DEGRADATION OF ASCORBIC ACID IN A DEHYDRATED MODEL FOOD SYSTEM DURING STORAGE BY Daniel B. Dennison 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 1978 ABSTRACT DEGRADATION OF ASCORBIC ACID IN A DEHYDRATED MODEL FOOD SYSTEM DURING STORAGE BY Daniel B. Dennison The storage stability of reduced and total ascorbic acid in a dehydrated model food system designed to simulate a ready-to-eat breakfast cereal was studied as a function of water activity, moisture content, oxygen and temperature. Sorption isotherm data obtained at 10, 20, 30 and 37°C for the model system were used to calcu- late the Brunauer, Emmet, Teller (BET) monomolecular moisture content at these respective temperatures. The model system, which contained 11.25 mg of ascorbic acid per 100 g dry weight basis (25% Recommended Dietary Allowance, RDA) was equilibrated at water activities below, at and above the water activity corresponding to the calculated BET monomolecular moisture content for the adsorption isotherm. The samples were sealed in thermal death time (TDT) and 303 cans and stored under isothermal conditions to prevent any shift in water activity. The TDT cans were filled with the model system Daniel B. Dennison so that no headspace remained and limited gaseous oxygen was available due to the inter and intrastitial spaces. The 303 cans were filled with an equal mass of model sys- tem as the TDT, thus providing a large excess of gaseous oxygen in the headspace. The effect of air and moisture vapor transmission on the storage stability of ascorbic acid in the model system was studied by packaging the model system in l-oz. paperboard boxes with waxed liners. The samples were stored at 30°C and 10, 40 and 85% rela- tive humidities. Reduced, dehydro and total ascorbic acid concen- trations were determined as a function of storage time by an automated o-phenylenediamine fluorometric assay pro— cedure. Under all storage conditions, reduced and total ascorbic acid losses could be satisfactorily described by first-order kinetics. Ascorbic acid destruction in the model system stored in cans or boxes was linearly dependent on the water activity, exhibiting maximum stability below the BET monomolecular moisture content. These results con- tradict the hypothesis that the BET monomolecular moisture content should represent the water activity of greatest stability. Comparison of the rate constants for total and reduced ascorbic acid degradation in TDT and 303 cans at identical conditions of water activity and storage Daniel B. Dennison temperature showed a dependence on the availability of gaseous oxygen. The results were interpreted in terms of the consumption of dissolved oxygen in the degradation of ascorbic acid and the transfer of gaseous oxygen into the product moisture which would be governed by the equil- ibrium constant K = (02)d/(02)g, where (02)d and (02)g represent the concentration of dissolved and gaseous oxygen present in the system, respectively. The influence of riboflavin and vitamin A on the degradation of ascorbic acid in the dehydrated model system was studied as a function of water activity and storage temperature. Ascorbic acid degradation was found to be unaffected by the presence of either vitamin. The catalytic influence of trace minerals (Fe, Cu, Zn, Ca) on the rate of ascorbic acid degradation was studied as a function of water activity. No catalysis by the added metals was observed as 0.10 and 0.40 aw, except for copper at 0.40 aw. This is interpreted as a lack of metal ion mobility and/or insolubility at low water activities. At 0.65 aw, which was in the capillary region of the adsorption isotherm, a 2-3 fold increase in degradation rate over the nonfortified system was observed for each form of added trace mineral with the exception of zinc oxide, which did not exhibit catalysis. These results are explained of the basis of Fe and Cu Daniel B. Dennison ion mobility in the capillary region of the isotherm, whereas Ca and Zn catalysis are interpreted as altering the activity coefficient of oxygen. Activation parameters were calculated for the degradation of ascorbic acid in the model system stored in metal cans. The free energy of activation remained constant at all aw, thus it was concluded that the degra— dation of ascorbic acid followed only one mechanism. ACKNOWLEDGMENTS The author wishes to express special gratitude to Dr. James R. Kirk. His advice, personal direction and expert counsel provided invaluable guidance through- out this investigation and during the writing of this dissertation. The author also expresses grateful appreciation to the members of his graduate committee, Drs. J. R. Brunner, D. R. Heldman, L. L. Bieber and R. W. Luecke for their review of the manuscript. A special note of appreciation is extended to the members of the Department of Food Science and Human Nutrition, Michigan State Uni- versity, for their continual interest and encouragement. The author expresses his most sincere appreciation to his wife, Judy, for constant inspiration and encourage- ment throughout his graduate studies. ii . TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . LIST OF FIGURES . . . . . . . . . . INTRODUCTION . . . . . . . . LITERATURE REVIEW . . . . . . . . . Sorption Phenomena . . . . . . . Water Activity and the Sorption Isotherm Theoretical Treatments of the Sorption Isotherm . . . . . . . . . . Physical State of Water . . . . Thermodynamic Activation Parameters. . . Reactions in Dehydrated Foods. . . . . Microbiological. . . . . . . . . Enzymatic. . . . . . . . . . . Chemical . . . . . . . . . . . Ascorbic Acid . . . . . . . . . . Structure and PrOperties. . . . . . Reactivity in Solution . . . . . . Stability in Food Products . . . . . Predictions of Product Storage Stability . EXPERIMENTAL PROCEDURES . . . . . . . Model Food System Composition. . . . Model Food System Preparation. . . . Addition of Other Nutrients to the Model System . . . . . . . . . . . Model Food System Equilibration . . . . Model System Packaging and Storage . . . Moisture Content Measurement . . . . . Ascorbic Acid Determination . . . . . Data Analysis . . . . . . . . . . Experimental Design . . . . . . . . iii Page vii 17 20 22 23 24 25 26 29 35 38 42 42 42 43 44 48 49 49 50 51 Page Ascorbic Acid Stability in TDT Cans . . . . 52 Ascorbic Acid-—Vitamin Interactions in TDT Cans . . . . 53 Ascorbic Acid Stability in Paperboard Boxes . . . . . . 53 Ascorbic Acid Stability in 303 Cans . . . . 53 Ascorbic Acid--Mineral Interaction in 303 Cans . . . . . . . . . . . . 54 RESULTS . . . . . . . . . . . . . . 55 Ascorbic Acid Stability in TDT Cans. . . . . 55 Ascorbic Acid--Vitamin Interactions in TDT Cans . . . . . . 62 Ascorbic Acid Stability in Paperboard Boxes . . 65 Ascorbic Acid Stability in 303 Cans. . . . . 69 Ascorbic Acid--Minera1 Interaction in 303 Cans . . . . . . . . . . 72 Thermodynamic Activation Parameters. . . . . 76 DISCUSSION. . . . . . . . . . . . . . 80 Ascorbic Acid Stability in TDT Cans. . . . . 80 Ascorbic Acid--Vitamin Interaction in TDT Cans . . . . . . 84 Ascorbic Acid Stability in Paperboard Boxes . . 86 Ascorbic Acid Stability in 303 Cans. . . . . 91 Ascorbic Acid-~Mineral Interaction in 303 Cans . . . . . . . . . 103 Thermodynamic Activation Parameters. . . . . 107 SUMMARY AND CONCLUSIONS . . . . . . . . . 118 REFERENCES. . . . . . . . . . . . . . 122 iv LIST OF TABLES Table Page 1. Composition of model food system . . . . . 43 2. Mineral supplement, source and %RDA employed in mineral fortification study . . . . . 45 3. Rate constants and half-lives for TAA and RAA loss as a function of water activity and storage temperature in ascorbic acid fortified dehydrated model food system packaged in thermal death time cans . . . 58 4. Activation energies for TAA and RAA loss in ascorbic acid fortified dehydrated model food systems packaged in TDT cans . . . . 62 5. Rate constants and half-lives for TAA and RAA loss as a function of water activity and storage temperature in multivitamin dehydrated model food system packaged in thermal death time cans . . . . . . . 64 6. Activation energies for TAA and RAA loss in multivitamin fortified dehydrated model system packaged in TDT cans . . . . . . 65 7. Rate constants and half-lives for TAA and RAA loss at 30°C as a function of water activity in ascorbic acid and multivitamin fortified dehydrated model food systems packaged in paperboard boxes. . . . . . 68 8. Rate constants and half-lives for TAA and RAA loss as a function of water activity and storage temperatures in a dehydrated model food system packaged in 303 cans . . . . 71 9. Activation energies for TAA and RAA loss in ascorbic acid fortified dehydrated model food system packaged in 303 cans . . . . 72 Table 10. ll. 12. 13. 14. 15. 16. Page Rate constants and half-lives for TAA loss as a function of mineral fortification and water activity in a dehydrated model food system stored at 30°C in 303 cans . . . . 74 Rate constants and half-lives for RAA loss as a function of mineral fortification and water activity in a dehydrated model food system stored at 30°C in 303 cans . . . . 75 Activation parameters for TAA and RAA degra- dation in ascorbic acid fortified model system stored in TDT cans . . . . . . 77 Activation parameters for TAA and RAA degra- dation in multivitamin fortified model system stored in TDT cans. . . . . . . 78 Activation parameters for TAA and RAA degra- dation in ascorbic acid fortified model system stored in 303 cans. . . . . . . 79 Linear regression of rate constant vs. water activity for TAA and RAA degradation in model system stored in 303 and TDT con- tainers . . . . . . . . . . . . 94 Activation parameters for reduced ascorbic acid degradation in an intermediate moisture model system . . . . . . . . 115 vi Figure 1. 10. LIST OF FIGURES General moisture sorption isotherm for foods (A) and effect of hystersis on moisture sorption isotherm (B) . . . . . . . Stability profile of dehydrated foods (from Labuza, 1976). . . . . . . . . . Molecular structure of biologically active forms of ascorbic acid. . . . . . . Fraction of ascorbic acid species in solution as a function of pH. . . . . . . . Plots loglo k/kO vs. /I for ionic reactions of various types. The lines are drawn with slopes equal to ZAZB (Laidler, 1965) Adsorption isotherm for the dehydrated model food system at 20°C (from Bach, 1974). . . Block diagram for equilibration of dehydrated model food system (inner and outer chamber represent Aminco-Aire unit) . . . . . . Fraction of TAA remaining vs. time for selected aw at 30°C in the ascorbic acid only fortified model system stored in TDT cans . . . . . . . . . . . Fraction of RAA remaining vs. time for selected aw at 30°C in the ascorbic acid only fortified model system stored in TDT cans . . . . . . . . . . . . Fraction of DAA remaining vs. time for 0.10 aw at selected temperatures in the ascorbic acid only fortified model system stored in TDT cans . . . . . . . . . vii Page 21 27 28 34 46 47 56 57 60 Figure Page 11. Arrhenius plot for the activation energy of TAA loss in the ascorbic acid only forti- fied model system stored in TDT cans as a function of aw . . . . . . . . . 61 12. Fraction of TAA remaining vs. time at selected relative humidities at 30°C in the ascorbic acid only fortified model system packaged in paperboard boxes. . . 67 13. Fraction of TAA remaining vs. time as a function of (a) water activity at 20°C and (b) temperature at 0.40 aw for the ascorbic acid only model system stored in 303 cans . . . . . . . . . . 70 14. Moisture content equilibration during storage at selected relative humidities in dehydrated model systems packaged unequilibrated in paperboard boxes (from Purwadaria, 1977). . . . . . . 87 15. Relationship of the rate of TAA degradation vs. water activity at constant storage temperature in 303 (A) and TDT (B) cans (O 10, o 20, A30, 0 37°C) . . . . 93 16. Ratio of (02)/ (AA) as a function of water activity . . . . . . . . . . . 99 17. Relationship of the rate of ascorbic acid degradation vs. water activity at con- stant temperature for an intermediate moisture model food system (data from Lee and Labuza, 1975) . . . . . . . 101 18. Dependence of entropy of activation (A) and enthalpy of activation (B) on water activity for TAA loss in the dehydrated model system stored in TDT cans ( £5) and 303 cans ( O ) . . . . . . . . 109 19. Dependence of entropy of activation (A) and enthalpy of activation (B) on water activity for RAA loss in the dehydrated model system stored in TDT cans ( A) and 303 cans ( O ) . . . . . . . . 110 viii Figure Page 20. Isokinetic relationship of TAA degradation in dehydrated model system stored in TDT cans (A)and303cans(‘). . . . . . . 111 21. Isokinetic relationship of RAA degradation in dehydrated model system stored in TDT cans (A)and303cans(‘). . . . . . . 112 22. Isokinetic relationship for ascorbic acid degradation in an intermediate moisture model system (data from Lee and Labuza, 1975) . . . . . . . . . . . . . 116 ix INTRODUCTION Fortification of food products is not a recent idea. As noted by Bauernfeind (1971), Boussingault, in 1833, first proposed the fortification of table salt with iodine in South America to prevent goiter. The initial impetus for nutrient fortification in the United States - came principally from an executive World War II order in 1941 requiring the addition of thiamin, riboflavin, niacin and iron to white bread and flour to combat beriberi, ariboflavinosis, pellagra and iron deficiency anemia. From the initial addition of vitamins to flour, the practice of nutrient fortification of most foods has been commonplace and, in many instances, expected by the consumer. The principle consideration that should govern any fortification program is public health. The reso- lution of nutrient deficiencies in humans must involve a multi-phase approach. Aylward and Morton (1971) outlined six areas for the improvement of the nutritional status of the general population: (1) improvement in the initial quality of plant or animal product, (2) changes in dietary habits and household practices, (3) improved methods of processing, preservation and storage, (4) biological enoblements, (5) physical treatment and (6) addition of supplements. The changing food patterns and the increased use of manufactured foods have led to the acceptance of food enrichment policies in more than thirty countries. Bauernfeind (1971) has presented an extensive compilation of fortified foods utilized in various countries. Extensive fortification of cereals, flours and other low moisture products has been employed for a number of years in the United States. Addition of nutrients above the label claim is generally required in order to compensate for incomplete distribution, nutrient degradation during processing and storage and analytical error (Borenstein, 1971). Labuza (1968) and Borenstein (1971) have noted that little kinetic data are available describing the stability of vitamins during processing and storage of foods. The influence of factors such as pH, moisture content, oxygen, product composition, trace metals on vitamin stability are unknown or difficult to extrapolate from one product to another. Ascorbic acid is frequently not found in fortified products such as flour and ready-to-eat breakfast cereals because of its instability. Ascorbic acid destruction is known to be affected by pH, moisture content, oxygen and trace metals, but a detailed knowledge of the physio- chemical properties governing the loss and mechanism of ascorbic acid deterioration during storage of dehydrated foods is negligible. The purpose of this research was to investigate the decomposition of ascorbic acid in dehydrated foods as a function of selected product and storage variables. A model food system was chosen to simulate the composition of a ready-to-eat breakfast cereal. The goals were to obtain kinetic data describing the degradation of ascorbic acid as influenced by water activity, temperature, moisture content, oxygen availability and interaction with other nutrients, including trace metals. It was anticipated that thermodynamic activation parameters could be calcu- lated and thus provide information regarding the degrada- tive process of ascorbic acid. LITERATURE REVIEW Sorption Phenomena The common goal of food preservation is the extension of the product shelf-life to permit convenient storage and distribution. The most important consider- ation in food preservation is controlling the growth of microorganisms in the product. One method of preservation is reduction of the availability of water to the micro- organisms by dehydration. However, other physical and chemical reactions may still occur in dehydrated foods which adversely affect food quality. An understanding of the relationship between the state of water and these processes will contribute to an improved quality of the product following storage. Water Activity and the Sorption Isotherm Water is prevalent throughout nature, as would be anticipated, and is very important to the structure, texture and chemical interactions of foods. The thermo— dynamic state of water in a food system is given by the chemical potential (Labuza, 1976) u = no + RT ln aw (1) where no = standard state chemical potential for water, R = gas constant, T = absolute temperature and aw = thermo- dynamic activity of water. The activity of water may be defined as a z; =—___. (2) where: P0 = vapor pressure of pure water, P = vapor pressure of water in the food product and % ERH = per cent relative humidity at which the system neither gains nor loses water. The removal of water from a food system and its relationship to the water activity are depicted in Figure 1. During the initial stages of water removal, the water activity of the product remains close to unity (Figure 1A). On continued removal of water, the product undergoes a relatively rapid decrease in water activity per unit loss of moisture content (9 HZO/g solid). This phenomenon is typical of food systems and the curve is referred to as the moisture sorption isotherm. Equation 2 would predict the ratio of the moisture content to water activity to be independent of the determination of the sorption isotherm. That is, whether the equilibrium moisture is obtained via the removal of water (desorption) or the addition of water (adsorption). Close examination of the sorption isotherm .Amv EHOLDOmw :oflumuom wusumAOE co mfimumum>n mo vomumw pom Awuom waamowmoHofln mo wusuosuum umasomaoz .m muzmflm pwom oflnuoommoupzswplq Odom oflnuoomMIA N o _ oflo\ No-0: old/\o 0-0-0: / \ : \/z o 10 o: 28 .mm mo GOHDOCDM 6 mm coflusaom :fl mmflommm pflom canuoomm mo cofluomum In N. o. m o c N .v musmflh SEIOBdS :IO NOILOVHJ 29 dehydroascorbic acid. The planar structure of L-ascorbic acid was confirmed by x-ray crystallographic analysis (Cox et al., 1932). Reactivity in Solution Ascorbic acid is stable in the dry crystalline state. In solution, it reacts readily giving a variety of products depending upon the reaction conditions. Oxygen, pH, metal ion catalysis and temperature are reported as the principle factors determining the decom- position of ascorbic acid. Although extensively studied by numerous investigators, uncertainty still exists as to the mechanism of degradation. A discussion of the importance of these factors is most easily presented in terms of the presence or absence of oxygen. The anaerobic decomposition of ascorbic acid in aqueous solutions has been reported to be pH dependent. Finholt et a1. (1963) found that the maximum rate of anaerobic degradation of ascorbic acid occurred at pH 4.0 at 96°C. Huelin et a1. (1971) reported an increase in destruction rate as a function of pH at 50 and 100°C, reaching a maximum at pH 2.3, decreasing to a minimum at pH 4 and increasing again to pH 6, the highest pH of the study. The discrepancy between these two studies is difficult to rationalize; however, Finholt et a1. (1963) did find a catalytic effect due to the buffer type and concentration used to maintain the pH. No primary salt 30 effect on the anaerobic degradation rate was found. Fin- holt et a1. (1963) suggested that the rate maximum at pH 4 was due to the complexation of ascorbic acid and the monoionic species of ascorbic acid as a reactive intermediate. Various reaction products from ascorbic acid degradation have been reported in the literature. Huelin et a1. (1971) measured carbon dioxide and furfural for- mation during the anaerobic degradation of ascorbic acid. One mole of CO2 was formed per mole of ascorbic acid over -the pH range 2.2—6.0, but the yield of furfural decreased with increasing pH. Investigations of Coggiola (1963) identified, 2,5-dihydro-2-furoic acid as an end product of the anaerobic degradation of ascorbic acid and Kurata and Sakurai (1967) have reported 3-deoxy-L-pentosone as a probable intermediate breakdown product. The anaerobic catalysis of ascorbic acid by cupric ion at pH 2.1 - 3.5 has been reported by Shtamm et al.(1974). The reaction was second-order with respect to c0pper (II) concentration, first-order in ascorbic acid and inversely proportional to hydrogen-ion concen- tration. Additionally, Huelin (1953) and Huelin et a1. (1971) have reported the catalytic properties of fructose, sucrose, fructose-l-phosphate and fructose-1,6-diphosphate in the anaerobic decomposition of ascorbic acid. The fructofuranose form of fructose was found to be the 31 catalytic agent. Sucrose catalysis was attributed to its partial hydrolysis in the reaction medium. The aerobic oxidation of ascorbic acid has received intensive investigation. The reaction rate is known to depend on oxygen, pH, metal catalyst, buffer and temperature. The rate of the spontaneous oxidation of ascorbic acid increases with pH, showing maxima at pH 5.0 and 11.5 (Csuros and Petro, 1955). This pH dependence has been confirmed by Miller and Joslyn (1949) and Khan and Martell (1967) for the pH range 2 - 6. Weissberger et a1. (1943) found no rate maxima, although the same pH dependence of the reaction from pH 4.7 - 9.2 was observed. The monoionic and diionic species of ascorbic acid have been interpreted as being more reactive in the presence of oxygen than the neutral species. The reported order of ascorbic acid oxidation with respect to hydrogen-ion varies, -0.5, -l.0 and -2.0, Miller and Joslyn (1949), Khan and Martell (1967) and Dekker and Dickinson (1940), respectively. The oxygen dependence of ascorbic acid oxidation has been reported to be first-order above 0.2 atm oxygen (Khan and Martell, 1967). Recently, Shtamm and Skurlator (1974a) have reported a half-order dependence on oxygen, which has been supported by the results of Jameson and Blackburn (1975; 1976a,b). Jameson and Blackburn also 32 suggest that the oxygen dependence data of Khan and Mar- tell (1967) more adequately fits half-order kinetics. Reports of the catalytic properties of transition metals, notably copper and iron, on the oxidation of ascorbic acid abound in the literature. In many cases the rates are found to increase linearly with electrolyte concentration. Existing theories do not account for the rate enhancement exhibited by ion-molecule reactions upon the addition of neutral electrolytes. It is generally assumed that all rate changes resulting from the presence of neutral electrolytes may be described by the Bronsted equation (Bronsted, 1922): Y Y k = k __.__A B (15) O Y+ where YA' and y* are the activity coefficients for the YB reactants and the transition state, respectively. Once the applicability of this equation was accepted, the major problem in developing a theory of salt effects was determining the relationship of the activity coefficients of the reactants to the added electrolyte concentration. The Debye-Huckel theory provided the first relationship of this type in the form of the Debye-Huckel limiting law, _ 2 Log Y — -QZ /I-, (16) where Q is a constant for a given solvent, Z is the valency of the ion and I is the ionic strength. The 33 limiting law is valid only below concentrations of 0.01M. Following the procedure of Scatchard (1922), equation (16) may be substituted into the rate equation (15) giving Log k = Log k0 + 2QZAZB /I— . (17) For aqueous solutions at 25°C, the value of Q is approxi- mately 0.51 (Laidler, 1965). This equation has been shown to describe salt effects for a large number of ion- ion reactions (Figure 5), even though the applicability of this equation is exceeded. According to equation (17) a plot of log k vs. /I-wi11 give a straight line of slope 1.022AAB. For the case where one reactant is a neutral molecule, the product ZAZB in equation (17) is zero and no salt effect would be predicted. The fact that salt effects do exist for ion-molecule reactions has often been explained by assuming that the solutions were not dilute enough to allow application of the Debye-Huckel limiting law. To surmount this shortcoming several relationships using higher powers of I and purely emperi- cal parameters have been developed. Weissberger and LuValle (1944) reported that only the monoionic ascorbic acid moiety was susceptible to Cu(II) catalysis. More recently, studies by Khan and Martell (1967) have shown the oxidation of ascorbic acid to be linearly dependent on the concentration of Cu(II) and Fe(III) ions. Ogata et a1. (1968) found that the 34 A O 0.6 ’ 'o . O O B 0.4 ' f>° o o: C o 0 . O , O .. . o S 0.2 . ...o . Q g; " . . ., . , , D .9 '. -: . o . o o ’0.2 ’ O 0 0 O .o o E -O.4 P F O O _ ‘ . 0.1 0.2 0.3 /I Figure 5. Plots loglo k/ko vs. /I for ionic reactions of various types. The lines are drawn with slopes equal to Z Z (Laidler, 1965). A B 2+ 2+ _ A Co(NH3)5Br + Hg zAzB - 4 B s o 2' + I' z z = 2 2 8 A B c Co(OC2H5)N:N02 + OH zAzB = 1 D (c (UREA) )3+ + H o z z - o r 6 2 A B ‘ CH3COOC2H5 + OH zAzB = o + _ E H + Br + H202 ZAZB - -1 F Co(NH ) Br2+ + 0H" 2 z — -2 3 5 A B ’ 2+ 3- 35 catalytic ability of Cu(II) ions was dependent on the anion of the salt in the order CuC12>Cu(NO3)~CuSO4. Shtamm and co-workers (1974a,c) and Jameson and Black- burn (1976a,b) have also reported the catalytic proper- ties of Cu(II) ions in ascorbic acid degradation. Pek- karinan (1974) has reported evidence supporting Fe(III) catalysis of ascorbic acid. Other catalytic agents which have been reported include metal chelates, vanadyl and uranyl ions, and iron (III) chelates of aminopolycarboxlyic acids (Khan and Martell, 1968a,b, 1969). The mechanistic interpretation of metal catalyzed ascorbic acid degra- dation is varied, ranging from an ascorbate-metal-oxygen complex involving a one electron transfer to oxygen (Khan and Martell, 1967) to the formation of a metal-metal dinuclear-ascorbate-oxygen complex with a two electron transfer to oxygen (Jameson and Blackburn, 1975). Stability in Food Products The availability of kinetic data pertaining to the destruction of ascorbic acid in food products is much more limited than for destruction in model solutions. Singh et a1. (1976) studied the disappearance of ascorbic acid in a liquid infant formula as a function of initial oxygen concentration and light intensity at 7.2°C. The rate of ascorbic acid loss was found to be dependent on the presence of oxygen and obeyed overall second-order kinetics. If the oxygen level was maintained, pseudo 36 first-order loss of ascorbic acid was observed. The rate of ascorbic acid loss also increased linearly with light intensity up to 1600 lux. A companion study by Mack et a1. (1976) showed a first-order relationship for the uptake of oxygen in the same infant formula. Stability of ascorbic acid in canned tomato juice was studied by Lee et a1. (1977) as a function of pH, Cu(II) and temperature. First-order reaction kinetics were observed in the anaerobic system. The reported activation energy was 3.3 kcal/mole, which was less than that found by Pope (1972) for a similar tomato juice system. Lee et al. (1977) reported that the degradation rate of ascorbic acid reached a maximum at pH 4.06, and was linearly dependent on the concentration of copper at each pH. Karel and Nickerson (1964) reported the destruc- tion rates of ascorbic acid in dehydrated orange juice crystals increased with increasing aw, but no dependence on the storage atmosphere (air or vacuum) was observed. The authors concluded that there was no value of aw below which destruction of ascorbic acid ceased. Vojnovich and Pfeifer (1970) studied the sta- bility of reduced ascorbic acid in a corn-soy-milk (CSM) blend at three moisture contents. Lee and Labuza (1975) also studied the rate of ascorbic acid destruction in an intermediate moisture (0.32-0.84 aw) glycerol, corn oil 37 and microcrystalline cellulose model system. Both of these studies showed increased ascorbic acid degradation rates with increasing aw. Lee and Labuza (1975) also examined the effects of moisture sorption hystersis on ascorbic acid stability and found greater degradation of ascorbic acid on the desorption leg of the isotherm. However, activation energies calculated for ascorbic acid degradation were the same (~20 kcal/mole) for both legs of the isotherm. Recently, Waletzko and Labuza (1976) studied ascorbic acid stability in an intermediate moisture food system (aw=0.85) in an accelerated shelf-life test. They reported that ascorbic acid degraded rapidly by first- order kinetics whether packed in an air or Nz/H2 atmos- phere. Ascorbic acid was more stable, however, in the product packaged in the NZ/HZ atmosphere. The retention of ascorbic acid in tomato juice crystals as a function of aw was reported by Riemer (1977). First-order kinetics were obeyed for the loss of reduced ascorbic acid and dehydroascorbic acid. Activation energies ranging from 16.2 - 24.6 kcal/mole were calcu- lated. As with the study of Karel and Nickerson (1964) no effect of oxygen was observed. Ascorbic acid stability in products with high acidity may reflect a pH stabilizing effect. 38 The stability of ascorbic acid in foods is gen— erally improved at storage temperatures below 0°C. How- ever, Grant and Alburn (1965) reported the rate of ascorbic acid oxidation to be greater at —11°C than at 1°C in 0.02M acetate buffer of pH 5.0 and 5.5. Freeze concentration or the increase in concentration of the solutes in the frozen state over the unfrozen state has been presented as the most probable rationale for this phenomenon (Pin- cock and Kiovsy, 1966). The detailed study by Thompson and Fennema (1971) on the effect of freezing on ascorbic acid stability has shown the discrepancy to be a function of the solute concentration which governs the solubility of oxygen. Thus, the higher the solute concentration, which would be the normal pattern on freezing of foods, the less dissolved oxygen and the slower the oxidation rate of ascorbic acid at subfreezing temperatures. Predictions of Product Storage Stability The prediction of the storage stability of food products has received increased attention recently in view of nutritional labelling, cost and time requirements for conventional stability studies. Numerous mathematical techniques have been proposed for the prediction of pro- duct quality and deterioration rates of specific reaction types (e.g., nonenzymatic browning, lipid oxidation). 39 The most commonly used model is that which depends on a detailed understanding of the kinetics of the reac- tion being modeled. This requirement necessitates the experimental determination of those parameters affecting the reaction kinetics. Kwolek and Bookwalter (1971) developed a mathe— matical model for storage stability based on time-temper- ature data. The Arrhenius equation was satisfactory in predicting flavor and peroxide values. Mathematical models proposed for predicting the stability of space foods in semi-permeable packages, where lipid oxidation was the only deteriorative mechanism of concern, were sufficiently accurate to aid in package design (Karel and Labuza, 1969; Simon et al., 1971). Mizrahi et a1. (l970a,b) and Karel et al. (1971) developed a computer model for the prediction of the storage life of dehydrated cabbage as a function of non- enzymatic browning. The feasibility of using accelerated storage tests for determining the necessary kinetic param- eters have been evaluated. Conditions of high storage temperature and high moisture content provided experi- mental data in a short time that could be used to predict the degree of browning at ambient storage conditions. Correlation analysis between accelerated shelf life test data and long-term storage studies indicated that accel- erated storage tests could be utilized in the development of prediction equations. 40 The storage stability of potato chips that undergo loss of quality resulting from moisture adsorption and oxidative rancidity was studied by Quast et a1. (1972) and Quast and Karel (1973). A mixed model based on reaction kinetics and empirical data fitting was used to describe the deterioration process. The agreement between the actual storage tests and the computer prediction was good, although the procedure was less accurate for accel— erated test studies. Wanniger (1972) proposed a mathematical model for prediction of ascorbic acid storage stability using the equation ln k = -Ea/RT + a 1n H20 + b, where a and b are constants and the other symbols possess their normal meaning. Utilizing the data of Vojnovich and Pfeifer (1970), he reported excellent correlation between the experimental and predicted results. A basic computer model for simulation of nutrient stability in semi-permeable packages was postulated by Heldman (1974). The loss of the nutrient was considered to be a function of oxygen and water activity; storage temperature was not incorporated into the model. Lee et a1. (1976) developed a computer simulation for determining ascorbic acid stability in canned tomato juice. Copper‘ion concentration, pH and storage temper— ature were the parameters incorporated into the model. The experimental and predicted results were in excellent 41 agreement (i3%). The simulation also contained the addi- tional capacity to calculate the effect of seasonal variations as a result of storage temperature fluctuations using the Fourier transform series. Recently, Mizrahi and Karel (1977) proposed a "no model" accelerated test method for the determination of the deterioration rate of dehydrated products due to moisture sensitive reactions. This technique is based on theoretical considerations that the deterioration rate is inversely proportional to the rate of moisture gain at any given moisture content. This procedure was success- fully applied to the prediction of ascorbic acid loss in tomato power and browning in dehydrated cabbage. From this overview of methods for predicting product stability, it is obvious that presently used models generally rely on one index of quality, e.g. browning, lipid oxidation, nutrient loss, organoleptic properties. To date, modeling proposals have lacked general acceptance and utility. The reason for this appears to be their inability to consider simultaneous deteriorative reactions. EXPERIMENTAL PROCEDURES Model Food System Composition The composition of the dehydrated model food sys- tem is given in Table 1. The model system was designed to simulate a ready-to-eat breakfast cereal. Model Food System Preparation The ingredients, except coconut oil, were mixed dry in a ribbon blender, water was then added to give a slurry of approximately 40% total solids. The slurry was heated to 60°C, the coconut oil was then added and the system was homogenized in a Manton-Gaulin homogenizer at 2000 psig (lst stage, 1500 psig: 2nd stage, 500 psig). The pH of the homogenized slurry was 6.8. The model system for all studies was fortified with USP reduced ascorbic acid at a level of 25% NAS/NRC RDA (11.25 mg ascorbic acid) per 1009 model system (dry weight basis). The ascorbic acid was thoroughly mixed into the homogenized model system slurry after it had cooled to ambient temperature. Following fortification, the model system slurry was layered onto stainless steel freeze-drying trays, 42 43 placed in a Virtis Model FFD 42 WS Freeze-Dryer and frozen at a platen temperature of -40°C. The model sys- tem was then dried to 5p absolute pressure at a platen temperature of 110°C. Table 1. Composition of model food system Component %d Proteina 10.2 Fat 1.0 Carbohydrateb 76.6 Reducing SugarC 5.1 Sucrose 5.1 Salt 2.0 aSoya protein--Promine D. Central Soya. bFood Grade Powdered Starch--A.E. Staley, Inc., and Corn Syrup Solids 15 D.E., American Maize. cSupplied by the corn syrup solids, % dry weight basis. dCalculated on dry weight basis. Addition of Other Nutrients to the Model System Fortification of the dehydrated model food system with other nutrients for the interaction studies was accomplished in a manner similar to the addition of ascorbic acid. In the case of the multivitamin system, vitamin A, as retinyl acetate, was added to the model system slurry 44 prior to homogenization using the coconut oil as the carrier. Riboflavin and ascorbic acid were then added as aqueous solutions to the homogenized model system slurry after it had cooled to ambient temperature. The model system was then freeze—dried as described above. Both vitamin A (retinyl acetate, Sigma) and riboflavin (Sigma) were added at a level of 25% NAS/NRC RDA (0.287mg and 0.450mg, respectively) per 100g model system (dry weight basis). For the mineral fortification study, the indi- vidual mineral was added at the levels indicated in Table 2 to the homogenized model system slurry. All salts were of reagent grade quality. The model system was then freeze-dried as described above. Model Food System Equilibration Water activities for the model system were adjusted using equilibrium moisture content isotherm data (Figure 6) determined by Bach (1974) for the freeze- dried model system at 10, 20, 30 and 37°C using the method of Palnitkar and Heldman (1971). All experiments in this study were performed on the adsorption leg of the sorption hystersis loop. Equilibration was accom— plished by placing thin slabs of the freeze-dried model system into an equilibration chamber and forcing con- ditioned air of the desired aw and temperature from an Aminco-Aire unit through the closed system (Figure 7). 45 Table 2. Mineral supplement, source and %RDA employed in mineral fortification study. Mineral mg per Supplement source %RDA 1009 d.b. Control -- —- -- FeSO4 ' 7H20 Mallinckrodt 10 9.0 25 22.4 FeCl2 ' 4H20 Mallinckrodt 10 6 4 25 16 0 Fe Fisher 10 1.8 25 4.5 ZnCl2 Mallinckrodt 10 3.1 25 7.8 ZnSO4 ' 7H20 Mallinckrodt 10 6.6 25 16.5 ZnO Fisher 10 1.9 25 4.7 CaCO3 Mallinckrodt 10 300.3 25 750.5 CuCl ° 2H 0 Baker 2.7 2 2 6.7 CuSO ° 5H 0 Mallinckrodt 24 BO (3 MOISTURE_CONTENT an ID 46 Figure 6. 0.8 LO Adsorption isotherm for the dehydrated model food system at 20°C (from Bach, 1974). .Auflcs wufldnoocfl8¢ usmmwumwu Hwnfimno Hmupo paw umccflv Emumhm pOOM Hmpofi pmumuphzmp mo coflumunflaflsqm MOM Emummflc xOOHm .h muomflm mmmZfimU WV mmm2 mmmfimo 1‘ rlll 20351qu3on I mmHmHnHzommo 48 A dehumidifier and cooling coil were placed in the closed system when relative humidities were required which were lower than could be provided by the Aminco-Aire unit alone. Samples were equilibrated in approximately twenty-four hours. Model System Packaging and Storage Following equilibration of the model system to the desired equilibrium moisture content, the product was immediately packaged in either enameled metal containers (TDT-208x006 or 303- 303x406) or 1-oz. paperboard boxes. The metal containers prevented the transfer of air and moisture vapor into or out of the package. All containers were filled with the same mass of model system, approxi- mately 15g. This amount of the model system filled the TDT cans leaving no headspace in the container, whereas the 303 cans permitted a large headspace in the container. Paperboard l-oz. cereal boxes (3cm x 7cm x 10.3 cm) containing waxed liners (thickness 0.009cm) were packaged with unequilibrated freeze-dried model system. Moisture transfer coefficient for the liner and paperboard box 5 gHZO—cm/mz-h-mmHg) as plus liner were equal (7.25 x 10- reported by Purwadaria (1976). The packaged model system was then stored in constant temperature cubicles of the appropriate temper- atures. The cubicle temperatures were estimated to be il°C of the set temperatures of 10, 20, 30 and 37°C. 49 Relative humidities in the cubicles were estimated to vary no more than 12% of the 10, 40 and 85% RH storage conditions. Moisture Content Measurement Moisture content (dry weight basis) of equili— brated model system was determined by drying the samples in a vacuum oven at a vacuum of 28 inches Hg at the same temperature at which the product was equilibrated and stored. An ethanol-dry ice cold trap was inserted in the line between the vacuum oven and the vacuum pump to aid in the transfer of moisture from the product. Dry air was admitted into the vacuum oven at a rate of 15-20 ml/min to aid in the displacement of water vapor from the drying chamber. All samples were dried until they reached a constant weight. Using this method, the water activity of each sample could be determined to ensure that the aw had not changed during storage. The validity of this method for water activity determination of a large number of samples was confirmed by the vapor pressure manometric method of Sood and Heldman (1974). Ascorbic Acid Determination Ascorbic acid was measured by the continuous flow o-phenylenediamine micro-fluorometric procedure described by Kirk and Ting (1975). Ascorbic acid was extracted from the sample matrix with a 3% m-phosphoric acid--3% 50 glacial acetic acid solution. Samples were filtered and the extract introduced to a Technicon autoanalyzer. Dehydroascorbic acid was permitted to condense with o-phenylenediamine forming a fluorophor which was detected fluorometrically with an excitation wavelength of 360nm and emmission wavelength of 436nm. Total ascor- bic acid was measured as dehydroascorbic acid following oxidation of reduced ascorbic acid with 2,6-dichloro- indOphenol. Boric acid was utilized as the blanking reagent. Reduced ascorbic acid was determined from the difference of total minus dehyroascorbic acid. Data Analysis The loss of ascorbic acid was analyzed by the first-order kinetic equation for all storage conditions studied: _ d (C) _ —_dt_— - kC (18) where (C) = molar concentration of ascorbic acid; t = time (days); and k = first-order rate constant (days-l). Ascorbic acid levels for zero time storage were deter- mined after each aliquot of the dehydrated model system was equilibrated to the desired water activity. Ascorbic acid determinations were performed at preset intervals for a minimum of two half-lives, except for samples stored at 10°C in TDT cans which exhibited an extremely long half-life. 51 The temperature dependence for ascorbic acid degradation was analyzed according to the Arrhenius equation: k = A exp (-Ea/RT) (19) where k = first-order rate constant; A = Arrhenius pre- exponential; Ea = activation energy (cal/mole); R = gas constant (1.987 cal/OK-mole) and T = absolute temperature (OK). The reaction rate constants and the activation energies were calculated by linear regression analysis and by a computer program, KINFIT. The KINFIT program differs from the usual least squares techniques in that numerical integration pro- cedures are used to provide a fit to the desired dif- ferential equation (Dye and Nicely, 1971; Singh et al., 1975). 'This method of calculation assists in accounting for errors in vitamin assays and small variations in storage times. This program is specifically written for chemical reactions. Thermodynamic activation parameters for the destruction of ascorbic acid were calculated based on the theory of absolute reaction rates. The appropriate equations are outlined in the literature review. Experimental Design Various factors such as temperature, pH, oxygen, metal ion concentration and water aCtivity have been 52 reported as influencing the degradation rate of ascorbic acid in solution or dehydrated systems. As noted by Labuza (1968) and Borenstein (1971), a dearth of kinetic data exists on the storage stability of ascorbic acid and other nutrients in food systems. It was, therefore, the purpose of this study to examine some of these factors and obtain satisfactory kinetic data for ascorbic acid degradation in a dehydrated model food system. Ascorbic Acid Stability in TDT Cans The model food system was prepared, fortified with ascorbic acid only and equilibrated as previously described. Equilibration conditions were 0.10, 0.24, 0.40, 0.50 and 0.65 aw at 10, 20, 30 and 37°C. Following equilibration, the model system was then sealed in TDT cans and stored at the respective equilibration temperatures. This provided a total of twenty conditions with which to examine the effects of water activity and temperature on the storage stability of ascorbic acid. The 0.10 aw represents a condition below the BET monomolecular moisture content, 0.24 aw is the experimentally determined BET monomolecular moisture content, 0.40 and 0.50 aw are conditions in the monolayer region and 0.65 a approaches the capillary W region of the adsorption isotherm. 53 Ascorbic Acid--Vitamin Inter— actions in TDT Cans The model system was prepared, fortified with ascorbic acid and vitamins A and B2 and equilibrated as previously described. Equilibration conditions were 0.10, 0.24, 0.40, 0.50 and 0.65 aw at 10, 20, 30 and 37°C. Following equilibration, the model system was then sealed in TDT cans and stored at the respective equilibration temperatures. This provided twenty conditions with which to examine the effects of vitamin interactions, water activity and temperature on the storage stability of ascorbic acid. Ascorbic Acid Stability in Paperboard Boxes The model food system was prepared, fortified with ascorbic acid only and with ascorbic acid, vitamins A and B2 and packaged unequilibrated in 1—02. paperboard boxes with waxed liners. The packages were stored in cubicles at 10, 40 and 85% RH at 30°C. This permitted an examination of the effects of moisture vapor and air transmission on the storage characteristics of ascorbic acid. Ascorbic Acid Stability in 303 Cans The model food system was prepared, fortified with ascorbic acid only and equilibrated as previously described. Equilibration conditions were 0.10, 0.40 and 54 0.65 aw at 10, 20, 30 and 37°C. After equilibration, approximately 15g of the model system was then sealed in 303 cans and stored at the respective equilibration temperatures. This provided twelve conditions for the examination of the influence of a large gaseous oxygen reservoir on the storage stability of ascorbic acid. Ascorbic Acid--Mineral Inter- action in 303 Cans The model system was prepared, fortified with the minerals in Table 2 and equilibrated as previously described. Equilibration conditions were 0.10, 0.40 and 0.65 aw at 30°C. After equilibration, approximately 15g of the model system was then sealed in 303 cans and stored at 30°C. Zero, 10 and 25% RDA fortification levels were employed to evaluate the effects of mineral concentration at selected aw on the storage character- istics of ascorbic acid. In addition, the effect of the anion of the mineral employed could also be evaluated. RESULTS Ascorbic Acid Stability in TDT Cans The effects of water activity, moisture content and storage temperature on the stability of ascorbic acid were determined by studying the destruction of reduced (RAA), dehydro (DAA) and total (TAA) ascorbic acid in a dehydrated model food system. The model system was equilibrated to 0.10, 0.24, 0.40, 0.50 and 0.65 aw at 10, 20, 30 and 37°C, packaged in TDT cans, sealed and then stored isothermally at their respective equilibra- tion temperatures. The dependence of TAA and RAA destruction on aw at 30°C are shown in the first-order kinetic plots of Figures 8 and 9. At all aw and storage temperatures, the loss of TAA and RAA could be satisfac- torily described by first-order kinetics. Correlation coefficients associated with these plots were > 0.95. The first-order rate constants and half-lives for TAA and RAA degradation are presented in Table 3. The experimentally determined TAA and RAA concentrations at zero time storage were in good agreement with calcu- lated ascorbic acid levels at t = 0 computed by the 55 56 A 1.0 w .40 .8 014 o . ‘LJ‘O .6 - . 0.50 . . . 055 .4» C) .2- o ' A .' 1 4 1 n 20 40 60 80 Figure 8. DAYS Fraction of TAA remaining vs. time for selected at 30°C in the ascorbic acid only fortified model system stored in TDT cans. 57 0J0 0.24 AAO 0.50 0.65 .IL I J J 1 20 40 60 80 DAYS Figure 9. Fraction of RAA remaining vs. time for selected aw at 30°C in the ascorbic acid only fortified model system stored in TDT cans. 58 Table 3. Rate constants and half-lives for TAA and RAA loss as a function of water activity and storage temperature in ascorbic acid fortified dehydrated model food system packaged in thermal death time cans. Temp TAA RAA °C 3w b b a c a c k 0 t% k 0 t% 10 0.10 0.31 0.03 224 0.43 0.07 161 0.24 0.37 0.04 187 0.45 0.05 154 0.40 0.42 0.05 165 0.47 0.09 147 0.50 0.49 0.03 141 0.58 0.05 119 0.65 0.50 0.02 139 0.55 0.05 126 20 0.10 0.45 0.06 154 0.65 0.01 107 0.24 0.95 0.12 73 1.34 0.21 52 0.40 1.28 0.11 54 1.69 0.22 41 0.50 1.12 0.08 62 1.30 0.13 53 0.65 1.44 0.28 48 1.93 0.54 36 30 0.10 0.91 0.08 76 1.11 0.15 63 0.24 1.78 0.23 39 2.30 0.31 30 0.40 3.13 0.26 22 3.84 0.42 18 0.50 3.99 0.34 17 4.63 0.50 15 0.65 4.77 0.42 15 5.29 0.51 13 37 0.10 0.98 0.11 71 1.23 0.20 56 0.24 5.01 0.36 14 4.44 0.16 16 0.40 7.03 0.24 10 7.87 0.28 9 0.50 9.24 0.59 8 8.90 0.57 8 0.65 15.74 0.66 4 16.85 1.11 4 aFirst Order Rate Constant, x 10".2 days_1 (KINFIT analysis). b cHalf-Life, days. Standard Deviation, x 10- 2 59 KINFIT program. The standard deviation of the rate con- stants calculated by the KINFIT program was, in general, less than 10% of the respective rate constants. Data in Table 3 show that the stability of RAA and TAA in the model system stored in TDT cans decreased with increasing aw and storage temperature. Rate con- stants for RAA destruction are slightly greater than those describing TAA losses at the same conditions. Little significance is placed on this rate difference because of the standard deviations associated with the rate constants describing RAA losses are in part a function of the RAA determination (TAA-DAA=RAA). The stability of DAA showed a strikingly dif- ferent pattern than that observed for RAA and TAA. Dur- ing the initial portion of the storage study, the con- centration of DAA increased at 10°C, 0.10 and 0.24 aw and 20°C, 0.10 aw followed by a decrease after approxi- mately twenty-five days of storage (Figure 10). No increase in DAA levels were observed at the other storage conditions and DAA loss at the other storage conditions appeared to fit a first—order equation. The rate con- stants for the loss of DAA, however, were not calculated for reasons which will be presented in the discussion. The effect of temperature on the rate constants describing the loss of TAA as a function of aw at 10, 20, 30 and 37°C is shown in Figure 11. A similar temperature 6O .mcwo Boa cfl pmuoum Ewumwm Hmpoe pwflmauuow waco pflom cannoomm on» :w moms» umummswu mmmuoum pwuomawm um 03 . o... at A A. ... 1B 983% can. “Vb 3m oa.o How mafia .m> mcflcwmfimu < 0.95 and the standard deviations of the rate constants as calcu- lated by the KINFIT computer program were less than 10%. The degradation rate of TAA and RAA in the multivitamin fortified model system increased with increasing aw, similar to that found for the model system containing only ascorbic acid. There was no significant difference in rate constants for TAA and RAA loss. Model system containing vitamins A, B2 and C, which was equilibrated at aw of 0.10 to 0.65 and stored at 10°C in TDT cans showed apparent increases in RAA and TAA stability under these conditions (Table 5) when compared to the model system containing only ascorbic acid and stored under similar conditions (Table 3). The pattern of small initial accumulation of DAA at low aw and storage temperatures found in the ascorbic acid model system (Figure 10) was also observed for the 64 Table 5. Rate constants and half-lives for TAA and RAA loss as a function of water activity and storage temperature in multivitamin dehydrated model food system packaged in thermal death time cans. Temp a TAA RAA 0C w ka 0b tC ka 0b tc 8 15 10 0.10 0.18 0.05 385 0.14 0.04 495 0.24 0.31 0.03 224 0.33 0.12 210 0.40 0.43 0.03 161 0.41 0.11 169 0.50 0.46 0.06 151 0.52 0.22 133 0.65 0.45 0.04 154 0.54 0.07 128 20 0.10 0.49 0.10 141 0.84 0.14 83 0.24 1.03 0.13 67 1.75 0.34 40 0.40 1.46 0.21 47 2.46 0.78 28 0.50 1.18 0.11 59 2.43 0.37 28 0.65 1.31 0.24 53 2.88 0.22 24 30 0.10 1.03 0.69 67 1.70 0.38 40 0.24 3.63 0.28 19 3.43 0.24 20 0.40 3.46 0.30 20 3.81 0.52 18 0.50 3.28 0.30 21 4.40 0.54 16 0.65 5.96 0.77 12 7.38 0.85 9 37 0.10 0.98 0.19 71 1.21 0.09 62 0.24 3.14 0.19 22 -- -- -- 0.40 5.20 0.55 13 6.01 0.73 12 0.50 5.08 0.49 14 6.64 1.52 10 0.65 8.46 0.78 8 10.64 0.95 7 aFirst Order Rate Constant, x 10-2 days-1 (KINFIT analysis). CHalf-Life, days. bStandard Deviation, x 10- 2 65 multivitamin fortified model system. At higher temper- atures, the concentration of DAA decreased steadily with time as noted previously. Activation energies calculated from the kinetics data describing the destruction of TAA and RAA in the multivitamin model stored in TDT cans are presented in Table 6. The values are similar to those calculated for the ascorbic acid fortified system with the exception of the 0.10 a condition (Table 4). Activation energies for W aw equal to or above the monomolecular moisture content averaged 16.5i2.8 kcal/mole for TAA loss and 16.0:3.l kcal/mole for RAA loss. Table 6. Activation energies for TAA and RAA loss in multivitamin fortified dehydrated model system packaged in TDT cans. Activation Energy a (kcal/mole) w TAA RAA 0.10 11.5 13.1 0.24 13.5 11.8 0.40 16.8 16.6 0.50 15.9 16.1 0.65 20.0 19.3 Ascorbic Acid Stability in Paperboard Boxes Two model systems were used in this study, one fortified with ascorbic acid only and the other with 66 vitamins A, B2 and ascorbic acid. Both model systems were packaged, unequilibrated, in paperboard boxes con— taining waxed liners and stored in cubicles at 10, 40 and 85% relative humidity at 30°C. This study was designed to evaluate the influence of air and moisture vapor trans- mission on the storage stability of ascorbic acid. Destruction of TAA in the ascorbic acid only model system stored in boxes at 10, 40 and 85% RH are described by first-order kinetics (Figure 12). Similar conformity to first-order kinetics were found for the loss of RAA in the ascorbic acid only fortified model system and the loss of both TAA and RAA in the multi- vitamin fortified model systems. It is recognized that the moisture content of the boxed model systems changed during storage and the effect of this change on the destruction rate of ascorbic acid is discussed later. The calculated first-order rate constants and half—lives for the degradation of TAA and RAA in the ascorbic acid and multivitamin fortified dehydrated model systems packaged in paperboard boxes are presented in Table 7. Higher relative humidities were associated with greater destruction rates of TAA and RAA. Little difference is noted in the destruction rates for TAA versus RAA in the ascorbic acid only system. There is a slightly greater rate of loss of RAA in the multivitamin fortified system. 67 LO 6 .2 a o < .l a \ < .06 %RH a 010 A40 .02 085 l2 24 36 DAYS Figure 12. Fraction of TAA remaining vs. time at selected relative humidities at 30°C in the ascorbic acid only fortified model system packaged in paperboard boxes. 68 Table 7. Rate constants and half-lives for TAA and RAA loss at 30°C as a function of water activity in ascorbic acid and multivitamin fortified dehydrated model food systems packaged in paperboard boxes. Model TAA RAA System % RH b b a c a c k 0 t15 k 0 t8 Ascorbic 10 2.45 0.19 28 2.66 0.21 26 Acid Only 40 3.63 0.16 19 3.77 0.25 18 85 7.01 0.45 10 7.06 0.46 10 Multi- 10 2.01 0.34 34 2.93 0.24 24 vitamin 40 3.00 0.21 23 3.48 0.33 0 20 85 10.88 0.19 6 13.44 1.91 5 aFirst Order Rate Constant, x 10-2 days-1 (KINFIT analysis). b cHalf-Life, days. Standard Deviation, x 10- 2 69 Ascorbic Acid Stability in 303 Cans The degradation of total, reduced and dehydro ascorbic acid in dehydrated model food system packaged in 303 cans was studied as a function of water activity and storage temperature. The model system was equili- brated to 0.10, 0.40 and 0.65 aW at 10, 20, 30 and 37°C, sealed in 303 cans and stored isothermally at their respective equilibration temperatures. The experimental results for the loss of TAA stored in 303 cans at 0.10, 0.40 and 0.65 aw at 20°C and at 10, 20, 30 and 37°C at 0.40 aw are presented in Figure 13. These data conformed to the first—order kinetic function as did the destruction data at all aw and storage temperatures. Figure 13 shows the dependence of TAA stability on aw as well as temper- ature. A similar first-order dependence was found for the experimental data for RAA destruction in the model system stored in 303 cans. The first-order rate constants and half—lives for TAA and RAA degradation in the model system stored in 303 cans are presented in Table 8. These data show an increase in the rate of TAA and RAA loss with increasing water activity at a constant temperature. No significant dif- ference is apparent in the rate constants for TAA and RAA destruction at the same storage conditions of temperature and water activity. 70 .mcmo mom cw pmuoum Emumwm H0605 Mano pwom UHnHoomm on» now So ov.o um mupumhwaEwu Any can Doom um mufi>wuom kum3 “my mo :ofluocpw 6 mm mfifiu .m> mcflcflmfimu dds «0 cowuomum .MH musmflm m> 0.93 and the standard deviation of the rate constants were approximately 10% of the rate constants. The experimentally determined rate constants and half-lives are presented in Tables 10 and 11 for TAA and RAA degradation, respectively. No difference in TAA and RAA destruction rates was observed, and a similar dependence on water activity was found as for the other studies. No rate enhancement in ascorbic acid destruction over the nonfortified system due to any of the minerals studied was observed for the 0.10 and 0.40 aw storage conditions. The only exception to this is the slight catalysis in the CuCl2 and CuSO4 fortified systems at 0.40 aw. At 0.65 aw, which is in the capillary region of the adsorption isotherm, a 2-3 fold increase in the degradation rate of TAA and RAA over the nonfortified 74 .msmn .mAAHumHm: n .H|m>mp~ ca x .ucmumcoo muwu Hmpuolumufimm m oo.a AH hm.m ow mm.H v ca oo.m ma m~.m am hH.H Omsu o hm.HH ma om.v om mm.H N ma Hm.m ma mm.m Hm mm.H HOSO v ¢m.>a um m>.m he wv.H mm m h hm.oa mm hm.~ we me.a ca oomu ma Hm.m mm om.m me mw.H mm mm mH.m Hm m~.~ mm vm.H oa ocw v mm.ha om om.~ on mh.H mm v v ow.mH em Hm.~ mv qv.a OH omcu m mo.m~ mm he.~ me mo.H mm N e mm.ha om mm.m om ov.a CH HUCN v m~.>a Hm m~.~ mv mm.H mm m mw.m om mm.~ mm Hm.H OH on A H¢.m em va.~ we mm.H mm N b mv.oa hm hm.~ mm mm.H OH Howm o mm.HH mm HH.~ mm Hm.H mm v w ~N.HH Hm o~.~ mm Hm.H oa omwm «a mo.m om om.m om Nm.H s: 0:02 x x x ucmewamasm <0”; 36 mw.o m ov.o m oa.o Hmuwcfiz .mcmo mom cw Doom um pmuoum Ewumwm poom H0005 pwumupxnmp a Ca >uw>fiuom uwumz 0cm noduwowmwuuow Hmumcwfi mo cofiuocsm 6 mm mmoH dds new mm>wanmamn paw mucmumcoo wumm .OH manna 75 .msmo .mmfifiumammn . imamleoH x .ucmumcoo mum“ nopuolumuHmm H m mH.m NH Na.m mm om.H v mH Nh.v pH oo.v om ov.H Omso m Ho.NH NH mm.m ow m>.H N NH em.m MN Ho.m em mN.H Hugo v Nn.>H NN HH.m mm Hm.H mN m n Hv.oH HN om.m mv Ho.H 0H ovmo SH om.m mN mv.N mv mm.H mN HN hm.m mN Nm.N oe eh.H CH OCN v mm.mH mN No.m mm >5.H mN v w MH.mH vN mm.N hm hm.H OH omCN m mH.mN mN hv.N mm MH.N mN N v oo.>H 5N mm.N mm mm.H CH Hucu v Hm.>H Hm ON.N Nv mm.H mN b mm.m wN vw.N Nm Nm.H OH on b hm.a wN aw.N ov Nn.H mN N b no.OH mN hm.N He on.H 0H Humm o mn.HH wN vv.N Nm wH.N mN v m m¢.HH mN bv.N ov vh.H OH omwh VH Ho.m 0N mv.m mm vo.N In 0:02 x x x an ax nu mx 2» mx Huom Hmum3 can coHumoHMHuHOM HwnmcHE mo coHuocsu 0 mm mmoH (dz HON mm>HHImHms can mucmumcoo mumm .HH mHnt 76 system is noted for each trace mineral added, regardless of form. The only exception is zinc oxide which did not exhibit catalysis. Thermodynamic Activation Parameters The thermodynamic activation parameters associated with the degradation reaction for ascorbic acid were calculated from the first-order rate constants for TAA and RAA destruction. The calculated entropy of activation (AS+), enthalpy of activation (AH+) and free energy of activation (AGI) for TAA and RAA degradation for the three temperature dependent studies are presented in Tables 12, 13 and 14. The free energy of activation for TAA and RAA loss in each study remains constant at the various water activities, whereas an increase in both the entropy and enthalpy of activation is noted with increasing water activity. 77 Table 12. Activation parameters for TAA and RAA degradation in ascorbic acid forti- fied model system stored in TDT cans. Ea AH+ AS+ AG+ aw ______ (kcal/mole) (kcal/mole) (e.u.) (kcal/mole) TAA 0.10 8.1 7.5 -43 20.2 0.24 15.9 15.3 -15 19.8 0.40 17.6 17.0 - 8 19.5 0.50 19.2 18.6 - 3 19.4 0.65 19.2 18.6 - 2 19.2 RAA 0.10 7.1 6.5 -46 20.1 0.24 14.2 13.6 -20 19.7 0.40 17.8 17.2 - 7 19.4 0.50 18.1 17.5 - 6 19.3 0.65 19.3 18.7 - 2 19.2 78 Table 13. Activation parameters for TAA and RAA degradation in multivitamin fortified model system stored in TDT cans. Ea AH+ AS+ AG+ aw ______ (kcal/mole) (kcal/mole) (e.u.) (kcal/mole) TAA 0.10 11.5 10.9 -31 20.2 0.24 13.5 12.9 -22 19.4 0.40 16.8 17.2 - 7 19.4 0.50 15.9 15.3 -14 19.5 0.65 20.0 19.4 1 19.1 RAA 0.10 13.1 12.5 -25 19.9 0.24 11.8 11.2 -28 19.4 0.40 16.6 16.0 -11 19.4 0.50 16.1 15.5 -13 19.3 0.65 19.3 18.7 - 1 19.0 79 Table 14. Activation parameters for TAA and RAA degradation in ascorbic acid forti- fied model system stored in 303 cans. Ea AHI AS+ AG+ (kcal/mole) (kcal/mole) (e.u.) (kcal/mole) TAA 0.10 10.7 10.1 -33 20.0 0.40 16.0 15.4 -13 19.4 0.65 18.3 17.7 - 5 19.2 RAA 0.10 10.7 10.1 -33 20.0 0.40 15.6 15.0 -15 19.5 0.65 17.0 16.4 - 9 19.2 DISCUSSION The storage stability of ascorbic acid in low moisture dehydrated food products was studied utilizing a model food system designed to simulate a ready-to-eat breakfast cereal. The model system provided a product of known composition with which to conduct the study. Model system preparation, as previously described, was relatively simple and ensured a homogeneous distribution of vitamins and minerals. Ascorbic Acid Stability in TDT Cans The degradation of total (TAA) and reduced (RAA) ascorbic acid in the low moisture dehydrated model food system are adequately described by first-order kinetics with correlation coefficients 2 0.95 (Table 3). This first-order dependence was observed at all storage con— ditions of water activity and temperature. Data in Table 3 show that the storage stability of TAA and RAA in the model system stored in TDT cans decreased as the water activity increased from 0.10 to 0.65 at each storage temperature. These data contradict the hypothesis 80 81 of Salwin (1959, 1962) that the BET monomolecular moisture content should represent the equilibrium moisture content for maximum storage stability of the product. The BET monomolecular moisture content has been determined to be equal to an aw of 0.24 in the dehydrated model system (Bach, 1974), yet ascorbic acid degradation was signifi— cantly reduced at 0.10 aw. Ascorbic acid degradation at a measurable rate at 0.10 aw also conflicts with the view of Fennema (1976), that reactions which depend on solva- tion would not be measurable at water activities below the monomolecular moisture content. The dependence of RAA loss on water activity has been observed by other investigators (Karel and Nickerson, 1964; Vojnovich and Pfeifer, 1970; Lee and Labuza, 1975) but their studies did not include aw below the BET monomolecular moisture content. The dependence of TAA loss on water activity and its conformity to first-order kinetics has not been reported. The rate constants for RAA loss are, in general, slightly greater than the rate constants for TAA loss at corresponding storage conditions. It is important to note that little significance can be attached to the difference in rate constants for TAA and RAA losses. Bauernfeind and Pinkert (1970) have summarized the possible degradation pathways for reduced ascorbic acid. A major route for RAA degradation at neutral pH 82 is through dehydro (DAA) ascorbic acid. Following the scheme shown in equation (20) it is evident that if DAA is more stable than k k RAA ef=ié DAA ——39 products (20) -1 RAA (k2 < kl)’ one would expect the rate constant describ— ing RAA loss to be significantly greater than that for TAA loss. In fact it would be fortuitous if TAA loss could be described by first-order kinetics. In this study, the rate of loss of TAA and RAA were essentially equal at corresponding conditions, suggesting that either RAA does not degrad via DAA, or that the degradation rate for DAA is greater than for RAA (k2 > k1). This latter view of ascorbic acid degradation is supported by the investigations of Khan and Martell (1967) in which they utilized the 2, 4-dinitrophenylhydrazone method of Roe (1943) to measure the formation of dehy- droascorbic acid with time. Examination of the data in Figure 10 shows that the concentration of DAA during the initial stages of the storage study increased followed by a subsequent decrease in DAA concentration at 10°C, 0.10 and 0.24 aw and 20°C, 0.10 aw. At all other conditions a constant decrease in DAA levels with time was noted. This is interpreted as further evidence that RAA degrades 83 principally via DAA (equation 20). Kinetic treatment of the data describing the destruction of DAA with time was not attempted because it had not been independently confirmed that RAA degraded solely via DAA in the dehy- drated model system. The data for DAA loss does suggest, however, that the rate of DAA degradation was faster than the rate of RAA degradation. The temperature dependence of TAA and RAA degra- dation in dehydrated model system stored in TDT cans was described by the Arrhenius equation (Table 4). The activation energies calculated for aw equal to and above the monomolecular moisture content (aw = 0.24) were in the range of 15-19 kcal/mole, and showed a slight depen- dence on the water activity. These values were not sig- nificantly different from the activation energies reported by Lee and Labuza (1975) for the destruction of RAA in an intermediate moisture model food system at aw 0.32-0.84 on both the adsorption and desorption legs of the sorption isotherm. The activation energy for TAA and RAA destruc- tion in the model system stored at 0.10 aw in TDT cans was 7-8 kcal/mole. This value was determined by the t-test to be significantly different from the activation energies at the other water activities. No stability studies of ascorbic acid in systems at water activities below the BET monomolecular moisture content have been reported. Lee et a1. (1977), however, reported the 84 activation energy for the anaerobic destruction of ascorbic acid in canned tomato juice (pH = 4.06) to be 3.3 kcal/mole. The observed change in activation energy for ascorbic acid destruction in the model system at the 0.10 aw storage condition could be interpreted as a change in degradative mechanism. Based on these limited data such a conclusion may be tenuous. Ascorbic Acid--Vitamin Interaction in TDT Cans Kinetics data were obtained for TAA and RAA loss in the same model system which was fortified with vitamins A and B2 and then stored in TDT cans. The calculated first-order rate constants and half-lives for TAA and RAA loss are presented in Table 5. Cor- relation coefficients for the determination of the rate constants were 2 0.95. In general, the stability of TAA and RAA in this multivitamin study showed a similar dependence on water activity as was found for the system containing only ascorbic acid. The degradation rate for TAA and RAA loss in the multivitamin fortified system increased with aw. As was found in the preceding study, the BET monomolecular moisture content did not prove to be the aw offering the greatest stability to ascorbic acid. There was no difference in degradation rates between TAA and RAA loss in the multivitamin study. DAA 85 levels were measurable and followed a pattern similar to that observed in the model system containing only ascorbic acid. The measurement of DAA concentration with time during the multivitamin study presents further supporting evidence that the principal degradative path- way for RAA in the model system was via DAA. As in the preceding study, the loss of DAA in the multivitamin fortified system could be inferred to be faster than the loss of RAA. Comparison of TAA and RAA degradation constants between the multivitamin model system and the model sys- tem containing only ascorbic acid showed no difference, except at 10°C and low aw where an apparent increase in ascorbic acid stability was observed in the multivitamin system. This difference may be due to the relatively long storage period which did not permit assays beyond one-half life at these conditions. From these data, it is concluded that neither vitamin A nor B2 significantly interacted with ascorbic acid to either catalyze or inhibit the degradation of ascorbic acid. The excep- tional stability of riboflavin in dehydrated products reported by Borenstein (1971) and Dennison et a1. (1977) support this conclusion. The temperature dependence of TAA and RAA degra- dation in the multivitamin fortified model system stored in TDT cans was described by the Arrhenius equation 86 (Table 6). The activation energies calculated for water activities equal to and above the BET monomolecular moisture content were 15-19 kcal/mole. These Ea values were not significantly different from the activation energies found for TAA and RAA degradation in the model system containing only ascorbic acid and stored in TDT cans. The higher Ea value at 0.10 aw for the multivitamin system over the model system with ascorbic acid only probably reflects the experimental imprecision associated with the long storage period. Ascorbic Acid Stability in Paperboard Boxes The experimental data for the degradation of ascorbic acid in model systems, which were fortified with ascorbic acid only and ascorbic acid, riboflavin and vitamin A, and stored in paperboard boxes, was treated by first-order kinetics (Table 7). The standard deviations associated with the rate constants were rela- tively large and the correlation coefficients were approx- imately 0.90. The model system was packaged unequili- brated and therefore the moisture content and water activity changed during storage. Experimental data in Figure 14 reported by Purwadaria (1976) using the same model system and boxes demonstrated the time dependence required for the unequilibrated model system to reach moisture content equilibrium with the storage atmosphere 24 5; I MOISTURE CONTENT E t Figure 14. 87 A {A ‘afdrf” A A A A % RH . ‘0 10 u 40 A 85 . ’ ° ' o ' ' 4 8 l2 l6 2O WEEKS Moisture content equilibration during storage at selected relative humidities in dehydrated model systems packaged unequilibrated in paperboard boxes (from Purwadaria, 1977). 88 at 30°C. The rate of ascorbic acid degradation in the previous studies was found to be dependent on aw, there- fore, the rate of TAA and RAA loss in the model system packaged in paperboard boxes would be expected to increase with time until the moisture in the model system achieves equilibrium with the storage atmosphere. Treatment of the experimental data for TAA and RAA loss in the model system stored in paperboard boxes at 10% RH by first—order kinetics is believed to result in little error in the calculated destruction rates because of the short storage period required for moisture equilibration at this condition (Figure 14). Determi- nation of the rate constants for ascorbic acid destruc— tion in the boxed model system stored at 40 and 85% RH present a more complex situation. For simplicity, the experimental data for TAA and RAA destruction in the paperboard boxes at 40 and 85% RH were analyzed by the first-order rate function in order to compare these data with similar rate data obtained for the model system stored in TDT cans at constant water activity. The experimental first-order rate constants and half-lives for TAA and RAA loss in both model systems stored at 10, 40 and 85% RH at 30°C in paperboard boxes are presented in Table 7. The stability of ascorbic acid was dependent on the relative humidity of the storage atmosphere and there was little difference in 89 degradation constants for TAA and RAA loss either in the ascorbic acid only fortified model system or the multi- vitamin fortified model system. These results are similar to those found in the previous studies for ascorbic acid stability in TDT cans. There was a marked difference in the half-lives for TAA and RAA loss in the model system packaged in paperboard boxes at 10% RH and 30°C versus the TDT cans at 0.10 aw and 30°C (28 vs. 76 days, respectively). This observed difference in half-lives decreases as the rela- tive humidity of the storage atmosphere increases. The use of the paperboard boxes permitted relatively free transmission of atmospheric gases and moisture vapor across the waxed liner as shown by the moisture vapor 5g H20 - cm/m2 - h mmHg) transfer coefficient (7.25 x 10- reported by Purwadaria (1976). Since the product was packaged unequilibrated at a very low moisture content, the driving force for moisture equilibration would be into the product resulting in a gradual increase in the moisture content. Under these conditions, the observed rate of TAA and RAA loss at 40 and 85% RH should be less if the system had been equilibrated to these aw conditions prior to storage. The increasing moisture content and aw (Figure 14) with storage time are inter- preted to account for the differences in the observed destruction rates of TAA and RAA in the model system 9O stored at 40 and 85% RH and 30°C in the boxes and the model system packaged in TDT cans at 0.40 and 0.65 aw and stored at 30°C. The dramatic differences observed for TAA and RAA stability in the model food system equilibrated to 0.10 aw and stored in TDT cans at 30°C versus the model system packaged in boxes and stored at 10% RH at 30°C cannot be accounted for on the basis of moisture content or aw differentials. Equilibration of the boxed model system at 10% RH is achieved in about two weeks (Figure 14), yet the rate of loss of TAA and RAA in the model system is 2-3 times faster than in the TDT containers. Since the model systems were identical in composition, essentially of equal aw, and the transfer of atmospheric gases could occur only in the boxed model system, these data suggest the involvement of oxygen in the stability of TAA and RAA in the dehydrated food system at neutral pH. The influence of oxygen on the stability of ascorbic acid has been reported by Waletzko and Labuza (1976) in an intermediate moisture model system at 0.85 aw. Karel and Nickerson (1964), however, reported no significant difference in the rate of ascorbic acid destruction in dehydrated orange juice crystals stored in air or vacuum. Data from this latter study probably reflects the stabilizing influence of low pH and/or the chelation of metal ions 91 by the acids present in the soluble solids of the dehy- drated orange juice crystals. Ascorbic Acid Stability in 303 Cans The comparison of the destruction rate at 30°C for TAA and RAA at 10% RH in the boxed model system ver— sus the 0.10 aw condition in the TDT cans suggested the involvement of oxygen in the degradative process. This led to a storage study in which a large reservoir of gaseous oxygen could be provided with a constant water activity. The 303 can was selected as the storage con- tainer and was filled with the model food system forti- fied with ascorbic acid equal to the mass (”15g) utilized in the TDT can study. The model system stored in 303 and TDT cans at all storage conditions was identical in composition and source of ingredients. The water activi- ties and moisture contents of the two were identical within experimental error; therefore, there should be no effect of viscosity, reactant mobility or new catalytic sites on the stability of ascorbic acid. The only vari- able in the two systems is the relative amount of gaseous oxygen at any condition of temperature and water activity. The first-order degradation constants and half- lives for TAA and RAA loss in the dehydrated model system packaged in the 303 metal cans are presented in Table 8. The dependence of the rate constants for TAA and RAA 92 degradation in the 303 cans on water activity and storage temperature was similar to that found for ascorbic acid destruction in the TDT cans. The experimental rate con- stants for TAA and RAA loss in the 303 cans were found to be greater than the corresponding rate constants in the TDT cans. The relationship between the rate constant and aw was shown to be linear between 0.10-0.65 aw and is valid for water activities below the capillary region of the adsorption isotherm for the model system at all temper- atures studied (Figure 15). Mathematically, this function takes the form kobs = k0 + (slope) aw (21) where k0 is the intercept at zero water activity. This relationship was also applied to the rate constants for ascorbic acid degradation in the TDT can study. Although the rate constants for TAA and RAA loss in the TDT can are less than the corresponding rate constants for the same product in 303 cans, they exhibit the same linear vs. a (Figure 15). relationship of kobs w Calculation of the slopes obtained for the observed rate constant vs. aw function (equation 21) at constant temperature are presented in Table 15 for TAA and RAA destruction in the model system stored in 303 93 .Auohm D .om 4 .ON 0 .OH 0v memo Amy Hoe pcm A5 mom :H musumuwmswu wmmuoum ucmuwcoo um >UH>Huom kumz .w> :oHumpmuqmp «NB NC mumu mcu mo QHnmcoHuMme 34 mo To Nd mo To Nd 11X? b\\\|q|\|\\# . .N . .¢ 1 .m . .m A .0. mA 4 .m. .mH wusmHm le>1 2. ( sfiop) 94 . «BAH. 55.4- . mmmH. m~.mu nm\aoa Hm o mmsH. hm.H mm o HASH. mv.H- nm\mom . mane. sm.¢ . mmso. mm.H om\eoe mm 0 mass. -.m me H «who. mo.m om\mom mH.H HmHo. hm.o mm.H moHo. sm.v om\aoe mmmo. vo.m mHmo. mH.m o~\mcm . omoo. om.m . smoo. om.~ OH\aoa Hm m mmoo. Hm.~ mm H mwoo. mm.~ OH\mom oHumu onHm mIOH x x oHumu onHm mIOH x ox Uo\chHmucoo a $9 .mumchucoo 909 can mom :H pmuoum Emumwm HmGOE CH :oHumpmHmmp <0H>Huom nwuw3 .m> ucmumcoo mama mo conmmHmmu HmwcHH .mH mHnme 95 and TDT cans. The ratio of the slopes for the 303 to TDT containers decreased with increasing temperature, and at 37°C the difference in slopes are within experimental error. This relationship parallels the decrease of oxygen solubility in water with increasing temperature. The meaning of k0, which is obtained from equation 21 would be the degradation constant at 0.0 aw at the respective temperatures. The relatively large standard deviations associated with the ko values are a function of the limited number of data points and the experimental precision and are believed to be the reason for the dis- parity in these values at the various temperatures. The 303-TDT storage containers represent a closed system, mass may neither enter nor leave the container; however, mass may transfer across phase boundaries within the container. The transfer of gaseous oxygen into the product moisture at a given water activity and temperature would be governed by the equilibrium constant K = (22) where (02)d and (02)g represent the concentration of dis- solved and gaseous oxygen, respectively. As the dis- solved oxygen in the moisture of the dehydrated model system is consumed via the destruction of ascorbic acid, 96 the concentration of dissolved oxygen would be maintained in accordance with this equilibrium constant. The model system used to study the stability of TAA and RAA was prepared as described in the methods section, equilibrated to the appropriate aw and packaged in 303 cans and TDT cans, each with an equal mass of product (15g). This left no headspace in the TDT con- tainer, although a limited amount of gaseous oxygen was present in the intra and intersitial spaces of the pro- duct. The maximum concentration of dissolved oxygen that could be present in any of the equilibrated model system used (303 or TDT can studies) was estimated assuming a moisture content of 10%, which corresponds to an aw of 0.65 at 37°C. Using the solubility of oxygen in water at 0°C (4.89 cm3/100cm3), even though this temperature will yield an artificially high value, the maximum oxygen content from air which could be present in the moisture of the model system would be 4.36 x 10'7 moles of oxygen/cm3 of water. Assuming complete solubility, calculation of the moles of ascorbic acid in the fortified model system yielded 6 moles of ascorbic acid/cm3 of water in the 6.10 x 10' model system at 0.65 aw and 37°C (based on 11.25 mg ascorbic acid/100 9 model system, dry weight basis). These data show that the maximum dissolved oxygen content was approximately a factor of 10 less than the 97 theoretical stoichiometric ratio of one mole of oxygen/ mole of ascorbic acid reported by Hand and Greisan (1942) for the oxidation of ascorbic acid in a system with pH = 7.0. Thus, all equilibrated samples used in this study, whether packaged in cans or boxes contained dis- solved oxygen at a concentration less than that required for a l to 1 mole ratio of oxygen to ascorbic acid. No headspace was left in the TDT can after filling it with model system. Thus, the moles of gaseous oxygen which could be present in the TDT cans would vary depend- ing upon the inter and intrastitial spaces in the packaged model system. Using estimates that 20 to 40% of the total volume of the TDT can could be occupied by air, the total moles of gaseous oxygen were calculated and 5 to 5.4 x 10-5. This would repre- ranged from 2.8 x 10- sent approximately a lO-fold excess of moles of gaseous oxygen/mole of ascorbic acid in the 159 of model system in the TDT can. The moles of gaseous oxygen present in the 303 can were estimated to be 5.1 x 10-3, provid- ing an approximately 1000 fold excess of gaseous oxygen/ mole of ascorbic acid in the 303 can. Thus, for an equal mass of product, the dissolved oxygen concentration would be more nearly maintained at its initial level in the container with the larger headspace. Oxygen solubility in pure water is known to decrease with an increase in temperature. Joslyn and 98 Supplee (1949) have reported that the solubility of oxygen decreases as a function of soluble solids in pure sugar solutions at constant temperature. There- fore, as a function of temperature and soluble solids the dissolved oxygen content in the moisture of the model food system would be expected to be significantly lower than that calculated for pure water. Assuming that the ascorbic acid content (moles/ml) in the moisture of the food product can only decrease in concentration with increasing water activity and that the concentration of dissolved oxygen (moles/ml) would be constant regard- less of the water activity, the initial ratio of dissolved oxygen to ascorbic acid would increase as a function of water activity. This relationship was approximated by assuming oxygen solubility in the model system to be comparable to that in pure water, and plotting the (Oz)/(RAA) ratio against aw (Figure 16). A sharp increase in the slope of the curve is noted in the region cor- responding.to capillarity of the adsorption isotherm and suggests an increase in the rate of ascorbic acid degradation not in accord with equation 21. This hypothe— sis is supported by the ascorbic acid stability data reported by Lee and Labuza (1975) for an intermediate moisture food system equilibrated on either the adsorption or desorption leg of the equilibrium moisture sorption isotherm. Lee and Labuza (1975) found an increase in 99 .HAH>Huom umumz mo coHuocsu 6 mm H<Huom umuw3 co Amy :oHum>Huom mo amHmcucm can Afiv coHum>Huom mo amonucm mo mocmpcwmoo .mH musmHm ..< 0.0 0.0 ¢.0 N.0 0 0.0 0 In J O—I O a J O. 4ON: V V H S 3- of d 0 .im_ nAunmI Q 0 16 0 low low- 0 m 1mm 4 Low. 110 .H 40 mCmo mom pCm HOV mCmo Boa CH pmuoum Emumwm Hmpoa pmumupwnmp OCH CH mmoH «dz How >HH>Huom Hmum3 :0 Amy CoHum>Huom Ho aaHmcqu UCm AHuom mo amouqu mo mocmpcwmwo .mH musmHm :4. Rb mYo vAV N0. 0 0A0 0H0 va _N0 0 O a O 40 40.- c .mv- gAUMwI V V H S .8- .8. .0. 40m: 4 EON 10v- 0 m .mm 4 .Om- 111 20 - l6- 4‘ 4» I b d l2' 8 . H—I 44L HI .j_l J -40 -20 O # AS Figure 20. Isokinetic relationship of TAA degradation in dehydrated model system stored in TDT cans (A) and 303 cans (A). 112 Figure 21. i 1 -4O {-20 AS Isokinetic relationship of RAA degradation in dehydrated model system stored in TDT cans (A) and 303 cans (fi). 113 free energy of activation indicates the degradation reaction of ascorbic acid in the model system, in all likelihood, follows one mechanism. The data presented earlier for the involvement of oxygen in the degradation of ascorbic acid suggest a possible rationale for the observed dependence of AS+, AH+ and ACT on aw. The entropy of activation is a measure of the randomness of the activated complex rela- tive to the reactants. A negative A81 represents a loss in entropy or loss of freedom as the reactants proceed through the transition state, whereas a positive AS+ suggests a loosely bound activated complex and a gain in freedom. A decrease in AS* with increasing water activity may be due to increased solvation of the reac- tants and the activated complex or a decreased effective charge on the transition state. At 0.10 aw in the model system, the ASI of -43 e.u. obtained for TAA degradation in the TDT can (Table 12) indicates a loss in entropy and the formation of a tightly bound activated complex. As the system approaches 0.65 aw, the entropy of acti- vation approaches zero, suggesting little or no dif- ference in the internal degrees of freedom of the tran- sition state versus the reactants. The experimental data for ascorbic acid degradation in the model system show a dependence on the availability of oxygen. Thus, 114 a possible reaction intermediate would be the ascorbic acid-oxygen complex postulated by Khan and Martell (1967). Since the free energy of activation is constant, the change in A81 could be interpreted as an increase in degree of solvation of both the reactants and transition state, as the water activity increases to 0.65. In a like manner, the change in AHI as a function of aw reflects increased solvation of the reactants and tran- sition state with increased water activity, but the degradation of ascorbic acid in the model system is entropy controlled. Little experimental data are available from the literature which may be treated in a similar manner. The activation parameters of Lee and Labuza (1975) for the degradation of ascorbic acid in an intermediate moisture system (aw = 0.32 - 0.84) have been calculated and are presented in Table 16 and Figure 22. The free energies of activation were about 18 kcal/mole compared to the 19 - 20 kcal/mole found in the present study. The enthalpy and entropy of activation were also constant at approximately 18 kcal/mole and zero e.u., respectively. The activation parameters calculated for the adsorption and desorption studies did not indicate a dependence on the sorption hystersis phenomenon. These calculations suggest that the degradation of ascorbic acid in the 115 .AmhmHv munnmq pCm 004 m0 mump Eoum pmpmeono muwumfimnmm CoHum>HuU¢m e.uH o.>H ¢+ v: m.mH m.mH vm.o m.nH m.>H MH+ m+ m.HN m.wH mh.o m.mH H.mH ml H+ h.mH ¢.mH no.0 u.wH >.mH HH+ o o.NN m.mH .Hm.o H.mH o.mH m+ H+ m.mH «.mH Nm.o CoHHQHOmmp m.>H v.5H N+ H+ w.hH w.hH vm.o m.hH m.>H w: NI m.mH N.5H mb.o v.mH «.mH 0 ml v.mH H.mH no.0 m.mH h.mH HH+ HI o.NN v.mH Hm.o N.mH H.mH 5+ m+ m.HN m.om Nm.o COHHQHOmpm N Con H CCH N CCH H Con N CCH H can 3 m HmHoa\Hmoxv +o< 1.6.mc +m< HmHoB\Hmoxv +m< m.Emum>m Hobos musumHOE mumemEHmuCH Cm CH CoHumpmHmmp pHom UHQHoomm pmoscmu How mumumfimumm COHum>Huo< .mH mHnt 116 20 no -20 :5 1..” AS Figure 22. Isokinetic relationship for ascorbic acid degradation in an intermediate moisture model system (data from Lee and Labuza, 1975). 117 intermediate moisture model system of Lee and Labuza (1975) followed a similar pathway as found in the present study. Limited data are available for the calculation of activation parameters for the degradation of ascorbic acid and other nutrients in low moisture food systems. However, information of this type may lead to a better understanding of the role of water in nutrient stability, nonenzymatic browning and lipid oxidation. SUMMARY AND CONCLUSIONS The storage stability of total, reduced and dehydroascorbic acid in a low moisture dehydrated model food system was investigated as a function of storage temperature, water activity, oxygen availability, vitamin interactions and trace mineral catalysis. The destruction of TAA and RAA was dependent on the water activity and storage temperature, and its loss could be adequately described by first-order kinetics. Maximum stability of TAA and RAA was observed at low storage temperatures and 0.10 aw, which was below the BET monomolecular moisture content. A linear relation- ship between the degradation constants for TAA and RAA loss with water activity was found, which contradicts the hypothesis of Salwin (1959) that the BET mono- molecular moisture content should correspond to the water activity of greatest stability. The data pre- sented tends to support the suggestion by Karel and Nickerson (1964) that there is no water activity below which the degradation of ascorbic acid ceases. 118 119 Dehydroascorbic acid was measurable during the storage study. DAA concentration decreased at storage temperatures and water activities above 20°C and 0.24 aw, respectively. An exception was noted at 10°C and low aw in which there was a small initial increase in DAA con- centration followed by a gradual decrease after 25 days. It was concluded that the RAA degraded to DAA, which underwent further destruction. Oxidation reactions in dehydrated food systems have been shown to be a function of reduced viscosity promot- ing mobility of reactants, dissolution of precipitated catalysts and swelling of solid matrices exposing new catalytic sites (Labuza et al., 1970). These factors certainly explain to some extent the increase in reaction rates for TAA and RAA in dehydrated food systems. How- ever, the identical model system was prepared and equil- ibrated at the same aw, and stored in TDT and 303 cans so that the only variable was the amount of gaseous oxygen available. Factors other than oxygen would have been internally compensated for in this study. The increased destruction rates for ascorbic acid in the 303 cans was attributed to the presence of the large gaseous oxygen reservoir in the 303 can and the main- tenance of the initial dissolved oxygen concentration according to the equilibrium constant. 120 The effects of added riboflavin and vitamin A on the storage stability of ascorbic acid was studied using the same model system. Comparison of degradation rates for TAA and RAA loss in the multi-vitamin fortified model system versus the ascorbic acid fortified model system showed no significant effect. The catalytic properties of added trace minerals (Fe, Cu, Zn, Ca) on the degradation of ascorbic acid was investigated at 30°C and three water activities. At and below 0.40 a no rate enhancement of ascorbic acid w' degradation was observed over the nonmineral fortified model system due to the presence of these minerals, except for copper at 0.40 aw. At 0.65 aw, which is in the capillary region of the adsorption isotherm, a 2-3 fold increase in degradation rate over the nonmineral fortified system was observed for each form of the added trace mineral. The lone exception was zinc oxide, which did not exhibit catalysis due to its inherent insolubility. The catalytic effect of Fe and Cu ions in the capillary region of the adsorption isotherm were explained on the basis of the mobility of these ions. Catalysis by soluble zinc electrolytes and calcium ions was attributed to the electrolyte effect on increasing the activity coefficient of oxygen, thus increasing the rate of ascorbic acid degradation. 121 Activation parameters were calculated for the degradation of ascorbic acid in the model system stored in TDT and 303 cans. Although a significant change in the energy of activation was noted as a function of water activity, the free energy of activation remained constant. It is concluded that the degradation of ascorbic acid in the dehydrated model system followed the same mechanism at all water activities studied. REFERENCES REFERENCES Acker, L. 1962. Enzymatic reactions in foods of low moisture content. Adv. Food Res. 11:263. Acker, L. 1969a. Enzyme activity at low water contents. Recent Adv. Food Sci. 3:239. Acker, L. 1969b. Water activity and enzyme activity. Food Tech. 23:1257. Acker, L. and R. Weise. 1972a. Lipase behavior in systems of low water content. I. Influence of the physical state of the substrate on enzymic lipolysis. Lebensmittel - Wiss. u. Technol. 5:181. Acker, L. and R. Weise. 1972b. Behavior of lipase in systems of low water content. II. Enzymic lipolysis in the range of extreme low-water activity. Z. Lebensmittelunters. u. - Forsch. 159:205. Adamson, A. W. 1976. Physical Chemistry of Surfaces, Wiley and Sons, New York. Aylward, F. and I. D. Morton. 1971. Vitamin fortifi- cation of foods. Proc. 3rd. Int. Congr. Food Sci. Technol. 192. Bauernfeind, J. C. and D. M. Pinkert. 1970. Food pro- cessing with ascorbic acid. Adv. Food Res. 18:219. Berlin, E., P. G. Kliman and M. J. Pallansch. 1966. Surface areas and densities of freeze-dried foods. Ag. Food Chem. 14:15. Blackadder, D. A. and C. Hinchelwood. 1958a. Kinetics of the rearrangement and oxidation of hydra- zobenzene in solution. (1). Rearrangement and spontaneous oxidation. 3. Chem. Soc. 2720. 122 123 Blackadder, D. A. and C. Hinchelwood. l958b. Kinetics of the rearrangement and oxidation of hydrazo- benzene in solution. (11). Catalyzed oxidation. 3. Chem. Soc. 2728. Bronsted, J. N. 1922. Theory of the rate of chemical reactions. Z. Physik. Chem. 102:169. Brunauer, S. 1945. The Adsorption of Gases and Vapors. Princeton University Press. Princeton, New Jersey. Brunauer, S., P. H. Emmet and E. Teller. 1938. Adsorp- tion of gases in multimolecular layers. J. Am. Chem. Soc. 66:309. Caurie, M. 1970. A new model equation for predicting safe storage moisture levels for optimum sta- bility of dehydrated foods. J. Fd. Technol. 5:301. Caurie, M. 1971a. A practical approach to water sorption isotherms and the basis for the determination of optimum moisture levels of dehydrated foods. J. Fd. Technol. 6:85. Caurie, M. 1971b. A single layer moisture absorption theory as a basis for the stability and avail- ability of moisture in dehydrated foods. J. Fd. Technol. 6:193. Coggiola, I. M. 1963. 2,5-dihydro—2-furoic acid: A product of the anaerobic decomposition of ascorbic acid. Nature, London. 200:954. Cox, E. G. and T. H. Goodwin. 1936. Crystalline structure of the sugars (III) ascorbic acid and related compounds. J. Chem. Soc. 769. Csuros, Z. and J. Petro. 1955. Autoxidation of ascorbic acid as a function of temperature. Acta Chim. Acad. Sci. Hung. 1:199. de Boer, J. H. and C. Zwikker. 1929. Adsorption as a result of polarization-adsorption isotherm. Z. Phys. Chem. 133:407. Dekker, A. D. and R. G. Dickinson. 1940. Oxidation of ascorbic acid by oxygen with cupric ion as catalyst. J. Am. Chem. Soc. 63:2165. 124 Dennison, D., J. Kirk, J. Bach, P. Kokoczka and D. Heldman. 1977. J. Food Process. Preserv. 1:43. Duden, R. 1971. Enzymic reactions in foods with very low water activity. Lebensmitte1.-Wiss. u. Technol. 6:205. Dye, J. L. and V. A. Nicely. 1971. A general purpose curve-fitting program for class and research use. J. Chem. Edu. 66:443. Duckworth, R. B. 1971. Differential thermal analysis of frozen food systems. I. The determination of unfreezable water. J. Fd. Technol. 6:317. Duckworth, R. B. and G. M. Smith. 1963. Diffusion of solutes at low moisture levels. Recent Adv. Food Sci. 6:230. Eichner, K. 1976. The influence of water content on non- enzymic browning reactions in dehydrated foods and model systems and the inhibition of fat oxi— dation by browning intermediates, in Water Relations of Foods: ed. R. B. Duckworth. Academic Press, New York. Exner, O. 1964. Enthalpy—entropy relation. Coll. Czech. Chem. Commun. 12:1094. Exner, O. 1973. The enthalpy-entropy relationship. Prog. Phys. Org. Chem. 11:411. Eyring, H. 1935. Activated complex and the absolute rate of reactions. J. Chem. Phys. 6:107. Feeney, R. E., G. Blankenhorn and H. B. F. Dixon. 1976. Carbonyl-amine reactions in protein chemistry. Adv. Prot. Chem. 16:135. Fennema, O. R. 1976. "Water and ice" in Principles of Food Science, pt. 1, Food Chemistry. ed. 0. R. Fennema, Marcel Dekker, Inc., New York. Finholt, P., R. B. Paulssen and T. Higucki. 1963. Rate of anaerobic degradation of ascorbic acid in aqueous solution. J. Pharm. Sci. 61:948. Fox, K. K., V. H. Holsinger, M. K. Harper, N. Howard, L. S. Pryor and M. J. Pallansch. 1963. Measure— ment of the surface areas of milk powders by a permeability procedure. Food Tech. 11:127. 125 Frenkel, Y. I. 1946. Kinetic Theory of Liquids. The Clarendon Press, Oxford. Grant, N. H. and H. E. Alburn. 1965. Fast reactions of ascorbic acid and H202 in ice, a presumptive early environment. Biochemistry 6:1913. Hand, D. B. and E. C. Greisen. 1942. Oxidation and reduction of vitamin C. J. Am. Chem. Soc. 66:358. Harkins, W. D. and G. Jura. 1944. A vapor adsorption method for determination of the area of a solid without assumption of molecular area. J. Am. Chem. Soc. 66:1366. Heldman, D. R. 1974. Computer simulation of vitamin stability in foods. Paper presented at seminar on Stability of Vitamins in Foods for the Associ- ation of Vitamin Chemists, Chicago, IL. Henderson, 8. M. 1952. A basic conception of equilibrium moisture. Agr. Eng. 66:24. Huelin, F. E. 1953. Studies on the anaerobic decompo— sition of ascorbic acid. Food Res. 16:633. Huelin, F. E., I. M. Coggiola, G. S. Sidhu and B. H. Kennet. 1971. The anaerobic decomposition of ascorbic acid in the pH range of foods and in more acid solutions. J. Sci. Fd. Agric. 66:540. Jameson, R. F. and N. J. Blackburn. 1975. The role of Cu-Cu dinuclear complexes in the oxidation of ascorbic acid by 02. J. Inorg. Nucl. Chem. 37:809. Jameson, R. F. and N. J. Blackburn. 1976a. Role of copper dimers and the participation of copper (III) in the copper catalyzed autoxidation of ascorbic acid. Part II. Kinetics and mechanism in 0.100 mol dm'3 potassium nitrate. J. C. S. Dalton 534. Jameson, R. F. and N. J. Blackburn. 1976b. Role of copper dimers and the participation of copper (III) in the copper catalyzed autoxidation of ascorbic acid. Part III. Kinetics and mechanism in 0.100 mol dm‘3 potassium chloride. J. C. S. Dalton 1596. 126 Joslyn, M. A. and H. Supplee. 1949. Effect of sugars on conversion of dehydroascorbic acid. Food Res. 16:216. Karel, M. 1973. Recent research and development in the field of low-moisture and intermediate foods. CRC Crit. Rev. Food Tech. 6:329. Karel, M. and J. T. R. Nickerson. 1964. Effect of relative humidity, air and vacuum on browning of dehydrated orange juice. Food Technol. 18:104. Karel, M., T. P. Labuza and J. F. Maloney. 1967. Chemical changes in freeze-dried foods and model systems. Cryobiology, 6:288. Karel, M., S. Mizrahi and T. P. Labuza. 1971. Computer prediction of food storage. Modern Pack. 66:54. Khan, M. M. Taqui and A. E. Martell. 1967a. Metal ion and chelate catalyzed oxidation of ascorbic acid by molecular oxygen. I. Cupric and ferric ion catalyzed oxidation. J. Am. Chem. Soc. 66:4176. Khan, M. M. Taqui and A. E. Martell. 1967b. Metal ion and metal chelate catalyzed oxidation of ascorbic acid by molecular oxygen. II. Cupric and ferric chelate catalyzed oxidation. J. Am. Chem. Soc. 66:7104. Khan, M. M. Taqui and A. E. Martell. 1968a. The kinetics of the reaction of iron(III) chelates of amino- polycarboxylic acids with ascorbic acid. J. Am. Chem. Soc. 66:3386. Khan, M. M. Taqui and A. E. Martell. 1968b. Kinetics of metal ion and metal chelate catalyzed oxi— dation of ascorbic acid. III. Vanadyl ion catalyzed oxidation. J. Am. Chem. Soc. 66:6011. Khan, M. M. Taqui and A. E. Martell. 1969. Kinetics of metal ion and metal chelate catalyzed oxidation of ascorbic acid. IV. Uranyl ion catalyzed oxidation. J. Am. Chem. Soc. 61:4668. Kirk, J. R. and N. Ting. 1975. Fluorometric assay for total vitamin C using continuous flow analysis. J. Food Sci. 66:463. 127 Kurata, T. and Y. Sakurai. 1967. Degradation of L-ascorbic acid and mechanism of non-enzymatic browning reaction. II. Nonoxidative degradation of L-ascorbic acid including formation of 3- deoxy-L-pentosene. Agric. & Biol. Chem. 61:170. Kurpianoff, J. 1959. "Bound water in foods," Funda- mental Aspects of the Dehydration of Foodstuffs, the Macmillian Company, New York, NY. Kwolek, W. F. and G. M. Bookwalter. 1971. Predicting storage stability from time-temperature data. Food Tech. 66:1025. Labuza, T. P. 1968. Sorption phenomena in foods. Food Tech. 66:263. Labuza, T. P. 1971. Kinetics of lipid oxidation in foods. CRC Crit. Rev. Food Tech. 6:355. Labuza, T. P. 1976. "Interpretation of sorption data in relation to the state of constituent water" in Water Relations of Foods. ed. R. B. Duck- worth. Academic Press, New York. Labuza, T. P., J. F. Maloney and M. Karel. 1966. Autoxi- dation of methyl linoleate in freeze-dried model systems. II. Effect of water on cobalt catalyzed oxidation. J. Food Sci. 61:885. Labuza, T. P., S. R. Tannenbaum and M. Karel. 1970. Water content and stability of low-moisture and intermediate moisture foods. Food Tech. 24:543. Labuza, T. P., S. Cassil and A. J. Sinsky. 1972. Sta- bility of intermediate moisture foods. 2: Microbiology. J. Food Sci. 66:160. Laidler, K. J. 1965. Chemical Kinetics, second edition, McGraw—Hill, New York. Langmuir, I. 1918. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 66:1361. Lee, S. H. and T. P. Labuza. 1975. Destruction of ascorbic acid as a function of water activity. J. Food Sci. 66:370. 128 Lee, Y. C., J. R. Kirk, C. L. Bedford and D. R. Heldman. 1977. Kinetics and computer simulation of ascorbic acid stability of tomato juice as functions of temperature, pH and metal catalyst. J. Food Sci. 66:640. Leffler, J. E. 1955. The interpretation of enthalpy and entropy data. J. Org. Chem. 61:533. Loncin, M., J. J. Bimbenet and J. Lenges. 1969. Influence of the activity of water on the spoilage of foodstuffs. J. Food Technol. 3:131. Mack, T. E., D. R. Heldman and R. P. Singh. 1976. Kinetics of oxygen uptake in liquid food. J. Food Sci. 61:309. Maloney, J. F., M. Karel, and T. P. Labuza. 1966. Autoxidation of methyl linoleate in freeze-dried model systems. I. Effect of water on the auto- catlyzed oxidation. J. Food Sci. 61:878. Mapson, L. W. 1945. Influence of halides on the gxi- dation of ascorbic acid. 2. Action of Cl on the cupric-cuprous system. Biochem. J. 66:228. Martin, M. F. 1958. Factors in the development of oxi- dative rancidity in ready to eat oat flakes. J. Sci. Food Agric. 6:817. Matz, S., C. S. McWilliams, R. A. Larsen, J. H. Mitchell, Jr., M. McMullen and B. Layman. 1955. The effect of variations in moisture content on the storage deterioration rate of cake mixes. Food Tech. 6:276. Miller, J. and M. A. Joslyn. 1949. Effect of sugars on the oxidation of ascorbic acid. III. Influence of pH and type of buffer. Food Res. 14:354. Mizrahi, S., T. P. Labuza and M. Karel. 1970a. Computer aided predictions of extent of browning in dehydrated cabbage. J. Food Sci. 66:799. Mizrahi, J., T. P. Labuza, and M. Karel. 1970b. Feasi- bility of accelerated tests for browning in dehydrated cabbage. J. Food Sci. 66:804. 129 Mizrahi, S. and M. Karel. 1977. Accelerated stability tests of moisture sensitive products in permeable packages by programming rate of moisture con- tents. J. Food Sci. 66:1575. Mousseri, J., M. P. Steinberg, A. I. Nelson and L. S. Wei. 1974. Bound water capacity of corn starch and its derivatives by NMR. J. Food Sci. 66:114. Multon, J. L. and A. Guillot. 1976. "Water activity in relation to the thermal in activation of enzymic proteins," in Water Relations of Foods. ed. R. B. Duckworth, Academic Press. New York. Ogata, Y., Y. Kosugi and T. Morimato. 1968. Kinetics of cupric salt-catalyzed autoxidation of L-ascorbic acid in aqueous solutions. Tetrahedron 66:4057. Palnitkar, M. P. and D. R. Heldman. 1971. Equilibrium moisture characteristics of freeze-dried beef components. J. Food Sci. 66:799. Pekkarinan, L. 1974. The mechanism of the autoxidation of ascorbic acid catalyzed by iron salts in citric acid solution. Finn. Chem. Lett. 233. Pincock, R. E. and T. E. Kiovsy. 1966. Kinetics of reactions in frozen solutions. J. Chem. Edu. 43:358. Pitt, J. I. 1976. "Xerophilic fungi and the spoilage of foods of plant origin," in Water Relations of Foods. ed. R. B. Duckworth. Academic Press, New York. Polanyi, M. 1928. Anwendurg der langmuirschen theorie auf die adsorption von gasen an holzkohle. Z. Physik Chem. A138:459. Pope, G. G. 1972. Effect of time, temperature and for- tification level on the retention of ascorbic acid in fortified tomato juice. Ph.D. thesis, Ohio State University, Columbus, Ohio. Purwadaria, H. K. 1977. Simulation of nutrient sta- bility in dry foods during storage. M. S. Thesis, Michigan State University, East Lansing, MI. Quast, D. G., M. Karel and W. M. Rand. 1972. Development of a mathematical model for oxidation of potato chips as a function of oxygen pressure, extent of oxidation and equilibrium relative humidity. J. Food Sci. 66:673. 130 Quast, D. G. and M. Karel. 1973. Simulating shelf-life. Modern Pack. 66:50. Reimer, J. 1977. Prediction of vitamin C. retention in a stored dehydrated food. Ph.D. thesis, Massachusetts Institute of Technology, Boston, MA. Reynolds, T. M. 1963. Chemistry of nonenzymic browning. I. The reaction between aldoses and amines. Adv. Food Res. 16:1. Reynolds, T. M. 1965. Chemistry of nonenzymic browning. II. Adv. Food Res. 16:167. Reynolds, T. M. 1969. Nonenzymic browning: sugar-amino interactions, in Symposium on Foods: Carbohy- drates and Their Roles, ed. H. W. Schultz, R. F. Cain and R. W. Wrolstad. Avi. Publishing Company, Inc. Westport, Conn. Rockland, L. B. 1957. A new treatment of hygroscopic equilibria. Food Res. 66:604. Rockland, L. B. 1969. Water activity and storage sta- bility. Food Tech. 66:1241. Salwin, H. 1959. Defining minimum moisture contents for dehydrated foods. Food Tech. 16:594. Salwin, H. 1962. The role of moisture in deteriorative reactions of dehydrated foods. Proceedings of Conference on Freeze-Drying of Foods. National Academy of Sciences National Research Council. Salwin, H. 1963. Moisture levels required for stability in dehydrated foods. Food Tech. 11:1114. Scatchard, G. 1922. Statistical mechanics and reaction rates in liquid solutions. Chem. Rev. 16:229. Schaleger, L. L. and F. A. Long. 1963. Entropies of activation and mechanism of reactions in solutions. Prog. Phys. Org. Chem. 1:1. Schultz, H. W., E. Day and R. Sinnhuber. 1962. Lipids and their oxidation. Avi. Publishing Company, Inc. Westport, Conn. Scott, W. J. 1957. Water relations of food spoilage organisms. Adv. Food Res. 1:83. 131 Shanbhag, S., M. P. Steinberg and A. I. Nelson. 1970. Bound water defined and determined at constant temperature by wide-line NMR. J. Food Sci. 35:612. Shtamm, E. V. and Yu. I. Shurlator. 1974a. Catalysis of the oxidation of ascorbic acid by copper(II) ions. I. Kinetics of the oxidation of ascorbic acid in the copper(II) - ascorbic acid - molecular oxygen system. Zhur. Fiz. Khin. 66:1454. Shtamm, E. V. and Yu. I. Skurlator. 1974b. Catalysis of the oxidation of ascorbic acid by copper(II) ions. IV. Kinetics of the oxidation of ascorbic acid in the copper(II) - ascorbic acid - hydrogen peroxide system. Zhur. Fiz. Khin. 66:1857. Shtamm, E. V., A. P. Purmal and Yu. I. Skurlator. 1974a. Catalysis of the oxidation of ascorbic acid by copper(II) ions. II. Anaerobic oxidation of ascorbic acid by copper(II) ions. Zhur Fiz. Khin. 66:2229. Shtamm, E. V., A. P. Purmal and Yu. I. Skurlator. 1974b. Catalysis of the oxidation of ascorbic acid by copper(II) ions. III. Mechanism of oxidation by molecular oxygen. The role of copper(I) ions. Zhur. Fiz. Khin. 66:2233. Simon, I. B., T. P. Labuza, and M. Karel. 1971. Computer aided prediction of food storage stability: oxidative deterioration of a shrimp product. J. Food Sci. 66:280. Singh, R. P., D. R. Heldman and J. R. Kirk. 1976. Kinetics of quality degradation: ascorbic acid oxidation in infant formula during storage. J. Food Sci. 61:304. Sood, V. C. and D. R. Heldman. 1974. Analysis of a vapor pressure manometer for measurement of water activity in nonfat dry milk. J. Food Sci. 66:1011. Stevens, H. H. and J. B. Thompson. 1948. The effect of shortening stability on commercially produced army ration biscuits. II. Development of oxi- dation during storage. J. Amer. Oil Chemists Soc. 66:389. 132 Szent-Gyorgi, A. 1928. Observations on the function of peroxidase systems and the chemistry of the adrenal cortex. Description of a new carbohy- drate derivative. Biochem. J. 66:1387. Thompson, L. U. and O. Fennema. 1971. Effect of freezing on oxidation of L-ascorbic acid. J. Agr. Food Chem. 16:121. Tillmans, J., P. Hirsch and W. Hirsch. 1932a. Reduction capacity of plant foodstuffs and its relation to vitamin C(I) reducing substances of lemon juice. Z. Untersuch. Lebensm. 66:1. Tillmans, J., P. Hirsch and F. Siebert. 1932b. Reduction capacity of plant foodstuffs and its relation to vitamin C(II) reducing substance of lemon juice as a stabilizer for the true vitamin. Z. Unter- such. Lebensm. 66:21. Tillmans, J., P. Hirsch and J. Jackisch. 1932c. Reduction capacity of plant foodstuffs and its relation to vitamin C(III) quantities of reducing substance in various fruits and vegetables. Z. Untersuch. Lebensm. 66:241. ' Toledo, R., M. P. Steinberg and A. 1. Nelson. 1968. Quantitative determination of bound water by NMR. J. Food Sci. 66:315. Vojnovich, C. and V. F. Pfeifer. 1970. Stability of ascorbic acid in blends with wheat flour, CSM and infant cereals. Cereal Sci. Today 16:317. Waletzko, P. and T. P. Labuza. 1976. Accelerated shelf- 1ife testing of an intermediate moisture food in air and in oxygen-free atmosphere. J. Food Sci. 61:1338. Wanninger, L. A., Jr. 1972. Mathematical model pre- dicts stability of ascorbic acid in food pro- ducts. Food Tech. 66:42. Weissberger, A., J. E. LuValle and D. S. Thomas, Jr. 1943. Oxidation processes. XVI. The autoxi- dation of ascorbic acid. J. Am. Chem. Soc. 66:1934. Weissberger, A. and J. E. LuValle. 1944. Oxidation pro- cesses. XVII. The autoxidation of ascorbic acid in the presence of copper. J. Am. Chem. Soc. 66:700. 133 Zilva, S. S. 1923a. Influence of reaction on the oxi- dation of the antiscorbutic factor in lemon juice. Biochem. J. 11:410. Zilva, S. S. 1923b. Note on the conservation of the potency of concentrated antiscorbutic extracts. Biochem. J. 11:416. Zilva, S. S. 1924a. Note on the conservation of the potency of concentrated antiscorbutic prepara- tions. II. Biochem. J. 16:186. Zilva, S. S. 1924b. Antiscorbutic fraction of lemon juice. Biochem. J. 16:632. Zilva, S. S. 1925. The antiscorbutic fraction of lemon juice. III. Biochem. J. 16:589. Zilva, S. S. 1927. Note on the precipitation of the antiscorbutic factor from lemon juice. Biochem. J. 61:354. Zsigmondy, R. 1911. Uber die struktur des gels der kieselsaure theorie der entwasserung. Z. Anorg. Chem. 11:356. MICHIGAN smTE UNIV. LIBRQRIES llHIWW“WIW”WWWmlWIII‘IHHIWIHII 31293008337416