1"“ ”g ’ ’ mm millTlTIITIITTIT 3 1293 00620 73 ll LIBRARY Michigan State University This is to certify that the dissertation entitled A GENERALIZED VISCOSITY MODEL FOR THE COOKING EXTRUSION OF STARCH BASED PRODUCTS presented by Kevin Lewis Mackey has been accepted towards fulfillment of the requirements for PhD. Food Science degree in Date February 13, 1989 MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE EUE DATE DUE DATE DUE 9;... a \{K ‘T~r “Ow—— ‘ 44. q 8) "v "' I‘ ~9- A is"? \ no” M); 10 um I 1% it MSU Is An Affirmative AetiorVEquel Opportunity Institution A GENERALIZED VISCOSITY MODEL FOR THE COOKING EXTRUSION OF ST ARCH BASED PRODUCTS By Kevin Lewis Mackey A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Food Science and Human Nutrition 1989 ABSTRACT A GENERALIZED VISCOSITY MODEL FOR THE COOKING EXT RUSION OF STARCH BASED PRODUCTS By Kevin Lewis Mackey A generalized model has been developed for predicting the extrudate viscosity of low to intermediate moisture content starch based products during cooking extrusion. The model incorporates the effects of shear rate, temperature, moisture content, time- temperature history and main history. The model was tested using doughs of potato flour, corn starch and whole wheat flour. Equipment used included an Instron Capillary Rheometer attached to a Model 4202 Instron Universal Testing Machine and an APV Baker MPF 50 D/25 co-rotating twin screw extruder. Die lengths of the capillary rheometer were 6.35 x 103 m and 2.54 x 10" or while the diameters were 3.18 x 10'3 m and 1.59 x 10‘3m. Extruder dies of 6.35 x 10" m in length and diameters of 2.54 x 10‘2 m and 3.18 10'3 m were used. Experiments with potato flour were conducted at temperatures of 25, 50, 65, and 95°C. Cook times within the capillary rheometer were 0 to 12 minutes after compression. Moisture contents of 0.282, 0.507, and 0.772 g water per g solids (22.0%, 33.7%, and 43.6% wb, respectively) were used in the capillary rheometer, while the moisture con- tents in the extruder were 0.67 and 1.0 g water per g solids (40% and 50% wb). Shear rates ranged from 1-10000 s“. In potato dough, shear rate was described by the power law model. Time- temperature history and strain histOry did not influence viscosity. The final predictive model incorporates shear rate, temperature and moisture content and yielded an R2 of 0.95 1. Corn starch dough moisture contents were 0.359, 0.476, and 0.572 g water per g starch (26.4%, 32.0%, and 36.4% wb, respectively) for capillary rheometer tests and 0.5 and 0.6 g water per g starch (33.0% and 37.5% wb) for extrusion tests. In the capillary rheometer, barrel temperatures were held at 50, 55, 60, 75, 85, 95, and 110°C at cook times of 1, 2, 3, 6, 12, and 24 minutes. In corn starch dough, slip analysis was inconclusive. Shear rate was modelled by the equation proposed by Ofoli et a1. (1988). Viscosity was found to be a function of cook temperature and moisture content, but not cook time. Observed versus predicted viscosities gave an R2 of 0.95 after accounting for shear rate, temperature, moisture con- tents and time-temperature history in the capillary rheometer. Contrary to tests for porato doughs, extrusion tests indicated that strain history was important for highly puffed corn starch exu'udates (R’=0.79 before strain history con'ection). Strain history was modelled as a function of mechanical energy input (shaft work). After correction for strain history, the fit of the model improved to an n2 of 0.85.. Whole wheat flour tests were conducted at the same temperature and cook times as those for corn starch, but at moisture contents of 0.333, 0.337, 0.385, and 0.426 g water per g solids (25.0%, 25.2%, and 29.9% wb, respectively). As obseved for corn starch slip conection gave inconclusive results. Shear rate was best modelled by the Herschel- Bulkley model. Overall fit of the model steadily improved with corrections for tempera- ture, moisture content and time-temperature history. Viscosity of cooked doughs were found to be a function of moisture content, cook time and temperature. The fit (R’=0.56) was not as good for whole wheat flour as was observed for corn starch (R’=0.95) and potato flour (R2=0.95). The lower R2 may have been due to the presence of flour compo— nents such as bran, gluten (protein) viscoelastic effects and lipids which were not accounted for by the model. Incorporation of gelatinization kinetics alone in the time-temperature effects was inadequate to totally model the viscosity. ACKNOWLEDGMENTS I would like to thank the United States Department of Agriculture for their provid- ing me with a fellowship which enabled me to pursue my studies. Also many thanks to the Departments of Food Science and Human Nutrition and Agricultural Engineering for their financial and adminstrative support while I was working on my degree. Many thanks to Dr. Ronnie Morgan for his initial guidance, encouragement, and advice in deciding on a dissertation t0pic and selecting of my coursework and committee. Thank also to Drs. Edwards, Markakis, and Uebersax for their willingness to serve on my guidance committee and for the many fine suggestions they have provided. Special thanks to Dr. Jim Steffe for his help in teaching me more about rheology, for having confidence in my abilities as a scientist and teacher and for his advice on career choices. I would also like to thank him for his excellent editorial skills. To Dr. Bob Ofoli I give my heart felt thanks for his advice, for his willingness to become my advisor after the departure of Dr. Morgan, and for all the hard work he did in helping me prepare this dissertation. Finally, I wish to acknowledge the support of my family and friends for their encouragement. I especially want to thank Julie, my wife, for her love, support, and understanding during this long and sometimes difficult process. Without you and your love this dissertation would have been much more difficult. iv Preface This dissertation was written in an alternative format to aid the author in preparing portions of the dissertation for publication. The first three chapters feature an introduc- tion, literature review and the model development. Chapters 4, 5, and 6 consist of time papers to be submitted for publication while chapters 7 and 8 are the conclusions and suggestions for future research. I hope that writing the dissertation in this format will not confuse the reader and will, perhaps, allow the reader to understand and apply the infor- mation in an easier manner. Kevin L. Mackey December, 1988 - TABLE OF CONTENTS LIST OF TABLES ............................................................................................ ix ° LIST OF FIGURES .......................................................................................... x NOMENCLATURE .......................................................................................... xii CHAPTER 1 INTRODUCTION AND OBJECTIVES ......................................... 1 J 1.1. Introduction ............................................................................................ (i, 1.2. Objectives. ............................................................................................... 2 CHAPTER 2 LITERATURE REVIEW ................................................................. 3 2.1. Starch ........................................................................................................ 3 v 2.1.1. Introduction ................................................................................ (3) 2.1.2. Structure of the starch granule ........................................................ 6 2.1.2.1. External structure of the starch granule. .................... 6 2.1.2.2. Internal structure of the starch granule. .................................... 6 2.1.3. Physical and Chemical Changes ...................................................... 7 2.1.3.1. Physical changes ..................................................................... 8 2.1.3.2. Chemical changes .................................................................... 10 2.1.4. Measurement Methods .................................................................... 13 2.1.4.1. Light Microscopy .................................................................... 13 2.1.4.2. Polarized Light Microscopy .................................................... 14 2.1.4.3. Electron Microscopy ............................................................... 15 2.1.4.3.1. Scanning Electron Microscopy (SEM) ........................... 15 2.1.4.3.2. Transmission Electron Microsc0py (TEM) .................... 16 2.1.4.4. Viscometers ............................................................................. 16 2.1.4.5. Thermal Analysis .................................................................... 17 2.1.4.5.1. Differential Scanning Calorimetry (DSC) ...................... l8 2.1.4.5.2. Thermal Mechanical Analysis (TMA) ............................ 18 2.1.4.6. Others ...................................................................................... 19 2.1.4.6.1. Enzyme ........................................................................... 19 2.1.4.6.2. X-ray diffraction ............................................................. 19 vi 2.1.4.6.2. Nuclear Magnetic Resonance ......................................... 20 2. 2. Rheology ................................................................................................... 21 / 2.2.1. introduction ..................................................................................... .2112 2. 2. 2. Models For Rheological Modelling of Food .................................... 21 2.3. Extrusion .................................................................................................. 25 - V 2.3.1. introduction ...................................................................................... <25; 2.3.2. Viscosity Models Used in Extrusion ............................................... 27 CHAPTER 3 MODEL DEVELOPMENT ............................................................. 29 3.1. Effect of Shear Rate on Viscosity ............................................................. 29 3.2. Effect of Temperature .............................................................................. 30 3.3. Effect of Moisture .................................................................................... 31 3.4. Efi‘ect of Time-Temperature History ....................................................... 31 3.5. Effect of Strain History ............................................................................ 36 3.6. A procedure for determining the parameters in Eq.[3.20]. ....................... 40 CHAPTER 4 EXTRUSION MODELLING OF PREGELATINIZED POTATO FLOUR ................................................................................................................... 42 4.1. Abstract .................................................................................................... 42 v4.2. Introduction .............................................................................................. <4L 4.3. Model Development ................................................................................. 45 4.4. Materials and Methods .............................................................................. 47 4.5. Results and Discussion ............................................................................. 50 4.6. Conclusions .............................................................................................. 57 4.7. Nomenclature ............................................................................................ 58 CHAPTER 5 EXTRUSION MODELING or CORN STARCH .......................... 59 5.1. Abstract ..................................................................................................... 5% 1/ 5.2. Introduction ............................................................................................... 5 L602 5.3. Model Development .............................................................. , ................... 62 5.4. Materials and Methods ............................................................................. 65 5.5. Results and Discussion ............................................................................. 71 5.6. Conclusions .............................................................................................. 92 5.7. Nomenclature ............................................................................................ 94 CHAPTER 6 A RHEOLOGICAL MODEL FOR NATIVE WHOLE WHEAT FLOUR DOUGH EXTRUDATES ........................................................................ 97 6.1. Abstract ................... 9‘1 \/6. 2. Introduction ............................................................................................... Q98 : 6. 3. Materials and Methods ............................................................................. 6. 4. Results and Discussion .............................................................................. 105 6.5. Conclusions .............................................................................................. 119 6.6. Nomenclature ........................................................................................... 121 vii CHAPTER 7 SUMMARY AND CONCLUSIONS ............................................... 123 7.1. Overall Summary ..................................................................................... 123 7.2. Conclusions ............................................................................................... 126 CHAPTER 8 SUGGESTIONS FOR FUTURE RESEARCH ............................... 127 8.1. Effects of non-starch ingredients. ................ ’. ........................................... 127 8.2. Development of methods to quantify time-temperature history effects. .. 128 8.3. Development of equipment for measuring strain history. ........................ 129 8.4. Investigation of slip or friction within the capillary die. .......................... 129 LITERATURE CITED ........................................................................................... 130 APPENDICES ........................................................................................................ 138 Appendix 1. Temperatme, moisture, cook time, shear stress, shear rate, and observed viscosity of potato flour doughs. ........................................................ 138 Appendix 2. Moisture content, temperaturre, cook time, die LID, shear stress, and shear rate of corn starch .................................................................... 140 Appendix 3. Temperature, moisture content, cook time, shear stress, shear rate, observed viscosity of whole wheat flour doughs. ..................................... 169 viii LIST OF TABLES Table 2.1. Properties of whole granular starches. .................................................. 5 Table 2.2.1. Rheological models and their parameters used to characterize fluid food behavior. ......................................................................................................... 23 Table 3.1. Model parameters and what they describe. .......................................... 39 Table 3.2. Flow chart for determining the model parameters. .............................. 41 Table 4.1. Screw configurations of APV-Baker twin- screw extruder. ................. 49 Table 4. 2. Summary of model parameters and comparison with values from the literature. ................................................................................................................. 56 Table 5.1. Screw configuration for extrusion test on APV Baker 50 mm twin screw extruder. ....................................................................................................... 70 Table 5.2. Temperature andmoisture parameters as determined by linear regres- sion with viscosity measured at a shear rate of 100 s" ............................................ 74 Table 5.3. Values of constants a and B as a function of temperature. ................... 80 Table 5.4. Stepwise regression of observed viscosity versus temperature, shear rate, moisture content and time-temperature history corrections for corn starch doughs. .................................................................................................................... 83 Table 5.5. Data from twin screw extrusion tests on corn starch doughs. ............... 85 Table 5.6. List of model parameters for Eq. 5.22. ................................................. 91 Table 6.1. Screw configuration for extrusion tests on an APV Baker 50 mm twin screw extruder. ........................................................................................................ 104 Table 6. 2. Temperature and moisture correction parameters for whole wheat flour at 100 s. ........................................................................................................ 106 Table 6.3. Results of stepwise regression for whole wheat flour doughs. ............ 115 Table 6.4. List of parameters for Eq. 6.1. .............................................................. 118 LIST OF FIGURES Figure 4.1. Corrected viscosity of potato flour doughs (adjusted to 35% MC and shear rate of 100 s ) versus reciprocal temperature. .............................................. 51 Figure 4. 2. Corrected viscosity of potato flour dough (adjusted to T=50 C and shear rate of 100 ) versus moisture content. .......................................................... 52 Figure 4.3. Viscosity versus shear rate for cook times of 0 to 12 minutes at 95 C. ............................................................................................................................ 53 Figure 4.4. Observed versus predicted viscosity via Eq. [4.5] for capillary rheometer (CR) and extruder dies ('1‘ SE) for potato flour doughs. ......................... 55 Figure 5.1. Capillary rheometer assembly. ............................................................. 66 Figure 5.2. Observed versus predicted viscosities of corn starch doughs after shear correction for all temperatures, moisture contents and time-temperature histories. .................................................................................................................. 72 Figure 5.3. Apparent observed and predicted viscosities of corn starch doughs versus shear rate at a moisture content of 0.476 g water/ g corn starch. ................ 73 Figure 5.4. Observed versus predicted viscosities, of corn starch doughs after shear rate and temperature corrections, for all moisture contents and time- temperature histories. ............................................................................................. 75 Figure 5.5. Observed versus predicted viscosities of corn starch doughs, after shear rate, temperature and moisture corrections, for all time-temperature histo- ries. ......................................................................................................................... 77 Figure 5.6. Average apparent viscosity versus cook time for cook temperatures of 75, 85, 95, and 110 C (MC=O.476 g/g db) for corn starch doughs. ................... 78 Figure 5.7. Observed versus predicted viscosities corrected for shear rate, temperature, moisture, and time-temperature history for corn starch doughs. ....... 82 Figure 5.8 Observed versus predicted viscosities corrected for shear rate, temperature, moisture, and time-temperature history in extruder dies for corn starch. ...................................................................................................................... 86 Figure 5.9. Observed viscosity at 150 s", moisture content of 0.476 g/g and temperature of 50 C for corn starch doughs. .......................................................... 89 Figure 5.10. Observed versus predicted viscosities corrected for shear rate, temperature, moisture, time-temperature, and strain history in extruder dies for corn starch. ............................................................................................................. 90 Figure 6.1. Capillary rheometer assembly. ............................................................ 101 Figure 6.2. Observed versus predicted viscosities corrected for shear rate for all temperatures, moisture contents and time-temperature history for whole wheat flour doughs. ........................................................................................................... 108 Figure 6.3. Observed versus predicted viscosities corrected for shear rate and temperature for all moisture contents and time-temperature history for whole wheat flour doughs. ................................................................................................ 109 Figure 6.4. Observed versus predicted viscosities corrected for shear rate, tem- perature, and moisture content for all time-temperature histories for whole wheat flour doughs. ........................................................................................................... 110 Figure 6.5. Observed versus predicted viscosities corrected for shear rate, temperature, moisture, and time-temperature history for whole wheat flour 1 3 doughs. ................................................................................................................... 1 Figure 6.6. Observed versus predicted viscosities corrected for shear rate, temperature, moisture, and time-temperature history in extruder dies for whole flour doughs. ........................................................................................................... 1 16 xi NOMENCLATURE A, A’, material constant describing effects of cook temperature and product moisture A: b Cp cn Cs gnu BF‘FFfoB‘ content on viscosity at \v = co, dimensionless rate constant of effect of moisture content on viscosity, dimensionless protein concentration, decimal dry basis capillary rheometer starch concentration, wet basis decimal rate constant of effect of strain history ((9) on viscosity, s capillary diameter, in mechanical energy supplied by the extruder, Jm" Planck’s constant, consistency coefficient, Pa 3 consistency coefficient (Mizrahi-Berk), Pa“ sIII reaction transmission coefficient, s" “K" Boltzman’s constant, transmission coefficient, capillary length, m power index (Mizrahi-Berk), dimensionless moisture content, dry basis decimal reference moisture content, dry basis, decimal power indices, dimensionless volumetric flow rate, m’ls universal gas constant, 1.987 cal/g mole capillary radius, or xii time, 3 temperature, 'K reference temperature, ‘K twin screw extruder material constant describing polymer entanglement, dimensionless material constant describing effect of strain history on viscosity at 41 = co, dimen- sionless material constant describing effects of temperature on gelatinization, dimension- less material exponent for describing effect of moisture content on protein denatura- tion, dimensionless ‘ shear rate, 3" shear rate at the wall, 3'1 free energy of activation, kcal/g mole energy of gelatinization, kcal/g mole pressure drop, Pa apparent viscosity, Pa 3 normalized apparent viscosity ratio, dimensionless time-temperature history, s 'K strain history, dimensionless shear stress, Pa shear stress at wall, Pa yield stress, Pa yield suess, Pa xiii K“ yield stress-related parameter (Mizrahi—Berk), Pa °" 11 Newtonian viscosity, Pa s ll. high shear limiting viscosity, Pa s 1].. change in apparent viscosity due to gelatinization at \v = on, Pa 3 no high shear limiting viscosity (Bingham plastic), Pa 8 Subscripts ofn pre y,T,MC ¢ iT -‘y,r,Mc i.T.MC.v i.T.MC.v.¢ apparent viscosity corrected for shear rate, Pa 3 contribution to viscosity by gelatinization predicted apparent viscosity, Pa 3 viscosity at \v = co, Pa 3 apparent viscosity before gelatinization, Pa 8 apparent viscosity at q) = «t, Pa s ' apparent viscosity conected for shear rate and temperature, Pa 3 apparent viscosity corrected for shear rate, temperture and moisture con- tent, Pa 3 apparent viscosity corrected for shear rate, temperture, moisture content and time-temperature history, Pa 8 apparent viscosity corrected for shear rate, temperture, moisture content, time- -temperature history and strain history, Pa 3 xiv CHAPTER 1 INTRODUCTION AND OBJECTIVES 1.1. Introduction. The predominant ingredient in extruded snack and ready to eat (RTE) cereals is starch (Harper, 1986). Many of these products do not use purified starch but rather deg- errrred cereal grits which also contain some protein, fat, and fiber. The starch granule can undergo many changes leading to increased molecular rearrangement during extrusion, including loss of crystallinity, degradation and formation of amylose lipid complexes. These changes almost always will significantly affect viscosity. Therefore, temperature, moisture, shear, and feed composition are important factors accounting for changes in starch structure during extrusion (Harper, 1986). Extrusion cooking of starch based foods is increasing each year due to the versatil- ity, high productivity, low production costs, high product quality, increased energy effi- ciency and lack of product effluent of extrusion processes (Harper, 1981). However, the full economic exploitation of extrusion is currently constrained by a lack of understanding of the efi'ects of ingredients and process conditions on product quality, and difficulties in process design and scale-up. A general viscosity model would be useful for process engineering analyses. Harper (1986) states that there is no single model that incorporates starch gelatinization, viscoelastic behavior, density and thermal properties as functions of time, temperature, shear rate, and moisture content. Modeling in cooking extrusion needs more research, since models can be used to simplify complex processes in a cooking extruder, predict changes in product quality, and obtain a basis for new product development (Janssen, 1986). Most models proposed so far are empirical and limited to observed experimental conditions. 1.2. Objectives. The overall goal of this research is to obtain an understanding of the interactions of shear rate, temperature, moisture content, time-temperature history and strain history on the viscosity of extruded starch-based doughs. To achieve this goal, the specific objec- tives of the study are to: 1. Develop a generalized viscosity model incorporating effects of time-temperature history, moisture content, temperature, strain history, and shear rate during extrusion of low moisture starch based foods. 2. Evaluate the model with data obtained from capillary viscometry and twin screw extrusion for three dough systems: corn starch, potato flour and wheat flour. 3. Assess the effects of starch gelatinization during extrusion on rheological prop- erties. CHAPTER 2 LITERATURE REVIEW 2.1 Starch 2.1.1. Introduction Starch is the major storage form of glucose (energy) in plants. In cereal grains, starch contents range from 60 to 70% of the weight of the grain (Hoseney, 1986). Besides being an excellent source of energy for plants and animals, starch also provides important physical properties to foods. Pr0perties that are most important to the food sci- entist include gelling of puddings, thickening of gravies, setting of cakes, breads and other baked goods, and as a source of glucose for many sweeteners. In higher plants (eg. cereal grains), starch is formed in the plastids. All starch is stored in the form of granules. Starch granules made in the chloroplasts of leaves are considered transitory because at nightfall, the granule is broken up and the glucose is transferred to other portions of the plant. Reserve starch granules are usually formed in amyloplasts, though starch can be formed in chloroplasts which have lost their lamellar structure and therefore begin to produce large amounts of storage starch (Shannon and Garwood, 1984). In general, one plastid produces one starch granule except for rice and oat plants which produce compound granules (Hoseney, 1986). If examined under polarized light, intact starch granules show a distinct birefrin- gence pattern consisting of a Maltese cross. The presence of birefringence is an indica- tion of a high degree of order within the granule. Gelatinization is often defined as the loss of birefringence. Starch granules are also made up of semicrystalline materials as indicated by the classical studies of Katz (1928) and his co-workers using X-ray diffraction. Katz (1928) showed that intact starch granules gave three types of X-ray diffraction patterns (A, B, or C) depending on the source. Cereal starches (i.e corn, wheat, barley) give the A pattern, tuberous starches (potato) give the B pattern and smooth starches (pea and bean) give the C pattern. The C pattern is thought to be a mixture of both A and B patterns. Gelatiniza- tion of starches result in a loss of crystallinity and the appearance of the V type X-ray diffraction. Starch consists mainly of two large polysacchanide molecules, amylose and amy- lopectin, consisting of long chains of glucose. Amylose is considered to be primarily a linear ar(1-4) glucan with a molecular weight of 150,000-1,000,000 depending upon the source. Early research indicated that amylose was linear, as determined by enzyme studies (fit-amylase) but later research indicates that the [i-amylase used contained some a-amylase and/or debranching enzymes. Present research indicates that some limited branching is present in the amylose molecule (Banks and Greenwood, 1975). Amylose content ranges from 20-30% among biological sources (Table 2.1.), however, there are genetic hybrids in which there is an increase or decrease of the amylose content available. Amylopectin differs from amylose in both its molecular weight and its highly branched structure. The amylopectin molecule is made up of glucose units with a—(1-4) glycosidic, a-(1-6) glycosidic or both a-(1-4) and or-(l-6) glycosidic bonds. The average chain length is 20-25 glucose units and random branches give this molecule its high molecular weight. Evidence is mounting that amylopectin is the principal crystalline component of the starch granule (Lineback, 1984). Birefiingence and X-ray diffraction patterns of waxy maize starches containing no amyloseare similar to those of normal maize starch. High amylose maize starches are less birefringent and show less crystallin- ity than normal starches. Lineback (1984) also suggests that amylopectin is responsible for the crystalline structure, because amylose is released into solution when starches are exposed to gelatinization temperatures while amylopectin is not. Table 2.1. Properties of whole granular starches; Source Barley Triticale Wheat Rye Oats Potato Corn Waxy maize Broad bean Sorghum Rice High- amylose maize Gelatinization Granule shape Granule Temperature range ('0 51-60 55-62 58-64 57-60 53-59 59-68 62-72 63-72 64-67 68-78 68-78 67-90 round or elliptical round lenticular or round round or lenticular polyhedral oval round or polyhedral round oval round polygonal round irregular Amylose Size Content (run) (%) 20-25 22 2-6 19 23-24 20-35 26(23-27) 2-10 28 27 3-10 23-24 40(15-100) 23 15 28 15(5-15) 1 30 24 35 25(23-28) 3-8 17-19 25 52 Source: Lineback (1984). 2.1.2.8tructure of the starch granule 2.1.2.1. External structure of the starch granule. Starch granules from different plant sources have different morphologies (Table 2.1). The morphology depends on the structure and biochemistry of the starch producing chloroplast or amyloplast (Badenhuizen, 1969). These differences enable trained person- nel to determine which plant produced the starch. The granule grows via apposition, i.e., addition of material to the outer layers. This growth gives the characteristic ring structures seen in potato starch. These growth rings can also be seen in other starches after acid treatment of the starch granules. Scanning electron microscopy indicates that undamaged, unmodified granules have relatively smooth surfaces free of pores, cracks or fissures (Whistler et a1. 1984). 2.1.2.2. Internal structure of the starch granule. While the use of optical and scanning electron microscopy has enabled researchers to determine the general surface morphology of starch granules, the internal structure has been much more difficult to determine. The development and use of transmission elec- tron microscOpy (TEM) has provided researchers with greater resolution. Acid treatment (7% HCl for 35 days) has shown corn and waxy maize starch granules to have similar well-defined lamellar structures with alternating concentric electron-dense and electron- transparent rings (Mussulman and Wagoner, 1968). Rings occur at irregular intervals of 1200-4000 A and suggest that the overall length of individual amylopectin molecules are of this length (Y amaguchi et a1. 1979). Yamaguchi et a1. (1979) proposed that a single amylopectin molecule starts at one growth ring and finishes at the next, with each mole- cule made up of many 70 A clusters. It is interesting to note that high amylose corn starches do not reveal ring structures after acid or enzyme treatment, but instead show a heterogeneous internal su'ucture. Hood (1982) states that the molecular structure of starch corresponds with the crys- talline structure of the starch, and that it is well accepted that starch molecules are radially oriented within the granule. Pairs of the outer chains of amylopectin can form double helices which also contribute to the crystalline properties (Kainuma and French, 1972). Kassenbeck (1978) used TEM and ultrathin sections of starch granules to present ftu'ther evidence that radial orientation of amylopectin molecules exists. He also pres- ented three possible types of molecular organization: (1) a radially aligned, fibrillar arrangement of amylose in ordered regions; (2) amylose in amorphous regions where the amylose was degraded by the preparative method and appears in the form of precipitates;, and (3) an arrangement of amylopectin in crystalline regions where it appears as periodi- cally ananged block-like areas in radial sections, with crystallites in tangential lamellae. French (1984) states that there is no sharp line separating the crystalline and amor- phous sections of the starch granules and that some or all of the starch molecular chains continuously change from one phase to another. He also states that there is no definite evidence that amylose contributes to the crystalline structtue of the starch granule. Amy- lose not contributing to crystallinity would explain why amylose molecules are seen out- side the granule early in the gelatinization process and amylopectin molecules are not. 2.1.3. Physical and Chemical Changes Increase in viscosity of dilute solutions of starch and water is attributed to the starch granule taking up water and a concurrent swelling of the granule (Hoseney, 1986). As the granule swells, starch molecules are leached out causing further increases in vis- cosity. Gelatinization minimally entails (1) loss of crystallinity of the granule as measured by loss of birefringence and its X-ray diffraction patterns; (2) an uptake of heat as the conformation of starch changes; and (3) hydration of starch accompanied with granule swelling (Donovan, 1979). A more detailed summary of gelatinization is given by Olkku and Rha (1978) and is presented as follows: tion: (1) Granules hydrate and swell to several times their original size. (2) Granules lose their birefringence. (.3) Clarity of the mixture increases. (4) Marked, rapid increase in consistency occurs and reaches a peak. (5) Linear molecules dissolve and difi‘use from ruptured granules. . (6) Mixture retrogrades to a paste-like mass or gel. Atwell et al. (1988) recently proposed the following definition of starch gelatiniza- Starch gelatinization is the collapse (disruption) of molecular orders within the starch granule manifested in irreversible changes in properties such as granular swelling, native crystallite melting, loss of birefringence, and starch solubilization. The point of initial gelatinization and the range over which it occurs is governed by starch concentration, method of observation, granule type, and heterogeneities within the granule population under observation. This definition best describes what is occurring during the process of gelatinization and will be used in this work. 2.1.3.1. Physical changes In the presence of sufficient water and at temperatures below gelatinization, water is slowly and reversibly taken up by the granule (Olkku and Rha, 1978). Once gelatiniza- tion temperatures are reached, the granules begin to swell and become distorted (Sterling, 1978; Christianson et al., 1982). The granule’s ability to polarize light and diffract X-rays is lost at this time (Sterling, 197 8; Hoseney, 1986; Olkku and Rha, 197 8; Dono- van, 1979). Sterling (1978) reports that a three to six fold increase in volume is seen when starch granules begin to gelatinize. This increase is confirmed by Christianson et al. (1982), where a three and a half fold increase in com starch granule size was seen. The starch granule swells in a tangential direction during the gelatinization process. When temperatures increase beyond gelatinization temperatures, the granule continues to swell and may increase up to 25-30 fold, followed by a gradual collapse (Sterling, 197 8). Christianson et a1. (1982) presented excellent SEM photographs of corn starch granules at various stages of gelatinization. As the temperature increases, the granules begin to swell radially (65'C), ridges are formed at the surface (67'C), the granules become more angular in structure (70'C) and, at higher temperatures, the granules melt into thin flat disks. Different sections of the granule appear to swell at different times, and Sterling (197 8) postulated that differences in molecular structure cause this phenom- enon. Christianson et al. (1982) confirmed this when the amorphous regions were seen to swell at lower temperatures than the regions of the more highly ordered portion of the granule. Along with this swelling, solubilization of starch molecules is occurring and by the time a temperature of 70'C has been attained, approximately 10% of the starch has left the corn starch granule (Christianson et al., 1982). The granules seem to pass a transition at 80‘C where melting or softening of the granule occurs (Christianson et al., 1982). Twenty per cent of the starch is soluble at 85 'C and the granule appears smooth and more fluid than what is seen at lower temperatures. Wheat starch granules show a more ordered swelling as described by both Bowles et al. (1980) and Miller et al. (1973). It is interesting that both Miller et aL (1973) (wheat starch) and Christianson et al. (1982) (corn starch) have reported that maximum viscosity of starch suspensions heated 10 in the presence of excess water occurs after the majority of granule swelling has stopped. This is an indication that granule swelling is not the only reason an increase in viscosity is seen. Christianson et al. (1982) also compared viscosity data and SEM micrographs of starch dispersions of varying concentration and exposrue to different temperatures. At ‘65'C and under low shear, granules swelled but retained birefringence and exhibited dila- tant behavior. When exposed to high shear, shear thinning occurred After loss of bire- fringence (67-70'C) softening of the granules occurred so that shear thinning occurred at all shear rates. In their review on the gelatinization of starch, Olkku and Rha (197 8) discussed the effects of various ingredients on gelatinization. Proteins are believed to inhibit escape of soluble portions from the granule but do not appear to affect swelling. The same conclu- sion is drawn for the effects of oils and surface active agents. Pentosans are thought to compete for water and therefore inhibit gelatinization and granule swelling. Addition of salt can increase the granule’s ability to swell, while addition of sucrose retards hydration of the starch granule (Olkku and Rha, 1978; Bean and Yamazaki, 1978). 2.1.3.2. Chemical changes Along with the physical changes occurring in the starch granule, the chemical structure is also changing during the gelatinization process. The inuamolecular hydrogen bonds are disrupted upon heating and the structural integrity of the granule is lost. Once disruption of hydrogen bonds occur, the amorphous regions of the granule are the first to hydrate (Hood, 1982). This swelling of the amorphous phase can contribute to the dis- ruption of crystalline regions by pulling molecules away from the crystals (French, 1984). Blanshard (1979) proposed that in addition to diffusion of water and swelling, a hydration-facilitated helix-coil transition which is a melting process also occurs. 11 Miller et al. (1973) showed that amylose leached from the granule during heating and formed an extra granular network which contributed to increased viscosity. Approx- imately 10% of the starch has been solubilized at 70'C and by 80'C, 20% of the starch has been solubilized (Christianson et al., 1982). Using proton magnetic resonance (PMR) Jaska (1971) found water mobility to reversibly decrease with increased temperature until gelatinization temperatures are attained. This decrease in mobility indicates adsorption of water to starch molecules. When gelatinization occurs, an increase in water mobilization occurs which indicates a transfer of soluble starch (about 90% of total) into solution (Jaska, 1971). Jaska (1971) concluded that some of the starch is in solution inside the granule before it leaves the granule and that soluble starch solubilizes over a narrow tem- perature range. Various methods of thermal analysis have been used to study starch gelatinization (Hoseney, 1984; Nakazawa et al. 1984; Wootton and Bamunuarachchi, 1979a; Wootton and Bamunuarachchi, 1979b; Stevens and Elton, 1971; Biliaderis et al., 1980; Ghiasi et al., 1982a; Donovan, 1979), the effects of low moisture conditions (Wootton and Bamu- nuarachchi, 1979; Takahashi, et al., 1982; Burt and Russell, 1983; Sweat et al. 1984; Eliasson, 1980; Collison and Chilton 1974; Biliaderis et al., 1986), protein concentration (Eliasson, 1983), lipid and lipid-like substances (Harbitz, 1982; Kugimiya et al., 1980; Ghiasi et al. 1982b), and solute concenuation on gelatinization (Ghiasi et al., 1983; Spies and Hoseney, 1982; Evans and Haisman, 1982; Oosten, 1982). These methods of ther- mal analysis include difi'erential thermal analysis, differential scanning calorimetry, and thermal mechanical analysis. When gelatinization of starch is measured by differential scanning calorimetry (DSC) under low moisture conditions, two endotherms are observed. When excess water is present, the DSC curve shows only one endotherm indicating that water content 12 directly affects gelatinization kinetics. Various models have been proposed for this phe- nomenon. Donovan (1979) suggests that upon hydration/swelling of the amorphous por- tions of the granule and due to coupling with crystallites, melting of the latter occurs as long as water is present in excess. He postulated that when the amount of water is insufficient for complete melting, the remaining crystallites melt at a higher temperature and thus support the second DSC peak. Another hypothesis for starch gelatinization at low to intermediate moisture con- tents is proposed by Evans and Haisman (1982). They propose that two endothermic peaks reflected two types of melting: granules containing less stable crystallinity melt first; upon melting, the polysacchanide chains absorb more water making less water available for the remaining ungelatinized granules, causing these granules to melt at higher temperatures. Biliaderis et a1. (1986) have proposed a third explanation for starch gelatinization under low to intermediate moisture conditions. They postulate that the DSC curve is not representative of the initial crystallite profile, but rather the composite thermal effect of several processes that occur simultaneously during heating: melting, annealing and crys- tallization. They proposed that the order-disorder phase transitions of starch-water mixes are analogous to those of semicrystalline synthetic polymers. The thermomechanical properties of starch polymers can be altered by the presence of small amounts of water as a plasticizer. In the presence of a plasticizer. the glass transition temperatures of the gra- nule’s amorphous portion are decreased which in turn facilitates melting or reorganiza- tion of the starch crystallite and the amylose-lipid complexes to occur at lower temperatures. ‘ Wootton and Bamunuarachchi (1979b) and Eliasson (1980) report a linear relation- ship between moisture content and enthalpy of gelatinization when measured with a DSC. This change in enthalpy may be caused by a change in enu'opy and not in the 13 energy of activation, in accordance with the proposed mechanism of gelatinization by Biliaderis et al. (1986). Eyring and Steam (1939) also report that the change in enthalpy seen at difierent moisture contents of protein materials was due to differences in entropy and not due to a change in activation energy. Eliasson (1980) also reports that the tem- perature of the first endotherm does not vary significantly with water content. Thermal analysis of starch-water mixtures with varying levels of solute concentra- tions (salt or sucrose) shows increased thermal transition temperatures and a narrowing of the gelatinization temperature range. It is believed that these constituents compete with starch for water and therefore make water unavailable for gelatinization (Ghiasi et al., 1983; Spies and Hoseney, 1982; Evans and Haisman, 1982; Oosten, 1982). Eliasson (1983) indicates that increased levels of gluten lowered the gelatinization enthalpy and increased the gelatinization temperature. This is similar to results reported by Ghiasi et al. (1982a). It is believed that prorein inhibits migration of water into the starch granule. Gelatinization of starch in the presence of lipids or surfactants indicate a lowering of gel- atinization temperatures and increase in gel strength (Harbitz, 1983; Kugimiya et al., 1980; Ghiasi et al. 1982b). This is most likely due to the formation of an amylose lipid complex. 2.1.4. Measurement Methods Many methods of measuring starch gelatinization are discussed in the literature. These methods include light microscopy, polarized light microscopy, electron micros- copy, viscosity changes, X-ray diffraction, thermal analysis, and enzymatic methods. 2.1.4.1. Light Microscopy The use of a microscope allows the researcher to observe swelling duration, degree of swelling, swollen granule integrity and size (Zobel, 1984). Freke (1971), Collison and Chilton (1974) and Bean and Yamazaki (1978) all utilized a microscope to study starch 14 gelatinization. Both Freke (1971) and Bean and Yamazaki (1978) used a microscope equipped with a stage heater and a camera loaded with high contrast black and white film, while Collison and Chilton (197 4) exposed the starch solution to heat and then observed the results under a microscope. When heating on a microscope, care must be taken to prevent dehydration. This can be done by sealing the cover slop with mineral oil (Bean and Yamazaki, 1978) or silicone grease (Freke, 1971). Starch concentration of less than 1% were used by Freke (1971) and Bean and Yamazaki (1973) to reduce the number of granules in the field of observation. Snyder (1984) discussed the use of vari- ous stains (iodine, methylene blue, etc.) to aid in the determination of granule morphol- ogy. Chlorazol violet R on heat treated starch was used by Collison and Chilton (1974) to aid in determining damaged starch granules. Bean and Yamazaki (1978) also measured granule swelling by photographing a scale and then superimposing this scale. 2.1.4.2. Polarized Light Microscopy As previously mentioned, ungelatinized starch granules exhibit birefringence (a Maltese cross) when viewed under polarized light. Use of a polarized microscope to determine gelatinization is described by Snyder (1984), Burt and Russell (1983), Ghiasi et al. (1982a), Bean and Yamazaki (1978), Lelievre (1973) and Miller et al. (1973). Two methods are discussed: (1) use of a stage heater with the microscope while taking pictures at selected temperature intervals (Bean and Yamazaki, 197 8; Lelievre, 1973) and (2) controlled heating of the starch to a desired temperature, rapid cooling and then observation under a microscope (Ghiasi et al. 1982a; Burt and Russell, 1983; Miller et al. 1973). Heating rates fi'om l'C per 30 min (Lelievre, 1973), to l'C per min (Bean and Yamazalci, 197 8) to lO'C per min (Burt and Russell, 1983, Ghiasi et al. 1982a) have been used. As in unpolarized high microscopy the attachment of a camera and use of high contrast black and white film aid in the analysis. 15 2.1.4.3. Electron Microscopy Use of electron microscopy to study changes in granule structure during gelatiniza— tion has advantages over light microscopy techniques. These advantages include a greater depth of field and a higher resolution (about 70 A). Two types of electron microscopy have been used in observing starch granules: (1) Scanning (SEM) and (2) Transmission (TEM). SEM has been used to observe surface changes during gelatiniza- tion and TEM has been utilized to help discern the internal structure of the starch granule. 2.1.4.3.1. Scanning Electron Microscopy (SEM) Preparation of the samples for study by SEM is very important The samples must be dry to enable coating of the surface with a heavy metal (usually a 60:40 mix of gold- palladium). Gelatiniud or partially gelatinized starch samples have been prepared in an amylograph (Holmes and Soeldner, 1981; Hill and Dronzak, 1973; Miller et al., 1973; Christianson et al., 1982) or similar instrument, or by gently stirring (100 RPM) in a water bath (V arriano-Marston et al., 1985). Freeze drying the samples after heat treatment has been used extensively; however, the methods of preparation vary widely. Some authors directly froze the starch pastes in liquid nitrogen (Holmes and Soeldner, 1981; Miller et al. 1973) while others centrifuged the paste and decanted the supernate (Christianson et al., 1982; Varriano-Marston et al., 1985). Christianson et al. (1982) washed these pastes several times to remove minor amounts of exudate. Freeze-drying methods also vary from freezing inside a round-bottom flask using an ethanol-dry ice bath (Christianson et al., 1982) and freezing in a -20'C freezer, to immersion in liquid nitrogen or liquid isopentane (V arriano-Marston et al., 1985). Varriano—Marston et a1. (1985) report that rapid freezing followed by rapid freeze drying caused more damage than slow or rapid freezing and slow freeze drying. 16 Upon drying, starch granules are then placed on double stick tape and coated with a gold or goldzpalladium (60:40) mix to a thickness of 200-300 A. The coated samples are then placed in the SEM and pictures are taken following the manufacturer’s instructions. 2.1.43.2. Transmission Electron Microscopy (TEM) While giving higher resolution capabilities, sample preparation for TEM is much more difficult and can yield artifacts resulting in atypical morphologies. This occrn's because the granules are not easily infiltrated by the embedding media (Lineback, 1984). Because of these difficulties, partial degradation of the granules using acid or enzymes, staining, or freeze fracturing have been used (Mussulman and Wagoner, 1968; Chabot et al. 1978; French, 1984). Stains that react specifically with certain components of the granule are called posi- tive stains while those that provide an electron-dense outline of an otherwise electron- transparent object are called negative stains. These stains must be of high electron density and they must be non-crystalline (phosphotungstic acid or uranyl acetate) in order to be usable. Freeze fracture consists of rapidly freezing the sample in liquid nitrogen and then impact fracturing and evaporation at low temperatures. Coating of freeze fractured sam- ples is done first with carbon and metal. The carbon and metal films are cleaned by destroying the original object by chromic acid. 2.1.4.4. Viscometers Gelatinization of starches and effects of different treatments on starch or the pres- ence of various food ingredients (eg. fats, protein, sugar, salt) on gelatinization can be measured by changes in viscosity. The use of the Brabender Viscoamylograph is a traditional method for measuring starch gelatinization. Steffe et a1. (1988) have pro- posed an alternative to the Brabender Viscoamylograph using a Brookfield viscometer 17 with a small sample adapter, mixer paddle, and two water baths. This alternate method gave similar cru'ves compared to the Viscoamylograph but required significantly lower sample sizes and data collections. Amylograph procedures are described extensively in the Amylograph Handbook (Shuey and Tipples, 1980). Briefly, the amylograph measures the torque required to bal- ance the viscosity that is deve10ped when a starch slurry is exposed to a programmed heating and cooling cycle. Depending on the type of starch, a known weight (eg. 40 g wheat, 15 g potato) is added to 500 ml of water and placed in the amylograph. The slurry is equilibrated to 50°C and then the heating cycle is started. The sample is heated at 1.5'C/min to 95'C, held at 95'C for 1 hour and then cooled at 1.5'Clmin to 50'C. From this curve, the gelatinization temperature, pasting (i.e. maximum viscosity) peaks, ability to withstand shear (shear thinning) and extent of setback or gel formation upon cooling can be quantified. Addition of carboxymethylcellulose to the starch slurry permits detection of initial gelatinization temperatures which are difficult to see in some starch slur-ries. Viscosity of starch pastes using standard concentric cylinder, cone and plate, and parallel plate geometries have been reported by DeKee et al. (1980), Christianson and Bagley (1983), Christianson and Bagley (1984), Evans and Haisman (1979) and many others. Generally, the starches are gelatinized in other apparatus and placed in the vis- cometer to measure the viscosity. 2.1.4.5. Thermal Analysis Use of thermal analysis to study gelatinization has become increasingly popular in the last ten years, primarily because small sample sizes are required and gelatinization temperatures and enthalpies of transition can be determined easily. 18 2.1.4.5.1. Differential Scanning Calorimetry (DSC) Differential scanning calorimetry (DSC) has been used by numerous authors to study the efl‘ects of moisture levels, protein content, salt (N aCl) content, sucrose content, and lipids or surfactants on gelatinization (Eliasson, 1983; Lund, 1983; Lelievre, 1976; Donovan, 1979; Nakazawa et al., 1984; Wootton and Bamunuarachchi, 1979a; Stevens and Elton, 1971; Biliaderis et al. 1980; Wootton and Bamunuarachchi, 1979b; Sweat et al., 1984; Eliasson, 1980; Biliaderis et al., 1986; Evans and Haisman, 1982; Ghiasi et al., 1982b; Spies and Hoseney, 1982; Kugimiya et al., 1980; Harbitz, 1983; Colonna and Mercier, 1985; Hoseney, 1984). Approximately 20 mg of a starch-water solution is placed in a tared DSC pan. The starch may be input dry and water added with a syringe, or a starch slurry may be used. The pans are then hermetically sealed and equilibration is allowed to occur, usually over one hour. A DSC scan is then performed using an empty pan as reference at the rate of 10‘C per min from 20'C to a maximum of 150'C. Temperatures greater than 150'C require a pressurized system because of the possibility of pan failure. Once the endo- therm is obtained, the peak corresponding to gelatinization (50-85'C) is then integrated to give the enthalpy of the transition. 2.1.4.5.2. Thermal Mechanical Analysis (TMA) Thermal mechanical analysis (TMA) measures the volume expansion of materials as they are heated at a certain rate. Biliarderis et al. (1986) used TMA to examine the expansion of 50% starch-water mixtures. Two hundred and fifty milligrams of rice starch were firmly packed in the TMA’s quartz cell and covered with a thin layer of sand- paraffin oil (5:1, w/w) and placed in the TMA. A heating rate of TC per min was used over a temperature range of 25 to 97 'C. Weight loss due to evaporation of water and l9 normalization to a standard weight were then performed using the TMA’s software. TMA indicated a two-stage swelling pattern, the first associated with the onset of gela- tinization, the second with the melting of starch crystals. 2.1.4.6. Others Other important, but relatively minor methods (in terms of use) have also been used to measure starch gelatinization. These include X- ray diffraction, Nuclear magnetic resonance (NMR), and enzyme digestibility. X-ray diffraction and NMR are used infre- quently because of the cost of equipment and difficulty of use. Enzyme digestibility can be used to quantify starch gelatinization, but it gives little information about the exact structural and physical changes occurring. 2.1.4.6.]. Enzyme Chiang and Johnson (1977a, 1977b) report a method used to quantify the amount of gelatinized starch in flour or starch containing foods. Based on the fact that digestion of starch is easily performed by the enzyme glucoamylase to form glucose, this method uses spectrophotomenic methods to determine the total and gelatinized starch. The method is as follows: 1) disperse 20 mg of the sample in a 50 ml centrifuge tube and 3-5 ml water, 2) add 25 ml of glucoamylase solution and incubate 30 minutes at 40’C, 3) add 2 ml of 25% nichloracetic acid to inactivate the enzyme and precipitate pro- teins, 4) centrifuge for 5 minutes (16000 X g’s), 5) place 0.5 ml in test tubes containing 4.5 ml 0-toluidine reagent, 6) boil for 10 minutes then cool, 7) add 5 ml of glacial acetic acid and measure absorbance at 630 nm. Other methods are reported by Zobel (1984) and will not be discussed here. 2.1.4.6.2. X-ray diffraction X-ray diffraction was first used by Katz (1928) to measure changes of starch from native to gelatinized to retrograded forms. Hellman et al. (1954) and Yang et al. (1985) 20 have also used X—ray diffraction to measure changes in the starch granules. By exposing starch granules and starch gels to X-rays, the researcher can determine the crystalline structure of the material. Yang et al. (1985) used small angle X-ray scattering to measure changes in birefringence of starch granules during heating, cooling and storage. French (1984) also lists many researchers who utilized X-ray diffraction and their findings about the native starch granules. 2.1.4.6.3. Nuclear Magnetic Resonance Nuclear magnetic resonance (NMR) measures starch gelatinization at a molecular level using either high-resolution or wide-line proton magnetic techniques. Jaska (1971) followed the hydration and solubilization of starches using wide-line techniques. Lelievre (1975) used pulsed NMR to measure water mobility during gelatinization. It was determined that mobility decreased as hydration of the starch occurred at the onset of gelatinization, and as heating progressed beyond gelatinization temperatmes, the starch chain mobility increased. Both of these studies suggest that gelatinization is a melting process. Actual procedures depend upon the equipment being used. 21 2.2. Rheology 2.2.1. Introduction Rheology is the study of deformation and flow of matter (New American Heritage Dictionary, 1982). Measurement of rheological properties for food products is important because they can be used to describe basic physical properties of the food material, there- fore eliminating subjective terms such as gumminess or crunchiness, and enabling better process and quality control. Szczesniak (1987) stresses that the measurement of rheological properties should be performed in an instrument where the nature and magni- tude of the acting forces and the sample dimensions are well known, and the results can be expressed in fundamental units. 2.2.2. Models For Rheological Modelling of Food In an ideal system, there would be just two types of materials: solids and liquids. The solids would follow Hooke’s law when stress is applied, and liquids would follow Newton’s law of constant viscosity. However, many food materials exhibit properties of both solids and liquids, making it diffith to measure rheological properties of foods. Foods can be either viscoelastic solids, elastico-viscous (elastic) liquids or inelastic flu- ids. Walters (1975) defines viscoelastic solids as materials that do not continually change their shape when subjected to stresses, and elastic liquids as those which change their shape continually when subjected to stress, irrespective of how small the stress may be. Prentice (1984) expands Walters’ (197 5) definitions by defining viscoelastic foods as those which behave as if they had a solid structure (i.e. elastic), but whose deformation is modified by viscous behavior; and elastico-viscous foods as those which show some kind of elastic behavior (such as recoil) when flow ceases. 22 There are many models used to describe viscosity of fluid and semi-fluid foods. The simplest model which describes viscosity is Newton’s law (Table 2.2.1). However, there are very few foods which exhibit Newtonian behavior (among them fluid milk, and dilute corn syrup solutions); therefore more complex models are necessary. Common rheological models which are used to describe viscosity of fluid foods are the power law or Ostwald-de Waele (Reiner, 1949), Casson (Casson, 1959), Bingham (Bingham, 1922) and Herschel-Bulkley (Herschel and Bulkley, 1926). Of these, the model that is most widely used is the power law model because of its simplicity and ease of parameter estimation. Simple log-log transformation of shear stress and shear rate data allows one to estimate the parameters via linear regression provided a value for yield stress is given. The power law model has some drawbacks, however. F‘u'st, as shear rate increases towards infinity, viscosity approaches a value of zero. A second drawback is that as shear rate approaches zero the viscosity becomes infinite. However, within these limitations the power law has been found to adequately describe many food products. Food products may exhibit a yield stress. Yield stress is defined as the minimum mess that must be applied to initiate flow. Barnes and Walters (1985) recently ques- tioned if there was a yield stress or whether it is the inability of the instrumentation to measure flow. However, over the short time frame a producer or consumer is concerned with, many foods do exhibit a yield stress (e g. Mayonnaise, ketchup, molten chocolate). The simplest model which includes a yield stress is the Bingham model (Bingham, 1922). This model assumes that once the yield stress has been exceeded. normal Newto- nian behavior occurs. Butter and margarine are examples of Bingham foods, but in gen- eral the Bingham model is of little use because very few foods behave this simply. 23 Table 2.2.1 Rheological models and their parameters used to characterize fluid food behavior (Ofoli et al., 1987). Model Shear Stress Apparent viscosity Newtonian o = ll 11 = p, Powerlaw 0:19?- n=KfYr-l Bingham o = on + a = "'1 Herschel- = ° Bulkley O O°+KY| =--+K'Y"l .5 .5 ° .5 2 ‘Y if?” °”=K«M “W n=tK.Mt*"K.t"°‘r’ Heinz- = ' a ' 1 Cl 5111 0” 024-0107) [[00]! n]' n = . -" + “'0 Y Ofoli “1 0 =c?+u.i" ,, '3: THU-0.1] +tt_"y"-"] 24 The Herschel-Bulkley (Herschel and Bulkley, 1926) model is used with a wide range of food products. The model is a generalized power law model, with a yield stress added. The same drawbacks that are observed with the power law model at high shear rates are also seen with the Herschel-Bulkley model. Measurement of the value of the yield stress generally presents difficulties. The most common method is to extend shear stress versus shear rate curves to zero shear rate. Another method is to estimate the value of the yield stress, utilizing some form of curve fitting computer program. Osorio and Steffe (1985) have proposed a back extrusion method to quantify the yield stress for fluid foods. Given the limitations, the Herschel- Bulkley model is useful in determining rheological properties of the food material. Another model which has found wide use in the confectionery chocolate industry is the Casson model (Casson, 1959). Casson developed the model for use with pigment suspensions. It is important to note that the drawbacks seen with other models (zero vis- cosity at infinite shear rates, and infinite viscosity at zero shear rates) do not apply to.the Casson model. However, due to the fixed exponent of all the terms, the equation may not accurately describe all sections of the shear stress versus shear rate curve. Another modi- fication of the Casson model is the Mizrahi-Berk model (Mizrahi and Berk, 1972). Ofoli et al. (1987 ) have proposed another model which allows for a yield stress, variable shear-thinning and limiting viscosity at high shear rates. Disadvantages are the mathematical complexity of the model and the difficulty in obtaining model parameters without sophisticated computer programs. 2.3. Extrusion 2.3.1. Introduction Ram or piston type extruders were used to stuff sausage casing and other processed meats early in the history of food processing. An early application of the single screw exn'uder was as a continuous pasta press in the 1930’s (Harper, 1981). Single screw extruders were developed in the 1940’s for the purpose of making puffed snacks from cereal flours or grits (Harper, 1986). By the late 1950’s, extrusion cooked pet food had replaced other methods of production (Harper, 1981). In the 1960’s, ready to eat (RTE) breakfast cereals and textured soy protein products were being produced in cooking extruders (Harper, 1981). Reasons for the increased use of cooking extruders are listed below and discussed in detail by Harper (1981): l. Versatility: many foods can be produced with the same or similar equipment. 2. High productivity: extruders provide a continuous processing system, having greater production capacity. 3. Low cost: labor and floor space requirements are reduced. 4. Product shapes: a wide variety of shapes are available which are not possible with other production methods. 5. High product quality: cooking extrusion is basically a high temperature short time process. 6. Energy efficient: extruders operate at lower moisture contents, reducing heat requirements for drying. 7. Production of new foods: modification of proteins, starches and other food materials can be achieved in exn'uders to produce new food products. 8. Low effluents: little or no waste is produced, therefore costly waste treatment is avoided. Two types of extruders are used by the food industry today. The one that has been in use the longest is the single screw extruder. In the last few years, twin screw extruders 26 have been used increasingly in the food indusn'y. While single screw extruders have the ability to manufacture a wide range of food products (eg. breakfast cereals, pet foods, puffed snacks, etc.), use of twin screw machines is increasing because of their expanded range of applications and operational capabilities (Harper, 1986). Single screw extruders consist of a flighted screw rotating in a closely fitting bar- rel. The barrel is usually equipped with heating/cooling jackets in order to add or remove heat to the product as it passes through the extruder. The raw material is fed into the screw and as the material is conveyed through the transition section, the screw flights and barrel wall work and mix the material. Through viscous energy dissipation, the food material is heated and cooked. In the metering section of the extruder, pressure is built up before the material exits the die(s). The screw is a single piece with varying compression ratios and pitches, or a splined shaft which allows screw elements of differing configurations to be used (Harper, 1986). The splined shaft design allows for more flexibility in terms of maintenance and product variation. Single screw extruders often require pro-conditioners or pre-mix equipment to mix the dry and liquid ingredients and allow an equilibrium to be reached before feeding the material into the extruder. Twin screw extruders, as their name implies, consist of two screws inside a barrel which can be heated or cooled. There are two types: co-rotating or counter rotating. The co-rotating type is the most widely used by the food industry. The screws feature many geometries and can be used in any number of configurations. Unlike single screw exu'uder elements, the screws and paddles are self-wiping; therefore, there is less build-up of material. The high degree of mixing in twin screw extruders make them well suited for use as heat exchangers for highly viscous food materials (Harper, 1986). 27 Twin screw exn'uders have a higher initial cost than single screw exn'uders. The overall operating costs may be less, however, due to lower floor space requirements, little or no need for premixing equipment, ability to run at lower moisture contents (lower post-process drying requirements), and greater flexibility in types of products made. These factors give twin screw extruders an advantage over single screw extruders. 2.3.2. Viscosity Models Used in Extrusion Several rheological models for extrusion of flour or protein doughs have been reported in the literature. One of the earliest rheological models was by Harper et al. (1971), who modeled the viscosity as a function of inverse temperature, moisture, and shear rate as described by the power law model. Bhattacharya and Hanna (1986a) proposed an empirical model derived statistically for corn gluten meal and soy protein concentrates. Dough viscosity Was modeled as a function of moisture content and shear rate; temperature, time-temperature history and strain history were not incorporated. Therefore, the model is machine-and ingredient- specific and of limited general use. Effects of moisture content, barrel temperature, shear rate, residence time and shear strain on extrusion cooked waxy and non-waxy corn grits were studied by Bhattacharya and Hanna (1986b), who presented statistical models which indicated that moisture content and temperature are the most significant variables. How- ever, the authors stated that higher moistures reduce the degree of gelatinization, in con- trast to research results of several starch chemists studying starch gelatinization (Donovan, 1979; Eliasson, 1980; Biliarderis et al., 1986). The reduction in degree of gelatinization as moisture content increases is also in conflict with research by Chiang and Johnson (1977a) and Owusu-Ansah et al., (1983) both of which indicate that degree of gelatinization increases with moisture content. This 28 illustrates a problem with relying on statistical models for predicting viscosity changes in extrusion. Bhattacharya and Hanna (1986b) did n0t include strain or time-temperature history in their model. Jao et al. (1978) and Cervone and Harper (1978) also developed empirical models incorporating moisture content, temperature and shear rate, but time-temperature and shear history effects were not considered. Two models which included time-temperature history have been proposed by Rem- sen and Clark (1978) and Morgan et al. (1988). The model proposed by Remsen and Clark (197 8) included parameters for moisture content, shear rate and time-temperature history. Remsen and Clark (1978) did not study time-temperature history at low to inter- mediate moisttne concentrations, but assumed that relative increases in viscosity due to cooking were the same as for suspensions containing 70-75% moisture. Their model also exponentially approaches infinity for large time-temperature histories. This is in contrast to Morgan et al. (1988) who provided experimental data indicat- ing that protein dough viscosities approach some finite maximum value for time- temperature histories and shear rates encountered in extrusion. The model by Morgan et al. (1988) includes time-temperature history as well as shear rate, moisture content, and temperature. J anssen (1986) incorporates shear rate, temperature and time-temperature history in a model utilizing the power law relationship. However, moisture content and strain history were not incorporated. 29 CHAPTER 3 MODEL DEVELOPMENT 3.1. Effect of Shear Rate on Viscosity Many of the studies discussed previously use the power law model to characterize shear rate afi‘ects on the viscosity (Bhattacharya and Hanna 1986; Janssen 1986; Cervone and Harper 1978; Jao et al. 1978; and Remsen and Clark, 1978). While data shows good fit using the power law equation, the model itself has several drawbacks. The most important one is that it does not accommodate a yield stress. Other serious drawbacks are that it predicts unlimited decreasing apparent viscosity with increasing shear rate, and for a given shear thinning index it predicts zero or infinite viscosities at zero and infinite shear rates, respectively. Barnes and Walters (1985) contend that there is no such thing as a yield stress. Theoretically this may be true, but relative to the process time of most extrusion opera- tions, dough does not flow unless some minimum force is applied. Therefore, a yield stress may be defined as a material property in terms of the relative time frame of the process system involved. For example, cookie dough does not flatten out when put on a table before other quality parameters such as microbial acceptability are exceeded. Luxenburg et al. (1985) modeled soy doughs using Bingham plastic and Herschel- Bulkley models with good results. The Bingham plastic model does not allow for vari- able shear thinning index, and the Herschel-Bulkley model has a variable shear thinning index but approaches zero viscosity for large shear rates. This may present problems since shear rates encountered in extrusion may span 3 to 4 decades. 30 Morgan et al. (1988) used the Casson model to characterize soy protein doughs extruded in a capillary rheometer. The Casson model worked well, but it does not allow for a variable shear thinning index. Ofoli et al. (1987) have proposed a generalized rheo- logical model for the characterization of inelastic fluid foods. The model incorporates a variable shear thinning index, yield stress, and high shear limiting viscosity: L H .1 .. "o as": n-IITY] +lr_y ] [3.1] Advantages of this model are that it incorporates a yield stress, variable shear thin- ning indices and high-shear limiting viscosity. Equation [3.1] is the model used in this study to quantify shear rate effects. 3.2. Effect of Temperature Viscosity of Newtonian and non-Newtonian fluids are usually affected by tempera- ture. In general, as temperature increases, viscosity decreases. The relationship between viscosity and temperature can be represented by the equation of Glasstone et al. (1941): i(1"‘-T:') nr =nte ’ [3.2] where AB, is defined as the molar "free energy of activation" in a stationary fluid (Bird et al. 1960). The free energy of activation (AE,)is related to the amount of energy required for a molecule to escape its surroundings and move into an adjoining molecular site. Morgan et al. (1988) proposed that a shear rate dependency (such as shown in Eq. 3.1) could be combined with Eq. 3.2, if Metzner’s (1959) assumption that the shear thinning index is not affected by temperature is applied. In addition, one must assume that n2 is not afi‘ected by temperature. The combined equation is 31 I‘l “r i r_ r le.r=[[;;-] +u_,-an-~t] ash" 77) [3.3] 3.3. Effect of Moisture The effect of moisture content on viscosity has been described using a logarithmic mixing rule (Cervone and Harper, 1978; Harper et al., 1971; and Morgan et al., 1979). The logarithmic model is given by fluent/Wan) [3“] Morgan et al. (1988) combined Eqs. 3.3 and 3.4 to give a relationship which combines shear rate, temperature and moisture content effects on apparent viscosity: JI- ‘1 . _ T'T‘J'," +b(MC-MC,) “mac =[[%—:| +1.1.le H] 9 ( ) [3.5] 3.4. Effect of Time-Temperature History When starch and water are heated together, an increase in viscosity can be observed on an amylograph (Shuey and Tipples, 1980). The extent of this increase depends on the temperature and the time of exposure of the starch granule to that temper- ature. If starch granules are exposed to temperatures below the temperature required for gelatinization to start, no change in viscosity is seen until the threshold temperature is exceeded. Therefore incorporation of time-temperature history requires use of an activa- tion or threshold temperature. Increases in viscosity of dilute starch solutions is attributed to the starch granule taking up water and a concurrent swelling of the granule (Hoseney, 1986). As the gran- ule swells, starch molecules are leached out causing further increases in viscosity. Gomez and A guilera (1984) found that this classic gelatinization model of granule swelling and release of starch polymers was inadequate for high-shear extrusion cooking. 32 This was based on a previous work (Gomez and Aguilera, 1983) which found that, as extrusion moisture content decreased, "dextrinization" (i.e. shortening of polymer chains due to mechanical breakdown) appeared to become the predominant mechanism of starch degradation. They proposed a model that assumes the coexistence of three ptue states: raw, gelatinized and dexninized starch. In theory, the actual state of degradation is the sequence: raw -> gelatinized -> dextrr’nr’zed. A more complex model includes intermediate states of mechanically damaged granules, free polymers, and oligosaccharides and sugars. Different states are due to time-temperature history and shear-strain history in the extruder. When gelatinization of starch is measured by differential scanning calorimetry (DSC) under low moisture conditions, two endotherms are observed. When excess water is present, the DSC curve shows only one endotherm indicating that water content directly affects gelatinization kinetics. . Various models have been proposed for this phenomenon. Donovan (1979) sug- gests that, upon hydration and swelling of the amorphous portions of the granule and due to coupling with crystallites, melting of the latter occurs as long as water is present in excess. He postulated that when the amount of water is insufficient for complete melting the remaining crystallites melt at a higher temperature and thus support the second DSC peak. Another hypothesis for starch gelatinization at low to intermediate moisture con- tents is given by Evans and Haisman (1982). They proposed that two endothermic peaks reflected two types of melting: granules containing less stable crystallinity melt first, 33 and, upon melting, the polysacchanide chains absorb more water, making less water available for the remaining ungelatinized granules thus causing these granules to melt at higher temperatures. Biliaderis et al. (1986) have proposed a third explanation for starch gelatinization under low to intermediate moisture conditions. They provided evidence showing that the DSC curve is not representative of the initial crystallite profile, but rather the composite thermal effect of several processes that occur simultaneously during heating: melting, annealing and crystallization. They proposed that the order-disorder phase transitions of starch-water mixes are analogous to those of semicrystalline synthetic polymers. The thermomechanical properties of starch polymers can be altered by the presence of small amounts of water as a plasticizer. In the presence of a plasticizer. the glass transition temperatures of the granule’s amorphous portion are decreased which in turn facilitates melting or reorganization of the starch crystallite and the amylose-lipid complexes at lower temperatures. In this study, it is assumed that starch granules are analogous to semicrystalline synthetic polymers, as proposed by Biliaderis et al. (1986). Completely crystalline poly- mers follow first order kinetics for melting while purely glassy polymers follow second- order kinetics. A pseudo-first order kinetic model will be used in this study for time-temperature effects of starch gelatinization on the apparent viscosity. Harper et al. (197 8) and Morgan et al. (1979) used this approach with heat setting bovine plasma pro- tein suspensions and soy doughs, respectively. Pseudo first order polymerization assumes that concentration of one reactive spe- cies will remain constant and predicts disappearance of the monomer species. Starch gel- atinization is a much more complex reaction than first order; however, this simplification is used to approximate the "average-overall-viscosity" effect caused by gelatinization. Further development of this model is based on Gomez and Aguilera’s (1984) proposed 34 model. They assumd that increases in viscosity are due to the swelling granule and/or leaching of starch polymers outside the granule. This swelling and leaching is assumed to be similar to the viscosity increases caused by an increase in the "effective polymer molecular weight" during a plastic polymerization process (Morgan et al., 1988). Development of a time-temperature history function based on pseudo first order plastic polymerization processes was done by Morgan et al. (1988) for protein based doughs. Modification of this function for starch based doughs yields an equation for the effects of time-temperature history on viscosity due to gelatinization: na=B.(Cs)“(1—e"v)“ ‘ [3.61 Morgan et al. (1988) assumed that viscosity of a denaturing protein dough could be described by ' n =nt,r,nc+ne [371 Where “in“: represents the undenatured viscosity and no represents the increase in vis- cosity due to gelatinization of the starch granules. Use of Eq. 3.6 is only valid at constant temperature. In extrusion, the temperature will increase to a maximum with time and may decrease before exiting the die. There- fore, there is a need to incorporate variable time-temperature histories into Eq. 3.6. Mor- gan et al. (1979) proposed the use of an integrated time-temperature history, w, defined by 3 fl 3 “1:1. T(t)e" 0dr [3'8] 0 when T(t) is greater than the threshold temperature. Equation 3.8 is then related to Eq. 3.6 by k, = katv [3.9] 35 where k.=(k,k.,/h) and represents a reaction transmission coefficient for the material source, =Boltzman’s constant, k,=transmission coefficient, and h=Planck’s constant. Combining Eqs. 3.6 and 3.9 yields na=B.(Cs>°‘(t-e"-")° [3.101 which relates the change in viscosity to time-temperature history. Morgan et al. (1988) assumed that if the protein dough is held long enough at vari- ous temperatures, each treatment will approach the same viscosity. Kubota et al. (1979) studied the gelatinization rate of rice and potato starches and found the maximum viscosity to be a function of time. Similar results were found by Bakshi and Singh (1980) for gelatinization in rice kernels. Using differential scanning calorimetry (DSC) Lund and Wirakartusumah (1984) developed a model for starch gelatinization. Comparison of . enthalpy values versus cook time at different temperatures gave results similar to both Bakshi and Singh (1980) and Kubota et al. (1979). Kubota et al. (1979), Bakshi and . Singh (1980) and Lund and Wirakartusumah (1984) worked with high moisture systems. Dolan et al. (1988) found that viscosities collected at different time-temperature histories could be normalized by 11 " Tlo “"‘m-n. [3.111 It is important to note that all viscosities are specific to the cook temperature after either correction or normalization to the same temperature. In other words, the infinite viscosity term is the viscosity after infinite time at one cook temperature; if the cook tem- perature is changed, then a different infinite viscosity must be used. Dolan et al. (1988) also worked under high moisture (about 90%) conditions. 36 Dolan et al. (1988) stated that the change in viscosity caused by gelatinization is described by no =A(l - e""')° [3.12] Combining Eqs. 3.7 and 3.12 gives: “w = "in: 1 +A(l - e4‘VJa] [3.13] iftv -) co then Eq. 3.13 becomes: n- = “71.11%“ + A] [3.14] and therefore A becomes: A = n- - 1 [315] “tram: By combining Eqs. [3.5], and [3.13] we get a model that incorporates temperature, shear rate, moisture content, and time-temperature history: JI- ‘1 __ 1'1 r"-r," +bwc-uc, a fly,r,uc,y=[[1—;] +1.67”: 1"] e ‘ ( ) [1+A(l—e4‘v)] [3-15] 3.5. Effect of Strain History With increasing time of shear, the viscosity of dispersed starch solutions decreases. As starch granules swell they become more susceptible to mechanical degradation. There is much evidence that there is degradation of the starch granule during extrusion (Mercier, 1977; Colonna et al., 1984; Gomez and Aguilera, 1983; Gomez and Aguilera, 1984; Davidson et al., 1984; Diosady et al., 1985; Doublier et al., 1986). In each case the products of starch breakdown have been measured. Two different models for mechanical degradation of wheat starch have been proposed by Davidson et al. (1984) and Diosady 37 (1985). A serious drawback to both models is that they require the measurement of the extent of starch degradation by gel permeation chromatography (Davidson et al., 1984) or the degree of cook (Diosady, 1985). Anather drawback is that both techniques measure degradation and do not allow for predicting what the degradation may be. What is needed is an equation that can be used regardless of system geometries, conditions and raw materials. Pinto and Tadmor (1970) have proposed the following for quantifying total strain- history effects on transport properties of polymers: = [0‘th [3°17] Morgan et al. (1988) proposed a similar model : ' Viscosity will decrease as the value of 4) increases until a finite limiting viscosity is obtained. The value of the finite limiting viscosity depends on both the shear rates obtained within the extruder and the length of time at each shear rate. An equation describing strain history effects on viscosity was proposed by Morgan et al. (1988) and is given by n=1-B(1-e“’) [3.19] When this equation is incorporated into Eq 3.16, the final equation now incorporates shear rate, temperature, moisture content, time-temperature history, and strain history: 1. n “r , 1:, ”5..., ;3(r‘-r';‘)+bwc -Mc,) “ture.”: TY” l+llJY e [1+A(1_e*-*)°][1-B(1—e“’)] [3.20] 38 There are two important differences between Eq. [3.20] and the one proposed by Morgan et al. (1988): 1) the time-temperature history term has been modified to reflect the differences between protein denaturation and starch gelatinization; and 2) the general- ized Casson model was replaced with the model proposed by Ofoli et al. (1987) for describing shear rate effects. Table 3.1 lists the individual terms of the model. Limitations of the model include: 1. Difficulty in obtaining a yield stress value and the lack of an adequate non- linear regression computer program to determine the power indices and high shear limit- ing viscosity for the shear rate term. 2. The order in which one determines the moisture content and temperature effects on viscosity may yield different values for the respective parameters. 3. Determination of the moisture content and temperature effects above gelatiniza- tion temperatures may yield different numerical values than if quantified below gelatini- zation temperatures. 4. Evaluation of time-temperature history parameters is difficult if the heat transfer through the material is rapid; yet, this must be accomplished to allow comparison of data collected at different cook times and cook temperatures. 5. Strain history parameters may be functions of time-temperature history as well as shear rate and time. 39 Table 3.1. Model parameters and what they describe. Shear rate effects Temperature effects Moisture content effects Time-temperature history effects Strain history effects t " .- 11= 7‘] +1147” b(MC -uc,) e [1+A(1_ [5‘0“] [13(1 - e'“)l l. ‘r 40 3.6. A procedure for determining the parameters in Eq.[3.20]. Choose a reference temperature (below the gelatinization temperature) and mois- ture content and then collect shear stress versus shear rate data to obtain a relationship to describe the effects of shear rate on viscosity. Next, vary the temperature (again below gelatinization temperature) and hold the shear rate and moisture content constant to obtain the effects of temperature on viscosity. Vary moisture content at a temperature below the gelatinization temperature and quantify the effects of moisture on viscosity. Now vary both cook time and cook temper- ature and measure the effects of time-temperature history on viscosity. Finally, vary the strain history and quantify the effects of strain history on viscosity. A flow chart for determining the model parameters is shown in Table 3.2. 41 Table 3.2. Flow chart for determining model dependencies. Shear rate dependency Hold T=Tr, MC=MC,, \ll=0 and ¢=0. Collect viscosity versus shear rate data to determine shear rate effects. Establish 11 =f(y) See Eq. [3.1] Temperature dependency Hold at MC=MC,; \|I=0; and 41:0. . Vary temperature and collect viscosity data with f(y) given by [3.1]; Establish n = f(y, T) See Eq. [3.3] Moisture dependency Hold w=0 and ¢=0. . Vary moisture and collect viscosity data with f(y, T) given by [3.3] Establish n =f(y,T,MC) See Eq. [3.5] Time-temperature history dependency Vary cook time and temperature and collect viscosity data with f('°y, T, MC) given by [3.5] and ¢=o. Establish 1] =f(‘°y,T,MC,\|I) See Eq. [3.16] Strain history dependency Vary strain history and collect viscosity with fly, T,MC,\V) given by [3.16] Establish n =f(y,T,MC,\]I, 4)) See Eq. [3.20] 42 CHAPTER 4 EXTRUSION MODELLING OF PREGELATINIZED POTATO FLOUR 4.1. Abstract A generalized model is proposed for predicting the effects of shear rate, tempera- ture, moisture content, time-temperature history and strain history on the apparent viscos- ity of low to intermediate moisture starch-based doughs during cooking extrusion. The model was evaluated for potato flour doughs using an Instron Capillary Rheometer and a 50 mm APV Baker co-rotating twin screw extruder for all effects except strain history. The power law model adequately described shear rate effects in the range 10-10000 sec“. The generalized model fit observed data for temperatures of 25-95'C and moisture con- tents (wet basis) ranging from 22 to 50%. Time-temperature history variables were not quantified because the potato flour was pregelatinized. Strain history had no significant influence on the viscosity. 43 4.2. Introduction The predominant ingredient in most extruded snacks and ready to eat (RTE) cereals is starch (Harper, 1986). During cooking extrusion, starch granules undergo many changes leading to increased molecular rearrangement, including loss of crystallinity, degradation and formation of amylose lipid complexes. Changes in the starch granule during gelatinization may significantly affect viscosity. Therefore, the role of tempera- ture, moisture, shear, and feed ingredient composition are important factors involved in changing starch structure during cooking extrusion (Harper, 1986). Commercialization of extrusion processes is significantly constrained by lack of adequate scale-up information. Two important elements needed for improving food extrusion scale-up and design are effective process engineering analysis methods, and adequate rheological models. Currently, there is no single rheological model which incorporates starch gelatinization, viscoelastic behavior, density and thermal properties as functions of time, temperature, shear rate, and moisture content (Harper, 1986). Several extrusion models for flour and protein doughs have been reported in the lit- erature (Bhattacharya and Hanna, 1986; Cervone and Harper, 1978; Janssen, 1986; Jao et al., 197 8; Remsen and Clark, 197 8). However, only the model proposed by Remsen and Clark (197 8) incorporates effects of temperature, shear rate, time-temperature history and moisture content. More recently, Morgan et al. (1988) proposed a generalized viscosity model to describe the rheological changes which occur during the cooking extrusion of protein doughs. This work presents the most comprehensive rheological model yet. The mathe- matical relationship predicts viscosity as a function of temperature-time history, strain history, temperature, shear rate and moisture content. 44 Dolan et al. (1988) used the model by Morgan et al. (1988) as the basis for predict- ing changes in starch viscosity during the gelatinization of high moisture starch solutions. The model was found to predict changes in viscosity due to different time-temperature histories when shear rate, moisture content, temperature and strain history are held con- stant. While the changes that occur in the starch granule are different for high moisture systems than low moisture systems, it is important to note that time-temperature history is an important factor contributing to viscosity changes in both systems. The purpose of this study is to extend the work of Morgan et al. (1988) and Dolan et al. (1988) to develop a generalized viscosity model for low to intermediate moisture starch-based foods for use in extrusion. The model will incorporate the effects of time- temperature history, moisture content, temperature, shear-strain history, and shear rate. 45 4.3. Model Development The model presented by Morgan et al. (1988), with some modifications, was used to describe the viscosity of extruded starch based products. The power law model was found to be adequate to describe shear rate effects on the potato flour dough. The effect of time-temperature history on viscosity was modified to account for starch gelatinization kinetics. The resulting expression is 7] =1+A(1_ 12""): . [4.1] Incorporating the above changes, the generalized model of Morgan et al. (1988) for extrusion of protein doughs takes the following form for starch-based doughs: figure»... = [eéfi'hlf’fibwc '"CJ (“.Y)[1 +A(1 _ ft”)? [1 - B(1 - £45] [4.2] The first term incorporates the effects of temperature and moisture content on vis- cosity. It is important to note that AEJR is not a measure of gelatinization kinetics but rather of how temperature affects the flow of the material. Lubricating effects of water are described by the term b(MC-MC,). The K?) term represents any apparent shear rate dependent viscosity function (power law, Casson, Herschel-Bulkley, etc.) of the dough at the reference temperature and moisture content and with \v = 4) = 0. The next term describes the effects of gelatinization on the final viscosity of the starch dough. Time- temperature history (w) can range from zero for temperatures below gelatinization tem- peratures to infinity for either very high temperatures, long exposure time or a combination of the two. Also included are: the energy of activation for gelatinization (A5,) which may be affected by moisture content and or, an indicator of granule swelling and molecular entanglement. The final term incorporates strain history (it), which is a 46 measure of the effects of irreversible shear thinning. As strain history increases, the vis- cosity approaches a finite value (11.). If any of the variables is constant, the term that incorporates that variable reduces to unity. 47 4.4. Materials and Methods Potato flour (Lamb-Weston, Portland, Oregon) was mixed at room temperature with tap water to 25, 35, and 45% moisture (wb) in a large institutional kitchen mixer. The doughs were allowed to equilibrate at 7’C overnight in Ziploc Bags (Dow Chemical, Indianapolis, Indiana). Final moisture content of the doughs were determined by drying in a vacuum oven overnight at 70’C and 686 mm Hg. An Instron Capillary Rheometer and a Model 4202 Instron Universal Testing Machine (Instron Corp., Canton, Massachusetts) were used to measure the apparent vis- cosity of the doughs. Die lengths of 50.8 mm (2 in) and 6.35 m (1/4 in) and diameters of 0.49 m (1/8 in) and 1.59 mm (1/16 in) were used, giving L/D values ranging from 2 to 32. Two replicates for each plunger velocity and L/D were performed. Force versus plunger displacement curves were collected and force at the die entrance was calculated by extrapolation of the force versus displacement curves to the die as described by Bin?- hom and Turetzky (1964). Barrel drag was then subtracted from the conected force. A correction for entrance effects was made using the technique described by Bagley (1957). Shear rate and shear stress were then calculated using the Rabinowitsch equation (Whorlow, 1980) . di 7.. = 3Q, + ‘t,[ 11] [4.3] “R, dot” and the standard expression for shear stress at the wall of a capillary: _ APR, 4 r, — 2L [ -4] Slip analysis was performed using the method described by Darby (1976). Temperatures used in this study were 25, 50, 65, and 95'C. At 50 and 65'C, doughs were compressed in the capillary barrel and were held for 10 minutes. Moisture 48 contents used were 25, 35 , and 45%, wet weight basis. Tests were conducted at 50'C for 25% (wb) moisture samples. Cook times of 2, 4, 6 and 12 min at 95'C were performed on 35% (wb) moisture samples after compression. The 45% moisture samples were cooked for 4 and 12 minutes after compression. Effects of temperature (A5,) on viscosity was determined by linear regression of log viscosity versus inverse temperature; the effect of moisture content (b) on viscosity was determined by linear regression of log vis- cosity versus moisture content at constant temperature. Experimental extrusion tests were conducted using an APV Baker MPF 50D (APV Baker, lnc., Grand Rapids, Michigan) co-rotating twin screw extruder with the screw configuration shown in Table 4.1. Moisture contents were 40 and 50% (wb) and temper- atures at the die ranged from 40 to 75°C depending on exn'uder operating conditions. Feed rates were 1.26 x 10", 2.00 x 102, and 8.69 x 10'3 kg/s; and screw speeds used were 100, 220, and 350 RPM. Die diameter was 3.17 x 103 m and die lengths were 4.00 x 10’, 1.50 x 10'2 and 2.60 x 102 m. Pressure drop and extrudate temperature at the die were recorded two minutes after extruder operating conditions had been changed, to allow equilibrium conditions to be attained. Equilibrium conditions were assumed to exist when die pressure and barrel zone temperatures stabilized. Viscosities of the exn'udates were calculated by measuring the pressure drop at the die. Plotting pressure drop versus the three different die L/D’s with constant temperature, moisture, and mass flow rate allowed correction for end effects as described by Bagley (1957) for capillary dies. Shear rate was calculated using the Rabinowitsch equation (Whorlow, 1980) and shear stress was calculated using Eq. [4.4] for each die, tempera-8 ture, moisture and mass flow rate. Temperature and moisture correction on exu'uder data was performed using AE, and b estimated from the capillary rheometer. 49 Table 4.1. Screw configurations of MPF-50D/25 APV Baker twin screw extruder. Exn'uder L/D:15 Screw Configuration Length (cm) Screw Type 17.2 FS Feed Inlet 7.6 30F 7.6 FS 5.1 30F 2.5 45F 5.1 FS 6.2 30F 5.1 FS 5.1 30F 12.7 SL Extruder die Key to Notation: FS Feed Screw 30F 30 degrees Forwarding Paddles 45F 45 degrees Forwarding Paddles SL ' Single Lead screw 50 4.5. Results and Discussion There was apparent slip at some moisture contents, temperatures and shear rates as indicated by the presence of a "shark skin" on the extrudate. Slip analysis was conducted but results were not meaningful. Therefore, no further attempt was made to correct for slip. Inability to correct for slip may be due to a friction coefficient between the wall of the capillary die and the food material which could not be measured. Another cause may be due to a combination of both slip and friction or "slip-stick" where the material sticks until the friction coefficient is exceeded and then slips. The effect of temperatme on viscosity of 0.507 moisture samples is shown in Fig. [4.1]. An Arrhenius relationship was followed over the temperatures covered by these experiments. Over the temperatures encountered during extrusion, a linear relationship between viscosity an inverse temperature is usually observed. Actual moisture content of the 25 , 35 , and 45% doughs was determined to be 0.282 (22% wb), 0.507 (33.7% wb), and 0.772 (43.6% wb) g water per g potato flour, respec- tively. Moisture content effects at 50’C and 100 s" are shown in Fig. 4.2. As illustrated, a simple logarithmic relationship may not be adequate for broad ranges of moisture content. This is similar to the data for defatted soy flour adjusted to 95'C from Morgan et al. (1988). The power law model was used to describe shear rate effects for this study. Appar- ent viscosity versus shear rate for potato flour cooked at 95'C for 2, 4, 6, and 12 minutes are plotted in Figure 4.3. Data for 2, 4, and 6 minutes show similar slopes (flow behavior indices) and apparent viscosities. The 12 minute cook data exhibits more shear thinning. This is, in all likelihood, due to degradation of overcooked starch. Therefore, the time- temperature parameter \[I of Eq. [4.1] was set to infinity. In the process of making potato flour the potato starch is pregelatinized and therefore it would be expected that cooking 51 .239.an3 .3232... 382, film Deployed Loose ecu 0: Ron 3 “.3338 .50: 332. .3. 3.333 ooaootoo .5. 0.53... 7v. .oeauSanmfi omega. 385 38.0 88.0 286 - L L L L L [— b L b b L l.- L b L [D [1r 00F 02 “a“ mo o 0: «En mo n o: 5.? mo 4 62 non mme . 0 oz so... mm» - no.9 o w m. a p . m 10 on: a O m. o no.4 \ - .\.\. o d o \\m\ rodeo. M \\\\w a M\ rrvo+moé 52 .uchoo 9.3305 0383 film 9.: do 30.. .605 0:0 oomnp 0a 032.203 canoe .50: 330a .0 33033 0302.00 .m.¢ 239.“. x .02 .2380 8322.. md m.o to ,n.o «.0 mo _ a L L L e L a [— t _V 0: «Nu mo o 0: 5.8 mo n T o: 5.3 m0 4 f O o: Roe mme o ... m 0: «on mm; .. to. m 0 a m. . 0. He. m 3 I 4 r m . «A. race. moo . _ S w¢o+m_ 53 .o no «0 «3:58 a. 3 o e0 35: x000 no. 30.. 50cm 523 20:00 .50: 330a .0 35035 .n.¢ 0.59.“. Tm .301 nomcm ¢o+up 000— 00— OF = L b L F :30 0552 o 0 380 3:52 a n x000 3552 e 0 :30 05:5 m o u/ :30 05:3 up a row Too— fiooo. i¢o+up s—od ‘Krgsoosm 54 of potato flour should show an increase in viscosity. Comparison of the "coo " data with uncooked viscosities adjusted to the same temperature and moisture verifies that the potato flour is pregelatinized. Therefore, the parameter A in Eq. 4.1, which is an indica- tor of the relative contribution of gelatinization to viscosity, is zero. Strain history did not significantly affect the viscosity of the potato flour and therefore this parameter was also set to unity and excluded from further analysis. All Instron Capillary Rheometer and twin-screw extrusion data plotted were then fit into the final model . _1 fl:-"(r"‘.1';‘)+b(1trc-rrc,) nm=KY e Equation 4.5 gives an R2 of 0.951. Best fit parameter values are given in Table 4.2 along [45] with the physical description of each term. As indicated by Table 4.2 the values for AE, free energy of activation, b, and shear thinning index (n) are within the ranges observed by other researchers for other cereal and soy products. Any variation observed in fiee energy of activation (A3,) and b is probably due to individual product characteristics. The lower value for the power law index compared to pregelatinized corn flour is to be expected because native potato starch is highly shear thinning. The consistency coefficient is significantly higher than observed for other pregelatinized flours. The greater value of the consistency coefficient is most likely due to the hygroscopic nature of the potato flour which causes potato flour doughs to be extremely sticky at higher moisture contents. The means of the predicted versus observed data aresplotted in Fig. [4.4] for the capillary rheometer and twin-screw extruder data. Note that the twin-screw exu'uder data means fit with the same degree of accuracy as the capillary rheometer data 55 6.5500 .50: 3300 .2 EMS 00.0 50030.0 000 9.3 030802.... b05000 .0. anH— .cm 03 3.00003 030.00.... 0:0..0> 0020000 .04 0.52... mlod .350005 0020000 no.7: sawm— oo_o_ 2.: oh. _ 0: «ma mo o . _ oz 5.3 me n 02 ”8.? mo a H d o: No... mm» o a v3 a 02 x8 mme u o . me... a. O r 1. .. 9 v02 0. . m S r O n m . ,..A n m v¢o+mp _ r S 56 ture Table 4.2. Summary of model parameters and comparison with values from the litera- Physical Symbol Meaning This Study Data from literature (various food doughs at similar moisture and temperatures) AE, Aetivation Energy b Moisture coefficient 11 flow behavior index K consistency coefficient 8729 kcal per g mol 8.63 0.25 34903 (Pa 5") 4967; cooked cereal dough 8723; pregelatinized com flour 7300; soy pits 6900; defatted soy flour dough 6.7; defatted soy flour 7.9; com flour dough 0.19; soy pits 0.24; defatted soy flour 0.34; soy pits 0.36; pregelatinized corn flour 0.51; cooked cereal dough 4880; cooked cereal dough 4982; pregelatinized corn flour 16930; soy pits 28800; soy pits * Harper (1981) 57 4.6. Conclusions A model extending the work of Morgan et al. (1988) and Dolan et al. (1988) for use in extrusion cooking of low moisture starch-based doughs has been presented. This model incorporates shear rate, temperature, moisture content, time-temperature history and strain history. The model was tested with capillary rheometry and twin screw extru- sion of low moisture potato flour doughs. Overall, the results from the capillary rheometer and twin screw extruder are simi- lar over a wide range of shear rates, moisture contents, temperatures, time-temperature and strain histories. The capillary rheometer may, therefore, be used to approximate conditions within an extruder to obtain parameters to describe shear rate, temperature, moisture and time-temperature history. Specifically: l. The model was experimentally tested with potato flour doughs over a'wide range of experimental conditions: 25-95’C, 028-077 g water per g potato flour (22-44% wb), shear rate= 10-10000 sec", ut=0 to 00 and ¢=0 to w. 2. Due to the pregelatinization of potato flour, time-temperature and strain history effects were absent. Future research must focus on assessing the kinetic relationship of time- temperature history to starch gelatinization and strain history for ungelatinized starches, measurement of strain history parameters, and addressing slip or friction correction for capillary and extruder dies. 58 4.7. Nomenclature A relative amount of viscosity increase due to gelatinization, dimensionless b rate constant of effect of moisture content on viscosity, dimensionless (1 rate constant of effect of strain history (it) on viscosity, 3 D capillary diameter, m AE, fiee energy of activation, kcal/g mole K power law consistency coefficient, Pa s’n L capillary length, m MC moisture content, dry weight basis, decimal MC reference moisture content, dry weight basis, decimal n flow behavior index, dimensionless R universal gas constant, 1.987 cal/g mole t time, s T temperature, 'K Tr reference temperature, ’K T] apparent viscosity, Pa s AE, Energy of gelatinization, cal/g-mol n," predicated viscosity, Pa s y shear rate, s‘1 (1 strain history, dimensionless \V time-temperature history, 'Ks AP pressure drop, Pa 59 CHAPTER 5 EXTRUSION MODELING OF CORN STARCH 5.1. Abstract A generalized viscosity model for predicting the extrudate viscosity of low to inter- mdiate moisture content starch based products is proposed. The model incorporates the effects of shear rate, temperature, moisture content, time-temperature history and strain history. The model was tested using corn starch dough at various moisture contents. Equip- ment used included an Instron Capillary Rheometer attached to a Model 4202 Instron Universal Testing Machine and an APV Baker MPF 50 D/25 co-rotating twin screw extruder. Die lengths for the capillary rheometer were 6.35 x 10" m and 2.54 x 10'2 m, with a diameter of 3.18 x 10" m. Exu'uder dies of 6.35 x 10'3 m diameter, and 2.54 x 10’2 m and 3.18 x 10'3 m lengths were used. Dough moistlne contents were 0.359 (26.4% wb), 0.476 (32.0% wb), and 0.572 (36.0% wb) g water per g starch for capillary rheometer tests and 0.5 (32% wb) and 0.6 (38% wb) g water per g starch for the extrusion tests. In the capillary rheometer, barrel temperatures were maintained at 50, 55, 60, 75, 85, 95, and 110'C for cook times of 1, 2, 3, 6, 12, and 24 minutes. Shear rate effects were characterized by the model proposed by Ofoli et al. (1988). Viscosity was found to be a function of cook temperature and moisture content but not cook time. Observed versus predicted viscosity gave an R2 of 0.95 after accounting for shear rate, temperature, moisture content and time-temperature history in the capillary rheometer. Extrusion tests indicated strain history was important for highly puffed extru- dates (R’=0.79 before strain correction). Strain history was modelled as a function of mechanical energy input (shaft work). After correcting for strain history effects the fit of the model improved to an R2 of 0.85. 5.2. Introduction The major ingredient in many extruded products is starch (Harper, 1981). Over the last ten years, many research papers have addressed the mechanisms that affect materials during extrusion. Some have measured the effects of screw speed, temperature, moisture content and retention time on the molecular changes seen in the structure of the starch granule (Gomez and Aguilera, 1983; Mercier and Colonna, 1983; Colonna et al., 1984; Doublier et al., 1986). Others have proposed models to describe the physicochemical changes that occur in the starch granule during extrusion (Owusu-Ansah et al., 1983; Gomez and Aguilera, 1984; Davidson et al. 1984). Owusu-Ansah et al. (1983) measured physicochemical changes in corn starch as a function of temperature (100, 128, 170, 212, 240’C), feed moisture (11, 13.4, 17, 20.6, 23% wb) and screw speed (50, 58, 70, 82, 90 RPM), and reported that temperature, moisture and screw speed could significantly affect gelatiniza- tion. Response surface diapams indicated maximum gelatinization occrured at the low- est temperatme (100'C) and highest feed moisture (23%), while the minimum gelatinization occurred at the same temperature but lower feed moisture (11%) and 90 RPM. However, at lower screw speeds a minimum level of gelatinization could not be observed except at points of high moisture and temperature. This was in contrast to expected increases in gelatinization with increasing values of temperature and moisture. They concluded that the anomaly was caused by a high compression screw profile which generated peater mechanical work on the starch and therefore allowed complete gelatini- zation at lower temperatures. By using a lower compression screw profile at higher tem- peratures and moisture, peater gelatinization occurred. Gomez and Aguilera (1984) proposed a physicochemical model for extrusion of corn starch. This model is based on the combination of raw, gelatinized and dextrinized 61 starch. They concluded that the term "gelatinization" should not be used to describe the starch when cooked at moisture contents less than 20% because the starch becomes dex- ninized. Rheological models can be used by the food technologist and process engineer to predict product quality, and in scale-up of new products and in process control (Harper, 1981). Many rheological models for protein and starch based products have been pro- posed. One of the earliest proposed models (Harper et al. 1971) suggested that viscosity is a function of inverse temperature, moistlne content and shear rate. Cervone and Harper (197 8) proposed a model for pregelatinized corn flour which included tempera- ture, moisture and shear rate effects. A model incorporating shear rate, temperature and time-temperature history was used by Remsen and Clark (197 8) for defatted soy flours. Bhattacharya and Hanna (1986) presented an empirical model for corn gluten meal and soy protein concentrate as a function of shear rate and moisture content: Morgan et al. (1988) presented a model for protein dough viscosity as a function of shear rate, tempera- ture, moisture content, time-temperature history and strain history. Of these studies, only Morgan et al. (1988) included all of the variables Harper (1986) suggests are necessary for process engineering analyses. While Morgan et al. (1988) have proposed a model for extrusion cooking of protein doughs, there is no similar rheological model for use with starch based products. The purpose of this paper is to extend the model by Morgan et al. (1988) for pro- tein doughs to starch based products and to test the model with capillary rheometer and twin screw extruder data. 62 5.3. Model Development Morgan et al. (1988) presented the following model for cooking extrusion of pro- tein doughs. at 1:, ' 7'(r'-17')+b(uc—uc,) “have.” = [[7] + lit] 9 [1 + B.[As(MC)‘Cpl°[l — e“-*)“] [1 - Ba - e"')] - [5.1] Several modifications were made to Eq. [5.1] to make it suitable for use with starch based products. First, the shear rate characterization was changed from the Heinz-Casson model to the generalized model of Ofoli et al. (1987): .1 i t . - 11=[[—;] +ll.,y" "J [52} This generalized model allows for variable power indices, a yield stress and a high .— shear limiting viscosity. However, it should be pointed out that this model requires sophisticated regression propams to obtain estimates for the parameters. In its place, any acceptable viscosity model could be used. The time-temperature history expression also required modification. Morgan et al. (1988) assumed that if the protein dough is held long enough at various temperatures, each samme will approach the same viscosity. Kubota et al. (1979) studied the gelatini- zation rate of rice and potato starches and found the maximum viscosity to be a function of time and cook temperature. Similar results were found by Bakshi and Singh (1980) for gelatinization in rice kernels. Using differential scanning calorimetry (DSC), Lund and Wirakartusumah (1984) developed a model for starch gelatinization. By comparing enthalpy values versus cook time and temperature, similar results to those of Bakshi and 63 Singh (1980) and Kubota et al. (1979) were obtained. Kubota et al. (1979), Bakshi and Singh (1980) and Lund and Wirakartusumah (1984) each worked with high moisture sys- tems. Burros et al. (1987) used DSC to measure gelatinization of corn meal under low moisture (25%) and high temperature (80-140’C) conditions. They found results similar to those of Kubota et al. (1979). The order of the reaction was found to be 0.8 and it was believed to be less than one because of the decreased moisture content. They also indi- cated that the activation energy for gelatinization depends on moisture content, and increased with decreasing moisture content. Dolan et al. (1988), working with high moisture systems (about 90%), found that viscosities at different time-temperature histories could be normalized by 'I ”a ll“ -_ 5. no. “0 [ 3] It is important to note that the infinite and initial viscosities are specific to the cook temperature after either correction or normalization to the same temperature. In other words, the infinite viscosity is the viscosity after infinite time at one cook temperature; if the cook temperature is changed, then a different infinite viscosity is reached. Morgan et al. (1988) and Dolan et al. (1988) modeled the change in viscosity caused by denaturation and gelatinization by 7] = Tlgmrc + no [5 .4] Where 1’10 is defined by: no=n'.(1—e"")“ [5.5] Combining Eq. [5.4] and [5.5] gives: it = mm. +n2(1- e""')° [5.6] As \y -> on Eq. [5.6] becomes: “fining”; [5.71 Combining Eqs. [5.6], and [5.7], n,=n,mll +A(i —e"-*)“ ' [5.81 from which it follows that A = 11. ‘ [5.9] ‘1ng Finally, the modified Morgan et a1. (1988) model becomes "inn.” = H: '11]. + 11.7" Faye???‘ 47‘)».wa 46') Y 1 [1"‘(l-e4‘vlah1‘l3‘1'ed71 ~ . [5.10] and can now be used for predicting the viscosity of low moisture starch based products. 65 5.4. Materials and Methods Native corn starch (Melogel, National Starch and Chemical Corp., Bridgewater, NJ) at 8% moisture content (wb) was mixed in an institutional mixer with water to give sufficient quantities for all capillary rheometer tests to be performd. Moisture contents were 26, 32, and 36% (wb). These moisture contents were cho- sen because less than 26% moisture caused force readings to exceed the maximum force the capillary rheometer load cell could handle, and moisture contents peater than 36% caused water to be squeezed out of the material giving an uneven moisture distribution in the rheometer barrel. After mixing with distilled water, the samples were allowed to equilibrate overnight at room temperature (24'C) and then smaller samples were placed in polyethylene bags and stored at -10’C until capillary rheometer tests were performed. Twenty four hours before performing the tests, each sample was thawed at 4'C overnight. Prior to testing, each sample was allowed to come to room temperature. An Instron Capillary Rheometer and a Model 4202 Instron Universal Testing Machine (Instron Corp., Canton, Massachusetts) were used to measure apparent viscosity of the doughs (Figure 5.1). Die lengths of 2.54 cm (1 in) and 6.35 x 10'1 cm (1/4 in), and diameters of 4.90 x 102 cm (1/8 in) and 1.59 x 10'1 cm (1/16 in) were used, giving up ratios ranging from 2 to 16. Three replicates for each plunger velocity, temperature, moistlu'e content, cook time, and LID were performed. Force versus plunger displace- ment curves were collected and force at the die entrance was calculated by extrapolation of the force versus displacement curves to the die as described by Einhom and Turetzky (1964). Barrel drag was then subtracted from the corrected force. A correction for entrance effects was made using the technique described by Bagley (1957). Barrel wall > “ Plunger Sample 0 O 00 00.60.00 00 00000000....nl Capillary die Figure 5.1. Capillary rheometer assembly. 67 Shear rate and shear stress were calculated using the Rabinowitsch equation (Whorlow, 1980) (Eq. 5.11), (1.9.. +1"[-fl] [5.11] and the standard expression for shear stress at the wall of a capillary: it. = -APR° 512] tw— 2L [. Slip analysis was performed using the method described by Darby (1976) using dies of same length but different diameter. Temperatrues used ill this study were 50, 55, 65,75, 85, 95, and 110’C . At tem- peratures less than or equal to 60'C, the materials were compressed in the capillary bar- rel and allowed to equilibrate for 6 minutes. Cook times of 1, 2, 3, 6, 12 and 24 min at 75, 85, 95, and 110'C were performed on each sample after compression. Shear rate, temperature, moisture and time-temperature history parameters were determined as follows. Holding temperature (50'C), moisture content (32% wb) and time-temperature history (\|I =0) constant, viscosity versus shear rate were fitted to the model of Ofoli et al. (1987), using the Marquardt compromise method in the non-linear repession propam of SAS (SAS Institute, Cary, NC). Shear stress and shear rate values from the capillary rheometer were then averaged for each capillary die and Instron cross- head velocity. Temperature correction was obtained in the following way: holding shear rate, moisture content (32% wb) and time-temperature history constant (\[I =0), the natlu'al log of the viscosity versus (T‘-T,“) was repessed linearly to obtain AE,/R for temperature correction. The reference temperature, T,, was set at 323.15 'K. 68 The moisture correction parameter was determined by holding time-temperature history constant (\|I =0) and linearly repessing the natural log of the viscosity versus (MC-MG) with MC,=0.476 (db). Finally, the time-temperature history correction was performed. Strain history within the capillary rheometer was assumed to be insignificant and set to unity. Experimental extrusion tests were conducted using an APV Baker MPF 50D (APV Baker, Inc., Grand Rapids, MI) co-rotating twin screw exn'uder with the screw configura- tion shown in Table 5.1. Feed rates were 1.01 x 10’2 and 1.64 x 10'2 kg/s, at 200 and 400 RPM. The die diameter was 3.17 x 10'3 m, with lengths of 6.40 x 10'3 and 2.60 x 10'2 m. The pressure drop across the die and extrudate temperature at the die were recorded two minutes after extruder operating conditions had been changed, to allow equilibrium con- ditions to be attained. Plotting pressure drop versus the two different die L/D’s with constant temperature, moisture, and mass flow rate allowed correction for end effects as described by Bagley (1957) for capillary dies. Shear rate was calculated using the Rabinowitsch equation (Whorlow, 1980) and shear stress was calculated using Eq. [5.4] for each die, tempera- ture, moisture and mass flow rate. Temperature and moisture correction on the extrusion data were performed using AE, and b estimated from the capillary rheometer. The average residence time within the extruder was determined by adding 1 ml of Red 40 dye to the feed, timing how long it took for the dye to appear at the exit, collecting the colored material at intervals of 20 8 (low feed rate) or 15 s (high feed rate) until no dye came out, and recording the total time the dye remained in the extruder. The collected samples were dried, pound and Hunter 69 Lab color ’a’ values were determined for each time interval (a(t)). The mean residence time (t) within the extruder was calculated by _ ‘ a(t) t- o tZadt d: [5.13] 7O Table 5.1. Screw configuration for extrusion test on APV Baker 50 mm twin screw extruder. 17.8 cm 8.9 cm 7.6 cm 6.4 cm 10.2 cm 8.9 cm 7.6 cm 7.6 cm Feed Screw Forwarding paddles 30 offset Feed Screw Forwarding paddles 45 offset Single Lead Forwarding paddles 30 offset Feed Screw Single Lead Feed Inlet Extl'uder die 71 5.5. Results and Discussion Equilibrium moisture contents of the corn starch samples on a dry weight basis were determined to be 0.359, 0.476, and 0.572 g water per g corn starch. Moisture con- tents were found to be constant throughout storage. A non-linear repession analysis of ungelatinized corn starch at 0.476 g water per g corn starch (32% wb) and a temperature of SO'C gave the shear rate correction for all the data: I 0.979 573 n,=[[189;34] +69351«’y*’~‘”] [5.141 A repession coefficient (R2) of 0.996 indicates a good fit of the data. Shear stress and shear rate values were averaged at each capillary rheometer plunger velocity for each cook time, temperature, and moisture. When the average observed apparent viscosities were plotted against the average predicted values for all moisture contents, temperatures and time-temperature histories using Eq. [5.14], an R2 of 0.88 and a slope of 6.02 was . obtained. Figure 5.2 is a plot of the observed apparent viscosity versus the predicted apparent viscosity after accounting for shear rate effects. A plot of both predicted (cor- rected for shear) and observed viscosities versus shear rate can be seen in Figure [5.3]. If the shear rate correction alone were adequate, one would expect the predicted viscosity values to match the observed viscosity values. However, both Fig. [5.2] and Fig. [5.3] indicate that correcting for shear rate alone is not enough. Corrections for temperature and moisture were, therefore, carried out (Table 5.2). Both values for temperature (free energy of activation) and moisture (b) are within the range seen in the literature for similar materials, temperature and moisture. Figure 5.4 is the plat of the observed versus predicted apparent viscosity, corrected for temperature 72 .00_._30E 0nzuon0QE0Ll0E_u_ 0:0 00:00:00 0._30_0E 0003309050 :0 nos. 00:00:00 .6030 5:0 020300 £830 500 v6 00200003 0020000 0:0..0> 02200.5 .NAu. 059m 0 on .360005 002080 mo+mo._ mo+mo._ ¢o+mog 0.82 0.09 A L L F _ 33 02 0000.0 o 33 02 $2.0 n \ 33 0: $30 a .. 4\x mnmduam 00.80"» 060_ 10.000— l¢o+moé lmQ+mo.— loo+moé fih0+moé s 0:] ‘KrlsooslA perolperd 73 00.30 0.00 02303 a 050.0 .0 E3000 0.30.00. 0 30 .30. .0000 0:0.0> 000500 00.30 0.00 00 00200003 030.00... 0.5 0020000 300.009. 6.0 050.... T0 .33. 53% 00m— 2.: 0.? P elem $300003 03200.1 0 — 0l0n. 0300003 0020000 0 new .to— d D J 9 00— mw < .t M S 3 m. 0000- l. 0. we 4 fi¢0+m— .0 D [00+mp 74 Table 5.2. Temperature and moisture parameters as determined by linear repession with viscosity measured at a shear rate of 100 s-l. Parameter Estimate Range from Literature AB, 8473.14 ’K 3000-8000 'K b -399 -15 - -0.19 75 00:30:. 0.30.0QE3I0E: 0:0 3:00:00 0.30.0... =0 .00 .0:o:00..00 0.30.0083 0:0 30. .0000 .0000 000000 00.00 :.00 00 .00_:0000_> 0020000 000.0> 03200.0 .0...“ 0.090 0 01 .30.0005 0020000 0300.. mo+mo.. eo+mo.. 0.08. 0.00. 0.0. _ _ _ [r L O.— 33 0.... 0000.0 o 33 0... $2.... a 33 0... 090.0 a 10.0. M \ 9 m. < 0.8. 0.. 9 O D. . A 0 to 000. a. O O . m. 0 i¢o+mo . W4 00 . < d \ tmo+mo.. 0 a\ Sons”. 5 xmm...on> imo+mo.. 76 and shear rate. Compared to Fig 5.2, the fit has improved (R2=0.91, slope=0.482) and all data points appear to fall along a more defined line. However, it is clear from the slope (0.482) of the regression line that further correction is necessary. Observed apparent viscosity versus predicted apparent viscosity corrected for tem- perature, moisture and shear rate can be seen in Figure 5 .5 . There is further improvement in fit as measured by R2 (0.93), however, the slope of the repession line has decreased significantly (0.347). As can be seen in propessing from Figures 5.2 to 5.5 the data points are becoming more alipred. The decrease in slope of the repession line is an indi- cation that the observed viscosity is greater than the predicted viscosity. This is to be expected since the shear, temperature, and moisture corrections were made on an uncooked material; therefore, applying these corrections to cooked material should result in predicted values less than those of cooked starch. When the observed viscosities cor- rected to a shear rate of 100 s’1 are plotted versus cook time at a moisture content of 0.476 g water per g starch (32.0%) (Figure 5.6) an increase in viscosity with time is seen only for the 7 5’C cook temperature. At the Other cook temperatures, no appreciable difference in viscosity is observed at all cook times. By a cook time of 3 minutes, the data for the 75’C cook temperature had reached a constant viscosity. This is consistent with the observations of Kubota et al. (1979), Bakshi and Singh (1980), Lund and Wirakartasu- mah, (1984), Burros et al. (1987), and Dolan et al. (1988). When an analysis of the temperature profile within the capillary rheometer barrel was made using the heat conduction equation for infinite cylinders (Ozisik, 1980), the center temperature of the starch material reached and exceeded the gelatinization temper- ature in one minute or less. Therefore, it was assumed that the product of k. and ‘l’ was very large, indicating very rapid gelatinization. [The term, exp(-k, 0!), therefore is zero for all cook temperatures and cook times. 77 00030.: 0.30.00E3I0E: =0 .00 05000.50 0.306;. 0:0 0.30.0903 .30. .0000 .33 .000300 00.30 :.00 00 00200003 0020000 000.0> 03000.1 0.0 0590 0 01 0300005 0020000 00+_m_0.. 00+.00.. 03.00.. 0H0. 0.00. 33 0... 0000.0 o 33 0.... 00.4.0 0 33 02 000.0 a vmdflnm x0¢n.ofl> 10.00— 10.000— [00+m0; 100+m04 [00+m0._ s 0:1 ‘KrisooslA perolpeld 78 .2080 0830 58 .3 30 0\0 03.0320 00: 0:0 .00 .00 .00 .0 00.30.0903 x000 .2 0E: x000 000.0> 360003 300.0000 0m.0.0>< .ofi 0.39... 00: OQDO 5:. .0E_._. x000 ON _ L MN OF L |-+<-H|HB—l l—°—-|H+fl-9-I [jfi—Tjjfi j c.000— ¢o+woé mo+woé s Dd ‘Kggsoosm eBDJaAV 79 The correction for time-temperature history then becomes: 11 = ni.T.uc[1+A1 [515] where A’ is defined as: I n" a A = —1=B,(Cs) [5.16] 717.131“: The variable A’ is, essentially, a ratio of the maximum viscosity due to gelatiniza- tion that can be attained (for a specific cook temperature and moisture content) to the ungelatinized viscosity corrected to the cook temperature. It is important to note that A’ relates to a particular moisture content and cook temperature. The specificity for mois- ture content is not due to the lubrication effects of water, but rather to the effect of mois- ture content on the gelatinization kinetics. Values of a and B, for Eq. 5.15 were determined by holding the temperature con- stant and regressing In A’ versus 1n Cs for all the data. As temperature increased, a decreased (except at llO‘C) while B, increased (Table 5.3). This was to be expected because the difference between the viscosity of the uncooked starch and the cooked starch also increased as the cook temperature increased. The exact cause of the high a value at llO'C is not known. However, one possible explanation is that at the high cook temperature, some molecular breakdown may have occurred causing more entanglement between molecules and therefore a greater value of a. To find a predictive equation for the constant A’, equations for at and B, as a func- tion of temperature had to be determined. An Arrhenius relationship for both a and B, was assumed and regression of In at or B, versus l/T was then performed. Equations [5.15] and [5.16] are the resulting equations for a and B,. a _ costs-3?) R2 = 0.952 [5-17] 80 Table 5.3. Values of constants 0t and B, as a function of temperature. Temperature at B, ('C) 75 2.53 4.90 85 1.90 5.57 95 1.70 1 1.30 1 10 2.05 21.96 81 4021*) 1' Bo = R2 = 0.947 [5-18] Combining Eqs. [5.16], [5.17], and [5.18] yields an equation for the constant A’. By combining Eqs. [5.10] and [5.16], the final predictive equation for viscosity as a function of shear rate, temperature, moisture content, and time-temperature is JI- ‘1 . _ ; r"-r;‘ +b(Mc—Mc, a “i,r,uc,y=l:[%] +0.3": 5‘] e ( ) 1:1 +A’(1 - 84.10] [5.19] A plot of observed versus predicted viscosity after shear rate, temperature, mois- ture content and time-temperature history corrections using Eq. [5 .19] is shown in Figure [5.7]. Clearly the goodness of fit has been improved (R2=0.95, slope=0.72). Further regression analysis using forward stepwise regression (Table 5.3) indicates that the shear rate is the most important variable. This is to be expected, since viscosity is very depen- dent on shear rate. Moisture content is the next contributor to the model in order of importance. Again, this is not surprising; it implies that at the moisture contents used in this study, the lubrication effects of moisture contribute more to the viscosity than the effects of temperature or time-temperature history. Time-temperature history was next in importance, followed by temperature effects. The temperature correction might have been expected to be more important because of the large value of AEJR which indicates a high degree of temperature thin- ning. At low moisture contents, moisture and time-temperature history should be of greater importance compared to temperature in contributing to viscosity. This is because moisture can have a greater effect on the friction between the material and the capillary wall and therefore contribute to lubrication effects. While time-temperature history effects are a function of time, temperature and moisture content are nm a function of 82 000300 00.30 0.00 .0. >.0.0_0 0.30.00.031003 000 0.30.0.0 0.30.0903 .30. .0000 .0. 0303.00 00500003 00>.0000 000.0> 030.005 Rh 0.09... 0 on. $000005 00>.000O 00+00. 00+00. 00.100. 0.000. 0.00. 0.0. L _ — _ b 0.0— 3302 0000.0 0 3302 002.0 0. 30.02 0000.0 4 0000 .0033xm o 0.00. 4 0.000. 0 -00+00. ,00+00. 00.000 X00003. 00.00.. 8 Dd ‘Misoosm 99101990 83 Table 5.4. Stepwise regression of observed viscosity versus temperature, shear rate, moisture content and time-temperature history corrections for native corn starch. SUMMARY OF STEPWISE REGRESSION PROCEDURE FOR DEPENDENT VARIABLE 11,, VARIABLE PARTIAL MODEL STEP ENTERED R2 R2 C(P) F PROB>F 1 7 0.9379 0.9379 216.1 8304.9 0.0001 2 MOIST 0.0142 0.9521 43.2 162.7 0.0001 3 A’ 0.0034 0.9555 3.1 42.13 0.0001 4 TEMP 0.0001 0.9556 4.0 1.07 0.3022 84 time. It is also important to note that while the model was developed assuming no inter- action between terms but in actuality the order in which each term is determined may affect the value of its parameters. Results of extrusion tests can be seen in Table 5 .5 . The moisture content and tem- perature ranges are similar to the conditions used in the capillary rheometer tests. Shear rate ranges are higher than those observed for capillary rheometer tests. Average residence times within the extruder ranged from 45 to 85 seconds. While it should be recognized that residence time within a section is highly dependent upon the screw con- figuration, at the present time there is no accurate method to quantify actual residence time within an extruder zone on the basis of screw geometry. Extrudates varied in appearance from opaque and slightly swelled to translucent and very highly puffed depending on the die length, screw RPM and mass flow rate. Figure [5.8] gives a plot of data from twin screw extrusion runs with all data cor- rected for shear, temperature, moisture and time-temperature history using Eq. [5.19]. Some of the predicted viscosities for the extrusion data are greater than the observed viscosities, indicating that a correction for strain history may be appropriate. Predicted viscosities which were greater than the observed viscosities were from extrusion tests that gave a product that was very highly puffed and translucent in appearance (compared with an opaque, marshmellow-like, non-puffed appearance). Vergnes and Villemaire (1987) found a semi-logarithmic relationship between the power law consistency coefficient and work input for low moisture molten corn starch. Therefore, the relationship between shaft work and viscosity of the extrudate may be use- ful in attempting to quantify the effects of strain history. Use of the mechanical energy input for characterizing the effect of strain history may also be more practical because quantification of shear rate within each zone of the extruder is not possible 85 Table 5.5. Data from twin screw extrusion tests on corn starch. Die Shear Shear Residence Moisture 1],, n- 11“, Temp. Stress Ralte time Content (Pa s) (Pa 3) (Pa 3) (Pa) (8') (8) (db) 83.9 245691 245 56.1 0.60 460 460 1004 84.4 82542 145 75.1 0.61 518 518 569 87.8 264421 220 59.6 0.67 247 247 1201 95.6 68523 215 53.1 0.60 360 360 317 96.1 55327 158 71.6 0.67 263 263 350 96.1 118835 140 84.5 0.65 387 387 849 96.1 35364 150 69.8 0.61 428 332 236 97.2 299127 231 58.8 0.45 903 903 1293 98.3 31395 239 51.0 0.58 374 290 132 100.6 104229 145 80.4 0.50 917 712 719 101.7 215059 139 75.4 0.49 967 967 1549 104.0 336694 225 64.9 0.49 672 672 1498 105.6 80833 144 71.2 0.50 878 682 561 106.7 119117 244 59.0 0.50 716 556 488 112.8 113813 143 66.8 0.49 867_ 673' 794 112.8 170817 226 45.7 0.48 698 698 757 86 .cououm 500 .3 we? Looabxo c_ ESE... oLBoEanuloE: uco .oLBmBE .2333an .39. Loocm Lo”. nouootoo mo_:moom_> oouofioua m:m..o> noiomno .m.m 333... m 0d .3_moom_> noiomno coo. com co. - . _ 0.8 \\. lode. \ \ \ \ a \\o \ D D D \\ D D \ D . \ D \ D D o\u\ on n. Swank $.82 xéodu» 3 od ‘KrgsoosgA parogpard 87 from currently available knowledge, except in some highly specific situations (Mohamed, 1988). Total shaft power for the APV Baker MPF SOD/25 twin screw extruder was deter- mined using the equation Es=4.75 x lO“(%Torque)(RPM)/Q [5 .20] When the conected viscosity versus mechanical energy supplied by the extruder shaft was plotted. a definite relationship is observed (Figure 5.9). N on-linear regression analy- sis of observed viscosities corrected to 150 s", SO‘C, and 0.47 6 g/g moisture versus mechanical energy gave 0. = 282019 e<"-°“°"E') + 4300 [5.21] with an R2 of 0.88. Using the value of the viscosity at infinite strain history (as deter- mined from regression), and correcting for shear rate, temperature and moisture, a conec- tion for strain history can be obtained. The final equation for predicting the viscosity of starch-based doughs in the twin screw extruderis ; 1, ' .15-.“ ‘1 ;(r‘—r;‘)+b(Mc-uc,) “mm”: '-' +9.7 7 e 7 r [1 +A(1 .. fi-Vfl [1 - 0(1 -e‘“)1 [5.22] A plot of the observed versus predicted extrudate viscosities conected for shear rate, temperature, moisture, time-temperature history and strain history using Eq. [5.22] is seen in Fig. [5. 10]. The data appears scattered but regression analysis indicates a good fit (R2=0.85). The observed versus predicted viscosities do appear to fall along one line 88 indicating that shear, temperature, moisture, time-temperature and strain history correc- tions may adequately be used to predict the viscosity of corn starch extrudates over the temperatures and moisture contents used. Table 5 .6. gives the value or relationship for each parameter in Eq. 5.22. 89 .mcmson 5805 500 L3 0 on L0 oLBEoQEB oco m\m owed Lo “coucoo 83m_oE .7m 00* Lo B_moom_> noiomno 6.0 939m mE\w J39: 35cm BoEocooE mo+mo.~ mo+moé _ L L L LL! mo+mmé m0+m0.® mO+mo.v _ 1— L L L L L L Lifllo.o L odoov odoow ¢o+mmé ¢o+moé ¢o+mo.m ¢o+m¢.m neon oocovccoo Rm I vww. H mm . m 11 3&2 65. c2662 6”. cum + Ax .. Amvmvaxm .. 3m 1 > I I l I I I I I I I l I I I l I I I I I I I I I I l I I I s Dd ‘KrgsoosIA peroeuog 9O .50..on 500 .2 $6 .25..on E 332: £0.36 vco .ocayULanouloE: .8329: .oLBEoQEB .80.. .62? .6. 0302.00 @2883 03060.5 mamLo> 82330 09.0 v.59... m om .3_moom_> voimmno 0.000. 0.00. _ 0.0. d J a 0... 3 1. m \. . \\ COOP IA. \ s \ o \ o D \\D S. \ 1+. D D \\ D D U MA \\D D D .00 D \\ DD D n.\n_ gonna ...o.ooo. s x:0.0H> 91 Table 5.6. List of model parameters for Eq. 5.22. Parameter Value or relationship 1:, 189434 Pa ll... 69351 Pa s 11, 0.979 112 0.356 A5, 8473.14 'K b -3.99 A’ ‘ 6070.1 (4.371 +£1.32) ( CS ),( us ) 1W °° B Dependent on T, MC, and 1].. d and n, -5.87 x 103 92 5.6. Conclusions A model to predict the viscosity of extruded starch-based products was developed using the model proposed by Morgan et al. (1988). This model incorporates shear rate, temperature, moisture content, time-temperature history and strain history. The model is flexible enough to allow its use under a wide range of operating conditions and extrusion equipment and yet simme enough that the parameters can be determined via capillary rheometry and standard non-linear and linear regression techniques. Viscosities of cooked and uncooked corn starch measured via capillary rheometry at moisture contents of 0.359, 0.476, and 0.572 g water per g starch (26.4%, 32.0% and 36% mc wb respectively), cook times of one to 24 minutes, temperatures of 50, 75, 85, 95 and 110'C and shear rates of one to 300 s" were found to fit the model with reason- able accuracy. When capillary rheometer data was corrected for shear rate a fit of the observed versus predicted viscosity gave an R2 of 0.88 and a slope of 0.15. Subsequent correction for temperature, moisture content, and time-temperature history improved the fit and the slope to a final R2 of 0.95 and a slope of 0.72. Data for capillary rheometer and twin screw extrusion tests fell along the same line under low strain history conditions. Extrusion tests indicated that when strain history is large, correction for strain history is necessary. Modification of the model proposed by . Morgan et al. (1988) for strain history to include shaft work input as proposed by Vergnes and Villemaire (1987) yielded a final equation that gave a reasonable fit (R2=0.85) for extrusion data. Further research is necessary to quantify strain history effects. One problem encountered in this research was the lack of rheological equipment to enable the indepen- dent measurement of the effects of strain history on low moisture and, hence, high viscos- 93 ity starch doughs outside the extruder. A combination shearing and capillary rheometer as described by Vergnes and Villemaire (1987) may be what is needed to perform this necessary work. Also, a redesign of the capillary rheometer so that it has a wider barrel so heat transfer to the center of the starch dough is slower, may be necessary to quantify the time-temperature effects at values less than infinity. Further investigation into the phe- nomena of slip and friction within the capillary die also needs to be done. W 94 5.7. Nomenclature A, A’, A, relative amount of viscosity increase due to gelatinization, dimensionless b rate constant of effect of moisture content on viscosity, dimensionless Cs starch concentration, wet basis decimal Cp protein concentration, dry basis decimal d rate constant of effect of strain history (0) on viscosity, s D capillary diameter, m AE, free energy of activation, kcal/g mole Es mechanical energy supplied by the extruder, Jm’3 L capillary length, m MC moisture content, dry basis decimal MC, reference moisture content, dry basis, decimal n,, power indices, dimensionless R. capillary radius, m universal gas constant, 1.987 cal/g mole t time, s T temperature, 'K T, reference temperature, 'K Q volumetric flow rate, tn3 s‘1 01 index of molecular weight effects on viscosity, dimensionless [50, [3' material constant describing effects of temperature and moisture on gelatinization, dimensionless [5 material constant describing effect of strain history on viscosity at 4) = co, dimensionless T] apparent viscosity, Pa 8 '00 '0. TL. “true.” “70,1140 119 95 contribution to viscosity by gelatinization, Pa s normalized apparent viscosity ratio, dimensionless viscosity at \V = co, Pa 3 apparent viscosity corrected for shear rate, temperature, moisture content, time-temperature history and strain history, Pa 8 apparent viscosity at time-temperature history \|I, Pa 3 viscosity at \y = co, Pa s change in apparent viscosity due to gelatinization at \y = co, Pa 3 apparent viscosity before gelatinization, Pa 3 apparent viscosity at (b = no, Pa 3 €9<~ 96 material exponent for describing effect of moisture content on protein dena- turation, dimensionless shear rate, s., strain history, dimensionless time-temperature history, 'K-s yield stress, Pa shear stress at wall, Pa high shear limiting viscosity (Casson), Pa 3” high shear limiting viscosity (Ofoli), Pa 5"” pressure drop, Pa 97 CHAPTER 6 A RHEOLOGICAL MODEL FOR NATIVE WHOLE WHEAT FLOUR DOUGH EXTRUDATES 6.1. Abstract A generalized viscosity model for predicting the extrudate viscosity of low to inter- mdiate moisture content starch based products is proposed. The model incorporates the effects of shear rate, temperature, moisture content, time-temperature history and strain history. The model was tested using whole wheat flour doughs at varying moisture con- tents. Equipment used included an Instron Capillary Rheometer attached to a Model 4202 Instron Universal Testing Machine and an APV Baker MPF 50 D/25 co-rotating twin screw extruder. Die lengths of the capillary rheometer were 6.35 x 10'3 m and 2.54 x10“2 m, with diameter of 3.18 x 10'3 m and 1.59 x103 m. Extruder dies of 6.35 x10'3 m diameter, and 2.54 x 10'2 m and 3.18 x 10'3 m lengths were used. Whole wheat flour dough moisture contents were 0.333, 0.337, 0.385 and 0.436 g water per g starch (25.0%, 25.2%, 27.8%, and 30.4%, wb respectively) for capillary rheometer tests. In the capillary rheometer, barrel temperatures were maintained at 50, 55, 60, 75, 85, 95, and 110'C at cook times of 1, 2, 3, 6, 12, and 24 minutes. Shear rate was modelled by the generalized model proposed by Ofoli et al. (1987). Overall fit of the model improved as temperature, moisture content and time-temperature history were accounted for. The fit was not as good for whole wheat flour as was observed for corn starch and potato flour (R2: 0.56). The apparent lack of fit may be due to the presence of flour components such as bran, protein, and lipids which the model does not account for and which may have altered the gelatinization kinetics of the starch. 98 6.2. Introduction The food industry is constrained from effective use of extruders because of lack of an adequate model which describes the effects of extrusion process variables on the extrudate viscosity. Many researchers have proposed rheological models for extrusion cooking of cereal products (Harper et al. 1971; Remsen and Clark, 1978; Cervone and Harper, 1978; Bhattacharya and Hanna, 1986; Morgan et al., 1988). While all these mod- els account for shear rate, moisture, and temperature, only Remsen and Clark (197 8) and Morgan et al. (1988) included time-temperature history effects. In Chapters 4 and 5, the equation below was used to model the viscosity of potato flour and corn starch: 1: i "'(r‘ 1") Marc MC) nm=[[§°] +115!” “‘] e' ' ' r [1 +14 ’(1 - e"-")“] [1 — 130- e‘“)] [6.1] However, it is commonly accepted that various components of cereal doughs (protein, lipids, and pentosans) or ingredients added to the product (sugar, salt, etc.) can greatly affect the extent of gelatinization (Eliasson, 1983; Lund, 1984; Olkku, 1978; Ghiasi et a1. 1982). The proposed model does not account for the effects of non-starch components on the gelatinization kinetics of starch. Since it was observed in Chapter 5 that starch gela- tinization significantly affects viscosity, it is reasonable to expect that the presence of dough components such as protein, salts, or lipids may affect the changes in viscosity during cooking extrusion. The efl‘ects of these dough constituents can, therefore, be eval- uated by applying the model to a product containing significant levels of one or more non-starch components. Native whole wheat flour provides an ideal material for this exercise. 99 The objective of this study is to assess the accuracy of the model above for predict- ing the viscosity of native whole wheat flour extrudates. 100 6.3. Materials and Methods Whole wheat flour (International Multifoods, Minneapolis, MN) containing l4»l4.5% protein (wb) provided the feed material for the extrusion runs. Moisture con- tents used were 0.333, 0.337, 0.385, and 0.436 g water per g flour (25.0%, 25.2%, 27.8%, and 30.4%, wb respectively). The indicated moisture contents were used because at higher moisture contents, rapid loading of the capillary rheometer was diffith due to the stickiness of the material, and at lower moisture contents damage to the Instron load cell or capillary rheometer plunger might occur. Flour and water were mixed in an institutional mixer and allowed to equilibrate at room temperature (24'C) overnight before being placed in plastic freezer bags and stored at 0'C. Prior to capillary rheometer tests, the sample was allowed to thaw for a minimum of 12 hours. Moisture content was determined by drying overnight at 100'C. No change in moisture content was observed during frozen storage. An Instron Capillary Rheometer and a Model 4202 Instron Universal Testing Machine (Instron Corp., Canton, Massachusetts) (Figure 6.1.) were used to measure apparent viscosity of the doughs. Dies of 2.54 cm (1 in) and 0.635 cm (1/4 in) length, and 4.90 x 102 cm (1/8 in) and 1.59 x 10'1 cm (1/16 in) diameter were used, giving up ratios ranging from 2 to 16. Three replicates for each plunger velocity, temperature, moisture content, cook time and MD ratio were performed. Force versus plunger displacement data were collected and force at the die entrance was calculated by extrapolation of the force versus displacement curves to the die as described by Einhom and Turetzky (1964). Barrel drag was then subtracted from the cor- rected force. A correction for entrance effects was made using the technique described by Bagley (1957). Pressure drop through the capillary die was assumed to be the force at the die entrance over the barrel diameter. 101 Barrel wall — 7" Plunger 0.0.0.000... O. Sample Capillary die Figure 6.1. Capillary rheometer assembly. 102 Shear rate and shear stress were then calculated using the Rabinowitsch equation, 0 . _3Q «15,? and the standard expression for shear stress at the wall of a capillary: _APR, [63] 1w- 2L " Slip analysis was performed using the method described by Darby (1976). Temperatures used in this study were 50, 75, 85, 95, and llO’C . At 50'C the materials were compressed in the capillary barrel and allowed to equilibrate for 6 min- utes. Cook times of 1, 2, 3, 6, 12 and 24 min at 75, 85, 95, and 110'C were performed at each moisture content after compression. Shear rate, temperature, moisture and time-temperature history parameters were determined as follows. Holding temperature (50'C), moisture (0.333% db) and time- temperature history constant (w=0), viscosity versus shear rate were fitted to the model of Ofoli et al. (1987), using the Marquardt compromise method in the non-linear regression program of SAS (SAS Institute, Cary, NC). Then holding moisture content (0.333% db) and time-temperature history (\y=0) constant, 1n 11 versus ('1‘ 1-T,.") with T,=323.15 ’K were regressed linearly to obtain AEJR for the temperature correction. The moisture cor- rection parameter was determined by holding time-temperature history constant (01:0) and linearly regressing 1n 1] versus (MC-MC), with MC,=0.333 db. Finally, correction for time-temperature was performed. Experimental extrusion tests were conducted using a APV-Baker MPF 50 D (APV Baker, Inc., Grand Rapids, MI) co-rotating twin screw extruder with the screw configura- tion shown in Table 6.1. Feed rates of 1.01 x 10'2 and 1.64 x 10’2 kg/s, were used at 200 and 400 RPM. Two dies 3.17 x 10" m in diameter and 6.40 x 103 and 2.54 x 10" m in 103 length were used. Pressure drop and extrudate temperature at the die were recorded two rrrinutes after extruder operating conditions had been changed to allow equilibrium condi- tions to be attained. 104 Table 6.1. Screw configuration for extrusion tests on APV Baker 50 mm twin screw extruder. 17.8 cm 8.9 cm 7.6 cm 6.4 cm 10.2 cm 8.9 cm 7.6 cm 7.6 cm Feed Screw Forwarding paddles 30 offset Feed Screw Forwarding paddles 45 offset Single Lead Forwarding paddles 30 offset Fwd Screw Single Lead Feed Inlet Extruder die 105 6.4. Results and Discussion Final equilibrium moisture content of the wheat flour samples were determined to be 0.333, 0.335, 0.385, and 0.436 g water per g wheat flour (25.0%, 25.2%, 27.8%, and 30.4%, wb respectively). No loss in water was observed during storage. The extrudates exhibited the appearance of slip, with a " shark skin" texture on the surface. However, analysis for slip was found to be inconclusive at all moisture contents and temperatures. The inability to correct for slip may be due to a "slip-stic " phenome- non where there is a combination of friction and slip occurring at the capillary wall. Because "slip-stick" rather than pure slip occurred, Darby’s (1976) method for slip correction did not yield meaningful results. It was either impossible to obtain a slip coef- ficient because of wide variations in the data, or a negative apparent shear rate was obtained. Another reason for the difficulties encountered in slip correction may be a varying friction coefficient within the die, depending on the plunger velocity, die geome- try or a combination of the two. Due to the non-cohesiveness of the raw material, it was impossible to correct for this friction. The reference moisture content was set at 0.333 g water per g (25.0% wb) flow. A yield stress of 93.8 kPa was estimated by extrapolating shear stress versus shear rate curves to zero shear rate. After determining the yield stress, the other parameters in the shear rate model were determined via non-linear regression. The fit of the non-linear regression parameters as measured by R2 (0.88) was not as good as that observed in Chapters 4 and 5 for potato flour and corn starch. This poor fit may be due to the presence of bran, protein and other components in the whole wheat flour which are not accounted for by the model. Temperature and moisture correction terms are listed in Table 6.2. Values for both are within the ranges seen by other researchers for similar materials. 106 Table 16.2. Temperature and moisture correction parameters for whole wheat flour at 100 s' Parameter Estimate Range of previous values . Temperature (A5,) 5354.38 2000-5000 ’K Moisture (b) -7.91 ~15.0 - -0.19 N0te: All estimates are significant at the 95% level. 107 Comparison of average observed versus predicted viscosities at all shear rates, temperatures, moisture contents, and time-temperature histories conected for shear rate only can be seen in Figure 6.2. Regression analysis (R2=0.56, slope=1.86) of observed versus predicted values indicate that shear rate- correction alone is inadequate to describe the viscosity of whole wheat flour. Figrne 6.3 shows observed versus predicted Viscosities corrected for shear rate and temperature at all moisture contents and time-temperature histories. Temperature correc- tion improves the fit of the line (R”0.635) and the slope approaches unity (slope=0.83). While the correction for temperature improves the fit, the data still appears to be somewhat scattered as seen in Figure 6.3. Correcting the predicted viscosities for moisture content gives fin'ther improvement in the fit (Rz=0. 638) of the average observed versus predicted viscosities. A plot of observed versus predicted viscosity can be seen in Figure 6.4. The slope of the line has decreased to 0.76 indicating that predicted viscosities corrected for moisture, temperature and shear rate are less than observed viscosities. These predicted viscosities are lower because most of the observed viscosities were either fully or partially cooked and there- fore the starch and protein molecules had undergone gelatinization or denaturation, respectively. It is commonly accepted that viscosity increases as starch gelatinizes and protein denatures, therefore the predictive model must incorporate a term to quantify the effects of gelatinization on viscosity. Comparison of the observed and predicted viscosities of wheat flour cooked at 75'C and 85'C at a moisture content of 0.337 g water per gram solids (25.2% wb) and 75'C with a moisture content of 0.385 g water per gram solids (27.8% wb) indicated that no increase in viscosity occurred at varying cook times and therefore the 108 .2950 So: 000:2, 06:2, no. 5302 83800:.8 I08: nco 8:00:00 830.08 .0838an8 :0 Lo 80.. .0030 00.. 0808.60 0036033 0020020 m:80> 080680 .m.0 839... m on. 5.60005 0050000 mo+mo.. mo+mo.. 21.00.. odwo. 3%: 0.0. 330: Rm... 6 . 330.2 003 a no 00. 330.... 8.8 a d l 8 m. 10.000. m. 8 Dr W. . x r¢o+mo. o m... 1. ”A $050.. moo moons”. s x4007...» 100+m0._ 109 050300 So: 300;; 0653 .o. 00:30:. 8380080705: 0:0 03:03:00 8306:. :0 30 0.380083 0:0 030. .0050 .o. 00308.60 00280.03 00>.0mno 0:0.0> 003208.”. .00 0.30.1. 0 00 0300005 0050000 9:00.. 0300.. 2:00.. 0.000. 0.00. 0.0. L _ _ _ L 0.0— 330... Rm... 6 3302 000.0 a 330... 004.0 n 10.00. 4 4 10.000. 4 reo+mo. 0 <0 .0 100+mo. 000.004.". 0.400.001; 0300.. 8 od ‘Krlsoosm perolpeld 110 000300 .30: .0003 0602. .0. 00:30.0 0.30.00E0LI0E: =0 .0 3:03:00 0.30.0... 0:0 0.30.0080. .0.0. .0000 .0. 00.08.00 00.00.0003 00>.000o 030.0> 030.080 6.0 830.0 0 00 03.00005 00>.000O 00+0. 00$. 00+“: 000. 00. 0. F _ _ _ L O.— .03 0... 00.0 o .00. 03 000.0 0 .03 03 00....0 n 10.0. 0.00. 0.000. 140+00. 100+mo. x050...» 00030.. 5 od ‘ArlsooslA perolpeld 111 time-temperature history variable was set to zero (01:0). The lack of differences between the predicted and observed viscosities may be caused by the low moisture contents and lower cook temperatures. Lund (1984) observed that the transition temperature on a differential scanning thermogram increased slightly as the moisture content decreased. This transition temper- ature is indicative of the onset of gelatinization. Presence of gluten, lipids and solutes can also compete with the starch for water and thus afl'ect the onset of gelatinization. The combination of low moisture contents and the presence of gluten in the whole wheat flour most likely caused the gelatinization temperature to shift higher, leading to the lack of noticeable differences between the observed and predicted viscosities corrected for shear, temperature, and moisture content. Using observed and predicted viscosities at a 24 minute cook time and at each tem- perature where a noticeable difference between the observed and predicted viscosity occurred, In A’ versus ln(Cs) were regressed for each cook temperature. A relationship between a and B, and the inverse temperature was then determined. An Arrhenius rela- tionship between inverse temperature and B, was observed. This is similar to what was observed for corn starch (Chapter 5.). However, unlike cornstarch, an Arrhenius relationship was not observed between or and the inverse temperature. The equations for orandB,are [6.4] B. = Ami?) 16.51 112 Viscosity was observed to be a function of cook time and therefore time- temperature history parameters were calculated according to the procedure described by Morgan et al. (1988) and Dolan et al. (1988). The parameter AE‘ was found to be unaffected by moisture content. This is in agreement with Dolan’s (1988) findings that there is little difference between AE, values for high moisture corn starch solutions of varying starch concentrations. Burros et al. (1987) estimated activation energies for corn starch. Over the moisture content ranges used in their study, no significant difference between activation energies was observed. The moisture content range for this study is within the range studied by Burros et al. (1987). In this study, the activation energy (A5,) was found to be 25100 cal/g-mole. The constant kI was found to be slightly affected by moisture content, however, this may be due to normal sample variations and not because of any moisture content dependency. Values for k, were 2.5 for 0.436 g water per g solids (30.4% wb) and 1.1 for 0.385 and 0.337 g water per g (27.8% and 25.2% wb, respectively) solids. Figure 6.5 is a plot of observed viscosity versus predicted viscosity corrected for shear rate, temperature, moisture content and time-temperature history. The fit of the data improved (R2=0.687) when compared to the fit after correcting for shear rate, tem- perature, and moisture content alone; the slope was also slightly improved (0.815). This improvement is not as dramatic as that observed for native corn starch (Chapter 5). Reasons for this may include a) "slip-Stick" phenomena or friction effects may have had a greater effect on the viscosity of whole wheat flour than observed for native corn starch, b) the presence of protein, bran and other flour components that were not accounted for in the model may have had an effect on the time-temperature parameters, and c) the low moisture content used in this study may have amplified either of these effects or made it very difficult for the capillary rheometer to measure the effects. 113 000300 .30: 6002. 060.5 .0: 330:. 0.3.0.00E03I0E: 000 .8306... 0.3060000. .80. .0000 .0: 00.08.00 00200003 00>.0000 030.0> 006680 .00 0.30.... 0 00 03.00005 00>.000O 00+00. 00+00. 0.000. 0.00.. 0.0. _ _ b L O.OF 33 0... 000.0 6 .03 0.... 000.0 0 .03 02 004.0 a .0 <20 00. o 0.00. a . p O m. C 0 mm. D. 00.000. \ 0. O O . .0. 100+00 . ”ms 0 D hmmdnnm Imo+mo.. S x..0.0.0n> 114 Another possibility is the likely presence of an elastic component due to gluten in the whole wheat flour. Gluten is developed by mixing flour and water together and then kneading (mixing) the dough. The shear rate term in the model does not account for energy losses due to elasticity of the dough and therefore this may be why the overall fit of the model is less for whole wheat flour than for corn starch. It should be recognized that gluten development is probably the major contributor to viscosity below starch gelatinization temperatures. In the capillary rheometer it may be assumd that gluten development was negligible because very little mixing occm's in the instrument. Therefore, gluten development occurred during sample preparation and not during capillary rheometer tests. Stepwise regression after log-log transformation gave results similar to those for native cornstarch where the order of importance for each model parameter was shear rate, temperature, time-temperature history and moisture content. Table 6.4 lists the stepwise results for whole wheat flour. Extrusion test results after correction for shear rate, temperature, moisture content and time temperature history are plotted in Fig. [6.6]. Extrudate properties varied from unpuffed and unswelled material to a highly puffed bread-like material which did not hold its puffed appearance upon cooling. Unlike the case for corn starch, strain history does not appear to affect wheat flour during extrusion. A relationship between observed viscosity and shaft work input, as was seen for native corn starch, could not be deter- mined and therefore the strain history correction term was set to unity. Regression through the origin for observed versus predicted viscosity corrected for shear rate, temperature, moisture content and time-temperature history indicates a better fit (R’=0.788) than the data from the capillary rheometer when one compares R2 values (0.788 versus 0.687). 115 Table 6.3. Results of stepwise regression for whole wheat flour doughs. SUMMARY OF STEPWISE REGRESSION PROCEDURE FOR DEPENDENT VARIABLE n... VARIABLE PARTIAL MODEL ENTERED 19.2 1?.2 C(P) F PROB>F «’1 0.4953 0.4953 669.14 1026.39 0.0001 TEMP 0.1391 0.6344 198.65 397.65 0.0001 A’ 0.0538 0.6882 17.79 180.31 0.0001 MOIST 0.0046 0.6929 4.00 15.79 0.0001 116 050300 .00: .0002. 06:2. .0. 00.0 .00:..x0 5 E302 0.30.00E00I0E: 0:0 0.30.0... 0.30.0000. .30. 50:0 .0. 00.00.80 00200003 00.0.00... 000.0> 0020000 0.0 0.39.. 0 0n. .>._0000_> 00.0.00.n_ 0.000. .000. 0.0. L L 0.0. 0 q S 3 .J m \ D. o o \ 00.00. \ W. 0 0 \ s \ 3 \o o o W. O\\ O O O 0 MM. \ o \06 o m. 0 O O O Q x....n> [0.000. s 00:78. 117 Inability to correct for strain history for whole wheat flour compared to corn starch may be attributed to the differences in the general make-up of the two materials. Corn starch is nearly 100% starch, whereas the whole wheat flour consists of starch, bran, wheat protein and some lipids. ‘ The inclusion of water and the mixing in the extruder most likely caused development of the gluten which may have contributed to increase in viscosity that was not counteracted by a lowering of the viscosity by mechanical break- down of starch. A list of the parameter values or the relationship used in Eq. 6.1 are given in Table 6.5. It is important to note that the order in which values were determined may affect the final results. Also, some of the values depend on either temperature or moisture content or both. 118 Table 6.5. List of parameter values for Eq. 6.1. Parameter Value or relationship 1:0 93.8 kPa p,_ 105000 Pa s n1 1.0 n2 0.40 AE, 5354.4 'K b -7 .91 A e(—59.99+2“r—‘5)(Cs)(-143.6+-’°$’-) AEx 25100 cal/g-mole. k, 2.5 s"-'K at 0.436 g/g db [1 - [3(1 -e‘“)] 1.1 s"-'K at 0.337 & 0.385 g/g db 1.0 119 6.5. Conclusions Capillary rheometer tests were performed on whole wheat flour at moisture con- tents of 0.333 to 0.436 g water per g solids (25.0% to 30.4% wb), cook times of one to 24 minutes, temperatru'es of 50, 55, 60, 75, 85, and 110'C and shear rates of one to 1000 s“ to obtain the variables needed use Eq. 6.1. Overall fit (R2=0.687) of the observed versus predicted viscosities for whole wheat flour was less than was observed for native corn starch (R2=0.950). This may be due to the effects of flour components (bran, protein, fat) which were not present in the corn starch and are not accounted for by the general model. No differences due to time-temperature history were observed for samples contain- ing 0.337 g water per g solids (25.2% wb) and cooked at 75'C and 85'C, but at the other moisture contents and cook temperatures, viscosity did change with cook time. The presence of flour components competing for water probably afi‘ected the gelatinization kinetics at the low moisture content and lower cook temperatures, causing no measurable change in viscosity at different time-temperature histories. ' Use of the parameters obtained from the capillary rheometer in Eq. 6.1 gave good agreement between observed and predicted viscosity (R2=0.788) for whole wheat extru- dates. Different strain histories did not appear to affect the twin screw extrusion data and therefore the strain history term was set to unity for all extrusion tests. Again the presence of the different flour components may have had an effect on the ability to cor- rect for strain history. Future research should include the effects of flour constituents (e. g. bran, protein, fat) or other additives (salt or sugar for example) on the model. Specifically, the effects of flour constituents on the time-temperature history and strain history correction should be determined. Further research into quantification of strain history within the extruder is 120 also necessary. The question of whether an accurate measurement of shear rate in each extruder zone is necessary to quantify strain history or if measurement of work input alone is adequate must be answered. 121 6.6. Nomenclature A, t" coca .3. ”.63 .. E C, wpogsa 9:1—1" "a: relative amount of viscosity increase due to gelatinization, dimensionless rate constant of effect of strain history (