PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 c:/CIRC/DateDue.p85-p.15 A COMBINED CONVECTION COOKING AND SALMONELLA INACTIVATION MODEL FOR GROUND MEAT AND POULTRY PRODUCTS by Adam Edward Watkins A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biosystems and Agricultural Engineering 2004 ABSTRACT A COMBINED CONVECTION COOKING AND SALMONELLA INACTIVATION MODEL FOR GROUND MEAT AND POULTRY PRODUCTS By Adam Edward Watkins A predictive model for moist-air impingement cooking of ground-and-formed meat and poultry products was developed. A coupled heat and mass transfer model incorporating the effects of fat transfer was combined with a model for Salmonella inactivation, to produce a complete prediction tool for meat and poultry processors. The model utilized the finite element method to numerically solve separate equations for heat, moisture, and fat transport. These equations were coupled through boundary conditions and interdependent thermo-physical property relationships. An enthalpy formulation for heat transfer was utilized to avoid discontinuities related to solid-to- liquid phase changes of water and fat within the product. Boundary conditions unique to moist-air impinging flow were incorporated into the model. These boundary conditions accounted for the additional heating effects of surface condensation that are common within moist air impingement systems. To complete the fat transport component of the model, laboratory experiments were conducted to determine the fat holding capacity of ground beef as a function of temperature and initial fat content. Fat holding capacities ranged from 0.05 to 0.6 g fat/ g nonfat dry matter, and a polynomial model was parameterized to those data. Additionally, laboratory-scale, moist-air convection cooking tests were conducted to confirm the importance of fat transport. Species and initial fat content significantly affected (P<0.05) cooking time, yield, and fat loss for ground beef, pork, and turkey. The heating time required to reach 85°C varied by as much as 217 seconds between different fat contents of the same species. Differences in cooking yield of up to 18% were measured between different fat contents. Fat transport was responsible for up to 28% yield loss in high fat products. Finally, cooking experiments using an industrial moist-air impingement cooking system were used to validate the temperature, moisture, and yield predictions of the complete cooking model. The cooking model predicted transient patty center temperatures with a standard error of prediction of 80°C for 54 cooking tests. At temperatures above 45°C, the standard error of the prediction was 5.8°C. The standard error for final moisture content predictions was 2.3% wet basis. Standard error for final cooking yield predictions was 6%. Additional comparisons were conducted between the cooking model and published data collected from moist-air impingement ovens. Data from published sources were used to perform a verification of the Salmonella inactivation predictions of the model. The standard error of prediction for Salmonella inactivation was 1.3 logs (CFU/ g). ACKNOWLEDGEMENTS Funding for this project was provided by the United States Department of Agriculture National Needs Graduate Fellowship Program. Additional assistance was provided by the United States Department of Agriculture CSREES National Food Safety Initiative. Access to the J SO-IV impingement oven was provided by FMC FoodTech of Sandusky, OH. I would like to thank Mr. Bob Swackhamer, Dr. Nahed Kotrola, and Mr. Todd Gerold for assistance in arranging for and conducting tests with the JSO-IV oven. Of course this project would not have been possible without the support of my major professor, Dr. Brad Marks, the members of my committee, the members of our research team, and all of the people that have helped me here or there along the way. Special thanks go out to my family and friends who stood by me through the past four years. iv TABLE OF CONTENTS LIST OF TABLES ................................................................................... viii LIST OF FIGURES .................................................................................... x KEY TO SYMBOLS .............................................................................. xviii 1 INTRODUCTION AND OBJECTIVES ............................................................. 1 1.1 Background .......................................................................................................... 1 1 .2 Objectives ............................................................................................................ 4 2 REVIEW OF LITERATURE .............................................................................. 5 2. 1 Introduction .......................................................................................................... 5 2.2 Introduction to cooking ........................................................................................ 5 2.3 Heat and mass transfer in meat during cooking ................................................... 6 2.4 Modeling the cooking process ............................................................................. 8 2.4. 1 Empirical models ...................................................................................... 10 2.4.2 Models based upon heat and mass transfer principles .............................. 12 2.4.3 Contact cooking ........................................................................................ 14 2.4.4 Frying ........................................................................................................ 20 2.4.5 Convection cooking .................................................................................. 23 2.4.6 Other types of cooking .............................................................................. 28 2.5 Thermal and physical properties required for modeling .................................... 31 2.6 Microbial models ............................................................................................... 33 2.6.1 Primary models ......................................................................................... 34 2.6.2 Secondary models ..................................................................................... 35 2.6.3 Tertiary models ......................................................................................... 37 2.7 Combined models .............................................................................................. 38 2.8 Limitations of models ........................................................................................ 40 2.9 Summary ............................................................................................................ 41 3 MATERIALS AND EXPERIMENTAL METHODS ....................................... 42 3.1 Overview ............................................................................................................ 42 3.2 Laboratory oven cooking tests ........................................................................... 42 3.2.1 Experimental procedure ............................................................................ 42 3.2.2 Statistical analysis ..................................................................................... 47 3.3 Measurement of fat holding capacity ................................................................. 48 3.3.1 Experimental procedure ............................................................................ 48 3.3.2 Statistical analysis ..................................................................................... 51 3.4 Industrial oven cooking tests ............................................................................. 52 3.4.1 Experimental procedure ............................................................................ 52 3.4.2 Statistical analysis ..................................................................................... 56 3.5 Cooking model validation .................................................................................. 56 3.5.1 Cooking model .......................................................................................... 56 3.5.2 Experimental data ..................................................................................... 57 3.5.3 Comparisons with literature data .............................................................. 58 3. 6 Salmonella inactivation model validation .......................................................... 59 4 MODEL DEVELOPMENT ............................................................................... 61 4. 1 Introduction ........................................................................................................ 61 4.2 Heat and mass transfer model ............................................................................ 64 4.2.1 Heat transfer solution ................................................................................ 64 4.2.2 Moisture transfer solution ......................................................................... 67 4.2.3 Fat transfer solution .................................................................................. 70 4.2.4 Heat and mass transfer coefficients - Array of slot nozzles ..................... 71 4.2.5 Heat and mass transfer coefficients — array of round nozzles .................. 74 4.3 Microbial inactivation model ............................................................................. 76 4.4 Finite element formulation ................................................................................. 77 4.4. 1 Introduction ............................................................................................... 77 4.4.2 Finite element basics ................................................................................. 77 4.4.3 Governing equations ................................................................................. 79 4.4.4 Finite difference time solution .................................................................. 86 4.4.5 Application of FEM solution .................................................................... 87 4.4.6 User interface ............................................................................................ 88 5 RESULTS AND DISCUSSION ........................................................................ 89 5.1 Overview ............................................................................................................ 89 5.2 Laboratory oven cooking tests ........................................................................... 90 vi 5.2.1 Cooking time ............................................................................................. 90 5.2.2 Cooking yield ............................................................................................ 95 5.2.3 Fat loss ...................................................................................................... 99 5.2.4 Volume change ....................................................................................... 103 5.3 Fat holding capacity experiments .................................................................... 107 5.3.1 Low fat samples (5.6% initial fat wet basis) ........................................... 108 5.3.2 High fat samples (15% initial fat wet basis) ........................................... 114 5.3.3 Summary ................................................................................................. 123 5.4 Industrial cooking tests .................................................................................... 124 5.4.1 Cooking time ........................................................................................... 124 5.4.2 Cooking yield .......................................................................................... 129 5.4.3 Fat loss .................................................................................................... 130 5.4.4 Volume change ....................................................................................... 132 5.5 Cooking model validation ................................................................................ 134 5.5.1 Finite element mesh ................................................................................ 134 5.5.2 Temperature profile-experimental data ................................................... 135 5.5.3 Temperature profile-published data ........................................................ 143 5.5.4 Moisture content- experimental data ...................................................... 145 5.5.5 Cooking yield — experimental data ......................................................... 146 5.5.6 Cooking yield - published data ............................................................... 148 5.6 Lethality model validation ............................................................................... 148 5.7 Illustration of model utility .............................................................................. 151 CONCLUSIONS .............................................................................................. 155 FUTURE WORK ............................................................................................. 158 APPENDICES ................................................................................................. 161 8.1 Model and experimental temperature versus time curves for moist-air impingement cooking of ground beef patties ................................................... 162 8.2 Derivation of cooking-air thermo-physical property equations ....................... 189 8.3 Product thermo-physical properties ................................................................. 199 8.4 Screen shots from Visual Basic cooking model user interface ........................ 201 8.5 Visual Basic model code .................................................................................. 203 BIBLIOGRAPHY ............................................................................................ 255 vii LIST OF TABLES Table 3.1. Treatment conditions utilized for model validation experiments ............... 54 Table 5.1 — Analysis of variance for cooking time of ground turkey patties as affected by temperature and fat content ...................................................................... 92 Table 5.2 - Analysis of variance for cooking time of ground beef patties as affected by temperature and fat content ..................................................................... 93 Table 5.3 - Analysis of variance for cooking time of ground pork patties as affected by temperature and fat content ..................................................................... 93 Table 5.4 — Analysis of variance of yield as a function of center temperature and fat content for ground turkey patties .............................................................. 95 Table 5.5 - Analysis of variance of yield as a function of center temperature and fat content for ground beef patties .................................................................... 96 Table 5.6 - Analysis of variance of yield as a function of center temperature and fat content for ground pork patties .................................................................... 98 Table 5.7 — Analysis of variance for fat loss as functions of center temperature and fat content for ground turkey patties .............................................................. 100 Table 5.8 — Analysis of variance for fat loss as functions of center temperature and fat content for ground beef patties ................................................................ 101 Table 5.9 — Analysis of variance for fat loss as functions of center temperature and fat content for ground pork patties ............................................................... 101 Table 5.10 — Analysis of variance for effects of temperature and initial fat content on volume change for ground turkey patties cooked in a laboratory convection oven. . 103 Table 5.11 - Analysis of variance for effects of temperature and initial fat content on volume change for ground beef patties cooked in a laboratory convection oven ......... 104 Table 5.12 — Analysis of variance for effects of temperature and initial fat content on volume change for ground pork patties cooked in a laboratory convection oven. . . ..106 Table 5.13 — Results from regression of fat holding capacity as fimctions of time and holding temperature for 5.6% fat ground beef. ........................................... 1 11 viii Table 5.14— Results from regression of fat holding capacity versus holding temperature for 5. 6% fat ground beef ............................................................ 112 Table 5.15 — Results from regression of fat holding versus time and holding temperature for 15% fat ground beef ............................................................ 117 Table 5.16 - Results of regression of fat holding capacity versus holding temperature for 15% fat ground beef ............................................................. 118 Table 5.17 — Linear regression of fat holding capacity as functions of heating temperature, holding time, and initial fat content .............................................. 121 Table 5.18 - Linear regression of holding capacity as a function of holding temperature and fat content ....................................................................... 122 Table 5.19 — Linear regression of patty center temperature as functions of oven temperature, steam content, cooking time, and oven airflow ................................ 128 Table 5.20 — Regression parameters for cooking yield as a function of oven temperature, steam content, cooking time, and airflow ....................................... 130 Table 5.21 — Regression parameters for cooking yield as a function of oven temperature, steam content, cooking time, and airflow ...................................... 131 Table 5.22 - Standard error of prediction for the entire trial (SEP) and for data above 45°C (SEP'D450C) for center temperature of beef patties .............................. 142 Table 5.23. Standard error of prediction for transient center temperature of ground chicken breast patties predicted by the model and by the regression equation of Murphy et al. (2001a) ............................................................................. 143 Table 5.24 — Difference between measured and predicted moisture content for each oven condition. Experiment numbers correspond to the conditions listed in Table 3.1..145 Table 5.25 — Difference between measured and predicted cooking yields for each oven condition. Experiment numbers correspond to the conditions in Table 3.1 . . . . . 146 ix LIST OF FIGURES Figure 3.1 — General arrangement of the laboratory convection oven showing directions of steam and airflow .................................................................... 43 Figure 3.2 - Interior of laboratory oven showing conditioning chamber ..................... 43 Figure 3.3 - Interior of laboratory oven showing fan, ducts, and sample chamber (located at left) ....................................................................................... 44 Figure 3.4 — (a) Picture and (b) schematic of jig used to place thermocouples into meat patties during cooking experiments ........................................................ 46 Figure 3.5 — Schematic of tube setup used for centrifuging ................................... 50 Figure 3.6 — Schematic of the centrifuge tube after centrifuging showing the two distinct layers that formed within the tube. In this schematic the heating tube has been removed ......................................................................................... 51 Figure 3.7 — Stein model J SO-IV moist-air impingement oven .............................. 53 Figure 4.1 - Illustration of the airflow within an impingement oven in relation to the product ............................................................................................ 62 Figure 4.2 — Geometry of a patty illustrating radial coordinate system ...................... 65 Figure 4.3 - Geometry of a ground meat patty with the modeled region and element mesh indicated ....................................................................................... 78 Figure 4.4 - Illustration of the finite element mesh utilized for the model .................. 78 Figure 4.5 — Illustration of a triangular element showing counterclockwise node numbering ............................................................................................ 81 Figure 5.1 - Center temperature as a function of cooking time for ground turkey patties cooked in a laboratory convection oven: means of 5 replicates ....................... 90 Figure 5.2 — Center temperature as a function of cooking time for ground beef patties cooked in a laboratory convection oven: means of 5 replicates..... . . . . . . . . . . . . ...91 Figure 5.3 — Center temperature as a function of cooking time for ground beef patties cooked in a laboratory convection oven: means of 5 replicates..... . . . . . . . . . . . . . ......91 Figure 5.4 — Yield as a function of endpoint center temperature for ground turkey patties of two fat contents: means of 5 replicates ................................................ 96 Figure 5.5 - Yield as a function of endpoint center temperature for ground beef patties of two fat contents: means of 5 replicates ................................................ 97 Figure 5.6 - Yield as a function of endpoint center temperature for ground pork patties of two fat contents: means of 5 replicates ................................................ 98 Figure 5.7 — Fat loss as a function of temperature for ground turkey patties of two fat contents: means of 5 replicates .................................................................. 99 Figure 5.8 - Fat loss as a function of temperature for ground beef patties of two fat contents: means of 5 replicates ............................................................... 100 Figure 5.9 - Fat loss as a function of temperature for ground pork patties of two fat contents: means of 5 replicates ............................................................... 102 Figure 5.10 — Relationship between product temperature and volume change for ground turkey patties cooked in a laboratory convection oven: means of 5 replicates ............................................................................................. 104 Figure 5.11 — Relationship between product temperature and volume change for ground beef patties cooked in a laboratory convection oven: means of 5 replicates ............................................................................................. 105 Figure 5.12 - Relationship between product temperature and volume change for ground pork patties cooked in a laboratory convection oven: means of 5 replicates ............................................................................................. 106 Figure 5.13 — Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 2 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................. 109 Figure 5.14 - Fat holding capacity as a function of temperature for 5.6% fat gr0und beef heated for 5 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................. 1 10 Figure 5.15 — Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 10 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................. 1 10 Figure 5.16 - Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 15 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................. 1 1 1 xi Figure 5.17 — Comparison of fat holding capacity calculated from regression model versus experimental values: means of 5 replicates ..................................... 114 Figure 5.18 — Fat holding capacity as a function of temperature for 15% fat ground beef heated for 2 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................ 1 15 Figure 5.19 - Fat holding capacity as a function of temperature for 15% fat ground beef heated for 5 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................. 1 -1 5 Figure 5.20 — Fat holding capacity as a function of temperature for 15% fat ground beef heated for 10 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................. 1 16 Figure 5.21 — Fat holding capacity as a function of temperature for 15% fat ground beef heated for 15 minutes at temperatures from 30 to 90°C: means of 5 replicates ............................................................................................. 1 16 Figure 5.22 - Comparison of fat holding capacity calculated from regression model versus experimental values (15% fat): means of 5 replicates ........................ 119 Figure 5.23 — Comparison of fat holding capacity calculated from regression model versus experimental values: means of 5 replicates ............................................. 123 Figure 5.24 —— Bottom and center temperature versus time for ground beef patties cooked at (a) oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s and (b) oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s .................................................................. 125 Figure 5.25 - Cooking yield as a function of endpoint center temperature for ground beef patties cooked in a Stein J SO-IV industrial moist air impingement oven. . . . . . . 129 Figure 5.26 - Yield loss not accounted for by moisture loss as a function of endpoint temperature for ground beef patties cooked in a Stein J SO-IV industrial moist-air impingement oven ................................................................................. 132 Figure 5.27 — Reduction in diameter during cooking as a function of cooking yield for ground beef patties cooked in a Stein J SO-IV industrial moist-air impingement oven .................................................................................................. 133 Figure 5.28 -— Example comparison of experimental temperature data with data generated by the model (oven temperature=121°C, steam content=50%, air velocity =11.4 m/s) ........................................................................................... 135 xii Figure 5.29 — Example comparison of experimental temperature data with data generated by the model (oven temperature=121°C, steam content=70%, air velocity =11.4 m/s) ........................................................................................... 136 Figure 5.30 - Example comparison of experimental temperature data with data generated by the model (oven temperature=121°C, steam content=8 8%, air velocity =11.4 m/s) ........................................................................................... 136 Figure 5.31 — Comparison between model Salmonella Senfienberg lethality predictions and data points published by Murphy et a1. (2002) .............................. 148 Figure 5.32 — Comparison between model Listeria innocua lethality predictions and data points published by Murphy et al. (2002) ................................................. 149 Figure 8.1 - Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment la) ............................................................... 162 Figure 8.2 — Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment lb) ............................................................... 162 Figure 8.3 — Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 6a) ............................................................... 163 Figure 8.4 — Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 6b) ............................................................... 163 Figure 8.5 — Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 8a) ............................................................... 164 Figure 8.6 - Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 8b) ............................................................... 164 Figure 8.7 — Oven temperature: 121°C, oven steam content 70% by volume, oven airflow: 11.4 m/s (Experiment 11a) .............................................................. 165 Figure 8.8 — Oven temperature: 121°C, oven steam content 70% by volume, oven airflow: 11.4 m/s (Experiment 11b) .............................................................. 165 Figure 8.9 — Oven temperature: 121°C, oven steam content: 70% by volume, oven airflow: 16.8 m/s (Experiment 13a) .............................................................. 166 Figure 8.10 - Oven temperature: 121°C, oven steam content: 70% by volume, oven airflow: 16.8 m/s (Experiment 13b) .............................................................. 166 Figure 8.11 — Oven temperature: 121°C, oven steam content: 70% steam volume, oven airflow: 21.8 m/s (Experiment 18a) ...................................................... 167 xiii Figure 8.12 — Oven temperature: 121°C, oven steam content: 70% steam volume, oven airflow: 21.8 m/s (Experiment 18b) ....................................................... 167 Figure 8.13 - Oven temperature: 121°C, oven steam content: 88% by volume, oven airflow: 11.4 m/s (Experiment 21a) .............................................................. 168 Figure 8.14 - Oven temperature: 121°C, oven steam content:88% by volume, oven airflow: 11.4 m/s (Experiment 21b) .............................................................. 168 Figure 8.15 - Oven temperature: 121°C, oven steam content: 78% by volume, oven airflow: 16.8 m/s (Experiment 23a) .............................................................. 169 Figure 8.16— Oven temperature: 121°C, oven steam content: 78% by volume, oven airflow: 16. 8 m/s (Experiment 23b) .............................................................. 169 Figure 8.17 -— Oven temperature: 121°C, oven steam content: 78% by volume, oven airflow: 21.8 m/s (Experiment 25a) .............................................................. 170 Figure 8.18 — Oven temperature: 121°C, oven steam content: 78% by volume, oven airflow: 21.8 m/s (Experiment 25b) .............................................................. 170 Figure 8.19 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment 30a) .............................................................. 171 Figure 8.20 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment 30b) .............................................................. 171 Figure 8.21 — Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 32a) .............................. 172 Figure 8.22 — Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 32b) .............................................................. 172 Figure 8.23 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 34a) .............................................................. 173 Figure 8.24 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 34b) .............................................................. 173 Figure 8.25 - Oven temperature: 177°C, oven steam content: 70% by volume, oven airflow: 11.4 m/s (Experiment 37a) .............................................................. 174 Figure 8.26 — Oven temperature: 177°C, oven steam content: 70% by volume, oven airflow: 11.4 m/s (Experiment 37b) .............................................................. 174 xiv Figure 8.27 - Oven temperature: 177°C, oven steam content: 83% by volume, oven airflow: 11.4 m/s (Experiment 47a) ............................................................. 175 Figure 8.28 - Oven temperature: 177°C, oven steam content: 83% by volume, oven airflow: 11.4 m/s (Experiment 47b) ............................................................. 175 Figure 8.29 - Oven temperature: 177°C, oven moisture content: 84% by volume, oven airflow: 16.8 m/s (Experiment 49a) ....................................................... 176 Figure 8.30 - Oven temperature: 177°C, oven moisture content: 84% by volume, oven airflow: 16.8 m/s (Experiment 49b) ...................................................... 176 Figure 8.31 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 16.8 m/s (Experiment 50a) .............................................................. 177 Figure 8.32 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 16.8 m/s (Experiment 50b) .............................................................. 177 Figure 8.33 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 21.8 m/s (Experiment 54a) .............................................................. 178 Figure 8.34 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 21.8 m/s (Experiment 54b) .............................................................. 178 Figure 8.35 — Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment 56a) .............................................................. 179 Figure 8.36 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment 56b) .............................................................. 179 Figure 8.37 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 58a) ............................................................. 180 Figure 8.38 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 58b) .............................................................. 180 Figure 8.39 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 63a) .............................................................. 181 Figure 8.40 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 63b) .............................................................. 181 Figure 8.41 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 11.4 m/s (Experiment 66a) .............................................................. 182 XV Figure 8.42 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 11.4 m/s (Experiment 66b) .............................................................. 182 Figure 8.43 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 16.8 m/s (Experiment 68a) .............................................................. 183 Figure 8.44 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 16.8 m/s (Experiment 68b) .............................................................. 183 Figure 8.45 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 21.8 m/s (Experiment 70a) ............................................................. 184 Figure 8.46 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 21.8 m/s (Experiment 70b) .............................................................. 184 Figure 8.47 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 m/s (Experiment 73a) .............................................................. 185 Figure 8.48 - Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 m/s (Experiment 73b) .............................................................. 185 Figure 8.49 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 m/s (Experiment 75a) .............................................................. 186 Figure 8.50 - Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 m/s (Experiment 75b) .............................................................. 186 Figure 8.51 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 16.8 m/s (Experiment 78a) .............................................................. 187 Figure 8.52 - Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 16.8 m/s (Experiment 78b) .............................................................. 187 Figure 8.53 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 21.8 m/s (Experiment 80a) .............................................................. 188 Figure 8.54 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 21.8 m/s (Experiment 80b) .............................................................. 188 Figure 8.55 - Latent heat of vaporization for water as a function of temperature (From tabular data: Geankoplis, 1993) ......................................................... 189 Figure 8.56 - Viscosity of air as a function of temperature (From tabular data: Geankoplis, 1993) .................................................................................. 190 xvi Figure 8.57 - Viscosity of steam as a function of temperature (From tabular data: Geankoplis, 1993) ................................................................................. 191 Figure 8.58 - Density of air as a function of temperature (from tabular data: Geankoplis, 1993) ................................................................................. 193 Figure 8.59 - Density of saturated steam as a function of temperature (From tabular data: Geankoplis, 1993) ........................................................................... 194 Figure 8.60 - Density of steam at 101.35 kPa as a function of temperature (From tabular data: Geankoplis, 1993) .................................................................. 195 Figure 8.61 - Thermal conductivity of air as a function of temperature (From tabular data: Geankoplis, 1993) ........................................................................... 196 Figure 8.62 - Thermal conductivity of steam as a firnction of temperature (From tabular data: Geankoplis, 1993) .................................................................. 197 Figure 8.63 — Input screen of cooking model user interface .................................. 201 Figure 8.64 — Output screen of cooking model user interface ............................... 202 xvii KEY TO SYMBOLS Lower-case letters “030068) D‘FWG QC- q '-1 3 xxx-wt:- “'1 .1 NN~<><2< Dry-basis fat content (g fat/ g dry matter) Initial dry-basis fat content (g fat/ g dry matter) F atzprotein ratio (g fat/g protein) Enthalpy (J/ g) Slot height (cm) Mass (g) Number of microorganisms (CFU/ g) Initial number of microorganisms (CFU/ g) Pressure (Pa) Saturation pressure (Pa) Partial pressure of water vapor (Pa) Ideal gas constant (cm3-Pa/g mol-K) N/No Relative humidity (%) Temperature (°C) Melting temperature (°C) Reference temperature (°C) Saturation temperature (°C) Volume (cm3) Slot width (cm) Mass fi'action (decimal) Dimensionless groups Nu Pr Re Sc Sh Greek letters ttQDPme Nusselt number (th/kr) Prandtl number (Cp’lJ/k'l‘) Reynolds number (D-v-p/u) Schmidt number (u/p-DAB) Sherwood number (hm'X/DAB) Thermal diffusivity (cmz/s) Emissivity (-) Error term Latent heat (J/ g) Density (g/cm3) Stefan-Boltzmann constant (5.676-10‘12 W/cmz-K4) Specific growth rate (l/s) Viscosity (Pa-s) xix 1 INTRODUCTION AND OBJECTIVES 1.1 Background Impingement cooking is used for manufacturing many fully and partially cooked meat and poultry products. Impingement ovens utilize a type of convection in which cooking air is directed normal to the product surface at high velocities (>10 m/s). This is generally accomplished by forcing the cooking air through an array of slots or nozzles. Impinging airflow allows for heat transfer rates an order of magnitude higher than those that occur in conventional convection ovens (Gardon and Akfirat, 1966). Moist-air impingement ovens combine impinging airflow with high humidity cooking air. In moist-air ovens, humidity can exceed 90% moisture by volume. The use of high moisture cooking air creates a condensing condition on the surface of the product. Condensing conditions dramatically increase the rate of cooking by taking advantage of the release of latent heat from steam in the cooking air. High moisture cooking air also suppresses product moisture loss during cooking. This results in cooking yields higher than those achieved when cooking with dry air. In meat cooking systems, a second cooking section with lower air moisture content is often utilized to aid in surface browning of the product. Development of a computerized model for moist-air impingement cooking of meat products is of interest for two major reasons. First, cooking models can be used as tools for optimizing cooking processes. Heat and mass transfer models can be utilized to optimize processes in terms of product temperature, yield, and quality. Optimizing the cooking process can have major economic significance. In the United States alone, there are 76 manufacturing facilities producing ground beef and poultry patties. The value of production from these plants exceeds $520 million per year (F818, 2001). Even modest increases in cooking yield stand to increase profits significantly. A one percent increase in yield could potentially increase annual profits by over $5 million. Reducing energy costs and wasted line capacity due to overcooking stands to improve profits even further. Reducing overcooking also stands to improve product quality characteristics such as color, flavor, and texture. The second and more important reason for modeling the cooking process is to ensure microbial safety of the cooked product. The Food Safety and Inspection Service (FSIS) has proposed regulations that would shift the emphasis of cooking regulations for ground meat and poultry products from a system of time-temperature standards to a system of performance standards (F818, 2001). Such regulations have already been enacted for whole muscle meat products (F818, 1999). The new performance standards are based upon a 7-10g10 reduction in viable Salmonella for ready-to-eat (RTE) poultry products and a 6.5-logro reduction in Salmonella for RTE beef, roast beef, and corned beef products. The new standards are designed to give processors the flexibility to develop customized, science-based processing strategies. However, it will be the duty of the processor to demonstrate that the process meets the required reductions in Salmonella. Unfortunately, adequate scientific tools are not currently available for processors to reliably and confidently verify compliance with the Salmonella performance standards. Processors need a tool that will allow them to determine the effects of processing on the reduction of Salmonella. Due to safety concerns, it is neither feasible nor desirable for manufacturers to conduct microbial challenge studies in the factory environment. Experiments can be conducted in the laboratory using microbial pathogens including Salmonella and Listeria monocytogenes in model food systems, but these results ofien do not correspond well with actual processing conditions. The use of mathematical models is an alternative to conducting microbial chellenge studies. Although mathematical models are not a substitute for the judgment of an experienced microbiologist, F SIS has indicated that mathematical models based on heat transfer equations can be used to demonstrate the effects of processing on bacterial inactivation (F818, 2002). Descriptive mathematical models for cooking processes would be of great value to processors and would allow for continuing innovation in the processed meat industry. Unfortunately, the development of cooking models is limited, with current models not accounting for the unique surface heat and mass transfer conditions present in moist-air impingement ovens (Chapter 2). Few existing models account for variations in product composition which can have significant effects on heat transfer and cooking yield. In most cases, yield loss is modeled solely as a function of moisture changes in the product during cooking. Although exceptions exist, most of the available cooking models have not been coupled with models for microbial inactivation. To be of maximum utility, cooking models should include all variables that affect microbial lethality, including product size, density, specific heat, thermal conductivity, product composition, humidity, and strain of the organism (F818, 2001). Models should be flexible and applicable to various cooking conditions without the need for generation of empirical data sets. Mass losses due to both moisture and fat loss during cooking should be considered. Finally, models should be validated against actual process data, including microbial inactivation data from inoculated challenge studies. 1.2 Objectives The overall goal of this study was to develop a cooking model that would meet the above criteria, as well as provide a graphical tool for illustrating the effects of processing to manufacturing plant personnel. The specific objectives of this study were to: 1. Develop a coupled heat and mass transfer model incorporating the transient effects of moisture and fat transfer during moist-air impingement cooking of ground beef patties. 2. Incorporate Salmonella inactivation models into the heat and mass transfer model and combine with a user interface to produce a tertiary cooking model. 3. Validate the heat and mass transfer model using data collected from an industrial moist-air impingement oven. 4. Validate the Salmonella inactivation models using data collected from inoculated challenge studies in a pilot-scale moist-air impingement oven. 5. Develop the user interface as a tool than can be used to illustrate the effects of cooking parameters on yield and microbial safety to non-technical personnel, including oven operators. 2 REVIEW OF LITERATURE 2.1 Introduction From the perspective of the processor, the most important output variables of cooking are temperature profile, cooking yield, and microbial inactivation. Cooking models can be used as a powerful tool for estimating these parameters. To model the reduction in pathogens such as Salmonella during cooking, it is first necessary to model the cooking process itself. The temperature history of the meat during cooking is the most important factor for modeling microbial reduction. As a result, heat transfer has been the focus of many models for cooking processes. However, in terms of profitability, the most important factor for processors to consider is cooking yield. Therefore, mass transfer is also an important factor to consider when modeling cooking processes. The first portion of this chapter summarizes previous research in the area of cooking models. The thermo-physical properties required for modeling are also discussed. The final portion of the chapter addresses microbial modeling and previous work combining microbial models with models for meat cooking. 2.2 Introduction to cooking Cooking and other heat treatments are among the most important unit operations in the food processing industry. Heat treatments during processing vary widely, depending upon the type of product. Foods may be fully or partially cooked, blanched to inactivate enzymes, dried to extend shelf life, pasteurized to kill unwanted microorganisms, or simply heated as part of the manufacturing process (Fellows, 1988). The types of devices used for cooking and heating are nearly as numerous as the number of products that can be produced. Examples include heat exchangers, deep-fat fryers, contact-cooking systems, convection ovens, microwaves, and infrared ovens. As a result of their ubiquity in the food industry, heating and cooking systems are of great interest to the food engineer. Ovens and other heating systems must be properly designed to ensure safe foods of the highest quality. Under-cooking may result in foods that are not safe to eat due to surviving pathogenic microorganisms. Over-cooked products may not have the quality characteristics demanded by customers. In addition to affecting product quality, overcooking can also be wasteful to the processor in terms of energy consumption and reduced product yield. Most engineering analyses of cooking processes are aimed at determining the temperature and/or moisture profiles of the product undergoing cooking. Meat products are generally cooked prior to packaging. During cooking of unpackaged products, heat is transferred into the product, resulting in increased temperatures and thermal gradients within the meat. At the same time, moisture is transported out of the product, resulting in moisture gradients and yield loss (with the exception of boiling, in which case moisture may be added to the product). Fat losses during cooking can also be significant (Young et al., 1991; Pan and Singh, 2001; Badiani et al., 2002). 2.3 Heat and mass transfer in meat during cooking Numerous studies have described heat transfer during cooking of meat (Dagerskog and Bengtsson, 1974; Dagerskog, 1979a and b; Skjoldebrand, 1980; Skjoldebrand and Hallstrom, 1980; Housova and Topinka, 1985; Thorvaldsson and Skjoldebrand, 1995; Pan and Singh, 2001; Shilton et al., 2002). Virtually every analysis presumes that the dominant mechanism of heat transfer within meat products is conduction. However, Shilton et al. (2002) proposed that for high fat ground meat, heat transfer may be related to both conduction and internal convection within the melted fat phase. Mass transport of water within meat products has also received considerable attention (Dagerskog and Bengtsson, 1974; Hung et al., 1978; Dagerskog, 1979a and b; Skjoldebrand, 1980; Skjoldebrand and Hallstrom, 1980; Mittal et al., 1982; Mittal et al., 1983; Hallstrom, 1990; Thorvaldsson and Skjoldebrand, 1995; Pan and Singh, 2001). Most cooking models assume that mass transport of water is primarily the result of molecular diffusion to a drying product surface (Huang and Mittal, 1995; Ngadi et al., 1997; Zanoni et al., 1997; Chen et al., 1999; Mittal and Zhang, 2000; Mittal and Zhang, 2001; Shilton et al., 2002). Hung et al. (1978) proposed a different mechanism for water transport during convection cooking of bovine semitendinosus muscles in which it was determined water loss during cooking of these muscles was most likely due to pressure forces caused by fiber shrinkage. Furthermore, water loss was only weakly dependent on oven temperature and appeared to depend primarily on the amount of muscle shortening during cooking. For samples cooked directly from the frozen state, an initial period of drip loss was observed. The water lost during this period was believed to be due to the melting of ice crystals within the muscle. During this initial drip period, sample orientation affected the amount of loss; with samples having muscle fibers oriented vertically producing the highest losses. In this period, gravity was believed to contribute to the water loss. For samples thawed before cooking, no such initial drip period existed. Thorvaldsson and Skjoldebrand (1995) also noted that water transport was up to 25% faster in the direction parallel to muscle fibers than across muscle fibers during oven roasting of bovine semitendinosus muscles. Some authors have taken to expressing moisture transport in meat in terms of "water holding capacity" (Dagerskog, 1979a and b; Pan et al., 2000). This represents the amount of water that a type of meat will contain, given a certain temperature history. Upon heating, the water holding capacity of meat products generally decreases, resulting in water losses. Although this type of analysis does not truly describe a mechanistic process of water transport, it coincides with the knowledge that protein releases water upon denaturing (Bodwell and McClain, 1978). Combined with the fiber squeezing mechanism described by Hung et al. (1978), the water holding capacity model provides a reasonable approximation of water transport during some cooking processes. However, this technique does not account for resistance to moisture transport within the product. This a significant weakness for most processes, since moisture diffusivity within meats is known to be low (Zanoni et al., 1997). Models that do not account for moisture diffusivity must be inherently empirical, as the mechanisms for transport within the product are not described based upon first principles. 2.4 Modeling the cooking process Models for cooking processes can be divided into two categories. They are either based solely on data from experimental studies (empirical models) or derived from theoretical formulas of heat and mass transfer (Hayakawa, 1970). Empirical models can be developed for cooking processes using data collected under the conditions of interest. These models require less rigorous mathematical analysis and can provide good predictive capability for a limited range of conditions. Additionally, empirical models do not generally require the researcher to have accurate thermal property data to get good results. The weakness of empirical models is that they are not generally applicable to situations that are different from the conditions for which the model was developed. Slight changes in the system, such as changes in product size or geometry, require an entirely new set of experimental tests to generate new model parameters. Although it is theoretically possible to construct empirical models with an unlimited number of input parameters, the number of inputs that can be meaningfully included in the model is limited in practice. Models based on theoretical principles are generally more robust than empirical models. Fundamental heat and mass transfer equations can be used to produce models that are much more general in nature than can be accomplished through experimental methods. By changing the boundary conditions of the model equations, a single model can often be used for several different types of cooking processes. However, models based on theoretical principles have their own drawbacks. Differential equations for heat and mass transfer quickly become too cumbersome to solve using analytical techniques. Thus, numerical methods are typically applied to cooking models. In order to produce models based upon transport equations, accurate knowledge of thermal and physical properties of the product must be known. In situations where properties are not accurately known, theoretical models end up being semi-empirical in nature, as the model constants must be adjusted to fit experimental process data. A number of different empirical and theoretical models are described in the following sections. 2.4.1 Empirical models Empirical models can be effective for predicting the temperature and moisture history of products during cooking. Bengtsson et al. (1976) developed a model for heat and mass transfer during oven roasting of meat. In the cooking process that was modeled, the meat was placed on a metal rack inside of an oven chamber with walls maintained at a fixed temperature. The mechanisms of heat transfer in this type of cooking are primarily natural convection and radiation. A plot consisting of a dimensionless temperature term (Tair-chnm)/(Tau-Tim,”) as a function of a dimensionless time term (kt)/(pcp12) on a semi-log plot was used to develop simulation equations. From the diagram, a prediction equation for meat temperature as a function of time was developed. Results of the comparison between predicted and experimental results were good; however, quantitative results were not given. Sarkin (1978) used a technique nearly identical to that of Bengtsson et al. (1976) to produce a computer cooking simulation for meat products. An equation for the linear portion of the heating curve was developed (Equation [2.1]). T - -T log[ 3" )=a+b-t [2.1] Tair - Tinitial 10 Coefficients a and b are constants and were determined by regression. The model had errors between the predicted and experimental temperatures of 02°C or less for cooking times of up to 10 hours. These results show the close predictions that are possible with empirical models. Erdogdu et al. (1999) developed an empirical model for predicting the yield loss of shrimp during immersion cooking. Shrimp were cooked isothermally for varying periods in water at temperatures between 65 and 95°C. Afier heating, the yield losses were measured, and an equation for yield loss was determined by multiple linear regression (Equation [2.2]). %YieldLoss=a+b-T+c-T2+d-t2+e-t+f-T-t [2.2] The terms a, b, c, d, e, and f were regression constants. The error between experimental and predicted yield losses ranged from 0.15 to 2.6% yield. This again shows the potential for empirical models to produce very close approximations of experimental data. The primary drawback to each of the empirical models is that they are only valid for a specific set of cooking conditions. Although the error of the models is very low, this is to be expected, as the same data that were used for model development were used for error calculation. The models are not applicable to conditions other than those used in the regression. The cooking models of Bengtsson et al. (1976) and Sarkin (1978) only allow for meat temperature to be predicted as a function of cooking time. No provision is made for predicting yield, and processing conditions are not considered, thereby limiting the usefulness of the models as tools for processors. The yield model of Erdogdu (1999) 11 allows yield to be calculated as a function of time and temperature, but temperature profile is neglected. All three models ignore the effects of product composition. In addition, the models are not readily adaptable to products of different sizes or geometries from those used for model development. The weaknesses illustrated by these models show the underlying reason that empirical models cannot be utilized as general models for the varying processing conditions that occur in industry. 2.4.2 Models based upon heat and mass transfer principles Most cooking models reported in the literature are based on theoretical heat and mass transfer equations. Models based on transport principles are generally more flexible than those developed through purely empirical methods. The equations that govern heat and mass transfer are not product specific and are generally accurate for a wide range of conditions. During cooking of most meat products, the predominant mechanism of heat transfer within the product is conduction. The governing equation for conductive heat transfer in rectangular coordinates is derived from an energy balance (Bird et al., 1960). 6T 5 0T a BI a BI O O —— = — k 0 — + _ k . — + — k . — 203 Various cooking processes can be simulated using the conduction equation. Most cooking processes transfer heat to the product via the food surface. As a result, different cooking types can be modeled by changing the surface boundary conditions as 12 appropriate for the type of cooking to be modeled. Models for conductive, convective, and radiative cooking can be constructed by including the appropriate heat flux boundary conditions. The flux equations for conduction, convection, and radiation heat transfer are well known and are given by Equations [2.4], [2.5], and [2.6] respectively. q 6T X = kT ' a surface [ 2'4] %= 111‘ '(Too — Tsurface) [ 2'5 ] q 4 4 X = 8 ' O ' (Tsurface '- Theat source ) I 2-6 I Most mass transfer models are based upon transport by diffusion. The governing equation for diffusion is similar in form to the equation for conduction (Bird et al., 1960). 99:3(Dm .§)+_6_[Dm .§]+£(Dm “QC—j [ 27] at 6x 6x 6y 8y 62 62 Treatments of surface mass transfer vary, but are typically modeled using a convective mass transfer boundary condition and/or a term for evaporation. The equation 13 for mass transfer by convection is analogous to convective heat transfer and is usually given by Equation [2.8]. n = hm '(Cequilibrium " C) I 28] Numerous authors have published theoretical models for meat cooking. Models have been developed for contact cooking (Dagerskog, 1979a and b; Ikediala et al., 1996; Pan et al., 2000), immersion frying (N gadi et al., 1997; Farkas et al., 1995a and b; Vijayan and Singh, 1997; Mittal and Zhang, 2001), convection cooking (Mittal and Blaisedell, 1982; Mittal et al., 1983; Holtz and Skjoldebrand, 1986; Huang and Mittal, 1995; Zanoni et a1, 1997; Chen et al., 1999; Mittal and Zhang, 2000), oven roasting (Singh et al., 1984), microwave cooking (Mallikarjunan et al., 1996), and infrared cooking (Shilton et al., 2002). The following sections describe models for each of these types of cooking. 2. 4.3 Contact cooking Contact cooking is one of the simplest mechanisms of cooking. During contact cooking, products are placed in direct contact with a heating surface, resulting in heat transfer by conduction. The rate of cooking is limited by the temperature gradient between the product and heating surface and by the surface resistance of the product/heating element interface. This type of cooking is common in the fast food industry, particularly for ground beef patties. 14 Dagerskog (1979a) developed an early model for contact cooking of meat patties. The model was based on a one-dimensional formulation of the conduction equation with a term added to incorporate the latent heat of evaporation of water. This equation was solved numerically using a finite difference technique. Transfer of moisture out of the meat was calculated based upon experimental measurements of water holding capacity. The capacity of meat to store water under varying times and temperatures was determined experimentally. Experimental determination of water holding capacity was conducted by heating 10 g samples enclosed in sealed plastic pouches in an isothermal water bath. Water content of the samples was measured afler they had been heated and allowed to drain for 1 minute on absorbent paper. Water holding capacities were then plotted as functions of temperature and time. For each time step in the model, it was assumed that the free water released due to the change in water holding capacity was transported out of the meat. Heat and mass transport equations were used to determine the amount of water that exited the meat due to evaporation. A second water loss equation was formulated using a mass balance based on changing water holding capacity. The difference in water loss due to changing water holding capacity and water lost by evaporation was attributed to drip loss. Differences between the experimental and simulated water losses ranged from 0.2 to 2.7% loss for a pan temperature of 140°C and from 0.3 to 2.9% loss for a pan temperature of 180°C. Predictions of the center temperature ranged from 0 to 2°C of the experimental values at cooking times of up to 6 minutes. The water holding capacity model has an advantage over other methods, in that it accounts for the varying water binding capacities of different products. It also has the 15 benefit of easily describing the drip loss phenomenon. Weaknesses of the water holding capacity model are that it is highly empirical and does not account for internal resistance to moisture transfer. Because it does not account for internal resistance to moisture transfer, the water holding capacity model is expected to decrease in accuracy with increasing product thickness. The water holding capacity model is also product specific, as different products have much different water binding properties. Ikediala et al. (1996) developed a model for single-side pan-frying of meat patties that was not based on the water holding capacity model, but rather a two-dimensional Fourier conduction equation formulated in radial coordinates with an added term for the evaporation of water. a -c -T M=£~§~£nkT~fl)+—a—(kT-§]+Np-M [2.9] r A contact heat transfer coefficient was used to describe heat transfer at the patty/ grill interface. The same thermal properties reported by Dagerskog (1979a) were used for calculations. The primary difference between the heat transfer component of this model and the model of Dagerskog (1979a) was the expansion of the conduction equation into two dimensions. This allowed for the temperature distribution in the patty during cooking to be modeled in two dimensions, resulting in a more complete description of the temperature profile. Deviations between experimental and predicted center temperatures were less than 4°C. 16 Ikediala et al. (1996) modeled moisture loss in the patty slightly differently than Dagerskog (1979a). The average moisture content of the patties was modeled using an exponential decay model. mave : mave,initial ' CXP(- (3 + b ' T) ' t) I 210] The constants a and b for the mass transfer equation were determined using regression of experimental values. The use of an exponential decay function for moisture transfer is completely empirical and represents a total deviation from first-principles. The result is that the model is specific for the product and conditions tested and lacks robustness. Pan et al. (2000) used a technique similar to that of Dagerskog (1979a) and Ikediala et al. (1996) to model cooking of hamburgers by two-sided contact cooking. Heat transfer was modeled using a one-dimensional formulation of the Fourier conduction equation. However, the conduction equation was written in terms of enthalpy, rather than temperature. fi=£[kT(H).flaifl) [ 2.11] Use of the enthalpy formulation allowed cooking from the frozen state to be modeled without discontinuities related to the phase change from ice to liquid water. Relationships between enthalpy and temperature were used to produce temperature history curves. l7 Enthalpy of frozen hamburger was calculated using a food property computer program (Singh and Mannapperuma, 1994). Enthalpy of unfrozen hamburger was calculated using an empirical equation (Equation [2.12]). Hnonfrozen =Href +p'I(l6OO+2600'XW +15'Xf 'T)'(T—Tref)} I 2-12] The changes in moisture and fat contents, independent of position, were determined using Equations [2.13] and [2.14]. = Thm ' (m - mequilibrium) I 2-13 I 9|? [2.14] 12L 2 —hf ' (F - 1:equilibrium) The model assumed that no transport of water or fat occurred below specified threshold temperatures. The temperature dependence of the water and fat equilibrium concentrations was given by Equations [2.15] and [2.16]. ( 6w(T-Tinma1)) [ 2.15] mequilibrium = minitial '6 18 Fequilibrium = Finitial 'e(_6F(T_Ti“m°l» [ 2-16 l The previous equations were solved at each node for every time step, giving moisture and fat contents as functions of temperature. A finite difference technique was used for solving the heat transfer and mass equations. As in the model by Dagerskog (1979a), this method treats water and fat content as "state" variables (i.e., functions of just temperature rather than results of transport processes). The 5“, and Stems are related to the water and fat holding capacities of the meat and were determined experimentally with whole patties. This model (Pan et al., 2000) is one of the most complete models for cooking available in the literature. Differences between predicted and center temperatures were small, although differences of up to 10°C occurred in the temperature range between 0 and 40°C. Differences between measured and predicted yields were less than 3%. Inclusion of fat transport as a separate mechanism sets the model apart from any of the prior contact-cooking models. In addition, utilization of the enthalpy formulation for heat transfer represents an important step in the simulation of temperature profiles for products that may originate in the frozen state. However, the empirical nature of the water and fat holding capacity models limits robustness. Water and fat holding capacity were determined using whole patties. In this way, internal resistance to moisture and fat transfer was accounted for indirectly. However, this approach limits the utility of the model to products of the same composition and geometry that was used to develop the model parameters. 19 2.4.4 F tying Frying is another operation that is commonly used to cook meat products. Immersion (deep fat) frying is typically modeled as a moving boundary problem (Singh, 2000). Products undergoing frying are described as consisting of an inner core region surrounded by a dry crust region. As cooking proceeds, the boundary between the crust and core regions moves towards the product center. Vijayan and Singh (1997) developed a model for heat transfer during flying of frozen foods. Frying was modeled as a moving boundary problem. Separate transport equations were utilized for the crust and core regions. Heat transfer in the crust was modeled using a one-dimensional formulation of the Fourier conduction equation. A convective boundary condition was used to describe heat transfer at the oil-product interface. The position of the crust-core interface at a given time was given by Equation [2.17]. A-t Sj+1=3j'*'chrust —CIcore]'r_ [2.17] pm) In the core of the product, heat transfer was modeled using an enthalpy formulation of the conduction equation (see Equation [2.11]). Differences between simulated and measured center temperatures ranged from 1 to 6°C over a range of 0 to 75°C. The model did not account for mass transfer. Therefore, although useful for predicting temperature, this model cannot be used for yield predictions. 20 A model for moisture transport during deep fat frying of chicken drums was developed by Ngadi et al.(1997). A two-dimensional diffusion equation formulated in radial coordinates was used to describe moisture transfer in the product (Equation [2.18]). 6K: 1 6 ac: a at: ___=._ _. .[, .__. +2.. I) .___ :218 at r arIr m ar] 62[ m 62) [ 1 The two-dimensional formulation of the conduction equation allows for modeling of non- spherical products. This is useful not only for chicken drums, but for other products such as chicken strips. An exponential function was used for the surface mass transfer boundary condition. This resulted in an empirical relationship similar to that used by Ikediala et al. (1996) C = Cinitial 'CXP((-a + b ° Ton ) ' t) l 2-19 1 The constants a and b were determined by regression of experimental data and were equal to -0.045 and 4.167-10“ respectively. The utilization of an empirical boundary condition greatly limits the robustness of the model. The finite element grid was broken up into elements representing the bone, bone marrow, cartilage, and muscle portions of the chicken drumstick. The moisture diffusivity of the muscle portion of the drumstick was modeled as a function of cooking oil temperature and moisture content using an equation developed in separate 21 experiments (Equation [220]); (Ngadi and Correia, 1995). Constant diffusivities were used for the other components. _ 2930 oil 1)m =8.35-IO'° ~exp[ —0.56l-C+O.092-C2] [ 2.20] This model represents a good tool for determining moisture distributions in a complex system undergoing cooking, although moisture predictions were off by up to 30% for some conditions. Unfortunately, this was not combined with a model for heat transfer. Nonetheless, the usefulness of the finite element method for modeling systems with multiple physical properties and complex geometries was demonstrated. Mittal and Zhang (2001) developed a model for deep fat frying using a different approach. An artificial neural network (ANN) was used to predict temperature, moisture, and fat content in meatballs during deep fat frying. Input parameters included fat diffusivity, moisture diffusivity, thermal diffusivity, heat transfer coefficient, fat conductivity, and oil temperature, as well as frying time, meatball radius, and initial temperature. The data used to train the artificial neural network were generated from validated mathematical models. Heat, moisture, and fat concentration were modeled using one-dimensional transport equations formulated in radial coordinates. The maximum errors between modeled and experimental temperature, moisture, and fat content were 1.9°C, 0.004% dry basis, and 0.016% dry basis, respectively, for the optimum ANN design. 22 Artificial neural networks have several advantages over other types of modeling for predicting temperature and moisture distributions during cooking. ANN ’s can "learn" from new data to increase accuracy of prediction. This allows experimental data to be combined with simulation data from models. Neural networks can often generate results faster than mathematical models, which is useful in optimization studies where many data sets must be analyzed. However, in many cases where artificial neural networks are generated from simulation data, the question exists whether the actual simulation data would be more accurate than the data generated from the ANN. In this case, it would seem to make more sense to use the data of the original simulation rather than the ANN. In addition, theoretical models can be more readily adapted to new conditions. 2.4.5 Convection cooking Forced-air convection cooking is an important method for commercial cooking of meat products. Huang and Mittal (1995) developed a computer model for forced convection cooking, broiling, and boiling of meatballs. Heat and moisture transfer were modeled using one-dimensional conduction and diffusion equations formulated in radial coordinates. An energy balance at the surface of the meatball was used for the heat transfer boundary condition. This equation accounted for convective heat gain, latent heat of evaporating water, and conduction at the meatball surface. 81‘ am kT :5; surface =hT °(Tair “Tsurface)+ Dm 'p'l'glsurface I 2-21] 23 Boundary conditions for the moisture transport equation were based upon moisture content equilibrium between the surface and the environment. The transport equations were solved using the finite difference method. Constant transport properties were used in the simulation. Surface heat transfer coefficients for each type of cooking were determined using analysis of the center temperatures of aluminum spheres heated under the conditions of interest. The thermal and mass diffusivity values were estimated by minimizing the root-mean-square deviations between the observed and predicted temperature and moisture histories. This was conducted using only one set of cooking data and validated using four other data sets. The average root mean square errors for temperature and mass prediction versus experimental results were 3-5.1°C and 0.04-0.19 g respectively. The strength of this model is that it was derived entirely from transport equations. Although an empirical technique was used to fit thermal and mass diffusivity values, a similar technique is often used to determine the published values that are available for those constants. Chen et al. (1999) developed a model for convection cooking of chicken patties. The finite element method was used to solve heat and mass transfer equations based on transport principles. Cooking was modeled using two-dimensional transport equations formulated in radial coordinates. The use of two-dimensional formulations allowed for complete description of the temperature and moisture distributions in the patties. 9.91: 1.9. r.k_T.§T_ +2 k_T.§_T_ [2.22] at r 6r c 8r 62 62 24 9.92: 1.13. r.l‘_m_.@ +3 5&2“. [2.23] at r 6r cm (it 62 dz The two-dimensional formulation of the heat and mass transfer equations made it possible to model cylindrical patties as axis-symmetric bodies. The boundary condition for the heat transfer equation consisted of both a convective heat transfer term and a latent heat term for surface evaporation. kT- =hT-(Tair—T)+Dm-p-K-—— [2.24] The boundary condition for the mass transfer equation was based upon convective mass transfer at the surface. = h... -(m... -m) I 2251 The equilibrium moisture term used in the mass transfer boundary condition was calculated using the equation developed by Huang and Mittal (1995). - 5222.47 -1 .0983 RH = ”W _— ' mequilibrium I 2-26 I R g ' Tair 25 Based on previous research, heat capacity and thermal conductivity were modeled as functions of temperature and moisture content respectively (Murphy et al., 1998; Murphy and Marks, 1999). This decreased standard error of prediction for center temperature from 5.5 to 37°C, as compared to using constant values for Cl, and k. cp =3017.2+2.05-T+0.24-T2 +0.002r3 [ 2.27] kT =0.194+0.436-m [ 2.28] Singh et al. (1984) developed a heat and mass transfer model for oven roasting of meat. Heat transfer within the meat was modeled using a two-dimensional conduction equation. 2 2 flza. £4.22 [2.29] A convective boundary condition with a term for evaporation of water was used at the product surface. 6T h ' (Tair — T): k ' hm '(Psurface " Pair )+ kT ' El surface I 2-30] 26 The finite difference method was used to solve the heat transfer equation. No attempt was made to describe the transport of water within the product. A linear relationship was used for the value of Psurfacc (Equation [2.31]). I)surface =3 +b'T [ 2.31] Although a term for evaporation was used to improve the heat transfer model, this model is not suitable for predicting yield during cooking. A different technique, was utilized by Mittal and Zhang (2000) to develop a model for convective thermal processing of fiankfurters. This model was based upon an artificial neural network. Input variables for the ANN were fat-protein ratio, initial temperature, initial moisture content, frankfurter radius, ambient temperature, relative humidity, and process time. The artificial neural network was trained using data from validated mathematical models. The models were based upon one-dimensional heat and mass transfer equations. Moisture diffusivity and equilibrium moisture content as functions of temperature, fat protein ration, and relative humidity were used in the model following the work of Mittal and Blaisdell (1982): Dm = Exp(— 8.6787+0.08468.FP—0.3614-RH~FP—4341.5/Tabs +8.55-C) [ 2.32] mequilibrium = —0.102- 1n(- Rg oFP - (T + 5.665) ln(RH)/1.132 .107) [ 2.33 ] 27 The ANN was trained using 13,500 data points generated from the mathematical models. Fat-protein ratio was found not to be an important factor in predicting moisture content or temperature. As discussed earlier, artificial neural networks pose some advantages over purely numerical models, and may be effective for use in optimization experiments. The primary advantage of artificial neural networks is that they can be utilized to produce results much faster than models based upon the finite difference or finite element methods. As a result, ANN ’s have promise for optimization studies, real- time control, and other situations where fast calculations are essential. However, ANN ’s are still limited in that they must be “trained” with data either from experimental or model sources. The theoretical basis of numerical models means that they can be more easily adapted to changing processing conditions without the need for “training”. 2. 4. 6 Other types of cooking Models are available for several other types of cooking. Mallikarjunan et al. (1996) developed a model for microwave cooking of shrimp. Heat transfer was modeled using a two-dimensional transport equation in axial coordinates. 2 2 Earl—a. .1..£+_6_I+6_T + Q [234] 1' a!- arz 2 p.cp where Q is the heat generated by the microwaves and is given 28 Q = Q0 .[exp[‘_(15£'_r)] + exp[:(_§_:i)fl [ 2.35 ] P P Moisture loss during cooking was calculated using the following equation: am 9 ' V ' cp 5T -—=hm-A-(P.at—Pw)+——- — [2.361 Heat transfer at the surface was modeled using a convective heat transfer equation with a term for water evaporation. Equation [2.36] is somewhat unusual in that the evaporation term is generally included in the heat transfer equation. The result is that Equation [2.36] does not account for internal resistance to moisture transport. Equations were solved using the finite difference method. Constant thermal properties were used. The surface heat transfer coefficient was calculated using the equation for natural convection over a horizontal cylinder: 0.25 AT] [ 2.37] h =1.3196- — T (Ad Simulation results were validated by collecting transient temperature and mass data during cooking in a household microwave oven. The model temperature predictions were within 6°C of the experimental values. Mass losses were within 0.8% of the predicted values. 29 Another type of cooking uses far-infrared radiation for heating. Shilton et al. (2002) developed a model for far-infrared cooking of beef patties. The model accounted for heat transfer and evaporative mass losses during cooking. Heat transfer was modeled using a one-dimensional Fourier equation with a term for the latent heat associated with water evaporation. Mass transport was modeled by a diffusion equation. The boundary condition for heat transfer at the surface was given by: (Yr 4 4 kT ' E‘— = 0" Theating element _ Tsurface I 238 I A convective boundary condition was used for the mass transfer equation. Dm 'E=hm '(Cair ‘Csurface) I 239] The mass diffusion coefficient was calculated using an equation by Maroulis et al. (1998). 1),, = a.exp(b/T)-exp(c/c) [ 2.40] Thermal conductivity and density were modeled as a function of temperature using equations by Choi and Okos (1985). 30 k=a+b-T—CoT2 [2.41] p=a—b-T—c-T2 [2.42] The authors noted that the heat transfer model was not accurate for patties containing high levels of fat. To more accurately model heat transfer in high-fat patties, an effective thermal conductivity was calculated. 2 k=(a+b-T—c-T )+(heffl+heff2T) [2.43] This technique allowed more accurate modeling of heat transfer for meat containing high levels of fat. However, it did not actually relate heat transfer to fat content or describe the transfer of fat out of the product. Addition of a more detailed fat transfer component would increase the utility of the model. 2.5 Thermal and physical properties required for modeling Before it is possible to create cooking models based on heat and mass transfer principles, accurate values for the thermal and physical properties of meat products must be known. Since these values may vary widely with temperature and composition, it is necessary to quantify these effects. Thermal and physical properties of a wide range of food products are available from food engineering textbooks and handbooks, such as those published by ASHRAE (Stroshine and Hamann, 1996; ASHRAE, 1998). Although 31 the properties of interest can often be obtained, care must be taken to scrutinize the conditions for which they are accurate. Numerous articles have been published on the thermal properties of meat products. These properties were reviewed extensively by Sanz et al. (1987). Experimental values for thermal conductivity, enthalpy, apparent heat capacity, and density of beef, pork, mutton, poultry, and fish were reported. Most of the property values listed were in the -30 to 30°C temperature range. Other articles related to thermal properties of meat and meat products have been written by McProud and Lund (1983), Perez and Calvelo (1984), Dincer (1996), and Tsai et al. (1998). Moisture diffusivity data for a wide range of foods were compiled by Zogzas et al. (1996). Enthalpy data for meat products have been published by Levy (1979), Skala et a1. (1989), and Tocci and Mascheroni (1998). The rate of moisture transport in meat products is considerably lower than the rate of heat transfer. Zanoni et al. (1997) reported that moisture diffusion affected only a 3 mm deep layer of the product surface. Chen et al. (1999) reported similar results. These results give insight into the development of element meshes for future cooking models. The region of highest activity for moisture transfer is confined to a layer close to the surface of the patty, and thus a fine mesh should be used in that area. Accurate values for thermal and physical properties are critical to development of cooking models based on transport equations. Small deviations in pr0perty values often result in large differences in model performance. Thus, care must be used when selecting property data from published sources. The product composition can often have a dramatic effect on physical properties. Thus, composition-dependent property equations, such as those developed by Choi and Okos (1986), are often utilized. 32 2.6 Microbial models Mathematical microbial models can provide important tools for predicting the growth or reduction of microorganisms in foods. When properly utilized, models can provide an initial estimate of microbial behavior without the investment in time and materials required for microbial challenge studies. The use of models is also less expensive. Although microbial challenge studies may still be required to verify model results, the model may be used to more efficiently design such studies by selecting the conditions to be tested. Models can be used to quickly determine the effects of process changes on microbial food safety. These capabilities are invaluable for the planning of hazard analysis critical control programs. Graphical depictions of changes in microbial counts can serve as educational tools. This is especially valuable for showing non- microbiologists the effects of processing on microbial safety. However, the limitations of the model must be considered whenever a model is used to predict microbial activity in food products (F818, 2002). Mathematical modeling of microbial activity in food products has been extensively reviewed in the literature (Whiting and Buchanan, 1994; Whiting, 1995; Roberts, 1997). Whiting (1995) classified microbial models based upon a three-level scheme-as primary, secondary, or tertiary. Primary models describe the changes in a microbial population as a function of time. Secondary models describe the primary model parameters as functions of environmental parameters. Tertiary models combine primary and secondary models with a user interface to produce a complete simulation tool. 33 2. 6.1 Primary models Primary models describe the number of bacteria in a population as a function of time. Mathematical equations are used to describe the growth or inactivation curve using a set of parameter values. Ideally, these parameters relate to descriptive terms such as lag time or generation time. Examples of primary models are linear models, exponential models, and models based on the Gompertz function (Whiting, 1995). A log-linear inactivation model is given by Equation [2.44]. logN=logN0+—Il)--(t—t0) [2.44] Equation [2.45] is an inactivation model based on first-order kinetics. =e ' [2.45] A primary model for microbial inactivation based on a Gompertz equation was parameterized by Van Impe et al. (1995). y 2: a - exp[— upturn—2"“: - (k — t)+ 1)] I 2-46] 34 Peleg and Cole (1998) suggested a primary inactivation model based on a Weibull distribution (Equation [2.47]). log10 3(1) = —b(T) . 1"”) [ 2.47 ] Each type of primary model has its own merits and inherent weaknesses. The log-linear model is the most commonly used inactivation model in both industry and academia. Large quantities of data have been amassed using the model parameters for numerous microorganisms and processing conditions. This makes implementation of the log-linear model the easiest of any of the primary models. Additionally, log-linear models have become ingrained as the “standar ” for microbial inactivation and are thus widely accepted in industry. However, microbial inactivation does not always follow log-linear kinetics (Peleg, 1997). Other types of inactivation models may be more suitable for modeling inactivation in these cases. Unfortunately, these equations are often more cumbersome to use, and suitable experimental data for model constants are not as widely available as for log-linear models. To develop modeling software that is both flexible and as widely acceptable as possible in nature, log-linear inactivation kinetics should be utilized. However, to develop the most accurate model for specific cases, other types of primary models may be superior. 2. 6.2 Secondary models Secondary models describe the changes in primary model parameters as a function of environmental conditions. These models can show the effects of variables 35 such as temperature, pH, water activity, and substrate composition. Examples of secondary models are Arrhenius relationship models, response surface models, and square root models (Whiting, 1995). An Arrhenius relationship for the inactivation constant, k, was given by Geankoplis (1993). -133 k=a-e R-T [2.48] Equation [2.49] is a square-root relationship for k (Whiting, 1995). JE=a(T—T0) [2.49] Mattick et al. (2001) parameterized the following log-logistic secondary models for Salmonella. The models are designed to be used with the primary model in Equation [2.47]. b(T) = 6.841/{1 + exp[(76. 14 — T)/4.204]} [ 2.50 ] n(T) = 0.670/{1+ exp[(T - 80.14)/3.785]} . [ 2.51 ] 36 Juneja and Eblen (1999) developed a response surface model for inactivation of Listeria monocytogenes (Equation [252]). The model described decimal reduction time as a function of temperature, pH, salt content, and sodium pyrophosphate content. ln(D) = c1+ c2 - (T) + c3 - (pH) + c4 - (salt) + c5 ~(phos) + c6 - (T) - (pH) + c7 - (T) - (salt) + cs - (T) - (phos) + c9 - (pH) - (salt) + 010 - (pH) - (phos) +c11-(salt)-(phos)+ en -(T)2 +c13 -(pH)2 +c14 -(sa1t)2 +c15 -(phos)2 [ 2.52] 2. 6.3 Tertiary models Tertiary models combine primary and secondary models with a user-interface to to produce a complete simulation tool. Two widely used tertiary models in the United States are the American Meat Institute Process Lethality Determination Spreadsheet (AMI, 2003) and the United States Department of Agriculture's Pathogen Modeling Program (USDA, 2004). The AMI model is based upon a Microsoft Excel spreadsheet (AMI, 2003). The spreadsheet uses a log-linear thermal death time model to calculate process lethality based on time/temperature data inputted by the user. The user must also input 2 and Tm]- values for the thermal death time models. A table of z and Trey values for common meat microorganisms is supplied to aid the user in choosing input values. Unfortunately, 2 values may be influenced by numerous conditions that are not accounted for by the chart of suggested values. Thus many users will likely choose z and Tref values that are 37 inappropriate for the cooking process of interest and therefore generate lethality data that are highly suspect. The Pathogen Modeling Program (version 7.0) is a menu-based program that is based upon a suite of models (USDA, 2004). The program provides tools for assessing microbial inactivation of Clostridium botulinum, Escherichia coli 0157:H7, and Listeria monocytogenes, but not Salmonella. The models for microbial inactivation and survival are based upon a logistic inactivation model (Equation [2.5 3]). + 5131+”) (17 all + e—bzflj N (1 log-N?)- = log 12+ 3131041)) + (1+ 61320-111) [ 2.53 ] 2.7 Combined models Several authors have attempted to combine cooking models with microbial inactivation models. Zanoni et al. (1997) linked inactivation models for Enterococcus faecium to a cooking model for Bologna sausage. Two different inactivation models were utilized. The first was a simple first order inactivation model. —=—k-N [2.54] The second was a model developed by Whiting and Buchanan (1994). 38 a-(1+exp(-b1-t1)) (1-a)°(1+°"p(-b2 OW] I 2.55] The model was validated by inoculated challenge studies. At log reductions below 6, both the first-order inactivation model and the Whiting and Buchanan model predicted reductions within 1- log of the experimental values. At reductions above 6-log, the first order model greatly overestimated inactivation. However, at log reductions above 6, the Whiting and Buchanan model was within l-log of the experimental data. Unlike the first- order model, the Whiting and Buchanan model incorporates a tailing effect at high levels of inactivation, which provided a better fit to the experimental data. This clearly illustrates how certain inactivation models may be more accurate than others under some conditions. The contact-cooking model developed by Pan et al. (2000) included a first-order inactivation model for E. coli OlS7:H7. dN - 2.303 a“ = Dre, .lorrref-Tonz ' [ 2'56] Temperature data from each time step were utilized to obtain the number of microbes at each nodal location. The total surviving population at each time step was then determined. However, the inactivation model was not validated using experimental data. 39 Mallikarjunan et al. (1996) included an inactivation model for Listeria monocytogenes with a model for microwave cooking of cocktail shrimp. The inactivation model assumed first-order reaction kinetics. .N [2.57] The inactivation model was validated by injecting a liquid inoculum into the geometric center of each shrimp. After cooking, the shrimp were tested for surviving bacteria. This methodology made it impossible to track the actual numbers of surviving Listeria as a function of time. Although useful for predicting the worst-case scenario, this technique probably overestimates the cooking required for most products where microorganisms are present either at the surface or dispersed throughout the product. 2.8 Limitations of models Although microbial models can be valuable tools, there are limitations inherent in any model. A model is generally only accurate for the range of conditions under which it was developed. Extrapolating outside of the ranges used for the model development may give misleading results. Models are generally microbe specific, and a model for one microorganism cannot be expected to produce accurate results for another microorganism, or even for a different product. 40 2.9 Summary The models contained in the previous sections describe cooking processes with varying degrees of complexity. Although each of the models discussed has its own merits, none of the published models provide a “complete” description of the cooking process. Ideally, a cooking model should be based entirely upon engineering first- principles, be flexible for a wide range of products and product conditions, and describe the interrelationships between all of the components of the cooking process including heat transfer, moisture transfer, and the transfer of fat. The goal of the following sections was to develop such a model. 41 3 MATERIALS AND EXPERIMENTAL METHODS 3.1 Overview The study was broken up into three major groups of experiments. In the first group of experiments, ground beef, ground pork, and ground turkey patties were cooked in a laboratory convection oven. The results of these experiments were used to illustrate and quantify the differences in cooking characteristics between species and fat content. In the second group of experiments, laboratory studies were conducted to determine the fat holding capacity of ground beef. The data generated in these experiments were later used to develop the fat transfer portion of the cooking model (Chapter 4). The third set of experiments involved cooking ground beef patties in an industrial moist-air impingement oven. These data were then analyzed to determine the effects of processing conditions on yield and temperature profiles. In addition, results from this group of experiments were utilized to validate the computer-cooking model that is described in Chapter 4. 3.2 Laboratory oven cooking tests 3. 2. 1 Experimental procedure A series of laboratory experiments was conducted to investigate the effect of meat species and initial fat content on cooking characteristics during convection cooking of ground beef, ground pork, and ground turkey patties. All cooking tests were conducted using a custom built, laboratory convection oven. 42 The laboratory convection oven consisted of three chambers (Figure 3.1). These included a large conditioning chamber (Figure 3.2), a steam generator, and the cooking chamber itself. Steam generator Conditioning chamber with air heaters Sample chamber with suspended sample Figure 3.1 - General arrangement of the laboratory convection oven showing directions of steam and airflow. Figure 3.2 - Interior of laboratory oven conditioning chamber. 43 The conditioning chamber was the largest portion of the oven and was used to heat and condition the cooking air to the desired temperature and moisture content. The dimensions of the conditioning chamber were approximately 83 cm in length by 56 cm in width by 51 cm in height. The conditioning chamber contained four 350-watt strip heaters (McMaster-Carr: Cleveland, OH). Steam was injected into the conditioning chamber from the steam generation unit. The steam generation unit contained water heated by a 750-watt immersion heater (Tempco: Wood Dale, IL). An electronically activated solenoid valve was utilized to inject steam into the conditioning chamber. The sample chamber was a small container located at the edge of the conditioning chamber. The dimensions of the cooking chamber were approximately 10 cm by 10 cm by 10 cm. Cooking air was drawn from the conditioning chamber by a 6-watt centrifugal fan (Dayton model 4C440: Niles, IL) and passed through the sample chamber by means of tubular ducts (Figure 3.3). Figure 3.3 - Interior of laboratory oven showing fan, ducts, and sample chamber (located at left). The oven was connected to a computer interface that continuously monitored and controlled oven temperature and moisture content. During cooking, oven dry bulb and wet bulb temperature were controlled within i 0.2°C. Dry bulb and wet bulb temperature within the conditioning chamber was monitored using a high temperature dry bulb/wet bulb humidity probe (V iasala model DMP 246: Viasala, Woburn, MA). The airflow in the cooking chamber was 1.3 m/s. Ground turkey, ground beef, and ground pork were purchased from a local grocery store. Additional ground pork was provided from the Michigan State University Meat Laboratory. Two fat contents of each meat species were utilized: 1.4 and 8.6% for ground turkey, 7.2 and 17.5% for ground beef, and 15.7 and 41.9% for ground pork. Fat contents were determined in triplicate by solvent extraction (AOAC method 991.36: AOAC, 2000). The moisture content of each species was determined in triplicate by oven drying (AOAC method 950.46: AOAC, 2000). Moisture contents were 74.8 and 73.0% wet basis for the 1.4 and 8.6% fat ground turkey, 71.5 and 63.3% wet basis for the 7.2 and 17.5% fat ground beef, and 64.1 and 43.6% wet basis for the 15.7 and 41.9% fat ground pork. Each meat type was formed into uniform patties by pressing into plastic petri dishes (52 mm diameter; 13 mm height) . The patties were then frozen and removed from the dishes prior to use in the cooking experiments. Before cooking, patties were tempered to 4°C by placing in a refrigerator for 2-3 h. Prior to heating, each patty was weighed to the nearest 0.01 g. The radius and thickness of each patty were then measured, to the nearest 0.1 m, using a digital caliper. 45 A 24-gauge type-K thermocouple (Omega: Stamford, CT) was then inserted into the geometric center of each patty using a placement jig (Figure 3.4). (a) (b) \ AA \ if Thermocouples Frame Screen Figure 3.4 — (a) Picture and (b) schematic of jig used to place thermocouples into meat patties during cooking experiments. Patties were then placed into the cooking chamber of the oven, where they were supported on a wire mesh screen. This allowed for airflow on all sides of the product. The patties were cooked one at a time in the convection oven to center temperatures of 45, 55, 65, 75, and 85°C at an oven temperature of 177°C and a wet bulb temperature of 82°C. Five patties were cooked at each condition. Alter cooking, the patties were removed from the oven and weighed to the nearest 0.01 g. The new thickness and diameter of each patty were measured using the caliper, and the patties were frozen at - 5°C pending further analysis. After 24 h, the patties were removed from the freezer, and the moisture content was measured by oven drying (AOAC Method 950.46: AOAC, 2000) 46 The amount of water lost during cooking was calculated for each patty using the initial and final moisture contents (Equation [3.1]). ANImoisture = Minitial 'Xw,initial “Mfinal 'Xw,final I 3-1 I where Minitial and MW. are the initial and final mass of the patty and XW is the wet basis moisture content of the meat. These data were used to calculate the component of yield loss attributed to moisture. For many of the samples, the mass lost due to moisture loss was considerably less than the total mass loss. The mass loss not accounted for by moisture was attributed to fat loss during cooking. These losses were calculated using Equation [3.2]. AM fat = AM total _ AM moisture I 3-2 I 3.2.2 Statistical analysis Analysis of variance (ANOVA) was used to evaluate the effects (or=0.05) of initial fat content, cooking time, and time-fat interaction on center temperature, cooking yield, fat loss, and volume change within each meat species. AN OVA was conducted using the Microsoft Excel Data Analysis Package (Microsoft Excel Version 2000:Redmond, WA). 47 3.3 Measurement of fat holding capacity 3.3.1 Experimental procedure Laboratory experiments were conducted in an isothermal water bath to determine the fat holding capacity of ground beef. The fat holding capacity of the meat is the maximum amount of fat bound in the meat after a given heat treatment. The fat holding capacity of the meat was determined as a function of temperature and time using a so- called “net test” (Barbut, 1996). Two lots of ground beef were acquired from the Michigan State University Meat Laboratory. The fat contents of the two lots were 5.6 and 15.0% fat by mass, determined in triplicate using solvent extraction (AOAC method 991.36: AOAC, 2000). Brass tubes were used to contain the meat samples during cooking and were chosen due to their high thermal conductivity. The tubes were cut from 0.36 mm thick brass tube stock (K&S Engineering: Chicago, IL). The tubes had an inside diameter of 7.9 mm and a length of approximately 122 mm. Prior to each heating test, a silicone stopper was placed in one end of each heating tube. Teflon tape (13mm width) was then wrapped around the silicone stopper to ensure that the closed end of the tube was watertight. The weight of the combined tube, stopper, and tape was then measured to the nearest 0.1 mg using an electronic balance. Approximately 3.5g of meat was then loaded into each tube by hand. During loading, the meat was packed into the closed end of the brass tube by firmly tapping the tube on the bench top. After loading, the total weight of the tube and meat was measured to the nearest 0.1 mg. 48 A 30 um nylon mesh (Spectrum Labs: Rancho Dominguez, CA) was then attached over the open end of each tube using laboratory tape. A type-T thermocouple probe (Cole Parmer: Vernon Hills, IL) was inserted through the silicone stopper and into the approximate center of the meat sample within the tube. Tubes were then placed into an isothermal water bath (N eslab: Newington, NH) for one of four holding times. The open end of each tube was held above the surface of the water by a test tube rack. The entire meat sample was located in the submerged portion of the brass tube. The time measurement was started when the temperature of the center of the sample was within 1.0°C of the water bath temperature. Samples were heated at water temperatures of 30, 35, 40, 45, 50, 55, 60, 65, 70, 75, 80, 85, and 90°C and heating times of 2, 5, 10, and 15 minutes. Five tubes were heated simultaneously for each time-temperature combination. After heating, the tubes were removed from the water bath, and the surface moisture was removed with a paper towel. The tubes were then placed, mesh-side down in centrifiige tubes (length: 102 mm, diameter: 14.7 mm: Fisher: Pittsburgh, PA). A brass spacer fabricated from the same stock used to produce the heating tubes was placed in the bottom of each centrifirge tube (Figure 3.5). The length of the brace spacers was approximately 27 mm. Prior to centrifuging, the silicone stoppers and Teflon tape were removed from each tube. The tubes were centrifuged at 25°C and 1000g for 15 minutes. 49 \I ‘— centrifuge tube ‘— brass tube containing meat ‘__. tape *— 30 um nylon mesh ‘— brass spacer Figure 3.5 — Schematic of tube setup used for centrifuging. After centrifuging, two distinct layers were observed in each tube. A solid layer of extracted fat was located above a liquid layer (Figure 3.6). For the purposes of this study, only the fat located in the solid layer was of interest, as this is the component most often overlooked by cooking models. To separate the solid fat from the liquid layer, the brass tube containing the meat sample was first removed from each tube. A syringe with an 18-gauge needle (BD: Franklin Lakes, NJ) was then used to withdraw the liquid layer from beneath the solid fat layer. The centrifuge tube containing the spacer and the fat layer was then heated in a 102°C convection oven for 24 hours to drive off any remaining water. After heating, the tube was removed from the oven and allowed to cool in a dessicator for one hour. The tube was then weighed to the nearest 0.1 mg, and the weight of fat in the tube determined by subtracting the weight of the empty tube and spacer from the weight of the tube containing the spacer and extracted fat. 50 solid layer liquid layer Figure 3.6 — Schematic of the centrifuge tube after centrifuging showing the two distinct layers that formed within the tube. In this schematic the heating tube has been removed. The liquid layer contains water and soluble proteins. The solid layer is composed of fat. The fat content remaining in each sample after centrifuging was determined by subtracting the mass of the extracted fat from the initial mass of fat in the sample prior to heating. It was assumed that no protein was present in the extracted fat. The fat content was expressed in terms of dry basis fat content by dividing the mass of fat remaining in the sample by the mass of non-fat dry matter in the sample. This fat content was considered the fat holding capacity of the meat under the specific conditions of temperature and time. 3.3.2 Statistical analysis Quadratic response-surface models were used to describe the fat holding capacity of each batch of meat as functions of time and temperature (Equation 3.3). F=BO+Bl-T+[32-T2+B3-t+B4-t2+BS-t-T+a [3.3] 51 The statistical significance (or=0.05) of each regression parameter was determined using t-statistics. The models were then adjusted accordingly to eliminate non-significant terms. A regression model was also fit to the combined FHC results of the low and high fat ground beef (Equation [3.4]). F=B0 +B1'T+B2°T2+B3't+B4't2+B5'F+B6't'T+B7'T'FO +B8'I'F0 [ 34] The statistical significance (or=0.05) of each regression parameter was determined using t-statistics, and the model was again adjusted to eliminate non-significant terms. 3.4 Industrial oven cooking tests 3. 4. 1 Experimental procedure A series of cooking tests were conducted using a commercial continuous-feed moist-air impingement oven (Stein model J SO-IV2FMC FoodTech, Sandusky, OH) located at the FMC FoodTech Technical Center (Sandusky, OH). The JSO-IV oven consisted of two 3.65 m long cooking sections with a 0.8 m intermediate section. The belt width of the oven was 1.02 m. The oven was equipped with 1.07 m long in-feed and discharge sections, which were not utilized during this study. 52 Figure 3.7 - Stein model JSO-IV moist-air impingement oven. Commercially produced, pre-formed, frozen hamburger patties (Gordon Foods, Grand Rapids, M1) were used for the cooking tests. The patties had a thickness of 1 cm and average diameters of ~12 cm. The average uncooked patty weight was 105 g. The labeled fat content was 10%. This fat content was confirmed using solvent extraction (AOAC method 991.36: AOAC, 2000). Tests were conducted using a 1/3 partial factorial design (Table 3.1) comprised of three cooking temperatures (121, 177, and 232°C), three oven steam contents (50, 70, and 90% steam by volume), and three oven fan speeds (50, 75, and 100% of full). For each cooking condition, three oven belt speeds were chosen to produce varying degrees of doneness (undercooked, fully cooked, and over cooked). Belt speeds were chosen based upon the experience of a trained operator and adjusted as required to achieve the desired patty temperatures. Two hamburger patties were cooked at each belt speed, for a total of 54 patties. 53 Table 3.1. Treatment conditions utilized for model validation experiments. Experiment Temp. (°C) Steam by Volume Air Flow Time (Min.) m/s 1 121 50 % 11.4 5 121 50 % 16.8 11 8 121 50 % 21.8 8 11 121 70 % 11.4 8 13 121 70 % 16.8 5 18 121 70 % 21.8 11 21 121 88 % 11.4 11 23 121 78 % 16.8 8 25 121 78 % 21.8 5 30 177 50 % 11.4 8 32 177 50 % 16.8 6 34 177 50 % 21.8 3 37 177 70 % 11.4 3 47 177 83 % 11.4 6 49 177 84 % 16.8 3 50 177 86 % 16.8 6 54 177 86 % 21.8 8 56 232 50 % 11.4 5 58 232 50 % 16.8 2 63 232 50 % 21.8 7 66 232 70 % 11.4 7 68 232 70 % 16.8 5 70 232 70 % 21.8 2 73 232 82 % 11.4 2 75 232 82 % 11.4 7 78 232 82 % 16.8 7 80 232 82 % 21.8 5 Prior to cooking, the frozen patties were tempered in a -3°C cooler (> 3 h) to provide a uniform initial temperature distribution. Before each cooking run, the patties were weighed to the nearest 0.1 g. The radius and diameter of each patty were measured to the nearest 0.1 m using a digital caliper. A 24-gauge type-K thermocouple (Omega: Stamford, CT) was then placed and held at the center of each patty using a placement jig 54 (Figure 3.4). Additional thermocouples were positioned at the top and bottom surfaces of the patties and held in place by the placement jig. The thermocouples were attached to an oven data logger (Datapaq model 9000: Datapaq, Wilmington, MA). that was able to pass through the oven with the patties during cooking. During the cooking process, patty and air temperatures were recorded every 1 second. Patties were placed on the center of the oven belt and allowed to cook for the designated time. The patties were then removed from the oven and immediately weighed to the nearest 0.1 g to determine post-cooking weight. The thickness and diameter of each patty was then re-measured to the nearest 0.1 mm. Afler weighing and measuring, the patties were frozen (-20°C) using an impingement freezer and placed into sealed plastic bags pending firrther analysis. Dwell time in the impingement freezer was 2 minutes. After each cooking run, the data were uploaded to a personal computer, and temperature versus time curves were constructed. The bulk moisture content of each patty was determined using oven drying (AOAC Method 950.46: AOAC, 2000). The cooking yield was calculated for each patty by dividing the mass of the cooked patty by the mass of the patty prior to cooking. Using the moisture content of the raw and cooked meat, the mass of water that exited the patty during cooking was calculated (Equation [31]). 55 3.4.2 Statistical analysis Multiple linear regression was conducted to quantify the effects of the cooking variables on the center temperature achieved in the patties during cooking. Equation [3.5] was used as the basis for the regression. Tcenter = '30 + Bl ' Toven + I32 ' Xsteam + B3 ' time + B4 ° Airflow + 8 I 3-5 I A second regression was performed in an attempt to quantify the effects of cooking parameters on cooking yield. Regression was performed using Equation [3.6] as a basis. Yield = [30 + [31 woven + [32 -xS,,,m + [33 . time + [34 - Airflow + e [ 3.6] 3.5 Cooking model validation 3.5.1 Cooking model A computer model was developed to simulate moist-air impingement cooking of ground-and-formed meat and poultry patties. Development of this model is detailed in Chapter 4. The model was based upon heat and mass transfer principles and did not require experimental data for its development, with the exception of the relationships utilized for the fat holding capacity. However, it was necessary to validate the model using experimental data to validate its performance. This was conducted using data generated in the industrial cooking tests (Section 3.4) and additional published data from a pilot-scale impingement oven. 56 3.5.2 Experimental data The finished computer model was validated using the data generated in the industrial impingement cooking tests described in Section 3.4. The computer model requires user input of oven temperature, steam fraction, air velocity, and cooking time. Additional inputs required are product initial temperature, product moisture content, product fat content, and meat species. The computer model was executed using the 27 sets of process conditions listed in Table 3.1. The input conditions for the ground beef were an initial temperature of —2 to 0°C, initial moisture content of 66% wet basis, and initial fat content of 10% wet basis. The element mesh was set up for a patty diameter of 12 cm and thickness of 1 cm (see Chapter 4 for element mesh). Output data were generated for the transient patty center and surface temperatures, final bulk moisture content, and final cooking yield. The output were compared to the temperature profiles generated during the cooking experiments as well as to the final yield and moisture values measured after the cooking tests. A standard error of prediction (SEP) between the transient model and experimental temperature profiles was calculated for each run (Equation [3.7]). SEP = J: (Tpredictecrll : Fll‘measured)2 [ 3.7 ] The differences between the predicted and measured yield, moisture content, and fat loss were determined for each model run. The aggregate standard error of prediction 57 was also calculated for the endpoint yield, moisture content, and fat loss for all of the model runs. 3. 5.3 Comparisons with literature data In addition to the model validation using experimental data from the industrial cooking tests, the cooking model was compared to published data for ground chicken breast patties (Murphy et al, 2001a and b). The purpose of this comparison was to demonstrate the versatility of the model. The published data were collected using a pilot- scale moist-air impingement oven (Stein model 102: Stein, Sandusky, OH). Murphy et al. (2001a) developed a regression model for cooking time as a firnction of center temperature, oven air velocity, and oven wet bulb temperature (Equation [3.8]). ln(t) = 8.8678 + 0.0278 - T — 2.0410 - ln(wa) — 0.2306 - Vair + 0.0481 - ln(wa °Vair) [3.8] The regression for cooking time had an R2 of 0.95. Murphy et al. (2001b) developed a second regression model for yield of ground chicken patties as a function of patty center temperature, oven air velocity, and oven air wet bulb temperature (Equation [3.9]). The regression model for yield had an R2 of 0.89. Yield = 75.4031 + 0.9309 - T — 1.2443 . Va], + 0.1047 . vai,2 - 0.0121 -r2 — 0.0102 - M .v,,, + 0.0027 - wa .r [3.9] 58 The geometry of the Stein 102 oven used by Murphy et al. (2001a and b) was different from the geometry of the Stein JSO-IV. The Stein 102 oven used an array of round nozzles rather than the slot nozzles utilized in the J SO-IV oven (Chapter 4). Therefore, it was necessary to modify the cooking model with an equation for an array of round nozzles as discussed in Chapter 4. Model simulations were conducted using the process conditions of Murphy et al. (2001a and b). These conditions were a dry bulb cooking temperature of 149°C, wet bulb air temperatures of 40, 70, 85, and 95°C, and air velocities of 1.53, 2.13 and 2.73 m/s. The product settings for the model were chicken at an initial moisture content of 80% wet basis, initial fat content of 0.2% dry basis, and initial temperature of 4°C. The element mesh was adjusted for a patty diameter of 127 mm and a thickness of 12.7 mm. Cooking model tests were run using each combination of dry bulb temperature, wet bulb temperature, and airflow giving a total of 12 cooking conditions. The cooking time of each simulation was adjusted to give a final patty temperature of 80°C. For each model run, the temperature profile from the model was compared to the temperature profile generated by the regression model at 1 second intervals for the temperature range between 55 and 80°C (the calibration range of the regression model). Transient standard error of prediction was calculated for each run using Equation [3.7]. The same procedure was used to calculate the SEP for transient yield value for each run. 3.6 Salmonella inactivation model validation The location of the impingement oven utilized for our experiments did not allow for inoculated challenge studies. Therefore, simulated Salmonella Senftenberg S9 inactivation results were compared to literature values for ground meat patties cooked in a pilot-scale moist-air impingement oven (Murphy et al., 2002). Simulated results for Listeria innocua were also compared to literature values (Murphy et al., 2002) to demonstrate the flexibility of the model. The model was run using the equations for the geometry of the Stein 102 oven (array of round nozzles). The cooking air conditions utilized for the model runs were an oven air temperature of 288°C, a steam content of 25% by volume, and an air velocity of 4 m/s. Although the patties used by Murphy et al. (2002) were composed of a mixture of ground beef and ground turkey, the ground beef setting was utilized for the model runs. This setting was chosen based upon the fairly high (20% wet basis) fat content of the meat. The product properties were an initial moisture content of 58% wet basis, initial fat content of 20% wet basis, and initial temperature of -2°C. The model was run once for inactivation of S. Senftenberg and once for inactivation of L. innocua. The D and z-values for the model runs were set to 312 seconds and 623°C for S. Senftenberg and 1,508 seconds and 490°C for L. innocua, at a reference temperature of 60°C (Murphy et al., 2003). The reference temperature for both sets of D and 2 values was 60°C. The experiments of Murphy et al. (2002) utilized ground meat with an initial inoculum of 107 CFU/ g. Therefore, the model was set to run with a limit of 7-log total reduction. Transient inactivation curves generated by the model for each organism were plotted on the same graphs as data from the published study and SEP’s calculated between the model and experimental data. 60 4 MODEL DEVELOPMENT 4.1 Introduction A computer model was developed for moist-air impingement cooking of ground and formed meat and poultry products. The computer model is composed of a coupled heat and mass transfer mode] combined with a model for Salmonella inactivation. The finite element method (FEM) was used to numerically solve the differential equations associated with the heat and mass transfer model. The model was programmed using Microsoft Visual Basic (Version 6.0 professional edition: Microsoft, Redmond, WA). A typical impingement oven has impinging jets located above and below the product (Figure 4.1). The cooking air enters the ductwork and is forced through slots normal to the product. Not shown in Figure 4.1 is the perforated belt that the product travels on through the oven. A typical impingement oven would contain many pairs of impingement jets arranged in series along the length of the belt. The computer model was developed for an oven with the following controllable variables: airflow, air temperature, steam content by volume, and cooking time. The distance between the impinging jets and the product surface must also be specified. For the purpose of modeling, it was assumed that the upper and lower impinging jets were located an identical distance from the product surface. 61 Input airflow Impinging airflow fed . through slots __ Product suspended on - perforated belt (belt not 4 shown) '1‘ \ — I. Input airflow _ ,t Figure 4.1 - Illustration of the airflow within an impingement oven in relation to the product. The coupled heat and mass transfer model was based upon three transport components. The three components were heat transfer, moisture transfer, and fat transfer. Incorporation of the fat transfer component represents one of the unique aspects of the cooking model as compared to most previous models that have been published. The three transport solutions were coupled through the boundary conditions and interdependent thermo-physical properties. The boundary conditions associated with moist-air impingement ovens are somewhat different than those in conventional convection ovens. One of the unique aspects of moist-air impingement ovens is the very high oven air moisture content that is often used for meat product cooking. The high fraction of water in the air results in a condensing boundary condition during the period in which the surface of the patty is below the dew point of the cooking air. This condensing condition increases the rate of 62 heat transfer and reduces the rate of moisture loss from the product. Inclusion of the condensing boundary condition is another unique aspect of this cooking model. Two different sets of boundary conditions were used to model the cooking process — one for the impinging flows on the horizontal surface of the patties and one for the vertical edges of the patties. The first set of boundary conditions was used to describe the impinging flow that is the predominant mechanism for heat and mass transfer in moist-air impingement ovens. The boundary conditions used to model impinging flow were based on an expansion of the technique of Millsap and Marks (2002) and depended on transport correlations published by Martin (1977). The second set of boundary conditions was used to describe the convection that occurs along the vertical surfaces (edges) of the patty. These conditions were based upon transport correlations for turbulent flow past a flat plate (Bejan, 1995). The following set of basic assumptions was utilized when developing the heat and mass transfer model. 1. Heat transfer occurs by conduction within the patty and by a combination of convection, condensation, and evaporation at the patty surface. 2. Moisture and fat transfer within the patty occur primarily by diffusion and capillary flow, respectively. 3. When the temperature of the patty is below the dew point, moisture condenses at the patty surface, and the moisture content of the patty surface does not change. When the surface temperature of the patty is above the dew point, water evaporates at the surface. 63 4. The phase change from solid to liquid water occurs at a single temperature. A phase change from solid to liquid fat occurs at a single fat melting point. 5. Mass transport of water within the patty does not occur at temperatures below the freezing temperature. Mass transport of fat does not occur below the melting point of the fat. 6. Fat transport within the patty is driven by the gradient between the local fat content and the value at the surface. Surface fat content is a function of time, temperature, and initial fat content. 7. The size and shape of the patty does not change during cooking. 8. Boundary conditions were assumed to be the same for the top and bottom of each patty. 4.2 Heat and mass transfer model 4. 2. 1 Heat transfer solution The heat transfer simulation was formulated using the Fourier equation written in terms of enthalpy. The enthalpy formulation was utilized to produce a model free from discontinuities caused by the phase changes of fat and water. Similar techniques for modeling heat transfer have been utilized by Volner and Cross (1981), Pan (1998), and Pan et al. (2000). Most ground beef and poultry patties are shaped like short cylinders (Figure 4.2). This geometry allows heat and mass transfer in formed patties to be modeled using two- dimensional axisymmetric solutions. Heat transfer by conduction within the patties was modeled using Equation [4. 1]. 17 '1 Figure 4.2 - Geometry of a patty illustrating radial coordinate system. $1.43,. kT 2! ,3 RT 95 [4.1] at rar cp-pdr dch-pdz Heat capacity (cp), thermal conductivity (k7), and density (p) were modeled using transient values based on product composition and temperature (Section 9.3). Patties were assumed to be at a uniform internal temperature prior to cooking. The initial product temperature was converted to enthalpy using an equation based on the technique of Voller and Cross (1981): (Equation [4.2]). rCPsfmzen O T T < Tfreezing H =< cp.frozen ' Tfreezing T = Tfreezin g cp,frozen ° Tfreezing + )‘w + cp ' (T ' Tfreezing) Tfreezing < T <= Tmelting,fat [cp,frozen ' Tfreezing + Aw + cp ' (T — Tfreezing )+ A'f T > Tmelting,fat [4.2] 65 Equation [4.2] converts temperature values into enthalpy based upon an absolute zero reference temperature. Provision was made for two phase changes within the product during cooking. The first was the phase change between ice and liquid water, and the second was the phase change between solid and liquid fat. An initial temperature equal to the freezing temperature of the meat was assumed to imply a completely frozen sample. A similar technique was used to convert enthalpy values back to temperature for display purposes (Equation [4.3]). The variables op; and cm in Equation [4.3] are the heat capacities of the frozen and thawed meat, Tf and Tan are the freezing temperature of water and the melting temperature of fat, and Agw and M; are the latent heats of fusion for water and fat respectively. Based on the results of the fat holding capacity experiments described in Section 3.3, a value of 45°C was used for the fat melting temperature. H S Cp,f °Tf Cp,f -Tf S H rH/c Tr pf Scpf 'Tf +kw (H—kw —273‘Cp,f)/Cp,t Cp’f 'Tf +1“, < H Scpf °Tf +7tf +Tf 'cp,t Tm cp,f-Tf+}tf+Tf-cp,t Cp,f -Tf +Af +Tm °cp,t +Af [4.3] 66 The boundary conditions for heat transfer were composed of a convection term and a moisture transport term (Equation [4.4]). 6T kT ' 5n- = hT ' (Tair - Tsurface ) + h m,water ° 9“vaporization ° (C air " Csurface) [4.4] During the period in which the surface temperature of the meat is below the dew point temperature of the cooking air, heat transfer due to condensation occurs, causing additional heat transfer into the patty. When the surface temperature of the meat exceeds the dew point temperature, evaporation occurs at the patty surface, limiting sensible temperature increases in the product. This has the effect of limiting the maximum surface temperature reached at the product surface as a function of oven conditions and surface moisture content. Until the surface of the patty reaches low moisture contents due to drying, the surface temperature never exceeds the wet bulb temperature. At the vertical and radial centerlines of the product, heat transfer was assumed to be zero due to product symmetry (Equations [4.5] and [4.6]). 6T _=0 :0 4.5 6r r l ] §I=0 Z=0 [46] dz 4. 2.2 Moisture transfer solution Moisture transfer within the product was modeled using a two-dimensional equation for diffusion in radial coordinates (Equation [4.7]). 67 am 1 a k m,water 6m 5 k m,water am —: —..— r.————o-——— +———- —————--—— [4.7] at r at cm,water ' p 51' 62 cm,water ' P 62 The units for mm... are decimal dry basis moisture content. Mass transfer at the product surface was modeled using a convective boundary condition (Equation [4.8]). 8’ [4.8] k m,water an = h m,water ' (C air _ Csurface) The concentration of moisture in the cooking air was calculated using Equation [4.9]. [4.9] Cair : psteam ' Xair X3], is the molar fraction of steam in the cooking air. This is equal to the percentage of steam by volume. The density of steam (psteam) was calculated using an equation derived from tabular data (see Section 9.2). pmam = —0.0002 - log(T)+ 0.0015 [ 4.10] The concentration of moisture at the patty surface was calculated using Equation [4.11]. 68 Csurface : ERH ' Csat I 4-11 I Equilibrium relative humidity (ERH) was calculated using an equation by Huang and Mittal (1995). _ . -l.0983 ERH =ex 5222.47 111 [ 4.12] 1.9818-T In Equation [4.12], m is dry basis moisture content, T is absolute temperature, and ERH is decimal equilibrium relative humidity. The saturation concentration (Csat) was calculated using Equation [4.13]. Equation [4.13] was developed using regression of steam table data as shown in Section 9.2. csat = 8.121 - 10'10 .r3 —3.520.10'8 -'r2 41.320.10‘6 .r +6.215-10-7 [ 4.13] At the radial and vertical centerlines of the patty, mass transfer was assumed to be zero due to product symmetry (Equations [4.14] and [4.15]). —=0 r=0 [4.14] 69 92:0 2:0 [415] dz 4. 2.3 Fat transfer solution Fat transfer was modeled using a two-dimensional formulation of Darcy’s law for diffusion of liquids through porous media (Datta, 2002). 6F 1 6 BF 6 6F —=["‘a—r'(r'Dcap,fat '_)+—Z‘(Dcap.fat “an I 4'16] Fat content was written in terms of dry basis fat content. Dcapfat is the capillary diffusivity of fat in the product. The fat content at the surface of the patty was modeled using an equation generated from experimental values as detailed in Section 3. Equation [4.17] was utilized to set the values of the fat content at each boundary node as a function of temperature and product composition. This equation was derived from experimental data (Section 5.3.) F = 0.7062 — 0.0193 - T + 0.0001 - T2 + 0.0069 . Emma, + 0.0002 - T . PM,l [ 4.17 ] At the radial and vertical centerlines of the patty, fat transfer was assumed to be zero due to product symmetry (Equations [4.18] and [4.19]). 70 ——=0 r=0 4.18 &_ l ] §E=0 2:0 [4J9] dz 4. 2.4 Heat and mass transfer coefficients — Array of slot nozzles Heat and mass transfer coefficients utilized in Equations 4.4 through 4.8 were modeled as functions of instantaneous process conditions using the technique of Millsap and Marks (2002). For the surfaces of the patty subject to impingement conditions (the top and bottom), correlations developed by Martin (1977) for an array of slot nozzles were used to determine heat and mass transfer coefficients. Prandtl number, Schmidt number, and Reynolds number were first calculated using Equations [4.20], [4.21], and [4.22] respectively. c . . . Pr: p,m1x “mix [4.20] kmix Se=——“ii’£— [ 4.21] Pmn'DAB v- . . .w Re: J“ pm“ [4.22] “mix 71 Values for the heat capacity (emu), viscosity (um), thermal conductivity (kmix), density (pm), and diffusivity (D5,) of the air-steam mixture were calculated using the relationships in Section 9.2. The (W) term in Equation [4.22] is the slot width of the impinging jets. The term, v, in Equation [4.22] is the jet exit velocity. Nusselt and Sherwood numbers were calculated using Equations [4.23] and [4.24], respectively (Martin, 1977). Z 3 3 Nu=3-Pr°-42-f0:i- Lif— [ 4.23] 3 % + 1% 0 Z 3 3 Sh=—:—-Sc0'42-f02- Z'Re [4.24] %.+f% where f0 is a filnction of the slot geometry, and is given by Equation [4.25]. f0 =(60+4-(%—2)2)_1/2 [ 4.25] 72 In Equation [4.25], H represents the spacing between the impinging jets and the product surface, and W represents the slot width of the jets. Heat and mass transfer coefficients were calculated from the Nusselt and Sherwood numbers using Equations [4.26] and [4.27], respectively. Nu-k - h =——-——-m1x 4.26 T W l ] Sh-D hm,water = w 5“ [ 4.27] For the vertical surfaces of the patty (the patty edges), correlations for turbulent flow past a flat plate (Bejan, 1995) were used to calculate heat and mass transfer coefficients. Turbulent flow was assumed due to the flow regime created by configuration of the oven ductwork and turbulent air coming off of the top and bottom surfaces of the patty. Reynolds number was calculated using Equation [4.28] where (E) is the edge height of the patty. Rezw [4.23] ”mix 73 Prandtl and Schmidt numbers were calculated using Equations [4.20] and [4.21], respectively. Nusselt and Sherwood numbers were calculated using Equations [4.29] and [4.30]. 4 1 Nu = 0.037 ~ Re/S-Pr/3 [ 4.29] 4 l Sh = 0.037 - Re/5-Sc/3 [ 4.30] Heat and mass transfer coefficients were then calculated using Equations [4.31] and [4.32], respectively. hT 2M [ 4.31 ] E Sh . D hm,water = E sa [ 4.32] 4. 2.5 Heat and mass transfer coeflicients — array of round nozzles In addition to the correlations for an array of slot nozzles, additional correlations were utilized to predict the heat and mass transfer coefficients for an array of round nozzles (Martin, 1977). Nusselt and Sherwood numbers were calculated using equations [4.33] and [4.34], respectively. 74 Nu = Pro-42. K A (HS,0t /D, r)- G(Hslot /D, f)- F(Re) [ 4.33 ] Sh = Sc0‘42 -K A (Hslot /D, f)- G(Hslot /D, f)- F(Re) [ 4.34 ] KA(Hs[m/D, f) is the array correction function and is given by Equation [4.35]. 6 —0.05 H /1) K H D,f=1+—S'°—t— [4.35] A( slot/ ) [0.6/J? where: f=0.785-(D/L)2 [ 4.36] In the above equations, D and L represent the nozzle diameter and nozzle spacing, respectively. G(f, H/D) is a geometric function (Equation [4.37]), and F(Re) is a function of the Reynolds number (Equation [4.38]). 1—2.2-~/f [4.37] G(f’HSIOt/D)= ZW/F. I+0.2'(Hslot/D-6)"/E 75 F(Re): 0.5-Re2/3 [ 4.38] From the values of Nusselt and Sherwood numbers, values for heat and mass transfer coefficients were calculated using Equations [4.26] and [4.27] respectively. 4.3 Microbial inactivation model A simple first-order inactivation model (Equation [4.39]) was combined with the heat and mass transfer model. N0 AI 10 — : — 4.39 810( N ] D l l where: D = Dmf .10(Tref‘T)/Z [ 4.401 The reduction in the number of Salmonella at each node was determined by calculating the fraction of survivors after each time step. The number of survivors after each time step was calculated by inserting the temperature data generated by the cooking model into Equation [4.41]. The overall reduction at the center of the patty was then 76 determined for the entire cooking process. The aggregate microbial reduction for the entire patty was calculated by calculating the volume average of all the nodes. At N=N0/IOD [4.41] 4.4 Finite element formulation 4. 4.1 Introduction The finite element method (FEM) was utilized to solve the cooking model equations. The finite element is a common numerical method for solving differential equations that are difficult to solve using analytical techniques. The finite element method has several advantages over other numerical techniques that make it very suitable for formulating cooking models. FEM is readily adaptable to irregular geometries, it can be applied to systems containing more that one set of physical properties, and it is readily applied using computers. The following sections briefly describe the finite element method and the techniques used to convert the model equations into computerized form. A much more detailed description of the finite element method was given by Segerlind (1984). 4. 4.2 Finite element basics To apply the finite element method, the geometry of the system must first be broken down into a finite number of discrete regions referred to as elements. Elements are typically triangular or quadrilateral, although other geometries can also be utilized. 77 The comers of each element are referred to as nodes. Adjacent elements share the nodes at the corners they have in common. When the finite element method is performed, a numerical solution is generated for each node in the element mesh. Increasing or decreasing the number of nodes controls the resolution of the model. Models with a large number of nodes produce finer resolution at the expense of increased computing time. Figure 4.3 shows the geometry of the system used for the computer model as well as the element mesh. Figure 4.4 is a view of the element mesh with the coordinate axis indicated. Figure 4.3 - Geometry of a ground meat patty with the modeled region and element mesh indicated. 7.1.5474: '51 '51 P.‘ P." '4 PA PA PA '5. IL: '3‘ '1'.‘ P.- Df.‘ IF.- '4 'A umaaaauauaaaaaaaaanu: mmmmmmmmmm: Figure 4.4 - Illustration of the finite element mesh utilized for the model. The radial symmetry of ground meat and poultry patties allowed patties to be modeled as two-dimensional axisymmetric bodies. For the purpose of the cooking model, the modeled region was broken down into 117 triangular elements. This resulted in 82 nodes. The density of the element grid was higher at the patty surface to better 78 model the larger moisture, fat, and temperature gradients present near the product surface. The 117-node element mesh was chosen to balance solution accuracy with computing time (see Section 5 .5.1). 4. 4.3 Governing equations As discussed in the previous sections, the cooking model was formulated using three sets of differential equations. The basic solution of the heat transfer, moisture transfer, and fat transfer equations using the finite element method is very similar. Therefore, solution of a generic transport equation will be shown with emphasis given to the differences between the three solutions. The basis for each transport solution was the time-dependent two-dimensional field equation expressed in radial coordinates (Equation [4.42]). 2 i§$=1[nri[r@)]+D,9—Il [ 4.42] at r or ar 522 Equation [4.41] is a generic form of the governing equations given by Equations [4.1], [4.7], and [4.16]. In Equation [4.42], the term (of) is the unknown variable and corresponds to the enthalpy, moisture content, and fat-protein ratio in Equations [4.1], [4.7], and [4.16], respectively. The terms (A) and (D) are combinations of thermo- physical property values. For the heat transfer solution, the term D is equal to the thermal conductivity, and the term 7). is equal to the product of the heat capacity and the density of the meat. For the moisture transfer solution, the term D is equal to the moisture conductivity, and the term it is equal to the product of the specific moisture holding 79 capacity and the density of the meat. For the fat transfer solution, the D term is equal to the capillary diffusivity, and the A term is not utilized. The finite element method is a weighted residual technique. An approximate solution is substituted into the governing equations, and the error term calculated. The product of this term and a weighting function is then reduced to zero to produce a numerical solution. The weighting function may take many forms. The method chosen was Galerkin’s method. In this method, the weighting function uses the same functions that were used for the approximate solution (Segerlind, 1984). Using Galerkin’s method for the axisymmetric field problem, the weighted residual equation is Equation [4.43]. {11(6)}: —][N]T[%§(r—;ij+ozg+x%]dv [ 4.43] The term [N], is a vector of shape functions. Triangular elements were used for the cooking model. This allowed the density of the element mesh to be increased near the boundaries without dramatic change in the aspect ratio of the elements. Proper aspect ratio is desirable, as elements with greatly uneven lengths and widths can contribute to loss of solution accuracy (Segerlind, 1984). Figure 4.5 is an illustration of a typical triangular element. The element contains three nodes that are labeled as nodes i, j, and k. 80 Figure 4.5 - Illustration of a triangular element showing counterclockwise node numbering. The shape functions for a triangular element are given by Equations [4.44] through [4.46]. l Nk = ‘2TA—(ak + bkr + ckz) where: ai = RjZk -Rij bi=Zj—Zk 81 Ci =Rk-R' [ 4.44] [ 4.45] [ 4.46 ] =RkZi—RiZk bj=Zk—Zi °j=Ri“Rk ak=RiZj—RJ-Zi bk =Zi—Zj Ck=Rj"Ri The terms R" and Z“ are the r and z coordinates of the nth node, respectively. After applying the product rule for differentiation and considerable manipulation (Segerlind, 1984), Equation [4.43] can be converted to the form of Equation [4.47]. or:[ineavfiaelvjal [[[N]T D, grim“) + 112 gsin 0de + {[[NF (A gijdv F [ 4.47] The first integral of Equation [4.47] is related to transport within the product and can be written as Equation [4.48], where [km] is the element stiffness matrix. ik”-—Ii VII) Qg—W a—[N]+ wz°[—:23%]dv [4.48] The second integral in Equation [4.47] is related to the derivative boundary condition and can be written as Equation [4.49], where “(0} is the inter-element vector. {1(6)}: [[N]T(D EcosfHDz gsin0]d [4.49] 82 The third integral in Equation [4.47] is a time-dependent capacitance term. Using Equations [4.48] and [4.49], Equation [4.47] can be re-written in the form: {1(0)}: {102)}, lk(°)]i¢(e)i+ [C(e)J[¢(e)} [ 4.50] where [Cm] is called the capacitance matrix and is given by Equation [4.51], and {(1)} represents the nodal values of 6CD/6t. [CM]: [3.[N]T[N]dv [ 4.51 ] Inserting the shape functions into the element stiffness matrix and integrating gives the form of the stiffness matrix shown in Equation [4.52]. ' 2 ‘ ’ 2 7 ) Zm-‘D bi bisz bibk Zm'D Ci CiCZj CiCk [1((e ]= 4A r bibj bj bjbk + 4A 2 CiCj Cj CjCk [ 4.52] 2 2 bibk bjbk bk CiCk CjCk Ck ‘ Expanding the capacitance matrix using a lumped solution gives the form of the capacitance matrix given by Equation [4.53]. The lumped solution assumes that the variation art/at is constant across each element. This is the solution that provides for the maximum operating range of the final model. 83 0 0 [C(e)]=-’f§0 1 0 [4.53] o 1 Treatment of the inter-element vector is somewhat more involved. A derivative boundary condition over a boundary F can be specified using Equation [4.54]. Drgc080+ngsm0=—M¢b +S [ 4.54] The right hand side of Equation [4.54] can be substituted into Equation [4.49] yielding Equation [4.55]. {19}: I[INIT(1\4[N]{“’)1- Sill“ I 4551 Upon expansion, Equation [4.55] produces two terms given by Equations [4.56] and [4.57]. IkM(c)I= lMlNlrlNldl" [4.56] r13C a): and [4,7, r13c 84 Equations [4.56] and [4.57] can be expanded into the form of Equation [4.58] and [4.59]. (3Ri+RJ-) (Ri+RJ-) 0 [kM(e)]=2”11‘2’IL (Ri+RJ-) (Ri+3Rj) 0 [4.58] 0 0 0 {fs(e)}= “6 Ri+2RJ~ [4-59] 0 For the convective heat transfer boundary condition, the term M is equal to the heat transfer coefficient h. The term S is equal to the heat transfer coefficient multiplied by the temperature of the bulk fluid. Since the heat transfer solution was formulated in terms of enthalpy, the equilibrium enthalpy related to the oven air temperature was calculated using Equation [4.2] and substituted in place of the oven air temperature in calculating S. For the convective moisture transfer boundary condition, the term M is equal to the convective mass transfer coefficient hm. The S term is equal to the convective mass transfer coefficient multiplied by the equilibrium moisture content for the meat under the given oven air temperature and moisture conditions. Equilibrium moisture content was calculated using Equation [4.60] derived from the work of Huang and Mittal (1995). 85 —l/l.0983 EM C = 1.9818 Tsurface ln(RH) [ 4.60 1 — 5222.47 where: RH =100 ' (Coven /Csat) I 4°61 I 4. 4.4 Finite diflerence time solution A finite difference time-solution was used to provide a FEM solution over a number of time steps. This resulted in an equation with a time-step term (Equation [462]). ([C]+ 0At[k]){}b = ([c]— (1 — 0)At[l<]){cr>}a + At((1 — 0){F}b + 0{F}a) [ 4.62 ] Setting the value of 0 equal to 1/2 (central difference method), results in Equation [4.63]. At At At [[c]+ 70.0415), = [[c]_ ?[K]]{}b = [P]{}a + {F *} [ 4.64] 86 where: [A] = [[c]+ 323M] [ 4.65 ] lpl=[lcl—é‘,llkl) [ 4.66] and {w}: 4‘0}. +114.) 1 4.67] 4.4.5 Application of FEM solution Equation [4.64] has a form that can be readily programmed into a personal computer. For the cooking model, three sets of element solutions were constructed. The first set of equations was related to the heat transfer portion of the model (Equation [4.1]). The second set of equations was related to mass transport due to moisture migration (Equation [4.7]). The third set of equations was related to fat transfer (Equation [4.16]). The finite element solution was programmed using Microsoft Visual Basic (Version 6.0 professional edition: Microsoft, Redmond, WA); (Section 9.5). 87 4. 4.6 User interface A Windows-based user interface for the model was developed using Microsoft Visual Basic. The user interface consisted of two primary window screens controlled by a menu toolbar. Illustrations of each interface screen are shown in Section 9.4. The first of the windows is the input screen. On this screen, the user inputs the oven air temperature, the oven steam content or wet bulb temperature, the air velocity, and the cooking time. The user also has the option of inputing a set of transient oven conditions using a file input. The input screen also requires the user to specify the initial temperature, fat content, and moisture content of the meat. Buttons are available for selecting between beef, pork, and turkey. The final component of the interface screen is the microbial inactivation input section. Default D, 2, and reference temperature values for a seven-strain Salmonella cocktail in beef (Smith et al., 2001) are incorporated into the model. However, these defaults may be replaced with values specified by the user. The second screen of the user interface is the output screen. On this screen, the output temperature, moisture, yield, and microbial inactivation are displayed in graphical form. The final temperature, moisture, yield, and lethality values are also given in digital form. A message is displayed in the inactivation output section indicating whether or not the required level of lethality was achieved for the product selected. 88 5 RESULTS AND DISCUSSION 5.1 Overview The experimental results for this study are grouped into three main sections. The first section (Section 5.2) describes the results of a set of experiments conducted using a laboratory convection oven. Patties of ground turkey, ground beef, and ground pork were cooked to various endpoint temperatures to determine the effects of meat species and initial fat content on cooking yield, heating rate, and shrinkage. These experiments were intended to provide insight on differences in cooking behavior between meat species and fat content and to evaluate the need for including fat transfer in meat cooking models. Section 5.3 describes the results of a series of experiments conducted to determine the fat holding capacity of ground beef as affected by isothermal heating. These data were used to develop an expression for fat holding capacity as a function of initial fat content and temperature. This expression was then included as a portion of the cooking model (Section 4.2.3). Section 5.4 describes the results of experiments conducted using an industrial moist-air impingement cooking system (Stein model JSO-IV). Ground beef patties were cooked in the oven using various cooking conditions. The effects of oven conditions on cooking time, yield, and volume change were determined. Additionally, data from this set of experiments were used to validate the heat and mass transfer components of the cooking model. 89 l 5.2 Laboratory oven cooking tests In these tests, small meat patties were cooked in a laboratory convection oven to determine the effect of species and initial fat content on fat loss for ground turkey, ground beef, and ground pork (Section 3.2). The effects of species and fat content on cooking time, yield, and volume change were also determined. 5. 2.] Cooking time For each cooking test, the time required to reach the target endpoint center temperature was measured. Using this information, temperature versus time plots were constructed for each species and fat content (Figures 5 .1-5.3) 100, 904} 6 E O a b 805 E i 0 :1 g 70E O. i 0 Cl 5 60, t— : U. Q) . ‘a 50-: -_ 5 E a 01.4%fat 40; 08.6%fat 30PJJJLLFALLJ#JIII%IL11%L141 3.00 5.00 7.00 9.00 1 1.00 13.00 Time (minutes) Figure 5.1 — Center temperature as a function of cooking time for ground turkey patties cooked in a laboratory convection oven: means of 5 replicates. 90 100 . A 90 a D 0 e 80 “F D E 70 -:— 8. : 1:1 9 8 60 «e t : 1: o ’a’ 50 E __ 8 E 0 07.2% fat 40 l p l7._5_j/._ fat 30“ =W‘ 3.00 5.00 7.00 9.00 11.00 13.00 Figure 5.2 — Center temperature as a function of cooking time for ground beef Time (minutes) patties cooked in a laboratory convection oven: means of 5 replicates. 100 _ A 90 a g :1 o o 80 e: a 2 n . £3 70 . 8. E n o E, 601: a E 1:1 0 3;; 50 t 8 L , 015.7% fat 40 ":8 041.9% fat 30_ igmlrilrIllr larlilk 3.00 5.00 7.00 9.00 11.00 13.00 Time (minutes) Figure 5.3 — Center temperature as a function of cooking time for ground pork patties cooked in a laboratory convection oven: means of 5 replicates. 91 The cooking rates of turkey differed between the 1.4% and 8.6% fat samples (Figure 5.1). The temperature of both the low and high fat patties increased as a function of time. However, the heating rate of the 1.4% fat samples was slightly higher. The 1.4% fat patties took 254 seconds to reach 45°C compared to 264 seconds for the 8.6% fat patties. The difference in heating time increased as the patties reached higher temperatures. The 1.4% fat patties reached 85°C in 513 seconds compared to 684 seconds for the 8.6% fat patties, a difference of almost 3 minutes. Analysis of variance confirmed the statistical significance (P<0.05) of differences in heating times between the two fat contents (Table 5.1). Table 5.1 — Analysis of variance for cooking time of ground turkey patties as affected by temperature and initial fat content. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 6618213 4 1654553 48.2258 <0.001 Fat Content 59305.68 1 59305.68 17.2860 <0.001 Interaction 57447.52 4 14361.88 4.1861 0.006 With ground beef, the heating rate was significantly higher for the 17.5% fat patties than for the 7.2% fat patties (Figure 5.2). The average time required to heat the 17.5% fat patties to 45°C was 260 seconds compared to 361 seconds for the 7.2% fat patties. The heating time increased with increasing patty temperature. The time required to reach 85°C was 619 seconds for the 17.5% fat patties compared to 836 seconds for the 7.2% fat patties. Analysis of variance showed both center point temperature and fat content to affect (P<0.05) heating time for ground beef patties (Table 5.2). 92 Table 5.2 - Analysis of variance for cooking time of ground beef patties as affected by temperature and initial fat content. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 7420335 4 185508.4 19.7617 <0.001 Fat Content 1208353 1 1208353 12.8723 <0.001 Interaction 29187.52 4 7296.88 0.7773 0.547 As with the beef patties, the higher fat content pork patties took less time to heat than the lower fat patties (Figure 5.3). The average time required for the 41 .9% fat patties to reach 45°C was 236 seconds compared to 285 seconds for the 15.7% fat patties. Heating temperature increased with increasing time. The time required for the 41 .9% fat patties for each 85°C was 608 seconds compared to 710 seconds for the 15.7% fat patties. Analysis of variance showed temperature and fat content to be significant (P<0.05) with respect to heating time (Table 5.3). Table 5.3 - Analysis of variance for cooking time of ground pork patties as affected by temperature and initial fat content. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 1036732 4 2591829 59.6434 <0.001 Fat Content 59512.5 1 59512.5 13.6951 <0.001 Interaction 7689.6 4 1922.4 0.4424 0.777 In the case of the ground beef and ground pork patties, samples containing higher levels of fat achieved higher rates of heating. This was most likely due to the lower moisture content associated with the higher fat products. As water contributes the 93 greatest source of heat capacity in the meat matrix, lowering the water content would be expected to increase the heating rate. It has also been hypothesized that inclusion of high fat contents could result in convective heating within the meat, thereby increasing the heating rate (Shilton et al., 2002). Unlike the ground pork and ground beef patties, the heating rate of the ground turkey patties was higher for the lower fat product. It should be noted that the compositions of the two ground turkey samples were more similar than those of the ground pork and ground beef samples. Specifically, the moisture content of the 8.6% fat ground turkey was only 1.8 w.b. moisture points lower than the moisture content of the 1.4% fat patties. This compares with differences in moisture content of 8.2 and 20.5 w.b. moisture points for the ground beef and ground pork, respectively. The smaller differences in moisture content for the ground turkey mean that the heat capacity of the two samples was nearly the same. Thus, moisture probably played a less significant role in differences in heating rate for the ground turkey patties. The difference in heating rate between different fat levels of ground turkey patties was largest at temperatures above 55°C. Interaction between fat content and temperature was significant. This differs from ground beef and ground pork, which had consistent differences (i.e., no significant fat-temperature interaction). It is therefore likely that the different heating characteristics of the turkey were related to other compositional characteristics of the meat. The 8% initial fat ground turkey patties included ground skin as a method for increasing fat content. This may have contributed to the different heating characteristics of the ground turkey as compared to the ground beef and ground pork. 94 5. 2.2 Cooking yield For each test, the cooking yield was calculated at each endpoint temperature (Figures 5.4-5.6). Large differences in yield were evident between the 1.5% fat and 8.6% fat ground turkey patties (Figure 5.4). The yields of the 1.4% fat patties were consistently higher than the yields of the higher fat patties. The yield of the 1.4% fat patties ranged from 90% at 45°C down to 81% at 85°C compared to yields of about 83% to 63% for the 8.6% fat patties. The decrease in yield was approximately linear with respect to center temperature for both fat contents. Analysis of variance confirmed the significance (P<0.05) of both center temperature and fat content on cooking yield (Table 5.4). Higher yield losses for the 8.6% fat patties were presumably due to fat loss during cooking. However, the results presented in the next section do not support the conclusion that fat transport was solely responsible for yield differences between the two fat contents of turkey. It is possible that interaction between the fat and water within the patty resulted in lower water binding capacity for the higher fat patties, resulting in higher rates of moisture loss during cooking. Further experimentation is needed to determine the mechanisms for fat and water binding during cooking. Table 5.4 — Analysis of variance of yield as a function of center temperature and initial fat content for ground turkey patties. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 1359.399 4 339.8498 10.5351 <0.001 Fat Content 1751.823 1 1751.823 54.3051 <0.001 Interaction 203.2676 4 50.8169 1.5753 0.200 95 100.0 95.0 o 1.4% fat 90-0 ° , Efiéf’éfflj % 233 1:; ° . . . >~ ’ E n E 75.0 “E 8 E '3 5 70.0 "E °‘ 650-; D a 60.0 55.0 50.0:5.......r,...,,....,.... 40 50 60 70 80 90 Center temperature (° C) Figure 5.4 — Yield as a function of endpoint center temperature for ground turkey patties of two fat contents: means of 5 replicates. For the ground beef, only minor differences in yield were detectable between the 7.2% and 17.5% fat samples at each cooking temperature (Figure 5.5). The yield of both products decreased from about 84% at 45°C to between 63 and 66% at 85°C. Analysis of variance indicated that yield was related (P<0.05) to center temperature but not to initial fat content for these data.. Table 5.5 - Analysis of variance of yield as a function of center temperature and initial fat content for ground beef patties. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 2135.567 4 533.8917 37.9595 <0.001 Fat Content 9.6439 1 9.6439 0.6857 0.413 Interaction 21.9899 4 5.4975 0.3909 0.814 96 100.0 : 95-0 '0 7.2% fat 90.0 i n 17.5% fat 85.0 e r ”“m 80.0 4; g 75.0 -E III 70.0 { 65.0 —E 60.0w 55.0— 50.0“wielxlxwllwwlwin» 40 50 60 70 80 90 Center temperature (° C) D Percent yield OCI ‘r—T—TT TTT'TTlTT Figure 5.5 - Yield as a function of endpoint center temperature for ground beef patties of two fat contents: means of 5 replicates. For ground pork, there were large differences in yield between the 15.7 and 41.9% fat patties (Figure 5.6). The yield of the lower fat patties was much higher at each center temperature. Cooking yield for the 15.7% fat patties ranged from about 90% at 45°C to about 71% at 85°C. Cooking yield for the 41.9% fat patties ranged from about 74% at 45°C to about 54% at 85°C. Cooking yield decreased linearly as a function of temperature for patties of both fat contents. Analysis of variance showed differences (P<0.05) in yield as a function of temperature, fat content, and temperature-fat interaction (Table 5.6). 97 100.0 : 95.0 90.0 o 85.0 , , 80.0 , 75.0 D 70.0 0 65.0 «. 60.0 D 0 55.0 50.0:“11wil 40 50 60 70 80 90 :isfifit La 4.13%!“ Percent yield f T777 0 D Center temperature (°C) Figure 5.6 - Yield as a function of endpoint center temperature for ground pork patties of two fat contents: means of 5 replicates. Table 5.6 - Analysis of variance of yield as a function of center temperature and initial fat content for ground pork patties. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 2132.37 4 533.0926 96.4285 <0.001 Fat Content 4153.887 1 4153.887 751.376 <0.001 Interaction 79.8724 4 19.9681 3.61 19 0.013 The differences in cooking yield appear to be affected by both meat species and fat content. No significant correlation between initial fat content and cooking yield can be made across the three meat species. Although the yield at 85°C was lowest for the highest fat product (41 .9% fat ground pork) and highest for the lowest fat product (1.4% fat ground turkey), the behavior of samples of in-between fat contents was not consistent. This could be caused by differences in protein conformation and fat composition of the 98 three meat species. It may also be due to differences in water and fat binding capacity of the meat proteins. Differences in particle size and porosity of the meat may also affect the cooking yields by changing the moisture and fat transport dynamics. 5.2.3 Fat loss The amount of fat lost during cooking as a function of temperature for 1.4% and 8.6% fat ground turkey patties is shown in Figure 5.7. The amount of fat lost during cooking of ground turkey patties was below 2% for every temperature. Analysis of variance indicated that neither temperature nor initial fat content had a significant (P<0.05) effect on fat loss during cooking (Table 5.7). 30.0 , 25.0 i : 1.4% fat 0 8.6% fat 20.0 {- 15.0 10.0 7 Fat loss (% of initial mass) .U' o l tTfi—r, r ,. inL—HIHQIJTIJIGII 4O 50 60 70 80 90 .9 o I Center temperature (° C) Figure 5.7 - Fat loss as a function of temperature for ground turkey patties of two fat contents: means of 5 replicates. 99 Table 5.7 — Analysis of variance for fat loss as functions of center temperature and initial fat content for ground turkey patties. Factor Sum of Degrees of Mean F-Value P-value Squares Freedom Square Temperature 29.9082 4 7.4770 1.2317 0.3184 Fat Content 4.4131 1 4.4131 0.7270 0.4006 Interaction 17.8231 4 4.4558 0.7340 0.5760 The amount of fat lost during cooking of ground beef patties was much larger than the amounts lost during cooking of ground turkey (Figure 5.8). Fat losses were higher for the 17.5% fat patties, ranging up to 6% of the initial mass of the meat. Analysis of variance showed that initial fat content had a significant (P<0.05) effect on the amount of fat lost during cooking (Table 5.8). A significant temperature effect on the amount of fat lost during cooking could not be shown. 30.0 , A : 0 g 25.0 07.2/6 fat 8 E D 17.5% fat :9: 20.0 E E. : «5 115.0 :— °\° E 1,; 10.0 -- .2 1 L3 5.0 .. a D 00 l’lwm‘: ’:1°1% 40 50 60 7O 80 90 Center temperature (°C) Figure 5.8 - Fat loss as a function of temperature for ground beef patties of two fat contents: means of 5 replicates. 100 Table 5.8 — Analysis of variance for fat loss as functions of center temperature and initial fat content for ground beef patties. Factor Sum of Degrees of Mean F-Value P-value Squares Freedom Square Temperature 12.8007 4 3.2002 0.9060 0.4730 F at Content 240.3399 1 240.3399 68.0423 <0.001 Interaction 35.0635 4 8.7659 2.4817 0.065 Fat losses for ground pork were the highest of the three meat species tested (Figure 5.9). For the 15.7% fat pork patties, the amount of fat lost during cooking was low and never exceeded 2% of the initial patty mass. However, for the 41 .9% fat patties, the amount of fat lost during cooking was much larger. The amount of fat lost during cooking was about 20.5% of the initial patty mass at 45°C. The amount of fat lost increased with increasing patty temperature up to about 28% at 85°C. Analysis of variance showed cooking temperature, fat content, and temperature-fat interaction to affect (P<0.05) fat loss (Table 5.9). Table 5.9 — Analysis of variance for fat loss as functions of center temperature and initial fat content for ground pork patties. Factor Sum of Degrees of Mean F-Value P-value Squares Freedom Square Temperature 68.6303 4 17.1576 14.3904 <0.001 Fat Content 5660.502 1 5660.502 4747.583 <0.001 Interaction 56.5569 4 14.1392 1 1.8589 <0.001 101 30.0 _ a 7,; f D D 3 25.0 _ D E 3‘ 20.0 D E E o 15.7% fat «5 15.0 ; c ~ 041.9% fat 8) .- __.2_.___ a 10.0 9; 2 . *-' i 63 5.0 o 00 H‘HTHLtMT ° W494 40 50 60 70 80 90 Center temperature (° C) Figure 5.9 - Fat loss as a function of temperature for ground pork patties of two fat contents: means of 5 replicates. The quantities of fat lost by the 17.5% fat ground beef patties and 41 .9% fat ground pork patties during cooking are of great significance for modeling. Cooking models that do not take fat loss into account will likely over predict cooking yields. For high fat products such a sausage, these errors in yield prediction could exceed 25%. Many models may account for fat losses indirectly by over predicting mass transfer coefficients. However, this method may restrict the utility of the models to products of similar composition. To produce robust cooking models, transport of both moisture and fat components should be considered. 102 5.2.4 Volume change Plots of volume change as a function of product temperature were plotted for each meat species (Figures 5. 10-5. 12). The volume of ground turkey patties as a function of temperature is shown in Figure 5.10. Volume of the 1.4% fat patties ranged between 74 and 70% of the initial volume at center temperatures between 45 and 85°C. Volume of the 8.6% fat patties decreased more dramatically as a fiinction of center temperature and ranged from 62% to 56% of the original volume at center temperatures of 45 and 85°C respectively. Analysis of variance showed that initial fat content had a significant effect on volume change during cooking (P<0.05) (Table 5.10). However, the relationship between center temperature and volume change was not statistically significant. Although the change in volume was significantly different over the measured temperature range, both products exhibited major decreases in volume between the raw state and 45°C. This shows that major volume changes occur during the initial stages of cooking, and thus may be impossible to avoid during convection cooking. Table 5.10 — Analysis of variance for effects of temperature and initial fat content on volume change for ground turkey patties cooked in a laboratory convection oven. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 253.4806 4 63.3702 1.0724 0.383 Fat Content 2910.062 1 2910.062 49.2471 <0.001 Interaction 122.8181 4 30.7045 0.5196 0.722 103 100.0 5 o E 90,0 —: o 1.4/o fat '3 r Layman 3. 80.0 ‘5 a f" .E : o . E0 70.0 "' . . Q 0 C “-0 - g 60.0 {- D n 8 : U U D E 50.0 E 40.0 lAA‘iJ"‘+“‘#Tlltlt:thih 4O 50 6O 7O 80 90 Center temperature (° C) Figure 5.10 — Relationship between product temperature and volume change for ground turkey patties cooked in a laboratory convection oven: means of 5 replicates. The ground beef patties exhibited a similar decrease in volume during cooking (Figure 5.11). Analysis of variance showed both center temperature and fat content to significantly affect volume (Table 5.11). Unlike the ground turkey and pork patties, the higher fat beef patties had less shrinkage during cooking than did the lower fat patties. Table 5.11 — Analysis of variance for effects of temperature and initial fat content on volume change for ground beef patties cooked in a laboratory convection oven. Factor Sum of Degrees of Mean F-Value P-Value Squares Freedom Square Temperature 1424.392 4 356.0981 11.5257 <0.001 Fat Content 631.248 1 63 l .248 20.43 14 <0.001 Interaction 133.973 4 33 .4933 1.0847 0.377 104 100.0 , : 0 o 90.0 ; 07.2/0 fat E : arts/oat g 80.0 ~; "g t '30 70.0 f El ’5 t , a .8 . o 3 60.0 . 0 CI 5 . 8 50.0 1 ° . o a. - 40.0F**‘*#r*4%;illi+gain:.rr. 40 50 60 70 80 90 Center temperature (° C) Figure 5.11 — Relationship between product temperature and volume change for ground beef patties cooked in a laboratory convection oven: means of 5 replicates. The largest changes in volume during cooking occurred for ground pork patties (Figure 5.12). Volume decreased as a function of cooking temperature for both the 15.7% and 41 .9% fat patties. The 15.7% fat patties ranged from about 74% to 60% of their original volumes at 45°C and 85°C, respectively. The 41.9% fat patties ranged from 70% to 42% of their original volumes at 45°C and 85°C, respectively. The large volume changes correlate with high yield losses during cooking. Analysis of variance indicated that both center temperature and fat content significantly (P<0.05) affected volume change (Table 5.12). 105 Table 5.12 - Analysis of variance for effects of temperature and initial fat content on volume change for ground pork patties cooked in a laboratory convection oven. Factor Sum of Degrees of Mean F-Value P-Value Smres Freedom Sjuare Temperature 2125.544 4 531.386 16.7989 <0.001 Fat Content 1506.957 1 1506.957 47.6400 <0.001 Interaction 451.4145 4 112.8536 3.5677 0.014 100.0 . t GS) 900 - O 15.7% fat % : 041.9% fat 3. 80.0 2 E . é" 70.0 -} 6 .2 : o ’ ° 600 n o E ' : D “2’ : 9 a: 50.0 ‘ 40.0'1~w~~w~x+~wwpll 40 50 60 70 80 90 Center temperature (°C) Figure 5.12 — Relationship between product temperature and volume change for ground pork patties cooked in a laboratory convection oven: means of 5 replicates. Changes in volume are a major concern in the development of cooking models. Slight changes in thickness greatly affect predictions of center temperature and may also have large effects on yield predictions. Due to the geometry of the products tested, most heat and mass transfer occurs in the axial direction. Therefore, changes in product thickness are of greater concern than changes in product radius. 106 Although the volume of all samples decreased with increasing product temperature, the same was not true for patty thickness. Most of the reduction in patty volume was caused by a reduction in patty diameter. The thicknesses of the 1.4 and 8.6% fat ground turkey patties, when cooked to a temperature of 85°C, were 120 and 110% of the initial values, respectively. The final thicknesses of the 7.2 and 17.5% fat ground beef patties were 97 and 120% of their initial values, respectively. The thicknesses of the 15.7 and 41.9% fat ground pork patties were 120 and 110% of their initial values, respectively. Therefore, in all but one case, patty thickness actually increased, while the radius decreased. These results indicate that changes in volume cannot be modeled by simply reducing the size of the element mesh in proportion to the reduction in product mass. Further information is needed to determine the mechanisms for patty shrinkage. This would be an excellent area for further study. 5.3 Fat holding capacity experiments Fat holding capacity (FHC) was determined as a function of temperature and initial fat content for two lots of ground beef with initial fat contents of 5.6 and 15% by mass (Section 3.3). The fat holding capacity of the meat was defined as the amount of fat remaining in the meat after heating and centrifuging for 15 minutes at 1000 g. The fat content of the centrifuged meat was expressed in terms of dry basis fat content using Equation [5. 1]. 107 F: Mfat /Mdry matter [ 5-1] where M dry mane, refers to the mass of meat that is neither water nor fat. The mass of dry matter in each sample was calculated using Equation [5.2]. Mdry matter : I\dtotal _ (M fat + Mwater) [ 5-2 ] Mm and Mwatc, were calculated from the initial wet basis moisture and fat contents of the meat. The amount of dry matter was assumed to be constant during heating and centrifugation. During cooking, small quantities of dry matter are released in the forms of soluble proteins. Future studies should be conducted to describe the relationships between meat proteins and moisture and fat holding capacities in a more fundamental manner. However, inclusion of such relationships was beyond the scope of this study. The fat content, P, was plotted as a function of heating temperature for each heating time (Figures 5.13-5.16, 5.18-5.21). 5.3.] Low fat samples (5. 6% initial fat wet basis) The relationship between fat holding capacity and temperature for heating times of 2, 5, 10, and 15 minutes can be seen in Figures 5.13-5.16. The initial fat content of the ground beef was 0.14 g fat/g dry matter. For each heating time, the fat holding capacity was markedly lower at 55°C. The fat holding capacity of samples heated to 90°C ranged from 0.03 to 0.05 g fat/g dry matter. 108 Linear regression was used to fit a model for fat/dry matter ratio as a function of heating temperature and holding time for the 5.6% fat (wet basis) ground beef samples. A quadratic response surface was chosen for the initial regression model (Equation [5.3]). F=BO+BI~T+Bz~T2+B3-t+B4-t2+BS-t-T+e [5.3] Regression coefficients for the response surface model are given in Table 5.13. .o .o \l 00 + l I 1 9’ o\ / g dry matter) 53 .o A u. H T £3(l3 T 3 a [4., 0.2 i L O O O Q Q 0 1 + o . - » o : e o o , . 0 _ _.L_L_JHL I 1.,1__.i_ l__+_1__.t__l _r_+_l_i__4__t_l,4__i__i__i_+_i_;t_i_+.1.4_t_t_+ 1 .L_J._L_T 20 30 40 50 60 70 80 90 100 Temperature (° C) Figure 5.13 - Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 2 minutes at temperatures from 30 to 90°C: means of 5 replicates. 109 0.6 0.4 (g fat / g dry matter) Temperature (° C) Figure 5.14 - Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 5 minutes at temperatures from 30 to 90°C: means of 5 replicates. 0.8 I *4 - _ - ; 0.7 I 75 0.6 7t E I E 0.5 | w 0.4 it» a 0.3 i 39 u. 02 if 0 o o o 0.1 j . O . Q 0 O Q . O _1_L_.1_L_+,_,1 1.0411 1 l 1 1 L1. 1 1% 1 1 1 l +_1___L_1._L_+_J.__1_L_J,+_JW1__L_1 20 3O 40 50 60 70 8O 90 100 Temperature (° C) Figure 5.15 — Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 10 minutes at temperatures from 30 to 90°C: means of 5 replicates. 110 9 DJ 1* F (g fat / g dry matter) .C N ”—F—fi' 00...... O ,l 1 Ln 1 T ; . 1 I 1.,h_i 1 191,4 L._1__l_+_ L..L_L_..l.. 1.4.4 1.,1._+.1,n__1__1 +44 1 ,LJ 20 30 40 50 60 70 80 90 l 00 Temperature (° C) Figure 5.14 — Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 5 minutes at temperatures from 30 to 90°C: means of 5 replicates. 0.8 g 0.7 I I» .39 MON t / g dry matter) 9 .5 r——'—r—'—rT-'Tt r—r—r-r .53 0.3 39 a. 0.2 .r 9 O O . 0'1 - o 9 . o o O o o O 14; 1.1 1 l 1-1-J,._*_LL_1_L_+_L_L_L..L+_J__J_J.J_+..LNL_1_4+L.J_A_J_+__1_L.L..L_1 20 30 40 50 60 70 80 90 100 Temperature (° C) Figure 5.15 - Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 10 minutes at temperatures from 30 to 90°C: means of 5 replicates. 110 .9 00 I I I I I I I I I .0 \l ._— WI ._T.. .._1_ v : Tno- on? T - - ‘7 v v v ‘ ‘-T—Tw‘d. T . T «-< I F (g fat / g dry matter) O .0 .O .O .0 .O N w A u. as O 5 O O O O O I I I I I I I N O b) O A O LI! 0 O\ O \l O 00 O \O O p—n O 0 Temperature (° C) Figure 5.16 — Fat holding capacity as a function of temperature for 5.6% fat ground beef heated for 15 minutes at temperatures from 30 to 90°C: means of 5 replicates. Table 5.13 - Results from regression of fat holding capacity as functions of time and holding temperature for 5.6% fat ground beef. (F=[3o +3, 'T+13¢ -T’ +0, ~t+04 -t’ +13, -t-T+e) Degrees of Sum of Mean F-Value P-Value Freedom Squares Square Regression 5 0.0703 0.0141 37.8124 <0.001 Residual 46 0.0171 0.0004 Total 51 0.0875 Factor Coefficient Standard t-Statistic P-Value Error [30 (g fat/g dry matter) 0.3459 0.0339 10.2141 <0.001 B1 (g fat/g dry matter)-°C'1 -0.0060 0.0011 -5.6101 <0.001 02 (g fat/g dry matter)-°C'2 3.28E-05 8.67E-06 3.7842 <0.001 83 (g fat/g dry matter)'s" -00057 0.0032 -1.8057 0.078 [34 (g fat/g dry matter)-s'2 0.0002 0.0002 1.3904 0.171 05 (gfat/gdrymatter)°°C’l-s'l 2.96E-05 2.86E-05 1.0368 0.305 111 Heating temperature was found to affect the fat holding capacity of the meat (P<0.05). The heating time and time-temperature interaction terms of the response surface model were not significant. A second regression was performed, neglecting the time and time-temperature interaction terms (Equation [5.4]). F=BO+BI.T+8,.T2+2 [5.4] The regression coefficients of the modified response surface model are given in Table 5.14. Table 5.14 — Results from linear regression of fat holding capacity versus holding temperature for 5.6% fat ground beef. (F=[i,,+[ll -T+[12-'l‘z +8) Degrees of Sum of Mean F-Value P-Value Freedom Squares Square Regression 2 0.0691 0.0346 92.2654 2.43E-l7 Residual 49 0.0184 0.0004 Total 51 0.0875 Factor Coefficient Standard t-Statistic P-value Error I30 (g fat/gdrymatter) 0.3178 0.0295 10.7649 <0.001 B: (g fat/gdrymatter)-°C" -0.0057 0.0011 -5.4687 <0.001 [32 (gfat/gdrymatter)-°C'2 3.24E-05 8.69E—06 3.7254 <0.001 The regression model shows the fat holding capacity of the meat to be inversely proportional to the heating temperature. The drop off in fat holding capacity between 50 and 55°C is likely related to melting of fat globules within the meat matrix. The melting 112 temperatures of the most prevalent saturated fatty acids in ground beef occur near this range, with the melting temperatures of myristic, palmitic, and stearic acids being 54, 63, and 70°C, respectively (Bodwell and McClain, 1978). Ground beef with a fat content of only 5.6% contains only small quantities of extra-muscular fat and thus should be expected to contain lower percentages of saturated fatty acids. This could explain the lack of further losses in F HC at temperatures above 60°C. The lack of time-significance seems to support the idea that fat holding capacity is largely related to the physical state of the fat itself. Since fat melting occurs over a short period, once the fat has melted, further heating should not affect the amount of liquid fat. If fat holding capacity changes were driven primarily by changes in protein conformation, denaturing, etc., longer heating times would be expected to produce increasingly lower levels of fat holding capacity. A comparison of the fat holding capacity values predicted by the regression equation and the experimental data resulted in an R2 of 0.79 (Figure 5.17). The root mean square error (RMSE) of the regression was 0.02 g fat/g dry matter. 113 0.25 E 0.2 (U E C o S 0.15 -2 ° ~ 9 o g” ’ o o 90 d3 01 00 o .. . £9 ' ’00.. .0 g “ «A: 0. *5 0.05 -- '6 . g I an O I I I I 0 0.05 0.1 0.15 0.2 0.25 Experimental F (g fat/ g dry matter) Figure 5.17 — Comparison of fat holding capacity calculated from regression model versus experimental values: means of 5 replicates. 5.3.2 High fat samples (15% initial fat wet basis) The effect of temperature on fat holding capacity was much more pronounced for the 15% fat meat than for the 5.6% fat meat (Figures 5.18-5.21). The initial fat content of the 15% fat meat was 0.6 g fat/g dry matter. Fat holding capacity decreased with temperature and ranged between 0.07 and 0.17 g fat/g dry matter for samples heated to 90°C. Multiple linear regression was again used to model fat content as functions of heating temperature and holding time, using a quadratic response surface equation (Equation [5.3]). The regression coefficients for the response surface equation are given in Table 5.15. 114 .o .o \l 00 I—f‘r—r r-v I F (g fat / g dry matter) 95:99.09 Nw-hUIO .— rrI-Tr o o o o o 9 ° , o . o o 1___L._+___L. __ _1__+_ J.__1..__L___L__..+_ _._ 1___L 40 60 80 100 Temperature (°C) Figure 5.18 — Fat holding capacity as a function of temperature for 15% fat ground beef heated for 2 minutes at temperatures from 30 to 90°C: means of 5 replicates. F (g fat / g dry matter) O O O O O O O 0 o o 9 ° . Q 0 O . o 0 o .1 .+_ 4 Al___.L___.l_ _1_ i 1 1___L.__l 40 60 80 100 Temperature (0 C) Figure 5.19 - Fat holding capacity as a function of temperature for 15% fat ground beef heated for 5 minutes at temperatures from 30 to 90°C: means of 5 replicates. 115 I I I I I I I I I I I I I J .9 ox I c I Q) [r . E I- ’ I s I 3" 0.4 I . E 0.3 I 9 89 r ’ LL. 0.2 5* 9 . I . . . 0.1 I “ o o 0 HI. 1‘; 4 ' __.L 4 4+. .1.______L_ 1__1 +4- «1._ __L I 20 40 60 80 100 Temperature (° C) Figure 5.20 — Fat holding capacity as a function of temperature for 15% fat ground beef heated for 10 minutes at temperatures from 30 to 90°C: means of 5 replicates. 0.8 I . - - - hi -~— I 0.7 I I . I ’ . 99 L110 +44 F (g fat / g dry matter) iota):- p O O O O O O O O 0 fl 0 .f ____1__L_L__J-__ +._____1_.__1____1__-L_ + 1 1 I 1 1 1 1 ___I 20 40 60 80 l 00 Temperature (°C) Figure 5.21 - Fat holding capacity as a function of temperature for 15% fat ground beef heated for 15 minutes at temperatures from 30 to 90°C: means of 5 replicates. 116 Table 5.15 — Results from regression of fat holding capacity versus time and holding temperature for 15% fat ground beef. (F=Bo+B1'T+B2 'T2 +33 't+B4’t2 +35 -t-T+£) Degrees of Sum of Mean F-Value P-Value Freedom Squares Square Regression 5 1.4071 0.2814 54.8295 <0.001 Residual 47 0.2412 0.0051 Total 52 1.6483 Factor Coefficient Standard t-Statistic P-value Error [30 (g fat/g dry matter) 1.3497 0.1237 10.9131 <0.001 01 (g fat/g dry matter)-°C'l -0.0322 0.0039 -8.1558 <0.001 [3; (g fat/g dry matter)-°C‘2 0.0002 3.19E-05 6.5391 <0.001 [3. (g fat/g dry matter)°s" 0.0155 0.0117 1.3221 0.193 [34 (g fat/g dry matter)°s'2 -0.0006 0.0006 -1 .0612 0.294 135 (g fat/g dry matter)-°C"-s" -0.0001 0.0001 -1.0881 0.282 Heating temperature again affected fat holding capacity (P<0.05). However, as with the lower fat beef, holding time and time-temperature interaction were not statistically significant. A modified regression was performed omitting the terms for holding time and time-temperature interaction (Equation [5.5]). [5.5] The regression coefficients of the modified regression are shown in Table 5.16. 117 Table 5.16 — Results of regression of fat holding capacity versus holding temperature for 15% fat ground beef. (11:13o +13, -T+p, -T’ +2) Degrees of Sum of Mean F-Value P-Value Freedom Squares Sguare Regression 2 1.3922 0.6961 135.9025 <0.001 Residual 50 0.2561 0.0051 Total 52 1.6483 Factor Coefficient Standar t-Statistic P-Value d Error [30 (g fat/gdrymatter) 1.4163 0.1085 13.0474 <0.001 8. (g fat/gdrymatter)'°C" -00329 0.0039 -8.5302 <0.001 112 (gfat/gdrymatter)-°C'2 0.0002 3.191305 6.4820 <0.001 The fat holding capacity of the 15% fat ground beef was inversely proportional to the heating temperature. The effect of heating temperature was much more pronounced for the 15% fat ground beef than for the 5.6% fat product. The coefficient for the linear temperature term in the regression equation was —0.0329 (g fat/g dry matter)/°C for 15% fat ground beef, compared to —0.0057 (g fat/g dry matter)/°C for 5.6% fat ground beef. Like the low fat samples, the largest change in fat holding capacity for the 15% fat samples occurred between 40 and 55°C. However, unlike the lower fat samples, fat holding capacity of the 15% fat samples continued to decrease at temperatures above 55°C. This was likely due to the inclusion of higher quantities of long chain saturated fatty acids. The lack of time-dependence of the fat holding capacity was similar to that exhibited by the lower fat samples. This further supports the idea that fat holding 118 capacity, as measured in this study, is governed primarily by the physical state of the fat globules within the meat matrix. This is of special interest when formulating models for mass transfer, as the physical state of the fat must be considered. Clearly, solid fat does not have the transportability of liquid fat. However, it may be possible to develop highly advanced fat transport models based upon knowledge of the individual fat constituents of a given species and cut of meat. This would allow for maximum flexibility of advanced processing models. A comparison of the fat holding capacity values predicted by the regression model (Equation [5.5]) to the experimental data resulted in an R2 of 0.84 (Figure 5.22). The RMSE of the regression model was 0.071 g fat/g dry matter. l 33‘ £08 000.6 . E I 0 39 __ o «o 1.1.. 0.4 00 . E m e o O O O :5 0.2—— o g I“... ”r 0* II ““+I““I 0 0.2 0.4 0.6 0.8 1 Experimental F (g fat/ g dry matter) Figure 5.22 - Comparison of fat holding capacity calculated from regression model versus experimental values (15% fat): means of 5 replicates. 119 After completing separate regression analyses for the 5.6% and 15% fat ground beef samples, a combined regression was performed to create a model for fat holding capacity as a function of heating temperature, holding time, and initial fat content. A response surface model was chosen for the regression (Equation [5.6]). No second- order terms for fat content were utilized in the regression equation, because data were only available from two initial fat contents. The coefficients for the regression model are listed in Table 5.17. F=Bo+131°T+Bz°T24433444344205'F0+I36't'T+B7'T'FO 15.61 +Bg°t'F+8 120 Table 5.17 — Linear regression of fat holding capacity as functions of heating temperature, holding time, and initial fat content. F=B0+Bl “T7432 'T2 +53 't'I'B4 't2 +65 'Fo +56 't'T'I‘B-I °T'Fo +Bs-t-Fo+a Degrees of Sum of Mean F-Value P-Value Freedom Squares Square Regression 9 2.1005 0.2339 65.0322 <0.001 Residual 95 0.3409 0.0036 Total 104 2.4414 Factor Coefficient Standard t-Statistic P-value Error [30 (g fat/g dry matter) 0.3334 0.1104 3.0193 0.003 B! (g fat/g dry matter)°C'l -0.0136 0.0027 -5.037 <0.001 [3; (g fat/g dry matter)°C'2 0.0001 1.90E-05 6.3768 <0.001 [3, (g fat/g dry matter)°s" -0.0042 0.0111 -0.3804 0.704 [34 (g fat/g dry matter)-s'2 -0.0002 0.0003 -0.4541 0.651 [35 (none) 0.0503 0.0080 6.3074 <0.001 8. (g fat/g dry matter)-°C"°s" 0.0001 0.0002 0.7967 0.428 [37 (°C") -0.0005 0.0001 -4.2171 <0.001 Ba (8") 0.0008 0.0008 0.9458 0.347 None of the terms containing holding time were significant (P<0.05). A modified regression was performed excluding the time terms (Equation [5.7]). [ 5.7 ] F=Bo+131°T+I32'T2+I33°F0+I34'T°F0+8 The coefficients of the modified regression are shown in Table 5.18. 121 Table 5.18 - Linear regression of fat holding capacity as a function of holding temperature and fat content. (F=11.+9. -T+11. -T’ +11. ~F.+11. -T-F.+e) Degrees of Sum of Mean F-Value P-Value Freedom Squares Square Regression 4 1.9767 0.4942 106.330 <0.001 Residual 100 0.4647 0.0046 Total 104 2.4414 Factor Coefficient Standard t-Statistic P-value Error [30 (g fat/g dry matter) 0.7062 0.0746 9.4719 0.006 B. (g fat/g dry matter)-°C'l -0.0193 0.0026 -7.3996 <0.001 [3; (g fat/g dry matter)-°C'2 0.0001 2.16E-05 5.0990 <0.001 [3; (none) 0.0069 0.0020 3.4355 <0.001 [3.. (°C") 0.0002 3.41E-05 7.1495 <0.001 A comparison of the fat holding capacity values predicted by the regression and the experimental data produced an R2 of 0.81 (Figure 5 .23). The RMSE for the regression was 0.068 g fat/g dry matter. 122 .6 .0 \l 00 1 0.6 I lllTlI-T‘Irfi .9 LII 1 0.3 I Q 9 : O 0.2 I; 9 Predicted F (g fat/ g dry matter) O .b o O p—A 1 1 1 4 g 1 r I I 0.2 0.4 0.6 0.8 Experimental F (g fat/ g dry matter) 0 O'I Figure 5.23 - Comparison of fat holding capacity calculated from regression model versus experimental values: means of 5 repetitions. 5.3.3 Summary The regression models presented for fat holding capacity provide a tool that can be utilized as a component of the cooking model. However, it should be noted that this results in an empirical fat transfer model. The equations developed for fat holding capacity are specific for ground beef. Further studies should be conducted to measure the FHC of other meat species. In addition, mechanisms of fat transport should be firrther studied to produce a theoretical model for FHC. The methods used to determine fat holding capacity were considered adequate for the purposes of this study. However, several factors should be taken into consideration for further experiments. The most probable source of error for the FHC experiments was in the centrifugation procedure. The centrifuge utilized for the study was held at 25°C. Although the centrifuge tubes were insulated from the rotor by plastic inserts, some re- 123 solidification of fat took place during centrifugation. To eliminate this potential source of error, a method of separating the free fat while maintaining the sample temperature should be determined. It may also be desirable to measure the fat content of the meat directly using a method such as solvent extraction rather than the method of mass balances utilized in this study. 5.4 Industrial cooking tests A set of cooking experiments was conducted using an industrial moist-air impingement oven (Stein Model .1 SO-IV:F MC Foodtech, Sandusky, OH). Ground beef patties were cooked under different cooking conditions to quantify the effects of process conditions on heating rate and cooking yield. The results of this set of experiments were also used to validate the computer-cooking model. 5. 4.] Cooking time For each cooking experiment, temperature versus time was plotted for both the center and surface temperatures of the patty (Figure 5.24). The complete collection of temperature-time plots can be found in Chapter 9. The general form of each time- temperature plot was similar. After entering the oven, the surface temperature of the patty quickly rose to a semi-equilibrium level. Calculations showed that this level was equal to the wet bulb temperature of the meat surface (Section 4). The temperature at the meat surface was limited by a number of factors, including oven temperature, oven steam content, and the surface moisture content of the meat. 124 (a) 100 . 9O Surface 805% W 70 I 60 If. f1 Center Temperature (°C) KI! o 20 I 10 . 0 I ‘ ‘ . 1 ' . II I 1 0 50 100 150 200 Time (s) (b) 100 . 90 Surface \100 OO 1 ttr ox o it 00 1 TYTTWYI'TIIII Temperature (° C) W -b M Center 20 g 10 0 F 1 I 1 . . . I 0 50 100 150 200 Time (s) Figure 5.24 - Example surface and center temperature versus time for ground beef patties cooked at (a) oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s and (b) oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s. 125 The center temperature of the meat increased at a much slower rate than the surface and approached the surface temperature asymptotically. For most cooking experiments, the center temperature of the patty was at least 10°C below the surface temperature at the end cooking. Several revelations about the cooking process can be inferred from the experimental data. The first is that increasing the oven temperature does not proportionally increase the heating rate of the patties. The wet bulb temperature always limits the surface temperature of the patty. With the exception of extremely low surface moisture contents, the temperature of the patty surface cannot exceed the wet bulb temperature of the cooking air. The moisture content at which surface temperature will exceed the wet bulb temperature is a function of the equilibrium relative humidity of the meat. Under the conditions tested, this moisture content was between 5 and 7% wet basis. No observations of surface temperature exceeding 100°C were made for any of the 54 experimental cooking trials. 126 Because the heating rate at the center of the patty is controlled by the thermal gradients between the center of the patty and the patty surface, the patty surface temperature is the limiting factor for cooking time. Multiple linear regression was used to describe center temperature as a function of oven temperature, steam content, cooking time, and airflow. Table 5.19 shows the results of the regression. As expected, steam content, and cooking time were significant (P<0.05). A positive correlation existed between center temperature and steam content and cooking time. Not-surprisingly, cooking time had the largest effect on center temperature. More interesting however, is that oven steam content had a much more pronounced effect on center temperature than did oven temperature. The effect of steam content resulted in a potential temperature difference of up to 9.8°C over the range of steam contents tested, as compared to a non- significant effect of oven temperature. This compares to a difference of only 3.4°C related to oven temperature, indicating that oven steam content has a much greater effect on the surface wet bulb temperature of the patty than oven temperature. This observation should be taken into account when working to optimize oven settings. 127 Table 5.19 - Linear regression of patty center temperature as functions of oven temperature, steam content, cooking time, and oven airflow. (T=BO+BI .Toven +32 .M+B3 't+B4.Valr) Degrees of Sum of Mean F-Value P-Value Freedom Squares Square Regression 4 6548.757 1637.189 35.6310 <0.001 Residual 50 2297.425 45.9485 Total 54 8846. 182 Factor Coefficient Standard t-Statistic P-value Error Bo °C 30.3786 8.0882 3.7559 <0.001 01°C/°C 0.0306 0.0228 1.3391 0.187 B2 °C°%Steam" 0.2583 0.0641 4.0275 <0.001 B3 °C-s" 4.1720 0.4146 10.0639 <0.001 8. °C-s-m" -0.0218 0.2194 -0.0993 0.921 Oven air velocity did not significantly affect the patty center temperature. The temperature gradients that drive conduction within the patty result from increases in the patty surface temperature. Although increasing the airflow has the effect of increasing the heat transfer coefficient at the patty surface, corresponding increases in the mass transfer coefficient result in increased evaporative cooling in the later stages of cooling. The result was a zero net gain in the heating rate of the patty for the conditions tested. During the early stages of cooking, increases in airflow may increase the rate of condensation at the patty surface, thereby temporarily increasing the heating rate. However, as illustrated in the heating curves, the surface temperature reaches equilibrium within about 10 seconds for the range of oven conditions tested. The marginal gains in 128 surface temperature heating rate caused by increasing airflow did not significantly increase the rate of heating at the patty center. 5. 4.2 Cooking yield Figure 5.25 shows the relationship between patty center temperature and cooking yield for ground beef patties cooked in the J SO-IV impingement oven. Cooking yields decreased as a function of patty center temperature. 1 I 100.0 2' ,- mam"- _- i“, 95.0 90.0 85.0 . 80.0 75.0 3: .00 70.0 . 65.0 a: o 60.0 7E . .9 .8 55.0 o 50.0 - 40 50 60 70 80 90 100 110 Y T Cooking yield (percent) Center temperature (°C) Figure 5.25 - Cooking yield as a function of endpoint center temperature for ground beef patties cooked in a Stein JSO-IV industrial moist air impingement oven. The results of linear regression are shown in Table 5.20. Oven temperature, cooking time, and oven steam content were each found to significantly affect patty yield (P<0.05). Cooking time, oven temperature, and oven steam content each had a negative effect on cooking yield. Cooking time had the strongest effect on yield. Oven 129 temperature and steam content each had an effect approximately one order of magnitude below the effect of cooking time. Oven temperature created a potential difference in cooking yield of up to 5.6% over the range of temperatures used. Steam content had an effect equal to 4.8% yield over the range of steam contents tested. Table 5.20 — Regression parameters for cooking yield as a function of oven temperature, steam content, cooking time, and airflow. (Y=Bo +Bl .Tovel +B2 .M+B3 .t+p4 .valr) Degrees of Sum of Mean F-Value P-Value Freedom Squares Square Regression 4 3836.668 959.1669 50.0733 <0.001 Residual 50 957.7628 19.1553 Total 54 4794.43 Factor Coefficient Standard t-Statistic P-value Error [30 109.6251 5.2223 20.9917 <0.001 [31 °C'l -0.0503 0.0147 -3.4133 0.001 B; °o Steam’l -0.1259 0.0414 -3.0397 0.004 [33 s'2 -3.4932 0.2677 .13.0507 <0.001 8. s-m" -00453 0.1416 -0.3201 0.750 5.4.3 Fat loss For each patty, the cooking loss not accounted for by moisture loss was calculated and assumed to be entirely due to fat loss. The amount of fat lost increased with center temperature (Figure 5.26) and ranged from less than 1% of the initial mass at 50°C up to almost 10% of the initial mass at 95°C. This accounts for nearly all of the fat initially present in the patty. The results of linear regression for fat loss as a function of oven 130 parameters are shown in Table 5.21. Only cooking time significantly affected the yield lost due to fat. Clearly, the effects of fat loss must be taken into account to accurately model yield losses during cooking. Table 5.21 - Regression parameters for fat loss as a function of oven temperature, steam content, cooking time, and airflow. (AYfat =fi0 +51 °Toven +32 'M+B3 't+B4 'valr) Degrees of Sum of Mean F-Value P-Value Freedom Squares Smiare Regression 4 88.1928 22.0482 7.7278 <0.001 Residual 22 62.7679 2.8531 Total 26 150.9607 Factor Coefficient Standard t-Statistic P-value Error 130 -3.0612 2.9207 -l.0481 0.306 13l (°C") 0.0070 0.0081 0.8722 0.393 132 (% Steam") 0.0464 0.0230 2.0188 0.056 B3 (s") 0.0118 0.0025 4.7997 <0.001 134 (s-m'l) -0.0161 0.0771 -0.2085 0.837 131 10.0 - >. r .o 9 9.0 7:: . H T. ‘2 T3 8.0 7E: .0 . 1:: >4 : 0 {=3 :9 70 a . o o 09. c f o 8 3 5 0 f O . (U a . .. H o 4:. g T. 4.0 . . "’ H J- E a 3.0 *7 . 9 2 5 2.0 1‘ 0 0 :. i; E 10 ‘? . .0 0.0 7' . . . . a 4 - 4 40 60 80 100 Center temperature (°C) Figure 5.26 - Yield loss not accounted for by moisture loss as a function of endpoint temperature for ground beef patties cooked in a Stein JSO—IV industrial moist-air impingement oven. 5. 4.4 Volume change Patties exhibited a reduction in volume during cooking. The volume change was primarily due to dramatic reductions in the diameter of the patties during cooking. The diameter reduction as a function of endpoint center temperature is shown in Figure 5.27. The diameter of the patties decreased linearly as a function of center temperature. The thickness of the patties did not exhibit the same behavior. The thickness of the patties remained relatively unchanged during cooking, with approximately half of the patties undergoing slight decreases in thickness, and the other half undergoing slight increases. The average absolute change in thickness was 0.76 mm. The mean change in thickness was —0.14 mm. 132 40 35 o 30 7 9’ O z ‘ . ° 25 ‘. .0. ~ 0 g . ‘2'. 20 o 15_ ... Percent reduction in diameter fl r 10- o _ o a _ 1 1 1 1 1 1 1 1 1 l 1 - 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 l 1 I I . l 40 50 6O 70 80 90 100 1 10 Center temperature (° C) Figure 5.27 - Reduction in diameter during cooking as a function of cooking yield for ground beef patties cooked in a stein J SO-IV industrial moist-air impingement oven. The fact that the diameter of the patties changes so drastically during cooking, while the thickness stays relatively unchanged, presents a challenge for modeling cooking. To model cooking as completely as possible, volume changes should be incorporated into the model. However, for patties such as those utilized in these experiments, heat and mass transfer occur primarily in the axial direction, due to the height/diameter ratio of the patties. Since patty thickness remains relatively unchanged, changes in volume do not affect heat and mass transfer dramatically from the standpoint of modeling. However, further research in this area could shed light on the changing physical conditions that occur within meat during cooking. 133 5.5 Cooking model validation Validation of the cooking model was conducted using the data collected from cooking experiments with a Stein J SO-IV moist-air impingement oven (Sections 3.4 and 5.4). Validation was conducted for temperature, moisture, and yield. Additional comparisons were made using yield and temperature data from published sources (Murphy et al., 2001a and b). 5. 5. 1 Finite element mesh The finite element mesh utilized for cooking model validation consisted of 116 triangular elements with a total of 78 nodes. The element mesh was shown graphically in Chapter 4. A total of 28 nodes were located along the convective boundaries of the patty geometry. Mesh density was lowest at the center of the patty and increased near the surface where the temperature and moisture gradients were expected to be highest. The element mesh utilized was chosenoto balance solution accuracy with computing time. Increasing the number of elements from 48 to 116 lowered the transient standard error of prediction for the center temperature of a test experiment from 11.6°C to 55°C. This increase in accuracy came at the cost of computing time. The 48 element model only took 17 seconds to complete, compared to 75 seconds for the 116 element model. Increasing the number of elements from 116 to 234 did not improve the accuracy of the temperature prediction. However, the computing time required to run each test was increased to 295 seconds. This clearly illustrates that increasing the mesh density comes at a major cost in computing time. This is of importance to users utilizing the 134 model to simulate large numbers of conditions for applications such as process optimization. The value of the time step selected for the model was 1 second. This value was utilized to aid in comparisons with experimental data, which was recorded at one-second intervals. Unfortunately in certain cases, the one-second value for the time step may result in numerical oscillations if a constant element mesh size is utilized. Automatic routines for mesh generation and time step optimization could eliminate this potential problem. 5 . 5 .2 Temperature profile-experimental data Validation of temperature profiles was conducted by comparing transient center temperature data collected during the experimental tests to corresponding predictions generated by the model. Example center temperature profiles are shown in Figures 5.28- 5.30. All experimental and predicted temperature results are graphed in Chapter 9. 135 100 1 1 Ittrrt—r—I—rrW—v—rTrt—rrrvv—T—r Temperature (° C) (I. o 30 ~—-- . 1 20 lg: I Expenmenta 10 1E —— Model 0 _ t 4 t ' 1 ‘ 1 1 . 1’7. 0 50 100 150 200 250 Figure 5.28 — Example comparison of experimental temperature data with data generated by the cooking model (oven temperature=121°C, steam content=50%, air velocity=11.4 m/s). 100 : 90 80 70 6O 50 40 Temperature (° C) I Experimental L".— Model Time (s) Figure 5.29 - Example comparison of experimental temperature data with data generated by the cooking model (oven temperature=121°C, steam content=70%, air velocity=11.4 m/s). 136 100 90 80 70 60 50 40 3O 20 10 Temperature (° C ) I Experimental —- Model 0 100 200 300 400 Time (s) Figure 5.30 — Example comparison of experimental temperature data with data generated by the cooking model (oven temperature=121°C, steam content=88%, air velocity=11.4 m/s). Figure 5.28 is typical of most comparisons in several respects. Noticeable deviation between the model and experimental data occurred during the early stages of cooking. During the later stages of cooking, profiles for the model and experimental data were very close. Although the level of agreement between the model and experimental data varied between cooking runs, the phenomenon of the predicted center temperature lagging behind the measured temperature during the early stages of cooking was common to all cooking experiments. These results were similar to those reported by Pan et al. (2000) who found that temperature values lagged below predicted values for contact cooking of hamburger patties. These deviations were largest between 0 and 40°C. Disparities in the sample cooking curve provided by Pan et al. (2000) were very similar to the predicted error shown in Figure 5.28. 137 Based on Figure 5.28, several conclusions can be made. First, it is possible that the deviation between the temperature profile predicted by the model and the experimental data was caused by inaccurate predictions of thermo-physical properties in the model. The thermal conductivity may have been underestimated at low temperatures. The model also may have overestimated the heat capacity of the product at low temperatures. A second possibility for the deviation between the model and the experimental temperature profiles is the effect of volume change. The model did not take into account volume change during cooking. If the thickness of the patty changed significantly during cooking, the rate of heat transfer to the center of the patty may have changed, even under conditions of constant thermal properties. However, as discussed in Section 5.4.4, the thickness of the patties remained constant during cooking, with significant shrinkage occurring only in the radial direction. Because heat transfer occurred primarily in the vertical direction, it is unlikely that shape changes contributed significantly to error in the model for the product tested. Figures 5.29 and 5.30 illustrate a more pronounced deviation between the model and experimental temperatures during the first minute of cooking. This type of deviation was seen in about half of the 54 cooking trials. Typically, the deviation consisted of an increase in measured temperature above the predicted value followed by a drop in temperature to a level consistent with the predicted value. This deviation typically occurred over a temperature range of 0 to 40°C. At temperatures above 40°C, the experimental temperature profiles then closely matched the values predicted by the cooking model. 138 One possibility for the disparity between the model and experimental data in the early stages of cooking is that the thermocouple measuring center temperature may have moved as the meat began to cook. Changes in product texture, along with diameter changes occurring during cooking, would explain why the thermocouple could move during the early portion of cooking. However, this explanation does not account for the resolution between the measured and predicted temperature values at higher temperatures. The possibility that the author considers most likely is that fat-related effects were the cause of these deviations. From the fat holding capacity experiments in Section 3.3, large changes in fat holding capacitywere seen in the temperature range between 40 and 50°C. The laboratory experiments in Section 3.2 also show that yield loss is roughly linear at temperatures above 45°C. Most fat is in a form that is available for transport at temperatures above 45°C. Prior to heating, ground beef consists of a mixture of ground fat and muscle particles. Unlike whole muscle products, the fat particles are dispersed throughout the meat, rather than in their naturally occurring structures. In this sense, the fat particles exist as a dispersed phase within a matrix of lean meat. When the fat particles are in solid form, physical forces serve to keep them in suspension within the meat. However, upon melting, the fat is free to exit the meat through capillary mechanisms. It is proposed that during the initial phases of cooking, changes in the meat structure occurred that resulted in the thermocouple not giving an accurate representation of the average temperature in the center of the patty. Due to the near-frozen condition of the meat before cooking, the meat had a very firm initial texture. As a result, when the 139 thermocouple was inserted into the patty, a minute void space may have been created around the thermocouple, which would have provided a channel for condensing water to penetrate to the center of the patty, and thereby increase the thermocouple temperature above that of the meat. At temperatures above 40-50°C, constriction of the meat and filling of internal voids by melting fat would result in more intimate contact between the thermocouple and the meat matrix, thereby resulting in more accurate temperature readings. The vertical orientation of the thermocouple may have contributed to this type of effect by providing an uninterrupted vertical channel between the bottom and center of the patty. This theory would explain the deviations seen in Figures 5.29 and 5.30. It is also highly likely that the deviation in Figure 5.28, although less pronounced than the others, was caused by a similar effect. In the temperature curves where significant deviations occur between the measured and predicted center temperature values, the measured center temperature value appears to temporarily move in the direction of the surface temperature of the patty, before dropping back down to a level more consistent with the predicted value. This seems to support the hypothesis that condensing water at the surface temperature may be penetrating the patty along the thermocouple “channel”. Unfortunately, accurate surface temperature values were not available for many of the cooking trials due to the difficulty in maintaining uninterrupted contact between the thermocouples and the patty surface during cooking. The high airflow of the oven, combined with changes in patty geometry, often separated the thermocouples from the patty surface. For this reason, surface temperature data are missing from many of the figures in Section 9.1. Due to inconsistentcies in the surface temperature data, it was not 140 considered meaningful to calculate transient SEP for the surface temperature of each model run. Instead, general observations of the predicted and measured surface temperatures were made. For each simulated cooking run, the surface temperature of the meat quickly rose to a semi-equilibrium value. This temperature reflects the wet bulb temperature of the oven air. This temperature never exceeded the boiling point of water for any of the patties tested. For patties in which accurate surface temperatures are available, the predicted surface temperatures corresponded closely with the experimental values. In many cases, the measured surface temperature values were considerably different from the predicted values. However, it is the opinion of the author that these cases represent situations in which the thermocouple was not in contact with the actual surface of the patty. In some cases the thermocouple may have been lodged slightly below the surface of the patty. These cases illustrate the difficulty inherent in measuring the surface temperature of meat patties within commercial convection ovens. Standard error of prediction was calculated for the center temperature data of each cooking trial, using temperature data collected at 1 second intervals from both the model and experiments. The SEP of the transient center temperatures for individual trials ranged from 2.1 to 139°C (Table 5.21). The overall SEP for all of the cooking trials was 80°C. The SEP for the final center temperature of all of the patties was also 80°C. The error in the temperature curves was generally concentrated in the early portion of the curves. The cooking experiments that were run for short times had the largest errors. As the patty center reached high temperatures, the differences between the model and experimental data became small. A second set of SEP data was generated for the portion of each 141 cooking run above 45°C (Table 5.22). These SEP’s ranged from 0.5 to 108°C, a large improvement from the values taken over the entire temperature range. The overall SEP for all data points above 45°C was 58°C. The trials with large SEP values for temperature were generally the experiments with the shortest cooking times. In these tn’als, deviations between the model and experimental data at low temperatures affected a larger percentage of the total cooking time, thus resulting in larger SEP between experimental and predicted data. 142 Table 5.22 — Standard error of prediction for the entire trial (SEP) and for data above 45°C (SEPNsoc) for center temperature of beef paties. EXP- SEP (°C) SEP1>45°C (°C) EXP- SEP (°C) SEPT>45°C (°C) # (data points) (data points) # (data points) (data points) la 5.5 (193) 1.2 (128) lb 4.7 (157) 5.1Q3) 6a 3.9 (416) 4.1 (330) 6b 3.3 (418) 3.1 (324) 8a 2.4 (300) 1.2 (214) 8b 6.9 (303) 3.4 (247) 11a 6.9 (294) 2.8 (220) 11b 8.9 (295) 3.6 (244) 13a 4.0 (187) 3.1(115) 13b 12.5 (189) 9.6 (152) 18a 4.1 (3%) 3.6 (340) 18b 7.6 (396) 1.9 (341) 21a 5.5 (368) 2.4 (298) 21b 7.1 (369) 1.3 Q94) 23a 3.0 (320) 3.g229) 23b 5.1 (301) 4.3 (214) 25a 5.1 (L41) 4.1 (60) 25b 8.6 (117) 6.3 (60) 30a 3.2 (276) 2.1 (177) 30b 6.8 (280) 3.3 (206) 32a 4.2 (224) 1.7 (126) 32b 14.7 (224) 12.3 (196) 34a 5.3 (125) 2.5 (29) 34b 12.5 (116) 9.1 (68) 37a 7.1 (105) 0.5 (25) 37b 23.8 (108) 21.5 (69) 47a 2.1 (264) 1.7 (187) 47b 9.74273) 5.9 (206) 49a 8.0 (114) 9.5 (26) 49b 16.5 (120) 8.0 (27L 50a 9.5 (240) 7.7 (125) 50b 3.6 (240) 2.5 (129) 54a 13.5 (304) 10.8 (198) 54b 7.3 (313) 6.9 (254) 56a 6.3 (188) 5.5 (112) 56b 17.8 (186) 12.4 (119) 58a 13.1 (118) 7.1 (21) 58b 5.5 (116) 8.0 (23) 63a 7.2 (259) 5.1 (190) 63b 4.5 (262) 2.1 (169) 66a 4.0 (256) 1.8 (177) 66b 5.4 (260) 5.7 (161) 68a 2.8 (186) 0.8 (112) 68b 7.5 (190) 8.1 (83) 70a 7.1 (75) * 70b 12.0 (70) 9.9 (11) 73a 10.7 (77) * 73b 8.8 (73) 2.4 (5) 75a 8.0 (259) 2.3 (195) 75b 13.2 (259) 12.4 (235) 78a 2.2 (191) 0.9 (98) 78b 2.7 (192) 2.2 (109) 80a 8.3 (191) 2.5 (114) 80b 11.4 (192) 8.9 (161) * Center temperatures did not exceed 45°C 5 .5 .3 Temperature profile-published data A total of 12 model runs were conducted to simulate the conditions utilized in the experiments of Murphy et al. (2001a). The center temperature profiles of each model run were compared to temperature predictions from the regression equation developed by Murphy et al. (see Section 3.5.3). Transient comparisons were made between the center 143 temperatures predicted by the model and the Murphy et al. regression equation for the temperature range between 55 and 80°C. The model and regression transient temperature data were compared at 1 second intervals. Standard error of prediction for transient center temperature was calculated for each model run (Table 5.23). Table 5.23. Standard error of prediction for transient center temperature of ground chicken breast patties predicted by the model and by the regression equation of Murphy et a1. (2001 a). Dry Bulb Steam by Airflow SEP Center Temperature (°C) Volume (m/s) Temperature (°C) 149 6 1.53 10.3 149 6 2.13 12.4 149 6 2.73 13.5 149 25 1.53 1.4 149 25 2.13 1.6 149 25 2.73 2.4 149 60 1.53 2.6 149 60 2.13 1.6 149 60 2.73 1.8 149 91 1.53 3.7 149 91 2.13 2.8 149 91 2.73 3.2 For the wet bulb temperatures of 70 to 95°C, the SEP ranged from 1.1 to 61°C. This was comparable to the SEP of the experimental data collected from the Stein J SO- IV oven. However, for the driest cooking air condition (wa=40°C), the SEP ranged from 13.3 to 157°C. This indicates that the cooking model may not be reliable for extremely dry air conditions. This was probably due to the equation for equilibrium relative humidity. However, the cooking model was designed for high moisture impingement 144 ovens, so deviations at extremely dry conditions are not critical for the intended use of the model. 5. 5.4 Moisture content- experimental data For each experimental cooking run in the Stein J SO-IV oven, the endpoint moisture content predicted by the model was compared with the experimental values (Section 5.4.2). The deviations between the model and experimental moisture contents ranged from —5.2% to 3.6% wet basis moisture, with an average deviation of —0.04% wet basis (Table 5.24). This indicates that the model moisture predictions are centered around the measured values with little bias. Standard error of prediction for the complete set of final moisture contents was 2.3% moisture w.b. 145 Table 5.24 — Difference between measured and predicted moisture content for each oven condition. Experiment numbers correspond to the conditions listed in Table 3.1. Experiment Difference between measured and Number predicted moisture content (% wet basis) 1 1.6 6 -4.7 8 -2.5 11 0.8 13 2.4 18 -1.2 21 -0.5 23 1.7 25 3.6 30 -5.2 32 -3.7 34 -0.3 37 0.0 47 0.0 49 -l.2 50 1.5 54 1.2 56 2.0 58 -O.8 63 -l.7 66 -l .3 68 0.6 70 0.0 73 1.8 75 1.2 78 3.5 80 4.0 5. 5.5 Cooking yield — experimental data For each cooking run in the Stein J SO-IV oven, the cooking yield predicted by the model was compared to the experimental value. Standard error of prediction for the complete set of cooking yields was 5.9%. The deviation in predicted yield ranged from — 146 10.2 to 10.5% with an average deviation of -—1.2% (Table 5.25), indicating that the model had a slight bias towards overpredicting yield loss. Table 5.25 — Difference between measured and predicted cooking yields for each oven condition. Experiment numbers correspond to the conditions in Table 3.1. Experiment Difference between predicted and Number measured cookingxield (% yield) 1 3.7 6 10.5 8 6.3 11 -3.4 13 -6.5 18 4.1 21 -2.3 23 -2.6 25 -5.7 30 9.3 32 0.9 34 4.1 37 -5.3 47 -3.6 49 -7.6 50 -10.2 54 -6.0 56 8.9 58 -4.7 63 -l.1 66 2.1 68 -0.4 70 2.4 73 -1.4 75 -4.3 78 -10.2 80 -8.4 147 5. 5. 6 Cooking yield — published data Yield predictions of the model were also compared to yield predictions from the literature (Murphy et al., 2001b). The published paper presented a regression model for the yield of ground chicken patties cooked in a Stein model 102 impingement oven. The model and published yield equation were compared using the same procedures described in Section 5.4.1.1. Like the temperature predictions, the yield data was a much closer fit for wet bulb temperatures between 70 and 95°C. The SEP in that temperature range varied from 1.1 to 6.1% yield. At a wet bulb temperature of 40°C, the SEP ranged from 13.3 to 15.7% yield. 5.6 Lethality model validation For each run of the computer model, two sets of Salmonella lethality data were generated. The first set of data was the inactivation profile at the center point of the patty, generated using a log-linear equation. For the second data set, total inactivation of Salmonella within the patty was determined. A volume averaging procedure was used to determine total inactivation from inactivation at all of the nodal points. This overall reduction in Salmonella is the value that would be determined experimentally when counting the number of surviving organisms in a whole ground beef patty. Due to facility constraints, it was not possible to perform inoculated challenge studies for Salmonella in the moist-air impingement oven used in this study. However, data exist in the literature for tests that were conducted in a pilot-scale impingement oven (Murphy et al., 2002). 148 In order to minimally validate the combined cooking/ inactivation model, simulations were conducted to compare values of microbial inactivation predicted by the model with those from the literature. The cooking model was run using the cooking conditions used by Murphy et al. (2002), as described in Section 3.6. The resulting inactivation profiles were plotted on the same graphs as data points reported in the published study (Figures 5.31 and 5.32). 10 t 9t 8 7 1:1 7 5 _ __._ ___._ ______ a: 6 t D ,/"/ B E E a 5 £5 /0/ D D £2” 41; ’ n B 3 If” 2~; 1 4“ 0‘»1...1..1,....;...-i...- 150 160 170 180 190 200 Time(s) Figure 5.31 - Comparison between model Salmonella Senftenberg lethality predictions and data points published by Murphy et al. (2002). 149 10 9 .Z 8 L - 1:1 :1 A 7 5 "~6- -- — € 6 7? Q E a 5 t; B E / 8° 4 4E a 2 !/ IE 1 4:. 0: 4 1H+1~fii~w 150 160 170 180 190 200 Time(s) Figure 5.32 — Comparison between model Listeria innocua lethality predictions and data points published by Murphy et al. (2002). The Salmonella Senftenberg inactivation curve predicted by the model was based on D and z-values for Salmonella Senftenberg heated in turkey (Murphy et al., 2003). For heating times of 150 and 160 seconds, the inactivation curve predicted by the model is within the bounds of the experimental data reported by Murphy et al. (2002). However, the model reaches a maximum reduction of 7-log (set by the initial inoculum level) approximately 15 seconds before the experimental data. At times above 180 seconds, both the model and experimental data indicate a reduction of 7—logs. Standard error of prediction for the Salmonella inactivation curve was 1.3 logs (n=21). A comparison of the Listeria inactivation curve predicted by the model and the experimental data of Murphy et al. (2002) indicated a close fit between the experimental 150 and predicted inactivation’s. The model inactivation curve was generated using D and 2- values for Listeria innocua in ground turkey (Murphy et al., 2003). At a heating time of 150 seconds, the model slightly underestimated the level of lethality, although the model predictions were within l-log of the lower end of the experimental data range. As heating time increased, the inactivation curve predicted by the model closely estimated the experimental data. Standard error of prediction for the L. innocua inactivation curve was 1.1 logs (n=21). These comparisons only give a rough verification of the ability of the cooking model to predict microbial inactivation. However, the temperature prediction capability of the model has been verified much more extensively. Combining the temperature prediction capabilities of the model with experimentally derived kinetic parameters for microbial estimation (D and z-values) should enable the cooking model to produce an acceptable first estimate of microbial lethality during cooking. In addition, due to the design of the cooking model, the effects of process changes on corresponding changes in microbial lethality are readily visible. This makes the combined cooking and Salmonella inactivation model a valuable tool for predicting the effects of processing parameters on microbial safety. 5.7 Illustration of model utility The utility of a combined cooking and inactivation model is several-fold. Some examples are listed below. 151 1. A processor uses a moist-air impingement oven to produce ready-to-eat ground beef patties. The oven is currently operated at a dry bulb temperature of 177°C, a steam content of 75% by volume, and an air velocity of 18 m/s. The dwell time in the oven is 3 minutes. The processor wishes to increase the throughput of the oven by adjusting the moisture content of the cooking air from 75 to 85% moisture by volume. The processor wants to know what the new oven dwell time will be to reach the same patty center temperature that was achieved by the previous process. In this case, determining the new dwell time in the oven could be determined by running the cooking model under the new set of process conditions. Running the model under the original cooking conditions, the processor finds that the original final center temperature was 787°C. The processor then runs the model under the new conditions and sees that the dwell time that will achieve 787°C is 170 seconds, a ten second improvement over the original conditions, which would translate into a 5.8% increase in throughput. 2. In the situation above, the processor is concerned that the faster cooking time will not be adequate to achieve the desired level of microbial inactivation. In this case, the processor can use the lethality prediction function of the computer model. Running the cooking model, the processor sees that the cooking process greatly exceeds the required lethality in both cases. However, based on a 6.5-log target 152 reduction, the higher steam content cooking air achieves the desired lethality 14 seconds earlier than the original condition. 3. The processor is not convinced that the current operating conditions are making the most efficient use of the oven system. Experience indicates that increasing the steam content of the oven will decrease cooking time. However, quantitative data are not available. The processor would like to develop a set of experiments to optimize the oven settings. Taking the oven system offline to perform optimization experiments is time consuming and expensive. In addition to the time and manpower spent conducting the experiments, the oven must be taken out of production for the time period of the experiment. To determine the effects of changing process conditions, the processor runs simulations using the cooking model. Examples 1 and 2 show how the model can be used to illustrate changes in process conditions. By running the model under many more sets of conditions, the processor can determine the ideal settings for the desired cooking results. 4. Product output from oven systems sometimes falls behind due to problems with equipment and backups during previous unit operations. Operators might try to catch up with production quotas by increasing the oven belt speed. The manager is concerned that the desired level of microbial lethality is not being achieved due 153 to increases in belt speed. It is difficult for the manager to illustrate to the operators that their actions may result in food that is potentially unsafe. This instance illustrates the benefit of the graphical temperature and lethality outputs of the model. The graphical outputs can be used to illustrate the effects of process changes to personnel with no knowledge of heat transfer or microbial inactivation kinetics. If the oven conditions utilized in Example 2 were in use, the oven would achieve a 6.5-log reduction in Salmonella in 129 seconds. Decreasing the dwell time by ten seconds (119 seconds cooking time) would decrease the predicted lethality to only 0.8-log. In this case, a very small adjustment in cooking time would result in a product that is potentially unsafe. 154 6 CONCLUSIONS The main conclusions of this study were: 1. Fat content had a significant effect on the cooking time for ground turkey, ground beef, and ground pork. Differences in the time required to reach 85°C between different fat contents of each meat species were 171, 217, and 102 seconds for ground turkey, ground beef, and ground pork patties, respectively. Higher fat samples of ground beef and ground pork cooked faster than the lower fat samples. The opposite was true for ground turkey. Fat content significantly affected the yields of ground turkey with initial fat contents of 1.4% and 8.6%, and ground pork of initial fat contents of 15.7% and 41.9% fat. Differences in yield between fat contents of ground turkey and ground pork were 18 and 17%, respectively, when cooked to 85°C. No significant differences in yield were measured between ground beef with initial fat contents of 7.2 and 17.5%. Fat transfer contributed up to 6% of the yield loss of ground beef and up to 28% of the yield loss of ground pork patties. 2. Fat holding capacity of ground beef was modeled as a function of initial fat content and heating time using multiple linear regression. Holding time did not significantly affect fat holding capacity. The regression equation for fat holding capacity had a coefficient of variation of 0.81 and a standard error of prediction of 0.068 g fat/g dry matter. This equation was an important step towards developing 155 a model for heat and mass transfer of ground beef products that includes the effects of fat transport. The effects of oven temperature, steam content, and airflow on heating time, yield, fat loss, and volume change of ground beef patties cooked in an industrial moist-air impingement oven were quantified. Multiple linear regression showed that oven steam content had the largest effect on both heating rate and cooking yield. Neither oven temperature nor air velocity had a significant effect on patty center temperature. Oven temperature, steam content, and cooking time all had significant affects on cooking yield. Fat loss increased roughly linearly, and patty diameter decreased roughly linearly with patty temperature. Patty thickness remained fairly constant during cooking. A finite-element method-based cooking and Salmonella inactivation model was developed. The model was an improvement over previously published models for the following reasons: a) The heat and moisture transport portions of the model were based completely on heat and mass transfer principles. Empirical correlations were only used for thermal property relationships. This allowed for maximum flexibility of the model over different cooking conditions. b) Fat transport was incorporated in the model. The incorporation of fat transport created a model that could predict yield losses for products containing a wide range of fat levels. 156 c) The model incorporated the transient effects of heat and mass transfer related to moist-air impingement, including the effects of condensation as a surface mass transfer process (rather than as an “effective heat transfer” effect). d) The model was validated with experimental data from an industrial moist- air impingement oven. The cooking model was validated using data collected from an industrial moist-air impingement oven. The transient standard error of prediction for temperature was 8°C. The model more accurately predicted temperatures above 45°C. At temperatures above 45°C, the SEP was reduced to 57°C. Predictions of moisture content had errors ranging from O to 5.2% wet basis. Predictions of cooking yield had errors ranging from 0.1 to 15.4% with an average deviation of 5.9%. Comparisons with temperature and yield data compiled by other researchers were also favorable. Standard error of prediction for center temperature of chicken patties at cooking air wet bulb temperatures between 70 and 95°C ranged from 1.4 to 3.7°C. Higher errors occurred at an air wet bulb temperature of 40°C. Standard errors of prediction for yield ranged from 1.1 to 15.7%. The microbial lethality prediction of the model was compared to published data. Standard error of prediction for inactivation of Salmonella Senftenberg in ground beef patties was 1.3 logs (CFU/g). Standard error of prediction for Listeria innocua was 1.1 logs (CFU/g). 157 7 FUTURE WORK There are a number of studies that should be conducted in order to further develop our understanding of the dynamics of meat and poultry cooking. The results of these studies could be utilized for the development of more precise and more robust cooking models. In addition, further understanding of the mechanics of cooking could improve our general understanding of the effects of processing on meat and poultry quality attributes. Recommendations for further studies include: 1. Experiments designed to determine the effects of protein and fat chemistry on fat and moisture transport should be conducted. Specifically, the effects of fat composition on transport properties should be investigated. Models for fat viscosity as functions of lipid composition and temperature could contribute greatly to understanding of fat transport mechanisms. Understanding of these mechanisms is needed to maximize the effectiveness of cooking models for a wide range of product compositions. 2. An analysis of the changes in meat microstructure during cooking should be performed. Such studies should include analysis of the changes in porosity during cooking as well as the relationships between meat microstructure and volume change. The dynamics of volume change during cooking should be investigated. 158 Studies should be designed to determine the driving forces for fat transport and moisture transport via drip loss. These phenomena are generally attributed to changes in water and fat holding capacities, but there is little understanding of the physical mechanisms behind such changes. Further knowledge of these driving forces is needed for the development of mechanistic models for water and fat holding capacities and transport. Extensive validation of microbial inactivation kinetics during impingement cooking of meat and poultry products should be performed. This should be done utilizing inoculated challenge studies with an actual moist-air impingement oven. Automatic mesh generation subroutines should be added to the cooking model. This would allow for much easier modeling of cooking for various product geometries. It would also allow for the cooking model to account for volume change during cooking. The model should be used for optimization of impingment cooking processes, in terms of safety, cooking yield, and oven throughput. The model should be adapted in the future to simulate cooking of whole-muscle products. 159 8 APPENDICES 160 8.1 Model and experimental temperature versus time curves for moist-air impingement cooking of ground beef patties. Surface Center Temperature (’ C) Time (s) 200 250 Figure 8.1 - Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment 1a). 100 a: 35:2, Surface 80 x )p’iw" ~ - o . ‘ my 3 60 a g 0 4 g 7' Center P“ 20 / — Ode] O . . . 4 . 4 + L4. 4 . 1 0 50 100 150 200 Time (s) Figure 8.2 - Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment lb). 161 100 G 80 Q/ 2 3 60 g 40 E“ D E" 20 O 1 O 100 200 300 400 500 Time (s) Figure 8.3 — Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 6a). 100 00 0 Temperature (’ C) 4s Ox o o N O 500 Time (s) Figure 8.4 - Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 6b). 162 40 Temperature (’ C) 0‘ O N O 0 50 100 150 200 250 300 350 Time (s) Figure 8.5 - Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 8a). p—a O O 00 O Temperature (J C) O\ o 20 O 50 100 150 200 250 300 350 Time (s) Figure 8.6 - Oven temperature: 121°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 8b). 163 100 , , Surface 6‘ 80 1 3 t E 60 37 Center g 40 8‘ 13 20 * —Model 0 , 44 4 4 4 0 50 100 150 200 250 300 350 Time (s) Figure 8.7 — Oven temperature: 121°C, oven steam content: 70% by volume, oven airflow: 11.4 m/s (Experiment 11a). 40 Temperature (’ C ) N C Figure 8.8 - Oven temperature: 121 airflow: 11.4 m/s (Experiment 11b). Center 150 200 Time (s) °C, oven steam content: 70% by volume, oven 164 100 , 00 O Center 1 Temperature (’ C) 4:. O\ o o — Model 1 J___l_1 141 1 l O 50 100 150 200 Time (s) Figure 8.9 — Oven temperature: 121°C, oven steam content 70% by volume, oven airflow: 16.8 m/s (Experiment 13a). 100 _ 80 ~: 60 at Center Temperature (’ C) 40* 20 ' ——Model 0 4 4 4 4 4 - 4% 0 50 100 150 200 Time (s) Figure 8.10 — Oven temperature: 121°C, oven steam content 70% by volume, oven airflow: 16.8 m/s (Experiment 13b). 165 31—. O O OO O 05 0 Temperature (’ C) h o 0 100 200 300 400 500 Time (s) Figure 8.11 - Oven temperature: 121°C, oven steam content: 70% steam volume, oven airflow: 21.8 m/s (Experiment 18a). — 00 O O O 11*r a O (D 1 Temperature (’ C) 0‘ o 40 20 8 —Model 0 4 4 4 4 4 4 44 4 4 4 4 O 100 200 300 400 500 Time (s) Figure 8.12 - Oven temperature: 121°C, oven steam content: 70% steam volume, oven airflow: 21.8 m/s (Experiment 18b). 166 100 Temperature C C) O 1 1 1 1 1r 1 1 1 1 i 1 O 100 200 300 400 Tirm (s) Figure 8.13 — Oven temperature: 121°C, oven steam content: 88% by volume, oven airflow: 11.4 m/s (Experiment 21a). 100 80 60 Temperature (’ C) 0 100 200 300 400 Time (s) Figure 8.14 — Oven temperature: 121°C, oven steam content: 88% by volume, oven airflow: 11.4 m/s (Experiment 21b). 167 100 _ a 80 7: a.’ _ E 60 ,5 Center 8 g. 40 4, ,2 E 20 if —Model 0 P 1 % 1 L 0 50 100 150 200 250 300 350 Time (s) Figure 8.15 — Oven temperature: 121°C, oven steam content: 78% by volume, oven airflow: 16.8 m/s (Experiment 23a). 100 6 80 w 8 60 a 8.40 E o . ‘" 20 ——Model 0 4444444444444444444fi44444444444444 0 50 100 150 200 250 300 350 Time(s) Figure 8.16 - Oven temperature: 121°C, oven steam content: 78% by volume, oven airflow: 16.8 m/s (Experiment 23b). 168 fl 0 O OO O Center Temperature C C) A O\ o O —— Model N O l 0 50 100 150 Time (s) C Figure 8.17 - Oven temperature: 121°C, oven steam content: 78% by volume, oven airflow: 21.8 m/s (Experiment 25a). p— O O OO O Center Temperature C C) A 04 o o N O — Model lLllllll T I I 0 20 40 60 80 100 120 140 Time (s) Figure 8.18 - Oven temperature: 121°C, oven steam content:78% by volume, oven airflow: 21.8 m/s (Experiment 25b). 169 .—-0 O O E Surface m C 1 Temperature (’ C) A O\ o o N O l tit O .L T I Time (s) Figure 8.19 — Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 11.43 mls (Experiment 30a). 100 Surface __ 6 80 -. b ,- 2 5 g 60 g. 4 ' Center 8. 40 ' E x 43 20 ' ——Model 0 " 4 4 44 4 4 4 4 4 4 4 4 0 50 100 150 200 250 300 Figure 8.20 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 11.43 mls (Experimen 30b). 170 ._- O O OO 0 Temperature C C) A ON C O N O 0 50 100 150 200 250 Time (s) Figure 8.21 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 16.8 mls (Experiment 32a). Temperature C C) N O 250 Time (s) Figure 8.22 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 16.8 mls (Experiment 32b). 171 Temperature (D C) 140 Time (s) Figure 8.23 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 34a). 6‘ 0.x E 4*" g Center E d) {-41 — —— Model 11111111111111111411 l T T l 60 80 100 120 140 Time (s) Figure 8.24 - Oven temperature: 177°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 34b). 172 .—-s O O OO O Temperature C C ) A as o o N O O 120 Time (s) Figure 8.25 - Oven temperature: 177°C, oven steam content: 70% by volume, oven airflow: 11.4 m/s (Experiment 37a). H O O 00 O Temperature (’ C ) A as o o N O Time (s) Figure 8.26 - Oven temperature: 177°C, oven steam content: 70% by volume, oven airflow: 11.4 mls (Experiment 37b). 173 p.— O O OO O .L VIITI Center Temperature (’ C) O‘ o 40 20 E —Model 0 44444444444 44 0 50 100 150 200 250 300 Time(s) Figure 8.27 - Oven temperature: 177°C, oven moisture content: 83% by volume, oven airflow: 11.4 mls (Experiment 47a). G 1.x E Center 4; g: : o _ {-4 : E! —Model 0 4 . . . 1 4 1 .1. 0 50 100 150 200 250 300 Tine (s) Figure 8.28 - Oven temperature: 177°C, oven moisture content: 83% by volume, oven airflow: 11.4 mls (Experiment 47b). 174 p—l O O OO O Center Temperature (’ C) O'\ O -— Model OJ: LlrllljiliiliTl 1711111T t O 20 40 60 80 100 120 140 Tine (s) Figure 8.29 - Oven temperature: 177°C, oven steam content: 84% by volume, oven airflow: 16.8 mls (Experiment 49a). ~ O O OO O Temperature (’ C) A ox o o N O Time (s) Figure 8.30 - Oven temperature: 177°C, oven steam content: 84% by volume, oven airflow: 16.8 mls (Experiment 49b). 175 .— O O OO O Surface Center Temperature C C) 04 o 40 2' 29 ": ——Model 0 i A T I L T JT 0 50 100 150 200 250 300 Time (s) Figure 8.31 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 16.8 mls (Experiment 50a). Temperature (’ C) Titre (s) Figure 8.32 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 16.8 mls (Experiment 50b). 176 p—a O O 00 Q g (D Temperature C C) A ox o o N O l O A III ~44— .4 0 50 100 150 200 250 300 350 Tine (s) Figure 8.33 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 21.8 mls (Experiment 54a). 5 o 0‘ m o O 1 O 4 0 5' N O 44 4 — Model 11111111111114111111 I I I O .4— _4... _. Temperature (’ C) 3 -\ 1““ . 0 50 100 150 200 250 300 350 Titre (s) Figure 8.34 - Oven temperature: 177°C, oven steam content: 86% by volume, oven airflow: 21.8 m/s (Experiment 54b). 177 1—0 0 O OO O Temperature (’ C ) Ox o 40 Center 20 44" i: —Model 0 ’ 1 ,_ . 1 1 1 1 1 1 1 0 50 100 150 200 Tine (s) Figure 8.35 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 11.4 m/s (Experiment 56a). H O O 4; O N 0 Temperature 6 C) O\ 00 o o Model 0 50 100 150 200 Time (s) Figure 8.36 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 11.4 mls (Experiment 56b). 178 # O O OO O 1 1 1 1 Temperature C C) A O\ o o N O O Time (s) Figure 8.37 — Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 58a). p—ua O O 00 O A O Center Temperature (’ C) O\ o 20 — Mode lllllLlllll I 0 20 40 60 80 100 120 140 Time (s) Figure 8.38 — Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 16.8 m/s (Experiment 58b). 179 y—n O O I ESIn’face A 80 4K 0 Lu 5 6O 4 £3 E Center g. 40 7 Q) at E“ 20 _ —Model O11411111)111111111+L141411111 0 50 100 150 200 250 300 Tirrre(s) Figure 8.39 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 21.8 m/s (Experiment 63a). 100 , E Surface ‘ G 80 .4%/{r K1 93- 60 44 a - E. 40 k E Q) l" 20 4 —Model 0 I— 1 L4 + 1 1 1 1 1 1 1 1 1 l 1 J #1 % 1 1 1 l l J 1 1 1 0 50 100 150 200 250 300 Time (s) Figure 8.40 - Oven temperature: 232°C, oven steam content: 50% by volume, oven airflow: 21.8 mls (Experiment 63b). 180 .—e O 0 Surface 4 - 00 O 1 Center Ch O Temperature (3 C ) h o — Model N O 11m111111 l I 0 50 100 150 200 250 300 Time (s) —I>— Figure 8.41 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 11.4 mls (Experiment 66a). )— O O OO O Center Temperature C C ) as o 40 20 ~ —— Model 0 44444444 4 4 e 4 4 0 50 100 150 200 250 300 Tim: (8) Figure 8.42 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 11.4 mls (Experiment 66b). 181 Temperature (’ C ) Time (s) Figure 8.43 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 16.8 mls (Experiment 68a). 4o . Center Temperature (J C) — Model 1 1 1 4L 1 1r 1 1 1 0 50 100 150 200 Tine (s) 4 O l 1- Figure 8.44 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 16.8 m/s (Experiment 68b). 182 Surface Temperature C C) 80 Time (s) Figure 8.45 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 21.8 m/s (Experiment 70a). 100 , F Ice V .......... 80 7: Temperature C C) 80 Tine (s) Figure 8.46 - Oven temperature: 232°C, oven steam content: 70% by volume, oven airflow: 21.8 mls (Experiment 70b). 183 H O O Temperature C C) A O\ 00 o o o N O O 1 00 Tina (s) Figure 8.47 - Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 m/s (Experiment 73a). 100 a 80 _ “““ “ 1.x a e _ B 60 I § 46: E O , E" 20 44 O 4 A 0 20 40 6O 80 Time(s) Figure 8.48 - Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 mls (Experiment 73b). 184 H O O Temperature 6 C) A ON 00 o o o N O 0 50 100 150 200 250 300 Titre (s) Figure 8.49 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 mls (Experiment 75a). 100 6 80 ~ Ru 2 a a 60 8 40 E ,2 20 1 ——Model 0 4 44 444444444 444 4 4 0 50 100 150 200 250 300 Tine (s) Figure 8.50 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 11.4 mls (Experiment 75b). 185 H O O OO O 1 O\ O Center Temperature (’ C) A o N O — Model 1 1 1 1 1 1 I 0 50 100 150 200 Time (s) III Figure 8.51 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 16.8 mls (Experiment 78a). y—a O O OO O Center Temperature 6 C) 0‘ o 40 20 i . ——Model 0 4 4 4 4 4 4 4 4 4 0 50 100 150 200 250 Time (s) Figure 8.52 — Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 16.8 mls (Experiment 78b). 186 Temperature C) C) N O O 50 100 150 200 Time (s) Figure 8.53 - Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 21.8 mls (Experiment 80a). _ O O Temperature (’ C) A as oo o o o N O l 1 1 1 1 04 0 50 100 150 200 250 Time (s) c—lh— Figure 8.54 - Oven temperature: 232°C, oven steam content: 82% by volume, oven airflow: 21.8 m/s (Experiment 80b). 187 8.2 Derivation of cooking-air thermo-physical property equations The oven-air thermo-physical properties used in the cooking model were modeled using non-steady-state relationships. The latent heat of vaporization for water as a function of temperature was modeled using linear regression of tabular data (Geankoplis, 1993); (Equation [8.1]). ,1 = -—2.429 - T + 2502.8 [ 8.1 ] The resulting regression had an R2 value of 0.998 (Figure 8.55). 2550 1 N U! C O 2450 f 2400 4 2350 f 2300 g 3.“, = -2.429T + 2502.8 Latent heat of vaporization (J/ g) 2250 1 R2 = 0.9998 2200 :44444 14.44 11.41.11 .. .. ,. 1. ., .. ,1 0 20 40 60 80 100 120 Temperature (° C) Figure 8.55 - Latent heat of vaporization for water as a function of temperature (From tabular data: Geankoplis, 1993). 188 The viscosity of air was modeled as a function of temperature using regression of tabular data (Geankoplis, 1993); (Equation [8.2]). 14144 = —0.0000000002 - T2 + 0.0000005 -T + 0.0002 [ 8.2 ] The resulting regression equation had an R2 value of 0.9998 (Figure 8.56). 0.0003 _ 0.00028 43 A 0.00026 .3 0.00024 l g/c .9 o o o N N W .8, 0-00013 11,, = 213101“2 + 513-071~ + 0.0002 g 0.00016 4: E R2 = 0.9998 0.00014 g 0 00001 P111111L11111m1111111111111111 . l T I I l 0 50 100 150 200 250 300 Temperature (° C) Figure 8.56 - Viscosity of air as a function of temperature (From tabular data: Geankoplis, 1993). The viscosity of steam was modeled as a function of temperature using regression of tabular data (Geankoplis, 1993); (Equation [8.3]). 41.4.... = —0.00000001-T2 + 0.00004 - T + 0.0089 [ 8.3 ] 189 The resulting regression had an R2 value of (Figure 8.57). 0.02 4 0.019 4;_ 0.018 1; 7g 0.017 g, 0.016 g 34 0.015 4; 0.014 f 0.013 ; 0.012 45 0.011 0.01:‘*“4“‘4444444441111... 50 100 150 200 250 300 Temperature (° C) Viscosi 11.4.4444 = are-081° + 413051“ + 0.0089 R2 = 0.9997 Figure 8.57 - Viscosity of steam as a function of temperature (From tabular data: Geankoplis, 1993). A mixture equation developed by Burrneister (1983) was utilized to calculate the viscosity of the air-steam mixture (Equation [8.4]). H air + k1 steam “mix = 1+¢,,-’%( 1+¢.-x%( where: [8.4] 190 . [44444-4 (44 4/ J l and . 4444,1444 /.,,..°'”ll 4 4 44w The density of air was modeled as a function temperature using regression of tabular data (Geankoplis, 1983); (Equation [87]). pair = -0.00000000002 ~ T3 + 0.00000001-T2 — 0.000004 . T + 0.0013 [ 8.7 ] The regression equation had an R2 value of 1 (Figure 8.58). 191 1- t _ 3 - - - 0.0005 pa1,—-2E—11T +11+3418T2 41+: 06T+0,()o13 R2=1 0.0003 ruralwrrUllelHL 0 50 100 150 200 250 300 Temperature (°C) Figure 8.58 - Density of air as a function of temperature (from tabular data: Geankoplis, 1993). The density of water vapor at temperatures below 100°C was modeled as a function of temperature using tabular data (Geankoplis, 1993); (Equation [8.8]). P vapor = 0.00000000003 T“ + 0.00000000025 - T3 - 0.0000000039 - T2 + 0.00000085 . T [8.8] The regression equation had an R2 value of 1 (Figure 8.59). 192 0.0007 _ 3 2 A 0.0006 A psteam— 8.12E-10T - 3.52E-08T +1.32E-06T+ "e 6.22E-07 3;; 0.0004 5 Q 0.0003 a O u: '5 0.0002 + 0 £‘ 0.0001 — 0 . 1 . . 0 20 40 60 80 100 120 Temperature (° C) Figure 8.59 - Density of saturated steam as a function of temperature (From tabular data: Geankoplis, 1993). The density of steam at atmospheric pressure and temperatures above 100°C was modeled as a function of temperature using regression of tabular data (Geankoplis, 1993); (Equation [8.9]). = —0.0002 - log(T)+ 0.0015 [ 8.9 ] p steam The regression equation had an R2 value of 0.999 (Figure 8.60). 193 0.0007 MA 0.0006 * E g 0.0005 .5 0.0004 5 0 0.0003 1 a psteam = -0.0002Ln(T) + 0.0015 m 0.0001 1 O l l l 90 140 190 240 Temperature (°C) Figure 8.60 - Density of steam at 101.35 kPa as a function of temperature (From tabular data: Geankoplis, 1993). The density of the air-steam mixture was assumed given by Equation [8.10] which assumes that the molar fraction of each component is equal to the volume fraction. pmix = pair .Xair + psteam . Xstcam I: 810] The heat capacity of the air-steam mixture was calculated as functions of temperature and composition using an equation given by Millsap (2002); (Equation [8.11]). . : (pair 'Xair 'cpair + psteam .xsteam .cpstcam) [ 811 ] P-m'x (pair . Xair + psteam . xstcam) C 194 The thermal conductivity of air was modeled as a function of temperature using regression of tabular data (Geankoplis, 1993); (Equation [8.12]). kair = 0.000000003 - T2 + 0.0000008 - T + 0.002 [ 8.12 ] The regression had an R2 value of 0.9999 (Figure 8.61). Thermal conductivity (W/cm K) 2 1 k = -7E-09T3 - 513-0n2 + 0.0076T + 2.4187 1'5 R2 = 0.9999 1’: 0.5: 0 UL“.11111.11.511............. O 50 100 150 200 250 300 Temperature (° C) Figure 8.61 - Thermal conductivity of air as a function of temperature (From tabular data: Geankoplis, 1993). The thermal conductivity of steam was modeled as a function of temperature using regression of tabular data (Geankoplis, 1993); (Equation [813]). 195 k = 0.0000009 - T + 0.0002 [ 8.13 ] steam The regression had an R2 value of (Figure 8.62). A 0.00045 ' M 3 0.0004 5 E) E 0.00035 8 +3 5 0.0003 - “7.: km, = 9E-07T + 0.0002 -: i-' 0.0002 . 1 1 T 50 100 150 200 250 300 Temperature (° C) Figure 8.62 - Thermal conductivity of steam as a function of temperature (From tabular data: Geankoplis, 1993). Thermal conductivity of the air-steam mix was modeled using an equation given by Burmeister (1983); (Equation [8.14]). k p = kair + ksteam [ 814] mm Xstcam xair 1+Aas.( Aair) 1+Asa ( /Xstcam) 196 where: Aas =-1-< 4 and: r A33. :11 1+ [ “steam 1+[ p air steam l’l steam 1 M air air ’1 air )4 M steam 1 0.75 1 + 8% air.K Sscam 1+ ! Aairx 0.75 l + Ss‘“%airx Ssair 1 + flairx 0.5 ‘ . 1 + [0.733 - (S... 6...... )°" l/TM 0.5 ‘ Sair 1+ fiairx [ 8.15] > . 1 + [0.733 . (sWm . 3.1, )0-51/TM 3. 1+ m Tair.K [ 8.16] The diffusivity of steam in air was modeled using an equation from Vargaflik (1966); (Equation [8. 17]). Dw=02 T 16.[ air,K 273 I. 197 [ 8.17] 8.3 Product thermo-physical properties Composition dependent values of product thermo-physical properties were used for the model calculations (Equations [8.18] and [8.19]). The heat capacity of meat was modeled as a function of the mass fractions of water, protein, and fat (Choi and Okos, 1986). Separate equations were utilized for the temperature ranges below (Equation [8.18]) and above (Equation [8.19]) freezing. Cp.frozcn = Cp.ice .xW +cp.protein .XP +cp.fat .XF [ 8'18] c = c -XW +Cp.protein -XP +cp‘fat ~XF [ 8.19] Thermal conductivity of the meat was calculated using a series model based upon the volume fraction and thermal conductivity of the water, protein, and meat fractions of the meat (Choi and Okos, 1986); (Equation [820]). kT=k ~X +k ~X +k X [8.20] water water protein protein fat . fat The density of the meat was calculated using a parallel model based upon the density of the water, fat, and protein components of the meat (Choi and Okos, 1986); (Equation [8.21]). 198 1 mwater/p water + mfat /p fat + mprotein /pprotein p: [8.21] The moisture diffusivity of the meat was modeled as a function of fat-protein ratio and temperature using an equation developed by Mittal and Blaisdell (1984); (Equation [822]). 4829.7 km = 0.003~exp(— 0.442+FP— +1155] [ 8.22] A constant value of 64.4 J/g was utilized for the latent heat a fusion of fat (Skala etal., 1989). The latent heat of fusion for water was set at 337.8 J/g (Geankoplis, 1993). A value of 0.003 g/g was utilized for the value of moisture capacity, cm (Chen et al., 1999). A constant value of 0.0008 g/s'cm2 was utilized for the fat conductivity, kf. 199 8.4 Screen shots from Visual Basic cooking model user interface kawmymm MSU Biosystems Engineering Convection Cooking Model Oven Conditions 0 Input Steady Conditions Oven Conditions Ewen Temperature [€039] '1) Owen Steam Comer” 1'13 100% h, xolun‘vel a - lemrev ah ne I2 I 0R I)ve" \Nel Bulb Tempélalute [C] u A11 Flow R11»: (U 530 W's-,1 Tempermurer 1 Cooking T1me (11; . a) 0 input Translenl Conditions From File Product Salmonella Kinetics Product "1““ Conditions -— O Uh? Orr-.4311 Model F’avarrwii—«S _ - . Input Model F‘Jvamelevs o 5.354 Temperature (259:4 I.) u . pmk Moisture Conteni 0.135%) a ‘ III » Value a Reference . . Temperature 0 Tuvkey Fat Content (075%) 1 ' Value Figure 8.63 — Input screen of cooking model user interface. 200 He Mm 01w Geometry Heb [that MSU Biosystems Engineering Convection Cooking Model Temperature Profile Yield Iloisture Content 1111— 11 ’ Mustang (Lumen: Fusl 1.11434'0l9rhp'1371l1" ‘ hm] genial 'Qfl'tCIx‘lalgle r41: 4 L E ‘+ Log ledm’lfin no! 311191.941] ‘Mwmun L Stunted log veduchm - i ”6:121:73:~ Figure 8.64 — Output screen of cooking model user interface. 201 8.5 Visual Basic model code Module — Main Version 1.0 - July 2004 By Adam E. Watkins and Dr. Bradley P. Marks Department of Biosystems and Agricultural Engineering Michigan State University Sub Main() 'Many of the modules associated with the basic FEM architecture were provided by or modified ' from modules provided by Dr. Larry Segerlind - Michigan State University. These modules ' have been identified in the module comment statements. ‘ FEM cooking and microbial inactivation model - Main Module ' This module controls the calculations for a two-dimensional heat and mass transfer problem. 'Open the main input file. Contains node data, element data, boundary nodes, etc. Open "c:\Tdeield47.txt" For Input As #1 'Open the main output file Open "c:\modeloutput.txt" For Output As #3 'Input setup information Input #1, NumNodes, NumEle, NudeyNodes, NumMatlSets, NumDeridey 'Set solution timestep to l s timestep = 1 'Format display Form2.MSChart1.Visible = True 'Show chart 1 Form2.MSChart2.Visible = True 'Show chart 2 Form2.MSChart3.Visible = True 'Show chart 3 'Input oven conditions from a file or from the input screen If Forml .Option2 = True Then Open "c:\OvenConditions.txt" For Input As #2 'Open oven conditions input file Input #2, NumTimeSteps 'Get cooking time from file ReDim OvenConditions(NumTimeSteps, 3) 'Dimension oven conditions matrix Call InputOvenConditions(OvenConditions(), NumTimeSteps) 'Get oven conditions from file Forrnl .Frame3.Visible = True 'Show oven conditions graph 202 Forml.MSChart1.Visible = True OvenConditions(O, 1) = ”Temperature C" OvenConditions(O, 2) = "Steam Content” OvenConditions(O, 3) = "Air Velocity m/s" Forml.MSChart1.ChartData = OvenConditions 'Close oven conditions file 'Oven conditions from form Close #2 Else NumTimeSteps = Forml .Text4.Text ReDim OvenConditions(NumTimeSteps, 3) For I = 1 To NumTimeSteps OvenConditions(I, l) = Forml .Textl .Text OvenConditions(I, 2) = Forml .Text2.Text 'Show oven conditions graph 'Labels on graph 'Labels on graph 'Labels on graph 'Plot oven conditions on graph 'Input cooking time 'Dimension oven conditions matrix 'Oven temperature C 'Oven steam content % by volume If OvenConditions(I, 2) = 0 Then OvenConditions(I, 2) = 1 OvenConditions(I, 3) = For‘ml .Text3.Text Next I Forml .Frame3.Visible = True Forml .MSChartl .Visible = True OvenConditions(O, l) = "Temperature C" OvenConditions(O, 2) = "Steam Content" OvenConditions(O, 3) = "Air Velocity m/s" Forml.MSChart1 .ChartData = OvenConditions End If ReDim Coord(NumNodes, 2) ReDim EleMatlData(NumEle) ReDim EleNodeData(NumEle, 4) ReDim SubDerivBC(NumDefidey + l, 2) ReDim BoundaryType(NumDeridey + 1) ReDim deNode(NudeyNodes) 'Input node,element, and derivative boundary data 'Oven air velocity cm/s 'Show oven conditions graph 'Show oven conditions graph 'Labels on graph 'Labels on graph 'Labels on graph 'Plot oven conditions on graph 'Dimension matrix of coordinate data ' Coord( ,1) X coordinate ' Coord( ,2) Y coordinate 'Dimension element material set 'Dimension matrix of element node data ' EleNodeData%( ,l) = Node I ' EleNodeData%( ,2) = Node J ' EleNodeData%( ,3) = Node K ' EleNodeData%( ,4) = Node M 'Subscripts of the derivative ' boundary condition 'Convection or impingement at each ' derivative boundary 'Vector of boundary nodes Call InputBasicData(NumMatlSets, NumNodes, NumEle, Coord(), EleMatlDataO, EleNodeDataO) Call InputDerivBC(NumDeridey, NudeyNodes, deNode(), SubDerivBC(), BoundaryTypeO) Close #1 'Close input data file 'Input type of product If Forml .Option3 = True Then Meat = 1 'Beef TR = 6.5 'Target reduction Form2.Labe12.Caption = "6.5" Elself Forml .Option4 = True Then 'Label on Salmonella reduction graph 203 Meat = 2 'Pork TR = 6.5 'Target reduction Form2.Label2.Caption = "6.5" 'Label on Salmonella reduction graph Elself Forml .OptionS = True Then Meat = 3 ' Turkey TR = 7 'Target reduction Forrn2.Labe12.Caption = "7" 'Label on Salmonella reduction graph End If 'Input meat temperature, moisture, and fat content InitialT = Forml .Text5.Text 'Initial meat temperature InitialM = Forml .Text6.Text 'lnitial meat moisture content If InitialM = 0 Then InitialM = 0.1 'Avoid zero moisture condition not allowed ' by model equations InitialF = Forml .Text7.Text 'Initial meat fat content 'Input Salmonella reduction equation coefficients Dvalue = Form1.Text9.Text 'D-value Salmonella Z = Forml .Text10.Text 'z-value Salmonella Tref = Forml.Text11.Text ’T-ref Salmonella 'Dimension vectors for calculation of inactivation ReDim AverageN(NumTimeSteps) 'Volume average of microbial concentration Dim Reduction As Variant 'Overall microbial reduction ReDim logreduction(NumNodes) 'Center microbial reduction Ninitial = 1000000000 'Value used for log-reduction calculation ReDim No(NumN odes) 'Used for log—reduction calculation ReDim NW(NumNodes) 'Used for log-reduction calculation ReDim Nnew(NumNodes) 'Used forlog-reduction calculation ReDim NnewW(NumNodes) 'Used for log-reduction calculation For X = 1 To NumNodes No(X) = Ninitial 'Create a nodal microbial concentration vector NW(X) = Ninitial Next X 'Dimension global solution vector ReDim GSV(NumNodes) 'Enthalpy ReDim GSV_M(NumNodes) 'Moisture ReDim GSV_F (NumN odes) 'Fat 'Vector of temperature values ReDim temperature(NumNodes) Temperature vector 'Calculate solid content of the product Solid = 100 - lnitialM - InitialF 'Initial solid material in meat (NOT fat or water) 'Load vectors of moisture and F? values 204 For K = 1 To NumNodes GSV_M(K) = InitialM 'Load vector of nodal moisture values GSV_F(K) = InitialF / Solid 'Load vector of nodal fat values temperature(K) = InitialT 'Load vector of initial temperature values Next K ’Calculate the initial therrnophysical properties of the product Call CalculatelnitialProperties(heat_capacity_f, Meat, Solid, InitialT, InitialM, InitialF, heat_capacity, latentW, frozenH, latentF) 'Formulate enthalpy table for reconversion to temperature LevelOne = frozenI-I LevelTwo = frozenI-I + latentW LevelFive = frozenH + latentW + (2 "' 3.35) LevelThree = frozenH + latentW + (45 * 3.35) + (2 * 3.35) LevelFour = frozenl-I + latentW + (45 * 3.35) + latentF + (2 * 3.45) 'Percent of water thawed at the exact value of freezing temperature PercentThawed = 1 'Convert product temperature to enthalpy If InitialT < -2 Then For K = 1 To NumNodes GSV(K) = (InitialT + 273) " heat_capacity_f Next K Elself InitialT = -2 Then For K = 1 To NumNodes GSV(K) = frozenH + PercentThawed * latentW Next K Elself InitialT > -2 And lnitialT <= 45 Then For K = 1 To NumNodes GSV(K) = ((2 + InitialT) " 3.35) + (frozenH + latentW) Next K Elself InitialT > 45 Then For K = I To NumNodes GSV(K) = ((2 + InitialT) "' 3.35) + (frozenH + latentW + latentF) Next K End If 'Calculate oven air moisture content from wet bulb temperature if specified in that manner If Form1.Checkl.Value = 1 Then wa = Forml .Text8.Text Tov = Forrn1.Text1.Text latent = (-2.429 "‘ Forml.Text1.Text + 2502.8) Cwb = 8.12078904E-10 "' wa " 3 - 0.000000035203247 "' wa " 2 + 0.00000131977474 "' wa + 0.000000621530732 ConcOvenAir = Cwb - ((0.000731 / latent) "' (Tov - wa)) End If 205 'Dimension output matrices ReDim Output(NumTimeSteps, 2) Temperature output ReDim Output_S(NumTimeSteps, 3) 'Salmonella linear lethality output ReDim Output_SW(NumTimeSteps, 2) 'Salmonella logistic lethality output ReDim AverageOutput(NumTimeSteps, 2) 'Yield and moisture output ReDim AverageFat(NumTimeSteps) 'Element average fat content 'Dimension matrix of Element Physical Data ReDim ElePhyData(NumMatlSets, 15) 'ElePhyData( ,1) Equation coef, Dx 'ElePhyData( ,2) Equation coef, Dy 'ElePhyData( ,3) Equation coef, G 'ElePhyData( ,4) Equation coef, Q 'ElePhyData( ,5) Equation coef, lamda 'ElePhyData( ,6) Equation coef, Dx_M 'ElePhyData( ,7) Equation coef, Dy_M 'ElePhyData( ,8) Equation coef, G_M 'ElePhyData( ,9) Equation coef, Q_M 'ElePhyData( ,10) Equation coef, lamda_M 'ElePhyData( ,l 1) Equation coef, Dx_F 'ElePhyData( ,12) Equation coef, Dy_F 'ElePhyData( ,13) Equation coef, G_F 'ElePhyData( ,14) Equation coef, Q_F 'ElePhyData( ,15) Equation coef, lamda_F 'Calculate Bandwidth Call Cachandwidth(NumEle, EleNodeDataO, BandWidth) 'Dimension element matrices 'Element force vector Dim EFQ(4) 'Enthalpy Dim EFQ_M(4) 'Moisture Dim EFQ_F(4) 'Fat 'Element force vector, derivative BC Dim EFS(2) 'Enthalpy Dim EFS_M(2) 'Moisture Dim EFS_F(2) 'Fat 'Element stiffness matrix Dim ESM(4, 4) 'Enthalpy Dim ESM_M(4, 4) 'Moisture Dim ESM_F(4, 4) 'Fat 'Element stiffness matrix, derivative BC Dim EKM(2, 2) 'Enthalpy Dim EKM_M(2, 2) 'Moisture Dim EKM_F (2, 2) 'Fat 'Element subscript values 206 Dim EleSub(4) 'Element Phi values (nodal unknown values) Dim ElePhiVal(4) 'Enthalpy Dim ElePhiVal_M(4) 'Moisture Dim ElePhiVal_F (4) 'Fat 'Element capacitance matrix Dim ECM(4, 4) 'Enthalpy Dim ECM_M(4, 4) 'Moisture Dim ECM_F(4, 4) 'F at 'Element capacitance matrix, derivative BC Dim ECQ(4, 4) 'Enthalpy Dim ECQ_M(4, 4) 'Moisture Dim ECQ_F (4, 4) 'F at 'Dimension the arrays for the global system of equations NumEleSub = 4 NumEleNodes = 4 'Vectors used in finite difference time solution Temporary solution vector ReDim Temp(NumNodes) 'Enthalpy ReDim Temp_M(N umN odes) 'Moisture ReDim Temp_F(NumNodes) 'F at 'Global force vector ReDim GFV(NumNodes) 'Enthalpy ReDim GFV_M(NumNodes) 'Moisture ReDim GF V_F(N umN odes) 'Fat ReDim GFV_A(NumNodes) 'Enthalpy ReDim GFV_M_A(NumNodes) 'Moisture ReDim GFV_F_A(NumNodes) 'Fat ReDim GF V_Star(NumN odes) 'Enthalpy ReDim GFV_M_Star(NumN odes) 'Moisture ReDim GFV_F_Star(NumNodes) 'Fat 'Global stiffness matrix ReDim GSM(NumNodes, BandWidth) 'Enthalpy ReDim GSM_M(NumNodes, BandWidth) 'Moisture ReDim GSM_F(NumNodes, BandWidth) 'Fat 'Global capacitance matrix 207 ReDim GCM(NumNodes, BandWidth) 'Enthalpy ReDim GCM_M(NumN odes, BandWidth) 'Moisture ReDim GCM_F(NumNodes, BandWidth) 'Fat 'Global A matrix used for FD time solution ReDim GAM(NumNodes, BandWidth) 'Enthalpy ReDim GAM_M(NumNodes, BandWidth) 'Moisture ReDim GAM_F(NumNodes, BandWidth) 'Fat 'Global P matrix used for FD time solution ReDim GPM(NumNodes, BandWidth) 'Enthalpy ReDim GPM_M(NumNodes, BandWidth) 'Moisture ReDim GPM_F(NumNodes, BandWidth) 'Fat 'Vector used for volume averaging ReDim Volume(NumEle) ' FEM / Finite difference time solution 'Zero the global matrices Call ZeroGlobalMatrices(NumNodes, BandWidth, GFV(), GSM(), GFV_M(), GSM_M(), GFV_F(), GSM_FO. GCMO. GCM_MO, GCM_F()) 'Show message indicating program is running Form6.Show 'Initialize the global matrices for the first time step of the time solution T = 1 'Build the banded system of equations For K = 1 To NumEle Call CalculateProperties(Meat, Solid, NumEle, EleNodeData(), E1ePhyData(), temperature(), GSV_M(), GSV_F(), moisture_capacity, density) Call ESMatrix2DField(EleMatlData(), EleNodeData(), Coord(), E1ePhyData(), EFQ(), ESM(), EFQ_M(), ESM_M(), EFQ_F(), ESM_F(), Volume(l. KK) Call EleCapMatrix(ElePhyData(), EleNodeData(), ECM(), ECM_M(), ECM_F(), Coord(), KK) Call EleSubscriptValues(EleNodeData(), KK, NumEleSub, EleSub()) Call BuildBandedSystem(NumEleSub, EleSub(), EFQ(), ESM(), GFV(), GSM(), ECM(), GCMO, EFQ_M(), ESM_M(), GFV_M(), GSM_M(), ECM_M(), (ECM_M(), EFQ_F(), ESM_F(), GFV_F(), GSM_F(), ECM_F(), GCM_F()) Next K 'Add in the derivative boundary conditions when they occur - direct stiffness method 208 NumEleSub = 2 For I = 1 To NumDeridey Call CalDerivBC(heat_capacity_f, frozenH, latentW, latentF, I, Coord(), SubDerivBC(), EleSub(), EKM(), EFS(), GSVO, EKM_MO, EFS_MO, GSV_M(), EKM_F(), EFS_FO, GSV_F(), OvenConditions(), T, BoundaryTypeO, InitialM, InitialF, moisture_capacity, density, temperature(), ConcOvenAir) Call BuildBandedSystem(NumEleSub, EleSub(), EFS(), EKM(), GFV(), GSM(), ECQ(), GCM(), EFS_MO, EKM_MO. GFV_M(), GSM_M(), ECQ_MO. GCM_M(), EFS_FO, EKM_F(), GFV_F(), GSM_F(), ECQ_FO, GCM_F()) Next I 'Set force vector values for next time step For X = 1 To NumNodes GFV_A(X) = GFV(X) GFV_M_A(X) = GFV_M(X) GFV_F_A(X) = GFV_F(X) Next X 'Begin time stepping solution For T = 1 To NumTimeSteps 'Construct global matrices using direct stiffness method For K = 1 To NumEle Call CalculateProperties(Meat, Solid, NumEle, EleNodeData(), E1ePhyData(), temperature(), GSV_M(), GSV_F(), moisture_capacity, density) Call ESMatrix2DField(EleMatlData(), EleNodeData(), Coord(), E1ePhyData(), EFQ(), ESM(), EFQ_M(), ESM_M(), EFQFO, ESM_F(), Volume(), KK) Call EleCapMatrix(ElePhyData(), EleNodeData(), ECM(), ECM_M(), ECM_F(), Coord(), KK) Call EleSubscriptValues(EleNodeData(), KK, NumEleSub, EleSub()) Call BuildBandedSystem(NumEleSub, EleSub(), EFQ(), ESM(), GFV(), GSM(), ECM(), GCM(), EFQ_M(), ESM_M(), GFV_M(), GSM_M(), ECM_M(), GCM_M(), EFQFO, ESM_F(), GFV_F(), GSM_F(), ECM_F(), GCM_F()) Next K 'Add in the derivative boundary conditions when they occur - direct stiffness method NumEleSub = 2 For I = 1 To NumDeridey Call CalDerivBC(heat_capacity_f, frozenH, latentW, latentF, I, Coord(), SubDerivBC(), 209 EleSub(), EKM(), EFS(), GSV(), EKM_MO, EFS_M(), GSV_M(), EKM_F(), EFS_F(), GSV_F(), OvenConditions(), T, BoundaryTypeO, InitialM, InitialF, moisture_capacity, density, temperature(), ConcOvenAir) Call BuildBandedSystem(NumEleSub, EleSub(), EFS(), EKM(), GFV(), GSM(), GCM(), EFS_M(), EKM_MO, GFV_M(), GSM_M(), ECQ_M(), ' GCM_M(), EFS_F(), EKM_F(), GFV_F(), GSM_F(), ECQ_FO, GCM_F()) Next I 'Build matrices for finite difference (time) solution Call BuildGAM(BandWidth, GAM(), GCM(), GSM(), GAM_MO, GCM_M(), GSM_M(), GAM_F(), GCM_F(), GSM_F(), timestep, NumNodes) Call BuildGPM(BandWidth, GPM(), GCM(), GSM(), GPM_M(), GCM_M(), GSM_M(), GPM_F(), GCM_F(), GSM_F(), timestep, NumNodes) Call ModifyGFV(GFV(), GFV_A(), GFV_StarO, GFV_M(), GFV_M_AO, GFV_M_StarO, GFV_F_A(), GFV_F_StarO, GFV_F(), timestep, NumNodes) 'Solution of equations Call MultpyBandMatrix(NumNodes, BandWidth, GPM(), GSV(), Temp()) Call MultpyBandMatrix(NumNodes, BandWidth, GPM_M(), GSV_M(), Temp_M()) Call MultpyBandMatrix(NumNodes, BandWidth, GPM_F(), GSV_F(), Temp_F()) For I = 1 To NumNodes Temp(l) = Temp(I) + GFV_Star(l) Temp_M(I) = Temp_M(l) + GFV_M_Star(I) Temp_F (I) = Temp_F(I) + GFV_F_Star(I) GFV_A(I) = GFV(I) GFV_M_A(I) = GFV_M(I) GFV_F_A(I) = GFV_F(I) Next I Call DecompBandMatrix(NumNodes, BandWidth, GAM()) Call DecompBandMatrix(NumNodes, BandWidth, GAM_MO) Call DecompBandMatrix(NumNodes, BandWidth, GAM_F()) Call SolveBandMatrix(NumNodes, BandWidth, GSV(), Temp(), GAM()) Call SolveBandMatrix(NumNodes, BandWidth, GSV_M(), Temp_M(), GAM_MO) Call SolveBandMatrix(NumNodes, BandWidth, GSV_F(), Temp_F(), GAM_F()) 'Convert enthalpy data back to temperature for display Call ConverthoTemp(heat_capacity__f, GSV(), GSV_M, GSV_F, temperature(), NumNodes, InitialM, Solid, frozenH, latentW, latentF, heat_capacity, LevelOne, LevelTwo, LevelThree, LevelF our, LevelFive) 'Calculate surface fat for next time step Call CalculateSurfaceFatContent(InitialF, T, GSV_F(), NudeyNodes, deNodeO, temperature(), T) 'Calculate Salmonella reduction Call CalculateSurvivors(NumNodes, timestep, temperature(), Ninitial, No(), NW(), Nnew(), 210 NnewW(), logreduction(), logreductionW, Dvalue, Tref, Z, TimeToLimit, TR) 'Volume averaging of Salmonella reduction Call CalculateElementAverage(Volume(), NumEle, EleNodeData(), Nnew(), WeightedAverage) AverageN(T) = WeightedAverage Reduction = -Log(AverageN(T) / Ninitial) / Log(10) 'Calculate the average moisture and FF of each element for overall yield and bulk moisture determination 'Moisture Call CalculateElementAverage(Volume(), NumEle, EleNodeData(), GSV_M(), WeightedAverage) 'F at Call CalculateElementAverage(Volume(), NumEle, EleNodeData(), GSV_F(), WeightedAverage) AverageFat(T) = WeightedAverage F = AverageFat(T) 'Calculate yield from moisture and fat percentages Fat = F " Solid ‘Mass fat based on 100 g initial Water = M * (Fat + Solid) / (1 - M) 'Mass water based on 100 g initial Total = Fat + Water + Solid Total mass ' OUTPUT SECTION 'Graph yield AverageOutput(T, 2) = Total 'Yield AverageOutput(O, 2) = "Yield" 'Graph label 'Graph moisture AverageOutput(T, 1) = WeightedAverage 'Moisture AverageOutput(O, 1) = "M" 'Graph label M = AverageOutput(T, 1)/ 100 'Graph temperature output Output(O, l) = "Center Temp" 'Graph labels Output(O, 2) = "Surface Temp" 'Graph labels Output(T, l) =temperature(1) 'Center temperature Output(T, 2) = temperature(4) 'Surface temperature 'Graph Salmonella reduction Output_S(0, l) = "Log reduction" 'Graph label 211 Next T Output_S(O, 2) = "Target reduction" 'Graph label Output_S(T, l) = logreduction(l) 'Graph Salmonella reduction from linear eqn. Output_S(T, 2) = TR 'Graph target reduction Output_S(T, 3) = logreductionW ‘Graph Salmonella reduction from logistic eqn. 'Write data to output file Write #3, temperature(l), temperature(Z), temperature(4), Total, M, logreduction( 1 ), logreductionW, Reduction 'Labels for Salmonella reduction graph If logreduction(l) < TR Then Form2.Labe13.Caption = "Log reduction not acheived!" Form2.Label4.Visible = False F onn2.Labe15.Visible = False Else F orm2.Label3.Caption = "Log reduction acheived atz" Form2.Label4.Visible = True Form2.Labe15.Visible = True Form2.Label4.Caption = TimeToLimit Form2.LabelS.Caption = "s" End If 'End time stepping 'Digital display of endpoint data Form2.Label8.Caption = Round(temperature(4), 1) 'Display surface temperature F orm2.Label9.Caption = Round(temperature( 1 ), 1) 'Display center temperature F orm2.Labe112.Caption = Round(Total, 1) 'Display yield Form2.Labell3.Caption = Round((M "‘ 100), 1) 'Display moisture content 'Remove run message box F orm6.Hide 'Display graphs Form2.MSChartl .ChartData = Output Temperature graph F orrn2.MSChart2.ChartData = AverageOutput 'Yield/Moisture graph Form2.MSChart3.ChartData = Output_S 'Salmonella graph Close #3 End Sub 212 Module — CacherivBC Sub CalDerivBC(heat_capacity_f, frozenH, latentW, latentF, I, Coord(), SubDerivBC(), EleSub(), EKM(), EFS(), GSV(), EKM_MO. EFS_M(), GSV_M(), EKM_F(), EFS_F(), GSV_F(), OvenConditions(), T, BoundaryTypeO, InitialM, InitialF, moisture_capacity, density, temperature(), ConcOvenAir) This subroutine calculates the element contributions resulting from 'derivative boundary conditions 'Assign the node numbers of the side to the array EleSub() EleSub(l) = SubDerivBC(I, 1) EleSub(Z) = SubDerivBC(I, 2) 'Grab oven conditions OvenAirT = OvenConditions(T, 1) OvenAirM = OvenConditions(T, 2) OvenAirV = OvenConditions(T, 3) "' 100 'Evaluate the coefficients in the element matrices XLength = Coord(SubDerivBC(I, 1), l) - Coord(SubDerivBC(I, 2), l) YLength = Coord(SubDerivBC(I, l), 2) - Coord(SubDerivBC(I, 2), 2) Ri = Coord(SubDerivBC(I, l), l) Rj = Coord(SubDerivBC(I, 2), 1) SideLength = Sqr(XLength " 2 + YLength A 2) 'Get temperature and moisture for each boundary ' T1 = temperature at node 1 of boundary ' T2 = temperature at node 2 of boundary T1 = l "' temperature(SubDerivBC(I, 1)) T2 = l * temperature(SubDerivBC(I, 2)) Tave = (Tl + T2) / 2 ' M1 = moisture at node 1 of boundary (decimal) ' M2 = moisture at node 2 of boundary (decimal) M] = GSV_M(SubDerivBC(I, 1)) M2 = GSV_M(SubDerivBC(I, 2)) Mave=(M1 ’1+M2*1)/2 If Mave < 0 Then Mave = 0.01 'Convert moisture content to dry basis MDB = (100 "‘ Mave) / (100 - Mave) 213 'Oven geometry Height = 6.35 'cm S = 4.6482 'cm E = 1.17 'cm W = 0.635 'cm F = W / (2 "‘ S) 'Martin Equation parameter fo = (60 + 4 * (Height / W - 2) A 2) A -0.5 'Martin Equation parameter 'Calculate air physical properties as functions of temperature and steam content 'Steam concentration g/cmA3 If OvenAirT <= 100 Then DensitySatSteam = 8.12078904E-10 “ OvenAirT A 3 - 0.000000035203247 * OvenAirT A 2 + 0.00000131977474 * OvenAirT + 0.000000621530732 Elself OvenAirT > 100 Then DensitySatSteam = -0.0002 * Log(OvenAirT) + 0.0015 'g/cmA3 End If If Forml.Checkl.Value = 1 Then OvenAirM = ConcOvenAir density_air = -0.00000000002 * OvenAirT A 3 + 0.00000001 "‘ OvenAirT A 2 - 0.000004 "‘ OvenAirT + 0.0013 'g/cmA3 X_S = OvenAirM / (OvenAirM + density_air) Else X_S = OvenAirM/ 100 End If 'Mass fraction air X_A = 1 - X_S 'Molar weight of steam and air M_steam = 18 M_air = 28 S_a = 79 'K S_s = 559.5 'K 'Latent heat latent = 1.2 * (-2.429 * Tave + 2502.8) 'J/g 'Air and steam viscosity viscosity_air = -0.0000000002 "‘ OvenAirT A 2 + 0.0000005 "‘ OvenAirT + 0.0002 'g/cm s viscosity_steam = -0.00000001 "’ OvenAirT A 2 + 0.00004 * OvenAirT + 0.0089 'g/cm 3 phi_as = 1 / 8 A 0.5 * ((1 + (viscosity_air / viscosity_steam) A 0.5 * (M_steam / M_air) A 0.25) A 2 / (1 + (M_air / M_steam)) A 0.5) '- 214 phi_sa = l / 8 A 0.5 * ((1 + (viscosity_steam / viscosity_air) A 0.5 * (M_air / M_steam) A 0.25) A 2 / (l + (M_steam / M_air)) A 0.5) '- ’Mixture viscosity viscosity_mix = viscosity_air / (1 + phi_as * X_S / X_A) + viscosity_steam / (1 + phi_sa * X_A/ X_S) 'g/cm s 'Air and mixture density density_air = 000000000002 * OvenAirT A 3 + 0.0000000] * OvenAirT A 2 - 0.000004 * OvenAirT + 0.0013 'g/cmA3 density_mix = density_air "' X_A + DensitySatSteam * X_S 'g/cmA3 'Heat capacity cpair = 1.01 cpsteam = 0.00000087429 * OvenAirT A 2 + 0.00018055 * OvenAirT + 1.8616 '1.888 '4.2 cpmix = (density_air * X_A "' cpair + DensitySatSteam * X_S * cpsteam) / (density_air "‘ X_A + DensitySatSteam * X_S) 'Diffusivity Dsa = (2.16 * 10 A -5 * ((OvenAirT + 2731/273) A 1.8) * 10000 'cmA2/s A_as = 0.25 "' (1 + ((viscosity_air * viscosity_steam) "‘ (M_steam / M_air) A 0.75 + (1 + S_a/ (OvenAirT + 273)) / (l + S_s / (OvenAirT + 273))) A 0.5) A 2 "' (1 + 0.733 * (S_a "‘ S_s) A 0.5 / (OvenAirT + 273)) / (l + S_a / (OvenAirT + 273)) '- A_sa = 0.25 * (1 + ((viscosity_steam / viscosity_air) "' (M_air / M_steam) A 0.75 * (1 + S_s/ (OvenAirT + 273)) / (1 + S_a / (OvenAirT + 273))) A 0.5) A 2 * (l + 0.733 "‘ (S_a " S_s) A 0.5 / (OvenAirT + 273)) / (1 + S_s / (OvenAirT + 273)) '- 'Thermal conductivity k_air = 00000000003 "' OvenAirT A 2 + 0.0000008 "' OvenAirT + 0.002 'W/cm C k_steam = 0.000000009 * OvenAirT + 0.0002 'W/cm C k_mix = (k_air / (1 + A_as " (X_S / X_A)) + k_steam / (1 + A_sa "' (X_A / X_S))) 'W/cm C 'Calculate Prandtl and Schmidt numbers Pr_mix = cpmix * viscosity_mix / k_mix - Sc_mix = viscosity_mix / (density_mix "' Dsa) '- 'Calculate oven steam moisture concentration If Forml .Check1.Value = 1 Then ConcOvenAir = ConcOvenAir Else 215 ConcOvenAir = DensitySatSteam * OvenAirM / 100 'g/cmA3 End If 'Calculate model parameters 'Convection condition If BoundaryType(l) = 1 Then 'Calculate transport coefi'rcients Re = OvenAirV "' density_mix "' E / viscosity_mix Nu = 0.037 " Re A 0.8 " Pr_mix A (l /3) SH = 0.037 "' Re A 0.8 * Sc_mix A (1 / 3) H_M=SH"'Dsa/E h=Nu*k_mix/E 'Calculate surface moisture concentration If Tave < 0 Then Csat = 0.000000621530732 Else Csat = 8.12078904E-10 "' Tave A 3 - 0.000000035203247 "' Tave A 2 + 0.00000131977474 * Tave + 0.000000621530732 End If 'Calculate effective relative humidity RH = 100 "‘ (ConcOvenAir / Csat) MDB = (Mave * 100) / (100 - Mave) ERH = Exp((-5222.47 “ (MDB A -1.0983)) / (1.9818 * (Tave + 273))) Cs = ERH * Csat 'Calculate energy fluxes Q_conv = h * (OvenAirT - Tave) If ConcOvenAir > Csat Then Q_cond = H_M * latent "‘ (ConcOvenAir - Cs) Q_evap = 0 Else Q_cond = 0 End If If Cs > ConcOvenAir Then Q_evap = H_M * latent * (Cs - ConcOvenAir) Q_cond = 0 Else Q_evap = 0 Endlf 216 Q_total = Qficonv + Q_cond - Q_evap 'Effective heat transfer coefficient based on total flux (convection, condensing, evaporation) heff = Q_total / (OvenAirT - Tave) 'Calculate equilibrium values under current conditions 'Enthalpy Hequilibrium = OvenAirT * 3.35 + frozenH + latentW + latentF + (2 * 3.35) 'Moisture If RH >= 97 Then EMCWB = InitialM Else EMC = ((1.9818 ‘ (Tave + 273) "' Log(RH/ 100)) / ~5222.47) A (l / -l.0983) EMCWB = (100 * EMC) / (100 + EMC) End If X = 1 "' InitialM If X <= (EMCWB * 1) Then EMCWB = X End If 'FEM boundary coefficients 'Heat transfer CoetM = heff "‘ SideLength / 6 CoefS = heff "' Hequilibrium "‘ SideLength / 2 'Moisture transfer CoetM_M = H_M "' density " moisture_capacity " SideLength / 6 'Moisture transfer CoefS_M = H_M * density "' moisture_capacity * EMCWB “ SideLength / 2 'Stein JSO-IV impingement condition Elself BoundaryType(I) = 2 Then 'Impingement 'Calculate transport coefficients Re = OvenAirV "‘ W "' density_mix / viscosity_mix Nu=(2/3)*Pr_mixA0.42* foA0.75 ‘((2"'Re)/(F/fo+fo/F))A(2/3) SI-I=(2/3)* Sc_mixA0.42 * foA0.75 * ((2 * Re)/(F/fo+fo/F))A(2/3) H_M=SH"'Dsa/W h=Nu"'k_mix/W 'Calculate surface moisture concentration If Tave < 0 Then Csat = 0.000000621530732 217 Else Csat = 8.12078904E-10 " Tave A 3 - 0.000000035203247 * Tave A 2 + 0.00000131977474 * Tave + 0.000000621530732 End If 'Calculate effective relative humidity RH = 100 * (ConcOvenAir / Csat) MDB = (Mave * 100) / (100 - Mave) ERH = Exp((-5222.47 * (MDB A -1.0983)) / (1.9818 * (Tave + 273))) Cs = ERH "' Csat 'Calculate energy fluxes Q_conv = h " (OvenAirT - Tave) If ConcOvenAir > Cs Then (Lcond = H_M * latent * (ConcOvenAir - Cs) Q_evap = 0 Else Q_cond = 0 End If If Cs > ConcOvenAir Then (Levap = H_M “ latent * (Cs - ConcOvenAir) Q_cond = 0 Else Q_evap = 0 End If Q_total = (Lconv + Q_cond - Q_evap 'Effective heat transfer coefficient based on total flux (convection, condensing, evaporation) heff = Qtotal / (OvenAirT - Tave) 'Calculate equilibrium values under current conditions 'Enthalpy Hequilibrium = OvenAirT "‘ 3.35 + frozenH + latentW + latentF + (2 * 3.35) '792.64 + 6.44 'Moisture If RH >= 97 Then EMCWB =Initia1M Else EMC = ((1.9818 "‘ (Tave + 273) "' Log(RH/ 100)) / -5222.47) A (1 /-1.0983) EMCWB = (100 * EMC) / (100 + EMC) End If X = 1 "' InitialM IfX <= (EMCWB * 1) Then EMCWB = x 218 End If 'FEM boundary coefficients 'Heat transfer CoefM = heff "' SideLength / 6 CoefS = heff "‘ Hequilibrium "' SideLength / 2 'Moisture transfer CoetM__M = H_M "' density * moisture_capacity * SideLength / 6 'Moisture transfer CoetS_M = H_M * density * moisture_capacity * EMCWB * SideLength / 2 Elself BoundaryType(I) = 3 Then 'Small oven W = 2.6 'Calculate transport coefficients Re = OvenAirV * density_mix * W / viscosity_mix Nu = 0.037 "' Re A 0.8 * Pr_mix A (1 /3) SH = 0.037 * Re A 0.8 "‘ Sc_mix A (1 /3) '0.037 * Re A 0.8 * Sc_mix A (1 l3) H_M=SH*Dsa/W h=Nu"'k_mix/W 'Calculate surface moisture concentration If Tave < 0 Then Csat = 0.000000621530732 Else Csat = 8.12078904E-10 "' Tave A 3 - 0.000000035203247 "‘ Tave A 2 + 0.00000131977474 "' Tave + 0.000000621530732 EndIf 'Calculate effective relative humidity RH = 100 * (ConcOvenAir / Csat) MDB = (Mave "' 100) / (100 - Mave) ERH = Exp((-5222.47 * (MDB A -1.0983)) / (1.9818 " (Tave + 273))) Cs = ERH * Csat 'Calculate energy fluxes Q_conv = h * (OvenAirT - Tave) If ConcOvenAir > Csat Then Q_cond = H_M * latent "‘ (ConcOvenAir - Cs) Q_evap = 0 Else Q_cond = 0 219 End If If Cs > ConcOvenAir Then Q_evap = H_M * latent * (Cs - ConcOvenAir) Q_cond = 0 Else Q_evap = 0 Endlf Q_total = Q_conv + Q_cond - Q_evap 'Effective heat transfer coefficient based on total flux (convection, condensing, evaporation) heff = Q_total / (OvenAirT - Tave) 'Calculate equilibrium values under current conditions 'Enthalpy Hequilibrium = OvenAirT "‘ 3.35 + frozenH + latentW + latentF + (2 "‘ 3.35) 'Moisture If RH >= 97 Then EMCWB = InitialM Else EMC = ((1.9818 "' (Tave + 273) "' Log(RH/ 100)) / -5222.47) A (1 /-l.0983) EMCWB = (100 "' EMC) / (100 + EMC) End If X = 1 "' InitialM IfX <= (EMCWB " 1) Then EMCWB = X End If 'FEM boundary coefficients 'Heat transfer CoefM = heff "' SideLength / 6 CoefS = heff "‘ Hequilibrium * SideLength / 2 'Moisture transfer CoefM_M = H_M * density "' moisture_capacity "' SideLength / 6 'Moisture transfer CoefS_M = H_M * density * moisture_capacity * EMCWB * SideLength / 2 Elself BoundaryType(l) = 4 Then 'Stein 102 oven 8 'height above belt 1.27 'nozzle diameter 6 h d L 'nozzle spacing 220 'F = 0.906 * (d / l) A 2 'offset rows configuration ' o o o o o o o F = 0.785 * (d / L) A 2 'even rows configuration ' o o o o o o o 0 'Calculate transport coefficients Re = OvenAirV * d "' density_mix / viscosity_mix KHDF = (1 + ((h / d) / (0.6 / Sqr(F))) A 6) A -0.05 GFHD = ((1 - (2.2 * Sqr(F))) / (1 + 0.2 * (h / d - 6) * Sqr(F))) * 2 * Sqr(F) FREARN = 0.5 * (Re A 0.66) Nu = (Pr_mix A 0.42) “ KHDF "' GFHD * FREARN SH = (Sc_mix A 0.42) * KHDF " GFHD "' FREARN H_M=SH*Dsa/d h=Nu"'k_mix/d 'Calculate surface moisture concentration If Tave < 0 Then Csat = 0.000000621530732 Else Csat = 8.12078904E-10 * Tave A 3 - 0.000000035203247 * Tave A 2 + 0.00000131977474 * Tave + 0.000000621530732 End If 'Calculate effective relative humidity RH = 100 * (ConcOvenAir/ Csat) MDB = (Mave "‘ 100) / (100 — Mave) ERH = Exp((-5222.47 "‘ (MDB A -1.0983)) / (1.9818 * (Tave + 273))) Cs = ERH * Csat 'Calculate energy fluxes Q_conv = h * (OvenAirT - Tave) If ConcOvenAir > Cs Then Q_cond = H_M * latent * (ConcOvenAir - Cs) Q_evap = 0 Q_cond = 0 Else End If If Cs > ConcOvenAir Then Q_evap = H_M * latent * (Cs - ConcOvenAir) Q_cond = 0 Else Q_evap = 0 End If 221 Endlf Q_total = Qconv + Q_cond - Q_evap 'Effective heat transfer coefficient based on total flux (convection, condensing, evaporation) heff = Q_total / (OvenAirT - Tave) 'Calculate equilibrium values under current conditions 'Enthalpy Hequilibrium = OvenAirT * 3.35 + frozenH + latentW + latentF '792.64 + 6.44 'Moisture If RH >= 97 Then EMCWB = InitialM Else EMC = ((1.9818 " (Tave + 273) "' Log(RH/ 100)) / -5222.47) A (l / -l .0983) EMCWB = (100 * EMC) / (100 + EMC) End If X = 1 * InitialM If X <= (EMCWB * 1) Then EMCWB = X Endlf 'F EM boundary coefficients 'Heat transfer CoefM = heff * SideLength / 6 CoefS = heff "' Hequilibrium * SideLength / 2 'Moisture transfer CoefM_M = H_M * density * moisture_capacity "' SideLength / 6 'Moisture transfer CoefS_M = H_M * density " moisture_capacity * EMCWB "' SideLength / 2 'Evaluate the element matrices EKM(1,1)=2"‘ 3.14/2 ‘CoefM*(3 * Ri+Rj) EKM(1,2)=2 * 3.14/2 * CoeflVl *(R1+Rj) EKM(2,1)=2 * 3.14/2 * CoefM . (Ri +Rj) EKM(2,2)=2 * 3.14/2 * CoefM *(Ri+3 * Rj) EKM_M(1,1)= 2 * 3.14/2 * CoefM_M * (3 * Ri +Rj) EKM_M(I, 2) = 2 * 3.14/2 * CoefM_M * (Ri + Rj) EKM_M(2, 1) = 2 * 3.14/2 * CoefM_M * (R1 + Rj) EKM_M(2, 2) = 2 * 3.14 / 2 * CoefM_M * (R1 + 3 * Rj) 222 EFS(1)=2*3.14/3‘CoetS*(2*Ri+Rj) EFS(2)=2*3.14/3‘CoefS*(Ri+2"'Rj) EFS_M(I) =2 * 3.14/3 * CoefS_M * (2 * R1 + Rj) EFS_M(2) = 2 * 3.14 / 3 * CoefS_M * (R1 + 2 * Rj) End Sub Module — MultplyBandMatrix Sub MultpyBandMatrix(NumNodes, BandWidth, GSM(), GFV(), ProdVectorO) This module provided by Dr. Larry Segerlind - Michigan State University This subprogram multiplies a symmetric banded matrix and a column vector ' The banded matrix is stored as a rectangular array and only the ' coefficients on and above the main diagonal are stored in the array. For I = 1 To NumNodes Sum=0! K=I-l For J = 2 To BandWidth M = J +1 - 1 If (M <= NumNodes) Then Sum = Sum + GSM(I, J) * GFV(M) End If If (K > 0) Then Sum = Sum + GSM(K, J) * GFV(K) K = K - 1 End If Next J ProdVector(I) = Sum + GSM(I, 1) * GFV(I) Next I End Sub 223 Module - EleSubscriptValues Sub EleSubscriptValues(EleNodeData(), KK, NumEleSub, EleSub()) 'This module provided by Dr. Larry Segerlind - Michigan State University ' This subprogram calculates the subscripts associated with the ' element. The subprogram allows the element to have one, two or ' three unknown values at a node. NumEleSub = 4 If (EleNodeData(KK, 4) = 0) Then NumEleSub = 3 EleSub(l) = EleNodeData(KK, 1) EleSub(2) = EleNodeData(KK, 2) EleSub(3) = EleNodeData(KK, 3) EleSub(4) = EleNodeData(KK, 4) End Sub 224 Module - ESMatrix2DField Sub ESMatrix2DField(E1eMatlData(), EleNodeData(), Coord(), E1ePhyData(), EFV(), ESM(), EFV_M(), ESM_M(), EFV__F(), ESM_F(), Volume(), KK) ' This module heavily modified from a module by Dr. Larry Segerlind - Michigan State University This subprogram calculates the element stiffness matrix and element ' force vector for the three node triangular and the four node ' bilinear rectangular element 'Evaluation of the parameters in the element stiffness matrix MatlSet = EleMatlData(KK) Thermal properties Dx = ElePhyData(MatlSet, l) Dy = ElePhyData(MatlSet, 2) G = ElePhyData(MatlSet, 3) Q = ElePhyData(MatlSet, 4) 'Moisture transfer properties Dx_M = ElePhyData(MatlSet, 6) Dy_M = ElePhyData(MatlSet, 7) G_M = ElePhyData(MatlSet, 8) QM = ElePhyData(MatlSet, 9) 'Fat transfer properties Dx_F = ElePhyData(MatlSet, ll) Dy_F = ElePhyData(MatlSet, 12) G_F = ElePhyData(MatlSet, 13) Q_F = ElePhyData(MatlSet, 14) 'Evaluate the element stiffness matrix If (EleNodeData(KK, 4) = 0) Then Triangular Element XI = Coord(EleNodeData(KK, 1), 1) Y1 = Coord(EleNodeData(KK, 1), 2) Xj = Coord(EleNodeData(KK, 2), l) Yj = Coord(EleNodeData(KK, 2), 2) Xk = Coord(EleNodeData(KK, 3), 1) Yk = Coord(EleNodeData(KK, 3), 2) Ai=Xj*Yk-Xk"Yj Aj=Xk"'Yl-XI*Yk Ak=XI*Yj-Xj"‘YI Bi=Yj-Yk Bj=Yk-Yl Bk=YI~Yj 225 Rbar * Rbar * Rbar * Rbar "' Rbar * Rbar 3.14 3.14 c1=xrc-x1 Q=XLXk Ck=Xj-Xl TwiceArea = Ai + Aj + Ak Rbar=(XI+Xj +Xk)/3 Volume(KK) = TwiceArea "' 3.14 * Rbar DxOver4A = Dx / (2 "' TwiceArea) DyOver4A = Dy / (2 "' TwiceArea) GAreaOver12 = G "' TwiceArea / 24 Dx_MOver4A = Dx_M / (2 " TwiceArea) Dy_MOver4A = Dy_M / (2 " TwiceArea) G_MAreaOverl2 = G_M " TwiceArea / 24 Dx_FOver4A = Dx_F / (2 "' TwiceArea) Dy_FOver4A = Dy_F / (2 "‘ TwiceArea) G_FAreaOverlZ = G_F "‘ TwiceArea / 24 'Calculate the element stiffness matrix - enthalpy ESM(1,1)= 2 * 3.14 “ Rbar "‘ (DxOver4A "' Bi "‘ Bi + DyOver4A * Ci * Ci) + 2 "‘ 3.14 * GAreaOver12 " 2 ESM(I, 2) = 2 "‘ 3.14 " Rbar "‘ (DxOver4A * Bi " Bj + DyOver4A * Ci * Cj) + 2 * 3.14 "' GAreaOver12 ESM(l, 3) = 2 "' 3.14 * Rbar “ (DxOver4A * Bi * Bk + DyOver4A * Ci * Ck) + 2 * 3.14 "' GAreaOver12 ESM(2, l) = ESM(l, 2) ESM(2, 2) = 2 " 3.14 " Rbar * (DxOver4A "' Bj * Bj + DyOver4A * Cj "' Cj) + 2 "' 3.14 * GAreaOver12 * 2 ESM(2, 3) = 2 "' 3.14 "' Rbar "' (DxOver4A * Bj * Bk + DyOver4A * Cj * Ck) + 2 * 3.14 " GAreaOver12 ESM(3, l) = ESM(l, 3) ESM(3, 2) = ESM(2, 3) ESM(3, 3) = 2 "' 3.14 "' Rbar "‘ (DxOver4A " Bk * Bk + DyOver4A * Ck * Ck) + 2 "' 3.14 * " GAreaOver12 "' 2 'Calculate the element stiffness matrix - moisture ESM_M(l, l) = 2 "' 3.14 * Rbar "' (Dx_MOver4A "' Bi "‘ Bi + Dy_MOver4A " Ci "' Ci) + 2 "' * Rbar * G_MAreaOverlZ * 2 ESM_M(l, 2) = 2 * 3.14 * Rbar * (Dx_MOver4A * 131 r 131 + Dy_MOver4A * Ci * Cj) + 2 * "' Rbar * G_MAreaOverlZ ESM_M(l, 3) = 2 * 3.14 "' Rbar "' (Dx_MOver4A * Bi * Bk + Dy_MOver4A "‘ Ci " Ck) + 2"I 3.14 "' Rbar * G_MAreaOverlZ ESM_M(Z, 1) = ESM_M(l, 2) 226 3.14 3.14* 3.14* 3.14* 3.14* 3.14* 3.14 Else ESM_M(2, 2) = 2 * 3.14 * Rbar * (Dx_MOver4A * Bj * Bj + Dy_MOver4A * Cj * Cj) + 2 * * Rbar * G_MAreaOver12 * 2 ESM_M(2, 3) = 2 * 3.14 "‘ Rbar " (Dx_MOver4A * Bj * Bk + Dy_MOver4A * Cj " Ck) + 2 3.14 * Rbar * G_MAreaOver12 ESM_M(3, l) = ESM_M(l, 3) ESM_M(3, 2) = ESM_M(2, 3) ESM_M(3, 3) = 2 "' 3.14 * Rbar * (Dx_MOver4A "‘ Bk * Bk + Dy_MOver4A * Ck * Ck) + 2 3.14 * Rbar "' G_MAreaOver12 " 2 'Calculate the element stiffness matrix - fat ESM_F(I, 1) = 2 * 3.14 * Rbar * (Dx_FOver4A * Bi "' Bi + Dy_FOver4A * Ci * Ci) + 2 * Rbar * G_FAreaOver12 * 2 ESM_F(l, 2) = 2 * 3.14 * Rbar " (Dx_FOver4A "' Bi * Bj + Dy_FOver4A * Ci * Cj) + 2 * Rbar * G_FAreaOver12 ESM_F(l, 3) = 2 "' 3.14 "' Rbar * (Dx_FOver4A * Bi * Bk + Dy_FOver4A * Ci * Ck) + 2 * Rbar * G_FAreaOver12 ESM_F(2, l) = ESM_F(I, 2) ESM_F(2, 2) = 2 * 3.14 " Rbar * (Dx_FOver4A * Bj * B] + Dy_FOver4A * Cj * Cj) + 2 "' Rbar "' G_FAreaOver12 "‘ 2 ESM_F(2, 3) = 2 "’ 3.14 " Rbar * (Dx_FOver4A * Bj * Bk + Dy_FOver4A * Cj * Ck) + 2 " Rbar "' G_FAreaOver12 ESM_F(3, 1) = ESM_F(I, 3) ESM_F(3, 2) = ESM_F(2, 3) ESM_F(3, 3) = 2 * 3.14 * Rbar "' (Dx_FOver4A * Bk * Bk + Dy_FOver4A * Ck * Ck) + 2 * "' Rbar * G_FAreaOver12 * 2 'Calculate the element force vector - enthalpy EFV(l) = (2 * 3.14 / 4) * (Q * TwiceArea/6) "' (2 * XI + Xj + Xk) EFV(2) = (2 "‘ 3.14 / 4) "‘ (Q * TwiceArea / 6) "' (XI + 2 * Xj + Xk) EFV(3) = (2 "‘ 3.14 / 4) "' (Q "' TwiceArea / 6) * (XI + Xj + 2 "‘ Xk) 'Calculate the element force vector - moisture EFV_M(1)=(2 * 3.14 / 4) "‘ (QM "' TwiceArea / 6) “ (2 "' XI + Xj + Xk) EFV_M(2) = (2 * 3.14 / 4) “ (QM "‘ TwiceArea / 6) "' (XI + 2 * Xj + Xk) EFV_M(3) = (2 * 3.14 / 4) * (QM "‘ TwiceArea / 6) "‘ (XI + Xj + 2 * Xk) 'Calculate the element force vector - fat EFV_F(l) = (2 * 3.14 / 4) "‘ (QF "' TwiceArea / 6) * (2 "' XI + Xj + Xk) EFV_F(2) = (2 * 3.14 / 4) "' (QF * TwiceArea / 6) * (XI + 2 "‘ Xj + Xk) EFV_F(3) = (2 "' 3.14 / 4) "‘ (QF * TwiceArea / 6) * (XI + Xj + 2 "' Xk) 'Rectangular element 227 XLength = Coord(EleNodeData(KK, 2), 1) - Coord(EleNodeData(KK, l), l) YLength = Coord(EleNodeData(KK, 4), 2) - Coord(EleNodeData(KK, 1), 2) Area = XLength * YLength Rbar = (Coord(EleNodeData(KK, l), 1) + Coord(EleNodeData(KK, 2), 1) + Coord(EleNodeData(KK, 3), l) + Coord(EleNodeData(KK, 4), 1)) / 4 Volume(KK) = Area " 2 * Rbar * 3.14 Donver6B = Dx "‘ YLength / (6 "' XLength) DyBover6A = Dy "' XLength / (6 * YLength) GAreaover36 = G "‘ Areal 36 Dx_Moner63 = Dx_M * YLength / (6 "‘ XLength) Dy_MBover6A = Dy_M "' XLength / (6 “ YLength) G_MAreaover36 = G_M * Area / 36 Dx_Foner63 = Dx_F * YLength / (6 "' XLength) Dy_FBover6A = Dy_F "' XLength / (6 * YLength) G_FAreaover36 = G_F "' Area / 36 'Calculate element stiffness matrix - enthalpy ESM(l, 1) = 2 * Donver68 + 2 * DyBover6A + 4 * GAreaover36 ESM(l, 2) = -2 * Donver63 + DyBover6A + 2 * GAreaover36 ESM(l, 3) = -Donver68 - DyBover6A + GAreaover36 ESM(l, 4) = Donver6B - 2 * DyBover6A + 2 "' GAreaover36 ESM(2, 1) = ESM(l, 2) ESM(2, 2) = ESM(l, l) ESM(2, 3) = ESM(l, 4) ESM(2, 4) = ESM(l, 3) ESM(3, 1) = ESM(l, 3) ESM(3, 2) = ESM(2, 3) ESM(3, 3) = ESM(l, 1) ESM(3, 4) = ESM(l, 2) ESM(4, 1) = ESM(l, 4) ESM(4, 2) = ESM(2, 4) ESM(4, 3) = ESM(3, 4) ESM(4, 4) = ESM(l, 1) 'Calculate element stiffness matrix - moisture ESM_M(], l) = 2 * Dx_Moner68 + 2 "' Dy_MBover6A + 4 * G_MAreaover36 ESM_M(l, 2) = -2 * Dx_Moner68 + Dy_MBover6A + 2 "' G_MAreaover36 ESM_M(], 3) = -Dx_Moner68 - Dy_MBover6A + G_MAreaover36 ESM_M(l , 4) = Dx_Moner68 - 2 * Dy_MBover6A + 2 * G_MAreaover36 ESM_M(2, 1) = ESM_M(], 2) ESM_M(2, 2) = ESM_M(], l) ESM_M(2, 3) = ESM_M(], 4) ESM_M(2, 4) = ESM_M( 1, 3) ESM_M(3, 1)= ESM_M(l, 3) ESM_M(3, 2) = ESM_M(2, 3) ESM_M(3, 3) = ESM_M(l, l) ESM_M(3, 4) = ESM_M(], 2) ESM_M(4, l) = ESM_M(l, 4) ESM_M(4, 2) = ESM_M(2, 4) ESM_M(4, 3) = ESM_M(3, 4) ESM_M(4, 4) = ESM_M(l, l) 228 End If End Sub 'Calculate element stiffness matrix - fat ESM_F(l, 1) = 2 * Dx_Foner63 + 2 "' Dy_FBover6A + 4 "' G_FAreaover36 ESM_F( 1 , 2) = -2 "' Dx_Foner68 + Dy_FBover6A + 2 * G_FAreaover36 ESM_F( l , 3) = -Dx_Foner63 - Dy_FBover6A + G_FAreaover36 ESM_F( 1, 4) = Dx_Foner6B - 2 "‘ Dy_FBover6A + 2 " G_FAreaover36 ESM_F(2, l) = ESM_F(], 2) ESM_F(2, 2) = ESM_F(], l) ESM_F(2, 3) = ESM_F(], 4) ESM_F(2, 4) = ESM_F(l, 3) ESM_F(3, 1) = ESM_F(I, 3) ESM_F(3, 2) = ESM_F(2, 3) ESM_F(3, 3) = ESM_F(l, l) ESM_F(3, 4) = ESM_F(l, 2) ESM_F(4, l) = ESM_F(], 4) ESM_F(4, 2) = ESM_F(2, 4) ESM_F(4, 3) = ESM_F(3, 4) ESM_F(4, 4) = ESM_F(], 1) 'Calculate element force vector - enthalpy EFV(1)= Q * Area/4 EFV(2) = EFV(l) EFV(3) = EFV(l) EFV(4) = EFV( 1) 'Calculate element force vector - moisture EFV_M(1)= Q_M * Area / 4 EFV_M(2) = EFV_M(l) EFV_M(3) = EFV_M(l) EFV_M(4) = EFV_M(l) 'Calculate element force vector - fat EFV_F(1)= Q_F "' Area / 4 EFV_F(2) = EFV_F( l) EFV_F(3) = EFV_F( 1) EFV_F(4) = EFV_F(l ) 229 Module - InputBasicData Sub InputBasicData(NumMatlSets, NumNodes, NumEle, Coord(), EleMatlData(), EleNodeData()) This module slightly modified from a module by Dr. Larry Segerlind - Michigan State University This subprogram inputs the nodal coordinates, the element material ' index, the element node numbers and the equation coefficients Dx, Dy, ' G and Q For I = 1 To NumNodes Input #1, Coord(I, 1), Coord(I, 2) Next I For I = 1 To NumEle Input #1, EleMatlData(I), EleNodeData(I, l), EleNodeData(I, 2), EleNodeData(I, 3), EleNodeData(l, 4) Next I End Sub 230 Module — InputDerivBC Sub InputDerivBC(NumDerivBC, NudeyNodes, deNodeO, SubDerivBC(), BoundaryTypeO) 'This module slightly modified from a module by Dr. Larry Segerlind - Michigan State University This subprogram inputs the subscripts and coefficients for the ' derivative boundary condition For I = 1 To NumDerivBC Input #1, SubDerivBC(l, l), SubDerivBC(l, 2), BoundaryType(I) Next I For I = 1 To NudeyNodes Input #1, deNode(I) Next I End Sub 231 Module — BuildGAM Sub BuildGAM(BandWidth, GAM(), GCM(), GSM(), GAM_MO, GCM_M(), GSM_M(), GAM_F(), GCM_F(), GSM_F(), timestep, NumNodes) 'This module calculates the A matrix for the finite difference time solution using the central ' difference method For X = 1 To NumNodes For Y = 1 To BandWidth GAM(X, Y) = GCM(X, Y) + (GSM(X, Y) "‘ (timestep / 2)) GAM_M(X, Y) = GCM_M(X, Y) + (GSM_M(X, Y) "' (timestep / 2)) GAM_F(X, Y) = GCM_F(X, Y) + (GSM_F(X, Y) "‘ (timestep / 2)) Next Y Next X End Sub 232 Module - SolveBandMatrix Sub SolveBandMatrix(NumNodes, BandWidth, GSV(), GFV(), GSM()) 'This module provide by Dr. Larry Segerlind - Michigan State University This subprogram solves the system of banded equations using the method ' of Gaussian elimination. The stiffness matrix has been decomposed ' prior to entering this subprogram using DecomposeBandMatrix. ' This subprogram decomposes the column vector GFV(NumNodalVal%) before ' solving the system using backward substitution 'Decompose the global force vector ForI= 1 To (NumNodes- 1) MJ =1 + BandWidth - 1 If (MJ > NumNodes) Then M] = NumNodes Endlf NJ = I + l L = 1 For J = N] To M] L = L + 1 GFV(J) = GFV(J) - GSM(I, L) * GFV(I) / GSM(I, 1) Next 1 Next I ' Solution by backward substitution GSV(NumNodes) = GFV(NumNodes) / GSM(NumNodes, 1) For K = 1 To (NumNodes - l) I = NumNodes - K MJ = BandWidth If ((I + BandWidth - 1) > NumNodes) Then MJ=NumNodes-I+l End If Sum = 0! For J = 2 To MJ ' n=I+J-1 Sum = Sum + GSM(I, J) "' GSV(n) Next J GSV(I) = (GFV(I) - Sum) / GSM(I, 1) Next K End Sub 233 Module — BuildBandedSystem Sub BuildBandedSystem(NumEleSub, EleSub(), EFV(), ESM(), GFV(), GSM(), ECM(), GCM(), EFV_MO, ESM_M(), GFV_M(), GSM_M(), ECM_M(), GCM_M(), EFV_FO. ESM_F(), GFV_F(), GSM_F(), ECM_F(), GCM_F()) This module modified from a module provided by Dr. Larry Segerlind - Michigan State University This subprogram incorporates the element force vector and the element ' stiffness matrix into the global force matrix and the global stiffness ' matrix. 'This subprogram ASSUMES that the global stiffness matrix is symmetrical ' and stored in a rectangular format. The direct stiffness method for a banded system of equations For I = 1 To NumEleSub II = EleSub(I) 'Force vector GFV(II) = GFV(II) + EFV(I) GFV_M(II) = GFV_M(II) + EFV_M(I) GFV_F(II) = GFV_F(II) + EFV_F(I) For J = 1 To NumEleSub JJ = EleSub(J) JJ = JJ - II + 1 If (JJ > 0) Then 'Stiffness matrix GSM(II, JJ) = GSM(II, JJ) + ESM(I, J) GSM_M(II, JJ) = GSM_M(II, JJ) + ESM_M(I, J) GSM_F(II, JJ) = GSM_F(II, JJ) + ESM_F(I, J) 'Capacitance matrix GCM(II, JJ) = GCM(II, JJ) + ECM(I, J) GCM_M(II, JJ) = GCM_M(II, JJ) + ECM_M(I, J) GCM_F(II, JJ) = GCM_F(II, JJ) + ECM_F(I, J) End If Next J Next 1 End Sub 234 Module — BuildGPM Sub BuildGPM(BandWidth, GPM(), GCM(), GSM(), GPM_M(), GCM_M(), GSM_M(), GPM_F(), GCM_F(), GSM_F(), timestep, NumNodes) 'This module calculates the P matrix for the finite difference time solution using the central ' difference method For X = 1 To NumNodes For Y = 1 To BandWidth GPM(X, Y) = GCM(X, Y) - (GSM(X, Y) " (timestep / 2)) GPM_M(X, Y) = GCM_M(X, Y) - (GSM_M(X, Y) ‘ (timestep / 2)) GPM_F(X, Y) = GCM_F(X, Y) - (GSM_F(X, Y) "' (timestep / 2)) Next Y Next X End Sub 235 Module — ModifyGFV Sub ModifyGFV(GFV(). GFV_A(), GFV_StarO, GFV_M(), GFV_M_AO. GFV_M_StarO, GFV_F_A(), GFV_F_StarO, GFV_F(), timestep, NumNodes) 'Modify the global force vector for the finite difference time solution For X = 1 To NumNodes GFV_Star(X) = (GFV_A(X) + GFV(X)) * (timestep / 2) GFV_M_Star(X) = (GFV_M(X) + GFV_M(X)) * (timestep / 2) GFV_F_Star(X) = (GFV_F_A(X) + GFV_F(X)) * (timestep / 2) Next X End Sub 236 Module — EleCapMatrix Sub EleCapMatrix(ElePhyData(), EleNodeData(), ECM(), ECM_M(), ECM_F(), Coord(), KK) ' Compose the element capacitance matrix If (EleNodeData(KK, 4) = 0) Then ' Triangular Element ' [c]=lamda*A/3*[ l 0 0 ' 0 1 0 ' 0 l 1 ] EleNodel = EleNodeData(KK, 1) EleNode2 = EleNodeData(KK, 2) EleNode3 = EleNodeData(KK, 3) ' Calculate the element area Ri = Coord(EleNodel, 1) Rj = Coord(EleNodeZ, 1) Rk = Coord(EleNode3, 1) Rbar=(Ri+Rj+Rk)/3 XLengthSidel = Coord(EleNodeZ, l) — Coord(EleNodel, l) YLengthSidel = Coord(EleNodeZ, 2) - Coord(EleNodel , 2) XLengthSide2 = Coord(EleNode3, l) - Coord(EleNodeZ, 1) YLengthSide2 = Coord(EleNode3, 2) - Coord(EleNodeZ, 2) XLengthSide3 = Coord(EleNode3, l) - Coord(EleNodel, l) YLengthSide3 = Coord(EleNode3, 2) - Coord(EleNodel, 2) SidelLength = Sqr(XLengthSidel A 2 + YLengthSidel A 2) Side2Length = Sqr(XLengthSideZ A 2 + YLengthSide2 A 2) Side3Length = Sqr(XLengthSide3 A 2 + YLengthSide3 A 2) Half? = (SidelLength + Side2Length + Side3Length) / 2 Area = Sqr(HalfP * (HalfP - SidelLength) * (HalfP - Side2Length) * (HalfP - Side3Length» Lamda = ElePhyData(l, 5) Lamda_M = ElePhyData(l, 10) Lamda_F = ElePhyData(l, 15) ECM(1,1)= 2 * 3.14 "' Rbar "' Lamda * Area/3 ECM(2, 2) = 2 "' 3.14 * Rbar * Lamda * Area/3 ECM(3, 3) = 2 * 3.14 " Rbar * Lamda "' Area/3 ECM_M(1,1)= 2 * 3.14 "' Rbar * Lamda_M * Area/3 ECM_M(Z, 2) = 2 "' 3.14 * Rbar "' Lamda_M * Area / 3 ECM_M(3, 3) = 2 * 3.14 * Rbar "' Lamda_M * Area / 3 ECM_F(1,1)= 2 * 3.14 "‘ Rbar * Lamda_F "' Area / 3 237 ECM_F(Z, 2) = 2 "' 3.14 " Rbar * Lamda_F "‘ Area / 3 ECM_F(3, 3) = 2 "' 3.14 * Rbar "' Lamda_F " Area / 3 Else ' Rectangular element ' [c]=1amda*A/4"'[ l 0 0 0 ' 0100 ' 0010 ' 0001] EleNodel = EleNodeData(KK, l) EleNode2 = EleNodeData(KK, 2) EleNode3 = EleNodeData(KK, 3) EleNode4 = EleNodeData(KK, 4) ' Calculate the element area Sidel = Coord(EleNode4, 2) - Coord(EleNodel, 2) Side2 = Coord(EleNodeZ, 1) - Coord(EleNodel, 1) Area = Sidel "' Side2 Lamda = ElePhyData(l, 5) Lamda_M = ElePhyData(l, 10) Lamda_F = ElePhyData(l, 15) For J = 1 To 4 ECM(J, J) = Lamda * Area / 4 ECM_M(J, J) = Lamda_M "' Area / 4 ECM_F(J, J) = Lamda_F * Area / 4 Next J Endlf End Sub 238 Module - ConverthoTemp Sub ConverthoTemp(heat_capacity_f, GSV(), GSV_M(), GSV_F(), temperature(), NumNodes, InitialM, Solid, frozenH, latentW, latentF, heat_capacity, LevelOne, LevelTwo, LevelThree, LevelFour, LevelFive) 'Convert enthalpy to temperature data For K = 1 To NumNodes If GSV(K) < LevelOne Then temperature(K) = GSV(K) / heat_capacity_f - 273 Elself GSV(K) >= LevelOne And GSV(K) <= LevelTwo Then temperature(K) = -2 Elself GSV(K) > LevelTwo And GSV(K) <= LevelThree Then temperature(K) = ((GSV(K) - (frozenH + latentW)) / 3.35) - 2 Elself GSV(K) > LevelThree And GSV(K) <= LevelFour Then temperature(K) = 45 Elself GSV(K) > LevelFour Then temperature(K) = ((GSV(K) - (frozenH + latentW + 1atentF))/ 3.35) - 2 End If Next K End Sub 239 Module — CalculateProperties Sub CalculateProperties(Meat, Solid, NumEle, EleNodeData(), E1ePhyData(), temperature(), GSV_M, GSV_F(), moisture_capacity, density) ForI = 1 To NumEle 'Calculate element average temperature, moisture, and fat If EleNodeData(I, 4) = 0 Then ElementAverageT = (1 "' temperature(EleNodeData(I, 1)) + 1 * temperature(EleNodeData(I, 2)) + 1 "' temperature(EleNodeData(I, 3))) / 3 ElementAverageM = (1 * GSV_M(EleNodeData(I, 1)) + 1 * GSV_M(EleNodeData(I, 2)) + 1 " GSV_M(EleNodeData(I, 3))) / 3 ElementAverageF P = (1 "' GSV_F(EleNodeData(I, 1)) + 1 * GSV_F(EleNodeData(I, 2)) + 1 * GSV_F(EleNodeData(I, 3))) / 3 Else ElementAverageT = (l "‘ temperature(EleNodeData(I, 1)) + 1 * temperature(EleNodeData(I, 2)) + l * temperature(EleNodeData(I, 3)) + 1 * temperature(EleNodeData(I, 4))) / 4 ElementAverageM = (1 " GSV_M(EleNodeData(I, 1)) + 1 * GSV_M(EleNodeData(I, 2)) + 1 "' GSV_M(EleNodeData(I, 3)) + 1 "' GSV_M(EleNodeData(I, 4))) / 4 ElementAverageFP = (l "' GSV_F(EleNodeData(I, 1)) + l * GSV_F(EleNodeData(I, 2)) + 1 * GSV_F(EleNodeData(I, 3)) + l "' GSV_F(EleNodeData(I, 4))) / 4 End If moisture = ElementAverageM/ 100 'Element average moisture, decimal wb MP = Solid 'Mass protein MF = ElementAverageF P * Solid 'Mass fat MM = (moisture * MF + moisture * MP) / (1 - moisture) 'Mass moisture Total = MP + MM + MF 'Mass total Fat = MF / Total 'Percent fat Protein = MP / Total 'Percent protein 'Beef If Meat = 1 Then If ElementAverageT <= -2 Then heat_capacity = 1.9 "‘ moisture + 1.711 * Protein + 1.298 "' Fat Else heat_capacity = 0.9 * (4.18 "' moisture + 1.711 " Protein + 1.298 "' Fat) '(4.l8 * moisture + 1.711 * Protein + 1.298 "' Fat) End If moisture_capacity = 0.003 'g/g 240 capacity capacity melting melting density = 1 / (moisture + Fat / 0.9 + Protein / 1.4) 'g/cmA3 Vmoisture = moisture / 1 'Volume moisture Vprotein = Protein / 1.4 'Volume protein Vfat = Fat / 0.9 'Volume fat Vtotal = Vmoisture + Vprotein + Vfat Total volume VFm = Vmoisture / Vtotal 'Volume fraction moisture VFp = Vprotein / Vtotal 'Volume fraction protein VFf = Vfat / Vtotal 'Volume fraction fat 'thermal conductivity ElePhyData(l, 1) = ((0.602 * VFm + 0.18 * VFf + 0.2 "' VFp)/ 100) ElePhyData(l, 2) = ((0.602 ’ VFm + 0.18 " VFf + 0.2 * VFp)/ 100) ElePhyData( 1, 5) = heat_capacity " density 'density‘heat 'Coefficient for Mittal moisture diffusivity equation A = 0.003 ElePhyData(], 6) = 10000 * A * (0.003 "' Exp(-0.442 * ElementAverageFP - 4829.7 (ElementAverageT + 273) + 11.55 * moisture)) * moisture_capacity "' density 'moisture conductivity E1ePhyData(l, 7) = 10000 "‘ A * (0.003 * Exp(-0.442 * ElementAverageFP - 4829.7 (ElementAverageT + 273) + 11.55 "' moisture)) " moisture_capacity * density 'moisture conductivity ElePhyData(], 10) = density "' moisture_capacity 'density‘moisture If ElementAverageT <= 45 Then ElePhyData(] , 11) = 0 'fat diffusivity r-direction zero below ElePhyData(], 12) = 0 'fat diffusivity z-direction zero below Else ElePhyData(], 11) = 0.0008 'fat diffusivity r-direction ElePhyData(], 12) = 0.0008 'fat diffusivity z-direction End If E1ePhyData(l , 15) = 1 'Variable not used for fat trasfer equations Elself Meat = 2 Then If ElementAverageT <= -2 Then heat_capacity = 1.9 " moisture + 1.711 * Protein + 1.298 "‘ Fat Else heat_capacity = (4.18 * moisture + 1.711 "' Protein + 1.298 * Fat) '(4.18 "' moisture + 1.711 "' Protein + 1.298 "' Fat) End If 241 capacity capacity melting melting Turkey moisture_capacity = 0.003 'g/ g density = l / (moisture + Fat / 0.9 + Protein / 1.4) 'g/cmA3 Vmoisture = moisture / 1 'Volume moisture Vprotein = Protein / 1.4 'Volume protein Vfat = Fat / 0.9 'Volume fat Vtotal = Vmoisture + Vprotein + Vfat Total volume VFm = Vmoisture / Vtotal 'Volume fraction moisture VFp = Vprotein / Vtotal 'Volume fraction protein VF f = Vfat / Vtotal 'Volume fraction fat 'thermal conductivity ElePhyData(l, 1) = ((0.602 " VFm + 0.18 "' VFf+ 0.2 * VFp)/ 100) ElePhyData(], 2) = ((0.602 * VFm + 0.18 "' VFf + 0.2 * VFp)/ 100) ElePhyData(l, 5) = heat_capacity "' density 'density‘heat 'Coefficient for Mittal moisture diffusivity equation A = 0.003 ElePhyData(l, 6) = 10000 "‘ A * (0.003 "‘ Exp(-0.442 "' ElementAverageFP - 4829.7 (ElementAverageT + 273) + 11.55 " moisture)) "' moisture_capacity "' density 'moisture conductivity 0.0000000129 ' ElePhyData(l, 7) = 10000 "' A * (0.003 "' Exp(-0.442 " ElementAverageFP - 4829.7 (ElementAverageT + 273) + 11.55 "' moisture)) * moisture_capacity * density ‘moisture conductivity 0.0000000129 ' ElePhyData(], 10) = density "' moisture_capacity 'density*moisture If ElementAverageT <= 45 Then ElePhyData(l, l 1) = 0 'fat diffusivity r-direction zero below ElePhyData(], 12) = 0 'fat diffusivity z-direction zero below Else ElePhyData(l, 11) = 0.00125 'fat diffusivity r-direction ElePhyData( l , 12) = 0.00125 'fat diffusivity z-direction End If ElePhyData( 1 , 15) = 1 'Variable not used for fat trasfer equations Elself Meat = 3 Then If ElementAverageT <= -2 Then 242 capacity capacity melting melting End If heat_capacity = 1.9 * moisture + 1.711 * Protein + 1.298 "' Fat Else heat_capacity = (4.18 * moisture + 1.711 "' Protein + 1.298 * Fat) '(4.18 * moisture + 1.711 " Protein + 1.298 "‘ Fat) Endlf moisture_capacity = 0.003 'g/g density = l / (moisture + Fat / 0.9 + Protein / 1.4) 'g/cmA3 Vmoisture = moisture / 1 'Volume of moisture Vprotein = Protein / 1.4 'Volume of protein Vfat = Fat / 0.9 'Volume of fat Vtotal = Vmoisture + Vprotein + Vfat Total volume VFm = Vmoisture / Vtotal 'Volume fraction moisture VFp = Vprotein / Vtotal 'Volume fraction protein VFf = Vfat / Vtotal 'Volume fi'action fat 'thermal conductivity ElePhyData(l, 1) = ((0.602 "‘ VFm + 0.18 "‘ VFf+ 0.2 "' VFp)/ 100) ElePhyData(] , 2) = ((0.602 " VFm + 0.18 " VFf+ 0.2 "' VFp)/ 100) ElePhyData(l, 5) = heat_capacity " density 'density‘heat 'Coefficient for Mittal moisture diffusivity equation A = 0.003 ElePhyData(], 6) = 10000 * A " (0.003 * Exp(-0.442 * ElementAverageFP - 4829.7 (ElementAverageT + 273) + 11.55 "' moisture)) "‘ moisture_capacity * density 'moisture conductivity ElePhyData(l, 7) = 10000 * A * (0.003 * Exp(-0.442 * ElementAverageFP - 4829.7 (ElementAverageT + 273) + 11.55 * moisture)) "' moisture_capacity * density 'moisture conductivity ElePhyData(] , 10) = density "' moisture_capacity 'density‘moisture If ElementAverageT <= 45 Then ElePhyData(], 11) = 0 'fat diffusivity r-direction zero below ElePhyData( 1 , 12) = 0 'fat diffusivity z-direction zero below Else ElePhyData(], 11) = 0.00125 'fat diffusivity r-direction ElePhyData(l , 12) = 0.00125 'fat diffusivity z-direction End If ElePhyData(l, 15) = 1 'Variable not used for fat trasfer equations 243 Next I End Sub 244 Module — CalculateElementAverage Sub CalculateElementAverage(Volume(), NumEle, EleNodeData(), Phi(), WeightedAverage) This module calculates a weighted volume average of the nodal values. ReDim VolumeRatio(NumEle) ReDim ElementAverage(NumEle) WeightedAverage = 0 For K = 1 To NumEle TotalVolume = TotalVolume + Volume(KK) 'Calculate the element average of the nodal values If (EleNodeData(KK, 4) = 0) Then Triangular element Nodel = EleNodeData(KK, l) Node2 = EleNodeData(KK, 2) Node3 = EleNodeData(KK, 3) ElementAverage(KK) = (Phi(Nodel) + Phi(Node2) + Phi(Node3)) / 3 Else 'Rectangular element Nodel = EleNodeData(KK, 1) Node2 = EleNodeData(KK, 2) Node3 = EleNodeData(KK, 3) Node4 = EleNodeData(KK, 4) ElementAverage(KK) =(Phi(Node1) + Phi(Node2) + Phi(Node3) + Phi(Node4)) / 4 Endlf Next KK 'Calculate the volume weighted average For K = 1 To NumEle VolumeRatio(KK) = Volume(KK) / TotalVolume ElementAverage(KK) = ElementAverage(KK) " VolumeRatio(KK) Next K For I = 1 To NumEle WeightedAverage = Wei ghtedAverage + ElementAverage(I) Next I End Sub 245 Module — InputOvenConditions Sub InputOvenConditions(OvenConditions(), NumTimeSteps) 'Input oven conditions from file For I = 1 To NumTimeSteps Input #2, OvenConditions(I, 1), OvenConditions(I, 2), OvenConditions(I, 3) If OvenConditions(I, 2) = 0 Then OvenConditions(I, 2) = 1 Next I End Sub 246 Module - CalculatesurfaceFatContent Sub CalculatesurfaceFatContent(InitialF, T, GSV_F(), NudeyNodes, deNode(), temperature(), timestep) ’This soubroutine calculates the fat content at the surface of the meat as functions ' of time and temperature. Experimental fat holding data was utilized to determine ' the relationship between time, temperature and fat content. Fat content is expressed ' in terms of Fat/Protein ratio. 'Calculate FP at each boundary node For I = 1 To NudeyNodes X = 0.23 - (0.0127 "' temperamre(deNode(I))) + (0.0001 "‘ (temperamre(deNode(I)) A 2)) + (0.0617 "' InitialF) - (0.0007 * InitialF "‘ temperamre(deNode(I))) If X > GSV_F(deNode(I)) Then GSV_F(deNode(I)) = GSV_F(deNode(I)) Else GSV_F(deNode(I)) = X End If If GSV_F(deNode(I)) < 0 Then GSV_F(deNode(I)) = 0 Next I End Sub 247 Module - CalculateSurvivors Sub CalculateSurvivors(NumNodes, timestep, temperature(), Ninitial, No(), NW(), Nnew(), NnewW(), logreduction(), logreductionW, Dvalue, Tref, Z, TimeToLimit, TR) 'Calculate the number of surviving microorganisms for eanch node at each time step 'Log-linear inactivation equation For X = 1 To NumNodes If logreduction(X) < 9 Then 'limit solution to 9 log reduction in order to avoid ' computer overload from low numbers as well as because a '9-log reduction is well above the legal requirement If ((Tref - temperature(X)) / Z) > -2 Then (1 = Dvalue "‘ 10 A ((Tref - temperature(X)) / Z) Else d = 0.05 End If Nnew(X) = No(X) / (10 A (timestep / d)) If Nnew(X) < 1 Then Nnew(X) = l No(X) = Nnew(X) logreduction(X) = -Log(No(X) / Ninitial) / Log(10) Else logreduction(X) = 9 End If If logreduction(X) > 9 Then logreduction(X) = 9 End If 'Weibull inactivation equation If logreductionW < 9 Then b = 0.000000000011047 "‘ Exp(0.4l758 * temperature(l)) '0.03 * (temperature(l)) A 2 _ (2.7 " temperature(l)) + 72.19 11 = 1.12 NnewW(X) = NW(X) "‘ (10 A (-b "' ((timestep / 60) A n))) If Nnew(X) < 1 Then Nnew(X) = 1 NW(X) = NnewW(X) logreductionW = -Log(NW(X) / Ninitial) / Log(10) Else logreductionW = 9 End If Next X If logreduction( 1) <= TR Then TimeToLimit = TimeToLimit + timestep End If End Sub 248 Module — CalculateInitialProperties Sub CalculateInitialProperties(heat_capacity_f, Meat, Solid, InitialT, InitialM, InitialF, heat_capacity, latentW, frozenH, latentF) 'Calculate the initial thermal properties of the product before time-stepping. moisture = InitialM / 100 'Moisture, decimal wb Fat = InitialF / 100 'Fat content, decimal wb Protein = Solid / 100 'Protein content, decimal wb 'Beef If Meat = I Then If InitialT <= -2 Then heat_capacity_f = 1.9 "‘ moisture + 1.711 "' Protein + 1.298 "' Fat frozenH = (1.9 * moisture + 1.711 * Protein + 1.298 "‘ Fat) "‘ 271 'Enthalpy tied up 'in frozen meat latentW = 337.78 * moisture 'Latent heat of water latentF = 64.4 "' Fat 'Latent heat of fat Else heat_capacity_f = 1.9 * moisture + 1.711 * Protein + 1.298 * Fat heat_capacity = (4.18 "' moisture + 1.711 * Protein + 1.298 * Fat) frozenH = (1.9 * moisture + 1.711 * Protein + 1.298 " Fat) * 271 'Enthalpy tied up ‘in frozen meat latentW = 337.78 "' moisture 'Latent heat of water latentF = 64.4 "' Fat End If ’Pork Elself Meat = 2 Then If InitialT <= -2 Then heat_capacity_f = 1.9 * moisture + 1.711 "' Protein + 1.298 "' Fat frozenH = (1.9 "‘ moisture + 1.711 * Protein + 1.298 "' Fat) * 271 'Enthalpy tied up 'in frozen meat latentW = 337.78 * moisture 'Latent heat of water latentF = 64.4 "' Fat 'Latent heat of fat Else heat_capacity_f = 1.9 * moisture + 1.711 * Protein + 1.298 "' Fat heat_capacity = (4.18 * moisture + 1.711 * Protein + 1.298 * Fat) frozenH = (1.9 * moisture + 1.711 "' Protein + 1.298 "' Fat) * 271 'Enthalpy tied up 'in frozen meat latentW = 337.78 * moisture 'Latent heat of water latentF = 64.4 * Fat 'Latent heat of fat End If Turkey 249 Elself Meat = 3 Then If InitialT <= -2 Then heat_capacity_f = 1.9 " moisture + 1.711 "' Protein + 1.298 "' Fat frozenH = (1.9 "' moisture + 1.711 "' Protein + 1.298 "' Fat) * 271 'Enthalpy tied up 'in frozen meat latentW = 337.78 "‘ moisture 'Latent heat of water latentF = 64.4 * Fat 'Latent heat of fat Else heat_capacity_f = 1.9 * moisture + 1.711 "' Protein + 1.298 * Fat heat_capacity = (4.18 "' moisture + 1.711 ‘ Protein + 1.298 "‘ Fat) frozenH = (1.9 * moisture + 1.711 "' Protein + 1.298 " Fat) * 271 'Enthalpy tied up 'in frozen meat latentW = 337.78 " moisture 'Latent heat of water latentF = 64.4 "' Fat 'Latent heat of fat End If End If End Sub 250 Module — Cachandwidth Sub Cachandwidth(NumEle, EleNodeData(), BandWidth) This module provided by Dr. Larry Segerlind - Michigan State University 'This subprogram evaluates the bandwidth for any group of elements. ' The subprogram assumes that triangular elements have as many data ' values as rectangular elements. The extra data values are zeros. 'Evaluate the bandwidth NumEleNodes = 4 MaxDiff = 0 For I = 1 To NumEle For J = 1 To (NumEleNodes - 1) JJ = EleNodeData(l, J) If (JJ = 0) Then Exit For For K = (J + 1) To NumEleNodes K = EleNodeData(I, K) If (K = 0) Then Exit For Diff = Abs(JJ - KK) If (Diff > MaxDiff) Then MaxDiff = Diff Element = I Endlf Next K Next J Next I BandWidth = (MaxDiff + 1) End Sub 251 Module — DecompBandMatrix Sub DecompBandMatrix(NumNodes, BandWidth, GAM()) 'This module provided by Dr. Larry Segerlind - Michigan State University This subprogram decomposes a symmetric banded matrix into an upper ' triangular form using the method of Gaussian elemination. The ' matrix is stored in the rectangular array GSM(NumNodalVal%,BandWidth%). ' Only the upper part of the banded matrix is stored. 'Decompose the global stiffness matrix stored in a rectangular format NumNodalVal = NumNodes * 1 For I = 1 To (NumNodalVal - 1) MJ = I + BandWidth - 1 If (MJ > NumNodalVal) Then MJ = NumNodalVal End If NJ = I + l MK = BandWidth If ((NumNodalVal - I + l) < BandWidth) Then MK = NumNodalVal - I +1 End If ND = 0 ForJ=NJToMJ MK=MK-1 ND=ND+1 NL=ND+1 For K = 1 To MK NK = ND + K GAM(J, K) = GAM(J, K) - GAM(I, NL) * GAM(I, NK) / GAM(I, 1) Next K Next J Next 1 End Sub 252 Module - ZeroGlobalMatrices Sub ZeroGlobalMatrices(NumNodalVal, BandWidth, GFV(), GSM(), GFV_M(), GSM_M(), GFV_F(), GSM_F(), GCM(), GCM_M(), GCM_F()) 'This module modified from a module by Dr. Segerlind - Michigan State University 'This subprogram fills the global stiffness matrix and force vector ' with zero values. This subprogram assumes that the global stiffness ' matrix is symmetrical and stored in a rectangular format. For I = 1 To NumNodalVal GFV(I) = 0 GFV_M(I) = 0 GFV_F(I) = 0 For J = 1 To BandWidth GSM(I, J) = 0 GSM_M(I, J) = 0 GSM_F(I, J) = 0 GCM(I, J) = 0 GCM_M(I, J) = 0 GCM_F(I, J) = 0 Next J Next I End Sub 253 9 BIBLIOGRAPHY AMI. 2003. AMI Process Lethality Determination Spreadsheet. American Meat Institute. Arlington VA. AOAC. 2000. Oflicial Methods of Analysis of A OAC International. AOAC International. Arlington, VA. ASHRAE. 1998. ASHRAE Handbook: Fundamentals. American Society of Heating, Refrigerating, and Air-conditioning Engineers. Atlanta, GA. Badiani, A., Stipa, S., Bitossi, F ., Gatta, P.P., Vignola, G., and R. Chizzolini. 2002. Lipid composition, retention, and oxidation in fresh and completely trimmed beef muscles as affected by common culinary practices. Meat Science. 60:169-186. Barbut, S. 1996. Determining water and fat holding. Methods of Testing Protein Functionality. Blackie Academic and Professional. New York. Bejan, A. 1995. Convection Heat Transfer 2nd Ed. John Wiley and Sons. New York. Bengtsson, N.E., Jakobsson, B., and M. Dagerskog. 1976. Cooking of beef by oven roasting: a study of heat and mass transfer. Journal of Food Science. 41:1047- 1053. Bimbenet, J .J ., Loncin, M., and H. Brusset. 1971. Heat and mass transfer during air drying of solids. Canadian Journal of Chemical Engineering. 49:860-865. Bird, R.B., Stewart, W.E., and EN. Lightfoot. 1960. Transport Phenomena. John Wiley and Sons. New York Bodwell, CE. and P.E. McClain. 1978. Proteins. The Science of Meat and Meat Proteins. Food and Nutrition Press. Westport, CT. Chen, H., Marks, BR, and R.Y. Murphy. 1999. Modeling coupled heat and mass transfer for convection cooking of chicken patties. Journal of Food Engineering. 42: 139-146. Choi, Y., and M. Okos. 1985. Effects of temperature and composition on the thermal properties of foods. Food Engineering and Process Applications, Transport Phenomena vol. 1. Elsevier. London. Dagerskog, M. 1979 a. Pan frying of meat patties I. A study of heat and mass transfer. Lebensmittel- Wissenschaft and T echnologie. 12:217-224. 254 Dagerskog, M. 1979 b. Pan frying of meat patties 11. Influence of processing conditions on heat transfer, crust color formation, cooking losses, and sensory quality. Lebensmittel- Wissenschaft and T echnologie. 12:225-230. Dagerskog, M. and NE. Bengtsson. 1974. Pan frying of meat patties - relationship among crust formation, yield, composition, and processing condition. Lebensmittel- Wissenschaft and T echnologie. 72202-207. Datta, A. 2002. Biological and Bioenvironmental Heat and Mass Transfer. Marcel Dekker. New York. pp. 199-215. Dincer, I. 1996. Modeling of thermal and moisture diffusions in cylindrically shaped sausages during frying. Journal of Food Engineering. 28:35-43. Erdogdu, F., Balaban, M.O., and K.V. Chau. 1999. Mathematical model to predict yield loss of medium and large tiger shrimp (Penaues monodon) during cooking. Journal of Food Process Engineering. 22:383-394. Farkas, B.E., Singh, RP, and T.R. Rumsey. 1996 a. Modeling heat and mass transfer in immersion frying. 1., model development. Journal of Food Engineering. 29:211-226. Farkas, B.E., Singh, RP, and T.R. Rumsey. 1996 b. Modeling heat and mass transfer in immersion frying. 11, model solution and verification. Journal of Food Engineering. 29:227-248. Fellows, P. 1988. Food Processing Technology: Principals and Practice. Ellis Horwood. New York. FSIS. 1999. Performance standards for the production of certain meat and poultry products. Federal Register. January 6, 1999. 732-749. F SIS. 2001. Performance standards for the production of processed meat and poultry products; proposed rule. Federal Register. February 27, 2001. 12590-12636. F SIS. 2002. Use of microbial pathogen computer modeling in HACCP plans. FSIS Notice 55-02. December 2, 2002. Gardon, R., and J .C. Akfirat. 1966. Heat transfer characteristics of impinging two- dimensional air jets. Journal of Heat Transfer. 88:101-108. Geankopolis, C.J. 1993. Transport Processes and Unit Operations 3rd Ed. Prentice Hall. Englewood Cliffs, NJ. 255 Hallstrbm, B. 1990. Mass transport of water in foods - a consideration of the engineering aspects. Journal of Food Engineering. 12:45-52. Hayakawa, K. 1970. Experimental formulas for accurate estimation of transient temperature of food and their application to thermal process evaluation. Food Technology. 24:89- 100. Holtz, E. and C. Skjoldebrand. 1986. Simulation of the temperature of a meat loaf during the cooking process. Journal of Food Engineering. 5:109-121. Housova, J. and P. Topinka. 1985. Heat transfer during contact cooking of minced meat patties. Journal of Food Engineering. 4: 169-188. Huang, E. and GS. Mittal. 1995. meatball cooking - modeling and simulation. Journal of Food Engineering. 24:87-100. Hung, C.C., Davis, E.A., Gordon, J ., and HT. Davis. 1978. Mechanisms of water loss of bovine semitendinosus muscle dry cooked from the frozen state. Journal of Food Science. 43:1191-1 195. Ikediala, J .N., Correia, L.R., Fenton, GA, and N. Ben-Abdallah. 1996. Finite element modeling of heat transfer in meat patties during single-sided pan- frying. Journal of Food Science. 61:796-802. J uneja, V.K. and BS. Eblen. 1999. Predictive thermal inactivation model for Listeria monocytogenes with temperature, pH, NaCl, and sodium pyrophosphate as controlling factors. Journal of Food Protection. 62:986-993. Levy, FL. 1979. Enthalpy and specific heat of meat and fish in the freezing range. Journal of Food Technology. 14:549-560. Mallikarjunan, P., Hung, Y.C., and S. Gundavarapu. 1996. Modeling microwave cooking of cocktail shrimp. Journal of Food Process Engineering. 19:97-1 l 1. Maroulis, Z.B., Kiranoudis, C.T., and D. Marinos-Kouris. 1998. Heat and mass transfer modeling in air-drying foods. Journal of Food Engineering. 26:113- 130. Martin, H. 197 7. Heat and mass transfer between impinging gas jets and solid surfaces. Advances in Heat Transfer. 13:1-60. Mattick, K.L., Legan, J .D., Humphrey, T.J., and M. Peleg. 2001. Calculating Salmonella inactivation in non-isothermal heat treatments from isothermal nonlinear survival curves. Journal of Food Protection. 64:606-613. 256 McProud, L.M. and DB. Lund. 1983. Thermal properties of beef loaf produced in foodservice systems. Journal of Food Science. 48:677-680. Millsap, SC. and RP. Marks. 2002. Modeling condensing/convective heat transfer to food products in moist air impingement ovens. ASAE Paper No. 026045. Mannapperuma, J .D. and RF. Singh. 1989. A computer-aided method for the Prediction of properties and freezing/thawing times of foods. Journal of Food Engineering. 9:275-304. Mittal, GS. and J .L. Blaisdell. 1982. Moisture mobility in fiankfurter during thermal processing: analysis of moisture profile. Journal of Food Processing and Preservation. 6:1 11-126. Mittal, GS. and J .L. Blaisdell. 1984. Heat and mass transfer properties of meat emulsion. Lebensmittel- Wissenschaft and T echnologie. 17:94-98. Mittal, GS. and J. Zhang. 2000. Prediction of temperature and moisture content of frankfurters during thermal processing using neural network. Meat Science. 55:13-24. Mittal, GS. and J. Zhang. 2001. Artificial neural network for the prediction of temperature, moisture, and fat contents in meatballs during deep-fat frying. International Journal of Food Science and Technology. 36:489-497. Mittal, G.S., Blaisdell, J.L., and FL. Herum. 1982. Moisture mobility in meat emulsion during cooking. Journal of Food Technology. 17:709-717. Mittal, G.S., Blaisdell, J .L., and FL. Herum. 1983. Moisture mobility in meat emulsion during thermal processing: analysis of slab moisture profile. Meat Science. 8:15-32. Murphy, R.Y., Marks, B.P., and J .A. Marcy. 1998. Apparent specific heat of chicken breast patties and their constituent proteins by differential scanning calorimetry. Journal of Food Science. 63:88-91. Murphy, R.Y., Johnson, L.K., Duncan, L.K., Davis, M.D., Johnson, M.G., and J.A. Marcy. 2001a. Thermal inactivation of Salmonella spp. and Listeria innocua in chicken breast patties processed in a pilot-scale air-convection oven. Journal of Food Science. 66:734-741. Murphy, R.Y., Duncan, L.K., Johnson, ER, and MD. David. 2001b. Process lethality and product yield for chicken patties processed in a pilot-scale air- steam impingement oven. Journal of Food Protection. 64: 1549-1555. 257 Murphy, R.T., Duncan, L.K., Johnson, E.R., Davis, M.D., and J .A. Marcy. 2002. thermal inactivation of Salmonella seftenberg and Listeria innocua in beef/turkey blended patties cooked via fryer and/or air convection oven. Journal of Food Science. 67: 1879-1885. Murphy, R.Y., Duncan, L.K., Beard, B.L., and K.H. Driscoll. 2003. D and 2 values of Salmonella, Listeria innocua, and Listeria monocytogenes in fully cooked poultry products. Journal of Food Science. 68: 1443-1447. Ngadi, MD. and LR. Correia. 1995. Moisture diffusivity during deep-fat frying of chicken drum muscle. J. CSAE. 37:339-344. Ngadi, M.O., Watts, KC, and LR. Correia. 1997. Finite element method modeling Of moisture transfer in chicken drum during deep fat frying. Journal of Food Engineering. 32:1 1-20. Pan, Z. 1998. Predictive modeling and optimization of hamburger patty contact- cooking process. PhD. Dissertation. University of California Davis. Davis, CA. Pan, 2., Singh, RR, and T.R. Rumsey. 2000. Predictive modeling of contact heating process for cooking a hamburger patty. Journal of Food Engineering. 4629-19. Pan, Z. and RP. Singh. 2001. Physical and thermal properties of ground beef during cooking. Lebensmittel- Wissenschaft and T echnologi. 34:437-444. Peleg, M. and MB. Cole. 1998. Reinterpretation of microbial survival curves. Critical Reviews in Food Science. 38:353-380. Perez, M.G.R. and A. Calvelo. 1984. Modeling the thermal conductivity of cooked meat. Journal of Food Science. 49: 152-156. Roberts, T.A. 1997. Microbial grth and survival: developments in predictive modeling. Food Technology. 51:88-99. Sanz, P.D., Alonso, M.D., and RH. Mascheroni. 1987. Thermophysical properties of meat products: general bibliography and experimental values. Transactions of the ASAE. 30:283-296. Sarkin, RS. 1978. Computerized cooking simulation of meat products. Journal of Food Science. 43:1 140-1 143. Segerlind, L]. 1984. Applied Finite Element Analysis. John Wiley and Sons. New York. 258 Sereno, A.M. and G.L. Medeiros. 1990. A simplified model for the prediction of drying rates for foods. Journal of Food Engineering. 12:1-11. Shilton, N., Mallikarjunan, P., and P. Sheridan. 2002. Modeling of heat transfer and evaporative mass losses during the cooking of beef patties using far-infrared radiation. Journal of Food Engineering..55:217-222. Singh, RP. 2000. Moving boundaries in food engineering. Food Technology. 54:44- 53. Singh, N., Akins, R.G., and LE. Erickson. 1984. Modeling heat and mass transfer during the oven roasting of meat. Journal of Food Process Engineering. 7:205-220. Skala, D., Bastic, M., Bastic, L.J., Remberg, G., and J. Jovanovic. 1989. Thermal behavior of different hog and cattle tissue lipids by DSC analysis. Proceedings. 35th International Congress of Meat Science and Technology. Roskilde, Denmark. Danish Meat Research Institute. Skjoldebrand, C. 1980. Convection oven fiying: heat and mass transfer between air and product. Journal of Food Science. 45: 1354-1362. Skjoldebrand, C. and B. Hallstrom. 1980. Convection oven fiying: heat and mass transport in the product. Journal of Food Science. 45:1347-1353. Smith, SE, Maurer, J.L., Orta-Ramirez, A., Ryser, ET, and D.M. Smith. 2001. thermal inactivation of Salmonella spp., Salmonella typhimurium DT104, and Escherichia coli 0157:H7 in ground beef. Journal of Food Science. 66:1164- 1168. Stroshine, R. and D. Hamann. 1996. Physical Properties of Agricultural Materials And Food Products. Thorvaldsson, K. and C. Skjoldebrand. 1995. Water transport in meat during reheating. Journal of Food Engineering. 29: 13-21 . Tocci, A.M. and R.MH. Mascheroni. 1998. Characteristics of differential scanning calorimetry determination of therrnophysical properties of meats. Lebensmittel- Wissenschaft und T echnologie. 31:418-426. Tsai, S.J., Unklesbay, N., Unklesbay, K., and A. Clarke. 1998. Thermal properties of restructured beef products at different isothermal temperatures. Journal of Food Science. 63:481-484. USDA. 2002.Pathogen Modeling Program Version 5.1. United States Department of Agriculture 259 Van Impe, J .F ., Nicoli, B.M., Schellekens, M., Martens, T., and J. DeBaerdemaeker. 1995. Predictive microbiology in a dynamic environment: a system theory approach. International Journal of Food Microbiology. 25:227—249. Vijayan, J. and RP. Singh. 1997. Heat transfer during immersion frying of frozen foods. Journal of Food Engineering. 34:293-314. Voller, V. and M. Cross. 1981. Accurate solutions of moving boundary problems using the enthalpy method. International Journal of Heat and Mass Transfer. 24:545-556. Whiting, RC. 1995. Microbial modeling in foods. Critical Reviews in Food Science and Nutrition. 35:467-494. Whiting, RC. and R.L. Buchanan. 1994. Scientific status summary: microbial modeling. Food Technology. 48:1 13-120. Young, L.L., Garcia, J.M., Lillard, H.S., Lyon, CE, and CM. Papa. 1991. Fat content effects of yield, quality, and microbial characteristics of chicken patties. Journal of Food Science. 56:1527-1528, 1541. Zanoni, B., Peri, C., Garzaroli, C., and S. Pierucci. 1997. A dynamic mathematical model of the thermal inactivation of Enterococcusfaecium during bologna sausage cooking. Lebensmittel- Wissenschaft und T echnologie. 30:727-734. Zogzas, N.P., Maroulis, 2.8., and D. Marinos-Kouris. 1996. Moisture diffusivity data compilation in foodstuffs. Drying Technology. 14:2225-2253. 260