20.3 i 5.8cf Geometric mean i standard deviation (n=3). Means in same column with different superscript are significantly different (p<0.05). *Ratio of lactic acid (LA) to acetic acid (AA). 103 Table 2.2. Antimicrobial activities of whey protein based edible films containing sorbic acid (SA) against 4 strains of L. monocytogenes Diameter of inhibition zone (mm) SA LA:AA* CWD 95 CWD 249 CWD 201 CWD 1503 WWW 0 1:0 0a 0° 0° 0° 7:3 08‘ 0a 0° 0° 1:1 0° 0° 0° 0° 0.50 1:0 19.0i2.3b 19.3 :1: 5.7“ 12.3 :21“ 13.7:38“ 7:3 18.0i1.3b 12.3 :31“ 21.3 :37“ 127:3.2a 1:1 17.0:13b 17.0i1.5b° 15.3 :0.1°“ 18.0i1.2ab 0.75 1:0 21.0 i 1.9“ 20.0 i 0.2“ 12.0 i 0.7b 23.0 i 05°“ 7:3 24.7 i 4.2“ 16.7 i 3.4“ 21.7 i 1.4“ 25.7 i 2.5“ 1:1 27.3 i 0.6“ 21.7 i 3.0“ 20.3 i 1.1“ 28.3 at 25° 1.00 1:0 27.0 i 3.3“ 23.3 i 1.9“ 26.3 i 12° 27.7 i 3.3“ 7:3 29.7 i 2.5"c 25.7 i 1.3“ 24.0 :r 0.8“ 28.3 i 1.5“ 1:1 30.0 _+. 2.8“ 25.7 i 3.2“ 23.3 i 1.2“ 30.0 i 38° 1.50 1:0 31.3 i 25° 30.3 1 04° 22.3 i 3.8“ 30.7 i 3.1° 7:3 27.3 i 3.1“ 13.7 i 0.9“b 20.7 :r 1.1“ 30.0 i- 2.0° 1:1 25.3 i 40" 30.0 2°. 48° 27.7 i 18° 32.0 : 38° Geometric mean i standard deviation (n=3). Means in same column with different superscript are significantly different (p<0.05). *Ratio of lactic acid (LA) to acetic acid (AA). 104 potassium sorbate was added to trypticase soy broth at pH 5.0. Films containing SA were generally more inhibitory to L. monocytogenes than films containing PABA. McDade et al. (1999) reported that growth of L. monocytogenes was inhibited on frankfurters during 2 to 3 weeks of storage at 4 °C by coating the frankfurters with a whey protein film- forrning solution that contained propionic/sorbic acid (pH 5.2). However, non-uniformity of the antimicrobial coating on frankfurters after dipping, draining and drying would likely produce a less effective antimicrobial barrier as compared to pre-casted films. In our study, all antimicrobial edible films were uniform in thickness. Consequently, these films would be better suited to inhibit post-processing surface contaminants such as L. monocytogenes. Film discs containing SA or PABA also were inhibitory to E. coli 0157:H7 with inhibition zones ranging from 0.7 to 13.3 mm and 5.3 to 21.3 mm, respectively (Table 2.3, 2.4). When used at concentrations of 0.5, 0.75, or 1.0%, PABA was more effective against E. coli than SA. Richards et al., (1995) and Tsai and Chou (1996) showed similar inhibition of E. coli 0157:H7 on laboratory media using PABA and SA, respectively. Using SA and PABA, inhibition zones for S. Typhimurium DT104 ranged from 0 to 12.2 mm and 3.0 to 16.3 mm, respectively (Table 2.5, 2.6). Films containing PABA inhibited all strains of S. Typhimurium DT104 on TSAYE, whereas film discs containing 0.5 and 0.75% SA and lactic acid:acetic acid (1 :0 and 7:3) failed to inhibit three strains of S. Typhimurium DT104 (G10127, G10931 and G10601). While laboratory media containing 0.2 or 0.5% SA is reportedly bacteriostatic to Salmonella at pH 5.5 (Restaino et al., 1981; Elliot and Gray, 1981), the amount of SA released from our film discs containing 0.5 or 0.75% SA was presumably too low to inhibit these three strains. 105 Table 2.3. Antimicrobial activities of whey protein based edible films containing p- aminobenzoic acid (PABA) against 3 strains of Escherichia coli 0157:H7 Diameter of Inhibition zone (mm) PABA LA: AA* AR (Acid Resistant) AD 305 AD 317 cow/v) 0 00 0a 0a 0a 23 0a 0a 0a 1 :1 0 a 0 a 0 a 050 00 57:L?° 27:0? 07:0? 23 20:0?“ 97:05“ r7:0? 01 53:1? 100:L§° 40:0? 0J5 00 73:2?“ 92:1.3b 30:1? 73 27::09 128:r?° 20:0? n1 25:0?“ 118:09“ 27:0? 100 10 108:0?f 125:0?° 30:0? 73 103:23“cf 123:3?° 5.3:01“b lfl 90:06“f 125:12“ 30:0? 150 10 132:2? 15.8:0?d 95:09“ 73 130:10 158:12“ 103:15“ 01 132:0.4f 213:L?° 100:Lw Geometric Mean :2 standard deviations (n=3). Means in same column with different superscript are significantly different (p<0.05). *Ratio of lactic acid (LA) to acetic acid (AA). 104 Table 2.4. Antimicrobial activities of whey protein based edible films containing sorbic acid (SA) against 4 strains of E. coli 0157:H7 Diameter of inhibition zone (mm) SA(%)(w/v) LA: AR (Acid Resistant) AD 305 AD 317 AA. 0 20 0° 0° 0° 23 0° 0° 0° 11 0° 0° 0° 050 10 17:0?b 07:03° 07:0? 73 17:05°b 23:03° 17:0? 11 23:0?“ 40:0?“ 40:02° 075 10 20:0?“ 27:0.2ab 30:13° 23 43:1?“ 43:0?“ 20:02° 14 23:0?“ 17:04° 27:0? 100 10 50:1?“ 50:11“‘ 30:0? 23 53:1?“ 45:0?“ 53:0?b 11 83:0?“ 75:1?° 30:0? 150 10 90:3?“ 8.3:21°d 95:09“ 73 113:1?° 100:0? 103:1?° 11 133:0? 123:0? 110:1? Geometric mean i standard deviation (n=3). superscript are significantly different (p<0.05). *Ratio of lactic acid (LA) to acetic acid (AA). 105 Means in same column with different In accordance with previously published data (Ahamad and Marth, 1989; Richards et al., 1995), acetic acid is more inhibitory to L. monocytogenes, E. coli 0157:H7, and Salmonella in laboratory media at the same pH value than lactic acid. Three ratios of lactic acid: acetic acid (1 :0, 7:3, or 1:1) were used to adjust the pH of our film solutions. We expected that the solution containing more acetic acid (1:1) would be most inhibitory based on the aforementioned studies. However, incorporating three different ratios of lactic acid:acetic acid in films containing PABA or SA did not synergistically alter inhibition of the three test pathogens on TSAYE at pH 5.2. The antimicrobial findings in this study are based on the measurement of clear inhibition zones surrounding film disks where growth of the pathogen was inhibited. Diffusion of antimicrobials from the film disc depends on the size, shape, and polarity of the diffusing molecule as well as the chemical structure of the film and the degree of molecular cross-linking (Guilbert, 1986). According to Michaels et al., (1962), the shape of the diffusing molecule (linear, branched or cyclic) may impact the diffusion rate. When Chen et al., (1996) measured diffusion rates for SA and benzoic acid from chitosan film in a water-glycerol solution (aw = 0.8), more SA (57%) was released than benzoic acid (65%). The different interactions of SA and PABA in our WPI-based edible films likely resulted in different diffusion rates leading to varying degrees of inhibition. 2.4.2. Mechanical properties Average film thickness was 127.11um (i 35.39) with no significant differences observed between films (Table 2.7). When SA and PABA concentrations increased from 0 to 1.5%, % E increased from 6.37% to 74.28 and 42.16%, respectively (Table 2.7). While TS of WPI films significantly decreased with increasing levels of SA (p<0.05) 106 Table 2.5. Antimicrobial activities of whey protein based edible films containing p- aminobenzoic acid (PABA) against 5 strains of Salmonella Typhimurium DT104 Diameter of Inhibition zone (mm) PABA LA: (310931 010127 (301074 G10601 G11601 (%)(v/w) AA“ 0 1:0 0 ° 0 ° 0 ° 0 ° 0 ° 7:3 0 ° 0 ° 0 ° 0 ° 0 ° 1:1 0° 0 ° 0 ° 0° 0° 0.50 1:0 6.8 : 03° 6.7 : 0.5“ 6.7 : 1.5“ 7.3 : 1.6“ 4.7 : 08° 7:3 9.2 : 0.?“ 8.3 : 04° 7.7 : 25°“ 6.8 : 24°“ 6.3 : 0.1“ 1:1 7.0 : 0.6“ 3.0 : 0.5“ 4.7 : 15° 7.7 : 1.? 6.2 : 1.6“ 0.75 1:0 9.7 : 15°“ 8.7 : 0.? 7.7 : 06°“ 8.5 : 20°“ 4.7 : 28° 23 11.3 : 15°“ 7.7 : 04° 8.3 : 1.5“ 10.7 : 0.5“ 8.5 : 0.5“ 1:1 8.5 : 23°“ 3.0 : 0.7“ 7.5 : 29°“ 9.2 : 2.0“ 7.8 : 1.3“ 1.00 1:0 12.0 : 28°“ 9.8 : 2.3“ 9.8 : 2.3“ 10.0 : 2.1“ 5.0 : 25° 7:3 13.7 : 1.2“ 10.2 : 3.6“ 10.2 : 3.6“ 11.2 : 1.3“ 10.5 : 26° 1:1 11.0: 15°“ 60:05“ 90:10“ 117:1.1“ 102:0.5° 1.50 1:0 16.3 : 10° 14.0 : 1.8d 15.3 : 2.5f 14.0 : 3.0f 14.0 : 2.4d 7:3 145:1?“ 137:3? 13.8: 1.1f 137:2.1f 13.7: 1.9“ 1:1 15.0 : 18°“ 10.3 : 0.6“ 13.0 : 32°f 15.3 : 0.6°f 14.3 : 0.1d Geometric mean 32 standard deviations (n=3). Means in same column with different superscript are significantly different (p<0.05). *Ratio of lactic acid (LA) to acetic acid (AA). 107 Table 2.6. Antimicrobial activities of whey protein based edible films containing sorbic acid (SA) against 5 strains of Salmonella Typhimurium DT104 Diameter of Inhibition Zone (mm) SA LA: (310931 (310127 (301074 (310601 G11601 (%)(v/w) AA" 0 10 0° 0° 0° 0° 0° 23 0° 0° 0° 0° 0° 11 0° 0° 0° 0° 0° 0.50 10 0° 0° 1.7 : 0.1“ 2.0 : 03“ 3.2 : 07“ 23 0° 0° 5.3 : 07°“ 4.0 : 08°“ 0° 11 3.3 : 0.1 ° 0.7 : 0.? 1.0 : 02“ 4.5 : 15°“ 3.0 : 0.5ab 0.75 10 4.9 : 04“ 0° 1.7 : 0.5°° 2.7 : 09°“ 1.7 : 03“ 7:3 6.3 : 05°“ 2.0 : 0.? 7.3 : 09°“ 7.7 : 23°“ 2.0 : 08“ 1:1 5.3 : 14°“ 2.7 : 0.5°° 5.7 : 14°“ 5.7 : 0.?“ 2.0 : 0.1“ 1.00 10 7.5 : 11°“ 5.0 : 0.2“ 5.3 : 03°“ 5.3 : 15°“ 6.2 : 0.4“ 7:3 5.3 : 09°“ 4.7 : 0.3“ 6.2 : 14°“ 7.2 : 2.7“ 6.0 : 1.2“ 1:1 8.3 : 1.5“ 8.0 : 18°“ 8.0 : 2.2“ 7.5 : 37°“ 6.5 : 24°“ 1.50 10 9.3 : 06° 9.7 : 0.6“ 7.3 : 27°“ 8.3 : 31°“ 9.7 : 1.7“ 7:3 8.0 : 0.3“ 9.3 : 1.2“ 9.7 : 1.4“ 9.7 : 1.9“ 8.0 : 0.8“ 11 11.0:1.5° 11.3:18d 11.6:14° 122:1.8f 110:1.8“ Geometric mean i standard deviations (n=3). Means in same column with different superscript are significantly different (p<0.05). “Ratio of lactic acid (LA) to acetic acid (AA). 108 (Table 2.7), TS of films containing 1.5% PABA (5.7 MPa) was similar to the control (5.92 MPa). Films containing SA exhibited lower TS and higher % E as compared to films containing PABA. The reason for this phenomenon could be that the straight chain of SA can more easily penetrate into WPI chains than PABA which has a benzene ring. Consequently, SA may have allowed more mobility between protein chains, thereby producing films of lower TS and greater flexibility. The various organic acid mixtures used to adjust the pH of the film solutions did not significantly alter % E or TS. Increasing the amount of additives other than cross-linking agents generally produced films with lower TS and greater elongation since these molecules insert between protein chains to form hydrogen bonds with amide groups of proteins (Guilbert, 1986; Kester and Fennema, 1986). Reduced interactions between these protein chains lead to increased flexibility and movement. In our study, SA, PABA, acetic acid and lactic acid might be functioning as plasticizers to increase elongation and decrease TS as was previously suggested for lactic acid (Krull and Inglett, 1971). CaC12 was incorporated into our film solution as a cross-linking agent to improve the mechanical and water vapor permeability properties of the low pH films as previously suggested by others (Guilbert, 1986; Avena-Bustillos and Krochta, 1993). As a divalent cation, calcium cross-links between negatively charged groups on proteins, thereby increasing cohesion between protein chains, reducing protein polymer segmental mobility and improving both the mechanical properties and water vapor permeability (Krochta et al., 1990). Jeyarajah and Allen (1994) reported that CaC12 induced a change in B- lactoglobulin conformation, which facilitated polymerization during heating. Calcium 109 ions also increased the reactivity of SH groups at low pH. Although the SH — S-S interchange reaction is not possible at pH < 6.5, aggregation of most whey proteins can still occur in the presence of calcium (de Wit, 1981). Whey protein films are formed by heat-catalyzed protein-protein interactions that involve disulfide, hydrogen, and hydrophobic bonds. Heating denatures the protein and exposes internal SH and hydrophobic groups (Watanabe and Klostermeyer, 1976; Shimada and Cheftel, 1998) which promote intermolecular S-S and hydrophobic bonding upon drying (McHugh and Krochta, 1994b). Film formation is favored in more alkaline film solutions since SH reactivity increases at pH > 8 (Kella and Kinsella, 1988; Banerjee and Chen, 1995). In the present study, the film solution was at pH 8.0 during heating at 90 °C, after which the pH was decreased to 5.2 using lactic and acetic acid. A low pH environment would likely prevent S-S bond formation in the protein matrix, thereby weakening the film structure. Thus, tensile strength of the low-pH film (5.92 MPa) was substantially lower than that reported for high-pH film (13.9 MPa) (McHugh and Krochta, 1994a). However, tensile strength of our low-pH film was higher than that reported for corn zein (0.4 MPa) (Aydt et al., 1991), soy protein (4.5 MPa) (Gennadios and Weller, 1991), and wheat gluten based edible films (1.9-4.4 MPa) (Gennadios et al., 1993) when tested at 23 °C / 50%RH. 2.4.3. Water Vapor Permeability Films containing 0, 0.5, 0.75, 1.0, or 1.5% PABA exhibited average WVP values of 27.24, 53.73, 53.90, 55.34, and 54.00 g.mm/m2.d.kPa, respectively (Table 2.7). Increasing the concentration of PABA from 0.5 to 1.5% did not significantly alter WVP llO Table 2.7. Thickness, Tensile strength (TS), Percent elongation (% E), and Water Vapor Permeability (WVP) of whey protein isolate-based films containing sorbic acid (SA) and p-aminobenzoic acid (PABA). Antimic. LAzAA Thickness (um) %E TS (Mpa) WVP (%)(w/v) * (g.mm/m2.day.kPa) Control 1:0 128.4 : 24.4° 6.4 : 33° 5.9 : 14° 27.2 : 1.? SA (0.50) 1:0 121.1 : 366° 20.0 : 14° 4.9 : 27° 27.3 : 98° 23 126.4 : 247° 23.4 : 11° 4.5 : 03° 21.3 : 32° 11 127.6:40.2° 31.6: 19° 4.6:07b 315:2? SA (0.75) 10 137.3 : 30.3° 26.6 : 32° 4.9 : 05° 28.6 : 6.? 23 134.6 : 231° 27.1 : 09° 4.8 : 14° 27.5 : 1.? 11 112.9 : 48.? 24.6 : 87° 4.4 : 14° 32.9 : 14° SA (1.00) 1:0 130.5 : 404° 67.8 : 64° 3.8 : 0.? 43.5 : 53° 23 132.9 : 22.? 73.5 : 14° 3.8 : 01° 41.8 : 42° 11 120.4 : 195° 70.7 :1? 3.9 : 03° 41.5 :14° SA(1.50) 1:0 118.9:43? 730:2.1° 31:05° 438:6.7° 23 123.7 : 235° 74.3 : 45° 2.6 : 10° 45.6 : 19° 1:1 133.4 : 27.? 73.3 : 53° 2.7 : 09° 44.1 : 21° PABA (0.50) 10 123.7 : 22.? 18.3 : 56° 5.4 : 09° 51.8 : 2.? 7:3 145.4 : 458° 19.9 : 65° 5.4 : 0.7° 55.2 : 31° 11 120.5 : 27.3° 16.2 : 39° 4.4 : 11° 54.3 : 1.? PABA (0.75) 10 134.1 : 235° 20.9 : 85° 5.2 : 28° 56.4 : 1.? 7:3 119.1 :45? 30.8 : 36° 52:34° 50.19:24° 11 136.3 : 56.? 28.5 : 152° 5.2 : 24° 55.08 : 16° PABA (1.00) 10 133.2 : 431° 34.7 : 63° 4.4 : 0.? 59.79 : 35° 23 111.8 : 29.3° 33.3 : 25° 5.3 : 2.? 47.08 : 1.? 1:1 120.7 : 39.3° 30.1 : 8.? 5.8 : 18° 59.17 : 26° PABA (1.50) 1:0 123.4 : 577° 34.9 :10.6° 5.3 : 0.? 56.11 : 21° 23 131.7: 349° 38.7 : 51° 5.8 : 14° 56.95 : 54° 11 130.0 : 431° 42.2 : 101° 6.0 :16° 48.96 : 2.? Mean i standard deviation (n=3). Means in the same column with different superscript are significantly different (p<0.05). * Ratio of lactic acid (LA) to acetic acid (AA). 111 (p>0.05). Average WVP values for films containing 0.5 and 0.75% SA were 26.69 and 29.70 g.mm/m2.d.kPa, respectively, and were not significantly different from the control (27.24 g.mm/m2.d.kPa) (p>0.05). However, addition of 1.0 and 1.5% SA significantly increased WVP to 42.29 and 44.46 g.mm/m2.d.kPa, respectively (p<0.05). WVP is a measure of the ease with which a material can be penetrated by water vapor. WPI edible films tend to be poor moisture barriers due to abundant hydrophilic groups in proteins. Their moisture barrier properties can be improved by adding non- polar compounds such as lipids (McHugh and Krochta, 1994b). We incorporated candelilla wax into the film solution to reduce WVP. In preliminary experiments, diffusion of SA and PABA as demonstrated by inhibition zones was similar for films prepared with and without candellila wax (results not shown). Adding SA and PABA to the film solution increased WVP because both antimicrobials are hydrophilic compounds. Addition of polar additives may increase the hydrophilic character and the solubility coefficient of the film (McHugh et al., 1994). Moreover, additives such as SA or PABA weaken chain packing in the film to produce a looser structure which increases water mobility. 2.5. CONCLUSION Incorporating 0.5 to 1.5% of SA or PABA into WPI films (pH 5.2) led to inhibition of L. monocytogenes, E. coli 0157:H7, and S. Typhimurium DT104 on TSAYE at pH 5.2. Addition of PABA and SA increased %E and WVP, but decreased TS. Given our current work involving ready-to-eat meat products which will be reported elsewhere, these films may prove useful for inactivating post-processing contaminants on ready-to-eat foods such as processed meats. 112 CHAPTER 3 INHIBITION OF LIS T ERIA MONOC Y T OGENES, ESCHERICHIA C 0L1 0157:H7, AND SALMONELLA TYPHIMURIUM DT104 ON BOLOGNA AND SUMMER SAUSAGE USING ANTIMICROBIALS EDIBLE FILMS Cagri, A., Ustunol, Z., Ryser, E.T. Journal of Food Science (In press) 113 3.1.ABSTRACT Whey protein isolate (WPI) films (pH 5.2) containing 0.5 to 1.0% p- aminobenzoic acid (PABA) and/or sorbic acid (SA) were assessed for antimicrobial activity and mechanical properties while in contact with sliced bologna and summer sausage that were inoculated with Listeria monocytogenes, Escherichia coli 0157:H7, and Salmonella Typhimurium DT104. WPI films containing SA or PABA decreased Listeria, E. coli and S. Typhimurium populations 3.4-4.1, 3.1-3.6, and 3.1-4.1 logs on bologna and sausage after 21 days at 4°C, respectively. Background flora on slices was inhibited compared to controls. Film tensile strength decreased while % elongation remained unchanged following 72 h of product contact. Consequently, these films may prove useful in extending the self-life of ready-to-eat meats. 3.2. INTRODUCTION Post-processing contamination of ready-to-eat meat products has emerged as a serious public health concern. In 1998, 101 cases of listeriosis in 22 states, including 21 fatalities were traced to consumption of hot dogs that became contaminated after processing (CDC, 1998). A Class I recall was subsequently issued for 35 million pounds of tainted product. During the second half of 2000, another outbreak involving 29 cases in 10 states (including 7 fatalities) prompted the recall of approximately 14.5 million pounds of turkey and chicken delicatessen meat because of probable contamination with L. monocytogenes (CDC, 2000). Listeria continues to threaten the processed meat industry with over 75 Class I recalls involving more than 75 million pounds of cooked ready-to-eat meats (approximately 40 million pounds of hot dogs) issued from April 1998 114 to October 2001 (U SDA-F SIS, 2001a). In 2000, 2,298 cases of foodbome listeriosis were reported in the United States at an estimated cost of $2.33 billion (~$ 1 million /case), making listeriosis the costliest food-borne disease (USDA-F SIS, 2001b). Two other meat-borne pathogens also have emerged as major public health concerns. Escherichia coli 0157:H7 has been highly publicized due to outbreaks of illness associated with ground beef (Ostroff et al., 1990; Bell et al., 1994, Mead et al., 1997, Tuttle et al., 1999) and dry-cured salami (Tilden et al.,l996). Transfer of E. coli 0157:H7 from contaminated meat or utensils to other foods such as fresh fruits and vegetables is also a major concern (CDC, 1995). The annual cost of the 62,458 E. coli 0157:H7 infections reported in 2000 was estimated at $659.1 million (~$10,500/case) (USDA-FSIS, 2001b). Salmonella Typhimurium DT104, a multiantibiotic resistant strain, is also emerging as a serious threat to public health (Glynn et al., 1998). Food-borne transmission of DT104 has been documented for several meat-related outbreaks; suspected vehicles included roast beef, ham, pork sausage, salami sticks, cooked meats, frozen sausage samples, and chicken legs (Davies et al., 1996, Anonymous, 1996). In England, 17% of 786 fresh or frozen sausage samples yielded Salmonella spp., including S. Typhimurium DT104 (Nichols and de Louvois, 1995). Sharma et al. (2001) also recovered S. Typhimurium DT104 from several points in pork production. These findings indicate that meat products pose a serious health risk if not cooked and handled properly. Product slicing and packaging operations are major points at which bacterial pathogens can be introduced into cooked ready-to-eat meat products. Bologna sausage batter contains a mixed microflora, including possible pathogens, which are destroyed during cooking; however, vacuum-packed bologna is typically recontaminated during 115 slicing. In commercial manufacturing facilities, bacterial loads in meat reportedly increased 0.5 to 2.0 logs CF U/g during slicing (Holley, 1997b). Several studies have demonstrated that antimicrobial edible films can reduce bacterial levels on meat products. Siragusa and Dickson (1992, 1993) showed that organic acids were more effective against L. monocytogenes, S. Typhimurium, and E. coli 0157:H7 on beef carcass tissue when immobilized in calcium alginate than when used as a spray or dip. In another study, application of edible corn starch film containing potassium sorbate and lactic acid inhibited growth of S. Typhimurium and E. coli 0157:H7 on poultry (Baron, 1993). Antimicrobial chitosan films containing acetic or propionic acid reportedly inhibited growth of Enterobacteriaceae and Serratia liquefaciens on bologna, regular cooked ham and pastrami (Ouattara et al., 2000) during 168 h of storage at 5°C. Natrajan and Sheldon (2000) also incorporated nisin and chelators into protein- and polysaccharide-based films to inhibit Salmonella on poultry skin. These results emphasize the potential for adding antimicrobial compounds to edible packaging materials. In our previous study, WPI films (pH 5.2) containing 0.5 to 1.5% SA or PABA inhibited the growth of L. monocytogenes, E. coli 0157:H7, and S. Typhimurium DT104 on acidified (pH 5.2) trypticase soy agar containing 0.6% yeast extract (Cagri et al., 2001). In this study, WPI films containing SA or PABA were assessed for their ability to retain their desirable antimicrobial and mechanical properties while in direct contact with bologna and summer sausage slices. 116 3.3. MATERIALS AND METHODS 3.3.1. Products Commercially produced all-beef bologna (dia. 9.6 cm) and fermented Thuringer summer sausage (dia. 5.4 cm) were obtained from the delicatessen counter of a local supermarket. Bologna was pre-sliced (~3 mm thick) while the summer sausage was sliced approximately 3-mm thick in the laboratory using a food slicer (Chefs Mark, Dallas, TX). All-beef bologna contained 39.3% fat, 3.6% corn syrup, 2.4% salt, 0.25% sodium nitrite, spices and sodium erythorbate as reported by the manufacturer. Similarly, Thuringer summer sausage contained 64% beef, 36% pork, 19.8% fat, 1.8% dextrose, 2.1% salt, 0.25% sodium nitrite, sodium erythorbate, a pediocin-producing starter culture, spices, and flavoring. When purchased, beef bologna and summer sausage had pH values of 6.0 and 4.6, respectively. Products used in each of three trails were from the same lot. 3.3.2. Film Preparation Whey protein isolate (WPI, Alacen 895) (New Zealand Milk Products, North America, Inc., Santa Rosa, CA) (5% w/v) and glycerol (Sigma Chemical Co., St. Louis, MO) (2% w/v) were dissolved in distilled water containing 0.04% CaClz (w/v) (Sigma). After mixing, the solution was adjusted to pH 8.0 with 1N NaOH (Sigma), and heated at 90°C for 30 min in a shaking water bath (Precision Scientific, Winchester, VA). Following the addition of candelilla wax (Stahl Pash, Inc., New York, NY) (0.4%, w/v) during the last 5 min of heating, the solution was homogenized for 2 min in a SD-45 homogenizer (Tekmar Co., Cincinnati, OH), filtered through cheese cloth and cooled to 117 23 1|: 2°C. After adding 0.75 or 1.0% (w/v) sorbic acid (SA) (Sigma), p-aminobenzoic acid (PABA) (Sigma), or 0.5% SA: PABA (1.0: 1.0), the pH was adjusted to 5.2 using 1N lactic acid (Sigma). Following degassing by vacuum for 20 min, the whey protein solution was cast by pipetting 75 ml of the solution into sterile Teflon® plates (20x30 cmz). The solutions were dried for approximately 24 h at 23 :t 2°C/50 i 5% RH, after which the films were peeled from the plates and stored at 23 i- 2°C/50 i 5% RH until used. 3.3.3. Bacterial Strains Based on results from our previous WPI antimicrobial edible film work (Cagri et al., 2001), the most resistant strain of each of the three pathogens to SA and PABA was chosen. Strains used in this study included Listeria monocytogenes CWD 95 and Escherichia coli 0157:H7 AR (previously obtained from C. W. Donnelly, Dept. of Nutrition and Food Sciences, University of Vermont, Burlington, VT) and Salmonella Typhimurium DT104 G10127 (previously obtained from B. Swaminathan, Centers for Disease Control and Prevention, Atlanta, GA). All strains were maintained at -70°C in trypticase soy broth (TSB) (Difco Laboratories, Detroit, MI) containing 10% (v/v) glycerol and subcultured twice in TSB containing 0.6% (w/v) yeast extract (TSBYE) (Difco) at 35°C/18-24 h before use. 3.3.4. Product inoculation and storage Bologna and summer sausage slices were separately inoculated with 0.1 ml of an appropriately diluted culture so as to contain log 6.0 (i 0.1) L. monocytogenes, E. coli 118 0157:H7 or S. Typhimurium DT104 CFU/g. The inoculum was evenly spread on both surfaces with a sterile glass rod. The slice was then placed in a sterile 150 mm-diameter petri dish to which a piece of edible film slightly larger than the slice was previously added (Figure 3.1). Thereafter, an identically inoculated slice was placed on top of the edible film with this process repeated until a stack of eight slices, each separated by a piece of edible film, was obtained. One final piece of edible film was laid on the top slice so that both faces of all slices were in contact with the film. These petri dishes were then covered and stored aerobically to give a worst case scenario for film testing at 4°C. Two controls were used - the first consisting of bologna and summer sausage slices separated by the identical edible film prepared without antimicrobials and the second containing slices stacked together without the use of edible film. 3.3.5. Microbiological analysis Slices of bologna and summer sausage were examined for levels of the inoculum, as well as mesophilic aerobic bacteria, lactic acid bacteria, and yeast/molds immediately after inoculation and again following 4, 7, 10, 14, and 21 days of refrigerated storage. Slices weighing 10-g each were added to 90 ml of sterile 0.1% (w/v) peptone water (PW) (Difco) and homogenized for 3 minutes in a Stomacher 400 (Tekmar Co., Cincinnati, OH). Appropriate dilutions in PW were surface plated on Modified Oxford Agar (Difco), McConkey Sorbitol Agar (Difco), McConkey Agar (Difco), Plate Count Agar (Difco), MRS Lactobaccillus Agar and Rose Bengal Agar (Difco) to quantify L. monocytogenes, E. coli 0157:H7, S. Typhimurium DT104, mesophilic aerobic bacteria, lactic acid 119 bacteria, and yeast/molds, respectively, as outlined in the FDA Bacteriological Analytical Manual (FDA, 1998). 3.3.6. Mechanical Properties Edible films placed between uninoculated slices of bologna and summer sausage were examined at 0, 24, 48, and 72 h for thickness, tensile strength and percent elongation after conditioning the films for 24 h at 55 i 5 % RH / 23 :t 5° C. A model M micrometer (Testing Machine Inc., Amityville, NY) was used to determine film thickness. Measurements were taken at five different locations and the mean value was used in further calculations for moisture barrier and mechanical properties. Films were cut into strips measuring 101.6 mm by 25.4 m using a Precision Sample Cutter (Thawing Albert Instrument Co., Philadelphia, PA). All films were conditioned for 48 h at 23 i 2°C / 50 i 5% RH before testing. Tensile strength (TS) and percent elongation at break (% B) were determined according to standard D-882-91 (ASTM, 1992). The test was run using the Instron Universal Testing Machine Model 2401 (Canton, MA) at 23 2 2° C / 50 i 5% RH with a static load cell of lkN and a cross head speed of 50.8 cm/min. TS was calculated in MP3 from the following equation: TS = load /( sample width x sample thickness) % Elongation at break was determined by the following equation: % E = (distance sample stretched / original length of sample) x 100 120 3.3.7. Statistical Analysis All experiments were replicated three times. All film thickness, TS, and %E data was analyzed by one-way analysis of variance (ANOVA) using the SAS Statistical Analysis System (SAS Institute Inc., 1990). Means were compared using the Duncan Grouping test at p=0.05. 3.4. RESULTS AND DISCUSSION 3.4.1. Antimicrobial Properties L. monocytogenes. Using WPI films containing 0.75% PABA, 1.0% PABA, 0.75% SA, 1.0% SA, or 0.5% SA: 0.5% PABA (1:1), populations of L. monocytogenes decreased 1.5, 2.2, 3.0, 3.4, and 2.8 logs, respectively, on bologna slices after 21 days of storage at 4°C (Figure 3.1A) with all films remaining intact. Listeria counts remained relatively unchanged using films without antimicrobials; whereas in the absence of films, populations on bologna increased 2.2 logs after 21 days of refrigerated storage. Control WPI films without SA or PABA were acidified to pH 5.2 with lactic acid. Hence, lactic acid present in control films would also be expected to retard the growth of L. monocytogenes on bologna slices. L. monocytogenes was inhibited on fermented summer sausage using WPI films containing PABA or SA (Figure 3.1B). All WPI films containing antimicrobial agents reduced L. monocytogenes populations 3.0 to 4.1 logs on summer sausage slices after 21 days of refrigerated storage. While Listeria populations also decreased initially using antimicrobial-free film and slices without film, substantial regrowth occurred following 10 days of refrigerated storage. 121 l O—‘NO’AO’IO’NCOO Iog1OCFU/g lw‘l oA-lslxlrcbtschorxloolclo log1OCFU/g . .’ r 0 Days 14 21 0.) 7'0 ’3‘ .7 0-6‘ o/x / .9 '0 6‘ 6‘ A "i, 494 4 75° ’6, 0 ‘7 4 Q? Figure 3.1. Inhibition of L. monocytogenes on bologna (A) and fermented sausage (B) slices with WPI edible films containing 0.75%, and 1.0% p-aminobenzoic acid (PABA); 0.75% and 1.0% sorbic acid (SA) and 0.5% SA: 0.5%PABA (1:1). Control + film (WPI edible film without SA or PABA) and control (no film used). 122 log 10 CFU/g log10 CFU/g 7 Days Figure 3.2. Inhibition of E. coli 0157:H7 on bologna (A) and fermented summer sausage (B) slices with WPI edible films containing 0.75%, and 1.0% p-aminobenzoic acid (PABA); 0.75% and 1.0% sorbic acid (SA) and 0.5% SA: 0.5%PABA (1:1). Control + film (WPI edible film without SA or PABA) and control (no film used). 123 1 l 1 O l \ l l ,3" / i; l L 1 L ,7 654.66.5064166454 log1OCFU/g Iog1OCFU/g 0 . 0 Co °00 1o 7 0.2 10s '5'06 ”"04 °’ Days 14 21 0’5 '0'949584 4 84:0 £770; A o 4194 4 484 Figure 3.3. Inhibition of S. Typhimurium on bologna (A) and fermented summer sausage (B) slices with WPI edible films containing 0.75%, and 1.0% p-aminobenzoic acid (PABA); 0.75% and 1.0% sorbic acid (SA) and 0.5% SA: 0.5%PABA (1:1). Control + film (WPI edible film without SA or PABA) and control (no film used). 124 Wederquist et a1. (1995) reported that potassium sorbate 0.26% (w/w) and sodium lactate 2% (w/w) were highly inhibitory to L. monocytogenes on bologna during 28 days of storage. In our most recent study, low pH WPI-based edible films (pH 5.2) containing 0.5 to 1.5% SA or PABA also inhibited L. monocytogenes on laboratory media (Cagri et al., 2001). Using antimicrobial films, greater inhibition of Listeria was observed on summer sausage (pH 4.6) compared to bologna slices (pH 6.0) due to higher levels of SA and PABA in the undissociated form. In the absence of antimicrobial WPI films, beef bologna also allowed more growth of L. monocytogenes than summer sausage. Other inhibitory factors in tested fermented summer sausage, such as pediocin produced by the starter culture (Spelhaug and Harlander, 1989), nitrite, salt, low pH and low aW would also serve to further inhibit L. monocytogenes. The combined effects of nitrite and low pH (~4.6) were likely responsible for inhibiting L. monocytogenes on control summer sausage slices (Buchanan and Philips, 1990), with Glass and Doyle (1989) also reporting that L. monocytogenes was unable to multiply in vacuum-packed summer sausage (pH 4.8 - 4.9) containing 3.0 - 3.4% salt. Their results agree with those of other authors (Berry et al., 1990; Buncic et al., 1991), who also observed a decline in numbers of L. monocytogenes in fermented sausage during extended refrigerated storage. E. coli 0157 :H7. Numbers of E. coli 0157:H7 on bologna slices decreased 2.7 to 3.6 logs after 21 days at 4°C using antimicrobial films containing SA or PABA (Figure 3.2A). However, the pathogen decreased only 2.1 and 1.4 logs, respectively, on control slices with antimicrobial-free film or without film. Grth of E. coli 0157:H7 also was inhibited on fermented summer sausage (Figure 3.2B). Using WPI films containing 0.75 and 1.0% PABA or SA or the combination of 0.5% SA and 0.5% PABA (1:1), 125 populations decreased 2.3 to 2.6, 2.7 to 3.1, and 2.9 logs, respectively. Numbers of E. coli 0157:H7 significantly decreased on both bologna and summer sausage at the end of storage using antimicrobial films (p<0.05) (Table 3.1). In our previous study, WPI edible films containing 0.5 to 1.5% PABA or SA inhibited E. coli 0157:H7 in laboratory media (Cagri et al., 2001). However, numbers of E. coli 0157:H7 on summer sausage decreased for the film-free and antimicrobial-free controls, with this inhibition likely due to the combined effects of chemical preservatives, nitrite, salt and a pH of 4.6. Several researchers (Pond et al., 2001, Riordan et al., 1998) also observed that 2.5% salt and 100 ppm sodium nitrite were inhibitory to E. coli 0157:H7 in fermented summer sausage. However, other studies have claimed that this pathogen is very hardy (Benjamin and Datta, 1995), being able to survive more than 51 days in laboratory media at 10°C containing starter culture (107 CFU/ml), dextrose (0.8%), NaCl (2%), and NaNOz (200 ppm) (Glass et al., 1992, Tomicka et al., 1997). When initially present at 104 CFU/g, E. coli 0157:H7 reportedly survived during manufacture and storage of fermented sausage regardless of whether a starter culture was used (Glass et al., 1992). S. Typhimurium DT104. Using WPI films containing 0.75 or 1.0% SA or PABA, populations of S. Typhimurium DT104 decreased 2.7 to 3.1 logs on bologna slices after 21 days of refrigerated storage (Figure 3.3A). In contrast, cell numbers on antimicrobial-free film and film-free control bologna slices decreased only 1.5 and 0.5 logs after 21 days of storage, respectively. WPI films containing 0.75 or 1.0% PABA, SA, or a combination of 0.5% PABA and 0.5% SA (1:1) decreased levels of S. Typhimurium DT104 3.5, 4.1, 2.8, 3.6 and 3.9 logs on fermented sausage slices after 21 days, respectively (Figure 3.4B). However, a decrease of only 1.5 logs was observed for 126 control slices. Cagri et al. (2001) previously concluded that WPI edible films containing 0.75 to 1.5% PABA or SA inhibited all strains of S. Typhimurium DT104 tested on TSAYE. WPI films also were prepared to contain 0.5% of both SA and PABA to test their synergistic effect against growth of the three pathogens on bologna and summer sausage. WPI films containing SA or PABA were more inhibitory to the three test pathogens than WPI films containing both SA and PABA, although several studies claimed that SA and benzoic acid were more effective when used together rather than alone (Sofos and Busta, 1983, Luck, 1980) Mesophilic Aerobic Bacteria. After 21 days of refrigerated storage, growth of mesophilic aerobic bacteria (MAB) on bologna slices was inhibited as much as 5 .0 and 5 .8 logs using films containing 1.0% SA or PABA as compared to the antimicrobial-free film and film-free controls, respectively (Table 3.2). However, Petaja et al.(1979) found that incorporating 0.25% potassium sorbate into cooked sausage had no effect on levels of aerobic bacteria when the product was stored at 7°C. Growth of MAB was generally inhibited on summer sausage slices by all WPI films; however, MAB populations increased 1.4 and 2.4 logs after 21 days on sausage slices with antimicrobial-free film or film-free controls, respectively (Table 3.5). Thus, WPI films can extend the shelf life of both bologna and summer sausage at 4°C. Lactic Acid Bacteria. Using antimicrobial films, lactic acid bacteria (LAB) decreased 1.0 to 3.0 logs on bologna slices after 4 days and generally remained at these levels throughout storage; whereas no reduction in LAB populations was observed in the controls (Table 3.2). Fermented summer sausage initially contained 6.0 log CFU/g of 127 LAB with populations remaining relatively unchanged during 21 days of storage regardless of the type of film used (Table 3.5). However, numbers of LAB were 1.6 to 2.2 logs higher in samples stored without film. Early studies suggested that sorbate concentrations of 1.0% were markedly inhibitory to Lactobacillus bulgaricus, Lactobacillus acidophilus and Streptococcus lactis (Hamdan et al., 1971). LAB predominate in bologna and are often the major species present in fully ripened fermented sausage. According to Holley (1997a), bacteria present on bologna slices stored at 7°C were almost exclusively LAB, with Lactobacillus sake dominating. Bacillus thermosphacta wasabsent and coliforrns were rarely detected. Fermented sausage also contained thermotolerant homofermentative lactobacilli which were quantified along with pediococci. However, growth of pediococci was limited by low temperature (<10°C) storage of the samples (Holley et al.,1988). Yeast/Mold. As expected, mold grth on both bologna and summer sausage was inhibited with WPI films containing SA or PABA during 21 days of refrigerated storage (Tables 3.2 and 3.4). However, mold counts eventually reached 1.3-1.4 and 1.7- 2.2 logs in the antimicrobial-free film and film-free controls, respectively. Using WPI films containing PABA, yeasts were not detected on bologna or summer sausage slices until day 7. Films containing 0.75 or 1.0% SA inhibited yeasts 7 to 21 days and 21 days on bologna and fermented summer sausage, respectively. The combination of 0.5% SA and 0.5% PABA (1:1) prevented growth of yeasts on bologna but not on summer sausage. In the antimicrobial-free film and film-free controls, yeast populations on bologna and summer sausage increased 3.5-3.6 and 5.0-5.6 logs, respectively, by the end of storage. Inhibition of yeasts and molds by sorbate is well documented (Liewen and 128 Marth; 1985, Stead, 1995). Dipping fermented sausage and raw ham in a solution of 10- 20% sorbate inhibited mold growth (Leistner et al., 1975). Application of 2% gelatin, 2% liquid smoke, 0.2% corn starch, and 1% potassium sorbate on the inside of regenerated cellulose casings also prevented mold growth during sausage curing (Rose and Turbark, 1969). 129 845422 .deWO .Noaaa ewoaw_ ._daea .moawm evdawa .moaqm eqaavm .voahm .eda_e .moawm enoanm .moahm mm m pecamd so; Ham eeoanm .aaawe emoaem .moaae encanm mm .moaqa emoawm aaoaam smoawm .moawm .noaem ewoaom m enoHQO .moaw_ .Woaqm ewoanm encamm spoons; smoaam mm Amodvav “avenge 3:805:me 8m flaw—omega E02056 515 5:28 2:8 E memo—2 .305 cozfl>ou 25:83 H :32 Lassa? o :542282o Leave- o eaa+3§8o 2 No 4 am no “no BMHHHH .odHE .o.oHo.V .odHSV .32.? .333 .o.oHo.V v .OHSV .oquv .odHQV .odHSV .odHSV .odHSV .33.? G .EH: 15H: .33.? .33? .odHSv .odHQV .33? a 1.32 .odHiv :SHSV .odHiv .odHSv .odHSV .33.? 3 .oquv .327 .33.? .33.? .o.oHo._v .o.oHo._v .33? 2 .3 H 3v .3 H 3v .3 H 3v .3 H 3v .od H 3v .3 H 3v .2. H 3v 5 202 .odHiv .odHSV .odHQV .odHiv .odHSV .odHSV .o.oHo.V H .35 H 3v .3 H 3v Hod H 3v .3. H 3v .3 H 3v .3 H 3v .3 H iv o 85 86 2 6o 2: who 8.— mg :8 8E - 3550 8E + 35:8 omAHed HhAHoAv pdemA Ho.~Ho.~v awAHoAv .m4HoAv idHoAV v .odHoAV HodHoAv HodHoAV no.oHo._v HodHQV HodHoAV HodHQV o .o.oHN.N nodeA Ho.oHo.—v nodHoAV .odHQV HodHoAv Ho.oHo._v fim oodeN no.oHc._ HodHoAV HodHQV Ho.oHo._v HodHoAv HodHoAV 3 nodHNA .3de; HodHoAv HodHoAv HcdHoAv HodHoAv Ho.oHo._v A: 3 od H 9N 9 06 H N; a od H oAV a o.o H oAv a od H oAv a od H G. V a od H o._v n 202 BodHaA ovAHvN HodHQV HodHoAv HodHoAV HodHoAV HodHoAV v HcdHoAv idHoAV Ho.oHo.~v HodHoAV HodHoAv .odHoAV HodHoAv o cod cod md nmd co; mud co; mud mme Emu - _obcoU Ezm + 3980 0.05) (Tables 3.6 and 3.7). Percent elongation at break (% E) and tensile strength (TS) of films in contact with bologna and summer sausage slices were measured after 0, 24, 48, and 72 hours of storage to determine their eventual suitability as sausage casings. TS of all films decreased (p<0.05) after 24 h of contact with bologna and summer sausage slices; whereas % E increased (p<0.05) after contact with either meat product (Table 3.3 and 3.4). No significant increases (p>0.05) in % E were observed between films measured after 24 and 48 or 48 and 72 h of storage at 4°C. In addition, no significant decreases (p>0.05) in TS were observed beyond 24 h. Chemical stability of the films was predictably altered while in contact with bologna and summer sausage slices because of expected moisture absorption from the meat slices and release of SA, PABA, and glycerol to reach a steady state. Moisture absorption by the film likely weakened the polymer chain bonds since water molecules insert between protein chains to form hydrogen bonds with amide groups of proteins, thereby reducing tensile strength (Guilbert, 1986; Kester and Fennema, 1986). Reduced interactions between these protein chains likely led to increased flexibility and movement, as evidenced by the marked increase in % E. 135 Amodvg Beam? huge—mama 0.8 898233 EEobE .33 5:28 2:8 E mane—2 .Amncv c2332. c.8983 H :32 SHAQO amoHno snquc DHQHS 15H? idHed S 2 H a ME A. S H ho a No H no .. to H 2 A. 3 H 3 a No H ad as @313 :5de .2233 n.QH Z n.2H3 pdeS 4m 3 H ,3 a 3 H 3 a 2 H 3 a S H 3 a no H? a 2 H 3 o 3%,: E, 3: Hum; 3.8ng 3.2 Hm; 3.335: 023.8 1.33.2: Q n: H 2 SH 3 02 H 3: s h: H mfi 2.0.2 H 32 s at H I” 21% H 3a was NS H .. is a 2: H h: ._ nfl H 1% .. SN H3,” 54.2 Ho: A. 2: H N3 Hm M: H a 0.: a 2: HSM a E H ”.8 a we H :0 a ed H 0.8 a 3 H «a o m as 3le “SH: “SH? “SH? a2H; 1.3? Q ,8 H a S a to H 3 s 2 H S a mo H OH H to H S a 2 H 3 we as H m 3s a no H 3 a to H 3 a we H 2 a S H 3 a S H ow «N no H a S a to H 2 a no H 3 a 2 H 3 a no H 3 a S H 2 o 2:5 £535 992 cc _ Ed cq n K o c mac: 89:8 E 233 A0.05). Using different casings for hot dogs did not significantly change the rancidity (p>0.05). A general decrease in TBA values was observed during storage. Addition of nitrite and phosphate to hot dog formulations will delay rancidity because of their action as antioxidants (Zubillaga et al., 1984). TBA values in meat are well-correlated with oxidative rancidity (Melton, 1985). However, malonaldehyde, a highly reactive secondary product of lipid oxidation, can react with other meat components such as amino acids and amines, forming complexes that are not detected by TBA analysis (Gokalp et al., 1983). Therefore, competition for malonaldehyde in a system containing protein could result in reduced color development and incomplete quantitation of malonaldehyde. Hung and Zayas (1989) also showed that TBA values decreased from 0.79 to 0.59 mg malonaldehyde/kg in frankfurters during 45 days of storage. 4.3.3. pH The pH of hot dogs decreased from ~62 to ~57 for all casings during refrigerated storage (Table 4.3) with none of these differences significant after 42 days (p>0.05). Paneras and Bloukas (1994) also found that the pH of frankfurters decreased from 6.4 to 5.8 during 5 weeks of storage 5°C with this pH decrease attributed to activity of lactobacilli. 155 Jill. Amodvav 388.an b38536 23 383895 «cobbmu 5:5 5:38 283 E 3:82 .Amuav 8838 2888 H :82 156 .33 H Q3 .3 H E3 .3 H Q3 .3 H Q3 3 .3H Q3 .3333 .333 .333 33 ..33 H :3 3.33 H 33 . 3 H 33 .33 H 333 2 3333 3.333 .333 .3H :3 : :3 .33 H 333 .3 HQ3 .3 H 33 .3 H 333 N3 .33 H Q3 .33 H Q3 .3 H Q3 .33 H Q3 3N . 3 H 333 .3 H 33.3 .3 H 333 .33 H E3 3: 3333: 0330222.... 3:3 .33 H 333 .3 H 33 .3 H53 .3 H 33 : <5 2.5.2 53.38 EB <33 .33. :33 .33 3580 mo on»... .0388 838339: mo m3: m3 macaw 3:88 #883: can £03.30 .53 A ES, 8385 3.8 Ho: no In 98 833.» 33mg. .m.v 0.3—ah. Table 4.4. Analysis of Variance on the effect of easing type on microbiological, chemical, and physical characteristics of hot dogs during 42 days of refrigerated storage (P-values for independent variables and interactions) Source of Variance Casings Time Interaction (Casings x Time) L. monocytogenes <0.0001* * * <0.0001* * * 0.0004* * MAB 0.019* <0.0001*** 0.31 LAB <0.0001*** <0.0001*** 0.0018“ Mold 0.64 <0.0001“ * 0.84 pH 0.95 <0.0001* ** 0.92 TBA 0.18 0.0014" 0.88 Purge 0.0015“ 0.22 0.54 Color-interior L“ (Lightness) 0.35 <0.0001*** 0.39 a* (redness) 0.85 <0.0001“ * * 0.40 b*(Yellowness) 0.89 <0.0001* * * 0.45 Color-exterior L* (Lightness) <0.0001"‘ * * <0.0001* * * 0.003" a* (redness) 0.13 0.05 0.18 b*(Yellowness) 0.0008*** 0.18 0.39 *significant at p<0.01; ** significant at p<0.001, *" significant at p<0.0001. MAB: Mesophilic aerobic bacteria, LAB: Lactic acid bacteria 157 4.3.4. Diffusion Coefficient After one day of refrigerated storage, 78.7 and 75.6% of the PABA and SA diffused from the WPI film into the hot dog with only 7.7 and 6.5 % of initial PABA and SA remaining in the WPI casings afier 42 days of storage, respectively. For these whey protein-based films diffusion coefficients for PABA and SA at 4°C were 1.7 x 10'13 mz/sec and 3.2x10'l3 mz/sec, respectively. Chen et a1. (1996) also studied diffusion of SA and benzoic acid from edible film and found that about 39% of SA and benzoic acid was released from methycellulose/chitosan films into a glycerol/water solution (aw=0.8) within 30 minutes at 4°C. Since the storage time was only 6 hours in their study, the maximum release of antimicrobials was only 49%. Raising the temperature to 25°C enhanced the diffusion rate for both antimicrobials from these films; however, increasing the pH from 3.0 to 6.0 had no effect on diffusion. 4.4.5. Antimicrobial analysis The numbers of L. monocytogenes on hot dogs with control (WPI without antimicrobials) and collagen casings increased 4.2 and 5.4 logs, respectively, after 42 days of storage at 4°C (Figure 4.1). Using casings containing 1.0% PABA, Listeria populations remained relatively unchanged; however, ntunbers of Listeria on hot dogs prepared with WPI casings containing 1.0% SA or 0.5% SA: 0.5% PABA (1:1) increased 2.8 and 2.1 logs after 42 days, respectively. WPI casings containing SA were less effective than WPI casings containing PABA for inhibiting Listeria growth on hot dogs even though SA-containing films decreased populations of Listeria on bologna and summer sausage slices in our previous study (Cagri et al., 2002). Acidity of WPI casings (pH 5.2) may decrease when used on hot dogs (pH 6.1) due to the diffusion of chemical 158 components between the hot dog batter and the casing (e.g. lactic acid). Sorbic acid (pKa 4.19) is most active at pH<6.0 since it is the undissociated form that possesses antimicrobial activity. However, PABA (pKa 4.8) still shows some effectiveness against L. monocytogenes at pH values as high as 6.0 (Richards et al., 1995). Overall, WPI casings containing 1.0% PABA were most inhibitory to Listeria and were further examined in detail. Listeria monocytogenes: After 42 days of refrigerated storage L. monocytogenes populations increased about 2.5 logs on hot dogs prepared with WPI-control (antimicrobial—free), collagen, and natural casings (Figure 4.2). In contrast, WPI casings containing 1.0% PABA inhibited the growth of L. monocytogenes throughout 42 days of refrigerated storage. In our previous work, WPI films containing 1% PABA were more inhibitory, decreasing numbers of L. monocytogenes 2.8 to 3.0 logs on both bologna and summer sausage slices during 21 days of refrigerated storage (Cagri et al., 2002). Stacking these slices in sterile petri dishes may have induced greater release of PABA from the film to the slices compared to the hot dog casings, thereby allowing greater inhibition. As mentioned earlier, increasing the initial pH (5.2) of WPI casings would tend to decrease the antimicrobial activity of PABA. While phenolic compounds from the smoking process and nitrite in the hot dogs may also enhance the inhibitory effect of WPI-1% PABA casings against Listeria (Lou and Yousef, 1999), such effects were not observed in our study. Previous findings (Richards and Xing, 1992; Richard, 1995) indicated that undissociated uncharged PABA diffuses through the cell membrane more freely and results in PABA exerting an increased activity as the pH decreases from 7.0 to 5.0 since the pKa value of 159 PABA is 4.8 (Terada, 1972). Inside the cell, the dissociation of PABA leads to the uncoupling of both substrate transport and oxidative phosphorylation from the electron transport system with growth inhibition resulting principally from the lost the cellular uptake of amino acids, phosphate and organic acids (Freese et al., 1973). However, even at pH values close to neutral (pH 6.5-7.0) PABA is still reportedly inhibitory to L. monocytogenes (Richards et al., 1995). This explains the inhibitory effect of WPI-PABA casings against Listeria on hot dogs even at above optimal pH value in our study. Incorporating a buffering agent in the WPI-PABA casing formulation may help stabilize the hot dog pH at a value that would be more optimal for inhibition of Listeria. Diffusion test results showed that while only 0.12% (w/v) PABA remained the day after manufacturing, this concentration was still sufficient to prevent Listeria grth during 42 days of storage. However, if diffusion of PABA could be delayed by modifying the WPI casing preparation (e.g. tightening the protein chains with crosslinking agent), WPI-1.0% PABA would be even more effective against L. monocytogenes. Mesophilic Aerobic Bacteria: Populations of mesophilic bacteria increased 2.6, 3.7, 4.2, and 5.1 logs on hot dogs prepared with WPI-1.0% PABA, commercial collagen, natural and WPI-control casings, respectively, during 42 days of storage at 4°C (Figure 4.3). Growth of mesophilic bacteria on hot dogs was suppressed 1.0 to 2.5 logs using WPI casings containing 1% PABA. Our previous study also showed that growth of mesophilic bacteria on bologna and summer sausage slices was retarded at 4°C using WPI films containing 0.75 and 1.0% PABA (Cagri et al., 2002). The various microbial reduction strategies used for processed meats typically select for specific groups of organisms, some of which can proliferate and cause spoilage during storage. Processed meats generally spoil due to facultative Gram-positive bacteria (lactic acid bacteria) that 160 .Amnav 2233323 93533 H 3:322 .3530 Beanbag 53> 333E 38 “on :o noanSAueaoE Q .«o 5.30.5 %an $58205 .3. 95w:— «>8 «v m... 8 a 3 s 3 I B I L I P O . P . ~ 3 1‘ .. v _Ol It 6 _l m and H m I w m - - - w . u I <9}. .3.-..§|¢I . 3835 mwov «o: :o «c223 0323 023336 Mo 5305 .mé «.5»; £60 «v mm mm 3 E. x. o n P - b n .n. c I. Bingo o‘601 - I flaw? i J . a 59.30:? H 33+ _. h ‘ ~flm>+ ‘ J‘i ‘Tl‘tilllalp I w 164 .Amunv :oumgov Ewwaflm H €08 250880 .3530 “scumbag 53> @8305 mwow Hon no «F503 Eon £83 mo 5265 6.9 0.5»:— gun «t mm mu 3 3 h c r n I L h P c Bingo m601 -_u._3uzl0| ’ coau=oolil “ __.. R93 $733+ L m h 165 .Amnav :ozamsu @8283 H "=on 050880 .mwfimwo “couobmv 5MB Banana mwou “on no 22: mo 5386 .m... 95»:— 3.5 «V mm mu 3 3 h c , “_ea‘uzlol‘ 1 1 ‘ - m . cans—30+ _Eslll 1 h ‘ E: 5:211 166 Mold/ Yeasts: Mold counts increased only 1.0 to 2.4 logs on hot dogs during 42 days due to the combination of vacuum packaging and storage at refrigeration temperature (Figure 4.5). Mold grth was least using WPI-1.0% PABA casings. Although benzoic acid is a well known inhibitor of yeasts and molds (Balatsouras et al., 1963), some mold growth was observed on WPI casings containing 1.0% PABA. Although all hot dogs were vacuum-packaged, some air diffusion through inadequateheat seals could explain the mold growth that was detected on a limited number of samples. No yeasts were found during 42 days of refrigerated storage. 4.4.6. Mechanical Properties Tensile strength of WPI casings remained unchanged after cooking and smoking (Table 4.5), whereas tensile strength of the collagen and natural casings decreased significantly from 34.8 and 43.3 to 10.2 and 9.6 MPa, respectively (p<0.05). Percent elongation of WPI casings containing 1.0% PABA was also unchanged afier cooking and smoking (Table 4.5); however, percent elongation increased from 12.4 to 23.5% in WPI casings without PABA and decreased from 41.9 and 56.7 to 24.7 and 21.4% in collagen and natural casings, respectively. In this study wet strength of casings were not measured. However, compared to natural and collagen casings, WPI casings required extra care to avoid breakage during stuffing and smoking. A hand stuffer was used for WPI casings adjusted to avoid breaking the heat seal or rupturing. In addition, the rate of air circulation and the humiditywere reduced during the first stage of smoking and cooking in order to keep the hot dogs on the smoke house tract. 167 4.4.7. Cook yields and purge loss Hot dogs prepared with WPI casings had cook yields of 88.9 -89.4% compared to 94.3 and 95.3% for hot dogs prepared with collagen and natural casings, respectively (Figure 4.6A). Purge loss for hot dogs prepared with WPI-PABA or collagen casings did not significantly change during 42 days (Figure 4.6). However, purge loss for hot dogs prepared with WPI and natural casings decreased and increased after 14 days, respectively. WPI and WPI-PABA casings exhibited higher purge losses (0.60 and 0.69%) than did hot dogs with collagen (0.24%) and natural casings (0.30%). WPI casings are hydrophilic even after incorporation of candellila wax to decrease water vapor permeability. This explains the higher liquid release through hydrophilic WPI casings compared to other commercial casings. 4.3.8. Shear Force Hot dogs are suitable for shear testing since they have a constant diameter and homogeneous structure. The resulting shear forces for hot dogs prepared with WPI-1% PABA (0.75 kg/g), WPI (0.80 kg/g) and collagen (0.59 kg/g) were similar (Figure 4.6B), whereas hot dogs prepared with natural casings had a significantly higher shear force (1.71 kg/g) (p<0.05). When consumed, hot dogs are normally first bitten through with front teeth and then ground on the molars (Boyar and Kilcast, 1986). This biting action was used as the basis of texture measure 168 Table 4.4. Tensile strength (TS) and percent elongation (% E) of WPI-p-aminobenzoic acid (PABA), WPI, collagen, and natural casings before and afier cooking and smoking of hot dogs. Casings Before cooking and smoking After cooking and smoking TS (MPa) WPI-1% PABA 19.5 i 02" 16.9 :I: 0.3 c WPI 6.5 i 0.4 a 5.4 i 0.6 a Collagen 34.8 i: 0.1d 10.2 i 0.4 b Natural 43.3 a 0.3 c 9.6 i 0.2 b % E WPI-1% PABA 40.7 i 0.2 c 39.8 i 0.3 c WPI 12.4:03a 23.6:01" Collagen 41.9 i 0.5 ° 24.7 a 0.3 b Natural 56.7 a 0.3 d 21.4 a 0.5 b Mean i standard deviation (n=3). Means with different superscript are significantly different (p<0.05). 169 98 -~ A B B 96 Al . § 94 *‘ A A ; ll l {bl 2 92 l i E .2 90 l l l l > as 1 l l z % § 86 . l F l l o 841 i l l l i ? 80 a _.L.a.._._al_ _ _ _ _L.,a..._.l_ _ _ .. .__l..__..r..a.._§_a T _ lama“: as, WPI-1% WPI Collagen Natural PABA Type of Casings 1.6 7 B B a. 1.4 l I 9 12 4 .2 ’ l A 3 1 't A T '_ l 1 E 0.8 1 ..=, l A , . g 0.4 l . ‘” 0.2 l I . _ , 0 if _ _,,, , .. . ,_.,‘H.,_ . ! __Tas - ,. . _, WPI-1% WPI Collagen Natural PABA Type of Casings Figure 4.6. Cook Yield (A) and Shear force (B) of hot dogs prepared with WPI- p- aminobenzoic acid (PABA), WPI, collagen, and natural casings. 170 the texture meat which most closely relates to human assessment. Consequently, our shear force results suggest that hot dogs prepared with WPI and collagen casings will be more chewable, softer and thus more acceptable to consumers than hot dogs prepared with natural casings. 4.3.9. Color The color of hot dogs plays an important role in consumer acceptance. L*, a*, and b* Hunter color values of interior and exterior hot dog surfaces are shown in Table 4.7. L* (lightness) and a* (redness) values significantly increased in interior samples from all hot dogs during storage (p<0.05); however, b* (yellowness) only increased for hot dogs encased in WPI-1.0% PABA after 28 days. Exterior yellowness of hot dogs with WPI casings was significantly higher than for hot dogs prepared with collagen and natural casings (Table 4.8). Exterior lightness (L*) of hot dogs prepared with WPI and collagen casings increased with storage time while lightness of hot dogs prepared with natural casing remained unchanged during 42 days of storage. Exterior redness (a*) and yellowness (b*) of WPI-1.0% PABA, collagen, and natural casings did not change throughout storage, whereas both of these attributes significantly increased for WPI-0% PABA casings. Based on analysis of variance, exterior lightness and yellowness of hot dogs significantly changed with casing type; however, storage time and casing type interaction only had effect on the lightness value (p<0.001) (Table 4.4). Lightness refers to the relation between reflected and absorbed light, without regard to a specific wavelength. Yellowness or redness results from differences l7l .Amcdvnc Ceasefire bagoumcwmm Pa Etofioasm Eastmw 5:5 39. oEmm 05 55$» mane—2 .Amucv :ocmgov wag—Sm H :32 172 a; H mad sod H mg a; H a; a; H a; .9 a; H as a; H mg a; H So .3. H 23 ”N .2332 1.335 .332 .SHSd E a; H o2 a 3 H a: a; H 8d .55 H $5 _ ax; mm2 amsa 5:32 58:8 :3 $5 go. 2%? a8 memmU mo ugh .0meon BEBwEB .«o 93¢ mv wage memwo REE: Ea .cowmzoo .53 A