INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the text directly from the original or copy submitted. Thus, some thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the copy submitted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and improper alignment can adversely afreet reproduction.

In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyright material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, beginning at the upper left-hand comer and continuing from left to right in equal sections with small overlaps. Each original is also photographed in one exposure and is included in reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced xerographically in this copy. Higher quality 6” x 9” black and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to order. UMI A Bell & Howell Infonnation Company 300 North Zed) Road, Ann Arbor MI 48106-1346 USA 313/761-4700 800/521-0600

CAROTENE THERMAL DEGRADATION PRODUCTS AND THEIR EFFECTS ON THE OXIDATIVE STABILITY OF SOYBEAN OIL

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree Doctor of Philosophy in the Graduate

School of The Ohio State University

By

Donald Frank Steenson, M.S.

*****

The Ohio State University 1999

Dissertation Committee: Approved by Professor David B. Min, Advisor

Professor Steven J. Schwartz Advisor Professor Grady W. Chism III Food Science and Nutrition Professor Sheryl A. Barringer Graduate Program ÜMI Number: 9931679

UMI Microform 9931679 Copyright 1999, by UMI Company. All rights reserved.

This microform edition is protected against unauthorized copying under Title 17, United States Code.

UMI 300 North Zeeb Road Ann Arbor, MI 48103 ABSTRACT

Soybean oil used in processed foods is susceptible to lipid oxidation. Carotenes are utilized as colorants in processed foods containing soybean oil. Though carotenes are easily degraded during thermal processing, little is known regarding the effects of thermally degraded carotenes on the oxidative stability of soybean oil.

Thermally degraded (3-carotene or lycopene solutions were added to soybean oil samples at concentrations of 0, 25,45, and 50 ppm, respectively. Each sample, as well as controls containing 5, 25, and 50 ppm o f unheated (3-carotene or lycopene, contained 3 ppm of chlorophyll to allow photosensitized singlet oxygen oxidation to occur. The vial containing each sample was sealed airtight and stored either at 25°C in a light box (1650 lumens) or in a dark oven at 60°C. The oxidative stability of each soybean oil sample was determined by measuring (every 4 hours for 24 hours) peroxide value and headspace oxygen depletion by thermal conductivity gas chromatography. Oxidative stability was further identified by first utilizing solid phase microextraction (SPME) fibers to adsorb volatile oxidation products in the headspace of the samples, then quantifying and comparing the volatiles’ total peak area from their respective gas chromatogram.

Soybean oil samples containing 50 ppm degraded (3-carotene displayed 11.5% higher peroxide values (under light) as well as higher headspace oxygen depletion values (in the dark) when compared with controls (p<0.01). Lycopene degradation products (50 ppm)

in soybean oil decreased peroxide values up to 10.5% under light, and significantly

decreased headspace oxygen depletion of samples in the dark (p<0.05). Over all

concentration ranges, headspace oxygen depletion values for samples stored under light

containing either P-carotene or lycopene degradation products did not differ significantly

from controls. After 30 days of storage in the dark at 60°C, samples containing 50 ppm

degraded p-carotene displayed a significantly higher (p<0.05) SPME-GC total volatile

peak area when compared with controls containing 50 ppm düLl-trans P-carotene. Under

similar conditions, samples containing 50 ppm degraded lycopene displayed a

significantly lower (p<0.05) SPME-GC total volatile peak area when compared with

controls containing only oil.

These results indicate that thermally degraded P-carotene can act as a prooxidant in

soybean oil exposed to elevated temperatures, which may cause a decrease in the

oxidative stability of thermally processed foods containing soybean oil. Thermally

degraded lycopene, however, may act as an antioxidant in soybean oil exposed to

elevated temperatures and therefore may actually increase the oxidative stability o f food systems containing soybean oil.

Ill Dedicated to my wife and family.

IV ACKNOWLEDGMENTS

I would like to thank my advisor. Dr. David B. Min, for his unequaled enthusiasm and dedication in helping me to achieve at the highest level possible in all my scholastic and research endeavors. To my dissertation committee. Dr. Grady W. Chism, Dr. Steven

J. Schwartz, and Dr. Sheryl A. Barringer: I would like to extend my sincere appreciation for the constructive criticism, comments, and suggestions afforded to me throughout my degree program. A special thanks goes out to Cathy Zirkle, A1 McRoberts, Carol Rogers,

Ed Zirkle, and Linda Burianek, members of the department support staff who provided a wide range of invaluable assistance and truly helped get ‘the necessary things’ done. I deeply appreciate the assistance with the GC-MS analyses that I received from Susan

Hatcher. I am of course grateful to all my friends and fellow colleagues in the department for their never-ending support and encouragement throughout the day-to-day nuances o f graduate school life, especially Sherri, Monica, Cindy, Kwok-Man, Steve,

Sarit, Joe, Julie, Minhthy, Mario, Keith, and Sinan. VITA

August 15,1970 Bom — Louisville, Kentucky

1992 B.S. Biology Bucknell University Lewisburg, PA

1994 M.S. Foods and Nutrition The Florida State University Tallahassee, FL

1994 - 1995 ...... Quality Control Chemist Mulberry Ethanol, L.P. Mulberry, FL

1995 - 1996 Laboratory Scientist P. E. LaMoreaux & Associates Lakeland, FL

1996 - 1997 ...... Graduate Research Assistant Dept, of Food Science and Technology The Ohio State University Columbus, OH

1997-1999 USD A Fellow Dept, of Food Science and Technology The Ohio State University Columbus, OH

VI PUBLICATIONS

1. Steenson, D. F., and Sathe, S. K. 1995. Characterization and digestibility of Basmati rice {Oryza sativa L. var. Dehraduni) storage proteins. Cereal Chemistry. 72 (3): 275- 280

2. Min, D. B., and Steenson, D. F. Crude fat analysis. In Food Analysis, 2"^* edition, Nielsen, S. S., ed. Gaithersburg: Aspen Publishers, Inc., 1998, pp. 201-215.

FIELDS OF STUDY

Major Field: Food Science and Technology

V ll TABLE OF CONTENTS

Page

Abstract...... ii

Dedication...... iv

Acknowledgments ...... v

Vita...... vi

List o f Tables...... xiii

List o f Figures ...... xv

Chapter 1 Introduction...... 1

Chapter 2 Literature Review...... 5 2.1 Vegetable oils ...... 5 2.1.1 Production and utilization...... 5 2.1.2 Soybean oil instability ...... 7 2.1.3 Free radical autoxidation...... 8 2.1.3.1 Prooxidants in oils ...... 9 2.1.3.2 Free radical scavengers...... 10 2.1.4 Photosensitized oxidation ...... 11 2.1.4.1 Type 1 and H processes ...... 12 2.1.4.2 Sensitizers in o ils ...... 14 2.2 Volatile soybean oil oxidation products ...... 15 2.2.1 Formation of volatile compounds ...... 17 2.2.2 Isolation o f volatile compounds ...... 19 2.2.2.1 Direct injection...... 19

viii 2.22.2 Static headspace analysis...... 20 2.2.2.S Dynamic headspace analysis ...... 21 2.2.2.4 Solid phase microextraction ...... 22 2.2.3 Correlation with flavor quality ...... 27 2.3 The chemistry of oxygen ...... 28 2.3.1 Nature of triplet and singlet oxygen ...... 28 2.3.1.1 Reaction rates...... 29 2.3.2 Singlet oxygen generation ...... 31 2.3.3 Singlet oxygen oxidation detection ...... 32 2.3.3.1 Headspace oxygen ...... 33 2.3.3.2 Peroxide value ...... 33 2.3.3.3 Gas chromatographic reactor...... 34 2.3.3.4 Photodiode detector ...... 34 2.4 Singlet oxygen-catalyzed oxidation ...... 35 2.4.1 Oxidative products in o il ...... 35 2.4.2 Singlet oxygen initiated biological damage...... 37 2.5 Quenching of singlet oxygen ...... 37 2.5.1 Effects of tocopherols...... 39 2.5.2 Effects of carotenoids ...... 41 2.5.3 Mechanisms and kinetics...... 45 2.6 Carotenoid chemistry ...... 49 2.6.1 Structure...... 50 2.6.1.1 Chromophore ...... 50 2.6.1.2 Carbon skeleton...... 52 2.6.2 Classification...... 53 2.6.2.1 Carotenes...... 53 2.6.2.2 Oxycarotenoids ...... 54 2.6.2.3 Apocarotenoids ...... 54 2.6.3 Stereochemistry ...... 54 2.7 Carotenoid distribution and function...... 58

IX 2.7.1 In nature...... 58 2.7.1.1 Plant contents...... 58 2.7.1.2 Animal contents...... 62 2.7.2 In processed foods ...... 63 2.7.3 As colorants and additives...... 64 2.8 Roles of carotenoids in human nutrition ...... 66 2.8.1 Metabolism ...... 67 2.8.1.1 Absorption ...... 68 2.8.1.2 Tissue distribution ...... 70 2.8.2 Vitamin A precursors ...... 72 2.8.3 Additional health effects ...... 74 2.8.3.1 Antioxidant protection ...... 75 2.8.3.2 Cancer prevention ...... 76 2.8.3.3 Cardiovascular disease...... 78 2.9 Carotenoid oxidation/degradation ...... 78 2.9.1 Effects of oxidizing agents ...... 78 2.9.2 Effects of thermal processes ...... 82 2.9.3 Effects of storage ...... 87 2.9.4 Effects of encapsulation in minimizing degradation ...... 88 2.10 Isolation and identification of carotenoids...... 89 2.10.1 Extraction procedures ...... 89 2.10.1.1 Solvents...... 90 2.10.1.2 Supercritical fluids...... 92 2.10.2 Quantitation...... 93 2.10.2.1 UV/Vis spectroscopy ...... 93 2.10.2.2 HPLC-PDA...... 94 2.10.3 Separation...... 94 2.10.3.1 HPLC...... 95 2.10.3.2 TLC...... 102 2.10.4 Structure elucidation...... 104 2.10.4.1 UV/Vis spectroscopy ...... 104 2.10.4.2 IR spectroscopy ...... 105 2.10.4.3 Mass spectrometry ...... 109 2.10.4.4 NMR spectroscopy ...... 112

Chapter 3 Materials and Methods...... 115 3.1 Materials...... 115 3.2 Lycopene extraction ...... 116 3.3 Thermal treatment of carotenes...... 116 3.3.1 B-carotene and lycopene ...... 116 3.3.2 Crude lycopene extract ...... 117 3.4 Sample preparation and storage for the chlorophyll-photosensitized singlet oxygen oxidation of soybean oil ...... 119 3.5 Sample preparation and storage for the thermally-induced oxidation of soybean o il ...... 123 3.6 Headspace oxygen analysis ...... 123 3.7 Solid phase microextraction (SPME) volatiles analysis ...... 124 3.8 Analysis o f carotene degradation products ...... 126 3.8.1 Separation ...... 126 3.9 Statistical analysis ...... 127

Chapter 4 Results and Discussion...... 128 4.1 Separation and identification of carotenes and their thermal degradation products...... 128 4.2 Effects o f degraded carotenes on chlorophyll-photosensitized singlet oxygen oxidation of soybean o il ...... 134 4.2.1 B-carotene...... 134 4.2.2 Lycopene ...... 142 4.2.3 Crude lycopene extract ...... 151 4.3 Effects o f degraded carotenes on thermally-induced oxidation of soybean oil ...... 158

XI 4.4 Separation and identification of soybean oil volatile oxidation products by SPME ...... 163 4.5 Effects of degraded carotenes on formation of volatile oxidation products in soybean oil ...... 172 Chapter 5 Conclusions...... 184

List of References...... 187

XU LIST OF TABLES

Page

1. Fatty acid composition and analytical constants of soybean oil ...... 6 2. Major volatile compounds responsible for off-flavors in oxidized soybean oil ...... 16 3. The carotenoid content of fresh/frozen vegetables (pg/lOOg ‘wet weight’) ...... 60 4. The contents o f lutein, lycopene, and P-carotene in 10 varieties of tomatoes (pg/lOOg) ...... 61 5. Carotenoid contents of several common -based food products compared with whole tomatoes (pg/lOOg) ...... 64 6. Reported human tissue levels (nmol/g) of P-carotene and lycopene ...... 72 7. Relative vitamin A activity o f some carotenoids commonly foimd in vegetables .....73 8. UV/Vis spectroscopic data for several common carotenoids ...... 106 9. Tukey’s Studentized Range Test for the effects of P-carotene and thermally degraded P-carotene combinations on headspace oxygen depletion after 24 hours of light exposure ...... 135 10. Tukey’s Studentized Range Test for the effects of P-carotene and thermally degraded P-carotene combinations on peroxide formation after 24 hours of light exposure ...... 140 11. Tukey’s Studentized Range Test for the effects of lycopene and thermally degraded lycopene combinations on headspace oxygen depletion after 24 hours of light exposure ...... 145 12. Tukey’s Studentized Range Test for the effects of lycopene and thermally degraded lycopene combinations on peroxide formation after 24 hours of light exposure ...... 149

xui 13. Tukey’s Studentized Range Test for the effects o f crude lycopene extract and thermally degraded crude lycopene extract combinations on headspace oxygen depletion after 24 hours o f light exposure ...... 153 14. Tukey’s Studentized Range Test for the effects o f crude lycopene extract and thermally degraded crude lycopene extract combinations on peroxide formation after 24 hours of light exposure ...... 155 15. Volatile compounds detected in the headspace o f oxidized soybean oil by SPME-GC-MS ...... 167

XIV LIST OF FIGURES

Page

1. United States annual soybean oil production, 1988-1998 ...... 6 2. Excitation and deactivation of photosensitizers ...... 12 3. Sensitized oxidation reaction pathways for type 1 and 11 processes ...... 13 4. Main thermal decomposition products of linolenate hydroperoxides ...... 18 5. Solid phase microextraction device ...... 24 6. Molecular orbitals of triplet oxygen ...... 30 7. Molecular orbitals of singlet oxygen ...... 30 8. Generation of 'Oa by photochemical, chemical, and biological systems ...... 32 9. Singlet oxygen-catalyzed oxidation and degradation o f linoleate ...... 36 10. Scheme for the quenching of singlet oxygen and triplet sensitizer ...... 38 11. Singlet oxygen quenching rates (kq) as a function o f the length of the conjugated polyene system ...... 42 12. Steady state kinetics graph of triplet sensitizer quenching ...... 47 13. Plot of slope/intercept of Figure 10 regression line vs. quencher concentration (triplet sensitizer quenching only) ...... 48 14. Steady state kinetics graph of singlet oxygen quenching ...... 48 15. Plot of slope/intercept of the regression line from Figure 12 vs. quencher concentration (singlet oxygen quenching only) ...... 49 16. Isoprene, the basic unit of the carotenoid molecule...... 50 17. Lycopene, the prototypical carotenoid structure ...... 51 18. Numbering of the positions of the atoms in the carbon skeleton of lycopene 52 19. Structures of common carotenes ...... 53 20. Structures of typical oxygenated carotenoids (xanthophylls) ...... 55

XV 21. Structures of typical apocarotenoids ...... 56 22. Pathways involved in the conversion o f (3-carotene to retinoids ...... 68 23. Pathways and processes involved in the metabolism of carotenoids ...... 69 24. Possible reaction pathway for photosensitized oxidation of lycopene ...... 81 25. Nonvolatile compounds often formed during heating of (3-carotene ...... 83 26. Reaction sequence for the formation o f volatile compounds during heat treatment o f lycopene ...... 86 27. Sample procedure for isolation of carotenoids firom a natural source ...... 91 28. Main features o f a Cig bonded-phase silica material...... 96 29. Schematic diagram of a tjqjical HPLC system ...... 97 30. Typical DR. spectrum for lycopene ...... 107 31. Main fragmentations of the molecular ion of canthaxanthin by FAB MS ...... 110 32. Olefrnic section o f the IH-NMR spectrum of a typical carotenoid measured at 250 MHz (top), 400 MHz (middle), and 600 MHz (bottom)...... 113 33. Increased degradation of (3-carotene (decreased P-carotene peak size) with increased thermal processing at 90°C ...... 118 34. Increased degradation of lycopene (decreased lycopene peak size) with increased thermal processing at 90°C...... 118 35. Schematic of P-carotene or lycopene heating and fractionation into solutions with varying degrees of thermal degradation ...... 119 36. Addition of degraded carotene solutions to soybean oil to create the final concentrations in the tested sample solutions ...... 120 37. Addition of unheated carotene solutions to soybean oil to create the final concentrations in the tested control sample solutions ...... 121 38. Diagram of mirrored light box for sample storage under light ...... 122 39. Extent of thermal degradation of each collected P-carotene sample solution as seen by RP-HPLC ...... 129 40. Extent of thermal degradation of each collected lycopene sample solution as seen by RP-HPLC ...... 130

XVI 41. Extent of thermal degradation of each collected crude lycopene extract sample solution as seen by RP-HPLC ...... 132 42. Effects of thermally degraded (3-carotene on the headspace oxygen depletion of soybean oil containing 3 ppm chlorophyll under light storage at 25°C ...... 134 43. Effects of thermally degraded P-carotene on the peroxide value o f soybean oil containing 3 ppm chlorophyll under light storage at 25°C ...... 139 44. Correlation between headspace oxygen depletion and peroxide value for soybean oil containing degraded p-carotene under light storage at 25°C ...... 44 45. Effects of thermally degraded lycopene on the headspace oxygen depletion of soybean oil containing 3 ppm chlorophyll imder light storage at 25°C ...... 144 46. Effects of thermally degraded lycopene on the peroxide value o f soybean oil containing 3 ppm chlorophyll under light storage at 25°C ...... 148 47. Correlation between headspace oxygen depletion and peroxide value for soybean oil containing degraded lycopene under light storage at 25°C ...... 150 48. Effects of thermally degraded crude lycopene extract on the headspace oxygen depletion of soybean oil containing 3 ppm chlorophyll under light storage at 25°C ...... 152 49. Effects of thermally degraded crude lycopene extract on the peroxide value of soybean oil containing 3 ppm chlorophyll under light storage at 25°C ...... 154 50. Correlation between headspace oxygen depletion and peroxide value for soybean oil containing degraded crude lycopene extract under light storage at 25°C ...... 157 51. Effects of thermally degraded (3-carotene on the headspace oxygen depletion of soybean oil stored in the dark at 60°C ...... 159 52. Effects of thermally degraded lycopene on the headspace oxygen depletion of soybean oil stored in the dark at 60°C ...... 161 53. Effects of thermally degraded crude lycopene extract on the headspace oxygen depletion of soybean oil stored in the dark at 60°C ...... 162 54. Relative sensitivity of four different SPME fibers to hexanal in the headspace o f a soybean oil/acetone model system ...... 165

xvu 55. SPME-GC-MS chromatogram of volatiles in oxidized soybean oil ...... 166 56. SPME-GC mass spectrum of peak 17 (retention time 18.219 min.) (top) and 2-undecenal (bottom)...... 169 57. SPME-GC mass spectrum of peak 18 (retention time 18.681 min.) (top) and cyclododecane (bottom) ...... 170 58. SPME-GC mass spectrum of peak 19 (retention time 18.988 min.) (top) and dodecanal (bottom)...... 171 59. Percentage comparison of total GC peak area from volatiles formed in oil samples containing P-carotene or P-carotene degradation products after storage in a light box at 25°C ...... 173 60. Percentage comparison of total GC peak area from volatiles formed in oil samples containing P-carotene or P-carotene degradation products after storage in an oven at 60°C...... 177 61. Percentage comparison of total GC peak area from volatiles formed in oil samples containing lycopene or lycopene degradation products after storage in a light box at 25°C ...... 179 62. Percentage comparison of total GC peak area from volatiles formed in oil samples containing lycopene or lycopene degradation products after storage in an oven at 60°C...... 181

xvm CHAPTER 1

INTRODUCTION

In recent decades soybean oil production has seen dramatic growth as a result of its

widespread availability and relatively low cost. Today soybean oil accounts for over

75% of total vegetable oil consumption in the U. S., with total production exceeding its

nearest competitor, cottonseed oil, by nearly tenfold (Weiss, 1983). Food processors note

soybean oil’s superior physical properties, which include a wide liquid temperature

range, the ability to be hydrogenated selectively for blending with other oils, and

retention of pourable characteristics after partial hydrogenation (Pryde, 1980). As a

result, soybean is the predominant oil used in products such as mayonnaise, margarines,

prepared salad dressings, and salad or table oils.

A common drawback to the use of soybean oil in processed foods, however, is its

degree of unsaturation and resulting susceptibility to lipid oxidation. A reduction in

flavor quality and nutritional quality, as well as increased health risks such as heart

disease, are often associated with lipid oxidation in foods. Singlet oxygen CO 2) is a

reactive oxygen species that is often a promoter of lipid oxidation in food products

exposed to light (Foote et al., 1982). In these photosensitized oxidation processes, unsaturated fats must be exposed to light in the presence of oxygen and a sensitizer

1 before oxidation of the lipid can occur. Pigments and dyes such as chlorophyll, flavins,

acridine orange, erythrosine, methylene blue, and rose bengai have been found to be

effective photosensitizers (Faria, 1983).

Photosensitized oxidation can proceed by one of two different mechanisms: type I

and type II reactions. In a type I reaction, the excited triplet sensitizer interacts with a

substrate molecule directly, and the radicals that are formed subsequently undergo further

oxidation to form hydroperoxides. In a type II reaction, the triplet sensitizer first interacts

with oxygen to form singlet oxygen (^Oz). The singlet oxygen formed then reacts further

with various substrate molecules, while the triplet sensitizer interacts with other oxygen

molecules (Krinsky, 1979; Foote, 1982). As a result of its unique electron distribution,

singlet oxygen is very electrophilic and therefore extremely reactive with electron-rich

compounds containing double bonds such as linoleic acid (Frankel, 1985). The reaction

rate o f singlet oxygen with linoleic acid is 1,500 times faster than that of ground-state triplet oxygen and linoleic acid (Frankel, 1989).

Carotenoids are a group of lipid-soluble pigments created by the linking of five- carbon isoprene units together to form a polyene skeleton. They are classified into three main groups: the hydrocarbon carotenes, oxygenated xanthophylls, and smaller apocarotenoids (Britton et al., 1995). The carotenes 6-carotene and lycopene both possess polyene systems consisting of eleven conjugated double bonds, and have been shown to be excellent quenchers o f singlet oxygen (Matsushita and Terao, 1980; Packer et al., 1981; Di Mascio et al., 1989; Lee and Min, 1990; Jung and Min, 1991). The deactivation of * 0 2 by carotenoids occurs primarily by way o f physical quenching, a process o f transferring excitation energy firom *0% to the carotenoid resulting in formation of ground state (^0%) oxygen and triplet excited carotenoid. The energy is then dissipated

through rotational and vibrational interactions between the excited carotenoid and the

solvent to recover ground state carotenoid (Di Mascio et al., 1992; Stahl and Sies, 1993).

Carotenoids can therefore be effective at minimizing singlet oxygen-catalyzed lipid

oxidation in food products.

The extended polyene structure of carotenoid pigments also causes them to be very reactive in the presence o f light and/or heat. Because carotenoids such as P-carotene are widely used in industry as food colorants, thermal degradation (and resulting loss of color) during processing is of great concern to food manufacturers and nutritionists.

Although p-carotene and lycopene can be easily degraded by thermal processing (Marty and Berset, 1988; Handelman et al., 1991), very little information has been reported regarding the effectiveness o f thermally degraded carotenes as antioxidants. Therefore, my first objective was to determine on an individual basis the singlet oxygen quenching abilities of P-carotene and lycopene that have been exposed to thermal treatments and then added to a soybean oil model system. Peroxide values and headspace oxygen depletion of the oil samples were measured to determine the rate of oxygen uptake into the oil during light exposure, which can then be indirectly correlated with the rate of singlet oxygen quenching.

The second objective was to determine the ability of degraded carotenes (P-carotene and lycopene) to stop firee radical autooxidation initiated by triplet oxygen. To accomplish this, headspace oxygen depletion of oil samples stored in the dark containing degraded P-carotene or lycopene were measured and then compared with controls. Whether oxidized carotenes exhibit prooxidant or antioxidant properties in soybean oil

was determined by comparison with dl\-trans carotene (untreated) controls as well as controls containing only oil.

The third objective was to identify the extent of oxidation and/or hydrocarbon degradation of the carotenes after each thermal treatment by high-performance liquid chromatography (HPLC) methods, and to correlate the extent of degradation with antioxidant capacity.

The final objective was to compare the relative quantities of volatile lipid oxidation products formed in soybean oil containing degraded P-carotene or lycopene during storage. Solid phase microextraction (SPME) was utilized to collect the volatile compounds, which were then separated and identified using gas chromatography-mass spectrometry. The total peak area of volatile oxidation products present in the headspace of soybean oil samples containing degraded p-carotene or lycopene was compared to that of oil only controls.

This research helped to determine how the addition of carotenoids to food products for coloration and/or antioxidant purposes pre-processing might affect the long-term oxidative stability o f the lipids in the product. CHAPTER 2

LITERATURE REVIEW

2.1 Vegetable oils

2.1.1 Production and utilization

The production of vegetable oil in the United States has traditionally come from a variety of sources, including soybean, cottonseed, com, palm, sunflower, and peanut plants. In recent decades soybean oil production has seen dramatic growth as a result of its widespread availability and relatively low cost (Figure 1). Today soybean oil accounts for over 75% o f total vegetable oil consumption in the U. S., with total production exceeding its nearest competitor, cottonseed oil, by nearly tenfold (Weiss, 1983).

In addition to its availability and low cost, soybean oil possesses other desirable properties that help make it the dominant oil in the U. S. market. It has a high linoleic acid content and a low saturated fatty acid content (Table 1), thus improving its nutritional value relative to the more saturated oils such as coconut or palm oil. Food processors note soybean oil’s superior physical properties, which include a wide liquid temperature range, the ability to be hydrogenated selectively for blending with other oils, and retention o f pourable characteristics after partial hydrogenation (Pryde, 1980).

Due to the many advantages the use of soybean oil provides over other vegetable oils, it 16,500 15,500 c 14,500 o 13,500 12,500 11,500 1988 1990 1992 1994 1996 1998 Year

Figure 1: United States annual soybean oü production, 1988-1998.

% Myristic acid 0.1 Palmitic acid 10.5 Stearic acid 3.2 Oleic acid 2Z3 Linoleic acid 54.5 Linolenic acid 8.3 Arachidonic acid 0.2 Eicosenoic acid 0.9 Iodine value 120 -141 Melting point -23° to -20°

Table 1: Fatty acid composition and analytical constants of soybean oil. is utilized in a wide variety of products in the U. S. After refining, bleaching, and

deodorization steps, a high quality food oil, or refined oil, is ready for incorporation into

various foodstuffs. Soybean oil is the major oil used in mayonnaise, prepared salad

dressings, and salad or table oils. Virtually all mayonnaise, imitation mayonnaise, and

prepared salad dressings produced commercially in the U. S. are prepared exclusively

with soybean oil. Salad and cooking oils often use a blend of partially hydrogenated

soybean oü and other oils such as cottonseed and/or com in their production. Most

margarines consist of partially hydrogenated soybean oil alone, or blended with minor

fractions of other oils such as com or cottonseed. Soybean oil is excellent in the

preparation of shortenings used in the production of bread, cakes, cookies, pie cmsts,

icings, and fillings. Its use has expanded to the point where it can even be found on the

labels of soup, pudding, waffle, and macaroni and cheese mixes, prepared spaghetti

sauces, pizzas, and popcorn (Brekke, 1980).

2.1.2 Soybean oil instability

In addition to improving nutritional quality, the large amount of unsaturated fatty

acids commonly found in soybean oil unfortunately also increases its susceptibility to

oxidation. Oxidative processes can easily occur in soybean oil due to the low energy of

activation necessary to cause double bond migration and hydroperoxide formation on the

individual fatty acids. When double bond migration does occur, essential fatty acids such as linoleic acid are often lost, lowering the oil’s nutritional quality. The high degree of unsaturation in soybean oü has been linked to increased oxidation of solubilized vitamin

A and other compounds present in the oil, as well (Budowski and Bondi, 1960).

7 2.1.3 Free radical autoxidation

Free radical autoxidation is a process that can destroy the unsaturated lipid components (such as soybean oil) in foods, causing the development of undesirable

flavors and odors. In some cases, toxic degradation products are created which can affect the overall safety of the food product (Frankel, 1989). The free radical autoxidation process involves the reaction o f unsaturated oils with oxygen to form hydroperoxides in three basic steps: initiation, propagation, and termination (Burton and Ingold, 1984). The initiation step (1) can take place by metal catalysis, ultraviolet irradiation, and/or thermal or photodecomposition of peroxides/hydroperoxides to form free radicals (R-)-

RH ------^ R- (1)

During the propagation steps (2 & 3) the free radical R- reacts with oxygen to form hydroperoxides (ROOM).

R- + O2 ------ROO- (2)

ROO + RH ------> ROOM (3)

Interactions between ROO- and R- can cause the formation o f nonradical products through the termination step (4, 5, & 6).

R- + R------RR (4)

R 0 0 -+ R------^ ROOR (5)

ROO- -f ROO------^ ROOR -i- O2 (6)

Hydrogen abstraction from an imsaturated fatty acid is selective for the most weakly boimd hydrogen, due to the slow rate of propagation step (3). As a result, the rate of autoxidation thus depends on the number of double bonds present. The relative rates of autoxidation of oleate:linoIeate:linolenate has been determined to be approximately

1:12:25 on the basis of peroxide development (Frankel, 1985).

2.1.3.1 Prooxidants in oils

Compounds known as prooxidants bave been shown to have the ability to quicken the

rate of autoxidation of unsaturated fatty acids. Transition metals having two or more

valence states with a significant oxidation-reduction potential between them have been

found to increase the rate of lipid oxidation by decreasing the length of the induction

period, and increasing the maximum rate o f oxidation. Metals of importance as catalysts

of fatty acid autoxidation include copper, iron, manganese, and nickel (Love, 1985).

Andersson and Lingnert (1998) reported that addition of only 0.07 ppm copper to rapeseed oil resulted in a doubling of the amount of bexanal formed in the headspace after 35 days of storage in air. Addition of 70 ppm copper to similar rapeseed oil samples exposed to air caused the headspace bexanal concentration to increase 70-fold. Jung and

Min (1992) added various concentrations o f oxidized a-, y-, and S-tocopherols to soybean oil, and found that in the dark all three tocopherols acted as prooxidants, with a- tocopherol exhibiting the greatest prooxidant effect. Prooxidant activity increased as oxidized tocopherol concentrations increased in the oil. In another study, thermally oxidized triglycerides at levels of up to 2% were found to exhibit prooxidant effects on the oxidative stabilities of refined, bleached, deodorized, and purified soybean oils (Yoon et al., 1988). Terao (1989) used the intense prooxidant effects of 2,2’-azoèw(2,4- dimethylvaleronitrile) (AMVN) to produce methyl linoleate hydroperoxides for antioxidant testing. Haila et al. (1996) examined the effects o f lycopene, lutein, annatto,

9 and y-tocopherol on autoxidized triglycerides. Surprisingly, results showed that lutein

and lycopene displayed prooxidant behavior when added in the dark to a model system of

oxidized rapeseed oil. Henry et al. (1998) also found that both P-carotene and lycopene

acted as prooxidants when added to safflower seed oil at concentrations >500 ppm.

2.1.3.2 Free radical scavengers

Lipid oxidation can be considered a “chain” reaction, initiated and propagated by the

formation of free radicals. Due to the undesirable results of free radical autoxidation in

lipids, i.e., oxidative rancidity, off-flavor formation, loss of essential fatty acids, etc.,

many free radical “scavengers” have been found that can either delay or eliminate

oxidative processes altogether. Scavengers usually work by removing or deactivating

free radicals before they can propagate the autoxidative process that leads to degradation

of lipids. The antioxidant action of the scavenger can be described by equation (7),

R- -t- AH ------^ RH + A- (7)

where R- is the free radical and AH indicates the antioxidant compound. Terao (1989)

investigated the free radical scavenging activity of the carotenoids P-carotene,

canthaxanthin, astaxanthin, and zeaxanthin on oxidized methyl linoleate by measuring the production of methyl linoleate hydroperoxides. Canthaxanthin and astaxanthin retarded hydroperoxide formation more efficiently than P-carotene and zeaxanthin, indicating the superior peroxyl-radical trapping abilities of the former compounds. Lutein, a common xanthophyll, was cited in another study as having excellent peroxy-radical scavenging ability when added to an oxidized methyl linoleic acid solution (Chopra et al., 1993).

10 Haila et al. (1996) examined the free radical quenching abilities of lycopene, lutein, annatto, and y-tocopherol on autoxidized rapeseed oil triglycerides, and found that annatto and y-tocopherol effectively inhibited hydroperoxide formation. The combination of lutein with y-tocopherol was found to be the most effective at minimizing lipid autoxidation.

2.1.4 Photosensitized oxidation

Most packaged edible oils in supermarkets are exposed to substantial light during the extent of their shelf lives, and are thus subject to photosensitized oxidation as well as autoxidation. The mechanism of photosensitized oxidation as a pathway for the production o f hydroperoxides in oils differs from that of free-radical autoxidation. In photosensitized oxidation, unsaturated fats must be exposed to light in the presence of oxygen and a sensitizer before oxidation of the lipid can occur. Photosensitized oxidation of unsaturated fats begins with the absorption of light by photosensitizers. Two systems of electronically excited states exist for photosensitizers: the singlet (^Sens*) and the triplet (^Sens*) (Foote, 1968; Foote, 1976). The singlet state of a sensitizer arises when upon absorption of visible or near UV light energy, an electron is raised to a higher energy level. The excited singlet sensitizer can then revert back to the ground state by emitting fluorescent light, or it can undergo intersystem crossing (ISC), a process by which the spin of the excited electron is reversed, and a change in the state of the molecule from singlet (*Sens*) to triplet (^Sens*) results (Figure 2). The excited triplet sensitizer slowly decays to the ground state by emitting phosphorescent light. Because

II Excited state Sens* ISC k=l-20xl07sec hv Sens*

k=l-3xl0 /sec Ground state Sens triplet-triplet annihilation

Figure 2; Excitation and deactivation of photosensitizers

the lifetime of ^Sens* is approximately 20,000x the lifetime of ^Sens*, photosensitized oxidations usually proceed by way o f ^Sens*. Therefore, the most effective sensitizers are those which offer the longest triplet state lifetime (Foote, 1968). Pigments and dyes such as chlorophyll, flavins, acridine orange, erythrosine, methylene blue, and rose bengal have been found to be effective photosensitizers (Faria, 1983). Unfortunately, soybean oil has been found to contain lipid oxidation-promoting sensitizers such as chlorophyll and pheophytin at levels as high as 15 ppb and 100 ppb, respectively, even after industrial refining and distillation o f the oil (Usuki et al., 1984; Frankel, 1989).

2.1.4.1 Type 1 and 11 processes

Photosensitized oxidation can proceed by one of two different mechanisms: type 1 and type II reactions. In a type 1 reaction, the excited triplet sensitizer interacts with a substrate molecule directly, usually transferring a hydrogen atom or electron in the

12 process. The radicals that are formed subsequently undergo further oxidation to form hydroperoxides (Figure 3). Type I processes are favored in conditions with high substrate reactivity and concentration, and low oxygen concentration (Foote, 1976).

In a type II reaction, the triplet sensitizer first interacts with oxygen to form an excited electronic state of oxygen known as singlet oxygen ('O 2). The singlet oxygen formed then reacts further with various substrate molecules, while the triplet sensitizer is free to interact with other oxygen molecules (Krinsky, 1979; Foote, 1982). Electron transfer directly to oxygen to form a superoxide anion also occurs in type II reactions, though these reactions make up less than 1% of the total sensitizer-oxygen collisions that occur in type II processes. Both the solubility and concentration o f oxygen present in the system largely decides the rate of type II reactions. Whether substrate or triplet oxygen, respectively, react with the excited triplet sensitizer is the major determinant of whether a type I or type II reaction will occur.

hv ISC Sens -► ^Sens* -► Sens*

Type I Type n + RH

R- + -SensH

+ RH

ROOH 02 •- + Sens •+ ROOH

Figure 3: Sensitized oxidation reaction pathways for type I and II processes.

13 2.1.4.2 Sensitizers in oils

Many compounds are known to act as efficient photosensitizers in vegetable oils by

generating singlet oxygen ('O 2) through transfer of energy in the presence of light and

atmospheric (triplet) oxygen (^Oi). Chlorophyll, which is often naturally present in

vegetable oils (Usuki et al., 1984; Tan et al., 1997), is perhaps the most commonly known

and researched photosensitizer. Fakourelis et al. (1987) determined that chlorophyll in

purified virgin olive oil acted as a photosensitizer for singlet oxygen formation under

light. At a concentration of 3.3 x 10'^ M, chlorophyll b was able to act as a

photosensitizer in the oxidation of soybean oil (Jung and Min, 1991). Kiritsakis and

Dugan (1985) determined that at a concentration of 6 ppm, chlorophyll functioned as a

photosensitizer in the rapid oxidation of bleached olive oil. In addition, chlorophyll b

was found to be twice as efficient as chlorophyll A at sensitizing oxidative reactions. Lee

and Min (1988) found that a dilute solution of soybean oil in methylene chloride

containing 4 ppm chlorophyll was effectively oxidized due to the photosensitizing

properties of the chlorophyll. In a similar study, 4.4 x 10'^ M chlorophyll initiated

photosensitized oxidation in soybean oil containing various carotenoids (Lee and Min,

1990).

The decomposition products of chlorophyll, including pheophytin, pheophorbide,

chlorophyllin, and protoporphyrin, have also been found to exhibit photosensitizer properties (Rahmani and Csallany, 1998). Endo et al. (1984) evaluated the photosensitizer activities of pheophorbide and pheophytin by subjecting safflower oil samples containing each of the chlorophyll derivatives to photooxidation at 0°C. Results showed that pheophytin and pheophorbide exhibited stronger prooxidant activities than

14 either chlorophyll A or B. Pheophytin was also found to be a more effective sensitizer

than chlorophyll A in the photooxidation of olive oil (Kiritsakis and Dugan, 1985) as well

as methyl linoleate (Usuki et al., 1984).

Yamauchi and Matsushita (1977) found the artificial color methylene blue to be an

effective sensitizer in the photooxidation of methyl linoleate. Both methylene blue and

chlorophyll were utilized by Faria and Mukai (1983) to act as sensitizers in the

photooxidation of linoleic acid and safflower oil.

Additional compounds have been shown to exhibit sensitizer properties. Seely and

Meyer (1971) utilized hypericin, a natural photosensitizer with a strong resistance to photooxidative destruction, to examine the products of oxidized (3-carotene. An acetone solution containing 2 x 10'^ M hypericin effectively catalyzed the oxidation o f (3-carotene into several epoxide compounds. Soybean oil itself was reported to act as a sensitizer in the photooxidation of 4,7-undecadiene (Clements et al., 1973).

2.2 Volatile soybean oil oxidation products

The oxidation of soybean oil, whether through autoxidative or photosensitized processes, usually results in the formation of lipid hydroperoxides. These unstable primary oxidation products often degrade to form a variety o f volatile, short-chain secondary oxidation products (Ullrich and Grosch, 1988). Many of these volatile compounds negatively affect the overall quality of soybean oil by producing imdesirable off-flavors that have been described as grassy, painty, or fishy in nature (Table 2). As a result, much effort has been made to both individually identify these volatile compounds and to determine possible mechanisms for how they might form.

15 Relative Threshold Major volatiles % value (ppm)

t, r-2,4-decadienal 33.7 0.10

t, c-2,4-decadienal 17.9 0.02

t, c-2,4-heptadienal 11.1 0.04

2-heptenal 5.6 0.2

t, r-2,4-heptadienal 4.5 0.1

n-hexanal 4.5 0.08

n-pentane 3.1 340

n-butanal 1.5 0.025

2-pentenal 1.2 l.O

l-octen-3-ol 0.9 0.0075

2-pentyl furan 0.8 2.0

n-pentanal 0.7 0.07

2-hexenal 0.7 0.6

n-nonanal 0.7 0.2

n-heptanal 0.6 0.055

l-penten-3-ol 0.5 4.2

2-octenal 0.5 0.15

Table 2: Major volatile compounds responsible for off-flavors in oxidized soybean oil.

16 2.2.1 Formation of volatile compounds

There are several catalysts that help initiate soybean oil oxidation that can result in

the formation of different volatile compounds, including light, increased temperatures,

and the presence of metal ions in the oil. Lee et al. (1995) reported that soybean oil

exposed to a 1937 lux fluorescent light source at room temperature favored the formation

of the volatile compoimds 2-heptenal and l-octen-3-ol when compared with samples

stored in the dark. Light exposure tests were conducted by Robards et al. (1988) on com

chips flried in various vegetable oils and stored in commercial polypropylene film

packaging. Results showed that exposure to light increased the formation of propanal,

pentanal, and hexanal, with propanal exhibiting the most pronounced increase.

Robards et al. (1988) also examined the effects of elevated temperatures on the

formation of volatile compoimds firom vegetable oil in com chips. While the control

chips stored at -20°C showed no changes in their volatiles profile after one year, identical

samples stored at 60°C in the dark showed significantly increased concentrations of

pentanal and hexanal after only 15 days. Snyder et al. (1985) found that polyunsaturated

oils such as soybean and safflower were particularly susceptible to oxidation when stored

at 60°C in the dark. The major volatile compounds apparently formed from linoleic acid

oxidation included pentane, hexanal, and 2-heptanal. The main volatile compounds formed by the thermally induced oxidative degradation of methyl linolenate hydroperoxides were described by Frankel et al. (1991) as 2,4,7-decatrienal, 2,4- heptadienal, and propanal (Figure 4).

Andersson and Lingnert (1998) examined the effects o f copper on the formation o f hexanal in rapeseed oil. The addition of 70 ppm copper to the rapeseed oil sample

17 exposed to air caused the hexanal concentration to increase 70-foId after 35 days of

storage compared to the sample without copper. The addition of only 0.7 ppm copper to rapeseed oil doubled hexanal concentrations.

OOH

2,4,7-decatrienal

OOH

\/ ------\ / i

3-hexenal OOH

2,4-heptadienal

HOO

16

propanal

Figure 4: Main thermal decomposition products of linolenate hydroperoxides.

18 2.2.2 Isolation o f volatile compounds

A major challenge in deterrnining which volatile compounds are responsible for the

off-flavors produced by the oxidation of soybean and other vegetable oils has been in the

attempt to isolate the volatiles from the original sample matrix. Even after isolation,

concentration o f the volatile compounds by adsorption processes is often necessary

before proper qualitative analysis by gas chromatography can occur. As a result, several

different methods have been developed in the attempt to properly identify and quantify

volatile compounds from a variety of sample matrixes.

2.2.2.1 Direct injection

One of the first methods developed for volatile compound isolation was direct

injection of the lipid sample into a gas chromatograph (GC). In this procedure, the glass

injection port liner is packed with glass wool, on top of which the oil sample is added

directly. The glass liner is then lowered into the heated injection port and sealed into the

GC carrier gas system, ready for analysis. Using the direct injection method, Snyder et al. (1988) isolated and identified over 20 different volatile soybean oil oxidation products, with pentane, as well as hepta- and decadienal isomers, being found in the largest quantities. The relatively higher concentration of carbonyl compounds was attributed to possible thermal decomposition of volatile precursors in the injector. Dupuy et al. (1985) reported acetone, pentane, 2,4-decadienal, 2-heptenal, and 2,3-octanedione as the major volatile oxidation products of soybean oil. The use of an external inlet device in the GC was shown to improve the direct injection procedure by decreasing the degradation o f volatile preciursors in the injector port.

19 2.2.2.2 Static headspace analysis

Static headspace (SHS) analysis is the simplest method for isolation of volatile compounds. SHS analyses take advantage o f the high volatility of typical oxidized compounds by relying on the equilibrium distribution of analytes between two coexisting phases (liquid and gaseous). Once the volatile compounds have equilibrated between both phases in a closed system such as a sealed vial, an aliquot of headspace is extracted from the vial with an airtight syringe. The syringe headspace gas is then immediately injected directly into a GC for analysis.

Though SHS analysis is the simplest, most inexpensive method for volatile analysis, it lacks the sensitivity of other methods due to dilution by various headspace gases

(Zhang et al., 1994; Field et al., 1996; Jelen et al., 1998). In addition, SHS sampling is generally only effective for compounds with low boiling points and/or high vapor pressures such as ethanol or ethyl acetate (Snyder et al., 1985; Chin et al., 1996). By focusing his studies only on similar low-molecular weight volatile compounds, Frankel

(1993) was able to effectively investigate by SHS analysis the thermal decomposition of various oxidized oils. Propanal was found to best characterize the oxidative degradation of omega-3 fatty acids, while pentane and hexanal characterized the decomposition of omega-6 fatty acids typical of soybean oil. Robards et al. (1988) used SHS to isolate and quantify the high vapor pressure compounds propanal, pentanal, and hexanal formed during the oxidative deterioration of vegetable oil in com chips. Snyder et al. (1988) revealed that for investigations involving isolation of volatiles from oxidized soybean oil, the SHS method favored a higher relative proportion of acrolein, propanal, and pentanal because of their higher vapor pressures. In a previous study, hexanal and pentanal were

20 among the major volatile compounds isolated from soybean oil stored for eight days at

60°C (Snyder et al., 1985).

2.2.23 Dynamic headspace analysis

The lack of sensitivity often associated with SHS analysis can be overcome by

performing dynamic headspace (DHS) analysis, also commonly known as piurge-and-trap

analysis. During DHS analysis volatile compounds are continuously purged out of the

sample while being concentrated onto a porous polymer adsorbent trap. The next and

final step in the analysis involves thermal desorption of the volatiles immediately prior to

GC separation. A DHS procedure was developed by Lee et al. (1995) for isolating the

volatile compoimds from oxidized soybean oil and trapping them on the adsorbent 2,6-

diphenyl-/7-phenylene oxide (Tenax TA®) that allowed only minimal decomposition of

hydroperoxides (50°C for 30 minutes; 75 mL/min He flow rate). Selke and Frankel

(1987) reported the principle components in the DHS-GC chromatogram profiles of

oxidized soybean oil as pentane, hexenal, 2-heptenal, and 2,4-decadienal from linoleate

hydroperoxides, and 2/3-hexanal and 2,4-heptadienal from linolenate hydroperoxides.

Andersson and Lingnert (1998) utilized DHS analysis to examine the effects of oxygen

and copper concentration on the rate and extent of lipid oxidation in rapeseed oU.

Morales et al. (1997) analyzed the flavor and off-flavor components o f virgin olive oü

using DHS techniques with Tenax TA® as the adsorbent. In this case, thermaUy desorbed volatiles from the Tenax TA® were condensed on a cold trap injector cooled to -110°C before being flash heated and injected into the capillary GC system.

21 Unfortunately, problems with DHS methods have also been discussed in the literature. DHS analysis has been described as expensive, time-consuming, labor intensive, and prone to methodological problems (Zhang et al., 1994; Song et al., 1997).

Park (1993) observed that during the thermal desorption step in DHS analysis in which trapped volatiles are released from the adsorbent and backflushed by hydrogen carrier gas, structural alterations occurred to 2-alkenals and 2,4-aIkadienais. Recommendations included replacing hydrogen with an inert gas such as helium for DHS-GC analyses involving quantification of volatile imsaturated compounds formed during lipid oxidation. Lower-boiling compounds may be lost during the DHS purge cycle while other less volatile compounds such as hexanal, heptadienal, and decadienal are concentrated on the trap, in effect altering the relative ratios of the volatiles observed on

GC chromatograms (Snyder et al., 1988). Park and Goins (1992) encountered frequent problems with volatile (nonanal and decanal) carryover from one sample to the next when utilizing DHS analysis. The requirement that blank water samples be run up to four times between each sample analysis to ensure complete elimination of nonanal and decanal carryover was reported as a serious drawback for the use of DHS analyses in attempts to isolate volatile compounds.

2.2.2.4 Solid phase microextraction

A relatively new technique for the isolation o f volatile compounds from sample matrixes is known as solid phase microextraction (SPME). First described by Arthur and

Pawliszyn (1990), SPME is rapidly gaining popularity as the technique of choice in the isolation and identification of volatile compounds because of its ability to integrate

22 sampling, extraction, concentration, and sample introduction in a single step. Both the

cost and time involved with sample analysis is reduced, while at the same time efficiency

and selectivity are increased (Zhang et al., 1994).

Headspace SPME (HS-SPME) utilizes a fused silica fiber coated with a polymeric

organic liquid that is connected to stainless steel tubing. The tubing acts as a syringe

needle, increasing the mechanical strength of the fiber assembly and allowing repeated

sampling without physical damage to the fiber (Figure 5). During HS-SPME, the fiber is

withdrawn into the protective syringe needle, which then punctures the septum of a

sealed vial. The fiber is pushed out of the syringe needle by a plunger, exposing the fiber

coating to the sample headspace containing volatile compounds. After a predetermined

exposure time to the headspace, the fiber is withdrawn back into the protective syringe needle, the needle is inserted into the injection port of a GC, and the fiber is pushed out of the needle. Analytes from the fiber coating are then thermally desorbed in the GC injector port and quantitatively analyzed (Zhang and Pawliszyn, 1993; Zhang et al.,

1994).

Originally, most SPME applications involved analysis o f aqueous samples for the existence of possible contaminants. Arthur and Pawliszyn (1990) examined the effectiveness of SPME analysis in isolating the common groundwater contaminants

1,1,1 -trichloroethane, trichloroethene, and perchloroethylene. SPME analyses were conducted to determine whether the soil fumigant methyl isothiocyanate (Gandini and

Riguzzi, 1997) or the pesticide procymidone (Urruty et al., 1997) were present in wine samples.

23 plunger

barrel Z-slot

hub-viewing window

adjustable depth gauge

needle guide septum piercmg needle fiber attachment needle coated SPME fused silica fiber

Figure 5: Solid phase microextraction device.

SPME methods have also been used to isolate flavor volatiles firom a wide variety of sample matrixes (Yang and Peppard, 1994), including orange (Steffen and Pawliszyn,

1996; Jia et al., 1998) and apple (Matich et al., 1996; Song et al., 1997) flavor volatiles.

Aroma profiles of commercial vodkas (Ng et al., 1996) and black and white truffles

(Pelusio et al., 1995) have even been determined by SPME analyses. Bicchi et al. (1997)

24 differentiated various roasted coffees by their chromatographic flavor profiles. SPME-

GC patterns for characteristic volatile compounds were distinctly different among various

Swiss cheese samples in a study conducted by Chin et al. (1996). Clark and Bunch

(1997) performed both qualitative and quantitative analyses of flavor additives to tobacco samples. The affinity of SPME fibers for volatiles crucial to proper beer aroma, including the essential oils in hops (Field et al., 1996) and higher alcohols and esters

(Jelen et al., 1998) have been determined. Pyrazines, important constituents in a wide variety of food flavors, were analyzed by HS-SPME-GC. Results showed SPME to be very effective in the isolation, concentration, and analysis of pyrazines in model systems

(Ibanez and Bernhard, 1996). Pan et al. (1995) reported that SPME-GC could isolate semivolatile fatty acids (Ce-Cto) directly firom aqueous samples. The addition of a derivatizing agent in the SPME fiber further increased the analysis sensitivity.

Specific detection ranges for SPME volatile analyses have been determined firom different sample matrixes. Arthur et al. (1992) reported limit o f detection ranges firom

0.3 to 3 pg/L for the BTEX compounds (benzene, toluene, ethylbenzene, and xylene) using a 56 pm polydimethylsiloxane (PDMS) coating and FID detector. Using SPME-

GC-MS analysis, 31 typical tobacco flavors were detected, identified, and quantitated firom tobacco matrixes in concentrations as low as 10 ng/g to 6 pg/g (Clark and Bunch,

1997). Ethyl esters of Cg to Gig fatty acids present ia commercial vodkas were detected by SPME at pg/L levels (Ng et al., 1996), while Pan et al. (1995) reported detection of Cs to Cio fatty acids at levels lower than 1 ppb. Zhang and Pawliszyn (1993) described the

25 detection limits of HS-SPME at the ppt level when ion trap MS was used in conjunction with GC analysis.

The type of stationary phase present on the silica fiber used for SPME analysis greatly affects the affinity for the adsorption of each specific volatile compound to the fiber. Chin et al. (1996) found that adsorption of cheese volatiles on a polyacrylate- coated fiber was 3 to 20 times greater than that with a polydimethylsiloxane (PDMS) fiber. Several studied indicated that the polar stationary phase polyacrylate was selective for the more polar analytes, while PDMS stationary phases had a higher affinity for the non-polar hydrocarbons (Pan et al., 1995; Steffen and Pawliszyn, 1996; Clark and Bunch,

1997; Jelen et al., 1998).

As was previously mentioned, SPME analysis for the isolation of volatile compounds has many advantages over more traditional techniques. SPME completely eliminates consumption of solvents and corresponding solvent disposal costs, while greatly reducing extraction time (Arthur and Pawliszyn, 1990; Yang and Peppard, 1994; Pan et al., 1995).

Urruty et al. (1997) reported that SPME-GC-MS analysis was slightly more sensitive than enzyme-linked immunosorbent assays (ELISA) for the detection o f procymidone residues in wine.

Unfortunately, SPME analysis also has its drawbacks. The primary difficulties

Arthur and Pawliszyn (1990) encountered were poor precision due to difficulty in achieving exact partition equilibrium, and inconsistent positioning of the SPME fiber in the injector port of the GC. Jelen et al. (1998) revealed similar problems with SPME repeatability as determined by peak area comparisons. Yang and Peppard (1994) found the reproducibility of SPME analyses lacking due to its sensitivity to experimental

26 conditions, which can in turn influence the amount of volatiles adsorbed onto the SPME

fiber. Quantification of higher molecular weight apple volatile compounds by SPME was

hindered by the slow transport o f analytes into the gaseous phase, resulting in longer

equilibrium times (Matich et al., 1996). Pelusio et al. (1995) also reported that HS-SPME

techniques were not suited for quantitative analyses due to the fiber coating’s preferential

adsorption of the more polar, volatile compounds found in truffle aroma.

2.2.3 Correlation with flavor quality

Fresh soybean oil has a pleasant, nutty flavor that is considered acceptable to most

consumers (Kao et al., 1998). When soybean oil undergoes considerable lipid oxidation,

however, some o f the volatile oxidation products that are created drastically alter the

flavor profile of soybean oil to what has been described as a grassy, beany, painty flavor.

Photooxidized oils may develop distinctly different flavors described as sour, metallic, or buttery (Warner et al., 1989). The actual flavors imparted by lipid oxidation in oils are difficult to assess because of the wide variations in both their sensory impact and the actual methods used for their determination (Frankel, 1991). Indeed, many of the methods (SHS, DHS, and SPME) used to isolate and identify volatile oxidation products from oils have been developed with the ultimate goal of accurately correlating the quantitative and qualitative off-flavor profiles with decreases in sensory oil flavor quality scores.

Lee et al. (1995) attempted to correlate the amount of volatile compounds formed with the flavor scores for oxidized soybean oil. Results suggested that the major volatile oxidation products that were measured by GC accounted for less than half of the flavor o f

27 the soybean oil samples. A synergistic interaction of volatiles that gave them a much

stronger flavor than would be expected was offered as an explanation for the data.

Robards et al. (1988) showed a good correlation between oxidized com chip flavor scores

and GC volatile peak data. Using GC methods to determine com chip rancidity was

suggested as a sensitive, relatively precise altemative to traditional sensory evaluation

methods. Wamer et al. (1988) also reported good correlation between GC volatile

analyses and sensory flavor evaluations o f cmde soybean oil stored for various time

intervals.

2.3 The chemistry o f oxygen

Oxygen is the most abundant element on earth, occurring as free 0% gas and

combined in the form of water and oxides of many kinds. Molecular oxygen is unique in

being a paramagnetic molecule, even though it has an even number of electrons. Of the

six electrons in oxygen’s outer shell, two are assigned to the 2S orbital, while the other

four electrons are assigned to the three 2P orbitals.

2.3.1 Nature of triplet and singlet oxygen

Molecular oxygen exists in more than one form due to differences in the arrangement of the electrons in the highest occupied pair of molecular orbitals, the n* orbitals. The outermost pair of electrons (ti*) in the common ground state form of molecular oxygen are located on different orbitals and spin in parallel (Figure 6). Spin multiplicity, which is used to describe electron arrangements in molecular orbitals, can be defined by the

28 simple equation 2S + 1, where S is the total spin quantum number. Because ground state oxygen contains two %* electrons with parallel spins, its spin multiplicity is 3 [2(1) +1].

For this reason, ground state oxygen is also known as ‘triplet’ oxygen (^Oa). Triplet oxygen follows Hund’s Rule, which states that the most stable arrangement of electrons is that with the maximum number of unpaired electrons, all with the same spin direction.

A much more rare type of oxygen known as singlet oxygen (^0%) has its two outermost electrons spinning oppositely one-another, resulting in a spin multiplicity of 1

[2(0) + 1]. The specific electron configuration of singlet oxygen can be seen in Figure 7.

Singlet oxygen exists in two states: the higher energy state, and the lower energy state, *A (Foote, 1968). The ^2) state oxygen contains two electrons with opposite spins in different orbitals, and exists 37.5 kcal/mole higher than ground state triplet oxygen

(Foote, 1968). The state is very unstable and only exists for picoseconds before decaying to the state. The *A oxygen state, however, exists for several microseconds at 22.4 kcal/mole above ground state triplet oxygen. It is responsible for most singlet oxygen reactions involving lipid oxidation, and is therefore usually the specific state generally described as singlet oxygen ('Oz) in the literature.

2.3.1.1 Reaction rates

The presence of a completely empty tc* molecular orbital in the electron configuration of singlet oxygen makes it a very electrophilic compound that actively seeks electrons to fill the vacant orbital. As a result, singlet oxygen (’Oz) is a very reactive species and reacts with electron-rich compounds such as linoleic acid at least

29 Molecular orbital

Atomic. orbital Q Atomic...... orbital.... 00(S) (S)00 2Px 2Py 2Pz ''0''— ^ .p. ^ 2Px 2Py 2Pz \ W /

2S W 2S

Energy IS IS

Figure 6: Molecular orbitals of triplet oxygen.

Molecular orbital

Atomic orbital o Atomic orbital

2Px 2Py 2Pz X; M 2Px 2Py 2Pz

2S W 2S

Energy IS IS

Figure 7: Molecular orbitals of singlet oxygen.

30 1,500 times faster than ground state triplet oxygen CO 2 ) to form hydroperoxides (Frankel,

1989). Because of its relatively long lifetime (several microseconds) in the state, singlet oxygen has time to diffiise through condensed phases, and is often as a result able to initiate lipid oxidation. Singlet oxygen can also initiate lipid oxidation by accepting an electron from a donor molecule, resulting in the formation of a reactive superoxide ion

(O2- ) and substrate radical cation (Saito et al., 1983).

2.3.2 Singlet oxygen generation

Singlet oxygen can be produced by both chemical and biological means (Figure 8).

Several methods have been described in the literature for the generation of singlet molecular oxygen (^0%). Di Mascio and Sies (1989) generated singlet oxygen by two different methods; the thermal decomposition of the water-soluble endoperoxide of 3,3’-

( 1,4-naphthylidene) dipropionate (NDPO 2) to NDP, as well as by a hypochlorite/H202 mixture. Approximately one-half of the oxygen liberated in the NDPO 2 —> NDP reaction was found to be in the singlet state. The use o f the thermodissociable endoperoxide of

3,3 ’ -( 1,4-naphthy lidene) dipropionate (NDPO 2 ) to generate singlet oxygen for use in determining carotenoid (Di Mascio et al., 1989) and tocopherol (Kaiser et al., 1990) quenching ability, respectively, has also been performed. Foote (1968) has also reported the high-yield production of singlet oxygen by the chemiluminescent reaction between

NaOCl and H2O2 .

Carlsson et al. (1976) determined that the exposure of oxygen at low pressure to a microwave discharge in the presence of mercury vapor provided a very reliable route for the generation of singlet oxygen (^ 0 2 ). In addition, it is noteworthy that singlet oxygen

31 RCOO- + RCOO • Enzymes RC=0 O2 + Sens*

Endoperoxides p^ductX ^^^^ H2 O2 + OCl

Products + C1-

Ozonides ------^ H 2 O 2 + O 2 OH- + OH

OH- + O2 -

H2 O2 + HO2-

O2- + O2- 0 2 - + Y+

Figure 8: Generation of *02 by photochemical, chemical, and biological systems.

can be generated not only by standard photoexcitation methods, but also via chemiexcitation or “photochemistry in the dark” (Cilento and Adam, 1988).

2.3.3 Singlet oxygen oxidation detection

Researchers have developed methods to identify and quantify the creation of singlet oxygen (* 0 2 ) and its oxidative effects in a model system. Headspace oxygen, peroxide value, and gas chromatographic reactor methods indirectly measure singlet oxygen production by the determination of oxygen uptake into a model system, while germanium photodiode detectors have been used to quantitatively measure * 0 2 emission.

32 2.3.3.1 Headspace oxygen

By exposing lipid samples containing a photosensitizer to light in an airtight sealed

serum bottle, rates o f singlet oxygen oxidation in the oil can be determined by measuring

the headspace oxygen content in the vial. Oxygen disappearance in the headspace of the

sample vial can then be measured directly by a gas chromatograph equipped with a

thermal conductivity detector. Fakourelis et al. (1987) measured the rate of singlet

oxygen oxidation in virgin olive oil samples that contained the sensitizer chlorophyll and

were exposed to a light intensity of 4,000 lux. Rates of singlet oxygen oxidation of

soybean oil exposed to 4 ppm chlorophyll sensitizer at 10°C were determined by the

measurement of headspace oxygen disappearance (Lee and Min, 1988). Lee and Min

(1990) determined the singlet oxygen quenching mechanisms of carotenoids in the

chlorophyll-sensitized photooxidation of soybean oil by measuring headspace oxygen

depletion using a thermal conductivity gas chromatograph.

2.3.3.2 Peroxide value

Singlet oxygen oxidation detection can be performed through the use o f peroxide

value measurements, specifically when testing conditions include the use o f a photosensitizer under lighted conditions. Peroxide value testing determines all substances (peroxides or other similar products of fat oxidation), in milliequivalents of peroxide per 1000 grams of sample, which oxidize potassium iodide under specific conditions (AOCS, 1980). Endo et al. (1984) used peroxide value measurements in conjunction with UV absorbance at 234 nm to determine the ability o f chlorophyll derivatives to initiate photosensitized singlet oxygen oxidation of methyl linoleate. Two

33 other studies involved the measurement of peroxide values to determine the effects of

carotenoids in inhibiting the chlorophyll-photosensitized oxidation of soybean oil (Lee

and Min, 1990; Jung and Min, 1991). Fakourelis et al. (1987) correlated peroxide values

with the rate o f singlet oxygen oxidation in virgin olive oil samples that contained the

sensitizer chlorophyll and were stored under lighted conditions. Kiritsakis (1985)

suggested that because the ratio of peroxide value to conjugated dienoic acid content

increased when chlorophyll was present in an olive oil model system, a singlet oxygen

oxidizing effect was present. As a result, singlet oxygen oxidation could be measured by

determining extent of peroxide value formation.

2.3.3.3 Gas chromatographic reactor

A gas chromatographic reactor (OCR) can be used to detect the singlet oxygen (^0%)

oxidizing activities of different light sources on various polyunsaturated lipid sources

(Faria and Mukai, 1983). The main parts o f the OCR include the reactor (oxidation

chamber), the condenser or trap, the thermal conductivity and flame ionization detectors,

and an optional humidifier. By using the OCR Faria and Mukai (1983) determined that

the rate of singlet oxygen oxidation in both safflower oil and pure linoleic acid was

dependent on the wavelength of the light used. The rate of oxidation increased in the

following order: control (dark) < fluorescent < incandescent < UV (black light).

2.3.3.4 Photodiode detector

The potential for singlet oxygen (^0%) oxidation in a system can also be determined quantitatively by the direct measurement of 'Oz generation and emission through the use

34 o f a germanium diode photodetector. Through the use of a liquid nitrogen-cooled

germanium photodiode detector, Di Mascio et al. (1989) was able to quantitate both

singlet oxygen production and quenching by various carotenoids. Di Mascio and Sies

(1989) measured both monomol and dimol light emissions due to the chemiluminescent

transition of singlet oxygen to the triplet ground state. Singlet oxygen detection was

performed by monitoring infrared photoemission at 1270 nm.

2.4 Singlet oxygen catalyzed oxidation

Exposure to light in the presence of oxygen and a photosensitizer can lead to the

formation of singlet oxygen and subsequent oxidation of many compounds. Compounds that are most susceptible to oxidation by singlet oxygen are those that are highly unsaturated in nature, including many vegetable oils.

2.4.1 Oxidative products in oil

When reacting with unsaturated fats, singlet oxygen reacts directly with double bonds by the “ene” reaction, causing oxygen to be inserted at either carbon of the double bond

(Frankel, 1985; Frankel, 1991). The double bond, in turn, is shifted one carbon to the left or the right, resulting in the formation of a hydroperoxide in the trans configuration

(Figure 9). Singlet oxygen oxidation of linoleate produces a mixture of conjugated (9 and 13) and nonconjugated (10 and 12) hydroperoxides, while linolenate also produces both conjugated (9, 12, 13, 16) and nonconjugated (10 and 15) hydroperoxides. Singlet oxygen photooxidation o f the unsaturated diene 4-cfr,7-cfr-undecadiene was shown to yield 4-hydroperoxy-5-/ra«^,7-cfr-imdecadiene and 5-hydroperoxy-3-frar?iS’,7-cfr-

35 H H H H , I I J* ... h v /sensitizer/ 0% y ____

H

+

H breakdown [Hoducts H H OOH H = \ + I I ^ Y >R H H aldehyde

Figure 9: Singlet oxygen-catalyzed oxidation and degradation of linoleate.

undecadiene as initial products (Clements et al., 1973). Upon photosensitized oxidation at 0°C, methyl oleate produced a 50-50% mixture o f 9- and 10-hydroperoxides, linoleate a mixture of 66% conjugated and 34% unconjugated hydroperoxides, and linolenate a mixture of 75% conjugated and 25% unconjugated hydroperoxides (Frankel et al., 1979).

Matsushita and Terao (1980) utilized mass spectrometry to accurately determine the isomeric compositions of monohydroperoxides produced by singlet oxygen oxidation of methyl oleate and methyl linoleate. Equal amounts of the 9- and 10- isomers were found to be present in methyl oleate, while a slight predominance of the 9- and 16- isomers were foimd in the isomeric composition of methyl linoleate.

Hydroperoxides formed through singlet oxygen oxidation in oils can further degrade through metal catalysts and/or thermal treatments to form both volatile and nonvolatile

36 secondary products. Decomposition o f linoleate hydroperoxides in air at 37°C for a week

produced many nonvolatile monomeric compounds, including di- and tri-oxygenated

esters (Frankel et al., 1985). Volatile compounds such as octane, methyl octanoate, 2-

decenal, and 2-undecenal were identified by gas chromatography from the

photosensitized degradation of methyl oleate, while oxidized products of methyl linoleate

included 2-heptenal, methyl 9-oxononanoate, and lO-oxo-8-decenoate (Frankel et al.,

1985).

2.4.2 Singlet oxygen initiated biological damage

In addition to catalyzing the oxidation of the polyunsaturated lipids found in foods such as vegetable oils, singlet oxygen-catalyzed oxidation can also cause damage in biological systems. Cerutti (1985) found convincing evidence that cellular prooxidant states with increased concentrations of singlet oxygen ('O 2) can promote initiated cells to begin neoplastic growth, resulting in chromosomal damage from DNA structure alterations. Di Mascio et al. (1989) determined that singlet oxygen-inflicted damage of plasmid and bacteriophage DNA included loss of biological activity (as measured by transforming capacity in E. coli) and DNA single-strand breakage.

2.5 Quenching o f singlet oxygen

Because o f the oxidative damage singlet oxygen can cause in both food and biological systems, singlet oxygen quenching is a subject of much practical interest. The phrase “quenching of singlet oxygen” includes both “chemical” quenching, in which singlet oxygen (^Oz) reacts with a quencher Q, to give a product, QO 2 (1), and “physical”

37 quenching, in which the interaction of quencher and product leads only to reduction of

to the ground state with no oxygen consumption or product formation ( 2 ).

'O2 + Q > QO2 (1) Kr

‘O2 + Q ------> "O2 + Q (2 ) Icq Figure 10 reveals a general scheme that holds for photooxidation of a substrate (A) which is known to react only with singlet oxygen (^Oz) at a rate constant kr in the presence of a quencher (Q). In this scheme, physical quenching of ^Oz occurs at a rate kq, while chemical quenching occurs at a rate kox-Q- The decay o f singlet Oz in the surrounding solvent occurs at a rate kj. Chemical, or “charge transfer” quenching, involves interactions between the very electrophilic singlet oxygen molecule and an electron donor to give a charge-transfer complex. Because singlet oxygen acts as the electron-accepting component, the most easily oxidized compounds are usually the best quenchers.

Compounds likely to quench by the charge transfer mechanism include tocopherols, amines, phenols, metal complexes, and other electron-rich compounds (Foote, 1979).

Sen ► 'Sen* Sen* > AOz

kq

Sen Sen

Figure 10: Scheme for the quenching of singlet oxygen and triplet sensitizer.

38 A second quenching mechanism is known as physical, or “energy-transfer”

quenching. The mechanism involves formation of triplet quencher and ground state

oxygen (^Oi) through an energy transfer from a triplet sensitizer to triplet oxygen. The

efficiency of the triplet quencher depends on its ability to be very near or below the

energy of singlet O 2 in the state, which is 22 kcal/mole. Carotenoid compounds such

as p-carotene (Foote and Denny, 1968; Foote, 1970; Foote, 1976; Krinsky, 1979;

Krinsky, 1989; Stratton and Liebler, 1997) and lycopene (Foote, 1970; Krinsky, 1979;

Krinsky, 1989; Di Mascio et al., 1992), metal complexes (Foote, 1982), bilirubin (Stocker

et al., 1987), ascorbyl palmitate (Lee et al., 1997), and other compounds with very

extensive conjugated systems have triplet energies below 2 2 kcal/mole, and therefore can

be effective singlet oxygen physical quenchers.

2.5.1 Effects of tocopherols

Vitamin E, or a-tocopherol, can quench singlet oxygen at a high rate in all solvents,

but it is also converted into oxidized products in polar solvents (Foote, 1982; Liebler et

al., 1990). D-a-tocopherol was found to be an effective quencher of singlet oxygen (^0%) molecules at a rate of k = 2.5 x 10* moF^ s'^ in pyridine. The quenching process proved to be mostly “physical” in nature, with a-tocopherol deactivating about 1 2 0 ^0 % molecules before being destroyed (Fahrenholtz et al., 1974). Kaiser et al. (1990) discovered that the relative singlet oxygen physical quenching efficiencies of the tocopherol homologs decreased in the following order: a > p > y > ô-tocopherol.

Chemical reactivity of the tocopherols with ^0% was much lower, accounting for only 0.1-

39 1.5% of physical quenching. Singlet oxygen quenching rates for a-, y-, and 5-tocopherols

in ethanol were estimated by Yamauchi and Matsushita (1977) to be 2.6 x 10* M'^ s'\ 1.8

X 10* M'^ s'% and 1 .0 x 10* M*' s'\ respectively. The most effective quencher among

the three tocopherols was a-tocopherol. Burton and Ingold (1981) also found a-

tocopherol to be the most reactive antioxidant of the four tocopherol homologs a-, P~, y-,

and Ô-. The antioxidant activity of all tocopherols depends on the free hydroxyl group in

position 6 on the chromane ring; substitution of this group with an ester or ether group

eliminates the antioxidant activity (Sies and Stahl, 1995).

The effects o f tocopherols in an unsaturated oil system are not always beneficial,

however. The singlet oxygen quenching effect of a-tocopherol was found in one study to

be offset by the oxidation of the vitamin itself into hydroperoxidic products. The authors

therefore concluded that natural phenols such as a-tocopherol cannot prevent the buildup

of peroxide species in oils exposed to prolonged light (Carlsson et al., 1976). In another

study, the antioxidant ability of tocopherol was found to be only one-third that of various

anthocyanin compounds (Wang et al., 1997). Cilliard et al. (1980) demonstrated that

though tocopherols were effective singlet oxygen quenchers at low concentrations, at

concentrations higher than 5 x 10'^ M, a-tocopherol exhibited a pronounced prooxidant

effect in oxidized linoleic acid. Frankel et al. (1959) also foimd that natural soybean oil

exposed to light had a higher rate o f oxidation than soybean oils that had been ‘stripped’ to reduce their tocopherol content, again indicating the prooxidant effect of tocopherols present at elevated levels in vegetable oil.

40 2.5.2 Effects of carotenoids

The deactivation of ^0% by carotenoids occurs primarily by way of physical

quenching, a process of transferring excitation energy from ^ 0 % to the carotenoid

resulting in formation of ground state (^Oz) oxygen and triplet excited carotenoid (3).

The energy is then dissipated through rotational and vibrational interactions between ^C*

and the solvent to recover ground state carotenoid (4) (Di Mascio et al., 1992; Stahl and

Sies, 1993; Edge et al., 1997).

'O2 + c ------^ ^02 + ^C* (3)

^C* ------> C + thermal energy (4)

The existence of these quenching reactions has been proven in laser photolysis experiments in which triplet anthracene is produced and quenched by oxygen, yielding

^ 0 2 which then sensitizes absorption due to ^C* (Farmilo and Wilkinson, 1973). These tests have established beyond doubt that most quenching of ^ 0 2 by P-carotene (>99.9%) is in fact due to electronic energy transfer, resulting in production of ^C*, dissipation o f thermal energy into surrounding media, and a return of the carotenoid back to its ground state (Sies and Stahl, 1995). Because carotenoids can be “recycled” in this way, one molecule of P-carotene is capable of quenching as many as 250 to 1000 molecules of * 0 2

(Frankel, 1985; Liebler, 1993). As a result, carotenoids such as P-carotene and lycopene are among the most effective natural * 0 2 quenchers known (Di Mascio et al., 1992).

As can be seen in Figure 11, an increasing number of conjugated double bonds are associated with a higher efficiency of quenching ability against singlet oxygen (Foote et al., 1970; Tsuchiya et al., 1992). In addition, it has been suggested that energy transfer from singlet oxygen to carotenoids is exothermic for carotenes with 11 or more

41 l.OE+11

l.OE+10

CO l.OE+09

l.OE+08

l.OE+07

C 0 njugated C= C in P o lyene C hain

Figure 11 ; Singlet oxygen quenching rates (kq) as a function of the length of the conjugated polyene system.

conjugated double bonds and endothermie for those with fewer than 11 conjugated double bonds (Foote, 1979). To determine relative quenching ability, quenching rate constants for several different carotenoids and xanthophylls have been identified. Lee and Min (1990) determined that the total singlet oxygen quenching rate of five carotenoids increased as their conjugated double bond count increased. Lutein, zeaxanthin, lycopene, isozeaxanthin, and astaxanthin, which contain 1 0 , 1 1 , 1 1 , 1 1 , and

13 conjugated double bonds, respectively, had increasing ^0% quenching rates of 5.72 x

10^, 6.79 X 10^, 6.93 x 10^, 7.39 x 10^, and 9.79 x 10^ M’^ sec"\ respectively. Di Mascio et al. (1989) found that lycopene, with 11 conjugated double bonds, was the most efficient biological carotenoid singlet oxygen quencher, with a quenching rate constant of

42 31 X 10^ M'* sec'\ over twice that of P-carotene (14 x 10^ M'* sec'*). Matsushita and

Terao (1980) revealed a quenching rate constant for p-carotene that was higher than that

reported for lycopene, at 1.5 x 10*° M'* sec'*, while Packer et al. (1981) reported the P-

carotene rate at a much lower 1.5 x 10^ MT* sec'*.

Variation in the functional groups of carotenoids such as lycopene and P-carotene

may contribute to differences in their quenching rates, with the opening of the P-ionone

ring to an open chain in lycopene increasing its quenching ability over p-carotene (Di

Mascio et al., 1991). Foote et al. (1970) suggested that significant cis-trans isomerization

of the carotenoids may accompany singlet oxygen quenching. Though different

structurally, both a- and P-carotene have been determined to act as effective singlet

oxygen quenchers (Kiritsakis and Dugan, 1985).

Foote and Denny (1968) first reported that singlet oxygen could be effectively

quenched by low concentrations of P-carotene: they found that 95% of the

photooxidation of 0.1 M 2-methyl-2-pentene was inhibited by IC^ M P-carotene.

Kellogg and Fridovich (1975) found that at concentrations as low as 10'^ M P-carotene

still inhibited the peroxidation of linolenate in a model system.

Another reaction that has been used to measure the quenching activity of carotenoids

is the chlorophyll-sensitized photooxidation o f soybean oil, monitored as oxygen consumption. Using this method, the *0% quenching rate constants of P-apo- 8 ’-carotenal,

p-carotene, and canthaxanthin were found to be 3.06 x 10° M'* sec'*, 4.60 x 10° MT* sec'*, and 1.12 x 10*° M'* sec'*, respectively (Jung and Min, 1991). Lee and Min (1988) examined the effects of 0, 5, 10, and 20 ppm P-carotene on the oxidation of a soybean

43 oil/methylene chloride model system containing 4 ppm chlorophyll. (3-carotene reduced

the oxidation o f soybean oil at every concentration, and most efiectively at the highest

concentration o f 20 ppm. Warner and Frankel (1987), however, found that at levels ^ 0

ppm, P-carotene contributed to poor flavor and color in soybean oil, while 5 to 10 ppm p-

carotene reduced the photosensitized oxidation of the oil without decreasing oil quality.

The question of whether carotenoids such as P-carotene could, in addition to

quenching singlet oxygen, also act as chain-breaking antioxidants has also been posed.

Though P-carotene does not have the structural features commonly associated with chain-

breaking antioxidants. Burton (1989) claimed that given low oxygen partial pressures, P-

carotene had the potential to act as a lipid-soluble chain-breaking antioxidant. Burton

and Ingold (1984) found that at oxygen pressures of 150 torr or higher, however, P-

carotene and related compoimds could actually act as prooxidants in a methyl linoleate

model system. Heinonen et al. (1997) reported similar P-carotene prooxidant effects in a

10% oil-in-water emulsion, and suggested protecting P-carotene from oxidative

destruction by adding tocopherols to the fat emulsion. Stratton and Liebler (1997) used

an isotope dilution assay to distinguish Type 1 vs. Type II lipid peroxidation reactions in a

lipid bilayer model system. Results suggested that singlet oxygen quenching, rather than

radical scavenging reactions, is responsible for the photoprotective actions o f P-carotene

in a lipid system.

The ability o f synthetic carotenoids to quench singlet molecular oxygen (^0%) has also been determined. Devasagayam et al. (1992) synthesized a Czg-polyene-tetrone that exhibited a quenching rate constant for 'O 2 of kq = 16 x 1 0 ^ sec'\ which was higher

44 than the quenching rate constants for both (3-carotene (5 x 10^ M’’ sec'^) and lycopene (9

X 10^ M** sec’*). The presence of two oxalyl chromophores at each end of the polyene chain was credited with enhancing the *0% quenching ability of the Czg-polyene-tetrone.

2.5.3 Mechanisms and kinetics

In the prevention of singlet oxygen oxidation, an effective quencher can interact with either a triplet sensitizer or singlet oxygen (see Figure 10). In the former case, a reaction between the triplet excited sensitizer (^Sen*) and the quencher (Q) occurs at a rate kq to yield the ground state sensitizer (Sen). In the ground state, the sensitizer can no longer react with triplet oxygen (^O^) to form singlet oxygen.

In the latter case, a quencher can prevent the reaction of singlet oxygen and substrate

(A) two different ways. First and foremost, singlet oxygen (^Oz) and a quencher (Q) interact at rate kq in a physical exchange of energy from oxygen to quencher that results in the formation of grotmd state triplet oxygen (^Oz) and a triplet excited quencher. The second interaction involves a chemical reaction between *0 % and quencher that proceeds at a rate kox-q and results in the formation of an oxidized quencher molecule (QO 2).

Singlet oxygen can also revert to triplet oxygen through decay in the surrounding solvent.

The actual type of quenching occurring (singlet oxygen or triplet sensitizer) can be determined by examining the steady state kinetics of the quenching mechanisms (Foote,

1979). The rate of formation of oxidized substrate (AO 2) due to singlet oxygen oxidation can be expressed as:

drAp2l = K fknlO?!} X ______(krlA]}______(5) dt {kq[(^ + ko[02]} {kr[A] + kq[Q ] + kox.q[Q] + ky}

45 with the reciprocal of equation (5) being equation ( 6 ):

fdFAO^ir' = K-‘ fknr01+k»r0?1! % fkrlAl + kglOl + k ^ ro i + k^\ (6) {dt} {ko[02Î} ■ {k,[A]>

where AO2, A, and Q are the concentrations of the oxidized substrate, the substrate, and

the quencher, respectively, K is the rate of formation of excited triplet sensitizer, ko, kr,

kq, and kox-q are the reaction rate constants of triplet sensitizer with triplet oxygen (^ 0 %),

substrate with singlet oxygen (^ 0 2 ), physical quenching of ^ 0 2 , and chemical quenching

of 'O2, respectively, and kd is the decay rate of ‘O 2 in a specific solvent.

Under specific conditions of constant irradiation time and quencher (Q)

concentration, a linear relationship exists between the reciprocal of the formation rate of the oxidized substrate and the reciprocal of the concentration of the substrate. When

prevention of photosensitized oxidation occurs solely as a result of interactions between the quencher (Q) and excited triplet sensitizer (^Sen*), equation ( 6 ) is simplified to the

following:

fdrAO,!}-' = K-' d + k n r o i} X (l+ k d > (7) {dt} {ko[02]} {kr[Af}

Quenching mechanisms and rates can be observed by graphing the reciprocals of each

level of quencher used. A graph plotting the reciprocals of [AO 2] vs. [A] at various quencher [Q] levels reveals a straight line with a different intercept and slope for each quencher level, which is indicative o f excited triplet sensitizer (^Sen*) quenching (Figure

1 2 ). In this case, slope = K'^ {kd(kq[QJ + ko[0 2 ])}/kr ko[0 2 ] and intercept = K‘^ {kq[Q] + ko[0 2 ])}/ ko[0 2 ], with the latter dependent upon both quencher (Q) and oxygen concentrations. The ratio of the slopes of the plots in Figure 12 to their intercepts vs.

46 [Qi]

[Q2]

1/[A02]

[Q3]

1 /[ A ]

Figure 12: Steady state kinetics graph of triplet sensitizer quenching.

quencher concentration (Q) results in a horizontal line because it is independent of both quencher and oxygen concentrations (Figure 13). Again, a horizontal line plot of this kind is indicative of excited triplet sensitizer (^Sen*) quenching.

When prevention of photosensitized oxidation occurs solely as a result of interactions between the quencher (Q) and singlet oxygen (^Oz), equation ( 6 ) is simplified to the following:

IdjAOall'* = K-' (l+(kg^+ kg)r01 + kd> (7) {dt} { k r [ A ] }

The plots of the reciprocals of [AO 2] vs. [A] at different [Q ] in this case yield plots of straight lines with different slopes = K'* {(k ox-q + k<,)[Q] + k d }/k r and a constant intercept

K*‘ (Figure 1 4 ) . Slope/intercept = (kox-q + k q ) [ Q ] + k j /k r for these plots, which is independent of oxygen concentration. Plotting slope/intercept vs. quencher concentration

47 S/I

[Q]

Figure 13: Plot of slope/intercept of Figure 12 regression line vs. quencher concentration (triplet sensitizer quenching only).

[Ql]

[Q2]

1/[A02] [Q3]

1/[A]

Figure 14: Steady state kinetics graph of singlet oxygen quenching.

48 [Q] results in a line with intercept = kj/kr and a slope = kox-q + kq/kr (Figure 15). Slope

changes as quencher [Q] concentration changes along with a constant intercept value are

indicative of singlet oxygen quenching in the system. In this case the concentration of

quencher [Q] is usually kept low enough to ensure that it acts as a physical quencher of

singlet oxygen, and not a quencher of excited triplet sensitizer.

2.6 Carotenoid chemistry

Carotenoids derive their name from a representative of their group, P-carotene, which

was first isolated firom carrots {Daucus carotd) by Wackenroder in 1831. They are

among the most widespread and important classes of pigments in nature, with over 600

known naturally occurring compounds, each with similar chemical structures.

S/I

[Q]

Figure 15: Plot of slope/intercept of the regression line firom Figure 14 vs. quencher concentration (singlet oxygen quenching only).

49 2.6.1 Structure

Carotenoids are isoprenoid polyenes formed by joining eight C 5 isoprene units

(Figure 16). The isoprene units are linked in a head-to-tail manner except in the center of

the molecule, where a tail-to-tail linkage provides symmetry for the molecule. The

resulting compound contains two methyl groups near the center of the polyene chain that

are separated by six carbon atoms, while the other methyl groups are separated by only

five carbon atoms. This structural arrangement is illustrated in lycopene (Figure 17),

which consists of the prototypical C 40 isoprenoid skeleton firom which all other

carotenoids can be derived by modifications such as cyclization, substitution, elimination,

addition, and rearrangement (Stahl and Sies, 1996).

2.6.1.1 Chromophore

Carotenoids owe their color to the absorption of light by a feature o f their molecular

structure known as the ‘chromophore’. In most carotenoids the chromophore consists

entirely of a series o f conjugated carbon-carbon double bonds, often referred to as the

‘polyene chain’. Though it is possible to have up to 15 conjugated double bonds in the

CH CH HC

Figure 16: Isoprene, the basic unit of the carotenoid molecule

50 Figure 17: Lycopene, the prototypical carotenoid structure.

chromophore of a C 40 carotenoid, structures with 7 to 11 such bonds are far more

common. A chromophore of seven or more double bonds, present in carotenoids such as

p-carotene and lycopene, conveys the ability to absorb light in the visible region so that

colors such as yellow/orange and red are observed, respectively. Likewise, the chromophores of phytoene (three conjugated double bonds) and phytofluene (five conjugated double bonds) are not long enough to reflect light and provide color (Britton etal, 1995).

Other properties of carotenoids are determined specifically by the chromophore, as well. First, each carotenoid is characterized by a specific electronic absorption spectrum based upon absorption of light by its chromophore. As a result, absorption spectroscopy is an important technique in carotenoid analysis. Second, the polyene chain renders the molecule extremely susceptible to oxidative degradation and geometrical isomerization by light, heat, or acids. Finally, the length of a carotenoid’s chromophore, in terms of number o f conjugated double bonds, has been found to be closely linked to its ability to

51 quench, reactive compounds such as singlet oxygen in vitro (Foote et al., 1970; Miller et al., 1996).

2.6.1.2 Carbon skeleton

The carbon skeleton o f many carotenoids consists of a C 40 backbone, although compounds with C45 and even C50 structures have been identified (Weedon and Moss,

1995). It has long been recognized that some carotenoids known as apocarotenoids also have carbon skeletons with fewer than 40 carbon atoms, though most of these have been found to be degradation products of a larger C 40 compoimd.

For convenience, the positions of the atoms in the carbon skeleton of a carotenoid are often numbered in a conventional manner as shown by Figure 18. Usually only key positions on the molecule are numbered in this fashion to focus attention directly to the area o f interest.

16’ 13’ 17) 4 15’

3 20 ’ 2

Figure 18: Numbering o f the positions of the atoms in the carbon skeleton o f lycopene.

52 2.6.2 Classification

Carotenoids can be generally classified into three major groups: carotenes, oxygenated carotenoids (xanthophylls), and the apocarotenoids.

2.6.2.1 Carotenes

The carotenes are comprised solely of carbon and hydrogen (C 40H56), and may or may not include a cyclical ring structure at one or both ends of the hydrocarbon. Common carotenes include (3-carotene, lycopene, and a-carotene (Figure 19).

(3-carotene

lycopene

a-carotene

Figure 19: Structures of common carotenes.

53 2.Ô.2.2 Oxycarotenoids

A second major group of carotenoids are the oxygenated derivatives of the carotenes,

also known as the xanthophylls. The oxygen functions most commonly observed in this

group are hydroxy (monols, diols, and polyols), methoxy, carboxy, oxo, aldehyde, and

epoxy (5,6- and 5,8-epoxides). Xanthophylls containing triple bonds are also known

(Pfander, 1992). When present, hydroxyl substituents are usually found at carbon 3 in

the 8 - or P-ring of the compound. In most cyclic xanthophylls the 5,6 and 5’, 6 ’ double

bonds are very susceptible to epoxidation, while the unconjugated double bond in the e-

ring does not undergo epoxidation. Examples of xanthophylls include canthaxanthin,

zeaxanthin, astaxanthin, and P-cryptoxanthin (Figure 20).

2.6.23 Apocarotenoids

A carbon skeleton containing less than 40 carbon atoms defines the group o f carotenoids known as the apocarotenoids. Most of the molecules in this group occur as the result of degradation of one end o f a C 40 carotenoid. One notable exception is vitamin A (retinol), a C20 compound which is a metabolite of P-carotene but is not considered an apocarotenoid. With the degraded end of the molecule usually comprised of an aldehyde or ketone group, typical apocarotenoids can be seen in Figure 21.

2.6.3 Stereochemistry

The stereochemistry of the carotenoids includes geometrical isomerism about the carbon-carbon double bond, as well as absolute configuration of an asymmetric carbon.

54 canthaxanthin

zeaxanthîn

astaxanthin

P-cryptoxanthin

Figure 20: Structures of typical oxygenated carotenoids (xanthophylls).

55 P-apo-8’-carotenal

P-apo-13-carotenone

retinal

P-apo-14’-carotenal

P-apo-lO’-carotenal

Figure 21: Structures of typical apocarotenoids.

56 Each of the disubstituted, and trisubstituted acyclic double bonds in the carotenoid polyene chain can exist in two forms known as geometrical isomers. The C=C double bond in a carotenoid is designated either cis or trans, with a cis double bond implying a configuration with the highest-priority groups on the same side, and a trans configuration having the groups on opposite sides. More recently, these designations have been largely replaced by Z (zusammen = together) and E (entgegen = opposite) terminology.

Typically, as the number of double bonds in a molecule increases, the number of possible stereoisomers increases accordingly. In carotenoids, however, the number of stereoisomers is restricted because o f steric hindrance from methyl groups along the chain. For example, of the 1056 theoretically possible cis-trans isomers o f lycopene, only 72 are sterically unhindered cis isomers; of 272 possible P-carotene isomers, only 20 are in the unhindered cis form. Carotenoids can often undergo cis-trans isomerization when exposed to high temperatures (Pfander, 1992; O’Neil and Schwartz, 1992) and/or light (Humbeck, 1990). Most carotenoids are found naturally in the dl\-trans (E) form due to its higher stability.

A second major characteristic of carotenoid stereochemistry is their possession of at least one asymmetrically substituted carbon, also known as a chiral center. This results in a molecule that can exist in two distinct stereoisomeric forms that are mirror images of each other. These isomers are known as optical isomers or enantiomers. Optical isomers have identical physical properties with the exception that one rotates polarized light to the right (dextrorotary or ‘d’ isomer), while the other rotates the plane of polarized light an equal amount to the left (levorotary or T’ isomer). Often the absolute configuration of chiral carotenoids is determined by chiroptical methods such as circular dichroism.

57 2.7 Carotenoid distribution and function

2.7.1 In nature

Carotenoids are widely distributed in nature, providing the well-known yellow-orange

color of flowers (sunflower and marigold), the orange-red coloring o f fruits (tomato, orange), and the orange roots of carrots. The greatest production o f carotenoids, however, occurs in the photosynthetic tissues of plants and algae where the abundance of chlorophyll often masks their presence.

2.7.1.1 Plant contents

Total annual natural plant production of carotenoids has been estimated in excess of

100 million tons (Pfander, 1992; Britton et al., 1995). Most of this production is in the form of four major carotenoids: fucoxanthin in marine seaweeds and algae, and lutein, violaxanthin, and neoxanthin in green leaves (Pfander, 1992).

Carotenoids are responsible for the colors of many ftmits, including pineapple, lemons, paprika, and rose hips. Mangos have been foimd to contain from 20 to 125 pg/g total carotenoids (depending on cultivar), with 60 to 76% of the total consisting of 13- carotene (Godoy and Rodriguez-Amaya, 1987). Khachik et al. (1989) found apricots, cantaloupe, and pink grapefruit to be excellent sources o f (3-carotene, with pink grapefimt containing significant quantities (>3000 pg/lOOg) of lycopene.

Hart and Scott (1995) found that among fresh/frozen vegetables, good sources (>1000 pg/lOO g) of lutein were broccoli, butterhead lettuce, parsley, peas, peppers, spinach, and watercress; of lycopene: tomatoes; and of (3-carotene: broccoli, carrots, greens.

58 butterhead lettuce, mixed vegetables, parsley, spinach, and watercress (Table 3).

Khachik et al. (1995) found vegetables from Fiji Island such as amaranthus leaves and

drumstick leaves to contain high levels of both lutein and p-carotene. In an earlier study,

Khachik et al. (1992b) reported the predominant carotenoids in raw green vegetables

(broccoli, spinach, and green beans) to be neoxanthin, violaxanthin, lutein epoxide,

lutein, a-carotene, and p-carotene. In addition, a comprehensive list of carotenoids in

raw tomatoes included lutein, 5,6-dihydroxy-5,6-dihydrolycopene, lycopene 1,2-epoxide,

lycopene 5,6-epoxide, lycopene, neurosporene, y-carotene, P-carotene, phytofluene, and

phytoene. Ben-Aziz et al. (1973) listed a series of oxygenated carotenoids isolated from

tomatoes which included apo- 6 ' -lycopenal, apo- 8 ’-lycopenal, and lycoxanthin. In a

similar study, Britton and Goodwin (1969) isolated a series o f carotenoids from ripe

tomatoes, which included phytoene 1 ,2 -epoxide, phytofluene, (^-carotene, and lycopene.

Carotenoids found in tomato varieties can be seen in Table 4.

The established functions of carotenoids in plants can often be related to their ability to absorb light in the visible spectrum. In photosynthetic tissues, carotenoids usually have two well-defined functions: 1) to act as accessory pigments in photosynthetic processes, and 2 ) to protect ‘photosynthetic apparatus’ against potential damage from visible light such as photosensitized oxidation (Goodwin, 1980). In the former case, carotenoids absorb light at wavelengths different from the chlorophylls and then transfer the absorbed fight energy to the chlorophylls with extreme efficiency that can nearly approach 100% (Siefermann-Harms and Ninnemann, 1982). In the latter case, carotenoids have been shown to exert a photoprotective effect in photosynthetic plant

59 Lutein Zeaxanthîn Lycopene a-carotene (3-carotene

Frozen Green beans 494 70 299

Broccoli 1614 800

Green cabbage 80 ------51

Carrots 283 ----- 3610 10800

Cauliflower Trace

Cucumber 670 ------2 2 2

Greens 3046 ------1663

Leeks 161 69

Iceberg lettuce 1 1 0 74

Butterhead lettuce 1611 1603

Mixed frozen veg. 882 84 ------1045 3670

Parsley 5812 3505

Frozen peas 1633 360

Green pepper 660 235

Frozen sweetcom 522 437 60 45

Spinach 5869 — --- 3397

Spring onions 255 1 1 2

Tomato 78 2937 415

Watercress 10713 4777

Table 3: The carotenoid content of fresh/frozen vegetables (fig/lOOg ‘wet weight’)

60 trans- Total trans-^- Lutein lycopene lycopene carotene

Red varieties Cherry 1 0 1 2686 3780 473 ‘large’ 6 8 1915 2270 349 ‘salad’ 78 2158 2547 509 Flavourtop 48 4958 5653 428 191 1223 1582 1702 Ida FI Hybrid 103 1324 1711 964 Shirley FI 79 2079 2347 771 Craig 149 2948 3907 1093 Moneymaker 59 3475 4255 427 Allicanti 91 3659 4037 525 Beefsteak 89 2729 4833 883

Yellow varieties Sungold 204 390 528 2232 Gold Sunrise 107 2 1 2 1 93

Table 4: The contents of lutein, lycopene, and (3-carotene in 10 varieties of tomatoes (pg/lOOg).

cells. Sistrom et al. (1956) first observed this in comparing a wild purple photosynthetic bacterium with a blue-green mutant strain that lacked carotenoids. When exposed to air and light, the mutant strain stopped growing because its cells and chlorophyll had been killed and destroyed, respectively, while the wild strain continued to grow.

61 2.7.1.2 Animal contents

Though commonly thought of merely as plant pigments, carotenoids are also often

found in microorganisms and animals. Carotenoids are often responsible for yellow,

orange, or red colors in non-phototrophic bacteria, molds, and yeasts. Animals are

sometimes colored by carotenoids from their diet. For example, the feathers of flamingos

are colored pink-red by the presence of ketocarotenoids, while astaxanthin in the skin and

flesh of goldfish and salmon provide these fish with their unique colors. Protein-

astaxanthin complexes provide the bluish-green pigmentation for many marine

invertebrate animals. However, when the animal (e.g. lobster) is subsequently cooked,

the astaxanthin is liberated from the protein, changing the animal’s color to the orange-

red color of the carotenoid (van Breemen, 1996). Even the skin of humans is capable of

taking on a yellowish-orange hue following ingestion of large amounts of carotenoid

supplements and/or carotene-rich foods.

The carotenoid lycopene has been found to be present in human blood samples (0.5

pmo 1/liter plasma) as well as tissues such as adipose, adrenals, and testes(1 nmol/g wet

wt. to 20 nmol/g wet wt.) (Stahl and Sies, 1996). Khachik et al. (1992a) identified

eighteen different carotenoids (as well as vitamin A) from extracts of human plasma,

including several isomers of (3-carotene, lutein, and zeaxanthin. Stahl et al. (1993)

separated five geometrical isomers of P-carotene and seven isomers of lycopene in human serum and tissues. The most prevalent P-carotene isomer in human serum was found to be 13-cw-P-carotene. In testes tissue, however, considerable amounts of 9-cis- and traces of IS-c/s'-P-carotene were also detected. Human liver, adrenal gland, and

62 testes tissues contain significantly higher amounts of carotenoids such as P-carotene and

lycopene than kidney, ovary, and fat tissues. In liver, kidney, adrenal gland, ovary, and

fat tissues p-carotene is the major carotenoid present, whereas lycopene is the

predominant carotenoid in testes tissue (Stahl et al., 1992).

2.7.2 In processed foods

The carotenoid content of foods has been proven to change both quantitatively and

qualitatively after exposure to industrial processing involving heat treatments (Khachik,

1998). Stewed tomatoes along with processed tomato paste were determined to have

lower epoxide concentrations than raw tomatoes (Khachik et al., 1992b). Tan et al.

(1988) also examined the carotenoid content of tomato paste, finding the four dominant

carotenoids to be (in increasing quantities) phytofluene, P-carotene, phytoene, and

lycopene. Several tomato-based food products revealed lycopene as the most abundant

carotenoid, with concentrations ranging firom 0.3 mg/lOOg in vegetable beef soup to 55

mg/lOOg in tomato paste (Table 5). The concentration of P-carotene ranged from 0.23

mg/lOOg in to 1.51 mg/lOOg in vegetable beef soup (Tonucci et al., 1995).

Khachik et al. (1992b) revealed that cooked (by microwaving, steaming, and boiling) green vegetables did not have significantly changed carotenoid profiles from their raw counterparts. However, many of the epoxycarotenoids were destroyed in long-term boiling processes. The carotenoid contents of processed carrots have also been examined, and were observed to contain up to 15% less “total effective carotenes” when compared with raw carrots (Ogunlesi and Lee, 1979).

63 Lutein Lycopene Phytoene y-carotene P-carotene

Tomato soup 90 10920 1720 1950 230

Vegetable beef soup 1 1 0 310 350 1510

Minestrone soup 150 1480 280 920

Vegetarian veg. soup 160 1930 600 1500

Tomato juice 60 10770 1900 1740 270

Vegetable juice 80 9660 1710 830

Catsup 17230 3390 3030 590

Spaghetti sauce 160 15990 2770 3020 440

Tomato paste 340 55450 8360 9980 1270

Tomato puree 90 16670 2400 2940 410

Tomato sauce 17980 2950 3170 450

Whole tomatoes 80 P270 1860 1500 230

Table 5: Carotenoid contents of several common tomato-based food products compared with whole tomatoes (pg/lOOg).

2.7.3 As colorants and additives

Carotenoids are mainly produced for use as colorants in foodstuffs as well as for pigmentation of animal products by administration in the feed. The first synthesis of 13- carotene was reported in 1950 by Inhofifen and coworkers. The Inhoffen synthesis was

64 soon developed into an industrial process that facilitated the first commercial production

of P-carotene in 1954. Since then, improvements in carotenoid synthesis have lead to a

total annual sale value of synthetic carotenoids in 1995 of approximately $300 million,

which is expected to pass the $500 million mark in less than five years. Current

worldwide production capacity of P, P-carotene alone is approaching 500 tons per year

(Britton et al., 1995). Hoffinann-La Roche AG and BASF AG, the two major industrial

producers of carotenoids, today produce six different compounds: P,P-carotene,

canthaxanthin, astaxanthin, P-apo-8 ' -carotenal, P-apo- 8 ' -carotenoic acid ethyl ester, and

citranaxanthin. Recent market prices per kilogram for stable dispersible powders

containing 5-10% active carotenoid were $600 for p-carotene, $900 for the P-apo-8 ’- carotenoids, $1300 for canthaxanthin, and $2500 for astaxanthin (Khachik et al., 1992b).

Commercial synthetic carotenoids are used both as pigments for food (egg yolks, chickens, farm-raised salmon) and for coloration of food products such as margarines and cheeses. Once the carotenoids are transformed into formulations suitable for industrial use, they can be utilized in several different applications. A microcrystalline dispersion o f p-carotene in edible fat is commonly used in the manufacture of margarines. Powders containing different carotenoids microdispersed in a hydrophilic protective colloid are used in aqueous applications such as fruit juices (Britton et al., 1995).

Natural plant pigments have also been used as food colorants. Annatto (bixin, a C 25 diapocarotenoid), paprika extracts (containing casanthin and capsorubin), alfalfa and tagetes extracts (containing xanthophylls such as lutein), tomato extracts (containing lycopene), and carrot extracts (containing a- and p-carotene) are among the natural

65 extracts being used in industry. P-carotene normally produces yellow to orange colors,

while the apocarotenoids and lycopene produce a reddish hue. Carotenoids from

marigold petals can be added to poultry feed supplements, resulting in enhanced color

quality of the poultry meat and eggs (Emenhiser et al., 1996). Kearsley and Rodriguez

(1981) noted that when added as a final action in processing, P-carotene could be used to

color certain foodstuffs such as boiled sweets.

The main drawback to the use of natural carotenoids as colorants in food is their lack

of stability. Nielsen et al. (1996) found canthaxanthin to be a more stable colorant than

P-carotene and therefore more attractive both as a coloring agent and as an antioxidant.

Autoxidation appeared to be the primary mode o f degradation o f both a- and P-carotenes

found in natural encapsulated carrot powders (Wagner and Warthesen, 1995). Chen et al.

(1995) reported that canning o f carrot juice significantly reduced its brightness, with the

overall color changing from a dark orange to a yellowish hue. Regardless, the possible

upcoming government restrictions involving FD&C synthetic colors in food products (as

well as the increasing health-consciousness of consumers) should allow natural

carotenoid colors to play a larger role as food colorants and additives in the future.

2.8 Roles of carotenoids in human nutrition

Carotenoids have long been recognized for their importance as vitamin A precursors, but more recently strong evidence has surfaced that they may be protective against certain types of cancer regardless of provitamin A status. In addition, biological functions of carotenoids such as chemical quenching of free radicals and singlet oxygen

66 have also been demonstrated in the laboratory, proving that carotenoids do indeed serve

multiple roles in human nutrition.

2.8.1 Metabolism

The metabolic fate of carotenoids from both food and supplemental sources has recently received increased attention due to their potential use in the prevention of chronic disease and vitamin A deficiency. The conversion of P-carotene (and other carotenoids that contain an unsubstituted P-ring) into retinol in the intestine is well documented. The primary pathway of all-tram-p-carotene metabolism to retinol (vitamin

A) in the intestine is first through central cleavage of the molecule by p-carotene-15,15’- dioxygenase to form retinal (Figure 22). The retinal is then presumably bound to cellular retinol-binding protein type II (CRBP II) and subsequently converted to retinol by a microsomal reductase (van Vliet, 1996). The 9-cis isomer of p-carotene is also converted to retinal by the enzyme P-carotene-15,15’-dioxygenase, though less efficiently than the

3l\-trans form.

In human intestinal tissue vitamin A formation has also been shown to occur through eccentric cleavage, resulting in P-apo-carotenal byproducts which can be converted through chain-shortening processes to retinal (Figure 22). In the liver carotenoids may also be metabolized to compoimds other than vitamin A, with these retinoid-like metabolites possibly affecting growth regulation and other cellular activities (Rock et al.,

1996). Van Vliet et al. (1996) found that increasing lutein consumption caused reduced retinal formation from P-carotene in the liver, while lycopene had no such effect.

67 eccentric all-frarts-P-carotene cleavage central (S ’) à cleavage (15-15’)

P-apo-8’-carotenal chain shortening retinal retinal reductase OH retinol

Figure 22; Pathways involved in the conversion of p-carotene to retinoids.

2.8.1.1 Absorption

An average intake of 6 mg/day of the five major carotenoids has been observed in adults in the United States (Rock et al., 1996). However, only a fraction of the total intake is utilized by the body due to the relatively low efficiency of carotenoid absorption

(10-30%). With increasing intakes of carotenoids, percent absorption is reduced even further. Dietary fiber such as pectin has also been shown to inhibit P-carotene absorption.

68 Carotenoid availability from food sources depends partially on their release from the physical matrix in which they are ingested. Heating of plant foods before ingestion has been shown to dramatically increase the bioavailability of carotenoid pigments (Gartner et al., 1997), resulting in higher plasma carotenoid concentrations (Rock et al., 1998).

Dissolution o f released carotenoids into a bulk lipid phase is also crucial for efficient intestinal utilization (Figure 23). For this reason, consumption of dietary fat along with carotenoids can significantly increase their absorption into the human body.

Food Bioodstti

LPL

omueoM /lympiy Livor />

cfiylOflUcron / / ^ nutobolHoo , — n-VLtX.1 P'

LFL

VLOL

: X X •vil A mombfrnmw * i

Figure 23: Pathways and processes involved in the metabolism of carotenoids.

69 Once carotenoids have been solubilized in bulk lipid droplets in the stomach or

intestine, bile salts and pancreatic lipases help to capture the carotenoids in micelles.

Duodenal mucosal cells absorb micelles containing the carotenoids by a passive diffusion mechanism similar to that o f cholesterol and triglyceride lipolysis products. The carotenoids are then incorporated into chylomicrons and released into the lymphatics system (Stahl and Sies, 1996).

2.8 .1.2 Tissue distribution

In order to reach the tissues of the body, carotenoids must first be transported by chylomicrons from the intestinal mucosa to the bloodstream via the lymphatic system.

The more non-polar carotenoids such as p-carotene and lycopene reside in the hydrophobic center of the chylomicron, while carotenoids with polar functional groups such as lutein and zeaxanthin are found closer to the outer surface. This differential orientation within the chylomicron molecule may affect the carotenoids’ transfer to lipoproteins during circulation, as well as their uptake by extrahepatic tissues during hydrolysis of chylomicron triglycerides.

In the bloodstream chylomicrons are broken down by lipolytic action due to lipoprotein lipase, which gives rise to chylomicron remnants. The liver clears the remnants from the blood and then can resecrete the carotenoids within lipoproteins for transport to other tissues. The distribution of P-carotene, a-carotene, and lycopene among the very low-density lipoproteins (VLDL), low-density lipoproteins (LDL), and high-density lipoproteins (HDL) are similar for all three carotenoids, with 58-73% in

LDL, 17-26% in HDL, and 10-16% in VLDL. The more polar dihydroxy carotenoids

70 lutein and zeaxanthin, however, are found predominantly in HDL (53%), with

significantly lower proportions in LDL (31%) and VLDL (16%), respectively (Parker,

1996). Van Vliet (1996) actually determined that hepatic resecretion of carotenoids such

as P-carotene occurs first by VLDL, upon which the carotenoids are then transferred to

LDL by delipidation. In short, the hydrocarbon carotenoids are transported primarily in

LDL, whereas the slightly polar carotenoids tend to be transported primarily in the more

hydrophilic HDL lipoprotein fraction (Stahl and Sies, 1996). Significant differences have

also been discovered when comparing distribution of the different isomeric forms of

carotenoids in the human body. Though the amount of sl\-trans P-carotene greatly

exceeds that of cw-P-carotene in the plasma, much increased proportions of cw-P-

carotene has been observed in peripheral tissues.

After transport throughout the body in the lipoproteins, carotenoids are primarily stored in adipose tissue in humans, although they have also been foimd in liver, lung, corpus luteum, and adrenal tissues (Rock et al., 1996). Stahl and Sies (1996) reported that the highest levels of P-carotene and lycopene have been found in liver, adrenal, and testes tissues, with lesser amoimts in lung and kidney tissues (Table 6). In addition, lycopene concentrations exceeded those of P-carotene in every tissue except the ovary.

Prolonged over-consumption of vegetable juices has the capability of increasing carotenoid levels so high that the skin and liver can gain a yellow-orange discoloration

(Stahl and Sies, 1996).

71 Tissue P-carotene lycopene

liver 1.82-4.41 1.28-5.72 kidney 0.31-0.55 0.15-0.62 adrenal 5.6-9.39 1.9-21.60 testes 2.68-4.36 4.34-21.36 ovary 0.45-0.97 0.25 - 0.28 adipose 0.38 0 .2 - 1.3 lung 0.12-0.35 0.22 - 0.57 colon 0.17 0.31 breast 0.71 0.78 skin 0.27 0.42

Table 6 : Reported human tissue levels (nmol/g) of (3-carotene and lycopene.

2.8.2 Vitamin A precursors

Perhaps the most important physiological function o f carotenoids is to act as vitamin

A precursors in animals. Most animal species are capable o f enzymatically converting carotenoids into the vitamin A compound retinol (see Figure 22). The provitamin A activity of a carotenoid depends on the presence of at least one unsubstituted P-ring. As a result, p-carotene, with its two P-rings, is the carotenoid with the highest provitamin A activity. Other carotenoids containing one P-ring and exhibiting (to a lesser extent) provitamin A activity include: a-carotene, y-carotene, 5,6- and 5,8-monepoxides of p- carotene, cryptoxanthin, and the P-apocarotenais (Table 7). Other factors contributing to

72 Carotenoid Activity (%)

all-tranj-P-carotene 1 0 0

9-c/j-P-carotene 38

13 -cw-p-carotene 53

all-/ra«j-a-carotene 53

9 -c/5-a-carotene 13

13-cfj-a-carotene 16 all-/ra«j-cryptoxanthin 57

9-c/j-cryptoxanthin 27

15-c/j-cryptoxanthin 42

P-carotene 5,6-epoxide 2 1

P-carotene 5,8-epoxide (mutachrome) 50

y-carotene 42-50

P-zeacarotene 20-40

Table 7: Relative vitamin A activity of some carotenoids commonly found in vegetables.

provitamin A activity are state o f isomerization {cis vs. tram), gastrointestinal stability, and digestibility (O’Neil and Schwartz, 1992).

‘Retinol equivalent’ was introduced in 1974 by the NAS-Recommended Dietary

Allowances Panel to describe vitamin A content in terms of international unit activity, with 1 lU = 0.3 |ag retinol. Therefore by definition:

retinol equivalent = 1 pg retinol

73 retinol equivalent = 6 qg p-carotene = 12 qg other provitamin A = 3.33 lU vitamin A activity from retinol = 10 lU vitamin A activity from p-carotene

The recommended daily allowance for vitamin A is 1000 retinol equivalents. However, it is difficult to measure total carotenoid content (and therefore vitamin A content) in foods due to the sensitive nature of carotenoids to oxygen and solvents common in present analytical techniques. There are several specific nutritional roles for vitamin A in humans (Semba, 1998). The involvement of retinal, the vitamin A aldehyde, as the chromophore of the visual pigments in the eye, is crucial to human vision. In underdeveloped locations around the world where vitamin A deficiencies frequently occur, xerophthalmia, blindness, and premature death are all too common, especially among children. Vitamin A is also crucial in maintaining growth and reproductive efficiency, as well as maintenance of epithelial tissues and prevention of their keratinization. Recently it has even been shown to improve nonheme iron absorption from fortified rice, wheat, and com flours (Garcia-Casal, 1998).

2.8.3 Additional health effects

In addition to serving as precursors of vitamin A, other important biological functions have been shown to exist for carotenoids, including antioxidant activity (singlet oxygen quenching), and prevention o f diseases such as cancer and cardiovascular disease.

74 2.8.3.1 Antioxidant protection

A biological antioxidant can be defined as “compounds that protect biological systems against the potentially harmful effects of processes or reactions that can cause excessive oxidations” (Palozza and Krinsky, 1992). The antioxidant activities of carotenoids have been demonstrated both in vitro and in vivo in several studies. Stahl et al. (1998) found that lycopene and lutein acted synergistically to increase the antioxidant effects of carotenoid mixtures containing tocopherols and carotene (a and P) in multilamellar liposomes. Martin et al. (1996) reported that carotenoid-loaded cells were partially or completely protected against oxidant-induced changes in lipid peroxidation, demonstrating that P-carotene and lutein (or their metabolites) protect HepG2 human liver cells in vitro against oxidant-induced damage independent of provitamin A activity.

However, Bast et al. (1996) found that P-carotene administered via the diet did not significantly influence liver microsomal lipid peroxidation in rats. P-carotene was shown to act synergistically with a-tocopherol as an effective radical-trapping antioxidant in liver microsome membranes (Palozza and Krinsky, 1992). In a similar study, P-carotene supplementation increased the induction period and decreased PC-OOH production in plasma during AAPH-induced lipid peroxidation. The findings suggested that dietary p-carotene may play an important role in the overall antioxidant defense system of plasma (Meydani et al., 1994).

In vivo studies have shown similar results. Lepage et al. (1996) concluded that p- carotene deficiencies in children with cystic fibrosis led to excessive lipid peroxidation.

Subsequent p-carotene supplementation reduced serum malonaldehyde concentrations,

75 indicating the carotenoid acted as an antioxidant in vivo. Ribaya-Mercado et al. (1995)

found that when human skin was subjected to UV irradiation in vivo, skin lycopene and

P-carotene concentrations were lowered in a manner consistent with the consumption of

free radicals through quenching processes. The researchers concluded that the

antioxidant actions of lycopene and p-carotene might be an important defense mechanism

against the adverse effects of UV irradiation on the skin.

2.8.3.2 Cancer prevention

Studies have consistently sho-wn that individuals with the highest intakes of carotenoid-rich fruits and vegetables have the lowest risks for cancers such as lung, oral cavity, stomach, and esophagus (Krinsky, 1994). Pool-Zobel et al. (1997) determined that men on a diet supplemented with tomato, carrot, or spinach products showed significantly lower endogenous levels of strand breaks in lymphocyte DNA. High serum levels o f carotenoids have also been correlated with decreased risks for certain cancers

(Peto et al., 1981; Bendich, 1993; van Poppel, 1996). P-carotene has been shown to be particularly effective against smoking-related cervical intraepithélial neoplasia and cervical cancer (Charleux, 1996). Garewal (1995) found evidence o f a chemopreventative role for the carotenoid P-carotene against oral cavity cancer, the sixth most frequent cancer in the world. In another study, P-carotene supplemented elderly men had significantly greater natural killer cell activity than elderly men receiving placebos. Because natural killer cells are crucial in the body’s fight against tumor

76 growth, the increased activity can be correlated with decreased cancer risks (Santos et al.,

1996).

Lycopene has shown perhaps the most anti-cancer potential of all the carotenoids.

Giovannuci et al. (1995) foimd that among the carotenoids p-carotene, a-carotene, lutein, lycopene, and P-cryptoxanthin, only lycopene intake was related to a lower risk of non­ stage A1 prostate cancer. O f the four vegetables or Bruits that were found in the study to be significantly associated with lower prostate cancer risk, three (tomato sauce, tomato juice, and pizza) happened to be primary sources of lycopene. Dorgan et al. (1998) determined that the risk of developing breast cancer significantly decreased as serum concentrations of lycopene and lutein/zeaxanthin increased in women. Nagasawa et al.

(1995) examined the effects of chronic ingestion o f lycopene on the development of spontaneous mammary tumors in SHN virgin mice. The lycopene treatment significantly suppressed the mammary tumor development compared with the non-lycopene-fed control group. Sharoni et al. (1995) stated that lycopene inhibited tumor gro wth both in vitro and in vivo. In vitro treatments of lycopene inhibited the growth of human skin fibroblasts, while lycopene treated rats developed fewer and significantly smaller mammary tumors in vivo when compared to the control or P-carotene treated rats.

Lycopene delivered in cell culture medium from stock solutions in tetrahydrofuran more strongly inhibited proliferation of endometrial, mammary, and lung human cancer cells than did either a- or P-carotene (Levy et al., 1995).

77 2.8.3.3 Cardiovascular disease

Carotenoids have been linked to reduced risk for the development of cardiovascular disease. Krinsky (1994) reported on findings showing a distinct inverse association between serum carotene levels and ischemic heart disease. A study of U.S. health professionals showed a similar association between dietary intake of (3-carotene and heart disease risk. Men with the highest dietary (3-carotene intakes had a 29% decrease in heart disease risk. Further analysis, however, determined that heart disease risk was reduced

70% in current smokers with high (3-carotene intakes and 40% in former smokers, with lifelong nonsmokers actually having no significant correlation between (3-carotene intake and risk of heart disease. In a parallel study o f women nurses, those in the highest one- fifth of the population in terms of (3-carotene intake showed a 22% reduction in heart disease risk (Charleux, 1996).

2.9 Carotenoid oxidation/degradation

Due to their conjugated polyene backbone, most carotenoids are fairly unstable molecules and as a result are very sensitive to light, oxygen, and elevated temperatures.

The presence of these factors can cause oxidative degradation of carotenoids, resulting in destruction of the parent compound and formation of a variety of oxidized by-products.

2.9.1 Effects of oxidizing agents

El-Tinay and Chichester (1970) used the radical initiator azo-bis-isobutyronitrile

(AIBN) to accelerate the formation P-carotene oxidation products. AIBN is capable of

78 thermal breakdown to a radical species (R-) that can rapidly react with oxygen to form

peroxyl radicals (ROO-)- The resulting initial oxidized product was reported as P-

carotene-5,6- and 5,8-epoxides, with subsequent decomposition to other products such as

p-carotene-5,6-5’,6’-diepoxide derivatives. Handelman et al. (1991) disputed the former

claim, suggesting rather that the AIBN-initiated radical attack on P-carotene occurred at

multiple sites on the molecule, creating a ‘series’ of apo-carotenals of different sizes such

as retinal, P-apo-14’-carotenal, P-apo-12’-carotenal, and P-apo-10’-carotenal.

Yamauchi et al. (1993) utilized a similar radical initiator, 2,2’-azobis(2,4-

dimethylvaleronitrile) (AMVN), to allow alkylperoxyl radical oxidation of P-carotene.

The major resulting oxidized P-carotene products included 12-formyl-l l-nor-P,p~ carotene, 15’-fbrmy 1-15-nor-P,P-carotene, 5,6-epoxy-5,6-dihydro-P,P-carotene, and 19- oxomethyl-lO-nor-p,P-carotene. Using perphthalate as a chemical oxidizing agent, Seely and Meyer (1971) reported P-carotene-5,6-monoepoxide as the principle product of oxidation. Copper stearate was also effective as a catalyst in the degradation of lycopene to smaller oxidized products (Cole and Kapur, 1957). Khachik et al. (1998) prepared the oxidative metabolites lycopene 1,2-epoxide and lycopene 5,6-epoxide by oxidizing all-

/ra«5 -lycopene with m-chloroperbenzoic acid (MCPBA), followed by acid hydrolysis.

Micro-Cel C, a common chromatographic adsorbent, was found to react with various carotenoids to yield hydroxides and epoxides when exposed in the presence of a nonpolar solvent, a-carotene was converted to 4-hydroxy-a-carotene, while p-apo -8 ’ -carotenal underwent hydroxylation at the allylic 4-position of the P-ring. Lycopene was

79 completely oxidized by the Micro-Cel C, leaving lycopene 5,6-epoxide, 6 ’-apoIycopenal,

and Iycopene-5,6-diol as degradation products (Ritacco et al., 1984; Ritacco et al., 1984).

Photosensitizers can also be effective initiators of carotenoid oxidation. Seely and

Meyer (1971) utilized hypericin, a po^verful photodynamic agent, in the photosensitized

oxidation o f (3-carotene to yield products such as mutatochrome and aurochrome. Lutein

and zeaxanthin were found to undergo the photooxidative process more slowly than 13-

carotene. By irradiating lycopene in an 0% atmosphere with the presence of methylene

blue as a photosensitizer, Ukai et al. (1994) identified several oxidized products of

lycopene, including 2-methyl-2-hepten-6-one and apo- 6 ’-lycopenal (Figure 24).

Holman (1949) first investigated the oxidation of (3-carotene in imsaturated oils,

reporting that in the medium of an oxidizing imsaturated lipid, an intermediate product of

fat oxidation stimulated (3-carotene oxidation. Similar results were found in a later study

that determined the addition of unsaturated oil resulted in a shorter induction period for

the autooxidation of (3-carotene and vitamin A in a paraffin solution. The prooxidant effect of the oil increased with an increasing iodine value and degree of imsaturation

(Budowski and Bond, 1960). Camevale et al. (1979) disputed these previous findings by stating that increased imsaturation o f oil offers protection against autooxidation of carotenoids, because a higher degree o f unsaturation in oil provides a substrate diversion away firom the (3-carotene molecule, resulting in a lower rate of carotenoid oxidation.

Oxygen has been used to initiate carotenoid oxidation. By passing a slow current of pure oxygen through a solution of lycopene in hexane. Cole and Kapur (1957) claimed the oxidative products of lycopene to be acetone, methylheptenone, and laevulinic aldehyde.

80 lycopene

O2, hv methylene blue

2 -methyl- 2 -hepten-6 -one

apo-6 ’-lycopenai

Figure 24: Possible reaction pathway for photosensitized oxidation of lycopene.

Ben-Aziz et al. (1973) later isolated a series of epoxides and apo-lycopenals from tomatoes, including 1,2-epoxy-l,2-dihydro-v|/,\|/-carotene and 5,6-epoxy-5,6-dihydro- v|/,vj/-carotene, which may have been early products of the oxidative degradation of lycopene due to tissue senescence and/or physical injury. Teixeira Neto et al. (1981)

81 examined the kinetics of P-carotene oxidation in a model system of “nonfat” dry milk,

AVTCEL microcrystalline cellulose, and crystalline P-carotene (simulating a dehydrated

food product). Carotenoid oxidation could be accurately predicted by colorimetric

assessment of the decoloration of P-carotene in the model system.

2.9.2 Effects of thermal processes

Carotenoid pigments possess an extended polyene structure that causes them to be

very reactive in the presence of light and/or heat (Minguez-Mosquera and Jaren-Galan,

1995). Because carotenoids such as p-carotene and annatto (bixin) are widely used in

industry as food colorants, thermal degradation (and resulting loss of color) during

processing is of great concern to food manufacturers and nutritionists. Common

degradation products from the heating of p-carotene can be seen in Figure 25.

Marty and Berset (1986) compared the degradation of all-/ra«j-P-carotene during two

thermal processes: heating in sealed glass tubes, and extrusion cooking. After heating

the all-frawi'-p-carotene in sealed glass tubes for 2 hours at 180°C, the main degradation

products were identified as P-carotene-5,6-epoxide, p-carotene-5,6,5’,6’-diepoxide, and

P-carotene-5,8-epoxide. Similar compoimds were found after extrusion cooking of all-

truMf-P-carotene, with P-carotene-5,6,5’,8’-diepoxide also being present. Marty and

Berset (1988) found that after extrusion cooking, only 8 % of the all-rra«j-p-carotene

remained, with the other 92% of the P-carotene degrading into one of six main groups: 1)

mono- or poly-cfj stereoisomers, 2) a diepoxide derivative, 3) five apo-carotenals, 4) a polyene ketone, 5) a dihydroxide derivative, or 6 ) a monohydroxide diepoxide derivative.

82 mutatochrome

5,6,5’,6’-diepoxy-P-carotene

aurochrome

luteochrome

Figure 25: Nonvolatile compounds often formed during heating of p-carotene.

The authors concluded that the resistance of all-trans-P-carotene to high temperatures depends largely on the processing conditions. Different thermal treatments resulted in all-/ra«5-P-carotene losses between 7.5% and 92%. Prolonged heating at 180°C caused

83 only limited breakdown o f the carotenoid molecule, but the presence o f other constituents such as starch and/or water combined with mechanical mixing favoring incorporation of additional oxygen lead to much higher losses of all-/ra«^-p-carotene (Marty and Berset,

1990).

Ouyang et al. (1980) identified the main decomposition products of P-carotene formed during a simulated commercial deodorization o f pahn oil to be P-13-apo- carotenone, P-15-apo-carotenal, and P-14’-apo-carotenal. When P-carotene was heated in glycerol at 210°C for 5 min., 15 min., 1 hour, and 4 hours (to simulate the time/temperature combinations seen in deep fat frying and edible oil deodorization), respectively, over seventy nonvolatile compounds were observed by ER/MS (Onyewu et al., 1986).

A study examining the effect o f microwave cooking on the stability o f carotenoid pigments in sweet potato leaves demonstrated that the epoxy-containing carotenoids were more susceptible to heat loss than other carotenoids. Two lutein dehydration products were identified in the sweet potato leaves after microwave processing: 3,4-didehydro- p,e-caroten-3’-ol, and 3,4-didehydro-P,P-caroten-3-ol (Chen and Chen, 1993). Khachik et al. (1992b) also found the epoxycarotenoids present in foods such as green vegetables and tomatoes to be more sensitive to the thermal processes involved in microwaving, boiling, steaming, and stewing than the hydrocarbon carotenoids such as neurosporene, a - and p-carotene, lycopene, phytofluene, and phytoene. Godoy and Rodriguez-Amaya

(1987) examined the effects of thermal processing on both mango slices and puree. The only significant change in the mango slices after heat treatments was an increase in the

84 measured luteoxanthin content. In the processed mango puree, processing at 80°C for 10 minutes resulted in a 13% decrease in P-carotene content, 33% decrease in violaxanthin, and an increase in the auroxanthin content due to a 5,6- to 5,8-epoxide transformation.

Cole and Kapur (1957) were among the first to examine the extent of lycopene breakdown as a result of exposure to elevated temperatures. The authors reported lycopene losses of 15% and 25% in 3 hours of thermal treatment at 65°C and 100°C, respectively. A heat treatment of 97°C in water produced several novel P-carotene thermal degradation products, including decanal, 4-ethylbenzaldehyde, and cetoisophorcne. The compound 5,6-epoxy-P-ionone was also shown to be an important reaction intermediate, acting as a precursor for various volatile compounds such as P- ionone and 2-hydroxy-2,6,6-trimethylcyclohexanone (Kanasawud and Crouzet, 1990).

Using a similar thermal process, Kanasawud and Crouzet (1990) determined the resulting degradation products of lycopene by GC/MS analysis. The main characterized products included 2 -methyl- 2 -hepten-6 -one and citral, as well as the previously uncharacterized compounds 5-hexen-2-one, hexane-2,5-dione, 6-methyl-3,5-heptadien-2-one, and geranyl acetate (Figure 26). AU-/raMS-lycopene was also found to partially isomerize to the cis- trans isomer as a result of the heat treatments. Henry et al. (1998) reported that in a safflower seed oil model system, the rates of thermal degradation for selected carotenoids between 75°C and 95°C were as follows: lycopene > all-p-aws'-P-carotene = cw-p- carotene > lutein.

85 aIl-/ra«5-P-carotene

Cs-Q C9 -C10 Cleavage Cleavage C7-C8 Cleavage

6-methyl-3,5-heptadien-2-one 2-methyI-2-hepten-6-one

CHO

geranial neral

Figure 26: Reaction sequence for the formation of volatile compounds during heat treatment of lycopene.

Chandler and Schwartz (1988) examined changes in the carotene content of sweet potatoes subjected to one of several different thermal processes. All-frnw-P-carotene was found to be more susceptible to isomerization reactions resulting in the formation of

CIS isomers such as 13- and I5-c/j-P-carotene than degradation reactions during most processing treatments. The extent of isomerization was largely related to the severity and

86 length of the individual heat treatment. Chen and Chen (1995) reported extensive

isomerization in carrot carotenoids during canning (121°C, 30 min.) and HTST heating

(120°C, 30 sec.) thermal processes. The formation of several cis isomers of P-carotene,

including 13-c/5-P-carotene, 13-c/j-lutein, and IS-c/^-a-carotene may have been

responsible for the carrot juice color change from orange to yellow during the intensive

heat treatments. Ogunlesi and Lee (1979) reported a substantial increase in the

concentration of cis isomers and a 25-35% decrease in dl\-trans isomers of P-carotene

after retorting processed carrots, resulting in a 15% decrease in the vitamin A value.

2.9.3 Effects of storage

In addition to changes incurred as a result of processing, carotenoids are also capable

of undergoing alterations in their composition and structure as a result of simple extended

storage. During storage of mango slices in lacquered or plain tin-plate cans, no

significant loss of p-carotene was observed after 10 months of storage. However, upon

extended storage of the slices, a 50% reduction in total p-carotene was observed after 14

months, with continued degradation resulting in a P-carotene loss of 84% after 24

months. Other carotenoids such as violaxanthin and luteoxanthin also decreased in

quantity during storage, while auroxanthin levels remained constant (Godoy and

Rodriguez-Amaya, 1987).

Kopas-Lane and Warthesen (1995) determined that light promoted carotenoid pigment losses in raw spinach, with 60% of the violaxanthin and 2 2 % of the lutein present in the spinach being degraded after only 8 days of storage. Storage of raw

87 spinach in the dark did not affect spinach carotenoid levels, except for an 18% loss of all-

/rawj-p-carotene. For raw carrots, however, neither hghted nor dark cold storage affected

the major carotenoids. In a similar study, degradation of lycopene in a vegetable juice

model system was about one-fifth that of a- and p-carotene after an 8 -day storage period

(Pesek and Warthesen, 1987). In addition, Wagner and Warthesen (1995) revealed that

degradation o f a- and p-carotene during storage at 37°C occurred at the same rate.

2.9.4 Effects o f encapsulation in minimizing degradation

Though the trend in the food industry is towards natural products as opposed to

synthetic additives, carotenoids are limited by application problems. For example,

creating acceptable water-soluble forms of carotenoids is extremely difficult due to the

hydrophobicity of pure carotene crystals. Encapsulation provides a method to transform

liquids such as solubilized carotenes into stable free-flowing powders that can be easily

incorporated into aqueous food systems.

Wagner and Warthesen (1995) determined that hydrolyzed starch of 36.5 DE was more effective than 25, 15, and 4 DE in improving carotene retention during storage, with encapsulated carotenes enjoying a predicted half-life of 450 days at 21 °C, compared with

2 days for the spray-dried carrot juice control. In another study, the stability of p- carotene encapsulated in 25 DE maltodextrin by spray drying, freeze drying, and drum drying was evaluated. After 15 weeks o f storage, drum drying gave the best p-carotene preservation o f all the encapsulation methods. Due to its smaller particle size and surface

88 carotenoid content, the spray-dried encapsulated p-carotene showed the fastest degradation, with 80% degraded after a seven-week storage period at 45°C.

2.10 Isolation and identification of carotenoids

Isolation of carotenoids from biological sources usually involves extraction, saponification, and separation processes. Once separated, identification of individual carotenoids is often based on a complex combination of spectrometric, chromatographic, and chemical tests. High performance liquid chromatography (HPLC), thin layer chromatography (TLC), Mass spectrometry (MS), Gas chromatography (GC), Nuclear magnetic resonance (NMR) spectroscopy, Ultraviolet/Visible (UV/Vis) spectroscopy, and circular dichroism (CD) are all routinely used for separation and/or structural elucidation of carotenoids and their oxidation/degradation products.

The analysis of carotenoids is further complicated due to their structural instability, tendency to stereomutate, photo- and thermolability, and propensity towards oxidation.

As a result, all analytical experiments must be carried out in dim light with inert (under nitrogen or vacuum) atmospheric conditions. In addition, solvents must be purified, environmental temperatures must be no higher than 40°C, and samples must be dried and stored at -20°C under nitrogen (Schiedt and Liaaen-Jensen, 1995).

2.10.1 Extraction procedures

Extraction of carotenoids firom biological materials must be done as rapidly as possible to minimize oxidative and/or enzymatic degradation. Often blanching of plant

89 tissue and addition of calcium carbonate (CaCOs) and antioxidants such as butylated

hydroxytoluene (BHT) is included before extraction to minimize enzymatic reactions,

acid hydrolysis, and oxidative degradation, respectively. To facilitate maximal extraction

yields material should be ground into small pieces. The lipophilic carotenoids require

organic solvents that are free of oxidizing compounds, acids, or halogens for efficient

extraction to take place (Figure 27).

Usually extractions are carried out in a blender so that grinding and extraction can

occur simultaneously. Hart and Scott (1995) prepared samples for carotenoid analysis by

freezing them in liquid nitrogen and then grinding them under liquid nitrogen with a

Waring blender. Homogenates can be suction filtered through a Buchner funnel lined with filter paper coated with Celite (Silveira, Jr., and Evans, 1995). The extraction process can be repeated several times until all visible pigment is extracted from the source material (Khachik et al., 1992b).

2.10.1.1 Solvents

Different solvents have been used with varying success in the extraction of carotenoids from biological sources. Hakala and Heinonen (1994) extracted 6 mg of lycopene from 10 grams o f tomato puree using both petroleum ether and acetone.

Recovery and purity o f the lycopene were both better in the sample extracted with petroleum ether compared to that extracted with acetone, due to a smaller portion of polar xanthophylls being extracted with the more nonpolar petroleum ether solvent.

Tetrahydrofuran (THF) was used as the solvent to extract carotenoids from several different raw and cooked vegetables for HPLC analyses. Complete extraction of the

90 WET BIOLOGICAL DEHYDRATED MATERIAL BIOLOGICAL MATERIAL

CRUDE EXTRACT

Extraction into ether/hexane Evaporation

DRY LIPID EXTRACT Partition hexane/aqueous 85% methanol

T i HYPOPHASIC EPIPHASIC NEUTRAL ACIDIC CAROTENOID CAROTENOID CAROTENOID CAROTENOID Partition

1 1 HYPOPHASIC EPIPHASIC CAROTENOID CAROTENOID

^ Chromatography CCC. TLC, HPLC) ^

INDIVIDUAL CAROTENOIDS Rechromatography Crystallization Recrystallization

PURE CAROTENOID Characterization

Figure 27: Sample procedure for isolation o f carotenoids from a natural source.

91 A wrist-action shaker containing a solvent mix of hexane-acetone-ethanol (50:25:25) was

utilized to extract carotenoids from tomato puree. Fifteen mLs of water was added to the

mix to allow improved separation into distinct polar (clear aqueous) and nonpolar

(lycopene-containing red) layers.

Craft and Soares, Jr. et al. (1992) reported the relative solubility, stability, and

absorptivity o f several carotenoids in eighteen different organic solvents. Results showed

that the solubility of both lutein and P-carotene was highest in tetrahydrofiiran (THF),

hexane exhibited the least solubility for lutein, and both methanol and acetonitrile

exhibited the least solubility for P-carotene. Cyclohexanone caused the most degradation

of carotenoid pigments after ten days o f storage, with only 37% of lutein and 32% of P-

carotene absorbance remaining, respectively.

2.10.1.2 Supercritical fluids

The increasing demand for natural p-carotene has resulted in a growing interest in rapid, cost-effective methods of carotenoid extraction from plant sources. Most extraction methods presently involve the use of organic solvents, which are generally undesirable due to exposure to potentially toxic compounds, as well as environmental concerns. Vega et al. (1996) reported a maximum 99.5% extraction of P-carotene from carrot pulp using supercritical carbon dioxide extraction with 1 0 % ethanol as a co­ solvent. Concentration of ethanol and temperature were determined to be the most important factors in determining extraction yield. High efficiency extractions of P-

92 carotene from natural sources such as carrot pulp were determined to be feasible by

supercritical CO2 + ethanol extraction methods.

2.10.2 Quantitation

Because carotenoids generally obey the Beer-Lambert law, their quantitative determination is often accomplished by spectrometric methods in which the absorbance of a known volume of carotenoid solution is read at the wavelength o f maximal absorption. Carotenoid content can therefore be determined in pg/g material by using the following equation:

pg carotenoid/g = A x V x 10^ A ‘’‘la. X 100 X G where V is the total volume (mL) containing G grams of sample, and A ”‘icm is the specific absorbance or extinction coefBcient. Specifically, the extinction coefficient is defined as the theoretical absorbance o f a 1% solution (w/v) in a 1 cm path-length cuvette. For colored carotenoids, extinction coefficient values are usually around 2500, so a solution with a carotenoid concentration of 1 pg/mL would give an absorbance (A) of approximately 0.25 (Davies, 1976).

2.10.2.1 UV/Vis Spectroscopy

The pursuit of accurate carotenoid quantitation is most commonly accomplished by using spectroscopy in the UV/Visible region of the spectrum. Kearsley and Rodriguez

(1981) utilized UV/Vis Spectroscopy to determine the content and stability of (3-carotene in solution after exposure to thermal treatments, light, and changes in pH.

93 2.10.2.2 HPLC-PDA

High performance liquid chrcmatography-photodiode array detection (HPLC-PDA) has also more recently been used to quantitate carotenoids from various sources. The use of a photodiode array detector to quantitate carotenoids in biological extracts depends upon calibration with authentic source substances, or standards. Analysis precision is greatly increased by the addition of internal standards early in the analysis. Peak height and peak area ratios of the compound o f interest vs. the internal standard can be utilized for quantitation. In the absence of standards only semiquantitative results can be compiled (De Leenheer and Nelis, 1992).

2.10.3 Separation

Chromatography is perhaps the most important single technique in the separation of carotenoid pigments. The separation process of chromatography is based upon two phases known as the stationary phase and the mobile phase. A mixture of compounds is added to the mobile phase that is subsequently carried through the chromatographic system. As the mobile phase passes through the stationary phase, each compound in the mixture reaches an equilibrium distribution at a specific point between the two phases, resulting in differential migration rates through the system (Pfander, 1995).

The compounds to be separated can interact with the two phases in two specific ways, partition and adsorption. Partitioning occurs if the sample mixture diffuses into the interior of a liquid stationary phase. The latter term is applicable if the sample mixture is attracted to the surface of a solid stationary phase. Many adsorbents are utilized in carotenoid chromatography, among them starch, CaCOs, MgCOs, AI 2O3, and silica gels.

94 However, these various absorbents can achieve separation o f compounds in different

ways. Materials such as CaCO] and CaCOs have affinities for double-bond systems, so

separation of carotenoids is determined by the number and type of double bonds in the molecule. Adsorbents such as alumina or silica, on the other hand, separate carotenoid groups of differing polarity such as hydrocarbons, monohydroxy carotenoids, and polyhydroxy carotenoids according to polarity, with the most polar carotenoids being most strongly adsorbed.

Silica and alumina are now the most widely used stationary phases for the separation of carotenoids, and are usually used in bonded-phase chromatography. Bonded phases are prepared by modifying the reactive groups on the surface of the stationary phase

(silica or alumina). To do this, the silanol groups (SiOH) are reacted with an alkylating agent such as dimethyloctadecylchlorosilane (ODS) to form a Cig bonded phase.

Trimethylchlorosilane can also be reacted with the silanol groups in a second alkylation process, known as endcapping, which minimizes the number of remaining silanol groups

(Figure 28).

2.10.3.1 HPLC

High-performance liquid chromatography (HPLC) is by far the method of choice for carotenoid analysis for many reasons. Its ability to distinctly separate many compounds in a relatively short time proves its efficiency. HPLC analyses can be highly sensitive, with the detection of small amounts of impurities, trace carotenoids, and/or geometrical isomers being possible. Finally, HPLC analyses with a photodiode array detector and computer-aided data processor can yield much information about the sample being

95 Si—OH Residual silanol group

CH3

Si1-—0—^i—-(CH2)i7CH3 Bonded ODS group

CH3

CH3

Si1—0 "—^ i—CH3 Endcapped silanol

CH3

Figure 28: Main features of a Cig bonded-phase silica material.

analyzed. High-performance liquid chromatography, which has also been referred to as

‘high-pressure liquid chromatography’, includes liquid chromatographic methods that used stationary phases of particle size not > 1 0 pm and pressures > 2 0 bar (300 psi).

An HPLC system is comprised of several important components that can be viewed schematically in Figure 29. The pump is responsible for producing the pressures necessary to force mobile phase through the stationary phase particles. A sintered metal frit is often placed after the solvent reservoir to prevent solid impurities from reaching the

96 I * _ : (— A 9

7 10 :

r-5-0:>-5SM' — — - __ ; ■III , 0 13

[irwri 14

Figure 29: Schematic diagram o f a typical HPLC system. 1. solvent reservoir, 2. sintered metal frit, 3. high-pressure pump, 4. pulse damper, 5. drain valve, 6 . pressure gauge, 7. pre-column, 8 . injection syringe, 9. injection valve, 10. column, 11. thermostatted oven (optional), 12. detector, 13. recorder/integrator/plotter, 14. fraction collector.

pump and damaging it. A pulse damper can lessen or even remove the pulsations that result from pump action. A pre-column is most commonly inserted between the sample injector and the separation column to protect the latter from impurities in the sample or mobile phase. The separation column itself is usually constructed from 316-grade stainless steel, which is resistant to the high HPLC pressures as well as most chemical corrosion. The detectors utilized in HPLC analyses are usually UV/Vis due to their sensitivity and ease in operation (Johnson and Stevenson, 1978). The development of photodiode-array (PDA) detectors has been very helpful in the analysis of carotenoids due to their rapid data acquisition and ability to store entire spectra for later comparison

(De Leenheer and Nelis, 1992; Pfander and Riesen, 1995). Most recently, electrochemical detection (BCD) has been shown to be a particularly useful alternative to

97 UV/Vis detection methods for LC analyses requiring extremely high (finol) sensitivity

(Ferruzzi et al., 1998).

Both UV/Vis and PDA detectors generate chromatograms that consist of a curved

peak for each separated compound from the sample. The chromatogram provides the

following information to the researcher: 1) the retention time of the compound(s), and 2 )

the area and/or height of the peak(s). Retention times of unknown compounds and

standards can be compared for tentative identification purposes. The area/height of a

peak can be used, with the help of a calibration curve, to estimate the relative amounts of

each compound present in the sample.

Carotenoid separations can be accomplished by either normal-phase or reverse-phase

HPLC. Normal-phase employs adsorptive phases such as silica and alumina, as well as

polar bonded phases such as alkylamine or alkylnitrile in combination with nonpolar

mobile phases. In this type of HPLC, polar sites on the carotenoid compete for

adsorptive sites on the stationary phase, resulting in the least polar carotenoids

(carotenes) eluting first, while the more polar oxygenated carotenoids (xanthophylls) are retained in the column longer. Reverse-phase HPLC includes nonpolar bonded phases

(Cg and Cig) and polymer phases in conjunction with polar mobile phases. During reverse-phase HPLC, xanthophylls are more induced to stay in the polar mobile phase and therefore elute first, while the carotenes partition preferentially into the stationary phase and elute later. Both normal-phase and reverse-phase HPLC can be used with the same mobile phase solvent compositions throughout the analysis (isocratic), or the solvent compositions can change during the analysis (gradient) (Craft, 1992).

98 The investigation of carotenoids by HPLC has been ongoing for almost 30 years, with a wide variety o f papers being published on the subject during that time. Handelman et al. (1991) utilized a mobile phase of 85% acetonitrile/15% methanol through a Cig column in a gradient elution process in order to separate the degradation products of 13- carotene. Ammonium acetate (0.01%) was added to the initial mobile phase to help increase HPLC recovery of carotenoids. Several apocarotenoids (autooxidized products of P-carotene) were recorded on chromatograms at a wavelength of 350 nm. In a similar study, epoxide products of P-carotene antioxidant reactions were separated by reverse- phase HPLC using a mobile phase of methanol-hexane (85:15 v/v) with a flow rate of 1.5 mL/min. Further resolution of epoxide products was successfully accomplished through the use of a cyano-column (Liebler and Kennedy, 1992). Marty and Berset (1990) separated thermally oxidized P-carotene compounds by using a HPLC elution system of n-hexane/diethyl ether (95:5 v/v) with a flow rate of 1 mL/min.

HPLC has been employed as an effective technique to separate and quantify various carotenoids present in human plasma samples. Eighteen different carotenoids, including vitamin A, were separated from extracts of human plasma by HPLC on reversed-phase

Cig silica-based nitrile-bonded columns (Khachik et al., 1992a). Epier et al. (1993) developed a HPLC method for quantitative measurement of the six major carotenoids foimd in human serum. The mobile phase consisted o f a mixture of acetonitrile, methanol, and ethyl acetate, each containing 0.05% triethylamine (TEA) to increase carotenoid recovery. In addition, ammonium acetate (0.05 M) was also added to the methanol to minimize carotenoid losses on the column. Stahl et al. (1993) separated five

99 geometrical isomers of P-carotene and seven of lycopene in human serum and tissues using improved reverse-phase HPLC methods. Mobile phases consisted of methanoI/acetonitrile/2-propanoI (54/44/2), or methanoI/acetonitrile/ 2 -propanoI/H2 0

(10/40/40/10) with a flow rate of 1.0 mL/min and detection at 460 nm.

The carotenoid content of human serum was first examined using a non-aqueous reverse-phase (NARP) HPLC method developed by Nelis and De Leenheer (1983).

NARP mobile phases used included mixtures of acetonitrile, dichloromethane, methanol, tetrahydrofuran (THF), and ethyl acetate. The NARP method showed superior sample solubility o f nonpolar carotenoid components when compared with conventional aqueous reverse-phase chromatography, mainly due to the fact that carotenoids typically are only sparingly soluble in partially aqueous solvents.

Several HPLC methods have been developed to determine the carotenoid contents of firesh or processed fiaiits and vegetables both qualitatively and quantitatively. The major carotenoid constituents of extracts firom several raw and cooked green vegetables

(broccoli, green beans, and spinach) (EChachik et al., 1992b), as well as firesh tomatoes and tomato paste (Khachik et al., 1992b; Tonucci et al., 1995) have been separated by

HPLC on a Cig-reverse phase column. An isocratic mixture o f acetonitrile (85%), methanol (10%), dichloromethane (2.5%) and hexane (2.5%) was used for the mobile phase, with chromatographic analyses simultaneously being monitored at several different wavelengths (Khachik et al., 1992b). Tan (1988) also identified the carotenoids present in tomato paste by using isocratic reverse-phase HPLC with several different

NARP solutions. Carotenoids firom apricot, peach, cantaloupe, and pink grapefiruit extracts were also separated and quantitated on a Cig-reverse phase column (Khachik et

100 al., 1989). Kopas-Lane and Warthesen (1995) developed a reverse-phase gradient HPLC

method on a Cig column for the separation of xanthophylls, carotenes, and cis P-carotene

isomers from raw spinach and carrots. Initial conditions for the mobile phase included

90% acetonitrile/5% water/5% methanol, with a linear gradient increasing the methanol

content to 100% in 15 minutes. An HPLC solvent system of

acetonitrile/methanol/chloroform/hexane (75:12.5:7.5:7.5 v/v/v/v) pumped at a flow rate

of 1.0 mL/min was employed to determine the stability of carotenoids in sweet potato

leaves to microwave cooking. Eluate was monitored at 440 nm and comparing retention times of separated peaks with those o f the reference standards (Chen and Chen, 1993)

identified carotenoids.

More recently novel polymeric C 30 stationary phase columns have been shown to be effective in HPLC separation of carotenoid compounds and even geometric isomers

(Sander et al., 1994). Using such a column, Emenhiser et al. (1995) was able to obtain superior resolution of several isomers of the asymmetrical carotenoids lutein, a-carotene, and P-cryptoxanthin, as well as the symmetrical carotenoids zeaxanthin and P-carotene.

A mobile phase of methyl-terf.-butyl-ether (MTBE) in methanol was used isocratically to achieve separations on the polymeric C 30 stationary phase. The same C 30 column was later used by Emenhiser et al. (1996) to distinguish the different geometrical carotenoid isomers present in human serum, carrots, algae extract, and a poultry feed supplement.

Again a methanol-MTBE mobile phase was utilized for chromatographic analysis.

Addition o f triethylamine (0.1% v/v) to the mobile phase was found to increase the recovery o f p-carotene from the C 30 column from 52% to 88%.

101 Another proposal to increase HPLC resolution between carotenoid isomers was made

by Schmitz et al. (1995). The use of a calcium hydroxide stationary phase in HPLC

analyses was advocated for the specific purpose of increasing the resolution between the

various members o f the acyclic and cyclic geometric carotenes containing 5, 7, or 11

aliphatic double bonds.

Depending on how HPLC methods are used in the separation and identification of

carotenoids, occasionally unanticipated results have been known to occur. Piretti et al.

(1996) found that when using a normal-phase cyano-amino HPLC column, lycopene

samples and standards both exhibited diSering numbers of peaks depending on the

solvent used to prepare the sample for injection. Scott et al. (1992) reported that in

addition to reactions between carotenoids, injection solvents, and the mobile phase, metal

surfaces such as the stainless steel of metal frits in HPLC systems may be damaging to

carotenoid compounds. In addition, changes in ambient temperature during HPLC

analyses of carotenoids were reported to cause dramatic differences in the data collected.

A reduction of 1 minute in elution time for every 1°C rise in temperature was described,

with optimum resolution occurring at 20-22.5°C (Scott and Hart, 1993).

2.10.3.2 TLC

Thin-layer chromatography (TLC) first gained popularity in the 1950’s as a replacement for paper chromatography. Though more recently TLC has tended to be superseded by the more efficient and sensitive HPLC systems, TLC is still a simple and inexpensive method that is often utilized for pilot screenings of carotenoid mixtures of

102 unknown composition. The results of TLC screenings often direct the choice of conditions for subsequent preparative chromatography and/or HPLC (Schiedt, 1995).

TLC is generally applied on a micro- or semimicro scale, which permits rapid and sharp separation of the various compounds present in the sample as well as detection of substances at the trace level. Usually silica gels are used as the adsorbent; basic oxides or carbonates such as MgO are also commonly used. Tentative identification of a compound by TLC is based on comparison of its Rf value (the distance moved by the solute divided by the distance moved by the mobile front) with that of an authentic standard.

Carotenoid thermal degradation products have often been separated by TLC methods.

Kanasawud and Crouzet (1990) used aluminum oxide TLC plates of 1.5 mm thickness with an acetone-petroleum ether (4:95 v/v) elution solvent to separate the various non­ volatile thermal degradation products of P-carotene. In a similar study, alumina TLC plates separated five nonvolatile compounds produced by heat treatment of lycopene.

Separation by TLC was followed by determination of absorption spectra by UV/Vis spectroscopy (Kanasawud and Crouzet, 1990). The degradation products of P-carotene from heating in sealed glass tubes and extrusion cooking were fractionated on an alumina

TLC plate using 10% diethyl ether in n-hexane as a solvent. Each colored band was recovered in methylene chloride and rechromatographed twice to ensure purity (Marty and Berset, 1986). Onyewu et al. (1986) also utilized TLC plates with 10% diethyl ether in n-hexane to fractionate P-carotene products degraded by heating in glycerol.

Vegetable carotenoid pigment separation has also been achieved by using column chromatography in conjunction with TLC. Tomato carotenoids were separated on an

103 alumina column, then each fraction rechromatographed on TLC plates consisting of silica

gel and MgO-kieselguhr (1:1) (Ben-Aziz et al., 1973). Ritacco et al. (1984) reported the

use of a MgO TLC plate with an eluting solvent of 20% acetone in petroleum ether for the efficient separation of Micro-Cel C carotenoid artifacts.

2.10.4 Structure elucidation

In order to either identify a known naturally occurring carotenoid or elucidate the structure of a previously unknown carotenoid, the application o f a variety of physical and chemical methods is usually required. The assignment of stereochemistry in terms of not only chirality, but also geometrical configuration, usually is the source of most of the complexity in carotenoid analyses. As a result, structural elucidation involves various spectrometric methods as well as chemical derivatization (Britton et al., 1995).

Before a carotenoid can be characterized, chromatographic purification is necessary.

Chromatography also provides information regarding the polarity of the carotenoid in question so that tentative assignment as a hydrophobic carotene, monool, or more polar carotenoid can be made (Liaaen-Jensen, 1995). After chromatographic analyses are completed, spectrometric methods are usually applied next.

2.10.4.1 UV/Vis Spectroscopy

The UV/Vis spectrum of a carotenoid is usually examined first to provide information regarding the chromophore present. Specifically, examination o f the spectral fine structure and position of the X^ax of the main absorption band can help determine the specific number o f conjugated C=C double bonds present in an aliphatic, monocyclic, or

104 dicyclic carotenoid (Britton, 1995). Values for Xmax for some of the most commonly

encountered carotenoids can be seen in Table 8 .

Yamauchi et al. (1993) determined the main UV/Vis spectral peaks of several

oxidized products o f P-carotene in order to help characterize their structures. For

example, three peaks at 405, 426, and 452 nm helped to define one of the oxidized

products as 13,15'-epoxyvinyleno-13,15'-dihydro-P,P-carotene. In a similar study El-

Tinay and Chichester (1970) also utilized UV/Vis spectroscopy to help determine the

products formed upon oxidation of P-carotene. Lycopene epoxides present in firesh

tomatoes were partially characterized through the use of UV/Vis spectroscopy. Fractions

were dissolved in petroleum and values were recorded for comparison with known

standards (Ben-Aziz et al., 1973). Britton and Goodwin (1969) also identified a lycopene

oxidation product (phytoene 1 ,2 -oxide) in ripe tomato fiaiits by comparing Xmax values

with those of phytoene.

Jensen et al. (1982) examined the fine structure of UV/Vis chromatograms of

photoisomerized all-frnw-P-carotene in order to identify exactly where and when cis-

trans isomerization occurred. Findings showed a significant hypochromic shift and a reduction in vibrational fine structure in both the 9- and 15-C/5' isomers of p-carotene in hexane when compared with the dl\-trans isomer.

2.10.4.2 IR Spectroscopy

IR spectra are most useful in carotenoid structural elucidation for their ability to reveal the presence or absence of particular functional groups in the molecular structure.

105 Carotenoid ^max (nm) Solvent

Astaxanthin 480 A 485 B,C 478 EtOH 468 P

P-carotene 429 452 478 A 435 462 487 B 435 461 485 C 450 476 EtOH 425 450 477 H,P

Lutein 432 458 487 B 435 458 485 C 422 445 474 EtOH 421 445 474 P

Lycopene 448 474 505 A 455 487 522 B 458 484 518 C 446 472 503 EtOH 444 470 502 P

Phytoene 276 286 297 H,P

Zeaxanthin 430 452 479 A 440 463 491 B 433 462 493 C 424 449 476 P

Table 8 : UV/Vis spectroscopic data for several common carotenoids. Solvents: A, acetone; B, benzene; C, chloroform; EtOH, ethanol; H, hexane; P, light petroleum.

106 Infrared (IR) radiation emitted from an IR spectrophotometer enhances the vibration of atoms in a molecule so that stretching of the bonds and variation in specific bond angles occurs. Energy is absorbed (and an absorption band formed) when the frequency of the

IR radiation matches the frequency of a particular vibrational mode. Therefore, compounds with many different atoms and fimctional groups will give rise to many different absorption bands in the IR spectrum (Figure 30). It is the specific shape, size, and position of the absorption band that lends information as to the type of functional group that it represents. The positions of the absorption bands are typically indicated by the frequency in wavenumbers (the reciprocal value of the wavelength).

Unknown purified compounds can be identified by a direct comparison of its IR spectra with that of a standard (Onyewu et al., 1986). IR spectroscopy is advantageous as a method for structural elucidation due to its simplicity and rapidity at relatively low cost.

The recent advent of Fourier-transform (FTIR) instruments has greatly improved both the

V. i*j>nr •fl

Figure 30: Typical IR spectrum for lycopene.

107 sensitivity and accuracy of traditional IR spectroscopy. FTIR spectrometers can measure

an IR spectrum with 10 to lOOx the sensitivity and in 1/000^ the time required to obtain a

similar spectrum with a traditional instrument. The main disadvantage to IR

spectroscopy, however, is the limited information given, especially for the hydrocarbon

carotenes which lack any type of functional groups (Bernhard and Grosjean, 1995).

The methylene and methyl groups present in carotenoids and their derivatives

typically give rise to bands at 2950-2850 cm'\ while the aUcene groups present in their

polyene structure gives rise to weak bands between 1650 and 1550 cm'* and a strong

band between 990 and 960 cm'* (Yamauchi et al., 1993). Onyewu et al. (1986) determined that an unknown carotenoid compound with absorption bands only at the methyl C-H (2950-2850 cm'*) and polyene C=C (1650 cm'*, 990 cm'*) wavenumbers must be a hydrocarbon carotene with no oxygenated groups present.

The presence of a Z (cis) double bond in a carotenoid compound results in the appearance of strong absorption in the region around 750-800 cm'*. Mangoon and

Zechmeister (1957) utilized this information in showing the stepwise process in which prolycopene was converted into all-trnw-lycopene. As the cis double bonds in prolycopene were converted to trans, the absorption band at approximately 750-800 cm'* became smaller until it disappeared in the IR spectrum o f all-/rn«^-lycopene.

Alcohol groups often show a very distinct broad absorption band in the 3300-3400 cm-1 range due to the stretching o f the 0-H —-O bonds, while aldehyde and ketone groups have a C=0 stretching band near 1715 cm'*. Lu et al. (1995) declared that due to the absence of absorption bands in the OH and C=0 regions, an oxygenated carotenoid compound they were examining must have an ether linkage. In this study, the

108 information obtained from the IR spectra helped to identify the structure o f a novel

carotenoid with an epoxyiridane skeleton. Ouyang et al. (1980) reported that many of the

structures of p-carotene degradation products from palm oil deodorization contained

aldehyde end groups, due to the presence of strong absorption bands in the 1700 cm'^

region. Ukai et al. (1994) reported similar absorption bands in the region of 1678 cm'*

for an oxidized lycopene product containing an aldehyde group (apo- 6 ’-lyocopenal).

2.10.4.3 Mass spectrometry

The largest contribution mass spectrometry (MS) provides for structural elucidation of carotenoids is identification of the molecular mass of the molecule. In addition, characteristic fragmentation seen in a mass spectrum lends additional information as to the functional groups present in the compound, i.e., 18 mass units (water) from alcohols,

32 mass units (methanol) from methyl ethers, 80 mass units for 5,6- and 5,8-epoxides, etc

(Figure 31). To obtain a mass spectrum only a few pg of pure sample are required, though high precision instruments may require slightly more (Enzell and Back, 1995).

MS employs several different ionization techniques for compound identification. The first, and still most popular, technique is electron impact (El) mass spectrometry. In this method, electrons are accelerated and directed at the sample, with the resulting impact causing the generation of the positively charged molecular ion M*". The particular intensity of the ions produced by El mass spectrometry is highly dependent upon specific conditions such as temperature and ionizing voltage. Marty and Berset (1988) used El mass spectroscopy to help identify the degradation products of tra« 5 -P-carotene produced during extrusion cooking. The El mass spectrum presented fragments specific for

109 M-138 M-191 -H i -H M-56

M-203

Figure 31 : Main fragmentations of the molecular ion of canthaxanthin by FAB MS.

carotenoids with 1 or 2 cyclohexenyl rings, as well as two fragments characteristic of 5,6-

epoxide or 5,8-epoxide functions.

A comparison by Ouyang et al. (1980) of sample spectra with the El mass spectra of pure 9-c/j-retinal showed identical relative abundance and fragmentation patterns, allowing positive identification of the sample p-carotene degradation product as 9-cis- retinal. Khachik et al. (1992a) elucidated the structures of eighteen different carotenoids and their oxidation products in human serum samples with the help of El mass spectroscopy. El MS has been used by several authors to help determine the molecular structures of products formed during the oxidation/degradation of P-carotene

(Kanasawud and Crouzet, 1990; Handelman et al., 1991) and lycopene (Kanasawud and

Crouzet, 1990).

Chemical ionization (Cl) generates mass spectra through ionic reactions. In this method a reagent gas such as ammonia or methane is introduced under pressure to the ion source, and then the gas is bombarded with electrons to yield ions and neutral molecules. 110 The reagent gas molecules then react with the vaporized sample molecules, resulting in a less fragmented and simpler spectra compared with El mass spectrometry. Marty and

Berset (1988) utilized CIMS with NH 3 as the reagent gas to help identify fourteen different (3-carotene thermal degradation products. In a similar study Cl mass spectrometry was used to confirm the parent ion (and therefore the molecular weight) of several P-carotene thermal degradation products (Onyewu et al., 1986).

Fast atom bombardment (FAB) mass spectrometry is a ‘soft’ ionization technique normally used for structure elucidation of either nonvolatile or thermally labile compounds. In this technique high-energy atoms such as xenon or cesium bombard a solution of the sample dissolved in an involatile matrix such as glycerol. Volatilization and ionization of the sample compound occurs by way o f dissipating kinetic energy from the bombarding atom beam, resulting in the formation of even-electron protonated molecular ion(s). Delocalization of the charge in the carotenoid extended polyene system prevents the usual protonization of the carotenoid sample from occurring, however, and results in the formation of odd-electron molecular ions. Subjecting carotenoids to FAB mass spectrometry typically yields only the radical ion MT in the molecular-ion region of the spectrum. An advantage o f FAB MS over El or Cl techniques is the lack of sample vaporization prior to volatilization, which reduces the likelihood o f carotenoid degradation during MS analysis itself. Schmitz et al. (1992) described interfacing FAB mass spectroscopy with HPLC to provide a powerful tool in the analysis of carotenoids.

A mixture of 3-nitrobenzyl alcohol (NAB) and glycerol (80:20 v/v) was described as an effective FAB matrix for analysis of the more nonpolar carotenoids due to NAB’s ability to interact with and disperse the carotenoid throughout the matrix. Using FAB-LC/MS,

111 van Breemen et al. (1993) determined the limits of detection to be as low as 5 ng and 15

ng for lutein and a-carotene, respectively. Lu et al. (1995) used FAB-MS to identify a

new hydrogen peroxide oxidation product from lycopene. Van Breemen (1996) reported

that though FAB mass-spectroscopy linked with HPLC can produce highly sensitive and

unambiguous molecular weight confirmations for both oxygenated xanthophylls and

hydrocarbon carotenes, it is limited by NAB fouling of the FAB probe that results in

lower sample throughput and decreased sensitivity.

Matrix-assisted laser desorption ionization (MALDI) mass spectrometry is one more

type of MS which has been recently applied to the analysis of carotenoids. MALDI was

originally introduced as an extremely sensitive method for mass analysis of very large

molecules such as proteins up to 300 kD in size. While other MS methods have difficulty

analyzing carotenoids due to their lack of volatility and thermal lability (El and Cl) or

lack o f solubility and surface activity in common liquid matrices (FAB), MALDI mass

spectroscopy methods have the potential to overcome most of these drawbacks

(Kaufinann et al., 1996).

2.10.4.4 NMR Spectroscopy

Nuclear magnetic resonance (NMR) spectroscopy is commonly referred to as the most powerful technique for overall structural elucidation. The reason for this is simple: proton (^H) NMR is capable of identifying the structural surroundings o f every hydrogen atom in a carotenoid, while carbon-13 (^^C) NMR identifies the degree o f saturation for each carbon atom (sp^, sp^, sp) and its structural surroundings. Detailed and NMR analyses alone can identify a carotenoid structure as well as its stereochemistry and

112 geometry of carbon-carbon double bonds. In addition, a typical *H-NMR spectrum requires only 100-200 pg of sample. As the magnetic field strengths of NMR instruments have increased to as high as 750 MHz, the dispersion of chemical shifts in samples has increased to the point that interpretation of ^H-NMR spectra has been greatly simplified. In turn, the amount of structural information that can be deduced firom the spectra has greatly increased as well (Figure 32).

13,

-.4 r u 1" 409 HH2

»,

«e- (.4 «U am ChOj

1îl

TM TJ r.C « iE tl4 tLZ jrn !

Figure 32: Olefinic section of the ^H-NMR spectrum of a typical carotenoid measured at 250 MHz (top), 400 MHz (middle), and 600 MHz (bottom).

113 Mordi et al. (1991) identified several oxidized products of P-carotene (5,6-epoxy-P- carotene, retinal, p-ionone, p-apo-14’-carotenal) using ^H-NMR spectroscopy. In a similar study, ^H-NMR was performed at 270 MHz, the ^H-^H chemical shift-correlated

(COSY) NMR technique was employed to assign shifts and couplings, and ^^C-NMR was run at 70 MHz with proton decoupling in order to identify the products formed by the peroxyl radical oxidation o f P-carotene (Yamauchi et al., 1993).

A 400 MHz ^H-NMR and a 270 MHz ^^C-NMR spectrophotometer were utilized in an analysis of the photosensitized oxidation products of lycopene firom tomato puree

(Ukai et al., 1994). Lu et al. (1995) used *^C-NMR, ^H-NMR, and COSY two- dimensional NMR experiments to help identify lycopene-oxidized products. With the application of ^H-NMR spectroscopy at 270 MHz and the help of *^C-NMR spectroscopy, Englert et al. (1979) was first able to report the complete assignment of stereochemistry to the carotene prolycopene as 7,9,7’,9’-tetra-c«-\j/,\|/-carotene.

114 CHAPTERS

MATERIALS AND METHODS

3.1 Materials

(3-carotene, lycopene, and chlorophyll standards were obtained from Sigma Chemical

Company (St. Louis, MO). Lycopene (>98% pure) was generously donated by LycoRed

Natural Products Industries (Beer-Sheva, Israel). HPLC grade methanol, methyl-tert- butyl ether, acetonitrile, and chloroform were purchased from Fisher Scientific

(Pittsburgh, PA). Helium for both gas chromatography and HPLC sparging, compressed air for gas chromatography and HPLC autosampler function, hydrogen for gas chromatography, and nitrogen for gas chromatography, sample drying, and storage were obtained from the Chemical Store at The Ohio State University. Analytical reagent grade methanol, acetone, and hexane (Fisher Scientific, Pittsburgh, PA) were used for lycopene extraction. All solid phase microextraction (SPME) fibers [polydimethylsiloxane

(PDMS), 100 pm; polydimethylsiloxane/divinyIbenzene (PDMS/DVB), 65 pm; polydimethylsiloxane/carboxen (PDMS/carboxen), 75 pm; carbowax/divinylbenzene

(CW/DVB), 65 pm; polyacrylate, 85 pm] and fiber assembly holders used in soybean oil volatile analyses were purchased from Supelco, Inc. (Bellefonte, PA). Volatile standards

115 r,r-2,4-decadienal, 2-heptenal, r,/‘-2,4-heptadienal, n-hexanal, and n-pentane were procured from Aldrich (Milwaukee, WI). Refined, bleached, and deodorized (RED) soybean oil was obtained from Abitec Corp. (Columbus, OH).

3.2 Lycopene extraction

To obtain 85-90% pure crude lycopene extract for experimental procedures, lycopene was extracted from supermarket-bought (Kroger brand) paste. Under dark conditions, 10 g of tomato paste were mixed with 25 mL of methanol and 4 g of Celite, with the mixture being tissumized for approximately 2-3 minutes. The homogenized sample was then filtered through #1 and #45 Whatman filter papers and the filtrant placed in a beaker with 50 mL of acetone/hexane (1:1). Again the mixture was tissumized for approximately 2-3 minutes and then filtered through the same #1 and #45 Whatman filter papers. The filtrate was poured into a separatory funnel, with the filtrant being discarded.

The filtrate was washed with 10 mL of distilled water 3 times in the fimnel, each time with removal of the bottom layer (waste) and retention of the top hexane layer. The top hexane layer was then transferred to 4 mL amber vials and dried under nitrogen gas until the lycopene extract was in a solid crystalline state. The vials were sealed, wrapped in aluminum foil to minimize light exposure, and stored at -4°C until needed.

3.3 Thermal treatment of carotenes

3.3.1 (3-carotene and lycopene

A solution of 500 ppm (3-carotene or lycopene solubilized in acetone was added to a glass vial that was sealed airtight with a Teflon-faced rubber septum and aluminum cap

116 (Supelco, Inc., Bellefonte, PA). A portion of the unheated solution was saved. The

re m a in in g portion in the glass vial was wrapped in aluminum foil to minimize light

exposure, then placed in a water bath set at 90°C to allow a specific amount of thermal

degradation of the carotene to occur. After the P-carotene or lycopene solution received

the proper length of thermal treatment (Figures 33 and 34, respectively), the sealed glass

vial was removed firom the water bath and allowed to cool to room temperature. Out of

the cooled glass vial one of three heated solutions containing either 250 ppm P-carotene +

250 ppm degraded products, 50 ppm p-carotene + 450 ppm degraded products, or 0 ppm

P-carotene + 500 ppm degraded products was collected (Figure 35), contents verified by

RP-HPLC, and stored in a fireezer at -4°C for further use, respectively.

3.3.2 Crude lycopene extract

A 40 mL solution o f 500 ppm lycopene solubilized in methyl-fgrf-butyl-ether

(MTBE) was added to a 100 mL glass vial that was sealed airtight with a Teflon-faced rubber septum and aluminum cap. A 13 mL portion of the unheated solution was saved.

The remaining portion in the glass vial was placed in a water bath at 90°C to allow thermal degradation of the carotene to occur. After receiving the proper length of thermal treatment, the sealed glass vial was removed from the water bath and allowed to cool to room temperature. Out of the cooled glass vial, a solution containing either 250 ppm lycopene + 250 ppm degraded products, 50 ppm lycopene + 450 ppm degraded products, or 0 ppm lycopene + 500 ppm degraded products was collected, contents verified with HPLC, then placed in a sealed glass vial and stored at -4°C.

117 2.00E+08 y = -504997% + 2E+08 1.50E-H)8 3 R^= 0.9469

.H l.OOE+08

s ft. 5.00E+07

O.OOE+00 100 150 200 250 Time (minutes at 90 deg Q

Figure 33: Increased degradation of P-carotene (decreased P-carotene peak size) with increased thermal processing at 90°C.

2.0E+08

1.6E+08 y = -208486% + 2E+08 P = 0.9983 S 1.2E+08 .1 ^ 8.0E+07 ft.2 4.0E+07

O.OE+OO 0 200 400 600 800 1000 Time (minutes at 90 deg C)

Figure 34: Increased degradation of lycopene (decreased lycopene peak size) with increased thermal processing at 90°C.

118 water bath ® 90°C

40 mLs 500 ppm mcreasmg B-carotene or heating time lycopene in acetone

15 mLs 8 mLs 8 mLs 8 mLs unheated 250 ppm + 50 ppm + 0 ppm + 500 ppm 250 ppm degraded 450 ppm degraded 500 ppm degraded + 0 ppm

Figure 35: Schematic of P-carotene or lycopene heating and fractionation into solutions with varying degrees of thermal degradation.

3.4 Sample preparation and storage for the chlorophyll-photosensitized singlet oxygen oxidation of soybean oil

To study the effects of thermal processing on the singlet oxygen quenching abilities of degraded carotenes, a 2.5 mL solution containing thermally processed p-carotene or

119 lycopene (250 ppm carotene + 250 ppm degraded products, 50 ppm carotene + 450 ppm degraded products, or 0 ppm carotene + 500 ppm degraded products) was added in triplicate to 30 mL glass serum vials containing 22.5 mL of purified soybean oil and 3 ppm chlorophyll sensitizer. The resulting 25 mL soybean oil solutions then contained either 25 ppm carotene + 25 ppm degraded products, 5 ppm carotene + 45 ppm degraded products, or 0 ppm carotene + 50 ppm degraded products (Figure 36).

25 ppm 250 ppm 2.5 mL + 250 ppm 25 ppm

5 ppm + 50 ppm + 45 ppm 450 ppm

0 ppm 0 ppm + 2.5 mL + ► 500 ppm 50 ppm

22.5 mL soybean oil 25 mL soybean oil + 3 ppm chlorophyll + carotene

Figure 36: Addition of degraded carotene solutions to soybean oil to create the final concentrations in the tested sample solutions.

120 Similarly, 2.5, 1.25, or 0.25 mL of 500 ppm unheated p-carotene or lycopene were added to vials containing identical quantities of soybean oil and chlorophyll to create 25 mL control solutions containing either 50, 25, or 5 ppm of either lycopene or P-carotene.

Other controls included samples containing only oil, or oil and chlorophyll (Figure 37).

50 ppm

0.25 mL 500 ppm 25 ppm + 0 ppm

5 ppm

oil + oil only chlorophyll

Figure 37: Addition of unheated carotene solutions to soybean oil to create the final concentrations in the tested control sample solutions.

121 Each vial was then sealed airtight (for headspace oxygen content determination only)

with Teflon-faced rubber septa and aluminum caps and stored at 25°C in a mirrored light

box (70 cm X 50 cm x 60 cm) for 24 hours (Figure 38). The light source itself consisted

of a 100-watt fluorescent lamp (Lights of America, Inc., Walnut, CA) with an intensity of

1,650 lumens. The oxidative stability of the sample was determined by measuring both

peroxide value (AOCS, 1980) and the headspace oxygen content of the sample vial by

thermal conductivity gas chromatography every 4 hours during the aforementioned 24

hour period.

m

Figure 38: Diagram o f mirrored light box for sample storage imder light.

122 3.5 Sample preparation and storage for thermally-induced degradation of soybean oil

To study the effects of degraded carotenes on the oxidative stability of thermally-

treated soybean oil, 2.5 mL acetone solutions containing either 500 ppm P-carotene or

lycopene, or 0 ppm carotene + 500 ppm degraded P-carotene or lycopene were added in

triplicate to 30 mL glass serum vials containing 22.5 mL soybean oil, respectively. Each

vial (including controls containing 22.5 mL soybean oü and 2.5 mL acetone) was then

sealed airtight with Teflon-faced rubber septa and aluminum caps and then stored at 60°C

in the dark for 8 days. The oxidative stability of the sample was determined by

measuring the headspace oxygen content of the sample vial by thermal conductivity gas

chromatography after 8 days of storage at 60°C.

3.6 Headspace oxygen analysis

A Hewlett Packard 5890 Gas Chromatograph equipped with a thermal conductivity detector (TCD) and a stainless steel molecular sieve packed column (13x, 80/100, '

Alltech, Deerfield, IL) was utilized to determine the extent of headspace 0% depletion

(and corresponding soybean oil oxidation). High purity (99.995%) helium gas was used as both the carrier and auxiliary gas at a rate o f 30 mL/min. The injector, detector, and oven temperatures were set to 120°C, 150°C, and 40°C, respectively. Every 4 hours for

24 hours 100 pL o f headspace air from each sealed vial was injected into the GO manually using an airtight syringe. Ambient air was used as a reference o f O 2 content to correct for day to day chromatographic variability. A Hewlett Packard HP 3396A integrator (Avondale, PA) was used to record the oxygen peak in electronic counts.

123 3.7 Solid phase microextraction (SPME) volatiles analysis

Volatile compounds created as a result o f thermal (stored at 60°C in oven) or light- induced (stored in light box, see Figure 35) oxidation of soybean oil containing either 13- carotene or lycopene were isolated by solid phase microextraction (SPME), then separated and tentatively identified by gas chromatography. All soybean oil/carotene samples exposed to either oxidizing treatment were prepared by adding the following to

30 mL serum vials: 1) 22.0 mL RED soybean oil; 2) 2.5 mL of acetone containing either

500 ppm P-carotene/lycopene, 500 ppm degraded P-carotene/lycopene, or acetone only

(control); 3) 500 pL pure acetone (for thermally treated samples) or 500 pL 150 ppm chlorophyll in acetone (for samples placed in light box). Resulting soybean oil samples therefore contained 50 ppm P-carotene/lycopene, 50 ppm degraded P-carotene/lycopene, or just oil and acetone (controls), respectively. In addition, all samples placed in the light box contained 3 ppm chlorophyll.

Three different samples (50 ppm P-carotene/lycopene, 50 ppm degraded P- carotene/lycopene, oil only control), each in triplicate, received either the oven storage treatment or the light box storage treatment for 0, 5, 10,20, or 30 days. At the end of the storage time, SPME headspace analysis of each vial was conducted by penetration of the vial septum and adsorption of volatiles present with a fijsed silica fiber coated with a stationary phase of polydimethylsiloxane (PDMS) (100 pm film). The adsorption time was exactly 40 minutes, occurring at a fixed temperature of 60°C. The fiber was then retracted firom the vial and immediately inserted and desorbed for 2 minutes into the inlet liner of a Hewlett Packard 5890 Series II Gas Chromatograph equipped with an HP-5

124 crosslinked 5% phenyl-methyl silicone fused silica capillary column (30 m x 0.32 mm

ID, 0.25 pm film) and flame ionization detector (FID). Each GC run consisted of a ramped oven temperature profile as follows: 40°C (0 minutes) to 240°C at 8°C/min., then 240°C for 5 minutes (total run time 30 minutes). The carrier gas was nitrogen at a column flow rate of 4 mL/min. The FID detector was set at 300°C, while the injector

(and desorption temperature) was set at 250°C. The GC chromatogram total peak area from the oil only control sample was set at a value of 100%, and the peak areas of samples containing dX\-trans or degraded carotenes were compared at 0, 5, 10, 20, and 30 days as a percentage of the oil only control.

Gas chromatography-mass spectrometry (GC-MS) was utilized to positively identify volatile oxidation products of soybean oil that were adsorbed onto a solid phase microextraction (SPME) polydimethylsiloxane (PDMS) fused silica fiber and then desorbed onto a GC capillary column. Specifically, a Hewlett Packard 5890 Gas

Chromatograph equipped with an HP-5 crosslinked 5% phenyl-methyl silicone fused silica capillary column (30 m x 0.32 mm ID, 0.25 p.m film) as well as a computer monitor interface containing Hewlett Packard 59970 MS Chemstation Software was utilized for volatile separation. A Hewlett Packard 5970 Series Mass Selective Detector equipped with a Hewlett Packard 59822 B Ionization Gauge Controller was used to collect a mass spectrum for each volatile compoimd. Structural identification of the main volatile peaks was made with the help of Wiley Mass Spectrum Library computer software.

125 Each sample prepared for GC-MS analysis was held at 60°C for twelve hours in a

water bath while a PDMS SPME fiber was exposed to the sample’s headspace. The fiber

was then immediately inserted into the GC that was equipped with a special septum and

inlet liner for SPME fibers. The fiber was desorbed in the inlet liner for 4 minutes at

250°C, then retracted. Each GC-MS analysis consisted of an overall run time of 35

minutes, with a 25 minute ramping program in the middle. The oven was held at 40°C

for 5 minutes at the beginning of each nm, then ramped at a rate of 8°C/min. from 40°C

to 240°C, upon which the oven temperatiure was held at 240°C for the last 5 minutes of

the run. The injector was set at 250°C, while the detector temperature was 300°C.

Helium was used as the carrier gas through the capillary column at a flow rate of

approximately 1.5 mL/min.

3.8 Analysis of carotene degradation products

3.8.1 Separation

P-carotene and lycopene, along with their degradation products, were separated using

a Hewlett Packard Series 1050 HPLC system and Hewlett Packard 3396A integrator.

Isocratic reverse-phase liquid chromatography was conducted utilizing an Hewlett

Packard CDS Hypersil analytical column (5 pm, 200 x 4.6 mm) and a mobile phase

solvent system of methanol: methyl-Zert-butyl-ether (90:10). The flow rate for each run

was 1 mL/min. with a total run time of 20 minutes. A UV/Vis detector was set at 460 nm

for P-carotene, 472 for lycopene, and 350 nm for degraded product analyses to maximize peak sensitivity.

126 3.9 Statistical analysis

Two sample t-test and one-way Analysis of Variance (ANOVA) were used to determine the presence of differences (a=0.05) between two samples and groups of samples, respectively. Tukey’s Studentized Range Test determined which particular sample groups were different at a=0.05. All the aforementioned statistical analyses were calculated on the Minitab Version 10.1 for Windows statistical software package.

127 CHAPTER 4

RESULTS AND DISCUSSION

4.1 Separation and identification of carotenes and their thermal degradation products

Isocratic nonaqueous reverse-phase high performance liquid chromatography (RP-

HPLC) was utilized to isolate, quantify, and determine the extent of degradation in heated solutions of aS\rtrans P-carotene, all-tranj-lycopene, and crude lycopene extract, respectively (Nelis and De Leenheer, 1983). Though hexane is often used to solubilize carotenes for HPLC analysis, it did not effectively solubilize carotenes at concentrations higher than 500 ppm (Craft and Soares, 1992). Acetone, which more effectively solubilized the carotenes, was therefore utilized as the solvent in each experiment.

After solubilization, heating a solution of P-carotene in sealed vials at 90°C resulted in a fairly rapid linear degradation, or “bleaching”, of the dl\-trans peak (Figure 39). El-

Tinay and Chichester (1970) reported a similar linear relationship between the loss of P- carotene and time. The A,max for the dl\-trans p-carotene peak was found to be 460 nm, so the UV/Vis detector was set at this wavelength to quantify (in AU) the remaining all- trans P-carotene at each stage of thermal degradation (Stahl et al., 1993). As seen in

Figure 39, after approximately 3.5 hours of thermal treatment, the size of the dl\-trans P-

128 1.8 X 10* AU 8.5 X 10^ AU

I I 460 nm 2.0 X 10^ AU I

■ 1 2.0 X 10^ AU JJL I

350 nm

jw A ji.

500 ppm 250 ppm 50 ppm 0 ppm P-carotene + + + + 0 ppm 250 ppm 450 ppm 500 ppm degraded products

Figure 39: Extent of thermal degradation of each collected P-carotene sample solution as seen by RP-HPLC (arrows indicate all-fruw-p-carotene peak).

carotene peak at 460 nm was halved, from 1.8 x 10* AU to 8.5 x lO’ AU. A detector wavelength of 350 nm has been shown to effectively reveal chromatographic peaks o f carotene thermal degradation products, due to the presence of conjugated trienes in these compounds (Handelman et al., 1991). Such P-carotene degradation peaks were clearly visible at 350 nm on the chromatogram showing the 250 ppm P-carotene + 250 ppm degraded products solution. After 8 hours of thermal treatment at 90°C, only one-tenth of 1% o f the ail-trans p-carotene peak remained (effectively 0 ppm p-carotene + 500 ppm

129 degraded products). At the point in which all of the a!L\-trans ^-carotene had been degraded, many of the peaks representing oxidized products seen in the corresponding chromatogram run at 350 run were reduced in size as well, indicating extensive degradation of the carotene conjugated double bond system.

Continued thermal treatment of lycopene solubilized in acetone also resulted in linear reduction o f the dl\-trans peak size (Figure 40). Quantitation of lycopene was conducted at a wavelength o f472 nm, the reported A,max for the ail-trans isomer (Mangoon and

1.7 X 10* AU 7.0 X 10^ AU I I 472 nm 1.4 X 10^ AU I 0 AU L L I 350 nm

I

rJkJu\_ U i 500 ppm 250 ppm 50 ppm 0 ppm lycopene + + + + 0 ppm 250 ppm 450 ppm 500 ppm degraded products

Figure 40: Extent o f thermal degradation of each collected lycopene sample solution as seen by RP-HPLC (arrows indicate all-traw-lycopene peak).

130 Zechmeister, 1957). The small peaks visible at 472 nm that eluted before the a[\-trans

peak in the unheated lycopene solution were most likely oxidized products of lycopene

formed unintentionally during sample storage, handling, and chromatography (Emenhiser

et al., 1995). Piretti and Diamante (1996) hypothesized that anomalous peaks seen in

commercial standards o f all-/ran5 lycopene using RP-HPLC could also be geometrical

isomers (13-cfj) o f the predominant dX\-trans form.

As was the case with P-carotene, great care was taken to heat the lycopene solution

for just the right amount of time necessary to obtain the proper proportion of dl\-trans

lycopene to degraded products. In a manner similar to the results seen in Figure 39,

chromatograms o f the 250 ppm lycopene + 250 ppm degraded products solution viewed

at 472 nm and then 350 nm in Figure 40 revealed a 50% reduction in the size (in AU) of

the all-trans lycopene peak, and a marked increase in the size of the oxidized/degraded

peaks, respectively. After 14 hours of thermal treatment at 90°C, the elY-trans lycopene

peak was completely gone, which indicated thermal degradation of the carotene

equivalent to the desired 0 ppm lycopene + 500 ppm degraded products solution.

Lycopene appeared more resistant to thermal degradation when compared with (3-

carotene (14 hours at 90°C to completely degrade vs. 8 hours, respectively). This observation did not agree with the findings of Henry et al. (1998), who reported that lycopene was most susceptible to thermal degradation, with a rate constant approximately double that of p-carotene. Differences in the carotenes’ surrounding matrixes (acetone vs. safflower seed oil) may have accounted for the observed differences in the relative thermal degradation rates of lycopene and P-carotene (Onyewu et al., 1986). In addition.

131 the relatively high pressures that were present in the sealed vials during heating of the

carotene solutions have also been shown to specifically increase the degradation rate of

P-carotene (Marty and Berset, 1990).

A crude lycopene extract from tomato paste was the third solution to be exposed to

extended thermal degradation at 90°C. HPLC chromatograms set at both 472 and 350 nm were run at each stage of thermal degradation o f the crude lycopene extract solution, from unheated to completely degraded (Figure 41). The unheated crude lycopene

1.5 X 10“ AU 8.4 X 10^ AU 472 nm

1.9 X 10' AU

1.4 X 10^ AU

11. I

350 nm

r A_ 500 ppm 250 ppm 50 ppm 0 ppm lycopene + extract 0 ppm 250 ppm 450 ppm 500 ppm degraded products

Figure 41 : Extent of thermal degradation of each collected crude lycopene extract sample solution as seen by RP-HPLC (arrows indicate all-P-nw-lycopene peak).

132 extract solution revealed, in addition to the predominant all-tra«j lycopene peak at

approximately 14.3 minutes, a minor peak at approximately 17.5 minutes that was

identified (by spiking the solution with a standard) as dl\-trans P-carotene. Previous

studies have also shown red tomatoes to contain a small quantity of P-carotene in

addition to the predominant lycopene pigment (Hakala and Heinonen, 1994; Hart and

Scott, 1995). However, ai\-trans lycopene typically elutes before a!L\-trans P-carotene

only when acetonitrile is used in the mobile phase (Nelis and De Leenheer, 1983; Stahl et

al., 1992; Scott and Hart, 1993). The reason why in this case lycopene actually eluted

before p-carotene without the use o f acetonitrile in the mobile phase could not be

determined.

The crude lycopene extract solution degraded at a significantly slower rate than either

the p-carotene or lycopene solutions. While P-carotene and lycopene only required 3.5

and 8 hours to lose half of their peak area (in AU), respectively, the crude lycopene

extract solution required 24 hours to become half degraded. To become completely

degraded, the crude lycopene extract solution required 275 hours, or over eleven days!

Lipid-soluble antioxidants such as tocopherols that were present in the tomato paste may

have been extracted with the lycopene and helped to prevent thermal oxidation and

degradation of the lycopene present by acting as free radical scavengers (Burton and

Ingold, 1989; Palozza and Krinsky, 1992; Heinonen et al., 1997). After 275 hours of thermal treatment, less than 1/10 o f 1% of the dM-trans lycopene peak area remained, effectively creating the 0 ppm lycopene extract + 500 ppm degraded products solution.

133 4.2 Effects of degraded carotenes on chlorophyll-photosensitized singlet oxygen oxidation of soybean oil

4.2.1 P-carotene

To determine how degraded carotenes might affect the oxidative stability of a soybean oil model system exposed to light, proper controls must first be created to account for any extraneous variables. For this reason, in addition to the three samples containing 3 ppm chlorophyll and varying concentrations of degraded carotenes (5 ppm,

25 ppm, or 50 ppm), six control samples were also tested. Figure 42 shows seven of

^— B-car 50 ppm

#— B-car 25 ppm B-car 5 ppm 25+25

5 +45 0+ 50

oil+chloro (light)

12 Time (hours)

Figure 42: Effects of thermally degraded P-carotene on the headspace oxygen depletion of soybean oil containing 3 ppm chlorophyll under light storage at 25 °C.

134 the nine soybean cil/carotene/chlorophyll samples that were tested for headspace oxygen

depletion over a 24 hour period by injection of 100 pL headspace air into a gas

chromatograph equipped with a thermal conductivity (TC) detector.

The control sample containing 50 ppm unheated al\-trans (3-carotene (B-car 50 ppm)

had significantly less (p<0.05) headspace oxygen depletion after 24 hours than all other

samples, which indicated that there was less oxygen uptake and subsequent oxidation of

the oil (Figure 42, Table 9). This result was not surprising considering the numerous

studies citing p-carotene as an effective quencher of singlet oxygen (Foote and Dermy,

Depleted pmoles Sample* Grouping** Oi/mL headspace

50 ppm ail-trans 3.35 A

25 ppm all-trans 4.00 B

5 ppm al\-trans 4.21 B

25 ppm all-tram + 25 ppm degraded 4.06 B

5 ppm all-tram + 45 ppm degraded 4.06 B

0 ppm all-tram + 50 ppm degraded 4.56 C

oil + chlorophyll (in light) 4.24 BC

* Indicates final concentration of P-carotene in soybean oil. ** Means with different letters are significantly different (p<0.05).

Table 9: Tukey’s Studentized Range Test for the effects o f P-carotene and thermally degraded p-carotene combinations on headspace oxygen depletion (pmoles Oi/mL headspace) after 24 hours of light exposure.

135 1968; Edge et al., 1997), particularly in soybean oil model systems (Warner and Frankel,

1987; Jung and Min, 1991). The control sample containing 25 ppm unheated al\-trans p- carotene (B-car 25 ppm) was very close in headspace oxygen depletion after 24 hours to the sample containing 25 ppm 2l\-trans P-carotene + 25 ppm degraded products (25 + 25)

(4.00 vs. 4.06 pmoles Oz/mL, respectively). Similarly, the control sample containing only 5 ppm unheated dl\-trans P-carotene (B-car 5 ppm) was very close in headspace oxygen depletion after 24 hours to the sample containing 5 ppm d!i\-tram p-carotene + 45 ppm degraded products (5 +45) (4.21 vs. 4.06 pmoles Oz/mL, respectively) (Figure 42).

None of the headspace oxygen depletion values for the 5 and 25 ppm controls and their corresponding degraded mixture samples (5 + 45, 25 + 25) were significantly different

(Table 9). The only dissimilarity between the 5 and 25 ppm aï\-trans P-carotene controls and their corresponding 5 + 45 or 25 + 25 degraded mixtures was the presence of degraded carotenes in the non-control samples. Therefore, the degraded P-carotene compounds in this case did not act as either prooxidants or antioxidants in the soybean oil model system, but rather had no overall effect on the soybean oil oxidative stability under the light.

Of particular interest was the relationship in headspace oxygen depletion between the sample containing 50 ppm zil-trans p-carotene (0 + 50), and the control samples containing 50 ppm all-n*ara p-carotene (B-car 50 ppm) and oil and 3 ppm chlorophyll

[oil + chloro (light)], respectively. The 50 ppm degraded P-carotene sample showed a significantly higher (p<0.05) level of headspace oxygen depletion after 24 hours than all the other samples (Figure 42) at 4.56 pmoles Oz/mL, except for the oil and chlorophyll

136 only control at 4.24 fimoles Oz/mL (Table 9). Obviously the degradation of the P~

carotene polyene system from 50 ppm dll-trans p-carotene to 50 ppm degraded products

caused a dramatic change in the effects o f the carotene on soybean oil stability. With the

polyene system intact in the 50 ppm dl\-trans sample (B-car 50 ppm), headspace oxygen

depletion was significantly reduced. After the thermal destruction of the polyene system

in the 50 ppm degraded sample (0 + 50), however, headspace oxygen depletion was much

higher. This indicated that the loss of the extensive conjugated double bonds present in

dl\-trans P-carotene by heating eliminated its ability to quench singlet oxygen and

prevent photosensitized oxidation of the soybean oil. Similarly, Lee and Min (1990)

reported a decrease in the singlet oxygen quenching rate of five carotenoids as their

respective conjugated double bond coimts decreased. The 50 ppm degraded P-carotene

sample not only provided no antioxidant activity in the form of singlet oxygen quenching,

but in fact caused a slight, though not significant, prooxidant effect when compared with

the control containing only oil and chlorophyll (Figure 42).

In order to prove that oxidation of the soybean oil was indeed coming from singlet

oxygen oxidation, two control samples were prepared, stored for 24 hours, and examined

for headspace oxygen depletion (data not shown). The first control contained only

soybean oil and was stored in the light at 25°C, while the second control contained soybean oil as well as 3 ppm chlorophyll and was stored in the dark at 25°C. After 24 hours of storage in the light, the headspace oxygen depletion value for the sample containing only soybean oil and no added chlorophyll was only 0.74 pmoles Oz/mL, significantly less than the control in light storage containing soybean oil and 3 ppm

137 chlorophyll (4.24 p.moles Oz/mL). These results reiterated the importance of chlorophyll

as a photosensitizer that can excite ground-state triplet oxygen to the more reactive

singlet oxygen state (Clements et al., 1973; Usuki et al., 1984; Fakourelis et al., 1987), as

well as supported singlet oxygen as the main source of soybean oil oxidation under light

storage. After 24 hours of storage in the dark, the second control sample containing

soybean oil as well as 3 ppm chlorophyll experienced almost no headspace oxygen

depletion (0.004 p.moles Oz/mL). The importance of light as an exciter of sensitizers in

photosensitized oxidation reactions has been well documented in the literature (Clements

et al., 1973; Warner and Frankel, 1987; Rahmani and Csallany, 1998). The controls

showed that in the presence of light and a sensitizer like chlorophyll (two requisites for

singlet oxygen oxidation), oxidation o f soybean oil occurred, while samples lacking

either light exposure or chlorophyll were oxidized to a much lesser extent. The data

therefore strongly supported singlet oxygen as the main source of soybean oil oxidation

under light storage.

Peroxide values were also determined for soybean oil samples exposed to light

containing the same combinations of dl\-trans and degraded P-carotene to better determine the relationship between carotenes and oil oxidative stability. In a manner similar to the headspace oxygen depletion data for P-carotene, control samples containing

50 ppm dl\-trans P-carotene had the lowest peroxide values after 24 hours of storage at

25°C under the light (Figure 43). Their average value of 8.83 meq HzOz/Kg soybean oil after 24 hours o f light storage was significantly lower than those of either the 25 ppm

(11.33 meq HzOz/Kg) or 5 ppm (13.67 meq HzOz/Kg) dl\-trans p-carotene samples

138 15 -, . B-car 50 ppm . B-car 25 ppm 12.5 - . B-car 5 ppm .25+25 .5+45

= 10 - .0+50 o . oil+chloro (light) DX)

2.5 -

8 12 16 Time (hours)

Figure 43 : Effects of thermally degraded P-carotene on the peroxide value of soybean oil containing 3 ppm chlorophyll under light storage at 25°C.

(Table 10). Previous studies have also reported the effectiveness of all-trans P-carotene at lowering peroxide values in soybean oil (Warner and Frankel, 1987; Jimg and Min,

1991).

Peroxide values for the 25 ppm all-trans P-carotene control sample were higher after

24 hours than the sample containing 25 ppm all-trans p-carotene + 25 ppm degraded products (Figure 43), but not significantly (Table 10), at 11.33 meq HiOz/Kg vs. 10.50 meq HiOi/Kg, respectively. Similarly, peroxide values for the 5 ppm all-trans p-

139 Milliequivalents Sample* Grouping** HzOz/Kg soybean oil

50 ppm edl-trans 8.83 A

25 ppm all-trans 11.33 B

5 ppm all-trans 13.67 CD

25 ppm all-trans + 25 ppm degraded 10.50 B

5 ppm all-trans 4- 45 ppm degraded 13.00 C

0 ppm all-trans + 50 ppm degraded 15.00 D

oil + chlorophyll (in light) 13.50 C

* Indicates final concentration of p-carotene in soybean oil. ** Means with different letters are significantly different (p<0.05).

Table 10: Tukey’s Studentized Range Test for the effects of P-carotene and thermally degraded P-carotene combinations on peroxide formation in soybean oil (milliequivalents HiOz/Kg soybean oil) after 24 hours of light exposure.

carotene control samples (B-car 5 ppm) were slightly higher than their 5 ppm dULl-trans P- carotene + 45 ppm degraded products sample (5 + 45) coimterparts (Figure 43), but not significantly (Table 10), at 13.67 meq HzOz/Kg vs. 13.00 meq HiOi/Kg, respectively.

The lack of a significant difference between the all-trans 25 and 5 ppm p-carotene controls and the samples containing equal amounts of all-trans P-carotene as well as 25 and 45 ppm of degraded products, respectively, indicated that the degraded P-carotene products were not acting in either a prooxidant or antioxidant fashion. Rather, peroxide values for the 25 ppm all-trans P-carotene + 25 ppm degraded products (25 + 25) and 5

140 ppm dl\-trans P-carotene + 45 ppm degraded products (5 + 45) samples were based more

on the concentration of the dl\-trans P-carotene present, with higher levels of sàl-trans

corresponding with lower peroxide values (Warner and Frankel, 1987). The degraded P~

carotene products, therefore, when combined in a sample with sl\-trans P-carotene,

showed no significant effect on the oxidative stability of soybean oil.

When present in a soybean oil model system sans dl\-trans P-carotene, degraded p~

carotene products seemed to have a negative effect on the oxidative stability of soybean

oil. The sample containing 50 ppm degraded P-carotene (0 -+- 50) produced the highest

peroxide values after storage under light for 24 hours (Figure 43). At 15.00 meq

HiOz/Kg, the 50 ppm degraded p-carotene sample had significantly higher (p<0.05)

peroxide values than the control containing only oil and chlorophyll (Table 10). Unlike

samples containing mixtures of degradation products and aW-trans p-carotene, this result

depicted P-carotene degradation products as being prooxidants when present alone in

soybean oil exposed to light.

Peroxide values for two other control samples were also calculated to verify the presence of singlet oxygen oxidation as the main source of peroxide formation in the soybean oil samples (data not shown). One control sample was stored under the light and contained only soybean oil with no added chlorophyll. The resulting peroxide value for this sample was very low, at 2.5 meq HzOz/Kg, which indicated significantly reduced soybean oil oxidation under the light when the model system lacked adequate concentrations o f photosensitizers like chlorophyll (Usuki et al., 1984). The second control sample also supported photosensitized oxidation by singlet oxygen as the main

141 source o f peroxide formation. Containing both oil and 3 ppm chlorophyll, the control was stored for 24 hours in the dark, and then upon testing showed a peroxide value of only 1.5 meq H^Oz/Kg soybean oil. In this case the importance of light energy in the process o f chlorophyll (sensitizer) excitation was revealed, because even with chlorophyll present, singlet oxygen could not form to oxidize the oil due to a lack light energy

(Rahmani and Csallany, 1998). Peroxide value results for both controls therefore supported singlet oxygen as the main source of soybean oil oxidation when stored in the light at 25°C.

When individual headspace oxygen values for the soybean oil/p-carotene samples stored under the light were correlated with the samples’ corresponding peroxide values, a fairly linear (R^ = 0.96) regression line was formed (Figure 44). The general trendline clearly revealed that as headspace oxygen depletion increased in samples, peroxide values tended to increase as well. This result suggested that the oxygen being depleted from the headspace was indeed oxidizing the soybean oil by bonding to fatty acids to form peroxides. In addition, the similarities between the data obtained by the two different methods (headspace oxygen depletion and peroxide value) strengthened the validity o f the overall conclusions reached regarding the effects of P-carotene and its degradation products on the oxidative stability of soybean oil stored under the light.

4.2.2 Lycopene

Greater than 98% pure dl\-trans lycopene solubilized in acetone was added to soybean oil samples stored under the light to create nine different oil/carotene mixtures,

142 5

G O 4 GÇJ a. £• « ■a 2 g 3

o çj c.« 2 2 en o •B S 2 3 S y = 0.3563x - 0.2497 1 = 0.9588

0 0 5 10 15 Peroxide value (meq BbOa/Kg oil)

Figure 44: Correlation between headspace oxygen depletion and peroxide value for soybean oil containing degraded p-carotene under light storage at 25°C.

each similar in concentration to the aforementioned p-carotene samples. The headspace oxygen depletion values (in pmoles Oz/mL headspace) that were recorded every 4 hours during 24 hours of storage in the light for seven of the nine samples can be seen in Figure

45. In a manner similar to the- previously discussed sample control containing 50 ppm all-trans P-carotene, the 50 ppm all-trans lycopene sample control possessed headspace oxygen depletion values that were lower than all other samples throughout the 24 hour time course. After 24 hours of light storage the 50 ppm all-trans lycopene sample control

143 4.5 -, . lyc 50 Rjm 4 - -lyc 25 ppm . lyc 5 ppm .25+25 lî .5+45 .0+50 If . oil+chloro (light) II CC E 3 E

Time (hours)

Figure 45: Effects of thermally degraded lycopene on the headspace oxygen depletion o f soybean oil containing 3 ppm chlorophyll under light storage at 25°C.

lost 2.62 pmoles Oi/mL headspace, which was significantly lower (p<0.05) than all other samples except for the 25 ppm aï\-trans lycopene sample control, which lost 3.21 pmoles

Oz/mL headspace, respectively (Table 11). These results agreed with several other studies which reported d\\-trans lycopene to be very effective in the reduction of photosensitized oxidation because o f its ability to quench singlet oxygen back to its less reactive triplet state (Di Mascio and Sies, 1989; Edge et al., 1997). As the all-trans

144 Depleted pmoles Sample* Grouping** O z /m L headspace

50 ppm all-trans 2.62 A

25 ppm all-trans 3.21 AB

5 ppm all-trans 3.65 BC

25 ppm all-trans + 25 ppm degraded 3.54 BC

5 ppm all-trans + 45 ppm degraded 4.07 BC

0 ppm all-trans + 50 ppm degraded 4.12 C

oil + chlorophyll (in light) 3.79 BC

* Indicates final concentration of lycopene in soybean oil. ** Means with different letters are significantly different (p<0.05).

Table 11 : Tukey’s Studentized Range Test for the effects of lycopene and thermally degraded lycopene combinations on headspace oxygen depletion (pmoles Oi/mL headspace) after 24 hours of light exposure.

lycopene concentration in the soybean oil control samples decreased firom 50 ppm to 5

ppm, headspace oxygen depletion increased significantly (p<0.05) from 2.62 pmoles

Oz/mL to 3.65 pmoles Oz/mL (Table 11). This indicated that even at concentrations as high as 50 ppm, lycopene was still the limiting factor in quenching singlet oxygen to prevent photosensitized oxidation of the soybean oil.

The control samples containing 25 ppm and 5 ppm, respectively, of 2l\-tram lycopene compared favorably with corresponding samples containing 25 ppm all-trans lycopene +

25 ppm degraded products (25 + 25) and 5 ppm all-trans lycopene + 45 ppm degraded products (5 + 45), respectively (Figure 45). Though not significant, headspace oxygen

145 depletion for the 25 +25 and 5 + 4 5 samples was higher than for their corresponding 25 ppm and 5 ppm dX\-trans lycopene controls (Table 11). The degraded lycopene products in the 25 + 25 and 5 + 45 samples may have exerted a slight prooxidant effect in the soybean oil samples when exposed to light over a 24 hour period of storage. Due to the lack of a significant difference at a = 0.05, however, the validity of the observed prooxidant effect could be questioned.

A slight though again insignificant (p>0.05) prooxidant effect was noted for the degraded lycopene products in the sample containing 0 ppm dl\-trans lycopene + 50 ppm degraded products (0 + 50) when compared with the control containing only oil and 3 ppm chlorophyll [oil + chloro (light)] (Figure 45). Headspace oxygen depletion values for the two samples were 4.12 pmoles Oz/mL vs. 3.79 pmoles Oz/mL, respectively

(Table 11). As was the case with P-carotene, thermal degradation of the polyene system found in dl\-trans lycopene led to a very significant increase in the headspace oxygen depletion of the soybean oil samples (Figure 45, Table 11). Apparently the degraded lycopene products did not possess enough conjugated double bonds to effectively absorb energy firom singlet oxygen and lower it to its ground-state triplet form. As a result, in samples containing only degraded lycopene products, singlet oxygen was able to more frequently abstract electrons from unsaturated free fatty acids typically found in soybean oil such as linoleic and linolenic. The result was increased peroxy-radical and peroxide formation (and higher headspace oxygen depletion) in the soybean oil sample.

Singlet oxygen was proven to be the primary source of lipid oxidation by the use of two controls: 1) a sample that contained only soybean oil and was stored in the light at

146 25°C, and 2) a second control that contained soybean oil as well as 3 ppm chlorophyll

and was stored in the dark at 25°C (data not shown). The oil only control displayed a

headspace oxygen depletion value of only 0.70 jamoles Oz/mL after 24 hours of light

storage, which indicated that little singlet oxygen was being created in the absence of

significant concentrations of chlorophyll sensitizer. The small quantity of oxidation that

did occur was probably due to the approximately 0.1 ppm of chlorophyll present naturally

in refined, bleached, and deodorized soybean oil (Usuki et al., 1984). The sample stored

in the dark for 24 hours containing 3 ppm chlorophyll experienced no headspace oxygen depletion (0 pmoles Oz/mL). This result strongly suggested the need for light energy to initiate the processes necessary for the oxidation and corresponding headspace oxygen depletion that occurred in the soybean oil samples. Headspace oxygen depletion of soybean oil samples containing lycopene and/or lycopene degradation products was therefore mainly the result of singlet oxygen oxidation.

A second method that was utilized to quantify the extent of lipid oxidation in samples containing lycopene and/or lycopene degradation products after 24 hours of storage in the light was peroxide value (AOCS, 1980). Figure 46 clearly shows the decrease in peroxide value as the concentration of all-trans lycopene increased in sample controls.

Actual peroxide values decreased from 10.83 meq HzOz/Kg oil for the 5 ppm all-trans lycopene sample to 5.83 meq HzOz/Kg oil for the 50 ppm all-trans lycopene sample

(Table 12). Higher concentrations of up to 50 ppm all-trans lycopene were therefore increasingly effective in preventing peroxide formation in the soybean oil. Samples containing 25 ppm all-trans lycopene + 25 ppm degraded products (25 + 25) and 5 ppm

147 16 -, lyc 50 ppm 14 - lyc 25 ppm -a— lyc 5 ppm 12 - _*_25+25 o -Û—5+45 SX) 1 0 L -o—0+50 -O oil+chloro (light) 8 -

6 .

4 -

2 -

0 12 16 20 24 Time (hours)

Figure 46: Effects of thermally degraded lycopene on the peroxide value of soybean oil containing 3 ppm chlorophyll under light storage at 25°C.

all-trans lycopene + 45 ppm degraded products (0 + 45) were not significantly different

(p>0.05) from corresponding controls containing only 25 ppm and 5 ppm all-trans lycopene, respectively (Table 12). In addition, the peroxide value for the sample containing 0 ppm all-trans lycopene + 50 ppm degraded products (0 + 50) was not significantly different (p>0.05) from the corresponding control containing only oil and 3 ppm chlorophyll [oil + chloro (light)]. These results indicated that lycopene degradation products, as measured by peroxide values, did not act either as prooxidant or antioxidants in the soybean oil model systems, regardless of the presence or absence, respectively, of

148 Milliequivalents Sample* Grouping** HzOz/Kg soybean oil

50 ppm dll-trans 5.83 A

25 ppm dl\-trans 7.83 B

5 ppm all-trans 10.83 C

25 ppm all-trans + 25 ppm degraded 9.17 B

5 ppm all-trans + 45 ppm degraded 10.33 BC 0 ppm all-trans + 50 ppm degraded 13.67 D

oil + chlorophyll (in light) 14.33 D

* Indicates final concentration of lycopene in soybean oil. ** Means with different letters are significantly different (p<0.05).

Table 12: Tukey’s Studentized Range Test for the effects o f lycopene and thermally degraded lycopene combinations on peroxide formation in soybean oil (milliequivalents HzOz/Kg soybean oil) after 24 hours of light exposure.

aï\-trans lycopene. Rather, under light storage the degradation products seemed to be fairly inert. The presence of both light and chlorophyll were crucial to the development of peroxides in the soybean oil samples (Usuki et al., 1984; Rahmani and Csallany,

1998). The control sample containing only soybean oil and no added chlorophyll had a peroxide value of only 2.67 meq HzOz/Kg oil, while the value for the control containing oil and 3 ppm chlorophyll that was stored in the dark was only 1.3 meq HiOi/Kg oil

(Table 12). This data once again pointed towards singlet oxygen as the main oxidizing agent in the soybean oil/lycopene samples.

149 Individual headspace oxygen depletion and peroxide values were correlated in Figure

47 for all samples containing lycopene and/or lycopene degradation products. The resulting regression line showed a linearity of = 0.90, which was slightly lower than the regression line for the degraded P-carotene samples. The trendline revealed that increased peroxide values correlated fairly well with increased headspace oxygen depletion values, which elevated the validity of each separate method in determining the extent o f oxidation for each individual soybean oil/lycopene sample.

6

5 0.367X-0.3117 o 'S ' C w = 0.9022 If 4 3

g o ♦ ♦ 2. « II 2 1

♦ ♦ 0 0 5 10 15 Peroxide value (meq ISaChlKg oil)

Figure 47: Correlation between headspace oxygen depletion and peroxide value for soybean oil containing degraded lycopene under light storage at 25°C.

150 4.2.3 Crude lycopene extract

Though lycopene itself has not been approved as a food additive in the United States, a crude lycopene extract (CLE) firom tomato paste could be used as a food ingredient simply labeled on products as “tomato paste extract”. Determination of the effects of such an extract on soybean oil stability can therefore be very relevant to food processors in the United States. For this reason, in addition to determining the effects of >98% pure lycopene thermal degradation products on the oxidative stability of soybean oil, CLE firom tomato paste was also degraded and added to soybean oil samples to examine its effects on oil oxidative stability.

A pattern similar to results seen for pure lycopene was observed for the CLE all-trans lycopene control samples. The headspace oxygen depletion values for the 50, 25, and 5 ppm CLE all-trans lycopene controls were all significantly different (p<0.05) (Figure

48). The 50 ppm samples (lyc 50 ppm) had the lowest depletion values (2.99 pmoles

Oz/mL), while the 5 ppm samples (lyc 5 ppm) displayed the highest depletion values

(3.78 pmoles Oz/mL), of the three controls, respectively (Table 13). The CLE 50 ppm all-trans lycopene control samples exhibited the greatest ability to quench singlet oxygen and minimize soybean oil oxidation. CLE samples containing mixtures of all-trans lycopene and degradation products (25 + 25 and 5 + 45) had headspace oxygen depletion levels virtually identical to their corresponding 25 ppm and 5 ppm CLE all-trans lycopene control samples, respectively, after 24 hours of storage at 25°C in a light box

(Figure 48). The degraded CLE compounds did not seem to exhibit either prooxidant or antioxidant activity in the soybean oil samples also containing all-trans lycopene.

151 . lyc 50 ppm . lyc 25 . lyc 5 ppm c .25+25 o CQV .5+45 CJeu S* .0+50 ■a 2 . oil+chloro (light) ClS JS X o s ejo <5 Q- S V} o -a E 2 s m

12 16 Time (hours)

Figure 48: Effects of thermally degraded crude lycopene extract on the headspace oxygen depletion of soybean oil containing 3 ppm chlorophyll under light storage at 25°C.

The sample with no remaining dl\-trans lycopene and only containing CLE degradation products (0 + 50) showed headspace oxygen depletion values very similar to those exhibited by the control containing only oil and 3 ppm chlorophyll [oil -t- chlorophyll (light)]. Headspace oxygen depletion values for the two samples measured

4.03 and 4.05 p.moles Oi/mL, respectively (Table 13). The 50 ppm degraded CLE products in the absence of BÏl-trans lycopene consequently did not affect the rate of

152 Depleted pmoles Sample* Grouping** Oz/mL headspace

50 ppm al\-trans 2.99 A 25 ppm zil-trans 3.44 B

5 ppm dh-trans 3.78 C

25 ppm all-trans + 25 ppm degraded 3.39 B

5 ppm all-trans + 45 ppm degraded 3.78 C

0 ppm ahrtrans + 50 ppm degraded 4.03 C

oil + chlorophyll (in light) 4.05 C

* Indicates final concentration of crude lycopene extract in soybean oil. ** Means with different letters are significantly different (p<0.05).

Table 13: Tukey’s Studentized Range Test for the effects of crude lycopene extract and thermally degraded crude lycopene extract combinations on headspace oxygen depletion (pmoles Oz/mL headspace) after 24 hours of light exposure.

headspace oxygen depletion that occurred in the sample. Extracts from tomato paste have been known to sometimes contain, in addition to lycopene, minor amounts of other fat soluble antioxidants such as tocopherols. Tocopherols present in lipid systems have been shown to effectively quench singlet oxygen (Fahrenholtz, 1974; Carlsson et al.,

1976). Because the degraded CLE did not inhibit headspace oxygen depletion relative to controls, tocopherols most likely were either not present at appreciable amounts in the extract, or were degraded to the extent that they could no longer effectively quench singlet oxygen.

153 Control samples, one containing no added chlorophyll and the other stored in the

dark, experienced very little headspace oxygen depletion (0.71 and 0.03 pmoles Oz/mL,

respectively). Singlet oxygen was therefore most likely to have initiated the soybean oil

oxidation responsible for the observed headspace oxygen depletion in these control

samples (Usuki et al., 1984; Rahmani and Csallany, 1998).

The resulting peroxide values for soybean oil samples containing CLE degradation

products and/or dl\-trans lycopene and their controls after 24 hours of storage in a light

box can be seen in Figure 49. Peroxide values decreased by almost 50% as the

. lyc SO ppm . lyc 25 ppm 30 - . lyc 5 ppm .25+25 25 - .5+45 a 0 15 .0+50 > 20

10 -

Tune (hours)

Figure 49: Effects of thermally degraded crude lycopene extract on the peroxide value of soybean oil containing 3 ppm chlorophyll under light storage at 25°C.

154 concentration of zïl-trans lycopene in sample controls increased from 5 ppm to 50 ppm

(28.00 meq HiOz/Kg to 15.67 meq HzOz/Kg, respectively). All-trans lycopene present in the CLE was effective at diminishing the formation of peroxides in the soybean oil.

Peroxide values after 24 hours for the 25 ppm CLE al\-trans lycopene control, at

21.33 meq HzOz/Kg, were not significantly different from those of the sample containing

25 ppm CLE al\-trans lycopene + 25 ppm CLE degradation products, at 20.33 meq

HzOz/Kg (Table 14). The 5 ppm CLE sl\-trans lycopene control, however, had significantly higher peroxide values (28.00 meq HzOz/Kg) than the 5 ppm CLE dl\-trans

Milliequivalents Sample* Grouping** HiOi/Kg soybean oil

50 ppm dl\-trans 15.67 A

25 ppm dhrtrans 21.33 B

5 ppm aX\-trans 28.00 C

25 ppm a!L\-trans + 25 ppm degraded 20.33 B

5 ppm al\-trans + 45 ppm degraded 25.33 D

0 ppm sl\-trans + 50 ppm degraded 31.50 E

oil + chlorophyll (in light) 31.33 E

* Indicates final concentration of crude lycopene extract in soybean oil. ** Means with different letters are significantly different (p<0.05).

Table 14: Tukey’s Studentized Range Test for the effects o f crude lycopene extract and thermally degraded crude lycopene extract combinations on peroxide formation in soybean oil (milliequivalents HzOz/Kg soybean oil) after 24 hours of light exposure.

155 lycopene + 45 ppm CLE degradation products (5 + 45) sample (25.33 meq HzOz/Kg)

(Table 14). The 45 ppm degraded products in this case appeared to act as antioxidants in the soybean oil model system by reducing photosensitized oxidation. As previously mentioned, the CLE may contain other fat soluble antioxidants such as tocopherols.

Even if lycopene degraded products alone could not reduce photosensitized oxidation, the possible presence of tocopherols in the 5 + 4 5 sample may have helped to quench singlet oxygen and reduce its final peroxide value (Fahrenholtz, 1974; Carlsson et al., 1976).

CLE degradation products could not be clearly defined as antioxidants in soybean oil samples, however, due to inconsistencies surrounding their overall effects in light storage. For example, the sample containing 50 ppm CLE degraded products (0 + 50) had a slightly higher peroxide value than the control sample containing only oil and 3 ppm chlorophyll (Table 14). If the degradation products truly acted in an antioxidant fashion, the samples containing the 50 ppm degraded products would have had significantly lower peroxide values when compared with the control.

Singlet oxygen could be named as the primary culprit for most of the oxidation occurring in the sample vials that led to increasing peroxide values over time. Controls stored without added chlorophyll in the light or with chlorophyll in the dark had relatively little peroxide development, which is a strong indicator for the necessity of singlet oxygen as an initiator of soybean oil oxidation.

Figure 50 reveals the linearity of a regression line drawn between the headspace oxygen depletion values for CLE samples and their corresponding peroxide values. The data formed a trendline with an value of 0.93. The data points were more linear than those of the >98% pure lycopene samples mainly due to a general lack of outliers in the

156 5

4 y = 0.1477X - 0.1933 o 'S' •■S W = 0.929 If 3 II 2 t l I

0 0 5 10 15 20 25 30 35 Peroxide value (meq HbOi/Kg oil)

Figure 50: Correlation between headspace oxygen depletion and peroxide value for soybean oil containing degraded crude lycopene extract under light storage at 25°C.

CLE sample results. The trendline showed good correlation between the headspace oxygen results and the peroxide value results for the CLE samples.

In summary, the effects of carotene thermal degradation products on the oxidative stability of soybean oil stored under the light differed depending on the specific carotene and the relative concentrations of al\-trans vs. degraded compounds. (3-carotene degradation products tended to exhibit a prooxidant effect at concentrations of 50 ppm.

When measuring headspace oxygen depletion this effect was not significant; however,

157 the prooxidant effect was significant (p<0.05) for the peroxide value data. For the most

part, both lycopene and crude lycopene extract degradation products did not elicit either a

prooxidant or antioxidant response in the soybean oil samples stored under the light, but

rather existed in the oil as relatively inert ingredients. These results suggested that P-

carotene may decrease soybean oil oxidative stability if it has been degraded due to

thermal processing. Thermally degraded lycopene, when stored under the light, should

not increase or decrease the oxidative stability of soybean oil or similar products

containing soybean oil.

4.3 Effects of degraded carotenes on thermally-induced oxidation of soybean oil

In addition to photosensitized oxidation that requires light and sensitizers like

chlorophyll, unsaturated lipid systems are also prone to autoxidation initiated by free

radicals, especially at elevated temperatures (Burton and Ingold, 1984; Frankel, 1989).

Therefore, degraded carotenes present in soybean oil may have a profound effect on the

stability of a lipid system exposed to such conditions.

The effects of P-carotene thermal degradation products on the oxidative stabihty of

soybean oil stored in the dark at 60°C can be seen in Figure 51. After 8 days o f storage, headspace oxygen depletion occurred in each of the three samples, indicating the presence of free radical autoxidation and peroxide formation within the soybean oil. The sample containing 50 ppm degraded P-carotene displayed significantly more headspace oxygen depletion (5.79 pmoles Oi/mL) than either the oil only control (5.58 pmoles

Oz/mL) or the 50 ppm dïi-trans P-carotene control (5.33 pmoles Oz/mL) (p<0.005). All

158 g 50 ppm unheated B-car g 50 ppm degraded B-car □ oil control

e o CJ w o 93 CL c . va • o 93 O S u ex) B o

s 6 ya CL % "O B 3 X

Figure 51 : Effects of thermally degraded P-carotene on the headspace oxygen depletion of soybean oil stored in the dark at 60°C.

-trans P-carotene has been reported to act as a free radical inhibitor in the absence of light

(Burton, 1989). In addition, because there was no visible loss o f color in the samples containing 50 ppm al\-trans P-carotene during the 8 days o f storage at 60°C, it could be assumed that most of the P-carotene did not imdergo significant polyene degradation. It was therefore not surprising that the samples containing dX\-trans p-carotene experienced the least headspace depletion. The obvious prooxidant effect o f the 50 ppm degraded P- carotene compounds in the soybean oil was not expected, however, based on headspace

159 oxygen depletion values obtained from photosensitized oxidation data that clearly showed degraded P-carotene as having no significant effect vs. controls. The high level of significance obtained between the samples containing 50 ppm degraded p-carotene and the oil-only control (p<0.005) strengthened the argument that in dark conditions at elevated temperatures P-carotene degradation products increase the rate of autoxidation in soybean oil.

Lycopene degradation products were also examined for their effects on the oxidative stability of heated soybean oil in the dark. Results were very difierent from those observed with P-carotene degradation products. Samples containing 50 ppm degraded lycopene products had the lowest headspace oxygen depletion values of the three samples

(5.34 pmoles OJvaL), which also included a 50 ppm all-tram lycopene control (5.36 pmoles Oz/mL) and an oil-only control (5.54 pmoles Oi/mL). Though the 50 ppm degraded products samples were not significantly different in headspace oxygen depletion values than the samples containing 50 ppm aH-trans lycopene, both were significantly lower (p<0.05) than that of the oil-only controls (Figure 52). At levels of 50 ppm, the lycopene thermal degradation products exhibited a definite antioxidant effect in the soybean oil samples when stored in the dark at elevated temperatures. Miller et al.

(1996) similarly reported lycopene as the most effective scavenger of free radical cations in a nonpolar solvent system when compared with (3-carotene and various xanthophylls.

Lycopene degradation products in soybean oil exposed to light were found to be largely inert, so again the results seen here were very dififerent from previous observations with photosensitized oxidation. Verified by HPLC analysis, the polyene

160 g 50 ppm unheated tyc

g 50 ppm degraded lyc

□ oil control • iî i f I

II

Figure 52: Effects of thermally degraded lycopene on the headspace oxygen depletion of soybean oil stored in the dark at 60°C.

system in z\i-trans lycopene had been substantially degraded by thermal treatments during the formation of the thermal degradation products. The compoimds created by the thermal degradation must have had some type of free radical scavenging ability that was approximately equal to that of the dl\-trans compound.

The free radical trapping ability of crude lycopene extract (CLE) degradation compounds was also investigated. As seen in Figure 53, samples containing 50 ppm degraded lycopene exhibited significantly less (p<0.05) headspace oxygen depletion after

8 days of storage than the oil-only control (4.67 pmoles Oz/mL vs. 4.85 pmoles Oz/mL,

161 5.0 . g 50 ppm degraded lyc extract □ oil control

Figiore 53 : Effects of thermally degraded crude lycopene extract on the headspace oxygen depletion of soybean oil stored in the dark at 60°C.

respectively). Similar to what was previously described with the degradation products

from the 98% pure ZL^l-trans lycopene source, a definite antioxidant effect was exerted on the soybean oil samples by the CLE degradation products. Al\-trans lycopene has been shown to effectively scavenge free radicals during autoxidative processes (Miller et al.,

1996), but little has been reported in the literature on the effects of degraded lycopene on oxidative stability. Crude lycopene extracts may contain traces of other compounds with antioxidant properties such as tocopherols, which have been described as excellent chain

162 breaking antioxidants (Burton and Ingold, 1981; Di Mascio et al., 1991; Palozza and

Krinsky, 1992; Heinonen et al., 1997). Therefore, the possible presence of tocopherols in

the CLE may have helped increase the antioxidant properties of samples containing CLE

degradation products. As a whole, the CLE degradation products effectively reduced

peroxide formation and increased the oxidative stability o f the soybean oil.

In summary, carotene degradation products behaved very differently in lipid systems

exposed to autoxidation through temperature abuse. P-carotene degradation products

acted as prooxidants by significantly decreasing soybean oil oxidative stability. Just as

importantly, lycopene degradation products had an antioxidant effect in soybean oil

systems by significantly decreasing headspace oxygen depletion vs. controls. Similarly,

CLE degradation products also significantly increased the oxidative stability of soybean

oil. These results seemed to indicate that compounds not capable of quenching singlet

oxygen and preventing photosensitized oxidation might still be very effective as peroxy

radical scavengers during autoxidative processes. The type of carotenoid, as well as the

conditions surrounding the oil sample, played a large role in the determination of how the

degradation products, when formed, affected the oxidative stability of the lipid system.

4.4 Separation and identification of soybean oil volatile oxidation products by SPME

In order to further understand the effects of degraded carotenes on the oxidative

stability o f soybean oil, a solid phase microextraction (SPME) method was developed to determine, both quantitatively and qualitatively, the volatile compounds present in soybean oil stored under various conditions. The first step in the development of the

SPME method was to determine which stationary phase would be the most sensitive,

163 reproducible, and durable for repeated adsorption of compounds present in the headspace

of heated soybean oil/acetone model systems. Different stationary phases have been

shown to work more effectively in certain applications, depending on the relative polarity

of both the volatile compounds expected and the stationary phase fiber itself (Chin et al.,

1996; Steffen and Pawliszyn, 1996; Jelen et al., 1998; Song et al., 1998). The types of

volatile compounds typically found in oxidized soybean oil have been documented in the

literature as being 5 to 10 carbon saturated, monounsaturated, and polyunsaturated

aldehydes, alcohols, and ketones, as well as cyclic compounds such as furans (Frankel,

1985). Therefore, a known concentration of hexanal, a typical volatile compound foimd

in soybean oil, was inoculated into sealed soybean oil/acetone model systems. Four different SPME fibers were then exposed to the headspace of the samples for 20 minutes, and immediately injected into a GO. The mean hexanal peak size of triplicate runs for each of the four fibers can be seen in Figure 54.

The carboxen/polydimethylsiloxane (CAR/PDMS) fiber showed the lowest sensitivity to the hexanal, with a peak size that was 40% smaller than the next largest peak. Polyacrylate, a relatively polar stationary phase, was significantly lower in peak size than the fiber with the next largest peak size. The two stationary phases with the highest relative sensitivity to hexanal in a soybean oil/acetone model system were polydimethylsiloxane (POMS) and Carbowax®/divinylbenzene (CW/DVB). The

CW/DVB fiber displayed a greater affinity for hexanal, due to its larger peak size, when compared with the PDMS fiber. These results agree with the findings of Song et al.

(1998), who reported the hexanal partition coefficients (K) of the PDMS and CW/DVB stationary phases to be 8.64 x 10^ and 2.66 xlO \ respectively. Though the CW/DVB

164 2.0E+06 ^

g CAR/PDMS 1.6E+06 - g Polyaciylate

=) □ PDMS 1.2E+06 - g CW/DVB .Ë% 2 G. a 8.0E+05 S s

4.0E+05 -

O.OE+00

Figure 54: Relative sensitivity of four diffèrent SPME fibers to hexanal in the headspace of a soybean oil/acetone model system.

fiber was the most sensitive to hexanal in the headspace, its reproducibility was not as good as the PDMS fiber, with a standard deviation that was more than twice as large as the PDMS fiber. In addition, the durability of the CW/DVB fiber stationary phase was not acceptable due to its tendency to chip off of the fused silica fiber after only a few adsorption/desorption cycles. Comparatively, the PDMS fiber had the lowest standard

165 deviation of the four fibers tested, the second highest sensitivity to hexanal, and the highest durability of the four stationary phases. Therefore, the PDMS stationary phase was chosen for use in all further SMPE experimental analyses.

Once the best fiber stationary phase was chosen, analyses involving the separation and identification of soybean oil volatile oxidation products by SPME-GC-MS could be conducted. A chromatogram showing the various peaks and relative sizes of volatile compounds adsorbed from the headspace of an oxidized soybean oil sample can be seen in Figure 55. All of the volatile compounds that were identified eluted within the first 20 minutes of the timed run. The use o f a 30 m capillary column as well as a narrow-bore inlet liner improved the sharpness o f many of the peaks, which in turn allowed easier peak scans by the mass spectrometer. The relative abundances of the compounds varied greatly, with f,c-2,4-decadienal (peak 16) being the most abundant compoimd identified, with 19.6% o f the total volatile peak area (Figure 55, Table 15). Both the trans and cis

2,4-decadienals have often been reported as the volatiles present in the greatest quantities

b . BE. t-G

II 14 ^ 3. an i-B 3 2 . a c t- G C l.Q E -t-B 0. 0E+0 IB IS 20 T I n e C m1n . )

Figure 55: SPME-GC-MS chromatogram of volatiles in oxidized soybean oil (see Table 15 for complete numerical listing of volatiles).

166 Peak number Retention time Relative Compound (Figure 54) (minutes) %

1 pentane 1.154 11.3

2 pentanal 2.060 3.9

3 hexanal 3.757 12.4

4 2-butanone 5.992 2.9

5 heptanal 7.560 2.6

6 2-heptenal 9.323 4.2

7 2-pentyl-fliran 10.323 2.9

8 2,4-heptadienal 10.495 1.4

9 octanal 10.648 2.0

10 r-2-octenal 12.044 4.9

11 nonanal 13.128 4.6

12 t-2-nonenal 14.326 1.5

13 f,f-nona-2,4-dienal 15.442 0.9

14 2-decenal 16.363 6.0

15 /,r-2,4-decadienal 16.959 9.9

16 f,c-2,4-decadienal 17.382 19.6

17 2-undecenal 18.219 5.9

18 cyclododecane 18.681 1.4

19 dodecanal 18.988 1.7

Table 15: Volatile compounds detected in the headspace o f oxidized soybean oil by SPME-GC-MS.

167 in oxidized soybean oil (Dupuy et al., 1985; Selke and Frankel, 1987). Linoleate

hydroperoxides have been described as the precursors for the formation of 2,4-

decadienals in oxidized soybean oil (Selke and Frankel, 1987). Because linoleic acid is

by far the most prevalent fatty acid found in soybean oil (« 55%), it was not surprising to

report 2,4-decadienal as a the most predominant volatile oxidation product of soybean oil.

Hexanal (peak 3) was the second most common volatile compoimd adsorbed onto the

PDMS fiber based on total peak area, at 12.4% of the total volatile compound area (Table

15). Due to its prevalence in oxidized oils, hexanal is often used as a marker for the extent of oxidation in lipid systems (Heinonen et al., 1997; Andersson and Lingnert,

1998). Pentane (peak 1) was the first identified compound to elute (retention time =

1.154 minutes), as well as the third most abundant volatile, with 11.3% of the total identified volatile peak area.

In all, 19 different oxidation products were identified by GC-MS (Table 15). When profiling the volatile compounds that were identifiable versus those reported in the literature, it became clear that there was an observable bias towards larger, more non­ polar compounds over more polar, smaller molecular weight volatiles. For example, small polar compounds reported to be in oxidized soybean oil such as propanal and butanal (Frankel et al., 1991; Frankel et al., 1993) were absent. At the same time, larger compounds over ten carbons in length that have not been widely reported in the literature such as 2-undecenal (peak 17), cyclododecane (peak 18), and dodecanal (peak 19) were tentatively identified by GC-MS. The mass spectrum o f the 2-undecenal peak (top) compared with a library spectrum of 2-undecenal (bottom) can be seen in Figure 56. The

Wiley Mass Spectrum Library assigned an 83% match to this peak that eluted at 18.228

168 SS300-5 3 SOS ij i G7 8 200i I 7a a a l ■ G 0 0 0 ^ 5 see I 4 690- 2000- 208 III [21 17? 228 I <1-3 1 000 ' if III nil 1 . I / X \ 0- J i l fO 'Z 5 2-Undecen a I C is s a s i 3 / |70 8000^ 7 000^

27 50001 ^ 4305^-' 2sse^[ 230031 11 i'SeS^i lUt y iL - Li I I / 153 :38 Mass^Chara e

Figure 56. SPME-GC mass spectrum o f peak 17 (retention time 18.219 min.) (top) and 2-undecenal (bottom).

minutes and the standard 2-undecenal mass spectrum. Each spectrum displayed matching

fragment peaks at MW 121, 111, 97, 95, 83, 70, 57, and 56. The next identifiable

compound that eluted at 18.681 minutes was cyclododecane, a cyclic saturated

hydrocarbon that also has not been widely reported in the literature. The mass spectrum of the cyclododecane peak (top) compared with the library spectrum of cyclododecane

169 (bottom) very well, with the Whey Mass Spectrum Library assigning a 91% match to the

two mass spectra (Figure 57). Similar fragment peaks were observed at MW 140,111,

95, 83, 69, and 55. The last volatile compound to be tentatively identified by GC-MS

from the oxidized soybean oil headspace was dodecanal, a twelve carbon aldehyde not

commonly reported as an oxidation product of soybean oil. The mass spectra o f this peak

!0S28q ■3 338^ 4 1 300Ü I I, 69 ! I 7300 1 .1 S3 buuai 5 8 9 0 1 4 0 0 8 1 .ZaK/», i I 11: 0888 "J / 153 / / / !! [LI III ! ! Id“Iliu m . LiÉ.! jLJL iiÎ /a “O Cvclododecaoe 10308-] 89 GS9 g| 5^1 / ( 8880 j I I 7000 ] saaal 50891 4 000-^ III / 3008-^ I 1S9 2 028 i Î4! \ i 0 0 0 \ I I I - t - r - a — 83 ISS 120 142 1SÎ ISÎ Mas-'v/Chara e

Figure 57. SPME-GC mass spectrum of peak 18 (retention time 18.681 min.) (top) and cyclododecane (bottom).

170 at 18.988 minutes received a 72% match with the Wiley library spectrum for dodecanal.

The two spectra can be viewed in Figure 58. Common peak fragments in the mass spectra included MW 140, 110, 82, 75, and 57.

Because a very non-polar stationary phase (PDMS) was used for the SPME fibers in the GC-MS analyses, non-polar compounds such as 2-undecenal and dodecanal were

13832^ 4 I c;? a

Dodecan a I 10038

QSC2^ / S 8 3 8 ] 70001 60081 82 50001 / 4900^ 3203| c 880 j I il Î IS 140 11 1 II I u < Il j / 1 i C.T I ! i / .11 1 . i i . j i . ■ t » 1 ■-■■1 t- i -r-r-T-r BS 103 122 149 183 223 M&sz/Chargc

Figure 58. SPME-GC mass spectrum o f peak 19 (retention time 18.988 min.) (top) and dodecanal (bottom).

171 probably adsorbed onto the fiber more readily than relatively polar compounds such as propanal and butanal (Chin et al., 1996). This could explain the apparent bias against short chain aldehydes in favor of longer chain aldehydes in the SPME-GC chromatograms. Though a bias based on stationary phase polarity may exist, as a qualitative method for determining volatiles in soybean oil, SPME-GC seems to work very well overall. Many of the GC peaks were positively identified so that quantitation of only known soybean oil volatile oxidation products could be conducted more accurately when assessing the effects of degraded carotenes on the oxidation o f soybean oil and resulting volatile compound formation.

4.5 Effects o f degraded carotenes on formation of volatile oxidation products in soybean oil

In addition to headspace oxygen depletion and peroxide value measurement, solid phase microextraction (SPME) was utilized as a method to help determine how degraded carotenes might affect the oxidative stability of soybean oil exposed to extended storage in either a light box at 25°C or in a dark oven at 60°C. The SPME polydimethylsiloxane

(PDMS) fiber was exposed to the headspace of a sample vial containing soybean oil and either 50 ppm of all-frnw P-carotene/lycopene, 50 ppm of degraded carotene, or only soybean oil. The total area for all the GC chromatogram peaks identified as an oil oxidation product was calculated for each sample. The total peak areas of soybean oil/p- carotene samples stored in a light box from 0 to 30 days were compared on a percentage basis with the total peak area determined for the oil only control (Figure 59).

172 g B-car 50 ppm 140 g 0 + 50

120

100 2 c wo

OC o CmO

0 days 5 days 10 days 20 days 30 days

Figure 59: Percentage comparison o f total GC peak area from volatiles formed in oil samples containing P-carotene or P-carotene degradation products after storage in a light box at 25°C (oil only control = 100%).

As expected, at 0 time (no storage treatment) there was no significant difference in the total volatiles area for the two samples as a percentage of the oil only control. The total volatiles area for the soybean oil samples containing either 50 ppm all-traw p- carotene (B-car 50 ppm) or 50 ppm degraded products (0 + 50) were within 1.5% of the oil only control. Similar results were observed after 5 days of storage in the light box

173 (Figure 59), in which the samples containing 50 ppm a!L\-trans (3-carotene as well as those

with 50 ppm degraded products had total peak areas within 2% of the oil only control.

After 10 days o f storage in a light box (1,650 lumens) at 25°C, the difference in total

volatile peak areas for the three samples began to widen, though not significantly. The

soybean oil samples containing 50 ppm al\-trans (3-carotene had total peak areas that

averaged 93.9% of the oil only control peak area. The samples containing 50 ppm

degraded products also revealed total peak areas that were less than the oil only control,

at 96.2%.

The al\-trans (3-carotene-containing samples contained only 89.3% of the total

volatiles GC peak area of the oil only control samples after 20 days of storage in the light

box. This result indicated a possible antioxidant effect o f P-carotene in the soybean oil,

probably as a singlet oxygen quencher due to the presence of light and chlorophyll in the

model system. Warner and Frankel (1987) also reported that 20 ppm ai\-trans p-carotene

reduced the size of GC volatile peaks formed in soybean oil that was stored under the

light for 24 hours. Samples containing degraded P-carotene revealed an average total

volatile peak area that was slightly higher (2.2%) than that of the oil only control after 20

days of storage. As indicated in previous studies conducted in this lab utilizing degraded

carotenes in light storage, the degradation products did not seem to retain the singlet

oxygen quenching capacity of the zl\-trans form, therefore allowing increased oxidation

in the oil system.

Results after 30 days of light box storage were similar to that of the 20 day samples, though with less difference between sample types (Figure 59). Samples containing all-

174 trans P-carotene had an average total headspace volatiles peak count that was 6.7% lower

than the oil only control. The 50 ppm degraded P-carotene samples had a slightly higher

total peak area than the oil only control (0.4%), however. Again, the sll-trans P-carotene

seemed to exhibit an antioxidant effect in the soybean oil that reduced the quantity of

volatile compounds being produced through oxidative processes. The degraded products

did not reduce the total volatiles produced under light storage. The differences in total peak areas for the three samples may have been greater after 20 days than after 30 days because of the concentration of oxygen present in the headspace of the samples. Though it has been reported that volatiles may continue to be created after all of the headspace oxygen has been consumed in a vial (Min and Wen, 1983), lower oxygen concentrations may have decreased the rate of volatile compound formation significantly.

Though the samples containing p-carotene did have reduced volatile peak areas when compared with the oil only control after 10, 20, and 30 days of storage in a light box, the difference was not significant mainly due to the relatively large standard deviations seen for some of the samples. SPME analyses have had reported problems with reproducibility (Ng et al., 1996), with some of the sources of variability including lot to lot variation in fibers (Arthur and Pawliszyn, 1990), temperature sensitivity (Yang and

Peppard, 1994), and the position of fibers in the injector during desorption (Arthur et al.,

1992; Arthur and Pawliszyn, 1990). Yang and Peppard (1994) stated that SPME fibers were extremely sensitive to experimental conditions, and that “any change o f experimental parameters, which affect the distribution coefficient and adsorption rate, will also influence the amount adsorbed on the SPME fiber and the corresponding

175 reproducibility.” The low distribution coefficient for lipid systems and the headspace around them further complicated matters by greatly decreasing adsorption of volatile compounds, which in turn decreased the total peak area of the volatile compounds.

Results showed that smaller total peak areas were subject to larger variability than the larger total peak areas.

SPME was also utilized as a method to help determine how degraded P-carotene might affect the oxidative stability of soybean oil exposed to extended storage in a dark oven at 60°C. As shown in Figure 60, at 0 days the total peak areas for volatiles firom the samples containing either 50 ppm al\-trans p-carotene or 50 ppm degraded P-carotene were within 2% of the average value for the oil only control. After 5 days of oven storage, both sample types (B-car 50 ppm and 0 + 50) had 102.3% and 105.4% o f the total volatile peak area of the oil only control, respectively. Five days of storage was apparently not enough time to see any antioxidant/prooxidant effects firom carotenes in the oil as measured by total volatile peak area.

The results after 10 days of storage at 60°C in a dark oven were similar to the 5 day results (Figure 60), except that the total peak area for the 50 ppm dl\-trans P-carotene samples was slightly lower (1.1%) that the oil only control. The total peak area for the samples containing 50 ppm degraded p-carotene were again slightly higher than the oil only control, though not significantly. The oil only control for the 10 day samples had a very large standard deviation mainly due to the lack of integration o f several large volatile peaks on one of the GC runs. After 20 days of storage at 60°C in the dark, the total peak area of the volatile compounds being formed in samples containing 50 ppm all-

176 g B-car 50 ppm 160 ^ * 0 + 50 □ oil only

2 s wo a o

o

0 days 5 days 10 days 20 days 30 days

Figure 60: Percentage comparison of total GC peak area from volatiles formed in oil samples containing P-carotene or P-carotene degradation products after storage in an oven at 60°C (oil only control = 100%).

trans P-carotene was only 88.1% of that of the oil only control, while the solutions of degraded carotenes had higher relative volatile peak areas, averaging 110.8% of the oil only control. This pattern was repeated in the samples stored for 30 days at 60°C. The samples containing dl\-trans P-carotene had a significantly lower (p<0.05) total peak area

177 than the samples containing degraded p-carotene (Figure 60). This indicated that the all-

trans P-carotene may have helped prevent oxidized volatiles from being formed by acting

as a chain-breaking antioxidant (Burton, 1989; Miller et al., 1996). Conversely, the

degraded carotenes may have acted as prooxidants in the model systems and increased

the total volatiles peak area to 108.1% of that of the oil only control. These results

agreed nicely with those of the headspace oxygen depletion values for the zï\-trans p-

carotene samples vs. degraded P-carotene samples as seen in Figure 51.

Al\-trans lycopene and degraded lycopene were added to soybean oil at 50 ppm in a manner similar to P-carotene in order to determine the effects of these compounds on oil oxidation and subsequent volatile compound formation. At zero time, peak sizes for both aJil-trans and degraded lycopene samples were 1.7% higher than that o f the oil only control (Figure 61). After 5 days of storage in a light box, the average total peak area for the oü only control had risen slightly higher than the areas for the dl\-trans and degraded lycopene samples. Figure 61 revealed that after 10 days of light storage, the zi\-trans lycopene samples had a total volatiles peak area that was only 87.2% o f the oil only control, while the samples containing degraded lycopene displayed peak areas that were

92.9% of the oil only control. The disparity between the three sample peak areas increased further upon 20 days of light storage, in which the average volatile peak area for the al\-trans lycopene samples was only 75.7% that of the oil only control, and the degraded lycopene samples were 84.1%, respectively. Results for 30 days of light box storage were similar, at 89.2% and 94.2% of the oil only control for the al\-trans and degraded samples, respectively.

178 g Lycopene 50 ppm 160 . g 0 + 50 □ oil only 140

B 80 r o 60

0 days 5 days 10 days 20 days 30 days

Figure 61 : Percentage comparison o f total GC peak area from volatiles formed in oil samples containing lycopene or lycopene degradation products after storage in a light box at 25°C (oil only control = 100%).

Soybean oil alone has been reported to be very sensitive to light-induced oxidation

(Warner et al., 1989; Lee et al., 1995). Lycopene, however, has been shown to reduce the extent of light-induced oxidation through singlet oxygen quenching mechanisms (Di

Mascio et al., 1989; Edge et al., 1997). The lycopene present in the samples exposed to light for 20 days or more seemed to exert an antioxidant effect by reducing the total

179 volatile peak area, though not significantly. These results agreed with previous experiments measuring headspace oxygen depletion (Figure 45) and peroxide value

(Figure 46), which showed a significant reduction in oxidation in samples containing 50 ppm al\-trans lycopene. Due to large standard deviations in the SPME analyses, however, once again the observed differences in peak areas were not significant.

The effects of 50 ppm al\-trans lycopene as well as 50 ppm degraded lycopene on the oxidative stability of soybean oil (as measured by total volatile peak area) was next determined for samples stored in a dark oven at 60°C from 0 to 30 days (Figure 62).

Relative to the oil only control, at 0 time the total peak areas for both the samples containing dl\-trans lycopene and degraded lycopene were slightly higher, at 102.4% and

103.3%, respectively. At 0 time the total peak areas were expected to be approximately equal, which was the case for the means of the three sample groups. However, large deviations from the mean for two o f the groups were observed (Figure 62).

After 5 days o f storage in the oven, both the samples containing dl\-tram lycopene and those containing degraded lycopene had total volatile peak areas that were lower, though still within 5%, of the oil only control. As the storage time was prolonged to 10 days, the samples containing either aX\-trans lycopene or degraded lycopene revealed, on average, lower overall volatile peak areas, at only 87.8% and 91.4% of the oil only control peak area, respectively (Figure 62). Similar results were observed after storing the samples for 20 days at 60°C in the dark, in which the aUl-trans lycopene and degraded lycopene samples displayed total volatile peak areas that were 89.9% and 97.2%, respectively, of the oil only control’s total volatile peak area. No significant differences

(p>0.05) were detected between the three sample groups after 20 days of storage.

180 I Lycopene 50 ppm 160 . , 0 + 50 □ cil only

s uo

OS ’o «M o

0 days 5 days 10 days 20 days 30 days

Figure 62: Percentage comparison of total GC peak area from volatiles formed in oil samples containing lycopene or lycopene degradation products after storage in an oven at 60°C (oil only control = 100%).

After 30 days o f storage, the samples containing degraded lycopene displayed a significantly lower (p<0.05) total volatile peak area when compared with the oil only control (Figure 62). Possessing only 82.3% of the peak area o f the oil only controls, the volatile peak area for the samples containing degraded lycopene were even slightly lower that of the dl\-trans lycopene samples, which were 83.3% of the peak area of the oil only

181 controls. PA\-trans lycopene has been reported to be more effective than P-carotene at

scavenging radical cations that may promote autoxidation in oils exposed to elevated

temperatures (Miller et al., 1996). In addition, Snyder et al. (1985) noted that heating

soybean oil sans antioxidants at 60°C for more than 10 days significantly increased both

the number and size of observed volatile oxidation products. It was therefore not

surprising to see reduced volatile peak sizes in samples containing 3l\-trans lycopene

when compared with the oil only controls. What was imexpected was the apparent ability

of degraded lycopene products to prevent oxidation and volatiles formation by

scavenging radicals in a manner at least as effective as the aïL-trans form. These results

did agree with previous experiments that also indicated an antioxidant effect of degraded

lycopene in soybean oil systems exposed to elevated temperatures over a prolonged

period in the dark (Figure 52).

In summary, SPME analyses were performed on soybean oil samples containing all-

trans or degraded P-carotene or lycopene that were stored under light or dark conditions.

Reproducibility of sample results was lacking at times due to the many variables

surrounding SPME methods, and obtaining significant differences for total volatile peak

areas between the different samples was therefore difficult. However, interesting trends were still observed. As was the case with previous experiments with headspace oxygen depletion and oven storage of samples, carotene degradation products behaved very differently in lipid systems exposed to autoxidation through elevated temperatures

(60°C). P-carotene degradation products tended to act in a prooxidant manner that increased the total volatile peak areas relative to the oil only controls. Lycopene

182 degraded products, on the other hand, acted as antioxidants that resulted in significantly decreased overall volatile peak areas vs. oil only control peak areas. The effects of degraded (3-carotene or lycopene in soybean oil samples stored in a light box were not as dramatic, with no major significant effects or trends observed.

183 CHAPTERS

CONCLUSIONS

From the results of the various research studies, the following conclusions can be

drawn:

1 . Heating ail-trans (3-carotene and dl\-trans lycopene solubilized in acetone at 90°C

resulted in a rate of degradation that was fairly linear as measured by decreases in

the HPLC alX-trans peak size.

2. Crude lycopene extract solubilized in methyl-Zerr-butyl-ether and heated at 90°C

degraded at a rate more than 19x slower than that of pure dl\-trans lycopene.

3. All-trans (3-carotene and all-/ra«r lycopene, at concentrations from 5 ppm to 50

ppm, were effective at reducing headspace oxygen depletion and peroxide values in

soybean oil stored in a light box. This indicated that both oüï-trans carotenes were

able to effectively quench singlet oxygen.

184 4. Degraded (3-carotene and degraded lycopene compounds at concentrations from 25

ppm to 50 ppm had no significant effect on soybean oil oxidative stability as

measured by headspace oxygen depletion and peroxide value when stored under the

light.

5. Controls either stored in the dark with 3 ppm chlorophyll or stored in the light with

no added chlorophyll displayed significantly lower headspace oxygen depletion and

peroxide values, which indicated singlet oxygen was responsible for soybean oil

oxidation under the light.

6 . A linear correlation existed between headspace oxygen depletion values and

peroxide values for soybean oil containing either all-P-mrj/degraded (3-carotene (R^

= 0.96) or all-traWdegraded lycopene (R^ = 0.90), respectively.

7. Al\-trans P-carotene and dll-trans lycopene (50 ppm) were effective at reducing

headspace oxygen depletion in soybean oil samples stored at 60°C for 8 days in a

dark oven. This indicated the carotenes possessed free-radical trapping ability.

8 . Soybean oil samples containing 50 ppm degraded P-carotene exhibited significantly

higher (p<0.05) headspace oxygen depletion than oil only controls after 8 days of

storage in the dark at 60°C, which suggested degraded p-carotene compounds act as

prooxidants in oil stored at elevated temperatures.

185 9. Both lycopene and crude lycopene extract degradation products showed an

antioxidant effect in soybean oil samples stored in the dark for 8 days at 60°C by

significantly decreasing headspace oxygen depletion vs. oil only controls. These

degradation products may therefore possess free-radical trapping ability.

10. SPME-GC-MS was effective at isolating, separating, and identifying volatile

compounds typically found in oxidized soybean oil, though the non-polar PDMS

fiber used showed bias for larger less polar aldehydes over smaller aldehydes.

1 1 . Soybean oil samples containing degraded P-carotene or degraded lycopene (50

ppm) and stored in a light box for 30 days showed no significant difference in total

volatile peak area by SPME-GC when compared with ail-trans P-carotene or

lycopene controls, respectively.

12. After 30 days of storage in a dark oven at 60°C, soybean oil samples containing 50

ppm degraded P-carotene had a significantly higher (p<0.05) total volatile peak area

by SPME-GC when compared with dM-trans P-carotene controls.

13. SPME-GC revealed a significantly lower total volatile peak area for soybean oil

samples containing 50 ppm degraded lycopene compounds stored for 30 days in the

dark at 60°C when compared with oil only controls.

186 LIST OF REFERENCES

Andersson, K., and Lingnert, H. 1998. Influence of oxygen and copper concentration on lipid oxidation in rapeseed oil. J. Am. Oil Chem. Soc. 75: 1041-1046.

AOCS. Official Method Cd 8-53. In Official and Tentative Methods o f the American Oil Chemists’ Society, vol 1. American Oil Chemists’ Society, Champaign, IL, 1980, pp. Cd 8-53.

Arthur, C. L., Killam, L. M., Buchholz, K. D., Pawliszyn, J., and Berg, J. R. 1992. Automation and optimization of solid-phase microextraction. Anal Chem. 64: 1960- 1966.

Arthur, C. L., and Pawliszyn, J. 1990. Solid phase microextraction with thermal desorption using fused silica optical fibers. Anal Chem. 62: 2145-2148.

Bast, A., van der Plas, R. M., van den Berg, H., and Haenen, G. R. M. M. 1996. Beta- carotene as antioxidant. Eur. J. Clin. Nutr. 50 (Suppl. 3): S54-S56.

Ben-Aziz, A., Britton, G., and Goodwin, T. W. 1973. Carotene epoxides of Lycopersicon Esculentum. Phytochem. 12: 2759-2764.

Bendich, A. Biological functions of dietary carotenoids. In Carotenoids in Human Health, vol. 691, Canfield, L. M., Krinsky, N. I., and Olsen, J. A., eds. New York: Annals o f New York Academy of Sciences, 1993, pp. 61-67.

Bernhard, K., and Grosjean, M. Infrared spectroscopy. In Carotenoids, Volume IB: Spectroscopy. Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 117-134.

Bicchi, C. P., Panero, O. M., Pellegrino, G. M., and Vanni, A. C. 1997. Characterization o f roasted coffee and coffee beverages by solid phase microextraction-gas chromatography and principal component analysis. J. Agric. Food Chem. 45: 4680-4686.

187 Brekke, O. L. Soybean oil food products—their preparation and uses. In Handbook o f Soy Oil Processing and Utilization, Erickson, D. R., Pryde, E. H., Brekke, O. L., Mounts, T. L., and Falb, R. A., eds. Champaign: American Oil Chemists’ Society, 1980, pp. 383- 437.

Britton, G. UV/Visible spectroscopy. In Carotenoids, Volume IB: Spectroscopy. Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 13-62.

Britton, G., and Goodwin, T. W. 1969. The occurrence o f phytoene 1,2-oxide and related carotenoids in tomatoes. Phytochem. 8 : 2257-2258.

Britton, G., Liaaen-Jensen, S., and Pfander, H. Carotenoids today and challenges for the future. In Carotenoids, Volume lA: Isolation and Analysis, Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 13-26.

Budowski, P., and Bondi, A. I960. Autoxidation of carotene and vitamin A. Influence of fat and antioxidants. Arch. Biochem. Biophys. 89: 66-73.

Burton, G. W. 1989. Antioxidant action of carotenoids. J. Nutr. 119: 109-111.

Burton, G. W., Ingold, K. U. 1981. Autoxidation of biological molecules. 1. The antioxidant activity of vitamin E and related chain-breaking phenolic antioxidants in vitro. J. Am. Chem. Soc. 103: 6472-6477.

Burton, G. W., and Ingold, K. U. 1984. (3-carotene: an unusual type of lipid antioxidant. Science. 224: 569-573.

Carlsson, D. J., Suprunchuk, T., and Wiles, D. M. 1976. Photooxidation of unsaturated oils: effects of singlet oxygen quenchers. J. Am. Chem. Soc. 53: 656-660.

Camevale, J., Cole, E. R., and Crank, G. 1979. Fluorescent light catalyzed autoxidation of (3-carotene. J. Agric. Food Chem. 27: 462-463.

Cerutti, P. A. 1985. Prooxidant states and tumor promotion. Science. 227: 375-381.

Chandler, L. A., and Schwartz, S. J. 1988. Isomerization and losses of iraw-P-carotene in sweet potatoes as affected by processing treatments. J. Agric. Food Chem. 36: 129-133.

Charleux, J.-L. 1996. Beta-carotene, vitamin C, and vitamin E: the protective micronutrients. Nutr. Reviews. 54: S109-8114.

Chen, B. H., and Chen, Y. Y. 1993. Stability of chlorophylls and carotenoids in sweet potato leaves during microwave cooking. J. Agric. Food Chem. 41: 1315-1320.

188 Chen, B. H., Peng, H. Y., and Chen, H. E. 1995. Changes of carotenoids, color, and vitamin A contents during processing of carrot juice. J. Agric. Food Chem. 43: 1912- 1918.

Chin, H. W., Bernhard, R. A., and Rosenberg, M. 1996. Solid phase microextraction for cheese volatile compound analysis. J. Food Soi. 61:1118-1122.

Chopra, M., Willson, R. L., and Thumham, D. I. Free radical scavenging o f lutein in vitro. In Carotenoids in Human Health, vol. 691, Canfield, L. M., Krinsky, N. I., and Olsen, J. A., eds. New York: Annals of New York Academy of Sciences, 1993, pp. 246- 249.

Cilento, G., and Adam, W. 1988. Photochemistry and photobiology without light. Photochem. Photobiol. 48: 361-368.

Cillard, J., Cillard, P., and Cormier, M. 1980. Effect of experimental factors on the prooxidant behavior of a-tocopherol. J. Am. Oil Chem. Soc. 80: 255-261.

Clark, T. J., and Bunch, J. E. 1997. Qualitative and quantitative analysis o f flavor additives on tobacco products using SPME-GC-Mass spectroscopy. J. Agric. Food Chem. 45: 844-849.

Clements, A. H., Van den Engh, R. H., Frost, D. J., Hoogenhout, K., and Nooi, J. R. 1973. Participation of singlet oxygen in photosensitized oxidation of 1,4-dienoic systems and photooxidation of soybean oil. J. Am. Oil Chem. Soc. 50: 325-330.

Cole, E. R., and Kapur, N. S. 1957. The stability of lycopene. I.-Degradation by oxygen. J. Sci. Food Agric. 8 : 360-365.

Craft, N. E. Carotenoid reversed-phase high-performance liquid chromatography methods: reference compendium. In Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 185-205.

Craft, N. E., and Soares Jr., J. H. 1992. Relative solubility, stability, and absorptivity of lutein and (3-carotene in organic solvents. J. Agric. Food Chem. 40:431-434.

Davies, B. H. Carotenoids. In Chemistry and Biochemistry of Plant Pigments, vol. 2, Goodwin, T. W., ed. London: Academic Press, 1976, pp. 38-165.

De Leenheer, A. P., and Nelis, H. J. 1992. Profiling and quantitation of carotenoids by high-performance liquid chromatography and photodiode array detection. In Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 251-265.

189 Desobry, S. A., Netto, F. M., and Labuza, T. P. 1997. Comparison of spray-drying, drum- drying and freeze-drying for P-carotene encapsulation and preservation. J. Food Sci. 62: 1158-1162.

Devasagayam, T. P. A., Werner, T., Ippendorf, H., Martin, H.-D., and Sies, H. 1992. Synthetic carotenoids, novel polyene polyketones and new capsorubin isomers as efficient quenchers of singlet molecular oxygen. Photochem. Photobiol. 55: 511-514.

Di Mascio, P., Kaiser, S., and Sies, H. 1989. Lycopene as the most efficient biological carotenoid singlet oxygen quencher, ^rc/z. Biochem. Biophys. 274: 532-538.

Di Mascio, P., Murphy, M. E., and Sies, H. 1991. Antioxidant defense systems: the role of carotenoids, tocopherols, and thiols. Am. J. Clin. Nutr. 53: 194S-200S.

Di Mascio, P., and Sies, H. 1989. Quantifrcation of singlet oxygen generated by thermolysis of 3, 3 '-(1,4-naphthylidene) dipropionate. Monomol and dimol photoemission and the effects of 1,4-diazabicyclo [2.2.2]octane. J. Am. Chem. Soc. Ill: 2909-2914.

Di Mascio, P., Sundquist, A. R., Devasagayam, T. P. A., and Sies, H. Assay o f lycopene and other carotenoids as singlet oxygen quenchers. In Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 429-438.

Di Mascio, P., Wefers, H., Do-Thi, H.-P., Lafleur, M. V. M., and Sies, H. 1989. Singlet molecular oxygen causes loss of biological activity in plasmid and bacteriophage DNA and induces single-strand breaks. Biochim. Biophys. Acta. 1007: 151-157.

Dorgan, J. P., Sowell, A., Swanson, C. A., Potischman, N., Miller, R., Schussler, N., and Stephenson Jr., H. E. 1998. Relationships of serum carotenoids, retinol, a-tocopherol, and selenium with breast cancer risk: results from a prospective study in Columbia, Missouri (United States). Cancer Cans. Cont. 9: 89-97.

Dupuy, H.P., Flick Jr., G. J., Bailey, M. E., St. Angelo, A. J., Legendre, M. G., and Sumrell, G. 1985. Direct sampling capillary gas chromatography of volatiles in vegetable oils. J. Am. Oil Chem. Soc. 62:1690-1693.

Edge, R., McGarvey, D. J., and Truscott, T. G. 1997. The carotenoids as antioxidants — a review. JPhotochem. Photobiol B: Biol. 41: 189-200.

El-Tinay, A. H., and Chichester, C. O. 1970. Oxidation of P-carotene. Site o f initial attack. J. Org. Chem. 35: 2290-2293.

190 Emenhiser, C., Sander, L. C., and Schwartz, S. J. 1995. Capability of a polymeric C 30 stationary phase to resolve cis-trans carotenoid isomers in reverse-phase liquid chromatography. J. Chrom. A. 707: 205-216.

Emenhiser, C., Simunovic, N., Sander, L. C., and Schwartz, S. J. 1996. Separation of geometrical carotenoid isomers in biological extracts using a polymeric C 30 column in reversed-phase liquid chromatography. J. Agric. Food Chem. 44: 3887-3893.

Endo, Y., Usuki, R., and Kaneda, T. 1984. Prooxidant activities of chlorophylls and their decomposition products on the photooxidation of methyl linoleate. J. Am. Oil Chem. Soc. 61:781-784.

Englert, G. NMR spectroscopy. In Carotenoids, Volume IB: Spectroscopy. Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 147-260.

Englert, G., Brown, B. O., Moss, G. P., Weedon, C. L., Britton, G, Goodwin, T. W., Simpson, K. L., and Williams, R. J. H. 1979. Prolycopene, a tetra-c/j carotene with two hindered cis double bonds. J. Chem. Soc. Chem. Commun. 279: 545-547.

Enzell, C. R., and Back, S. Mass spectroscopy. In Carotenoids, Volume IB: Spectroscopy. Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 261-320.

Epler, K. S., Ziegler, R. G., and Craft, N. E. 1993. Liquid chromatographic method for the determination of carotenoids, retinoids and tocopherols in human serum and in food. J. Chrom. 619: 37-48.

Fahrenholtz, S. R., Doleiden, P. H., Trozzolo, A. M., and Lamola, A. A. 1974. On the quenching of singlet oxygen by a-tocopherol. Photochem. Photobiol. 20: 505-509.

Fakourelis, N., Lee, E. C., and Min, D. B. 1987. Effects of chlorophyll and P-carotene on the oxidative stability of soybean oil. J. Food Sci. 52: 234-235.

Faria, J. A. P., and Mukai, M. K. 1983. Use of a gas chromatographic reactor to study lipid photooxidation. J. Am. Oil Chem. Soc. 60: 1-5.

Farmilo, A., and Wilkinson, F. 1973. On the mechanism of quenching o f singlet oxygen in solution. Photochem. Photobiol. 18: 447-450.

Ferruzzi, M. G., Sander, L. C., Rock, C. L., and Schwartz, S. J. 1998. Carotenoid determination in biological microsamples using liquid chromatography with a coulometric electrochemical array detector. Anal. Biochem. 256: 74-81.

191 Field, J. A., Nickerson, G., James, D. D., and Heider, C. 1996. Determination of essential oils in hops by headspace solid-phase microextraction. J. Agric. Food Chem. 44: 1768- 1772.

Foote, C. S. Light, oxygen, and toxicity. In Pathology o f Oxygen. Autor, A. P., ed. New York: Academic Press, Inc., 1982, pp. 21-42.

Foote, C. S. Quenching of singlet oxygen. In Singlet Oxygen. Wasserman, H. H., and Murray, R. W., eds. New York: Academic Press, Inc., 1979, pp. 139-171.

Foote, C. S. 1968. Mechanisms of photosensitized oxidation. Science. 162: 963-970.

Foote, C. S. Photosensitized oxidation and singlet oxygen: consequences in biological systems. In Free Radicals in Biology, vol. II, Prior, W. A., ed. New York: Academic Press, Inc., 1976, pp. 85-133.

Foote, C. S., Chang, Y. C., and Denny, R. W. 1970. Chemistry of singlet oxygen. X. Carotenoid quenching parallels biological protection. J. Am. Chem. Soc. 92: 5216-5218.

Foote, C. S., and Denny, R. W. 1968. Chemistry of singlet oxygen. VII. Quenching by P- carotene. J.Am. Chem. Soc. 90: 6233-6235.

Foote, C. S., Denny, R. W., Weaver, L., Chang, Y., Phil, D., and Peters, J. 1970. Quenching of singlet oxygen. Ann. N. Y Acad. Sci. 171: 139-145.

Frankel, E. N. Chemistry of autoxidation: mechanism, products and flavor significance. In Flavor Chemistry of Fats and Oils, Min, D. B., and Smouse, T. H., eds. American Oil Chemists’ Society, 1985, pp. 1-37.

Frankel, E. N. 1993. Formation of headspace volatiles by thermal decomposition of oxidized fish oils vs. oxidized vegetable oils. J. Am. Oil Chem. Soc. 70: 767-772.

Frankel, E. N. 1991. Recent advances in lipid oxidation. J. Sci. Food Agric. 54: 495-511.

Frankel, E. N. The antioxidant and nutritional effects of tocopherols, ascorbic acid and P- carotene in relation to processing of edible oils. In Nutritional Impact o f Food Processing. Somogyi, J. C., and Muller, H. R., eds. Basel: Karger, 1989, pp. 297-312.

Frankel, E. N., Evans, C. D., and Cooney, P. M. 1959. Tocopherol oxidation in natural fats. J. Agric. Food Chem. 7:438-441.

Frankel, E. N., Neff, W. E., and Bessler, T. R. 1979. Analysis of autoxidized fats by gas chromatography-mass spectrometry: V. Photosensitized oxidation. Lipids. 14: 961-967.

192 Gandini, N., and Riguzzi, R. 1997. Headspace solid-phase microextraction analysis of methyl isothiocyanate in wine. J. Agric. Food Chem. 45: 3092-3094.

Garcia-Casal, M. N., Layrisse, M., Solano, L., Baron, M. A., Arguello, F., Llovera, D., Ramirez, J., Leets, 1., and Tropper, E. 1998. Vitamin A and P-carotene can improve nonheme iron absorption from rice, wheat, and com by humans. J. Nutr. 128: 646-650.

Garewal, H. 1995. Antioxidants in oral cancer prevention. Am. J. Clin. Nutr. 62 (suppl): 1410S-1416S.

Gartner, C., Stahl, W., and Sies, H. 1997. Lycopene is more bioavailable from tomato paste than from fresh tomatoes. Am. J. Clin. Nutr. 66: 116-122.

Giovannucci, E., Ascherio, A., Rimm, E. B., Stampfer, M. J., Colditz, G. A., and Willett, W. C. 1995. Intake of carotenoids and retinol in relation to risk of prostate cancer. J. Natl. Cancer Inst. 87: 1767-1776.

Godoy, H. T., and Rodriguez-Amaya, D. B. 1987. Changes in individual carotenoids on processing and storage of mango (Mangifera indicd) slices and puree. Int. J. Food Sci. Tech. 22: 451-460.

Goodwin, T. W. Functions of carotenoids. In The Biochemistry o f the Carotenoids, vol. 1. Goodwin, T. W., ed. New York: Chapman and Hall, 1980, pp. 77-82.

Hakala, S. H., and Heinonen, I. M. 1994. Chromatographic purification of natural lycopene. J. Agric. Food Chem. 42: 1314-1316.

Haila, K. M., Lievonen, S. M., and Heinonen, M. I. 1996. Effects of lutein, lycopene, annatto, and y-tocopherol on autoxidation of triglycerides. J. Agric. Food Chem. 44: 2096-2100.

Handelman, G. J., van Kuijk, F. J. G. M., Chatteqee, A., and Krinsky, N. I. 1991. Characterization of products formed during the autoxidation of P-carotene. Free Rad. Biol. Med. 10:427-437.

Hart, D. J., and Scott, K. J. 1995. Development and evaluation of an HPLC method for the analysis of carotenoids in foods, and the measurement o f the carotenoid content of vegetables and fruits commonly consumed in the UK. Food Chem. 54: 101-110.

Heinonen, M., Haila, K., Lampi, A.-M., and Piironen, V. 1997. Inhibition of oxidation in 10% oil-in-water emulsions by P-carotene with a- and y-tocopherols. J. Am. Oil Chem. Soc. 74: 1047-1052.

193 Henry, L. K., Catignani, G. L., and Schwartz, S. J. 1998. Oxidative degradation kinetics of lycopene, lutein, and 9-cis and dX\-trans P-carotene. J. Am. Oil Chem. Soc. 75: 823- 829.

Henry, L. K., Catignani, G. L., and Schwartz, S. J. 1998. The influence of carotenoids and tocopherols on the stability of safflower seed oil during heat-catalyzed oxidation. J. Am. Oil Chem. Soc. 75: 1-4.

Holman, R, T. 1949. Spectrophotometric studies of the oxidation o f fats. VIH. Coupled oxidation of carotene. Arch. Biochem. 21: 51-57.

Humbeck, K. 1990. Photoisomerization o f lycopene during carotenogenesis in mutant C- 6 D of Scenedesmus obliquus. Planta. 182: 204-210.

Ibanez, E., and Bernhard, R. A. 1996. Solid-phase microextraction (SPME) of pyrazines in model reaction systems. J. Sci. Food Agric. 72: 91-96.

Jelen, H. H., Wlazly, K., Wasowicz, E., and Kaminski, E. 1998. Solid-phase microextraction for the analysis of some alcohols and esters in beer: comparison with static headspace method. J. Agric. Food Chem. 46: 1469-1473.

Jensen, N.-H., Nielsen, A. B., and Wilbrandt, R. 1982. Chlorophyll a sensitized trans-cis photoisomerization of a//-/ran 5 -P-carotene. Am. Chem. Soc. 104: 6117-6119.

Jia, M., Zhang, Q. H., and Min, D. B. 1998. Optimization of solid-phase microextraction analysis for headspace flavor compounds o f orange juice. J. Agric. Food Chem. 46: 2744- 2747.

Johnson, E. L., and Stevenson, R. Vll. Special techniques. InBasic Liquid Chromatography. Varian Associates, Inc., 1978, pp. 165-222.

Jung, M. Y., and Min, D. B. 1992. Effects of oxidized a-, y- and 6 -tocopherols on the oxidative stability of purified soybean oil. Food Chem. 45: 183-187.

Jung, M. Y., and Min, D. B. 1991. Effects of quenching mechanisms of carotenoids on the photosensitized oxidation of soybean oil. J. Am. Oil Chem. Soc. 6 8 : 653-658.

Kaiser, S., Di Mascio, P., Murphy, M. E., and Sies, H. 1990. Physical and chemical scavenging of singlet molecular oxygen by tocopherols. Arch. Biochem. Biophys. 277: 101-108.

194 Kanasawud, P., and Crouzet, J. C. 1990. Mechanism of formation of volatile compounds by thermal degradation of carotenoids in aqueous medium. 1. P-carotene degradation. J. Agric. Food Chem. 38: 237-243.

Kanasawud, P., and Crouzet, J. C. 1990. Mechanism of formation of volatile compounds by thermal degradation of carotenoids in aqueous medium. 2. Lycopene degradation. J. Agric. Food Chem. 38: 1238-1242.

Kao, J.-W., Hammond, E. G., and White, P. J. 1998. Volatile compounds produced during deodorization of soybean oil and their flavor significance. J. Am. Oil Chem. Soc. 75: 1103-1107.

Karrer, P., and Jucker, E. 1945. Oxides o f P-carotene: P-carotene monoepoxide, P- carotene diepoxide, mutatochrome, aurochrome, luteochrome. Helv. Chim. Acta. 28: 427- 436.

Kaufinann, R., Wingerath, T., Kirsch, D., Stahl, W., and Sies, H. 1996. Analysis of carotenoids and carotenol fatty acid esters by matrix-assisted laser desorption ionization (MALDI) and MALDI-post-source-decay mass spectrometry. Anal. Biochem. 238: 117- 128.

Kearsley, M. W., and Rodriguez, N. 1981. The stability and use of natural colours in foods: anthocyanin, P-carotene and riboflavin. J. Fd. Technol. 16: 421-431.

Kellogg in, E. W., and Fridovich, I. 1975. Superoxide, hydrogen peroxide, and singlet oxygen in lipid peroxidation by a xanthine oxidase system. J. Biol. Chem. 250: 8812- 8817.

Khachik, P., Beecher, G. R., Goli, M. B., Lusby, W. R., and Smith Jr., J. C. 1992a. Separation and identification of carotenoids and their oxidation products in the extracts of human plasma. Anal. Chem. 64: 2111-2122.

Khachik, P., Beecher, G. R., and Lusby, W. R. 1989. Separation, identification, and quantification of the major carotenoids in extracts of apricots, peaches, cantaloupe, and pink grapefruit by liquid chromatography. J. Agric. Food Chem. 37: 1465-1473.

Khachik, P., Beecher, G. R., and Smith Jr., J. C. 1995. Lutein, lycopene, and their oxidative metabolites in chemoprevention o f cancer. J. Cell. Biochem. 22: 236-246.

Khachik, P., Goli, M. B., Beecher, G. R., Holden, J., Lusby, W. R., Tenorio, M. D., and Barrera, M. R. 1992b. Effect of food preparation on qualitative and quantitative distribution of major carotenoid constituents of tomatoes and several green vegetables. J. Agric. Food Chem. 40: 390-398.

195 Khachik, F., Pfander, H., and Traber, B. 1998. Proposed mechanisms for the formation of synthetic and naturally occurring metabolites of lycopene in tomato products and human serum. J. Agric. Food Chem. 46: 4885-4890.

Khachik, P., Steck, A., Niggli, U. A., and Pfander, H. 1998. Partial synthesis and structural elucidation of the oxidative metabolites of lycopene identified in tomato paste, tomato juice, and human serum. J. Agric. Food Chem. 46:4874-4884.

BCiritsakis, A., and Dugan, L. R. 1985. Studies in photooxidation of olive oil. J. Am. Oil Chem. Soc. 62: 892-896.

Kopas-Lane, L. M., and Warthesen, J. J. 1995. Carotenoid photostability in raw spinach and carrots during cold storage. J. Food Sci. 60: 773-776.

Krinsky, N. 1989. Antioxidant functions of carotenoids. Free Rad. Biol. Med. 7: 617-635.

Krinsky, N. 1994. The biological properties of carotenoids. Pure Appl. Chem. 6 6 : 1003- 1010.

Krinsky, N. 1979. Carotenoid protection against oxidation. Pure Appl. Chem. 51: 649- 660.

Lee, E. C., and Min, D. B. 1988. Quenching mechanism of p-carotene on the chlorophyll sensitized photooxidation of soybean oil. J. Food Sci. 53:1894-1895.

Lee, I., Fatemi, S. H., Hammond, E. G., and White, P. J. 1995. Quantitation of flavor volatiles in oxidized soybean oil by dynamic headspace analysis. J. Am. Oil Chem. Soc. 72: 539-546.

Lee, K. H., Jung, M. Y., and Kim, S. Y. 1997. Quenching mechanism and kinetics of ascorbyl palmitate for the reduction o f the photosensitized oxidation of oils. J. Am. Oil Chem. Soc. 74: 1053-1057.

Lee, S.-H., and Min, D. B. 1990. Effects, quenching mechanisms, and kinetics of carotenoids in chlorophyll-sensitized photooxidation of soybean oil. J. Agric. Food Chem. 38: 1630-1634.

Lepage, G., Champagne, J., Ronco, N., Lamarre, A., Osberg, I., Sokol, R. J., and Roy, C. C. 1996. Supplementation with carotenoids corrects increased lipid peroxidation in children with cystic fibrosis. Am. J. Clin. Nutr. 64: 87-93.

196 Levy, J., Bosin, E., Feldman, B., Giat, Y., Münster, A., Danilenko, M., and Sharoni, Y. 1995. Lycopene is a more potent inhibitor of human cancer cell proliferation than either a-carotene or P-carotene. Nutr. Cancer. 24: 257-266.

Liaaen-Jensen, S. Combined approach: identification and structure elucidation of carotenoids. In Carotenoids, Volume IB: Spectroscopy. Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 343-354.

Liebler, D. C. Antioxidant reactions o f carotenoids. In Carotenoids in Human Health, vol. 691, Canfield, L. M., Krinsky, N. L, and Olsen, J. A., eds. New York: Annals o f New York Academy of Sciences, 1993, pp. 20-31.

Liebler, D. C., Baker, P. P., and Kaysen, K. L. 1990. Oxidation of vitamin E: evidence for competing autoxidation and peroxyl radical trapping reactions of the tocopheroxyl radical./. Am. Chem. Soc. 112: 6995-7000.

Liebler, D. C., and Kennedy, T. A. Epoxide products of 13-carotene antioxidant reactions. \n Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 472-479.

Love, J. Factors affecting lipid oxidation-metal catalysts and chelators. In Flavor Chemistry o f Fats and Oils, Min, D. B., and Smouse, T. H., eds. American Oil Chemists’ Society, 1985, pp. 61-75.

Lu, Y., Etoh, H., Watanabe, N., Ina, K., Ukai, N., Oshima, S., Ojima, P., Sakamoto, H., and Ishiguro, Y. 1995. A new carotenoid, hydrogen peroxide oxidation products firom lycopene. Biosci. Biotech. Biochem. 59: 2153-2155.

Mangoon, E. P., and Zechmeister, L. 1957. Stepwise stereoisomerization of prolycopene, a polycw carotenoid, to all-/ra/w-lycopene. Arch. Biochem. Biophys. 69: 535-547.

Martin, K. R., Pailla, M. L., and Smith Jr., J. C. 1996. (3-carotene and lutein protect HepG2 human liver cells against oxidant-induced damage. J. Nutr. 126: 2098-2106.

Marty, C., and Berset, C. 1986. Degradation of rm/w-|3-carotene during heating in sealed glass tubes and extrusion cooking. / Food Sci. 51: 698-702.

Marty, C., and Berset, C. 1988. Degradation products o f trans-P-carotene produced during extrusion cooking. J. Food Sci. 53: 1880-1886.

Marty, C., and Berset, C. 1990. Factors affecting the thermal degradation of all-trans-^- carotene. J. Agric. Food Chem. 38: 1063-1067.

197 Matich, A. J., Rowan, D. D., and Banks, N. H. 1996. Solid phase microextraction for quantitative headspace sampling of apple volatiles. Anal. Chem. 6 8 : 4114-4118.

Matsushita, S., and Terao, J. Singlet oxygen-initiated photo oxidation of unsaturated fatty acid esters and inhibitory effects o f tocopherols and P-carotene. In Autoxidation in Food and Biological Systems. Simic, M. G., and Karel, M., eds. New York: Plenum Press, 1980, pp. 27-44.

Meydani, M., Martin, A., Ribaya-Mercado, J. D., Gong, J., Blumberg, J. B., and Russell, R. M. 1994. P-carotene supplementation increases antioxidant capacity of plasma in older women. J. Nutr. 124: 2397-2403.

Miller, N. J., Sampson, J., Candeias, L. P., Bramley, P. M., and Rice-Evans, C. A. 1996. Antioxidant activities of carotenes and xanthophylls. FEES Lett. 384: 240-242.

Min, D. B., and Wen, J. 1983. Effects of dissolved free oxygen on the volatile compounds of oil. J. Food Sci. 48: 1429-1430.

Minguez-Mosquera, M. I., and Jaren-Galan, M. 1995. Kinetics of the decouloring of carotenoid pigments./. Sci. Food Agric. 67: 153-161.

Morales, M. T., Rios, J. J., and Aparicio, R. 1997. Changes in the volatile composition of virgin olive oil during oxidation: flavors and off-flavors. J. Agric. Food Chem. 45: 2666- 2673.

Mordi, R. C., Walton, J. C., Burton, G. W., Hughes, L., Ingold, K. U., and Lindsay, D. A. 1991. Exploratory study of P-carotene autoxidation. Tetra. Lett. 32: 4203-4206.

Nagasawa, H., Mitamura, T., Sakamoto, S., and Yamamoto, K. 1995. Effects o f lycopene on spontaneous mammary tumour development in SHN virgin mice. Anticancer Res. 15: 1173-1178.

Nelis, H. J. C. P., and De Leenheer, A. P. 1983. Isocratic nonaqueous reversed-phase liquid chromatography of carotenoids. Anal. Chem. 55: 270-275.

Ng, L.-K., Hupe, M., Hamois, J., and Moccia, D. 1996. Characterisation of commercial vodkas by solid-phase microextraction and gas chromatography/mass spectrometry analysis. Sci. Food Agric. 70: 380-388.

Nielsen, B. R., Mortensen, A., Jorgensen, K., and Skibsted, L. H. 1996. Singlet versus triplet reactivity in photodegradation of C^g carotenoids. J. Agric. Food Chem. 44: 2106- 2113.

198 Ogimlesi, A. T., and Lee, C. Y. 1979. Effect of thermal processing on the stereoisomerisation of major carotenoids and vitamin A value of carrots. Food Chem. 4: 311-318.

O’Neil, C. A., and Schwartz, S. J. 1992. Chromatographic analysis of cisltrans carotenoid isomers. J. Chrom. 624: 235-252.

Onyewu, P. N., Ho, C.-T., and Daun, H. 1986. Characterization of p-carotene thermal degradation products in a model food system. J. Am. Oil Chem. Soc. 63: 1437-1441.

Ouyang, J. M., Daun, H., Chang, S. S., and Ho, C.-T. 1980. Formation of carbonyl compounds from P-carotene during palm oil deodorization. J. Food Sci. 45: 1214-1217, 1222.

Packer, J. E., Mahood, J. S., Mora-Arellano, V. O., Slater, T. P., Willson, R. L., and WoLfenden, B. S. 1981. Free radicals and singlet oxygen scavengers: reaction of a peroxy-radical with P-carotene, diphenyl finran and l,4-diazobicycIo(2,2,2)-octane. Biochem. Biophys. Res. Comm. 98: 901-906.

Pan, L., Adams, M., and Pawliszyn, J. 1995. Determination of fatty acids using solid- phase microextraction. Anal. Chem. 67: 4396-4403.

Palozza, P., and Krinsky, N. I. Antioxidant effects of carotenoids in vivo and in vitro: an overview, hi Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 403-420.

Palozza, P., and Krinsky, N. I. 1992. p-cer'^tene a-tocopherol are synergistic antioxidants. .4rc/î. Biochem. Biophys. 297: 184-187.

Park, P. S. W. 1993. Loss of volatile lipid oxidation products during thermal desorption in dynamic headspace-capillary gas chromatography. J. Food Sci. 58:220-222.

Park, P. S. W., and Goins, R. E. 1992. Determination of volatile lipid oxidation products by dynamic headspace-capillary gas chromatographic analysis with application to milk- based nutritional products. J. Agric. Food Chem. 40: 1581-1585.

Parker, R. S. 1996. Absorption, metabolism, and transport of carotenoids. FASEB J. 10: 542-551.

Pelusio, F., Nilsson, T., Montanarella, L., Tilio, R., Larsen, B., Facchetti, S., and Madsen, J. O. 1995. Headspace solid-phase microextraction analysis of volatile organic sulfur compoimds in black and white truffle aroma. J. Agric. Food Chem. 43: 2138-2143.

199 Pesek, C. A., and Warthesen, J. J. 1987. Photodegradation of carotenoids in a vegetable juice system. J. Food Sci. 52: 744-746.

Peto, R., Doll, R., Buckley, J. D., and Spom, M. B. 1981. Can dietary beta-carotene materially reduce human cancer rates? Nature. 290: 201-208.

Pfander, H. Carotenoids: an overview. In Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 3-13.

Pfander, H. Chromatography: part I, general aspects. In Carotenoids, Volume I A: Isolation and Analysis, Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 109-116.

Pfander, H., and Riesen, R. Chromatography: part IV, high-performance liquid chromatography. In Carotenoids, Volume lA: Isolation and Analysis, Britton, G., Liaaen- Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 145-190.

Piretti, M. V., Diamante, M., Bombardelli, E., and Morazzoni, P. 1996. Anomalous behaviour of lycopene by HPLC on cyano-amino polar phase. Biomed. Chrom. 10: 43-45.

Pool-Zobel, B. L., Bub, A., Muller, H., Wollowski, L, and Rechkemmer, G. 1997. Consumption of vegetables reduced genetic damage in humans: first results of a human intervention trial with carotenoid-rich foods. Carcinogenesis. 18: 1847-1850.

Pryde, E. H. Soybean vs. other vegetable oils as sources of edible oil products. In Handbook of Soy Oil Processing and Utilization, Erickson, D. R., Pryde, E. H., Brekke, O. L., Mounts, T. L., and Falb, R. A., eds. Champaign: American Oil Chemists’ Society, 1980, pp. 1-11.

Rahmani, M., and Csallany, A. S. 1998. Role of minor constituents in the photooxidation of virgin olive oil. J. Am. Oil Chem. Soc. 75: 837-843.

Ribaya-Mercado, J. D., Garmyn, M., Gilchrest, B. A., and Russell, R. M. 1995. Skin lycopene is destroyed preferentially over P-carotene during ultraviolet irradiation in humans. J.Nutr. 125: 1854-1859.

Ritacco, R. P., Britton, G., and Simpson, K. L. 1984. A proposed mechanism for the hydroxylation of carotenoids on Micro-Cel C. J. Agric. Food Chem. 32: 301-304.

Ritacco, R. P., Rodriguez, D. B., Britton, G., Lee, T.-C., Chichester, C. O., and Simpson, K. L. 1984. Investigations of carotenoid reactions on Micro-Cel C. J. Agric. Food Chem. 32: 296-300.

200 Robards, K., Kerr, A. F., Patsalides, E., and Korth, J. 1988. Headspace gas analysis as a measure of rancidity in com chips. J. Am. Oil Chem. Soc. 65: 1621-1626.

Rock, C. L., Jacob, R. A., and Bowen, P. E. 1996. Update on the biological characteristics of the antioxidant micronutrients: vitamin C, vitamin E, and the carotenoids. J. Am. Diet. Assoc. 96: 693-702.

Rock, C. L., Lovalvo, J. L., Emenhiser, C., Ruffin, M. T., Platt, S. W., and Schwartz, S. J. 1998. Bioavailability of p-carotene is lower in raw than in processed carrots and spinach in women. J.Nutr. 128: 913-916.

Sadler, G., Davis, J., and Dezman, D. 1990. Rapid extraction of lycopene and P-carotene from reconstituted tomato paste and pink grapefruit homogenates. J. Food Sci. 55: 1460- 1461. Saito, I., Matsuura, T., and Inoue, K. 1983. Formation of superoxide ion via one-electron transfer from electron donors to singlet oxygen. J. Am. Chem. Soc. 105: 3200-3206.

Sandler, L. C., Sharpless, K. E., Craft, N. E., and Wise, S. A. 1994. Development of engineered stationary phases for the separation of carotenoid isomers. Anal. Chem. 66: 1667-1674.

Santos, M. S., Meydani, S. N., Leka, L., Wu, D., Fotouhi, N., Meydani, M., Hennekens, C. H., and Gaziano, J. M. 1996. Natural killer cell activity in elderly men is enhanced by P-carotene supplementation. Am. J. Clin. Nutr. 64: 772-777.

Schiedt, K. Chromatography: part III, thin-layer chromatography. In Carotenoids, Volume lA: Isolation and Analysis, Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 131-144.

Schiedt, K., and Liaaen-Jensen, S. Isolation and analysis. InCarotenoids, Volume lA: Isolation and Analysis, Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 81-108.

Schmitz, H. H., Emenhiser, C., and Schwartz, S. J. 1995. HPLC separation of geometric carotene isomers using a calcium hydroxide stationary phase. J. Agric. Food Chem. 43: 1212-1218.

Schmitz, H. H., van Breemen, R. B., and Schwartz, S. J. Fast-atom bombardment and continuous-ftow fast-atom bombardment mass spectrometry in carotenoid analysis. In Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 322-336.

Schuep, W., and Schierle, J. 1997. Determination of p-carotene in commercial foods: interlaboratory study. J. AOCS Inter. 80: 1057-1064.

201 Scott, K. J. 1992. Observations on some of the problems associated with the analysis of carotenoids in foods by HPLC. Food Chem. 45: 357-364.

Scott, K. J., and Hart, D. J. 1993. Further observations on problems associated with the analysis of carotenoids by HPLC-2: column temperature. Food Chem. 47: 403-405.

Seely, G. R., and Meyer, T. H. 1971. The photosensitized oxidation o f (3-carotene. Photochem. Photobiol. 13: 27-32.

Selke, E., and Frankel, E. N. 1987. Dynamic headspace capillary gas chromatographic analysis of soybean oil volatiles. J. Am. Oil Chem. Soc. 64: 749-753.

Semba, R. D. 1998. The role of vitamin A and related retinoids in immune function. Nutr. Rev. 56: S38-S48.

Sharoni, Y., and Levy, J. 1995. Lycopene, the major tomato carotenoid, inhibits tumor growth in vitro and in vivo. Cancer Det. Prevent. 19: 110.

Siefermann-Harms, D., and Ninnemann, H. 1982. Pigment organization in the light- harvesting chlorophyll a/b protein complex of lettuce chloroplasts. Evidence obtained from protection of the chlorophylls against proton attack and from excitation energy transfer. Photochem. Photobiol. 35: 719-731.

Sies, H., and Stahl, W. 1995. Vitamins E and C, (3-carotene, and other carotenoids as antioxidants, J. Clin. Nutr. 62: 1315S-1321S.

Silveira Jr., A., and Evans, J. M. 1995. Flash chromatographic separation and electronic absorption spectra of carotenoids. J. Chem. Educ. 72: 374-375.

Sistrom, W. A., Griffiths, M., and Stanier, R. Y. 1956. The biology o f a photosynthetic bacterium which lacks colored carotenoids. J. Cell Comp. Physiol. 48: 473-515.

Snyder, J. M., Frankel, E. N., and Selke, E. 1985. Capillary gas chromatographic analysies of headspace volatiles from vegetable oils. J. Am. Oil Chem. Soc. 62: 1675- 1679.

Snyder, J. M., Frankel, E. N., Selke, E., and Warner, K. 1988. Comparison of gas chromatographic methods for volatile lipid oxidation compounds in soybean oil. J. Am. Oil Chem. Soc. 65: 1617-1620.

Song, J., Fan, L., and Beaudry, R. M. 1998. Application of solid phase microextraction and gas chromatography/time-of-flight mass spectrometry for rapid analysis of flavor volatiles in tomato and strawberry fruits. J. Agric. Food Chem. 46: 3721-3726.

202 Song, J., Gardner, B. D., Holland, J. F., and Beaudry, R. M. 1997. Rapid analysis of volatile flavor compounds in apple fruit using SPME and GC/time-of-flight mass spectrometry./. Agric. Food Chem. 45: 1801-1807.

Stahl, W., Junghans, A., de Boer, B., Driomina, E. S., Briviba, K., and Sies, H. 1998. Carotenoid mixtures protect multilamellar liposomes against oxidative damage: synergistic effects of lycopene and lutein. FEES Letters. 427: 305-308.

Stahl, W., Schwartz, W., Sundquist, A. R., and Sies, H. 1992. cis-trans isomers of lycopene and P-carotene in human serum and tissues. Arch. Biochem. Biophys. 294: 173- 177.

Stahl, W., and Sies, H. Physical quenching of singlet oxygen and cis-trans isomerization of carotenoids. In Carotenoids in Human Health, vol. 691, Canfield, L. M., Krinsky, N. I., and Olsen, J. A., eds. New York: Annals of New York Academy of Sciences, 1993, pp. 10-19.

Stahl, W., and Sies, H. 1996. Lycopene: A biologically important carotenoid for humans? Arch. Biochem. Biophys. 336: 1-9.

Stahl, W., Sundquist, A. R., Hanusch, M., Schwartz, W., and Sies, H. 1993. Separation of P-carotene and lycopene geometrical isomers in biological samples. Clin. Chem. 39: 810- 814.

Steffen, A., and Pawliszyn, J. 1996. Analysis of flavor volatiles using headspace solid- phase microextraction. J. Agric. Food Chem. 44: 2187-2193.

Stocker, R., Yamamoto, Y., McDonagh, A. F., Glazer, A. N., and Ames, B. N. 1987. Bilirubin is an antioxidant of possible physiological importance. Science. 235: 1043- 1046.

Stratton, S. P., and Liebler, D. C. 1997. Determination of singlet oxygen-specific versus radical-mediated lipid peroxidation in photosensitized oxidation o f lipid bilayers: effect of P-carotene and a-tocopherol. Biochem. 36: 12911-12920.

Tan, B. 1988. Analytical and preparative chromatography of tomato paste carotenoids. J. Food Sci. 53: 954-959.

Tan, Y. A., Chong, C. L., and Low, K. S. 1997. Crude palm oil characteristics and chlorophyll content. J. Sci. Food Agric. 75: 281-288.

Teixeira Neto, R. O., Karel, M., Saguy, I., and Mizrahi, S. 1981. Oxygen uptake and P- carotene decoloration in a dehydrated food model. J. Food Sci. 46: 665-669, 676.

203 Terao, J. 1989. Antioxidant activity of P-carotene-related carotenoids in solution. Lipids. 24: 659-661.

Tonucci, L. H., Holden, J. M., Beecher, G. R., Khachik, F., Davis, C. S., and Mulokozi, G. 1995. Carotenoid content of thermally processed tomato-based food products. J. Agric. Food Chem. 43: 579-586.

Tsuchiya, M., Scita, G., Freisleben, H.-J., Kagan, V. E., and Packer, L. Antioxidant radical-scavenging activity of carotenoids and retinoids compared to a-tocopherol. In Methods in Enzymology, vol. 213. Academic Press, Inc., 1992, pp. 460-472.

Ukai, N., Lu, Y., Etoh, H., Yagi, A., Ina, K., Oshima, S., Ojima, F., Sakamoto, H., and Ishiguro, Y. 1994. Photosensitized oxidation of lycopene. Biosci. Biotech. Biochem. 58: 1718-1719.

Ullrich, F., and Grosch, W. 1988. Identification of the most intense odor compounds formed during autoxidation of methyl linolenate at room temperature. J. Am. Oil Chem. Soc. 65: 1313-1317.

Urruty, L., Montury, M., Braci, M., Fournier, J., and Doumel, J.-M. 1997. Comparison of two recent solventless methods for the determination of procymidone residues in wines: SPME/GC/MS and ELISA tests. J. Agric. Food Chem. 45: 1519-1522.

Usuki, R., Suzuki, T., Endo, Y., and Kaneda, T. 1984. Residual amounts of chlorophylls and pheophytins in refined edible oils. J. Am. Oil Chem. Soc. 61: 785-788. van Breemen, R. B. 1996. Innovations in carotenoid analysis using LC/MS. Anal. Chem. 6 8 : 299A-304A. van Breemen, R. B., Schmitz, H. H., and Schwartz, S. J. 1993. Continuous-flow fast atom bombardment liquid chromatography/mass spectroscopy of carotenoids. Anal. Chem. 65 : 965-969. van Poppel, G. 1996. Epidemiological evidence for p-carotene in prevention of cancer and cardiovascular disease. Eur. J. Clin. Nutr. 50 (suppl. 3): S57-S61. van Vliet, T. 1996. Absorption of P-carotene and other carotenoids in humans and animal models. Eur. J. Clin. Nutr. 50 (suppl. 3): S32-S37.

Vega, P. J., Balaban, M. O., Sims, C. A., O’Keefe, S. F., and Comell, J. A. 1996. Supercritical carbon dioxide extraction efficiency for carotenes firom carrots by RSM. J. Food Sci. 61: 757-759.

204 Wagner, L. A., and Wartheson, J. L. 1995. Stability of spray-dried encapsulated carrot carotenes./. FoodSci. 60: 1048-1053.

Wang, H., Cao, G., and Prior, R. L. 1997. Oxygen radical absorbing capacity of anthocyanins. / Agric. Food Chem. 45: 304-309.

Warner, K., and Frankel, E. N. 1987. Effects of P-carotene on light stability of soybean oil. /. Am. Oil Chem. Sac. 64: 213-218.

Warner, K., Frankel, E. N., and Moulton, K. J. 1988. Flavor evaluation of crude oil to predict the quality of soybean oil. / Am. Oil Chem. Soc. 65: 386-391.

Warner, K., Frankel, E. N., and Mounts, T. L. 1989. Flavor and oxidative stability of soybean, sunflower and low erucic acid rapeseed oils. J. Am. Oil Chem. Soc. 66: 558-564.

Weedon, B. L. C., and Moss, G. P. Structure and nomenclature. In Carotenoids, Volume lA: Isolation and Analysis, Britton, G., Liaaen-Jensen, S., and Pfander, H., eds. Basel: Birkhauser Verlag, 1995, pp. 27-70.

Weiss, T. J. Commercial oil sources. InFood Oils and their Uses, Weiss, T. J., ed. Westport: AVI Publishing Co., Inc, 1983, pp. 35-63.

Yamauchi, R., and Matsushita, S. 1977. Quenching effect of tocopherols on the methyl linoleate photooxidation and their oxidation products. Agric. Biol. Chem. 41: 1425-1430.

Yamauchi, R., Miyake, N., Inoue, H., and Kato, K. 1993. Products formed by peroxyl radical oxidation of P-carotene. J. Agric. Food Chem. 41: 708-713.

Yang, X., and Peppard, T. 1994. Solid-phase microextraction for flavor analysis. / Agric. Food Chem. 42: 1925-1930.

Yoon, S. H., Jung, M., and Min, D. 1988. Effects of thermally oxidized triglycerides on the oxidative stability of soybean oil. /. Am. Oil Chem. Soc. 65: 1652-1656.

Zhang, Z., and Pawliszyn, J. 1993. Headspace solid-phase microextraction. Anal. Chem. 65: 1843-1852.

Zhang, Z., Yang, M. J., and Pawliszyn, J. 1994. Solid phase microextraction. Anal. Chem. 66: 844A-853A.

205 IMAGE EVALUATION TEST TARGET (Q A -3 ) /

/

/ â

« C e y .x

%

liâ IM . I2j 1.0 yo “= ta 2.2 K 3.6 2.0 l.l 1.8

1.25 1.4

150mm

/A P P L IE D A IIVI/4GE . Inc 1653 East Main Street — • Rochester, NY 14609 USA Phone: 716/482-0300 ------Fax: 716/288-5989

O '993. Applied Image. Inc.. Ail Rights Reserved