minor components and their roles on lipid oxidation in bulk oil that contains association colloids

156
University of Massachuses - Amherst ScholarWorks@UMass Amherst Dissertations Dissertations and eses 5-1-2012 Minor Components and eir Roles on Lipid Oxidation in Bulk Oil at Contains Association Colloids Bingcan Chen University of Massachuses - Amherst, [email protected] Follow this and additional works at: hp://scholarworks.umass.edu/open_access_dissertations is Open Access Dissertation is brought to you for free and open access by the Dissertations and eses at ScholarWorks@UMass Amherst. It has been accepted for inclusion in Dissertations by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact [email protected]. Recommended Citation Chen, Bingcan, "Minor Components and eir Roles on Lipid Oxidation in Bulk Oil at Contains Association Colloids" (2012). Dissertations. Paper 540.

Upload: others

Post on 12-Sep-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

University of Massachusetts - AmherstScholarWorks@UMass Amherst

Dissertations Dissertations and Theses

5-1-2012

Minor Components and Their Roles on LipidOxidation in Bulk Oil That Contains AssociationColloidsBingcan ChenUniversity of Massachusetts - Amherst, [email protected]

Follow this and additional works at: http://scholarworks.umass.edu/open_access_dissertations

This Open Access Dissertation is brought to you for free and open access by the Dissertations and Theses at ScholarWorks@UMass Amherst. It hasbeen accepted for inclusion in Dissertations by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please [email protected].

Recommended CitationChen, Bingcan, "Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids" (2012).Dissertations. Paper 540.

Page 2: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

MINOR COMPONENTS AND THEIR ROLES ON LIPID OXIDATION IN BULK OIL THAT CONTAINS ASSOCIATION COLLOIDS

A Dissertation Presented

by

Bingcan Chen

Submitted to the Graduate School of the University of Massachusetts Amherst in partial fulfillment

of the requirements for the degree of

DOCTOR OF PHILOSOPHY

MAY 2012

The Department of Food Science

Page 3: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

© Copyright by Bingcan Chen 2012

All Rights Reserved

Page 4: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

MINOR COMPONENTS AND THEIR ROLES ON LIPID OXIDATION IN BULK OIL THAT CONTAINS ASSOCIATION COLLOIDS

A Dissertation Presented

by

BINGCAN CHEN

Approved as to style and content by:

________________________________________ Eric A. Decker, Chair

________________________________________ D. Julian McClements, Member

________________________________________ Matthew A. Holden, Member

_______________________________________ Eric A. Decker, Department Head Department of Food Science

Page 5: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

DEDICATIONS

To my parents and my wife,

who made all of this possible, for their endless encouragement and patience.

Page 6: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

v

ACKNOWLEDGMENTS

I would like to express my heartfelt thanks to my advisor, Dr. Eric Decker, for this

greatest opportunity he has provided, for his guidance, enthusiasm, support and incredible

talent on this research and in my pursuit of a Doctoral degree. It was such a great honor

and best fortune for me to be his student. My entire life has been changed because of his

kindness. I would like to thank Dr. Julian McClements for his advice and support on this

research and in my pursuit of a Doctoral degree.

Also, I would like to extend thanks to Dr. Matthew Holden for being a member of

my committee. I appreciate his work and his constructive advice towards this dissertation.

I would like to express my appreciation to Jean Alamed for her continuous

assistance in my research as well as her friendship. Many thanks to Kla, Get, Ashley,

Jeffery, Caitlin, Mette, Ricard, Dao, Mickaël, and Giovanni for their continuing and

unconditional friendship and support over the years; I could not have done this without

their helps. I would also like to thank all faculty, staff and lab members with special

thanks to Dr. Xiao, Dr. Park, Dr. Goddard, Fran and Dan. In addition, I would like to

thank my friends outside of the University and/or country who continue to support my

endeavors.

Most importantly, I would like to thank my family, especially my Mom and Dad,

who love me so much, and my wife Jiajia for her endless love, support and

encouragement!

Page 7: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

vi

ABSTRACT

MINOR COMPONENTS AND THEIR ROLES ON LIPID OXIDATION IN BULK OIL THAT CONTAINS ASSOCIATION COLLOIDS

MAY 2012

BINGCAN CHEN, B.S., SICHUAN UNIVERSITY OF SCIENCE AND ENGINEERING, SICHUAN, CHINA

M.S., CHONGQING UNIVERSITY, CHONGQING, CHINA

Ph.D., UNIVERSITY OF MASSACHUSETTS AMHERST

Directed by: Professor Eric A. Decker

The combination of water and surface active compounds found naturally in

commercially refined vegetable oils have been postulated to form physical structures

known as association colloids. This research studied the ability of 1,2-dioleoyl-sn-

glycerol-3-phosphocholine (DOPC) and water to form physical structures in stripped

soybean oil. Interfacial tension and fluorescence spectrometry results showed the critical

micelle concentration (CMC) of DOPC in stripped soybean oil was 650 and 950 μM,

respectively. Light scattering attenuation results indicated that the structure formed by

DOPC was reverse micelles. The physical properties of DOPC reverse micelles were

determined using small-angle X-ray scattering (SAXS) and fluorescence probes. These

studies showed that increasing the water concentration altered the size and shape of the

reverse micelles formed by DOPC.

The impact of DOPC reverse micelles on the lipid oxidation of stripped soybean oil

was investigated by following the formation of primary and secondary lipid oxidation

products. DOPC reverse micelles had a prooxidant effect, shortening the oxidation lag

Page 8: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

vii

phase of SSO at 55 °C. It also was not able to change the lipid oxidation of stripped

soybean oil compared with DOPC reverse micelles at same concemtration ( i.e., 950 μM).

1,2-dibutyl-sn-glycerol-3-phosphocholine (DC4PC) which has the shorter fatty acid than

DOPC was not able to form association colloids and did not impact lipid oxidation rates.

This indicated that the choline group of the phospholipid was not responsible for the

increased oxidation rates and suggested that the physical structure formed by DOPC was

responsible for the prooxidant effect.

The impact of the DOPC reverse micelles on the effectiveness and physical location

of the antioxidants, α-tocopherol and Trolox was also studied. Both non-polar (α-

tocopherol) and polar (Trolox) were able to inhibit lipid oxidation in stripped soybean oil

in the presence of DOPC reverse micelles. Trolox was a more effective antioxidant than

α-tocopherol. Fluorescence steady state and lifetime decay studies suggested that both α-

tocopherol and Trolox were associated with DOPC reverse micelle in bulk oil. Trolox

primarily concentrated in the water pool of reverse micelle since it quenched NBD-PE

fluorescence intensity with increasing concentrations. A portion of α-tocopherol was also

associated with the aqueous phase of the DOPC reverse micelles but this was likely at the

oil-water interface since -tocopherol is not water soluble.

The addition of ferric chelator, deferoxamine (DFO) to stripped soybean oil

significantly prevented the lipid oxidation caused by DOPC reverse micelles as the lag

phase was extended from 2 to 7 days. DFO was also found to increase the antioxidant

activity of both Trolox and -tocopherol. Trolox and -tocopherol were found to be

rapidly decomposed by high-valence Fe(III) while low-valence-state (Fe (II) was much

less reactive. Fe(III) was also consumed by both hydrophilic Trolox and lipophilic α-

Page 9: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

viii

tocopherol presumably though reduction to Fe (II). DOPC reverse micelles were able to

decrease antioxidants-iron interactions as evidence by a decrease in antioxidant depletion

by iron and a decrease in iron reduction by the antioxidants. These results suggested that

the ability of DFO to increase the antioxidant activity of -tocopherol and Trolox was

due to its ability to decrease free radical production and not its ability to decrease direct

iron-antioxidant interactions.

Overall, the results presented in this dissertation show phospholipids and water can

form reverse micelles in edible oils. These reverse micelles increase lipid oxidation rates

by increasing the prooxidant activity of iron. Free radical scavenging antioxidants can

inhibit oxidation promoted by the reverse micelles with polar Trolox being more effective

than non-polar tocopherol presumably because Trolox is more highly associated with

the reverse micelle. The reverse micelles produced by DOPC protected tocopherol and

Trolox from direct degradation by iron. The knowledge gained from this study will

improve our understanding of the mechanism of lipid oxidation in bulk oils which will

hopefully provide new technologies to improve the oxidation stability of edible oils. For

example, it may be able to use oil refining technologies to remove prooxidative minor

components that for physical structure in bulk oils.

Page 10: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

ix

TABLE OF CONTENTS

Page ACKNOWLEDGMENTS ....................................................................................................... v ABSTRACT ............................................................................................................................ vi LIST OF TABLES ................................................................................................................ xiii LIST OF FIGURES............................................................................................................... xiv CHAPER 1. INTRODUCTION ....................................................................................................... 1 2. LITERATURE REVIEW ............................................................................................ 4

2.1 Oil refining procedure ........................................................................................... 4

2.1.1 Storage .................................................................................................... 4 2.1.2 Preparation ............................................................................................. 5 2.1.3 Solvent extraction .................................................................................. 5 2.1.4 Oil refining ............................................................................................. 6

2.2 Minor components in bulk oil............................................................................... 8

2.2.1 Free fatty acid ......................................................................................... 8 2.2.2 Monoacylglycerols and Diacylglycerols ............................................ 11 2.2.3 Phospholipids ....................................................................................... 13 2.2.4 Tocopherol ............................................................................................ 14 2.2.5 Pigments ............................................................................................... 15 2.2.6 Miscellaneous compounds .................................................................. 15

2.2.6.1 Nonsaponifiable minor components ................................... 15 2.2.6.2 Polar triacylglycerol polymers ............................................ 16 2.2.6.3 Oxidized triacylglycerols ..................................................... 17

2.2.7 Water ..................................................................................................... 18 2.2.8 Oxygen ................................................................................................. 19

2.3 The effect of minor components on bulk oil oxidation ..................................... 21

2.3.1 The effect of oxygen on the chemical stability of bulk oil ................ 21 2.3.2 The effect of water on the chemical stability of bulk oil ................... 22 2.3.3 The effect of free fatty acid on the chemical stability of bulk

oil ....................................................................................................... 23 2.3.4 The effect of MAG and DAG on chemical stability of bulk

Page 11: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

x

oil ....................................................................................................... 24 2.3.5 The effect of phospholipids on the chemical stability of bulk

oil ....................................................................................................... 24 2.3.6 The effect of tocopherols on the chemical stability of bulk oil......... 25 2.3.7 The effect of pigments on the chemical stability of bulk oil ............. 26 2.3.8 The effect of miscellaneous compounds on the chemical

stability of bulk oil ............................................................................ 28 2.4 Association colloids and lipid oxidation in bulk oils ........................................ 29

2.4.1 Evidence for the presence of association colloids in bulk oil ........... 29 2.4.2 Do association colloids impact lipid oxidation? ................................ 30 2.4.3 The perspective researches on the mechanism of lipid

oxidation in bulk oil .......................................................................... 32 2.5 Methodologies to study the impact of minor components on the

physicochemical properties of bulk oil ........................................................ 33 2.5.1 Chemical analysis ................................................................................ 34

2.5.1.1 Primary lipid oxidation products ......................................... 34 2.5.1.2 Secondary lipid oxidation products ..................................... 34

2.5.2 Physical analysis .................................................................................. 35

2.5.2.1 Small/wide angle x-ray scattering (SAXS/WAXS)............ 35 2.5.2.2 Steady-state and time resolved fluorescence

measuremen .......................................................................... 36 2.5.2.3 Interfacial tension analysis .................................................. 37

3. THE IMPACT OF PHYSICAL STRUCTURES IN SOYBEAN OIL ON

LIPID OXIDATION .................................................................................................. 38 3.1 Introduction .......................................................................................................... 38 3.2 Materials and Methods ........................................................................................ 39

3.2.1 Materials .............................................................................................. 40 3.2.2 Preparation of stripped soybean oil .................................................... 40 3.2.3 Light scattering properties of oil samples ......................................... 41 3.2.4 Determination of critical micelle concentration (CMC) in

stripped soybean oil .......................................................................... 42 3.2.5 Small angle X-ray scattering (SAXS) ............................................... 43 3.2.6 Front-Face (FF) fluorimetric measurements ...................................... 44 3.2.7 Measurement of oxidation parameters ............................................... 45 3.2.8 Statistical analysis ................................................................................ 46

3.3 Results and discussion......................................................................................... 46

Page 12: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xi

3.3.1 Attenuation of the water/phospholipid/MCT system ......................... 46 3.3.2 Determination of the CMC of phospholipids in SSO and

MCT................................................................................................... 47 3.3.3 Detection and characterization of phospholipids physical

structures by small angle x-ray scattering ....................................... 51 3.3.4 Effect of water/phospholipid molar ratio on the physical

properties of DOPC association colloids ........................................ 54 3.3.5 Oxidation of stripped soybean oil containing association

colloids .............................................................................................. 58

4. ROLE OF REVERSE MICELLES ON LIPID OXIDATION IN BULK OILS: IMPACT OF PHOSPHOLIPIDS ON ANTIOXIDANT ACTIVITY OF α-TOCOPHEROL AND TROLOX .................................................................... 61 4.1 Introduction .......................................................................................................... 61 4.2 Materials and Methods ........................................................................................ 63

4.2.1 Materials .............................................................................................. 63 4.2.2 Preparation of stripped soybean oil .................................................... 64 4.2.3 Formation of DOPC reverse micelles in stripped soybean oil .......... 64 4.2.4 Measurement of oxidation parameters ............................................... 65 4.2.5 Small angle x-ray scattering (SAXS) study of bulk oil with

DOPC and antioxidants .................................................................... 66 4.2.6 Fluorescence measurement of bulk oil with DOPC and

antioxidants ....................................................................................... 67 4.2.7 Statistical analysis ................................................................................ 68

4.3 Results .................................................................................................................. 69

4.3.1 Impact of phospholipids on oxidative stability of SSO ..................... 69 4.3.2 The impact of phospholipids on the oxidative stability of

SSO in the presence of α-tocopherol ............................................... 71 4.3.3 The impact of phospholipids on the oxidative stability of

SSO in the presence of Trolox ......................................................... 75 4.3.4 The impact of antioxidants on the properties of DOPC

reverse micelles in SSO .................................................................... 79

4.4 Discussion ............................................................................................................ 83

5. IRON/ANTIOXIDANTS RELATIONSHIP IN BULK OIL OXIDATION .......................................................................................................... 87 5.1 Introduction .......................................................................................................... 87 5.2 Materials and Methods ........................................................................................ 89

5.2.1 Materials .............................................................................................. 89

Page 13: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xii

5.2.2 Formation of DOPC association colloids in stripped soybean oil or MCT......................................................................................... 89

5.2.3 Measurement of oxidation parameters ............................................... 90 5.2.4 Determination of Trolox and α-tocopherol concentrations ............... 91 5.2.5 Determination the concentration of ferric iron in MCT .................... 92 5.2.6 EPR spectroscopy ................................................................................ 92 5.2.7 Statistical analysis ................................................................................ 93

5.3 Results .................................................................................................................. 93

5.3.1 Effects of the metal chelator, Deferoxamine (DFO) on the

oxidative stability of stripped soybean oil ...................................... 93 5.3.2 Effects of Fe(III) and Fe(II) on the depletion of α-tocopherol

and Trolox in MCT ......................................................................... 100 5.3.3 Effects of Fe(III) and Fe(II) on the depletion of α-tocopherol

or Trolox in the presence of DOPC reverse micelles ................... 102 5.3.4 Measurement of Fe(III) loss by α-tocopherol and Trolox ............... 105 5.3.5 Effects of Fe(III) and Fe(II) on the formation of PBN spin

adducts in MCT .............................................................................. 109

5.4 Discussion .......................................................................................................... 110

6. CONCLUSION ................................................................................................... 113 BIBLIOGRAPHY ................................................................................................................ 117

Page 14: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xiii

LIST OF TABLES Table Page

2.1. Free fatty acid content in crude soybean oil as the function of bean storage time ......................................................................................................................... 9

2.2. Free fatty acid content in soybean oil as a function of processing step ................... 9

2.3. Free fatty acid content in retail oil ............................................................................ 11

2.4. Concentration of MAG and DAG in various oils as a function of refining ........... 12

2.5. Phospholipid composition (%) as extracted from commercial seeds ..................... 13

2.6. Phosphorous content (ppm) of soybean oil as a function of refining step ............. 14

2.7. Water content in freshly-opened, commercially available vegetable oils as determined by Karl Fischer Coulometer ............................................................ 18

2.8. Solubility of oxygen (ppm) in Marine oil ................................................................ 20

Page 15: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xiv

LIST OF FIGURES Figure Page

2.1. The schematic of oil refining process ......................................................................... 6

2.2. Schematic representation of the association colloids formed by minor constituents in bulk oil (A) a reverse micelle formed by oleic acid; (B) a reverse micelle formed by phospholipids; (C) Inverse lamella structure formed by monoacylglycerol; (D) a mixed reverse micelle formed by free fatty acids and phospholipids .................................................... 30

2.3. The essential parts of a small angle scattering system. The drawing shows the X-ray source X, the sample S, the scattering angle θ, the slits used to define the incident and scattered beams, and the detector D ........................ 35

3.1. Molecular structures of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) ............................... 39

3.2. Attenuation of medium chain triacylglycerols (MCT) samples containing 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and varying concentrations of water ....................................................................................... 47

3.3. Determination of critical micelle concentration (CMC) of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) in medium chain triacylglycerols (A)and stripped soybean oil (B) as determined by the absorbance (480 nm) of 7,7,8,8-tetracyano- quinodimethane....................................................... 49

3.4. Determination of critical micelle concentration of 1,2-dioleoyl-sn-glycero-3-phosphocholine in stripped soybean oil using interfacial tension ................ 51

3.5. Small angle x-ray scattering profiles of stripped soybean oil (SSO) and SSO + 950 μM 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3- phosphocholine (DC4PC) ..................................... 52

3.6. The pair distribution functions, p(r), calculated from the scattering profiles of stripped soybean oil (SSO) and SSO + 950 μM 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine(DC4PC) using GNOM analysis of the three small angle x-ray scattering profiles indicated in Figure 5. ........................................ 53

3.7. Small angle x-ray scattering profiles of stripped soybean oil with different water/1,2-dioleoyl-sn-glycero-3-phosphocholine molar ratios (WO). Inserted figures are the 2 dimensional small angle x-ray scattering pattern of each sample ......................................................................................... 55

Page 16: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xv

3.8. A postulated structural change of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) as phospholipid concentrations are increased to above the critical micelle concentrations (CMC) form a reverse micelle and then the water:molar ratio is increased to form an anisotropic structure with a non-spherical shape ............................................... 57

3.9. The fluorescence intensity of N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)-1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine, triethylammonium salt (NBD-PE) in stripped soybean oil with different water/1,2-dioleoyl-sn-glycero-3-phosphocholine molar ratios (WO). Inserted figure is the molecular structure of (NBD-PE) ....................... 58

3.10. Formation of lipid hydroperoxides (A), propanal (B) in stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine(DOPC) or 1,2-dibutyryl-sn-glycero-3- phosphocholine (DC4PC) and 200 ppm water at 37 °C. Data represent means (n =3) ± standard deviations. Some error bars are within data points .................................................................................................................... 59

4.1. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) during storage at 55 °C. Some of the error bars are within data points .................................................................................. 70

4.2. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) with or with 10 or100 μM α-tocopherol during storage at 55 °C. Some of the error bars are within data points ........... 74

4.3. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 μM of 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) with or with 10 or100 μM α-tocopherol during storage at 55 °C. Some of the error bars are within data points ........... 75

4.4. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) with or with 10 or 100 μM Trolox during storage at 55 °C. Some of the error bars are within data points ....................... 77

4.5. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 μM of 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) with or with 10 or100 μM Trolox during storage at 55 °C. Some of the error bars are within data points ....................... 78

Page 17: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xvi

4.6. The neutralized fluorescence intensity of NBD-PE at stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) with varying concentrations(0-100 μM) of α-tocopherol and Trolox ................................................................................................................... 81

4.7. The fluorescence lifetime decay of NBD-PE at stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) with varying concentrations(0-100 μM) of α-tocopherol (A) and Trolox (B) concentrations ............................................................................ 82

5.1. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 2 mM deferoxamine (DFO) and/or 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) during storage at 55 °C. Some of the error bars are within data points ........................................ 95

5.2. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 100 M α-tocopherol, 2 mM deferoxamine (DFO) and/or 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) during storage at 55 °C. Some of the error bars are within data points ............................................................................................................ 97

5.3. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 100 M Trolox, 2 mM deferoxamine (DFO) and/or 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) during storage at 55 °C. Some of the error bars are within data points ........... 98

5.4. Depletion of α-tocopherol and Trolox (100 M) in medium chain triacylglycerols (MCT) in the presence of (A) Fe(III); and (B)Fe(II)............ 101

5.5. Depletion of α-tocopherol and Trolox (100 M) in medium chain triacylglycerols (MCT) containing 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) in the presence of (A) 40 ppm Fe(III); and (B) 100 ppm Fe(III) ........................................................................................... 103

5.6. Depletion of α-tocopherol and Trolox (100 M) in medium chain triacylglycerols (MCT) containing 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) in the presence of (A) 40 ppm Fe(II); and (B)100 ppm Fe(II) ............................................................................................. 105

Page 18: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xvii

5.7. Fe(III) development in medium chain triacylglycerols (MCT) containing (A) α-tocopherol or Trolox; (B) 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC); (C) 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) after 24 h storage at room temperature. For each group, different letters on the top of columns represent significant differences (p < 0.05) ........................................................................................ 109

5.8. EPR detection of radical formations in the reaction consisting of (A) 0.1 M PBN in stripped soybean oil; (B) 0.1 M PBN in MCT; (C) 0.1 M PBN+100 μM α-tocopherol in MCT; (D) 0.1 M PBN+100 μM Trolox in MCT; (E) 0.1 M PBN+10 mM cumene hydroperoxides+40 ppm Fe(II) in MCT; (F) 0.1 M PBN+cumene hydroperoxides+40 ppm Fe(III) in MCT after 24 h storage at 55 oC ...................................................... 110

6.1. Proposed mechanism of bulk oil oxidation that contains reverse micelles ......... 115

LIST OF ABBREVIATIONS

Page 19: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xviii

BHA: Butylated hydroxyanisole

BHT: Butylated hydroxytoluene

CA: Calcein

CMC: Critical micelle concentration

DAG: Diacylglycerols

DC4PC: 1,2-dibutyl-sn-glycerol-3-phosphocholine

DFO: Deferoxamine

DOPC: 1,2-dioleoyl-sn-glycerol-3-phosphocholine

EPR: Electron paramagnetic resonance

FFA: Free fatty acid

MAG: Monoacylglycerols

MCT: Medium chain triacylglycerols

NBD-PE: N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)-1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine, triethylammonium salt

PAHs: Polycyclic aromatic hydrocarbons

PBN: N-t-butyl-phenylnitrone

PC: Phosphatidylcholine

PE: Phosphatidylethanolamine

PI: Phosphatidylinosital

PLs: Phospholipids

PS: Phosphatidylserine

SAXS: Small angle X-ray scattering

SSO: Stripped soybean oils

TAG: Triacylglycerols

Page 20: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

xix

TBHQ: tert-Butylhydroquinone

TCNQ: 7,7,8,8-tetracyano-quinodimethane

TGP: Triacylglycerol polymers

TROLOX: 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid

Page 21: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

1

CHAPTER 1

INTRODUCTION

Food oils from a wide variety of sources are important to human health and the food

industry since they are important components in a large variety of food products and

since they are an important cooking medium (1). However, most oils are susceptible to

oxidation as a result of their unsaturated fatty acids (2). The products formed in oxidized

oil include numerous free radical species, primary oxidation products like lipid

hydroperoxides and secondary oxidation products like aldehydes, hydrocarbons, ketones

and epoxides that negatively impact aroma (3). In addition, lipid oxidation can produce

toxic products that can adversely impact biological tissues. For instance, linoleic acid

hydroperoxides are toxic to wild type Saccharomyces cerevisiae at a low levels (0.2 mM)

(4). Oxidized palm oil has been shown to induce reproductive toxicity and organotoxicity

of the kidneys, lungs, livers, and heart of rats (5). High intake of a mixture of oxidized

lard and cod liver oil caused impaired fertility in female rats and an increased incidence

of morphologically abnormal spermatozoa in male rats (6). Oxidized lard, soybean oil,

and particularly sardine oil increased spontaneous liver tumor development and the

formation of 8-hydroxy-deoxyguanosine (8-OH-dG) in the liver DNA of mice (7). These

results along with others suggest that consumption of oxidized lipids should be avoided

whenever possible.

How to prevent or retard lipid oxidation in food oil is a major focus of lipid research.

In the past six decades, a lot of attention has been drawn on finding effective ways to

extend the shelf life of oils. The addition of antioxidants is one effective way to retard

lipid oxidation (8). The most widely used antioxidants are free radical scavengers that

Page 22: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

2

inactive free radicals formed in the initiation and propagation steps of lipid oxidation. A

number of natural or synthetic phenols can interact with hydroperoxyl (LOO) and

alkoxyl (LO) radicals, producing hydroperoxides and alcohols, respectively, and low

energy antioxidant radicals that do not readily promote the oxidation of unsaturated fatty

acids (3).

Among the free radical scavenging antioxidants, synthetic phenolic antioxidants

[e.g., butylated hydroxyanisole (BHA), butylated hydroxytoluene (BHT) and tert-

Butylhydroquinone (TBHQ) normally exhibit higher antioxidative activity than natural

antioxidants. However, the utilization of many effective synthetic antioxidants is being

limited by consumer concerns of their negative impact on health. Research showed that

large doses of both BHA and BHT (500 mg/kg body weight/day) result in certain

pathological, enzyme, and lipid alterations in both rodents and monkeys, and in some

cases BHT has been reported to have teratogenic and carcinogenic effects in rodents (9).

Over the past several decades, food scientists have been mostly unsuccessful in their

search for natural antioxidants which could possess similar antioxidant activity as

synthetic antioxidants (10). The most common natural antioxidants in foods are

tocopherol and rosemary extracts. Interestingly, rosemary extracts are not actually

approved by the FDA for use as an antioxidant but instead as a flavor extract (FDA 21

CFR 182.(11)). Since the number of antioxidant options has not significantly increased in

the past 30 years and in fact may be decreasing since food companies are reluctant to

utilize synthetic antioxidants, new approaches are needed to fulfill the goal of retarding

oil and foods from oxidation. The development of novel approaches to inhibiting lipid

oxidation is highly reliant on our understanding of lipid oxidation mechanisms in variable

Page 23: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

3

lipid systems.

The mechanism of lipid oxidation in bulk oil has been studied for many decades (12).

However, research has revealed that this reaction is extremely complex being a balance of

the activity of numerous prooxidative and antioxidative factors (12). On top of these

chemical effects, lipid oxidation is also impacted by physical properties of food and food

components that influence factors such as partitioning of antioxidants, diffusion of

oxygen and interaction of prooxidants with lipid substrates. The impact of physical

properties on lipid oxidation can be seen in heterogeneous food systems such as

emulsions where the characteristics of the emulsion droplet interface are an important

determinant in oxidative stability (13). Even bulk oils, which are often assumed to be a

homogenous liquid, have numerous physical structures that impact lipid oxidation. This

is because refined bulk oils contain numerous minor components that are amphiphilic,

such as monoacylglycerols (MAG), diacylglycerols (DAG), phospholipids, sterols, free

fatty acids (FFA), and polar products arising from lipid oxidation, such as lipid

hydroperoxides, aldehydes, ketones, and epoxides. These surface active compounds in

combination with water can form physical structures known as association colloids that

can physically impact lipid oxidation (14).

The objective of this research is therefore to better understand how these association

colloids impact the lipid oxidation in bulk oil. Besides, how the interactions between the

minor components naturally from bulk oils and association colloids could impact the lipid

oxidation of bulk oil will be investigated. By doing so, the results can not only enhance

the oxidative stability of bulk oil by avoiding the prooxidative minor components during

refining process, but also improve the antioxidative ability of antioxidants in the bulk oil.

Page 24: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

4

CHAPTER 2

LITERATURE REVIEW

2.1 Oil refining procedure

The progress of oil refining has been progressed from emphasizing one operation

independently in the early days toward the current integrating manufacturing facilities to

produce a more complete range of value-added products from the raw plant seeds to the

final products. Before getting the final oils, there are several procedures need to be

carried out, i.e., storage, preparation, mechanical extraction, solvent extraction,

degumming, deodorization, bleaching, etc.

2.1.1 Storage

The typical operation associated with storage includes receiving, sampling, drying,

storage, and cleaning prior to processing. All these procedures are to make sure the seeds

are suitable to extract oil. The following description is based on the industrial soybean oil

extraction and refining processing. Initially, soybeans received are sampled and analyzed

for moisture content, foreign matter, and damaged seeds. Then the seeds are weighed and

conveyed to large silo or metal tanks for storage prior to processing. When the facility is

ready to process the seeds, the seeds are removed from the containers and cleaned of

foreign materials and loose hulls if necessary. Screens typically are used to remove

foreign materials such as sticks, stems, pods, tramp metal, etc. An aspiration system is

used to remove loose hulls from the seeds. The seeds are passed through dryers to reduce

their moisture content to approximately 10 to 11 wt % and then are conveyed to process

bins for temporary storage and tempering for 1 to 5 days in order to facilitate dehulling.

Page 25: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

5

2.1.2 Preparation

The preparation processing is fairly well standardized consisting four principal

operations: cracking, dehulling/hull removal, conditioning, and flaking. Soybeans are

conveyed from the process bins to the mill by means of belts or mass flow conveyors and

bucket elevators. In the mill, the beans may be aspirated again, weighed, cleaned of tramp

metal by magnets, and fed into corrugated cracking rolls. The cracking rolls “crack” each

bean into four to six particles, which are passed through aspirators to remove the hulls.

Then, the cracked beans and bean chips are conveyed to the conditioning area, where

they are put either into a rotary steam tubed device or into a stacked cooker and are

heated to “condition” them (i.e., make them pliable and keep them hydrated).

Conditioning is necessary to permit the flaking of the chips and to prevent their being

broken into smaller particles. Finally, the heated, cracked beans are conveyed and fed to

smooth, cylindrical rolls that press the particles into smooth “flakes”, which vary in

thickness from approximately 0.25 to 0.51 mm. Soybean flakes allow the soybean oil

cells to be exposed and the oil to be more easily extracted.

2.1.3 Solvent extraction

The extraction process consists of “washing” the oil from the soybean flakes with

hexane solvent in a countercurrent extractor. Then the solvent is evaporated from both the

solvent/oil mixture and the solvent-laden, defatted flakes. The solvent in oil is removed

by exposing the solvent/oil mixture to steam (contact and noncontact). Then the solvent is

condensed, separated from the steam condensate, and reused. Residual hexane not

Page 26: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

6

condensed is removed with mineral oil scrubbers. The solvent removed oil, called “crude”

soybean oil, is stored for further processing or transportation.

2.1.4 Oil refining

Crude vegetable oils predominantly contain triacylglycerols and small amounts of

minor components that naturally occur in the plant component of their origin, such as

proteins, free fatty acids, phospholipids, metals, tocopherols, pigments and sterols. The

effective removal some of those minor components without sacrificing the loss of

antioxidants is necessary to achieve a finished oil quality with acceptable standards for

flavor, appearance and stability. The oil refining process which includes degumming,

neutralization, bleaching, and deodorization is therefore performed to produce oils

suitable for use by the food industry and the consumer (Figure 2.1).

Figure 2.1. The schematic of oil refining process

Oil seed extraction will produce a crude oil that contains lipids such as

phospholipids and sterols in addition to triacylglycerols. However, the extraction process

will also produce conditions where triacylglycerols can react with enzymes such as lipase

Page 27: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

7

and lipoxygenase to form hydrolytic products of triacylglycerols (e.g. monoacylglycerols,

diacylglycerols and free fatty acids) and lipid oxidation products such as hydroperoxides.

Oil refining is performed to reduce the concentration of these minor components as they

can negatively impact the quality of the oil. For instance, caustic alkali is used to remove

free fatty acids in neutralization step since free fatty acids cause foaming and decreases

the smoke point of oils (15). Phosphoric or citric acid is used to aid the removal of

nonhydratable and hydratable phospholipids during the degumming step since

phospholipids cause cloudiness and can form brown colors during heating (16).

Bleaching is performed with activated bleaching earth at relatively high temperatures

around 100°C to decrease the concentration of pigments (e.g. chlorophyll), trace metals

and other polar compounds that can accelerate lipid oxidation. Finally, since the harsh

conditions of these steps can produce off-flavors due to oxidation and since the oil can

contain pesticides from the original oilseeds, the oils undergoes a deodorization step at

high temperatures (200-260°C) under reduced pressure. Deodorization processing under

high temperature not only removes volatile off-flavors and pesticides but also

decomposes lipid hydroperoxides, a potential lipid oxidation substrate. Overall, refining

is designed to remove minor components that negatively impact oil quality. However,

steps such as deodorization can also remove factors that improve oil quality such as

tocopherols and can cause the formation of trans fatty acids. While refining significantly

decreases the concentration of the minor components that impact the organoleptic quality

of oil, one should realize that even at these lower concentrations, minor components can

still impact oil chemistry such as lipid oxidation.

Page 28: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

8

2.2 Minor components in bulk oil 2.2.1 Free Fatty Acid

The reaction between water and triacylglycerols (TAG) results in the formation of

free fatty acid (FFA) and diacylglycerol (DAG). TAG hydrolysis is accelerated by lipases,

extremes in pH and heat (17). The presence of FFA produces undesirable flavors,

foaming during mixing and heating and causes a decrease in the smoke/flash point of oil

thus reducing the maximum temperature to which they can be heated. FFAs in crude

vegetable oils will increase with the age of the oilseeds and can be abnormally high if the

oilseeds have been damaged or improperly stored. This is due to damage to cells in the

seeds allowing lipase to interact with TAG. For instance, in oilseeds stored from 0 to 47

days, the FFA content in the extracted crude oil increased from 0.88 to1.80 wt % (18).

Recent research from Alencar and coworkers showed that FFA concentrations will also

increase during crude oil storage (19).

There are several factors which can impact FFA concentrations in crude oils and

during the refining process, such as oilseeds storage time before extraction (Table 2.1),

the initial FFA content, water content and processing temperatures (Table 2.2), oil type,

etc. The presence of FFA can catalyze the further hydrolysis of TAG, thereby increasing

the total FFA concentration in oil.

Page 29: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

9

Table 2.1 Free fatty acid content in crude soybean oil as the function of bean storage time

Bean storage time(d) FFA(%)

5 0.88

11 0.60

22 1.28

35 1.40

47 1.80

The moisture in soybean seeds is 13-15 %

Table 2.2 Free fatty acid content in soybean oil as a function of processing step

Processing step FFA (%) Crude 0.61 Degummed 0.31 Refined 0.05 Deodorized 0.02

Some earlier results showed the rate of TAG hydrolysis is proportional to the initial

level of FFA (20). Another important factor is the water concentration since this is one of

the required substrates for the hydrolytic reaction. Sarkadi and coworkers showed that

increasing of water content as a result of steam deodorization of peanut oil at 180oC

increased FFA formation (21). Considering that both FFA and water are important factors

in TAG hydrolysis, the following equation was used to estimate the formation of FFA

over time (t) in the bulk oils containing water (20).

2

2

[ ][ ]' [ ][ ]Sat

H Od FFAk FFA

dt H O

(1)

Here, k’ is the rate constant, [FFA] is the free fatty acid concentration, [H2O] is the water

concentration, [H2O]Sat is the water concentration at saturation.

The FFA content in deodorized food oils is required to less than 0.05 % in the

United States (22, 23). FFA are typically removed from oils by chemical neutralization

Page 30: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

10

which normally involves the conventional alkali (sodium hydroxide) neutralization

process, precipitating the FFA as soap stock and being removed by mechanical separation

from the oil. Recently, some oil refineries are using physical refining. Physical refining

is a high temperature, vacuum process that can remove both FFA and phospholipids at the

same time thus effectively combining the neutralization and degumming steps. The

choice of physical or chemical refining is highly correlated to the initial FFA

concentration of the crude oil. When the FFA content is lower than 2.5 % in the crude oil,

physical refining can be employed without the neutralization step. In some cases physical

refining is used on oils with > 2.5 % FFA but neutralization must be performed first (12).

A major limitation of physical refining is that the high temperatures required can cause

the formation of trans fatty acids (24).

FFA are produced at almost each step of refining since they often involve high

temperatures and water (25). Despite this FFA concentrations as low as 0.005 wt % can

be obtained (26). However, one should realize that these FFA concentrations cannot be

maintained as even low FFA concentration can promote TAG hydrolysis. Therefore, the

FFA content of retail bottled edible oil is higher than the concentrations measured

immediately after refining process (Table 2.3). For example, retail canola oil has a FFA

concentration of approximately 0.1 wt%. Finally, free fatty acid concentrations can

increase in refined oils during operations such as frying where the oils are exposed to

high temperatures and water (27).

Page 31: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

11

Table 2.3 Free fatty acid content in retail oil

Oil type Free Fat acids (wt %)

Canola 0.1

Coconut 3.93

Corn 0.12

Olive 0.30

Extra virgin olive 0.28

Peanut 0.32

Palm kernel 2.49

Palm olein 0.15

Rapeseed 0.04

Rice bran 0.08

Safflower 0.03

Sesame 2.37

Soybean 0.05

Sunflower 0.09

2.2.2 Monoacylglycerols and Diacylglycerols

Monoacylglycerols (MAG) are monoester of glycerol in which one of the hydroxyl

groups is esterified with fatty acids. There are two isomeric forms of MAG, i.e., 1-MAG

and 2-MAG, depending on the position of the ester bond on the glycerol group.

Diacylglycerols (DAG) are esters of the glycerol in which two of the hydroxyl groups are

esterified with fatty acids. They can exist in three structural isomers namely, 1,2-DAG,

2,3-DAG and 1,3-DAG (28). DAG is a common precursor for the synthesis of both TAG

and phospholipids in the oilseeds and oil bodies (28). Therefore, DAG that exists

naturally in crude vegetable oils can be derived from the incomplete biosynthesis of TAG

Page 32: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

12

and phospholipids. Knowledge of the amounts of DAG naturally found in oilseeds is

scare due to the lack of in situ measurements.

MAG and DAG are also formed from the partial hydrolysis of TAG (29). The

amount of DAG and MAG reported in the crude oil are therefore composed by two parts,

inherent concentrations in the seeds and those formed after crushing the seeds and during

refining. The formation of DAG and MAG depends on oilseeds storage conditions,

temperature, and moisture as was discussed previously for FFA.

Table 2.4. Concentration of MAG and DAG in various oils as a function of refining (30, 31).

Refining

step

Soybean oil Palm oil

MAG(%) DAG(%) MAG(%) DAG(%)

Crude oil 0.11 1.10 0.26 6.60

Degummed 0.10 1.44 0.17 6.70

Bleached 0.06 1.25 0.17 6.70

Deodorized 0.07 1.05 0.08 6.9

Table 2.4 lists the amount of DAG and MAG in crude and refined soybean oil and

palm oil. After bleaching and deodorization, MAG concentrations decease while DAG

concentrations remain similar to crude oil or even increase (30-32). Farhoosh and

coworkers recently showed the DAG concentration in crude soybean oil (1.66 %)

decreased to 1.35 % after the neutralization step. However, after deodorization, the DAG

concentrations (1.74 %) increased and reached around the initial amount in the crude

soybean oils (25). They also observed the similar trend in the canola oil. Possible reasons

for the increased in DAG in these reports is the existence of lipase in the crude oil which

could form DAG or from TAG hydrolysis by high temperatures and water in the

deodorization step (33).

Page 33: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

13

2.2.3 Phospholipids

Phospholipids (PLs) are integral elements of all cellular membranes in living

organism and possess unique chemical structures containing both lipophilic and

hydrophilic groups. Crude oil contains substantially high amount of PLs due to the

solvent extraction of cell membranes. Phospholipids in oils include phosphatidylcholine

(PC), phosphatidylethanolamine (PE), phosphatidylserine (PS), phosphatidylinosital (PI).

Different species of seeds have different compositions of phospholipids, but in general

PC is predominant (Table 2.5).

Table 2.5 Phospholipid composition (%) as extracted from commercial seeds

Soybean Sunflower Corn Rapeseed PC 21.9 25.4 30.4 24.6 PE 13.6 11.0 3.2 22.1 PI 12.0 19.4 16.3 14.7

The amount of phospholipids in the oilseeds is small comparing to animal tissues

(34). PLs in food oils cause foaming, cloudiness and darkening during thermal processing.

Normally, there are two types of PLs in crude oil, hydratable and non-hydratable. The

degumming step removes hydratable PLs by washing them into the water phase. Non-

hydratable PLs are not affected by water washing alone so citric or phosphoric acid are

often added to the wash water to bind divalent cations thus making non-hydratable PLs

hydratable. Physical refining can also be used to remove phospholipids as discussed

above. Like other minor components, only trace amount of PLs will remain in the final

refined oil (Table 2.6) with concentrations ranging from 20-2000 ppm with variations

among different oilseeds species and refining methods (35, 36).

Table 2.6 phosphorous content (ppm) of soybean oil as a function of refining step

Page 34: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

14

Refining step Soybean Soybean palm Soybean corn

Physical Chemical Physical Chemical

Crude 600 510 9.1 327.8 33.2 109.7 7.5 Degummed 58 120 0.7 5.7 0.8 1.24 0.6 Refined 13.1 1 0.7 1.7 0.3 0.3 0.1 Deodorized 0.16 1 0.5 0.2 0.3 0.3 0.1

2.2.4 Tocopherols

Tocopherols are the major oil soluble vitamin in crude oil. When the oilseeds are

crushed, the tocopherols are extracted along with the oil. Natural tocopherols are

mixtures consisting of tocopherols. The volatility of tocopherols is above

free fatty acids but below triacylglycerols. Thus, during the refining processing,

tocopherols are partially removed during deodorization thus decreasing their

concentrations in the refined oil. Gogolewski and coworkers found that after the refining

of rapeseed oil the tocopherol losses amounted to 30 %. Two thirds of the loss resulted

from distillation and thermal degradation during deodorization, while one third was

caused by the combined effects of neutralization and bleaching (37). Rossi and coworkers

reported that steam deodorization removed 200-300 ppm of tocopherols out of the

original ~ 1000 ppm in crude palm oil (38).

The initial concentration and type of tocopherols in bulk oil vary with different

oilseeds species. For instance, soybean oil has a high tocopherol concentrations with

more than 65% -tocopherols and 20% -tocopherols while there is almost no -

tocopherols in grape seeds oil (39). During storage tocopherol concentrations decrease

which is accompanied by the generation of oxidized tocopherols. This is because

tocopherols act as hydrogen donors and react with free radicals.

Page 35: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

15

2.2.5 Pigments

The intensity of the bulk oils color mainly depends on the presence of coloring

pigments such as carotenoids and chlorophyll. Vegetable oils with a minimum color

index are considered to be more suitable for edible and industrial purposes (40). Thus,

pigments in crude oil are partially removed during the degumming, refining, and

bleaching steps of the refining process.

Carotenoids are pigments in many vegetable oils and particularly in palm oil. They

contain a long conjugated polyene chains and are yellow/orange/red in color. Crude palm

oil normally contains 500-700 ppm of carotenes. These are mainly -carotene (24-42 %

of total carotene) and -carotene (50-60 %) along with low levels of several other

carotenes (41).

Chlorophylls are also found in vegetable oils. Chlorophyll includes chlorophyll a,

chlorophyll b, and the Mg-free chlorophyll derivatives pheophytin a and pheophytin b.

Some studies have reported no detectable chlorophylls in soybean oil after deodorization

(42). However, other groups showed the existence of chlorophyll and pheophytins in

refined oils (43). Pheophytin was reported to be the predominant chlorophyll, being about

0.06 and 0.10 ppm in refined soybean oil and corn oil, respectively (43).

2.2.6 Miscellaneous compounds

2.2.6.1 Nonsaponifiable minor components

Nonsaponifiable minor components in food oils include sterols, polycyclic aromatic

hydrocarbons (PAHs) and hydrocarbons (e.g. alkanes, alkenes and squalene)(44). Most

vegetable oils contain 700-1100 mg/100g of sterols, partly as free and partly as esterified

Page 36: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

16

sterols. High levels are present in rapeseed oil and in corn oil (45). Sitosterol is generally

the major phytosterol (50-80% of total sterol) with campesterol, stigmasterol, and Δ5-

avenasterol also frequently attaining significant levels. Brassicasterol is virtually absent

from the major seed oils except for rapeseed oil where it comprises 10 % of the total

sterols.

Food oils can be contaminated with PAHs due to the wide distribution of PAHs in

the environment, their lipophilic nature, and the migration from contaminated packaging

materials (46). Sekeroglu and coworkers tested 40 vegetable oil samples from Turkey for

PAHs and found the total PAHs were above 25 ppb in most of the oils samples(47). More

significantly, total PAHs levels in virgin olive oil always were 54.4-110.8 ppb(48).

However, the refined food oils only occasionally exceeded the 25 ppb limit. Recently,

Guillen and coworkers confirmed the occurrence of PAHs during the sunflower oil

oxidation stored in the close container at room temperature for 112 months (49, 50).

Squalene (C30H50) is a highly unsaturated open-chain triterpene. Olive oil is the

major commercial oil with significant amounts of squalene (51). In refined food oils, new

nonsaponifiable compounds are also formed as a consequence of the reactions occurring

during the refining process (52). These compounds include steroidal, arising from the

dehydration of sterols, terpenic from terpenic alcohols, and other compounds deriving

from squalene isomerization.

2.2.6.2 Polar triacylglycerol polymers

Polar triacylglycerol polymers (TGP) are formed during vegetable oil refining (53,

54). TGP are generally formed during the high temperature refining steps, e.g. bleaching

Page 37: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

17

and especially deodorization. An early study on the evolution of TGP in vegetable oils

during physical refining reported a 1-2 wt % increase of TGP concentrations at physical

refining temperatures less than 240 °C. When physical refining temperatures were

increased to 270 °C, a sharp increase in TGP cocnentration to 3.0 wt % was observed.

TGP formation also depends on the degree of saturation of the oils. For more saturated

oils, such as palm oil, TGP increases were restricted to 2 wt %, even at temperatures of

270 °C (55). Gomes used high-performance size exclusion chromatography to quantify

the amount of oligopolymer in refined oils and found that refined olive oil had 0.7 wt %

of oligopolymer twice of that in refined olive pomace oil (56). The final amounts of TGP

found in food oils are generally ~1 wt % depending on the fatty acid composition, initial

quality and refining technology used (57).

2.2.6.3 Oxidized triacylglycerols

Oxidized triacylglycerols (ox-TG), sometimes referred to as core aldehydes, are

generally nonvolatile oxidation products left in the oil after deodorization. Initial reports

suggested that ox-TG concentrations ranged from 4.55 wt % in crude oil to 9.29 wt %

after oil refining (57). However, more recently Hopia investigated the impact of oil

refining on the ox-TG contents in three vegetable oils, i.e., sunflower oil, soybean oil and

low erucic acid rapeseed oil (58). The results showed that the concentration of ox-TG

ranged from 0.53 to 0.90 % oil in the crude oils. After refining, ox-TG concentrations

were decreased in bulk oil and ranged from 0.29 % for sunflower oil to 0.78 wt % for

soybean oil. The loss was attributed to the adsorption of ox-TG by the activated earth

during bleaching. Farhoosh and coworkers found ox-TG concentrations in crude soybean

Page 38: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

18

oil and canola oil at 4.2 and 2.8 wt %, respectively (25). After oil refining, ox-TG

decreased to 2.8 and 2.0 wt % in soybean oil and canola oil, respectively. They also

noticed the level of the ox-TG in the refined oils was proportional to its original level in

the crude oils.

2.2.7 Water

Since water and oil have opposite polarities and they tend to spontaneously phase

separate because of the hydrophobic effect one might not expect to find water in bulk oil

(59). However, as discussed in the proceeding sections, refined oils contain many surface

active compounds such as free fatty acids (FFA), monoacylglycerols (MAG),

diacylglycerols (DAG) and phospholipids (PLs). These surface active compounds have

the ability to emulsify water into bulk oils. The water content in freshly-opened,

commercially available vegetable oils range from 200-2000 ppm (Table 2.7) (14). Water

concentrations in oil can change after opening since oil can absorb water from

atmosphere or water in the oil can evaporate.

Table 2.7. Water content in freshly-opened, commercially available vegetable oils as determined by Karl Fischer Coulometer

Oil Water Content(ppm)

Sunflower 203

Canola 236

Peanut 221

Sesame 197

Extra virgin olive 865

Water in refined oil is derived in two possible ways. The first is from the original

water in the oilseeds. Sorption isotherms experiments showed water with three levels of

Page 39: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

19

affinity in soybean seeds: (i) a region of strongly bound water at moisture concentrations

below 8%; (ii) a region of weakly bound water at moisture concentrations between 8 and

24%; and (iii) a region of very loosely bound water at concentrations greater than 24%

(60). When soybeans are dried for storage, the moisture content is ~ 10 wt %. That water

is likely the strongly bound water in the oilseed but it is possible that some of this water

ends up in the crude oil. The second possible source of water is the water used to wash

oils and remove the free fatty acids and phospholipids during the neutralization and

degumming steps. The majority of this water is removed by centrifugation. The

remaining water is removed by a vacuum dryer that controls the moisture content of the

washed oil to below 1000 ppm, most often in the range of 500 ppm (61).

2.2.8 Oxygen

Oxygen is a primary reactant in lipid oxidation reaction which fuels the fatty acid

decomposition pathway that causes rancidity. There are two types of oxygen in bulk oil

systems, dissolved oxygen (DO) and non dissolved oxygen. Dissolved oxygen is

incorporated into food while non dissolved oxygen stays in the headspace (62). The

solubility of oxygen and the rate of its diffusion in the bulk oil are relevant to the rate and

extent of lipid oxidation. Oxygen is as about 3-10 times more soluble in bulk oil than

water (saturation occurs at 5-10 ppm in pure water at 20 C) (63, 64). Schrader and

coworkers determined that the diffusion coefficient of oxygen in soybean oil was 0.6×10-

9 m2.s-1 at 20 °C which is lower than that of olive oil (65). The impact of temperature on

the solubility of oxygen in different types of oil is listed in Table 2.8. Some researchers

suggested that oxygen solubility was related to fatty acid chain length with oxygen

Page 40: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

20

solubility decreasing with increasing the hydrocarbon chain length (66).

Table 2.8. Solubility of oxygen (ppm) in Marine oil

Oil type Temperature(oC)

20 40 60 80

Redfish oil 20.4 25.8 29.6 4.1

Capelin oil 19.5 26.6 34.4 8.5

Herring oil 35.8 42.2 45.2 25.8

Mackerel oil 13.6 25.6 28.7 3.7

Harp seal oil 29.1 35.7 37.9 10.1

Flounder oil 23.2 25.6 26.9 9.6

Cod liver oil 39.3 47.6 43.0 6.0

The pressure is 101.325 Kpa

Another form of oxygen associated with bulk oil is non dissolved oxygen or

headspace oxygen. The air space above the oil, for example in the oil tank or commercial

oil bottle, is where headspace oxygen located. The headspace oxygen content in olive oil

and marine oil at temperature lower than 60 °C increases with the increase of temperature,

which following the rule of the Henry’s Law, i.e., the solubility is proportional to the

partial pressure of headspace gas (67). In other words, as temperatures increase, oxygen

in the oil moves into the headspace. At temperatures higher than 60 °C there is a large

variation in the solubility of oil, especially in marine oil presumably due to heat

accelerated oxidation which consumes oxygen (Table 2.8.). The mass transfer between

headspace oxygen and oxygen in the oil is influenced by: (i) the rate of absorption; (ii)

the oxygen concentration in the headspace; (iii) the headspace surface area to volume

ratio; and (iv) agitation. With those factors being considered, an equation was developed

to describe the absorption rate of oxygen between bulk oil and oxygen in the headspace

Page 41: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

21

(20).

( )O

dC KAC C

dt V

(2) Here, Co is the equivalent solubility of oxygen in the bulk oil; C is the oxygen content in

bulk oil after t times; K is the mass transfer coefficient; A/V is the area to volume ratio.

2.3 The effect of minor components on bulk oil oxidation

2.3.1 The effect of oxygen on the chemical stability of bulk oil

Lipid oxidation is not a spontaneous reaction. Thermodynamically, oxygen cannot

react directly with double bonds because the spin states are different. Ground state,

atmospheric oxygen is in a triplet state, whereas the double bond of unsaturated fatty

acids is in singlet state. Quantum mechanics requires that spin angular momentum be

conserved in reactions. Therefore, interactions between triplet oxygen and unsaturated

fatty acids demands that either the double bond be excited into a triplet state or oxygen is

converted to a singlet state. The former seems impossible due to the requirement of

prohibitive amounts of energy. The triplet oxygen also cannot convert to singlet states

itself thus direct reaction occurs between the oxygen and double bonds in oil do not occur

unaided (68). However, under the assistance of other minor components in bulk oil,

several reactions can happen. For example, photosensitizers such as chlorophylls can

produce singlet oxidation by mechanisms 3-5 outlined below.

Sen + hv → Sen* (3)

Sen* + 3O2 → 1O2* (4)

1O2* + LH → LOOH (5)

In addition, oxygen along with iron can generate alkyl radicals (L) as shown in

reactions 6 and 7. Schafer and coworkers proposed that iron-oxygen complexes ([Fe2+-

Page 42: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

22

O2]) are one route by which iron can promote lipid oxidation in cell membrane (69).

Fe2+ + O2 → [Fe2+-O2] (6)

[Fe2+-O2] + LH → L (7)

Triplet oxygen is directly involved in lipid oxidation propagation through its ability

to react with alkyl radicals in diffusion controlled radical-radical reactions to form

peroxyl radicals as shown in reactions 8 and 9 (70):

L + O2 ↔ LOO (8)

LOO + RH → LOOH+ R (9)

Because of these multiple pathways that oxygen can be involved in lipid oxidation,

it is not surprising that many studies have shown that reduction of oxygen in packaged

bulk oil can extend shelf life (71, 72). An 18-months shelf life test, performed on virgin

olive oil indicated that the slowest oxidation rates occurred at the lowest initial dissolved

oxygen concentration (71). The same group also reported that removal of oxygen from

extra virgin olive oil by nitrogen purging lowered peroxide values (72). Packaging

technologies such as oxygen scavenging film can also improve the oxidative stability of

bulk oil (73, 74).

2.3.2 The effect of water on the chemical stability of bulk oil

Despite being present at relatively low levels, water could have an impact of the rate,

extent and mechanism of lipid oxidation in bulk oils since it may act as a solvent for

hydrophilic or amphiphilic antioxidants and prooxidants (such as transition metals, free

fatty acids, or lipid hydroperoxides

Page 43: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

23

2.3.3 The effect of free fatty acid on the chemical stability of bulk oil

Generally, FFA with unsaturated bonds showed a lower oxidative stability than their

corresponding methyl esters and triacylglycerols (75). In addition, FFA themselves have

been reported to accelerate the oxidation of triacylglycerols in vegetable oils. FFA have

been shown to have several different prooxidative tendencies. Yoshida and coworkers

demonstrated that FFA (i.e., caprylic, capric, lauric, myristic, palmitic or stearic acid) in

microwaved stripped soybean oil increased oxidation with shorter the chain length and

the higher the levels of FFA being more prooxidative (76, 77). Aubourg and coworkers

later showed similar results in marine oil with a higher degree of oxidation seen in the

presence of short chain (lauric and myristic) than long chain (stearic and arachidic) fatty

acids at 30 oC (78). Paradiso and coworkers recently reported that low amounts of FFA

caused an increase in the formation of oxidized triacylglycerol and triacylglycerol

oligopolymers again showing the prooxidant activity of FFA (79). The prooxidant activity

of FFA is associated with its free acid group since methyl esters of fatty acids do not

accelerate oxidation. Proposed mechanisms of FFA include their ability to accelerate the

decomposition of hydroperoxides and bind metals to make them more prooxidative (75).

Obviously, it is very critical to control the levels and formation of FFA in crude, refined

and stored oils. While most commercial oils have FFA concentrations less than 0.05 wt %,

there may be addition benefits if these levels could be even lower. Minimizing FFA

formation during storage can be accomplished by decreased water exposure, minimizing

temperatures, preventing contact with lipases and avoiding exposure to extremes in pH.

2.3.4 The effect of MAG and DAG on chemical stability of bulk oil

Page 44: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

24

Different impacts of MAG and DAG on the oxidative stability of bulk oils has been

reported. Min and coworkers found that 0-0.5 wt % of monostearin, distearin,

monolinolein, or dilinolein acted as prooxidants in soybean oil (80, 81). Colakoglu and

coworkers found that soybean oil containing 1 wt % monoolein increased the rate of

oxygen consumption (82). Wang and coworkers observed that randomized corn oil

contained higher levels of MAG and DAG (0.3 and 5.1 %, respectively) oxidized much

faster than natural corn oil with no detectable MAG and 1.4% of DAG(83). On the

contrary, Gomes et al found that1-3 wt % MAG decreased oxidation in stripped olive oil

at 60 °C (84). The effect of combinations of MAG or DAG with antioxidants in the bulk

oil has also been studied. MAG or DAG in combination with citric acid resulted in

greater antioxidant activity than the MAG or DAG itself (85). In addition, as the chain

length of the fatty acids on MAG or DAG was increased, the more effective citric acid

became.

2.3.5 The effect of phospholipids on the chemical stability of bulk oil

The impact of phospholipids (PLs) on bulk oil oxidation is also controversial.

Dipalmitoyl phosphatidylcholine (DPPC) and dipalmitoyl phosphatidylethanolamine

(DPPE) were reported to have poor antioxidant activity at 50oC and showed no

synergistic effect with α-tocopherol in methyl linoleate (86). Phosphatidylcholine (PC)

and phosphatidylethanolamine (PE) from egg yolk accelerated the oxidation of methyl

linoleate (86). Takenaka and coworkers also found that PC and PE promoted bonito oil

oxidation in the absence of α-tocopherol. However, they found that PE showed

synergistic antioxidant activity with α-tocopherol, while PC did not (87).

Page 45: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

25

Alternatively, Bandarra and coworkers showed that 0.5 % PC was an effective

antioxidant in sardine oil at 40oC. In addition, a high synergistic effect was observed in

the same system with a mixture of α-tocopherol and PE (88). Others have reported

similar results that PLs can increase the activity of tocopherols in bulk oils (89-92). For

the researchers who observed the antioxidant activity of PLs in bulk oil, metal chelating,

free radical scavenging and formation of Maillard reaction products were given as the

mechanisms by which PLs inhibit oxidation (93-95).

2.3.6 The effect of tocopherols on the chemical stability of bulk oil

Tocopherols are the most common free radical scavenging antioxidants found in

vegetable oils. Vegetable oils normally contain tocopherols concentrations in the range of

200-1000 ppm which originate from the seeds. After refining, almost 70 % of the

tocopherols remain in the bulk oil with 30 % being removed during deodorization.

Overall, tocopherols are probably the most important antioxidants in vegetable oils (for

review see refs (96, 97)). The antioxidant mechanisms of tocopherols (TOC) are shown

below.

LOO + TOC → LOOH +TOC (10)

LOO + TOC → LOO-TOC (11)

However, under certain conditions tocopherols can also act as prooxidants. This is

thought to be due to their ability to reduce endogenous transition metals naturally

occurring in the oil or metals resulting from contamination during processing. These

reduced metals can then promote the decomposition of pre-existing lipid hydroperoxides

as shown in pathways 12 and 13. Yoshida and coworkers found that both α-tocopherol

Page 46: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

26

(100 M) and α-tocotrienol (100 M) reduced cupric iron (300 M) and the

concomitant formation of α-tocopheryl quinone and α-tocotrienyl quinone was observed

by UV absorption spectrum and also by HPLC analysis in methyl linoleate. Interestingly,

the -, and -forms of tocopherols were not found to reduce cupric iron in the same

system (98)

TOC +Mn2+ → TOC + + Mn+ (12)

LOOH + Mn+ → LO +HO-+ Mn2+ (13)

The prooxidant activity of tocopherols has also been proposed to be due to the

ability of the tocopherol radical to promote fatty acid oxidation especially when

tocopherol radical concentrations are high (98-100).

LOOH + TOC → LOO + TOC (14)

LH + TOC → L + TOC (15)

Trace amounts of oxidized tocopherol exist in bulk oil after refining. Min and

coworkers reported that oxidized α-, γ- and δ-tocopherols promote the oxidation of

purified soybean oil in the dark at 55 °C (101, 102). Since it is possible that oxidized

tocopherols occur naturally in bulk oil and that tocopherol ingredients could contain

oxidized tocopherols, this could be a source of prooxidants in oils. More work is need to

determine if and how oxidize tocopherol could impact the antioxidant/prooxidant activity

of tocopherols.

2.3.7 The effect of pigments on the chemical stability of bulk oil

The two major pigments that can impact lipid oxidation in oils are cholorophyll and

carotenoids. Chlorophyll, a photosensitizer, is prooxidative when exposed to light due to

Page 47: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

27

its ability to promote the formation of singlet oxygen (103). As described previously,

singlet oxygen is formed by chlorophyll photosensitization by energy transfer from light

to a sensitizer and then to triplet oxygen (104). Singlet oxygen can then initiate lipid

oxidation because it can directly react with the double bonds of unsaturated fatty acids to

form lipid hydroperoxides (105). In the bulk oil, it is important to remove as much

chlorophyll as possible during refining since most retail oils are stored in transparent

plastic packages.

Unexpectedly, chlorophyll and its degradation product, pheophytin, have also been

reported to inhibit lipid oxidation in the rapeseed and soybean oils at 30 °C during

storage in the dark. This has been proposed to be due to their ability to scavenge peroxyl

and other free radicals (106).

Carotenoids can be efficient singlet oxygen and excited photosensitizers quenchers

that reduce oxidation by converting excited photosensitizers or singlet oxygen to their

less reactive states (107). Burton also proposed that under low oxygen partial pressure

conditions, β-carotene can act as a lipid soluble chain breaking antioxidant (108). They

also found that at oxygen pressures of 20 kpa or higher, β-carotene and related

compounds are prooxidants in a methyl linoleate model system possibly due to the

formation of carotenoids oxidation products (109). During high temperature oil refining,

carotenoids can also become thermally degraded. Thermally degraded carotenoids were

reported to accelerate lipid oxidation in soybean oil at concentration of 50 ppm (110).

More research is needed to elucidate how chemical changes of carotenoids impact the

oxidative stability of oils.

Page 48: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

28

2.3.8 The effect of miscellaneous compounds on the chemical stability of bulk oil

Small amounts of squalene (200 ppm) were reported to have a limited impact on the

oxidative stability of stripped olive oil 40 and 62 °C in the dark, whereas higher

concentration (7000 ppm) of squalene showed an antioxidative effect at the same system

(111). In rape seed oil, Malecka found that squalene (4000ppm) increased the oxidative

stability of oil heated at 170 °C for 10 h (112). No significant impact of squalene (0-

8000ppm) on stripped sunflower oxidation was observed by Mateos and coworkers (113).

Oppositely, squalene was found to accelerate oxidation of stripped olive oil in a

Rancimat apparatus at 100 °C (114).

The effect of sterols on the oxidation of olive oil was studied at 180°C by Gordon

(115). They found that Δ5-avenasterol and fucosterol were effective as antioxidants at

concentration of 0.1 wt %, while other sterols, including cholesterol and stigmasterol,

were ineffective. The proposed mechanism was that lipid free radicals could react rapidly

with sterols at unhindered allylic carbon atoms (115). Phytosterols have also been

reported to act as antioxidants in soybean, rice bran and sunflower (116-118).

Yoon and coworkers found that the oxidized triacylglycerol fractions obtained by

thermal oxidation of soybean oil acted as prooxidant when it was added to stripped

soybean oil (119). Gomes and coworkers found that oxidized triacylglycerol

oliogopolymers, a class of oxidation compounds present in refined vegetable oils, were

the most prooxidative in vegetable oil amongst all the oxidized triacylglycerol fractions

tested (120).

2.4. Association colloids and lipid oxidation in bulk oils

Page 49: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

29

2.4.1 Evidence for the presence of association colloids in bulk oil

As discussed briefly above, one of the major issues that is often overlooked in bulk

oil oxidation is the physical properties of minor components (121). Bulk oil minor

components, such as monoacylglycerols (MAG) and diacylglycerols (DAG),

phospholipids (PLs), sterols, free fatty acids (FFA), and polar products arising from lipid

oxidation are surface active compounds. These surface active molecules have the ability

to form physical structures in the bulk oils in the presence of the small quantities of water

(~300 ppm) typically found in refined oils (14, 121). These structures are known as

association colloids. Recent research suggests that association colloids could be lipid

oxidation reaction sites in oils.

Previous studies have demonstrated that surface active molecules can self-assemble

in non-polar solvents (e.g. benzene, isooctane, cyclohexane, toluene, free fatty acid and

triacylglycerols) in the presence of water to form a variety of association colloids such as

reverse micelles, micro-emulsions, lamella structures, and cylindrical aggregates (Figure

3) (122-125). For instance, Sosaku and coworkers formed reverse micelle systems using

soybean phospholipids as a surfactant with fatty acids or fatty acid ethyl esters as the

organic solvent (126). Sinoj and coworkers recently formed nutrient delivery systems

consisting of water-in-oil nanoemulsions using oleic acid embedded in canola oil (127).

Chen and coworkers identified ordered lamellar structures in hazelnut oil in the presence

of monoglycerols (MAG) (128, 129). These studies suggest that the amphiphilic

molecules in refined vegetable oils have the ability to form association colloids (130).

Page 50: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

30

Figure 2.2. Schematic representation of the association colloids formed by minor constituents in bulk oil (A) a reverse micelle formed by oleic acid; (B) a reverse micelle formed by phospholipids; (C) Inverse lamella structure formed by monoacylglycerol; (D) a mixed reverse micelle formed by free fatty acids and phospholipids. There are several lines of evidence that association colloids occur naturally in refined and

crude oils. The complete removal of water from refined oil is very difficult even at

temperatures up to 200 oC. This suggests that some of the water in oil is bound to polar

compounds possibly in association colloids. We have indirectly seen this in our research

since water content in oil is reduced by the removal of polar compounds. In addition, it

was observed by research groups using ultrafiltration (UF) that only a small portion of the

phospholipids in oils penetrates through UF membranes with a pore diameter of the order

of 4 nm (131-133). The proposed reason for the rejection of phospholipids by these non-

porous membranes was suggested to be due to the formation of phospholipids reverse

micelles, swollen in the presence of small quantities of water plus containing other minor

components such as pigments, which are then rejected by size exclusion. The amount of

rejection phsopholipids closely depended on the amount and the size of reverse micelles

formed during solvent extraction (134-136).

2.4.2 Do association colloids impact lipid oxidation?

Page 51: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

31

Many of the compounds involved in lipid oxidation reactions such as lipid

hydroperoxides, free fatty acids and antioxidants are surface active. Previous research

with oil-in-water emulsions has shown that the physiochemical property of the water-oil

interface plays a very important role in lipid oxidation chemistry since it can impact the

location and reactivity of both prooxidants and antioxidants (137). This suggests that the

water-oil interface of association colloids could also impact lipid oxidation chemistry by

acting as nano-reactors. Some of the first research in this area was by Koga and Terao

who observed that the presence of phospholipids enhanced the antioxidant activity of α-

tocopherol in stripped oil containing a trace amount of water (1%) (92). This study found

that the presence of phospholipid increased the degradation of α-tocopherol in the

presence of the water soluble free radical generator, 2,2-azobis(2-amidinopropyl)

dihydrochloride (AAPH) suggesting that the phospholipids increased the exposure of α-

tocopherol to the water phase. This could also be due to the fact that the reduction

potential of α-tocopherol is lower in polar environments which could make tocopherol a

more efficient free radical scavenger (138). Degradation of α-tocopherol by the water-

soluble free radicals decreased as the phospholipid’s hydrocarbon tail group size was

decreased, and thus the ability of the phospholipid to form association colloids was lost.

Koga and Terao also found that the antioxidant activity of α-tocopherol could be

increased in bulk oil when it was conjugated to the polar head group of

phosphatidylcholine (PC) (95). This increase in activity was again thought to be due to

the increased partitioning of the reactive portion of α-tocopherol into the water phase of

the association colloids in bulk oil (95). This pioneering work was the first to suggest that

the presence of physical structure in bulk oils could impact lipid oxidation chemistry by

Page 52: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

32

altering the activity of free radical scavenging antioxidants.

Wilailuk and coworkers evaluated the ability of water, cumene hydroperoxide, oleic

acid, and phosphatidylcholine to influence the structure of reverse micelles in a model oil

(n-hexadecane) system containing sodium bis(2-ethylhexyl) sulfosuccinate (AOT)

reverse micelles. This study found that water, cumene hydroperoxide, oleic acid, and

phosphatidylcholine can alter reverse micelle size and lipid oxidation rates (139, 140).

Kasaikina and coworkers found the decomposition of cumene hydroperoxides into free

radicals was accelerated by the existence of reserve micelle formed by cationic

surfactants in organic media (130, 141-144). In addition, they found the oxidation

stability of sunflower oil at 80oC in the presence of 0.1 mM 2,6-di-tert-butyl-4-

methylphenol (BHT) was decreased in the presence of fatty alcohols, such as 1-

tetradecanol, 1-octadecanol, and the MAG, 1-monopalmitoylglycerol. This result was

explained by the so-called "micellar effect": the relatively higher concentration of polar

species such as hydroperoxide and peroxyl radicals within and nearby the reverse

micelles formed in the presence of these fatty alcohols and MAG, leading to an increase

of the rate of chain initiation via an acceleration of the hydroperoxides decomposition

(145).

2.4.3 The perspective researches on the mechanism of lipid oxidation in bulk oil

Currently in the food industry, the trends of changing lipid profiles to contain more

polyunsaturated fatty acids and less hydrogenated fats result in food products with an

increased susceptibility to rancidity. In the meantime, the efficacy antioxidants available

in FDA “direct contact with food” category are limited. To overcome the challenge of

Page 53: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

33

developing low trans fatty acids products with nutritionally significant amounts of

unsaturated fatty acids, the food industry must develop new antioxidant technologies. The

overall objective of this research will be to gain a better understanding of how the

chemical and physical properties of edible oils impact the deterioration of

polyunsaturated lipids. In particular, focus will be on how nano-sized structures in bulk

oil (association colloids) impact the oxidative stability of lipids since above mentioned

evidences indicate that these physical structures are the site of oxidation reactions.

Understanding how the physical nature of edible oils impacts rancidity could lead to the

development of new antioxidant technologies and to the more efficient use of existing

antioxidant ingredients.

The framework of this research include:(1) producing association colloids in

stripped soybean oil using inherent minor component(s) in bulk oil. The well-defined

structure will be characterized by all sorts of techniques; (2) how the physical properties

of association colloids impact the chemical properties of bulk oil in a more complex

system after the addition of minor oil components, including different phospholipids,

antioxidants and prooxidant (i.e.,transition metals); (3) elucidate the physical location of

these minor components in the association colloids. The relationship of the physical

location and chemical reactivity of the minor components will then be compared to

determine their relative overall ability to inhibit or promote lipid oxidation in bulk oil. By

better understanding how physical structures in bulk oils impact reactions that cause

rancidity, it is hoped that new antioxidant technologies can be developed.

2.5 Methodologies to study the impact of minor components on the physicochemical

Page 54: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

34

properties of bulk oil

2.5.1 Chemical analysis

2.5.1.1 Primary lipid oxidation products

Lipid hydroperoxide in stripped soybean oil with or without association colloids will

be determined (146). Precisely weighted of oil will be mixed with 2.8 mL of methanol/1-

butanol (2:1). The reaction will be started by added 15 μL of 3.94 M ammonium

thiocyanate and 15 μL of ferrous iron solution. After 20 min of incubation at room

temperature, the absorbance will be measured at 510 nm using an UV-vis

spectrophotometer. Hydroperoxides concentrations will be determined using a standard

curve prepared from hydrogen peroxide.

2.5.1.2 Secondary lipid oxidation products

Headspace hexanal will be determined using a gas chromatograph (147). The

headspace conditions will be as follows:Sample (1 mL) in 10 mL glass vials capped with

aluminum caps with PTFE/silicone septa were pre-heated at 55 °C for 15 min in an

autosampler heating block. A 50/30 μmDVB/ Carboxen/PDMS solid-phase

microextraction (SPME) fiber needle from Supelco (Bellefonte, PA, USA) was injected

into the vial for 2 min to absorb volatiles and then was transferred to the injector port

(250 °C) for 3 min. The injection port was operated in split mode, and the split ratio was

set at 1:5. Volatiles were separated on a Supleco 30 m × 0.32 mm Equity DB-1 column

with a 1 μm film thickness at 65 °C for 10 min. A flame ionization detector was used at a

temperature of 250 °C. Concentrations will be determined using a standard curve made

from hexanal.

Page 55: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

35

2.5.2 Physical analysis

2.5.2.1 Small/wide angle x-ray scattering (SAXS/WAXS)

X-ray scattering is powerful method for unraveling structural details and molecular

shapes (148). The physical principles of SAXS/WAXS are similar. Firstly, the electric

field of the incoming wave induces dipole oscillations in the atoms, then the accelerated

charges generate secondary waves that add at large distanced (far field approach) to the

overall scattering amplitude. All secondary waves have the same frequency but may have

different phases caused by different path lengths.

Figure. 2.3. The essential parts of a small angle scattering system. The drawing shows the X-ray source X, the sample S, the scattering angle θ, the slits used to define the incident and scattered beams, and the detector D.

X-ray from the source is formed into a fine beam, often by slits, and strike the

samples. A small fraction of this beam is scattered in other directions which forms an

angle with the direction of the incoming beam. A detector is used to record the scattering

intensity which depends on the scattering angle. The ordered structure information of the

sample can often be obtained from the analysis of the scattering intensity at a sequence of

scattering angles.

sin4q

(16)

Page 56: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

36

Also, from the relation between the diffraction peaks, the nature of ordered materials,

such as liquid crystalline phase may be identified. When the composition of the sample is

known together with the nature of materials, X-ray diffraction data can be used to extract

information about the characteristic dimensions of materials. Also for systems without

long-range order, SAXS may provide useful information for micelles, liposomes and

other disperse systems, as well as on gel structure (149).

2.5.2.2 Steady-state and time resolved fluorescence measurement

Luminescence is the emission of light that can occur when a molecule excited to a

higher order electronic state relaxes back to the ground state. Depending on the electronic

nature of the excited state, singlet (S1) versus triplet (T1), luminescence is divided into

two categories, fluorescence and phosphorescence, respectively (150).

Fluorescence measurements can be typically made in two regimes, i.e., steady state

and time resolved measurements. Steady-state measurements are measured with constant

illumination and observation. Due to the nanosecond timescale of fluorescence, the

steady-state condition for a system is normally reached almost immediately after sample

is illuminated. In contrast, time-resolved measurements involve monitoring the temporal

dependence of a given fluorescence parameter, like emission intensity or anisotropy. For

these experiments the sample is exposed to a pulse of light that is shorter than the decay

time of the sample and measurements are then made using a high-speed detection system,

capable of making discrete observations within a nanosecond time regime. Although

time-resolved experiments are more complex, these measurements can provide novel

information that may not be observed in steady-state measurements due to averaging

Page 57: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

37

processes. For example, anisotropy decays of fluorescent macromolecules are frequently

more complex than single exponential. The precise shape of the anisotropy decay

contains information about the shape of the macromolecule and its flexibility. Time-

resolved measurements of intensity decays may also contain information that is obscured

in steady-state measurements due to averaging processes (151, 152).

2.5.2.3 Interfacial tension analysis

Interactions occur between the molecules of a liquid and those of any liquid which is

not soluble in the liquid. These result in the formation of an interface. Energy is required

to change the form of this interface. The work required to change the shape of a given

surface is known as the interfacial tension.

Surfactants consist of a hydrophilic head and a hydrophobic tail. If a surfactant is

added to oil then it will initially enrich itself at the surface; the hydrophilic head projects

from the surface. Only when the surface has no more room for further surfactant

molecules will the surfactant molecules start to form agglomerates inside the liquid.

These agglomerates are known as reverse micelles. The interfacial tension of system will

decline as the increasing of surfactant concentration. The steady lowest interfacial tension

of the system will reach when reverse micelle formed after which will not change. The

surfactant concentrations at which reverse micelle formation is known as the reverse

critical micelle formation concentration (CMC). By doing interfacial tension analysis, the

surface active properties of minor components are able to be investigated.

Page 58: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

38

CHAPTER 3

THE IMPACT OF PHYSICAL STRUCTURES IN SOYBEAN OIL

ON LIPID OXIDATION

3.1 Introduction

Bulk oil is often assumed to be a homogenous liquid, in contrast to more complex

multiphase systems, such as emulsions, which are regarded as heterogeneous liquids.

However refined oil, which is processed to remove many of the non-triacylglycerol

fractions, still contains numerous minor components that are amphiphilic, such as

monoacylglycerols, diacylglycerols, phospholipids, sterols, free fatty acids, and polar

products arising from lipid oxidation, such as lipid hydroperoxides, aldehydes, ketones,

and epoxides (121, 153, 154). In addition, refined oils also contain small but significant

amounts of water. Previous studies have demonstrated that surface active molecules can

self-assemble in the presence of water to form a variety of association colloids in non-

polar solvents (e.g. benzene, isooctane, cyclohexane, toluene, free fatty acid and

triacylglycerols), such as reverse micelles, micro-emulsions, lamella structures, and

cylindrical aggregates (122-124). For instance, Sosaku and coworkers formed reverse

micelle systems using soybean phospholipids as a surfactant in fatty acid or fatty acid

ethyl esters as the organic solvent (126). Sinoj and coworkers recently formed nutrient

delivery systems consisting of water-in-oil nanoemulsions using oleic acid embedded in

canola oil (127). These studies suggest that the amphiphilic molecules in refined

vegetable oils have the ability to form association colloids (130).

The mechanism of lipid oxidation in bulk oil has been studied for decades (155).

Research over the past two decades has shown that in emulsion systems, lipid oxidation

Page 59: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

39

is strongly influenced by the physical properties of the emulsion droplets and their

interfaces, which can impact pro-oxidant/lipid interactions and the location of

antioxidants (156). Most previous research has treated bulk oil as a homogeneous system

without considering how lipid oxidation reactions could be influenced by their physical

structure. If bulk oils are actually micro-heterogeneous systems containing association

colloids, it is highly likely that these physical structures will impact lipid oxidation as

they do in emulsion systems.

To better understand how microstructures impact the oxidation of bulk oil, it will be

necessary to develop model association colloid systems that reflect the types of physical

structures that would be expected to exist in refined vegetable oils. Therefore, in this

study, a model system consisting of different concentrations and types of phospholipids

[1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and 1,2-dibutyryl-sn-glycero-3-

phosphocholine (DC4PC), Figure 3.1] and water was used to characterize association

colloids in stripped soybean oil and to evaluate their impact on lipid oxidation.

Figure 3.1.Molecular structures of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC).

3.2 Materials and Methods

Page 60: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

40

3.2.1 Materials

Avanti Polar Lipids, Inc (Alabaster, AL) was the source of 1,2-dioleoyl-sn-glycero-

3-phosphocholine (DOPC) and 1,2-dibutyryl-sn-glycero-3- phosphocholine (DC4PC). N-

(7-nitrobenz-2-oxa-1,3-diazol-4-yl)-1,2-dihexadecanoyl-sn-glycero-3-

phosphoethanolamine, triethylammonium salt (NBD-PE) was acquired from Invitrogen

(Carlsbad, CA). Soybean oil was purchased from a local store and stored at 4 °C. Silicic

acid (100-200 mesh), activated charcoal (100-400 mesh), 7,7,8,8-TCNQ (7,7,8,8-

tetracyano-quinodimethane) and hexane were purchased from Sigma-Aldrich Co. (St.

Louis, MO). Medium chain triacylglycerols (MCT, Miglyol®) were purchased from

Sasol North America Inc (Houston, TX). All other reagents were of HPLC grade or purer.

Distilled and deionized water was used in all experiments.

3.2.2 Preparation of stripped soybean oil

Stripped soybean oil was prepared according to the method of Boon et al (157). In

short, silicic acid (100 g) was washed three times with a total of 3 L of distilled water and

activated at 110 °C for 20 h. Activated silicic acid (22.5 g) and activated charcoal (5.625

g) were suspended in 100 and 70 mL of n-hexane, respectively. A chromatographic

column (3.0 cm internal diameter × 35 cm height) was then packed sequentially with 22.5

g of silicic acid followed by 5.625 g of activated charcoal and then another 22.5 g of

silicic acid. Thirty grams of soybean oil was dissolved in 30 mL of hexane, which was

then loaded onto the column and eluted with 270 mL of n-hexane. To retard lipid

oxidation during stripping, the collected soybean oil was held on ice and covered with

aluminum foil to eliminate the impact of light. The solvent in the stripped oil was

Page 61: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

41

removed with a vacuum rotary evaporator (RE 111 Buchi, Flawil, Switzerland) at 37 °C,

and the remaining trace solvent was removed by flushing with nitrogen for at least one

hour. The colorless oil was kept at -80 °C for subsequent studies. Removal of minor

components was verified by spotting a sample on a silica gel G thin layer

chromatography (TLC) plate and developed in a tank with hexane/diethyl ether/acetic

acid (70:30:1 v/v/v) (158). Water content of the stripped oils was determined by Karl

Fisher (159).

3.2.3 Light Scattering Properties of Oil Samples

A series of samples were prepared to determine the DOPC and water concentrations

that produced association colloids without causing cloudiness in the oil. A stock solution

of DOPC was prepared at a concentration of 127 moles DOPC/mL chloroform, which

was blended into MCT over the concentration range of 0 to 1270 mole/kg lipid which is

in the range of phospholipid concentrations found in refined bulk oils (~ 2000 moles/kg

lipid). The mixture of MCT and DOPC was then titrated with water. The initial water

concentration in the samples was 0.67 mmoles/kg lipid (determined by Karl Fisher

analysis) and was increased to 560 mmoles/kg lipid by adding water. The vials were

tightly closed with Teflon-lined caps and placed on a stirrer motor at 25 °C for 24 hours

before measurements. The final water concentrations were verified by Karl Fisher

analysis. The clarity of the samples was measured by light scattering using an Agilent

7010 particle size spectrophotometer. The Agilent 7010 particle size spectrophotometer

measures the attenuation of a colloidal dispersion from 350-800 nm. The transparent

region was defined as samples with a measured attenuation 0.2 cm-1 (visibly clear),

Page 62: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

42

while the cloudy region was defined as samples with attenuation > 0.2 cm-1.

3.2.4 Determination of critical micelle concentration (CMC) in stripped soybean oil

The procedure described by Kanamoto and coworkers (160) was followed to

spectrophotometrically determine the CMC of the phospholipids in stripped oil. In short,

varying concentrations of DOPC or DC4PC were dissolved in either SSO or MCT by

mixing for 12 h followed by addition of 7,7,8,8-tetracyano- quinodimethane (TCNQ,

5mg) to 5 g of oil/DOPC in a small conical flask and the mixture was agitated using a

magnetic stirrer for 5 h at room temperature. After sedimentation of excess TCNQ by

centrifugation at 2000 g for 20 min, absorbance measurements at 480 nm were recorded

(Shimazu 2014, Tokyo, Japan). SSO oil or MCT without DOPC was used as a control.

The CMC was taken as the intersection point of straight lines extrapolated from low and

high DOPC concentrations in the curve generated from absorbance and DOPC

concentrations on a semi-log plot.

The CMC of DOPC in SSO was also determined by interfacial tension (IFT)

measurements using Drop Shape Analysis (DSA100, Krüss GmbH, Hamburg, Germany)

equipped with a pendant drop module. A needle with a diameter of 1.830 × 10-3 m was

used to create a pendent drop (containing the stripped soybean oil and varying

concentrations of DOPC). The pendant drop was extruded from the syringe into a quartz

cell containing distilled water. Droplet images were taken every 2 s and the drop shape

analysis program supplied by the instrument manufacturer (based on the Young-Laplace

equation) was used to determine IFT values.

Page 63: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

43

3.2.5 Small angle X-ray scattering (SAXS)

SAXS measurements were performed on the oil samples containing varying

concentrations of phospholipids and water using a Rigaku Molecular Metrology SAXS

instrument (Rigaku, Inc) at the W.M. Keck Nanostructures Laboratory at the University

of Massachusetts-Amherst. The instrument generates X-rays with a wavelength of 1.54 Å

and utilizes a 2-D multi-wire detector with a sample-to-detector distance of 1.5 m.

Samples were inserted into the 1 mm outer diameter quartz capillary (Hampton Research,

Aliso Viejo,CA, USA) which were then sealed by Duco® Cement and then enclosed in

an airtight sample holder. This assembly was then put in the X-ray beam path, and data

were collected on the samples for 3 hours.

The experimental SAXS intensity curves were corrected for background, sample

absorption and detector homogeneity, using MicroCal Origin 5.0™ software. The

scattering vector amplitude q is defined by equation (1), where θ is half of the scattering

angle and is the wavelength:

sin4q (1)

Fittings for the experimental SAXS curves were obtained using the GNOM program

(version 4.5, ATSAS, German) (161). For a set of monodisperse spherical particles

randomly distributed, the scattering intensity is given by the following equation (162):

(2)

where γ is a factor related to the instrumental effects; np corresponds to the number

of scattering species; Δρ is the electron density contrast between the scattering species

Page 64: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

44

and the medium; V is the scattering species volume; P is the normalized particle form

factor (P(0)=1); and S is the structure factor of the particle system, its value being ≈ 1 for

non-correlated systems. In this particular case, the intensity function I depends solely on

the particle form factor and according to theory, the pair-distance distribution function

p(r), can be obtained by Fourier inversion of the intensity function (163):

(3)

This function provides information about the shape of the scattering particles as well

as their maximum dimension, Dmax accounted for by the r (pair-distance) value where p(r)

goes to zero.

3.2.6 Front-Face (FF) Fluorimetric Measurements

Front-face fluorescence with surface active fluorescent probes was used to verify the

ability of DOPC and water to form association colloids. The surface active fluorescent

probe used was NBD-PE which is a phospholipid analog with a fluorescent functional

groups covalently attached to the choline head group. Steady-state emission

measurements were recorded with a PTI Spectrofluorimeter (PTI, Ontario, Canada) with

the sample holder held at 22 °C. To minimize any reflection of the excitation beam by the

cell window and by the underlying liquid surface of the sample into the emission

monochromator, samples are stored in triangular suprasil cuvette. The emission was

observed at 90° to the incident beam, i.e., 22.5° with respect to the illuminated cell

surface. A 2.0 nm spectral bandwidth was used for both excitation and emission slits were

employed for the NBD-PE excitation at 468 nm. The integration time was 1 s, and the

wavelength increment during emission spectrum scanning was 1 nm. The intensity of the

Page 65: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

45

spectra was determined as the emission signal intensity (counts per second) measured by

means of a photomultiplier. In order to avoid self quenching of probes, the concentrations

of NBD-PE was 9.5 M (1/100 of DOPC) (164).

3.2.7 Measurement of oxidation parameters

Lipid hydroperoxides were measured as the primary oxidation product using a method

adapted from Shanta and Decker (165). Accurately weighed oil samples were added to a

mixture of methanol/butanol (2.8 mL, 2:1, v:v) followed by addition of 15 μL of 3.94 M

thiocyanate and 15 μL of 0.072 M Fe2+. The solution was vortexed, and after 20 min the

absorbance was measured at 510 nm using a Genesys 20 spectrophotometer

(ThermoSpectronic, Waltham, MA). The concentration of hydroperoxides was calculated

from a cumene hydroperoxide standard curve.

Secondary oxidation products were monitored using a GC-17A Shimadzu gas

chromatograph equipped with an AOC-5000 autosampler (Shimadzu, Kyoto, Japan)

(147). Sample (1 mL) in 10 mL glass vials capped with aluminum caps with

PTFE/silicone septa were pre-heated at 55 °C for 15 min in an autosampler heating block.

A 50/30 μmDVB/ Carboxen/PDMS solid-phase microextraction (SPME) fiber needle

from Supelco (Bellefonte, PA, USA) was injected into the vial for 2 min to absorb

volatiles and then was transferred to the injector port (250 °C) for 3 min. The injection

port was operated in split mode, and the split ratio was set at 1:5. Volatiles were separated

on a Supleco 30 m × 0.32 mm Equity DB-1 column with a 1 μm film thickness at 65 °C

for 10 min. The carrier gas was helium at 15.0 mL/min. A flame ionization detector was

used at a temperature of 250 °C. Propanal concentrations were determined from peak

Page 66: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

46

areas using a standard curve prepared from an authentic standard.

3.2.8 Statistical Analysis

Duplicate experiments were performed for all studies. All data shown represents

mean values standard deviations (n = 3). Statistical analysis of lipid oxidation kinetics

was performed using a one-way analysis of variance. A significance level of p<0.05

between groups was accepted as being statistically difference. In all cases, comparisons

of the means of the individual groups were performed using Duncan’s multiple range

tests. All calculations were performed using SPSS17 (http://www.spss.com; SPSS Inc.,

Chicago, IL).

3.3 Results and discussion

3.3.1 Attenuation of the water/phospholipid/MCT system

Commercial refined vegetable oils are optically clear and yet potentially contain

physical structures such as association colloids. This is possible because many

association colloids have dimensions below 100 nm and thus they do not strongly scatter

visible light (166). The objective of this initial study was to define the

phospholipid/water/lipid concentrations that produce optically transparent oils,

corresponding to commercial refined oils. To do this, a partial phase diagram of the three

component system water/phospholipid/MCT was constructed to determine the

phospholipid and water concentrations that did not form structures that scattered light

strongly as determined by light attenuation measurements (Figure 3.2) using a method

previously reported for phospholipids and water in olive oil (167). In the presence of

Page 67: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

47

1.25, 2.5, and 5 mg DOPC/g MCT, a large increase in attenuation was observed at a water

concentration 20 mg/g MCT. At 20 mg/g MCT, the attenuation of the sample

containing the highest concentration of DOPC (5 mg DOPC/ g MCT) had the lowest

attenuation. This is likely due to the formation of a larger number of small non-light

scattering association colloids in the presence of high levels of DOPC. In the remainder

of the experiments we focused on those phospholipid/water/lipid concentrations that gave

optically transparent systems.

Figure 3.2. Attenuation of medium chain triacylglycerols (MCT) samples containing 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and varying concentrations of water.

3.3.2 Determination of the CMC of phospholipids in SSO and MCT

Phospholipids can self-assemble into a variety of structures in water or oil phases

because of their amphiphilic nature (168). The nature of the structures formed depends on

the phospholipid’s chemical structure and solvent type (169). In non-aqueous media, such

as oil, phospholipids normally form reverse micelles when their concentration just

exceeds the CMC (critical micelle concentration), with the polar head groups pointing

0 5 10 15 20 25 30 35 40 45 50

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6 DOPC=1.25 mg DOPC=2.5 mg DOPC=5 mg

Att

en

uati

on

at

600 n

m (

%)

Water content (mg)

Page 68: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

48

into the hydrophilic core of the reverse micelle and the non-polar tails pointing towards

the hydrophobic oil. Therefore, in the region of the phase diagram where the system is

clear (Figure 3.2), it is important to determine whether the DOPC concentration is above

or below its CMC so as to determine if any association colloid structures form. SSO can

rapidly oxidize resulting in the production of additional surface active oxidation products

which could alter the CMC. Therefore, the CMC of DOPC was measured in both SSO

and MCT to determine if SSO was a suitable solvent for analysis of association colloids.

One of the most important properties of reverse micelles is their ability to increase the

solubility of molecules. 7,7,8,8-TCNQ is not lipid soluble and does not absorb light at

480 nm when dispersed in non-polar solvents. 7,7,8,8-TCNQ can undergo charge-transfer

(170) interactions with surfactants when the surfactant concentration is above the CMC,

which can be observed by absorbance at 480 nm (160). The absorbance values of 7,7,8,8-

TCNQ solubilized in MCT were plotted as a logarithmic function of phospholipid

concentration (Figure 3.3 A). When the concentration of DOPC was increased, the

absorbance of TCNQ initially remained low, but there was a rapid increase in absorbance

at DOPC concentrations above 500 M. Based on the intercept of these two linear

regions, the CMC of DOPC was estimated to be ~ 550 M in MCT. This experiment was

repeated in SSO where the CMC of DOPC was found to be ~ 650 M (Figure 3.3 B).

Differences of the CMC in MCT and SSO could be due to differences in fatty acid chain

length and the presence of other minor surface active components in the oils that were not

completely removed by stripping (171). Since the CMC values for MCT and SSO were

fairly similar, SSO was used for all subsequent experiments. The CMC of DC4PC in the

SSO could not be determined by this method because DC4PC did not dramatically change

Page 69: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

49

the absorbance of the 7,7,8,8-TCNQ before the DC4PC reached its solubility limit in the

SSO.

Figure 3.3. Determination of critical micelle concentration (CMC) of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) in medium chain triacylglycerols (A)and stripped soybean oil (B) as determined by the absorbance (480 nm) of 7,7,8,8-tetracyano- quinodimethane.

The CMC of DOPC in SSO was also determined by interfacial tension

10 100 10000.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6A

b480

Phospholipid concentration (M)

550M

A

10 100 1000

0.0

0.5

1.0

1.5

2.0

2.5

Ab

480

Phospholipid concentration(M)

650M

B

Page 70: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

50

measurements. At low DOPC concentrations, the interfacial tension decreased with

increasing DOPC concentration, but at higher DOPC concentrations the interfacial

tension reached a fairly constant value indicating that reverse micelles had been formed.

A semi-logarithmic plot between interfacial tension and the concentration of DOPC in

SSO is shown in Figure 3.4. Based on the intercept of the two linear regions a CMC of ~

950 M was calculated. The CMC value obtained by interfacial tension measurements is

about 300 M higher than that obtained with the spectrophotometric technique, which

could be due to some of the DOPC partitioning into the aqueous phase during the

interfacial tension measurements. Alternatively, the spectrophotometric and interfacial

tension methods may have given different CMC values because they are based on

different physical principles. For the purposes of this study, we considered the CMC of

DOPC in SSO to be in the range of 650-950 M. The normal concentration of

phospholipids in refined vegetable oil is higher than 950 M, suggesting that commercial

oils could have association colloids formed by phospholipids.

10 100 1000

0

2

4

6

8

10

12

14

16

IFT

(m

N/m

)

Phospholipid concentration (M)

950M

Page 71: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

51

Figure 3.4. Determination of critical micelle concentration of 1,2-dioleoyl-sn-glycero-3-

phosphocholine in stripped soybean oil using interfacial tension.

3.3.3 Detection and characterization of phospholipids physical structures by small

angle x-ray scattering

As mentioned above, the specific structures formed by phospholipids in non-

aqueous solution depend on both concentration and phospholipid type. In consideration

of the complicated compositions of phospholipids in refined bulk oil, we first determined

how two types of phospholipids impacted association colloid structure in ternary systems

containing SSO, water, and either DOPC or DC4PC using. Figure 3.5 shows the SAXS

profiles in the SSO systems containing ~ 60 ppm water (i.e., intrinsic water content after

oil stripping) and either 950 M of DOPC or DC4PC. These phospholipid concentrations

were chosen as these samples did not strongly scatter light (Figure 3.2) but were equal to

or above the CMC of DOPC (Figure 3.3 and 3.4). As shown in Figure 3.5, the two

different types of phospholipids induced changes of the minor structures in stripped oil in

such a way that the scattering profiles are distinct and displaced towards different q

values and intensity. The Bragg space (d-spacing) and the intensity of the samples with

DOPC was similar to blank stripped oil, being 141 Å and 3 (a.u.) respectively. However,

the structures with DC4PC are quite different to the blank and the one with DOPC. Its

Bragg peak shifts towards a higher q value with Bragg space being 68 Å while the

intensity rises to 12 (a.u.). In addition, the structures with different types of phospholipids

are verified by pair distribution p(r) function analysis (Figure 3.6). The p(r) of DC4PC

had a long tail at high r, which is typical of a cylindrical structure. The r value at the

Page 72: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

52

inflection (~20 nm) usually represents the diameter of the cylinder (172). The p(r) of

DOPC had a symmetrical shape which represents a spherical structure, presumably a

reverse micelle. For the bulk stripped soybean oil, there was only a flat line indicating the

presence of no structures.

Figure 3.5. Small angle x-ray scattering profiles of stripped soybean oil (SSO) and SSO + 950 M 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3- phosphocholine (DC4PC).

0.0 0.4 0.8 1.2 1.6 2.0

0

1

2

3

0

1

2

3

0

5

10

q (nm-1)

SSO

68Å

52.1 Å

141 Å

51.3 Å

Inte

nsit

y

DOPC

DC4PC

141 Å

52.6 Å

Page 73: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

53

Figure 3.6. The pair distribution functions, p(r), calculated from the scattering profiles of stripped soybean oil (SSO) and SSO + 950 M 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine(DC4PC) using GNOM analysis of the three small angle x-ray scattering profiles indicated in Figure 3.5.

In the stripped oil, almost all the surface active minor components and the majority

of water had been removed by the silicic acid and activated charcoal sandwich

chromatography which is consistent to our, Karl Fisher, TLC and previous HPLC results

(173). Therefore it is reasonable that the bulk stripped oil did not contain any association

colloid structures (174). The structures formed by DC4PC are quite different to those

formed by DOPC. DOPC and DC4PC have the same phosphocholine hydrophilic head

group (Figure 3.1), but DOCP possesses cis- oleic fatty acids on sn-1 and sn-2 of

glycerol side chains, whereas DC4PC has 2 butyl fatty acids. As indicated, amphiphilic

molecules, like phospholipids, having a large tail area and a small head group area, can

self-assemble into reverse spherical micelles in highly non polar liquids due to a critical

packing parameter (CPP) values greater than 1 (175). There are no reported CPP values

of DC4PC. However, based on the critical packing parameter equation (176):

0 10 20 30 40 50

0.000

0.005

0.010

0.015

0.020

0.025

P(r)[

a.u

.]

r/nm

DOPC SSO DC4PC

Page 74: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

54

(4)

(5)

(6)

where ν is the volume of the hydrocarbon tail and n is the number of carbons in the

hydrophobic chain, the CPP of DC4PC in SSO should be lower than that of DOPC,

approximately less than ½ assuming no difference of their effective area of the head

group. This is because the CPP is mainly dependent on the hydrocarbon tail length. A

low CPP often corresponds to the formation of cylindrical structures.

3.3.4 Effect of water/phospholipid molar ratio on the physical properties of DOPC

association colloids

Experimental and theoretical approaches have shown that the key structural

parameter of reverse micelles is the water/surfactant molar ratio (Wo) (177, 178). On one

hand, water will serve to bridge the phosphate head groups between neighboring

phospholipids through hydrogen bonds (179). On the other hand, the water content will

determine structure size as well as the amount of the water that is strongly associated

with the phospholipid head groups (164). Oil-rich phospholipid solutions with micellar

morphology can change to lamellar and hexagonal structures as the Wo is raised in the

system (24). In order to characterize the physical structure of stripped soybean oil with

DOPC (950 M) as a function of Wo, a combination of SAXS and Front-Face

Fluorimetry were used.

Page 75: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

55

Figure 3.7. Small angle x-ray scattering profiles of stripped soybean oil with different water/1,2-dioleoyl-sn-glycero-3-phosphocholine molar ratios (WO). Inserted figures are the 2 dimensional small angle x-ray scattering pattern of each sample.

At high Wo, water is solubilized into phospholipid reverse micelles and structures

such as cylindrical micelles and spherical swollen micelles can form depending upon the

amount of water in the system (176). In our system there is a clear Bragg peak on all the

SAXS profiles, with d-space around 141 nm (Figure 3.7). The SAXS intensity is very

Page 76: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

56

weak (i.e., 3 a.u.) for the system with Wo = 0 (no added water), suggesting that DOPC

alone forms very little structure. When the Wo increased from 0 to 35 and 47, the

scattering intensity increased from 3.0 to 8.0, and 13.0, respectively, presumably due to

the formation of association colloids by DOPC. Increasing Wo above 100 results in the 2-

D X-ray pattern of the samples becoming aligned, indicating a transformation from an

isotropic to an anisotropic system. At the same time, the SAXS intensity increased greatly

to 749.6 a.u. In all samples, no other Bragg peaks were observed (Figure 3.7), indicating

the absence of cylindrical micelles and spherical swollen micelles. Lack of

transformation form reverse micelles to cylindrical micelles and spherical swollen

micelles was also reported in a phospholipid in hexane and soybean oil system (180). The

fact that the reverse micelles did not transform to these other colloidal structures in our

system could be due to: (1) the increase in Wo was too small in our experimental design

so that the transformation did not occur; or, (2) the concentration of the DOPC was too

low to allow transformation. While the anistropic transition at Wo values greater than 100

indicates that the structures were changing symmetry. This could be due to a stretching

of the symmetrical structure into more of an oblong structure such as shown in Figure

3.8.

Page 77: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

57

Figure 3.8. A postulated structural change of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) as phospholipid concentrations are increased to above the critical micelle concentrations (CMC) form a reverse micelle.

Fluorescence intensity of selected probes can also be used to provide information

about the micro-environment of association colloids as a function of Wo. In this study, we

used NBD-PE as a fluorescence probe and Front Face Fluoremetery to detect changes in

its spectra. This probe is a phospholipid analog grafted with a fluorophore on the

phosphate head groups (Figure 3.9 insert) that has previously been used as an indicator

of changes in the interfacial properties of reverse micelle systems (164). Theoretically,

NBD-PE possesses similar properties as DOPC as both are amphiphilic molecules with

the ability to participate in the formation of association colloids and reside at the water-

oil interface (164). The probe is useful because its fluorescence decreases when it is

exposed to water. Figure 3.9 shows the fluorescence intensity of NBD-PE decreased with

increasing water concentrations suggesting that the exposure of the probe to water

increased. The exposure of the probe to the water phase can also be confirmed by a

change in the wavelength of maximum fluorescence intensity of NBD-PE with a red shift

Page 78: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

58

occurring when the probe is exposed to a more polar environment (164). A red shift of 4

nm was observed for NBD-PE in the DOPC/stripped soybean system when the Wo was

increased from 11.7 to 46.8.

Figure 3.9. The fluorescence intensity of N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)-1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine, triethylammonium salt (NBD-PE) in stripped soybean oil with different water/1,2-dioleoyl-sn-glycero-3-phosphocholine molar ratios (WO). Inserted figure is the molecular structure of (NBD-PE).

3.3.5 Oxidation of stripped soybean oil containing association colloids

Studies of lipid oxidation in bulk oils have been carried out for several decades, but

many of the mechanisms involved in this reaction are still unclear, including whether

physical structures in the oil can impact oxidation kinetics. The ultimate purpose of these

experiments was to produce soybean oil, water, and phospholipid systems with different

physical structures and determine if these physical structures impact lipid oxidation

kinetics. Therefore, the lipid oxidation kinetics of stripped soybean oil containing 1000

μM of either DOPC or DC4PC and ~ 200 ppm water was studied as determined by lipid

hydroperoxides and propanal during storage at 37 °C (Figure 3.10). In the presence of

0 20 40 60 80 1000.0

500.0k

1.0M

1.5M

2.0M

2.5M

3.0M

H

O

O

O

O+HN

P

O-O

O

ONH

NO

N

N+

O

O-

(CH3CH2)3

CH3(CH2)14

CH3(CH2)14

Inte

nsit

y (

Co

un

ts)

Wo

Page 79: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

59

DOPC, the lag phase of lipid hydroperoxides (Figure 3.10A) and headspace propanal

(Figure 3.10B) were 8 and 13 days, respectively, compared to 13 and 17 days,

respectively, for the no phospholipid control. DC4PC had the same lag phase times for

lipid hydroperoxides and headspace propanal as the control (Figure 3.10A and B).

Figure 3.10. Formation of lipid hydroperoxides (A), propanal (B) in stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine(DOPC) or 1,2-dibutyryl-sn-glycero-3- phosphocholine (DC4PC) and 200 ppm water at 37 °C. Data represent means (n =3) ± standard deviations. Some error bars are within data points.

0 5 10 15 20 250

40

80

120

160

200

240

Lip

id h

yd

rop

ero

xid

es

(m

mo

l /K

g o

il)

Storage Time(day)

SSO SSO w/ 1000 microM DOPC SSO w/ 1000 microM DC4PC

A

0 5 10 15 20 250

40

80

120

160

200

240

Pro

pan

al

(m

ol

/Kg

oil

)

Storage Time(day)

SSO SSO w/ 1000 microM DOPC SSO w/ 1000 microM DC4PC

B

Page 80: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

60

Phospholipids have previously been reported to have protective effects in bulk oils.

The mechanisms for these protective effects have been postulated to be due to the

phospholipid’s ability to chelate metals, decompose lipid hydroperoxides, directly

scavenge free radicals, and increase the antioxidant activity of tocopherols (181). In this

study, DOPC and DC4PC were chosen since they both have identical choline head groups

so any impact of the phospholipids on alteration in lipid oxidation kinetics by chemical

pathways would be similar. Therefore, the major differences in the oil systems containing

DOPC and DC4PC would be in how these phospholipids produced different physical

structures. As discussed above, 7,7,8,8-TCNQ was an effective tool to monitor the

formation of structures by DOPC but this probe did not detect structures formed by

DC4PC. Using small angle x-ray scattering, structures could be seen for both DOPC and

DC4PC but these structures were very different with DOPC forming spherical and DC4PC

forming cylindrical structures (Figure 3.6). Therefore these results suggested that the

spherical structures formed by DOPC were prooxidative while the cylindrical structures

formed by DC4PC had no impact on oxidation rates. These results suggest that

phospholipids have the ability to form colloidal structures in vegetable oils and that these

structures could have an impact on the oxidative stability of food oils.

Page 81: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

61

CHAPTER 4

ROLE OF REVERSE MICELLES ON LIPID OXIDATION IN BULK OILS:

IMPACT OF PHOSPHOLIPIDS ON ANTIOXIDANT ACTIVITY OF

α-TOCOPHEROL AND TROLOX

4.1 Introduction

Lipid oxidation is one of the main factors limiting the shelf life of bulk oils, since it

adversely affects flavor and quality, and potentially produces toxic reaction products(182).

Preventing or inhibiting the oxidation of bulk oils is therefore of great importance to

consumers and the food industry. A variety of mechanisms have been proposed to be

responsible for the oxidation of bulk oils during processing and storage, with

photosensitized oxidation, metal-promoted and autoxidation being the most well-known.

Some factors that impact the oxidative stability bulk oils include: oil extraction and

processing conditions; light exposure; temperature; fatty-acid composition; antioxidant

composition; oxygen levels; and the presence of minor components (183). Manipulation

of these factors can be used to retard lipid oxidation in edible oils.

One of the most effective ways of inhibiting lipid oxidation in bulk oils is to

incorporate antioxidants (184). Depending on their mechanism of action, antioxidants can

be classified as either “primary” or “secondary” antioxidants (185). Tocopherols are the

most common primary antioxidant present in many vegetable oils, which may originate

naturally from the extracted oil itself or may be manually added after oil refining (186).

However, tocopherols may not be the most effective antioxidants in bulk oil systems.

Page 82: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

62

Indeed, research has shown that hydrophilic antioxidants (Trolox or ascorbic acid)

possess better antioxidant activity than their hydrophobic analogues (tocopherol and

ascorbyl palmitate) in some bulk oil systems (156). The greater tendency for hydrophilic

antioxidants to accumulate at the air-water interface where oxidation may be expected to

begin is one of the mechanisms proposed to account for their better antioxidant activity in

bulk oils (156). However, other mechanisms have also been proposed to account for the

ability of hydrophilic antioxidants to act as better antioxidants than their hydrophobic

analogs in bulk oils since air is more hydrophobic than oil and thus there is no driving

force to concentrate hydrophilic antioxidants at the oil-air interface (181, 185). Studies

have shown that the addition of phospholipids to bulk oils increased the antioxidant

activity of tocopherol (181). It was postulated that the phospholipids formed

microstructures, known as association colloids, within the bulk oil, which caused the

tocopherol molecules to accumulate in the phospholipid microstructures where lipid

oxidation primarily occurred. Other researchers have also demonstrated the ability of

phospholipids to act as antioxidants in various kinds of bulk oils (90, 187). Several

mechanisms were proposed to account for the antioxidant activity of the phospholipids,

including their ability to chelate metals, decompose lipid hydroperoxides, and scavenge

free radicals. Nevertheless, there is still a poor understanding of the importance and

contribution of the combination of phospholipids and tocopherols to the oxidative

stability of bulk oils.

In the previous chapter, we characterized the formation of reverse micelles in bulk

oils, and their impact on lipid oxidation (188). These reverse micelles were formed by

adding phospholipid (1,2-dioleoyl-sn-glycero-3-phosphocholine, DOPC) into stripped

Page 83: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

63

soybean oil (SSO) containing water levels similar to that found in commercial refined

oils. Using small angle x-ray scattering, this study showed that the combination of DOPC

and water resulted in the formation of reverse micelles in bulk oil. Alternately, when

phosphatidylcholine with short chain fatty acyl residues (1,2-dibutyryl-sn-glycero-3-

phosphocholine, DC4PC) was added to the same system, no reverse micelles were formed.

The lipid oxidation chemistry of these two systems was different with DOPC reverse

micelles demonstrating a prooxidative effect while DC4PC had no effect on oxidation

rates. This study suggests that the combination of phospholipids and water can form

physical structures in bulk oils and these structures can impact lipid oxidation chemistry.

In the present study, we examined if reverse micelles in bulk oil produced by

phosphatidylcholine(i.e., DOPC) were able to impact the activity of free radical

scavenging antioxidants. The antioxidants tested included alpha-tocopherol (non-polar)

and Trolox (polar), which are chemical analogs. In addition, the system had equal molar

concentrations of phospholipids with one system containing reverse micelles (DOPC) and

the other containing no measurable reverse micelles (DC4PC). We also measured how

the tocopherol analogs impacted the structure of the reverse micelles with the aim of

better understanding the location and properties of antioxidants in bulk oils. By better

understanding how physical structure in bulk oils impact the activity of antioxidants such

as tocopherols, it might be possible to design systems to improve the activity of these

important “natural” antioxidants.

4.2 Materials and methods

4.2.1 Materials

Page 84: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

64

1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and 1,2-dibutyryl-sn-glycero-3-

phosphocholine (DC4PC) were acquired from Avanti Polar Lipids, Inc (Alabaster, AL).

N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)-1,2-dihexadecanoyl-sn-glycero-3-

phosphoethanolamine, triethylammonium salt (NBD-PE, Cat.No. N-360) was acquired

from Invitrogen. Soybean oil was purchased from a local store and stored at 4 °C. Silicic

acid, activated charcoal, hexane, α-tocopherol, Trolox were purchased from Sigma-

Aldrich Co. (St. Louis, MO). All other reagents were of HPLC grade or purer distilled

and deionized water was used as needed.

4.2.2 Preparation of stripped soybean oil

Stripped soybean oil with DOPC reverse micelles were prepared according to the

method of Chen et al (157). Briefly, formation of reverse micelles was accomplished by

pipetting DOPC (1000 μM, final concentration) in chloroform into an empty beaker and

then flushing with nitrogen until the chloroform was evaporated. The appropriate amount

of medium chain triacylglycerols (MCT) or stripped soybean oil (SSO) was then added

followed by double distilled water at a final concentration of 200 ppm. The oil samples

were then stirred in a beaker at 1000 rpm in a 20 °C incubator room for a 24 h.

4.2.3 Formation of DOPC reverse micelles in stripped soybean oil

The formation of DOPC reverse micelles was done using our previous procedure

(188). Briefly, DOPC (1000 μM) in chloroform was pipetted into an empty beaker and

then chloroform was removed by flushing with nitrogen. The appropriate amount of

stripped soybean oil was then added followed by double distilled water to a final amount

Page 85: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

65

of ~ 200 ppm. The sample in beaker was magnetically stirred at the speed of 1000 rpm in

a 20 °C incubator room for a 24 hrs. Antioxidants were dissolved in ethanol, mixed with

stripped soybean oil and stirred for at least 12 hrs to obtain the homogenous samples. In

the lipid oxidation studies, a mixture of 75% of medium chain TAG and 25% SSO were

used due to the high amount of samples needed. Samples for the oxidation studies were

aliquoted into GC vials (1 mL/vial) stored at 55 °C in the dark. For fluorescence, the

NBD-PE probe was mixed into stripped soybean oil at the same time as DOPC and water.

4.2.4 Measurement of oxidation parameters

Lipid hydroperoxides were measured as the primary oxidation product using a method

adapted from Shanta and Decker (165). Accurately weighed oil samples were added to a

mixture of methanol/butanol (2.8 mL, 2:1, v:v) followed by addition of 15 μL of 3.94 M

thiocyanate and 15 μL of 0.072 M Fe2+. The solution was vortexed, and after 20 min the

absorbance was measured at 510 nm using a Genesys 20 spectrophotometer

(ThermoSpectronic, Waltham, MA). The concentration of hydroperoxides was calculated

from a cumene hydroperoxide standard curve.

Secondary oxidation products were monitored using a GC-17A Shimadzu gas

chromatograph equipped with an AOC-5000 autosampler (Shimadzu, Kyoto, Japan)

(147). Sample (1 mL) in 10 mL glass vials capped with aluminum caps with

PTFE/silicone septa were pre-heated at 55 °C for 15 min in an autosampler heating block.

A 50/30 μmDVB/ Carboxen/PDMS solid-phase microextraction (SPME) fiber needle

from Supelco (Bellefonte, PA, USA) was injected into the vial for 2 min to absorb

volatiles and then was transferred to the injector port (250 °C) for 3 min. The injection

Page 86: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

66

port was operated in split mode, and the split ratio was set at 1:5. Volatiles were separated

on a Supleco 30 m × 0.32 mm Equity DB-1 column with a 1 μm film thickness at 65 °C

for 10 min. The carrier gas was helium at 15.0 mL/min. A flame ionization detector was

used at a temperature of 250 °C. Propanal concentrations were determined from peak

areas using a standard curve prepared from an authentic standard.

4.2.5 Small angle x-ray scattering (SAXS) study of bulk oil with DOPC and

antioxidants

SAXS measurements were performed on the oil samples containing varying

concentrations of phospholipids and water using a Rigaku Molecular Metrology SAXS

instrument (Rigaku, Inc) at the W.M. Keck Nanostructures Laboratory at the University

of Massachusetts-Amherst. The instrument generates X-rays with a wavelength of 1.54 Å

and utilizes a 2-D multi-wire detector with a sample-to-detector distance of 1.5 m.

Samples were inserted into the 1 mm outer diameter quartz capillary (Hampton Research,

Aliso Viejo,CA, USA) which were then sealed by Duco® Cement and then enclosed in

an airtight sample holder. This assembly was then put in the X-ray beam path, and data

were collected on the samples for 3 hours.

The experimental SAXS intensity curves were corrected for background, sample

absorption and detector homogeneity, using MicroCal Origin 5.0™ software. The

scattering vector amplitude q is defined by equation (1), where θ is half of the scattering

angle and is the wavelength:

sin4q (1)

Page 87: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

67

Fittings for the experimental SAXS curves were obtained using the GNOM program

(version 4.5, ATSAS, German) (161). For a set of monodisperse spherical particles

randomly distributed, the scattering intensity is given by the following equation (162):

(2)

where γ is a factor related to the instrumental effects; np corresponds to the number of

scattering species; Δρ is the electron density contrast between the scattering species and

the medium; V is the scattering species volume; P is the normalized particle form factor

(P(0)=1); and S is the structure factor of the particle system, its value being ≈ 1 for non-

correlated systems. In this particular case, the intensity function I depends solely on the

particle form factor and according to theory, the pair-distance distribution function p(r),

can be obtained by Fourier inversion of the intensity function (163):

(3)

This function provides information about the shape of the scattering particles as well

as their maximum dimension, Dmax accounted for by the r (pair-distance) value where p(r)

goes to zero.

4.2.6 Fluorescence measurement of bulk oil with DOPC and antioxidants

Front-face fluorescence with surface active fluorescent probes was used to verify the

ability of DOPC and water to form association colloids. The surface active fluorescent

probe used was NBD-PE which is a phospholipid analog with a fluorescent functional

groups covalently attached to the choline head group. Steady-state emission

measurements were recorded with a PTI Spectrofluorimeter (PTI, Ontario, Canada) with

Page 88: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

68

the sample holder held at 22 °C. To minimize any reflection of the excitation beam by the

cell window and by the underlying liquid surface of the sample into the emission

monochromator, samples are stored in triangular suprasil cuvette. The emission was

observed at 90° to the incident beam, i.e., 22.5° with respect to the illuminated cell

surface. A 2.0 nm spectral bandwidth was used for both excitation and emission slits were

employed for the NBD-PE excitation at 468 nm. The integration time was 1 s, and the

wavelength increment during emission spectrum scanning was 1 nm. The intensity of the

spectra was determined as the emission signal intensity (counts per second) measured by

means of a photomultiplier. In order to avoid self quenching of probes, the concentrations

of NBD-PE was 9.5 M (1/100 of DOPC) (164).

The fluorescence decay of NBD-PE in the systems was carried out at 22 °C using a

PTI Laserstrobe fluorescence lifetime instrument with a PTI GL-3300 nitrogen laser and

a GL-302 tunable dye laser with C-500 laser dye, exciting the oil samples at 500 nm.

Each data point on a lifetime decay curve represents five laser flashes, and each decays

represents 100 of these data points spaced out over the collection time interval.

Data were analyzed using the commercial PTI software. Fluorescence decays were

fitted to sums of two and three exponentials, and the average lifetime was calculated.

4.2.7 Statistical analysis

Duplicate experiments have been performed with freshly prepared samples. All data

shown represents the mean values ± standard deviation of triplicate measurements.

Statistical analysis of lipid oxidation kinetics was performed using a one-way analysis of

variance. A significance level of p<0.05 between groups was accepted as being

Page 89: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

69

statistically difference. In all cases, comparisons of the means of the individual groups

were performed using Duncan’s multiple range tests. All calculations were performed

using SPSS17 (http://www.spss.com; SPSS Inc., Chicago, IL).

4.3 Results

4.3.1 Impact of phospholipids on oxidative stability of SSO

Commercial soybean oil is a highly complex liquid system containing a number of

different components that could potentially alter the rate and extent of lipid oxidation

such as tocopherols, chlorophyll, phospholipids, fatty acids, and water (185).

Consequently, stripped soybean oil (SSO) was used as a model oil in this study to reduce

any effects associated with these minor components. Previous research using small angle

x-ray diffraction showed that the combination of DOPC and water resulted in the

formation of reverse micelles in bulk oil while DC4PC did not form structures10. Initially,

we examined the influence of phospholipid type (DOPC and DC4PC at 1000 μM) on the

formation of lipid hydroperoxides and headspace hexanal in SSO during storage at 55 °C

(Figure 4.1A and B). In the absence of added phospholipids, the lag phase for both lipid

hydroperoxides and headspace hexanal lasted 2 days. Addition of DC4PC had little

impact on the lipid oxidation profile, but the addition of DOPC reduced the lag-period by

1 day, confirming that it acted as a prooxidant.

Page 90: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

70

Figure 4.1. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 μM of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) during storage at 55 °C. Some of the error bars are within data points.

Previous studies of the oxidative stability of bulk oils containing phospholipids

suggest that phosphatidylethanolamine (PE) improves their oxidative stability (189, 190).

On the other hand, studies on the impact of phosphatidylcholine (PC) have indicated

seemingly contradictory results. Prooxidant, antioxidant, and no effects of PC in bulk oil

0 1 2 3 4 5 6 70

100

200

300

400

500

600

Lip

id H

yd

rop

ero

xid

es (

mm

ol/ K

g o

il)

Storage time (day)

Blank (SSO) SSO w/ 1000 microM DOPC SSO w/ 1000 microM DC4PC

A

0 1 2 3 4 5 6 70

100

200

300

400

Hexa

nal (m

icro

mo

l/K

g o

il)

Storage time (day)

Blank (SSO) SSO w/ 1000 microM DOPC SSO w/ 1000 microM DC4PC

B

Page 91: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

71

oxidation have been reported in the literature (191, 192). In agreement with our study, a

prooxidant activity of DOPC was reported by Takenaka, et al (192) in bonito oil, which

was attributed to the unsaturated hydrocarbon chain of the PC used. In our study we used

phosphatidylcholine with oleic acid. DOPC was chosen because of its ease of handling

and dispersion into bulk oil due to its low melting point and due to its high oxidative

stability. In general, the oxidative stability of oleic acid is 10-20 times greater than the

linolenic and linolenic acids found in SSO and thus the fatty acyl residues in the added

DOPC would not be expected to decrease the oxidative stability of the oil. In addition, Le

Grandios, et al (193) showed that oleyl and linolenyl residues in PC are more oxidatively

stable than the same residues in triacylglycerols (TAG) again suggesting that the oleyl

residues in the DOPC added to the SSO would not be responsible for the observed

increase in oxidation rates. These data suggest that the prooxidant activity of DOPC in

comparison to DC4PC is not due to its unsaturated fatty acids.

It is possible that the impact of the phospholipids on the oxidation stability could be

due to the polar head groups. However, since both phospholipids tested had the same

head group at equal molar concentrations this is not likely. One major difference between

DOPC and DC4PC in SSO is their ability to form physical structure with DOPC forming

reverse micelles and DC4PC forming cylindrical structures (188). In addition, this study

showed that DOPC could form physical structures at much lower concentrations than

DC4PC due to its lower critical micelle concentration. Overall, these data suggested that

the prooxidant activity of DOPC could be related to its ability to form reverse micelles.

4.3.2 The impact of phospholipids on the oxidative stability of SSO in the presence

Page 92: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

72

of α-tocopherol

Previous studies of the influence of antioxidant polarity on lipid oxidation concluded

that polar antioxidants inhibit lipid oxidation in bulk oil systems more effectively than

non-polar antioxidants (156, 194, 195). A number of studies have previously examined

the effects of combinations of phospholipids and antioxidants on lipid oxidation in bulk

oils, such as α-tocopherol, flavonoids, etc (90, 196, 197). Some of these studies used

lecithin with a high proportion of PE as the phospholipid source, some of them used

commercial oils with unknown intrinsic antioxidant levels, and some used model systems

like triolein, methyl linoleate, or methyl laurate which may not accurately reflect

commercial bulk oil compositions (197). For this reason, we used a model system that

consisted of stripped soybean oil to mimic commercial oil triacylglycerol compositions

and well-defined phospholipids and antioxidants to provide a more fundamental

understanding of the complex physicochemical phenomenon involved.

The formation of lipid hydroperoxides (LH) and headspace hexanal in SSO

containing different levels of phospholipid (0 and 1000 μM DOPC) and the non-polar

antioxidant α-tocopherol (0, 10 and 100 μM) was measured during incubation at 55 °C

(Figures 4.2). As discussed earlier, in the absence of phospholipid and antioxidants the

lag phase of lipid hydroperoxides (LH) and hexanal formation for the control was 3 days

for the SSO. The incorporation of 1000 μM DOPC reduced the lag phase of LH and

hexanal formation to 2 days, again indicating that this phospholipid was prooxidative. In

the absence of DOPC, the duration of the lag phase increased with increasing α-

tocopherol concentration compared to the control. The addition of 10 and 100 μM α-

tocopherol increased the lag phase for LH to 9 and 27 days, respectively and headspace

Page 93: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

73

hexanal formation to 7 and 27 days, respectively. When DOPC was present, the impact of

α-tocopherol on lipid oxidation was more complex (Figures 4.2 A and B). DOPC

enhanced the antioxidant activity of 10 μM α-tocopherol by extending the lag phase of

LH formation from 9 to 11 days (Figure 4.2 A), and hexanal formation from 7 to 11 days

(Figure 4.2 B). The opposite trends were observed when 100μM α-tocopherol and

DOPC were added to the SSO. In this case, the lag time was shorter (less oxidatively

stable) in the presence of DOPC than in its absence (Figure 4.2 A and B). For example,

DOPC decreased the lag phase of LH formation from 27 to 23 days (Figure 4.2 A) and

hexanal formation from 27 to 20 days (Figure 4.2 B) in the presence of 100 μM α-

tocopherol.

Page 94: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

74

Figure 4.2. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) with or with 10 or100 M α-tocopherol during storage at 55 °C. Some of the error bars are within data points.

The impact of DC4PC on the antioxidant activity of α-tocopherol was also

determined (Figure 4.3A and B). As observed previously, DC4PC did not impact with

lipid hydroperoxide or hexanal formation rates compared to the control. α-Tocopherol

again decreased oxidation rates but in this case DC4PC decreased the effectiveness or α-

0 5 10 15 20 25 30 35 400

100

200

300

400

500

600

700

800

900 Blank (SSO) SSO w/ 1000 uM DOPC SSO w/ 1000 uM DOPC +10 uM TOC SSO w/ +10 uM TOC SSO w/ 1000 uM DOPC +100 uM TOC SSO w/ +100 uM TOC

Lip

id H

yd

rop

ero

xid

es (

mm

ol/ K

g o

il)

Storage time (day)

A

0 5 10 15 20 25 30 35 400

500

1000

1500

2000

2500

3000

3500

4000 Blank (SSO) SSO w/ 1000 uM DOPC SSO w/ 1000 uM DOPC +10 uM TOC SSO w/ +10 uM TOC SSO w/ 1000 uM DOPC +100 uM TOC SSO w/ +100 UM TOC

He

xa

na

l (m

icro

mo

l/K

g o

il)

Storage time (day)

B

Page 95: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

75

tocopherol at both 10 and 100 µM unlike DOPC which enhances antioxidant activity only

at 10 µM.

Figure 4.3. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 M of 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) with or with 10 or100 M α-tocopherol during storage at 55 °C. Some of the error bars are within data points.

4.3.3 The impact of phospholipids on the oxidative stability of SSO in the presence

0 5 10 15 20 25 300

100

200

300

400

500

600

700

800

900 Blank (SSO) SSO w/ 1000 uM DC4PC SSO w/ 1000 uM DC4PC +10 uM TOC SSO w/ +10 uM TOC SSO w/ 1000 uM DC4PC +100 uM TOC SSO w/ +100 uM TOC

Lip

id H

yd

rop

ero

xid

es (

mm

ol/ K

g o

il)

Storage time (day)

A

0 5 10 15 20 25 30

0

300

600

900

1200

1500 Blank (SSO) SSO w/ 1000 uM DC4PC SSO w/ 1000 uM DC4PC +10 uM TOC SSO w/ +10 uM TOC SSO w/ 1000 uM DC4PC +100 uM TOC SSO w/ +100 UM TOC

He

xa

na

l (m

icro

mo

l/K

g o

il)

Storage time (day)

B

Page 96: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

76

of Trolox

The impact of different concentrations of Trolox (10 and 100 μM), the polar analog

of α-tocoherol, on the oxidative stability of SSO with or without 1000 μM DOPC was

also investigated. Again, the formation of LH and hexanal were determined during

storage at 55 °C (Figure 4.4 A and B). Trolox at 10 µM increased the lag phase for both

lipid hydroperoxides and headspace hexanal formation from 3 (control) to 14 days. When

10 µM Trolox was added to the SSO in combination with 1000 μM DOPC the lag phase

for lipid hydroperoxides increased from 14 to17 days while headspace hexanal formation

was similar in the presence and absence of DOPC. Increasing the concentration of Trolox

to 100 μM further decreased lipid oxidation rates with no appreciable increase in lipid

hydroperoxides or hexanal after 37 days of storage (again more effective thanα-

tocopherol). However, like α-tocopherol, the presence of 1000 μM of DOPC decreased

the antioxidant activity of 100 µM Trolox with the lag phase of LH and hexanal

formation decreasing to 27 and 20 days, respectively.

0 5 10 15 20 25 30 35 400

100

200

300

400

500

600

700

800

900 Blank (SSO) SSO w/ 1000 uM DOPC SSO w/ 1000 uM DOPC +10 uM Trolox SSO w/ +10 uM Trolox SSO w/1000 uM DOPC +100 uM Trolox SSO w/ +100 mciroM Trolox

Lip

id H

yd

rop

ero

xid

es (

mm

ol/ K

g o

il)

Storage time (day)

A

Page 97: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

77

Figure 4.4. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) with or with 10 or 100 M Trolox during storage at 55 °C. Some of the error bars are within data points.

In the presence of DC4PC, the antioxidant activity of Trolox decreased (Figure 4.5A

and B). At a concentration of 10 M Trolox, DC4PC decreased the lag phase for

hydroperoxide formation from 9 to 7 days and hexanal formation from 9 to 5 days. At a

concentration of 100 M Trolox, DC4PC decreased the lag phase for hydroperoxide and

hexanal formation to 25 days (no LH or hexanal formation was observed for Trolox

alone).

0 5 10 15 20 25 30 35 400

500

1000

1500

2000

2500

3000

3500

4000 Blank (SSO) SSO w/ 1000 uM DOPC SSO w/ 1000 uM DOPC +10 uM Trolox SSO w/ +10 uM Trolox SSO w/1000 uM DOPC +100 uM Trolox SSO w/ +100 uM Trolox

Hexa

nal (m

icro

mo

l/K

g o

il)

Storage time (day)

B

Page 98: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

78

Figure 4.5. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 1000 M of 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) with or with 10 or100 M Trolox during storage at 55 °C. Some of the error bars are within data points

Overall, both α-tocopherol and Trolox inhibited lipid hydroperoxide and hexanal

formation with increasing concentrations from 10 to 100 µM increasing antioxidant

activity. The more polar Trolox was more effective than α-tocopherol which is similar to

trends reported by other researchers (156).

0 5 10 15 20 25 30 35 40

0

100

200

300

400

500

600

700

800

900 Blank (SSO) SSO w/ 1000 uM DOPC SSO w/ 1000 uM DC4PC +10 uM Trolox SSO w/ +10 uM Trolox SSO w/1000 uM DC4PC +100 uM Trolox SSO w/ +100 mciroM Trolox

Lip

id H

yd

rop

ero

xid

es (

mm

ol/ K

g o

il)

Storage time (day)

A

0 5 10 15 20 25 30 35 40

0

300

600

900

1200

1500 Blank (SSO) SSO w/ 1000 uM DOPC SSO w/ 1000 uM DC4PC +10 uM Trolox SSO w/ +10 uM Trolox SSO w/1000 uM DC4PC +100 uM Trolox SSO w/ +100 uM Trolox

He

xa

na

l (m

icro

mo

l/K

g o

il)

Storage time (day)

B

Page 99: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

79

4.3.4 The impact of antioxidants on the properties of DOPC reverse micelles in SSO

Aqueous self-assembly of amphiphilic surfactants to form association colloids are

recognized to be driven primarily by hydrophobic interactions (198). In non aqueous

systems, uncertain effects could also drive the self-assembly of amphiphiles into

association colloids and this is referred to as reverse self-assembly (199). Our previous

research illustrated 1000 µM of amphiphilic DOPC plus 200 ppm water resulted in the

formation of reverse micelles, a type of association colloid. Here, the physical structures

of the DOPC/water reverse micelles containing different concentrations of antioxidants

were measured using small angle x-ray scattering (SAXS) and the fluorescence steady

state and lifetime decay of the surface active fluorescent probe, NBD-PE. These studies

were not carried out with DC4PC as this phospholipid does not form association colloids

at the concentrations used in this study.

SAXS is a unique technique which has been widely used to distinguish structure

transitions in reverse micelles. In our previous study, SAXS was able to determined that

DOPC and water formed reverse micelles and that the size and shape of the reverse

micelles changed with varying water concentrations. The same technique was employed

in this study to determine if α-tocopherol and Trolox had any impact on the structure of

DOPC reverse. SAXS patterns revealed that the scattering patterns of DOPC reverse

micelles were not altered by either α-tocopherol or Trolox (data not shown) indicating

that the antioxidants did not have a major impact on the size and shape of the reverse

micelles. The only difference of SAXS between the samples was scattering intensity;

however, this difference did not increase with increasing antioxidants concentration

Others reported the phospholipids aliphatic chain is the main factor that impacts the

Page 100: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

80

scattering intensity of phospholipids structures (200). Therefore, the intensity differences

caused by the antioxidants could be due to their association with the phospholipid acyl

chains thus causing attenuation of the x-ray scattering.

Fluorophores have been reported to provide valuable insights into

microenvironmental changes in reverse micelles (164). Previously, we selected NBD-PE,

a phospholipid analog grafted with a fluorophore on the phosphate head groups to study

the interfacial properties of DOPC reverse micelle systems. NBD-PE emission intensity

increases in a polar environment. The lowest recorded emission of NBD-PE was

observed in the SSO control. The addition of DOPC and water in SSO increased the

emission to 8.5 ×105 counts, 30% higher than in SSO alone. Neither α-tocopherol nor

Trolox caused a shift in the fluorescence emission peak wavelength (data not shown). The

addition of 10 µM α-tocopherol decreased the emission intensity of NBD-PE in the

DOPC reverse micelles but further increases in α-tocopherol did not further decrease

NBD-PE fluorescence (Figure 4.6). Conversely, a concentration dependent decrease in

NBD-PE fluorescence occurred in the presence of Trolox.

Page 101: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

81

Figure 4.6. The neutralized fluorescence intensity of NBD-PE at stripped soybean oil containing 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine(DOPC) with varying

concentrations(0-100 M) of α-tocopherol and Trolox.

The low NBD-PE emission in the SSO control is likely due to the water

concentrations (< 60 ppm in pure SSO). When DOPC and water were added to the SSO,

NBD-PE was incorporated into the reverse micelles thus placing it in a more polar

environment thus increasing its emission intensity. The ability of increasing

concentrations of the water soluble Trolox to decrease fluorescence intensity suggests to

that Trolox can interact with water in the reverse micelle thus decreasing NBD-PE-water

interactions and thus fluorescence intensity. Additionally, the polar Trolox molecules

could directly interact with NBE-PE thus decreasing fluorescence intensity by decreasing

the ability of NBD-PE to become excited. Overall, this data suggests that both the NBD-

PE probe and Trolox are incorporated into the DOPC reverse micelles.

0 10 20 30 40 50 60 70 80 90 1000.5

0.6

0.7

0.8

0.9

1.0

No

rma

lize

d I

nte

ns

ity

Antioxidant concentration(M)

Toc Trolox

Page 102: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

82

Figure 4.7. The fluorescence lifetime decay of NBD-PE at stripped soybean oil containing 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine(DOPC) with varying concentrations(0-100 M) of α-tocopherol (A) and Trolox (B) concentrations.

Information about the properties of the antioxidants in the reverse micelles can also

be gained by measuring the lifetime decay of NBD-PE. Overall, neither α-tocopherol not

Trolox had a major impact on the fluorescence lifetime of NBE-PE (Figure 4.7A and B).

125 130 135 140 145 150 155 160

0

50

100

150

200

250

Inte

ns

ity

(A

.U.)

Time(ns)

SSO+NBD

SSO+NBD+DOPC 10M Toc 50M Toc

100M Toc IRF

A

125 130 135 140 145 150 155 160

0

50

100

150

200

250

Inte

nsit

y (

A.U

.)

Time(ns)

SSO+NBD SSO+NBD+DOPC 10M Trolox

50M Trolox 100M Trolox IRF

B

Page 103: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

83

However, there was a slight trend of Trolox decreasing the lifetime of NBE-PE in DOPC

reverse micelles. This again suggests that both the NBD-PE probe and Trolox are

incorporated into the DOPC reverse micelles.

4.4 Discussion

The main motivation of investigating the impact of DOPC based reverse micelles in

stripped soybean oil on the physicochemical properties of antioxidants was to better

understand the underlying phenomena of lipid and antioxidant interactions in bulk oil.

The parameters evaluated were phospholipids with the same polar head groups and

different fatty acyl chains (DC4PC and DOPC) that would result in formation of different

types of physical structures. Our previous study showed that DOPC but not DC4PC

formed reverse micelles in SSO10. In addition, DOPC reverse micelles accelerated lipid

oxidation while DC4PC, which does not form physical structure had no impact on lipid

oxidation rates. This data suggest that lipid oxidation in bulk oils is not only influenced

by the traditional chemical factors, such as lipid compositions, transition metals and

antioxidants, but is also related to the existence of physical structures of. Since one of the

most effective techniques employed for retarding the lipid oxidation is the addition of

free radical scavengers, this study was conducted as an extension of the previous study to

determine how physical structure impact the activity of antioxidant in bulk oils. In this

study, α-tocopherol and Trolox were chosen as hydrophobic and hydrophilic antioxidants

that have similar free radical scavenging functional groups (e.g same chromanol ring)

with different hydrocarbon tails (201).

DOPC by itself promoted lipid oxidation while DC4PC had no effect. This would

Page 104: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

84

suggest that the physical structure formed by DOPC increased lipid oxidation rates since

both DOPC and DC4PC have the same polar head groups and thus would impact the

chemistry of lipid oxidation in a similar manner. This is supported by the previous

observation that DOPC below its critical micelle concentration, where no reverse

micelles are formed, does not promote lipid oxidation (188). The prooxidative effect of

DOPC suggest that reverse micelles facilitate lipid oxidation.

One would expect the prooxidant activity of DOPC reverse micelles to decrease the

activity of the antioxidants. However, at low antioxidant concentrations, DOPC increased

the activity of both Trolox and α-tocopherol. Koga and Terao (181) have postulated that

physical structures formed by phospholipids can increase the antioxidant activity of

tocopherols by allowing them to concentration at the oil-water interface where lipid

oxidation is most prevalent. However, increasing α-tocopherol and Trolox concentrations

to 100 µM in the presence of a constant DOPC concentrations resulted in a reduction in

the effectiveness of the antioxidant. Antioxidants such as tocopherols in bulk oils have

been shown to lose effectiveness as their concentrations increase and this could help to

explain this observation. However, loss of effectiveness with increasing concentrations is

not typically seen at the α-tocopherol and Trolox concentrations used in this study (202).

A simple explanation for this observation is not available but it could be due to

differences in the number and/or type of association colloids found in this simple model

system compared to commercial oils that would contain a much greater variety of

endogenous surface active molecules. Another possible explanation is that the DOPC

reverse micelles used in this study would produce a negatively charged interface that

could attract prooxidative transition metals. Partitioning of high concentrations of α-

Page 105: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

85

tocopherol and Trolox into the same physical location as the transition metals could allow

the antioxidants to reduce the metals into their more prooxidative state and thus increase

lipid oxidation rates which would be seen as a decrease in the effectiveness of the

antioxidants. This pathway could be further increased by the accumulation of lipid

hydroperoxides, the substrate for metal promoted lipid oxidation, into these same

structures since they have been reported to partition into reverse micelle structures in oils

and act as the co-surfactant (145). The presence of DC4PC only decreased the activity of

α-tocopherol and Trolox indicating that it was not able to enhance the activity of the

antioxidants.

Trolox was a more effective antioxidant in the presence of the DOPC micelles than

α-tocopherol which is in agreement with previous research (156). The fluorescence of

NBD-PE was affected to a much greater extent by Trolox than α-tocopherol suggesting

that Trolox was partitioning into the DOPC reverse micelles differently than α-tocopherol.

This could be due to Trolox’s higher water solubility which would allow it to partition

into the water phase while α-tocopherol has virtually no water solubility and thus at best

would align at the oil-water interface. Differences in the physical location of Trolox and

α-tocopherol could alter the overall impact on lipid oxidation by influencing their

antioxidant (free radical scavenging) or prooxidant (reduction of transition metals)

activity.

This study showed that the antioxidant activity of both Tolox and α-tocopherol were

influenced by physical structures in bulk oils. Work such as this can provide important

data on how physical structures in bulk oils impact the activity of antioxidants. Since few

new antioxidants are available for food applications, this information might provide

Page 106: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

86

information on how to alter the physical structures in bulk oils such that we can utilize

the available food antioxidants more effectively.

CHAPTER 5

Page 107: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

87

IRON/ANTIOXIDANTS RELATIONSHIP IN BULK OIL OXIDATION

5.1 Introduction

Lipid oxidation is one of the major factors limiting the shelf life of bulk oils, since it

adversely affects flavor and quality by forming a very complex mixture of lipid

hydroperoxides, fatty acid chain-cleavage products, and polymeric materials (203,204).

In addition, lipid oxidation presents food safety concerns as low concentrations of lipid

oxidation products, such as 4-hydroxynonenal (4-HNE) and malonaldehyde (MDA) can

promote inflammation, atherosclerosis, neurodegenerative diseases and cancer (205).

The current dietary trends to consume healthier oils, such as omega-3 fatty acids is a

major challenge for food scientists since these oils are extremely susceptible to lipid

oxidation (206). Therefore, new strategies to prevent or inhibit the oxidation of bulk oils

are of major importance to consumers and the food industry. Based on the current

knowledge of lipid oxidation mechanisms, some approaches have been developed to

fulfill this goal. The application of antioxidants can retard lipid oxidation (8). But in

many foods the currently approved antioxidant are not sufficient to prevent rancidity.

Since very few new antioxidants are available to the food industry, new technologies are

needed to enhance the effectiveness of currently available food-grade antioxidants.

Bulk oil is a deceptively complex food system since it contains numerous minor

components, such as free fatty acid, sterols, antioxidants, phospholipids,

monoacylglycerol, and water (121,207). Recent studies from our group found that in bulk

oils some of these minor components are able to form physical structures known as

association colloids. These physical structures can increase lipid oxidation reactions and

alter the effectiveness of antioxidants, such as α-tocopherol, Trolox, and chlorogenic acid

Page 108: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

88

esters (188,208,209). For example, the more polar Trolox was able to associate with

association colloids more effectively than nonpolar -tocopherol. Since association

colloids increase lipid oxidation rates, the association of Trolox with these association

colloids could help explain by it’s a more effective antioxidant than -tocopherol.

The role of transition metals on lipid oxidation in food systems has been intensively

studied (210,211). Low valence-state metal ions, such as ferrous and cuprous, can

participate in the initiation and propagation steps of lipid oxidation by abstracting

hydrogen to directly form free radicals and decomposing lipid hydroperoxides into free

radicals such as the alkoxyl radical (212). The prooxidative effect of high-valence-state

metal ions is less clear although there is some evidence that they can promote oxidation

when they are reduced by antioxidants to form the more prooxidative low valence state of

the metal ions (213,214). For example, ascorbic acid and gallic acid can promote lipid

oxidation by reducing iron to its low valence-state (215,216).

Vegetable seed oils inevitably contain transition metals which originate from the

seed and/or manufacturing equipment and ingredients suggesting that they could be

important prooxidants in bulk oils. Tocopherols are important antioxidants in bulk oils

but they have been reported to have prooxidant activity in vegetable oils at high

concentrations (217,218). While association colloids in the bulk oil seem to act as

important nano-reactors in bulk oils, it is unclear what promotes this oxidation. In

addition, since molecules such as Trolox and tocopherols can also have prooxidative

activity, it is possible that association colloids could negatively impact lipid oxidation by

enhancing the prooxidant activity of antioxidants. Therefore, the present investigation

was undertaken to explore if iron was an important prooxidant in bulk oil containing

Page 109: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

89

association colloids. In addition, the ability of high or low valence state iron to interact

with α-tocopherol and Trolox was also determined. Iron-antioxidant interactions were

determined in a model system containing non-oxidizable medium chain triacylglycerol

(MCT) by measuring the depletion of antioxidants and the formation of free radicals.

5.2 Materials and methods

5.2.1 Materials

1,2-Dioleoyl-sn-glycero-3-phosphocholine (DOPC) was acquired from Avanti

Polar Lipids, Inc (Alabaster, AL). Soybean oil was purchased from a local store and

stored at 4 °C. Medium chain triacylglycerols (MCT, Miglyol®) were purchased

from Sasol North America Inc (Houston, TX). Silicic acid, activated charcoal,

calcein (CA), deferoxamine (DFO), methanol, hexane, alpha-tocopherol, Trolox,

and N-t-butyl-phenylnitrone (PBN) were purchased from Sigma-Aldrich Co. (St.

Louis, MO). Anhydrous ferric chloride and anhydrous ferrous chloride beads

(99.998 % purity) were purchased from Sigma-Aldrich Co.,(St. Louis, MO). All

other reagents were of HPLC grade or purer distilled and deionized water was used

as needed.

5.2.2 Formation of DOPC association colloids in stripped soybean oil or MCT

Stripped soybean oil with DOPC reverse micelles were prepared according to the

method of Chen et al (188). Briefly, formation of reverse micelles was accomplished by

pipetting DOPC (1000 μM, final concentration) in chloroform into an empty beaker and

then flushing with nitrogen until the chloroform was evaporated. The appropriate amount

Page 110: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

90

of medium chain triacylglycerols (MCT) or stripped soybean oil (SSO) was then added

followed by double distilled water at a final concentration of 200 ppm. The oil samples

were then stirred in a beaker at 1000 rpm in a 20 °C incubator room for a 24 h.

5.2.3 Measurement of oxidation parameters

For lipid oxidation kinetics studies antioxidants were incorporated into stripped

soybean oil by stirring at 300 rpm for at least 12 h at room temperature to obtain the

homogenous samples. Then, samples were aliquoted (1 mL/vial) into GC headspace vials

and stored at 55 °C in the dark.

Lipid hydroperoxides were measured as the primary oxidation products using a

method adapted from Shanta and Decker (165). Oil samples were weighed and recorded

before adding to a mixture of methanol/butanol (2.8 mL, 2:1, v:v) followed by addition of

15 μL of 3.94 M thiocyanate and 15 μL of 0.072 M Fe2+. The solution was vortexed, and

after 20 min the absorbance was measured at 510 nm using a Genesys 20

spectrophotometer (ThermoSpectronic, Waltham, MA). The concentration of

hydroperoxides was calculated from a cumene hydroperoxide standard curve.

The secondary oxidation product, hexanal, was monitored using a GC-17A

Shimadzu gas chromatograph equipped with an AOC-5000 autosampler (Shimadzu,

Kyoto, Japan). Samples (1 mL) in 10 mL glass vials capped with aluminum caps with

PTFE/silicone septa were heated at 55 °C for 15 min in an autosampler heating block

before measurement. A 50/30 μm DVB/ Carboxen/PDMS solid-phase microextraction

(SPME) fiber needle from Supelco® (Bellefonte, PA, USA) was injected into the vial for

2 min to absorb volatiles and was then transferred to the injector port (250 °C) for 3 min.

Page 111: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

91

The injection port was operated in split mode, and the split ratio was set at 1:5. Volatiles

were separated on a Supleco 30 m × 0.32 mm Equity DB-1 column with a 1 μm film

thickness at 65°C for 10 min. The carrier gas was helium at 15.0 mL/min. A flame

ionization detector was used at a temperature of 250°C. Hexanal concentrations were

determined from peak areas using a standard curve prepared from an authentic standard.

5.2.4 Determination of Trolox and α-tocopherol concentrations

For HPLC analysis, 100 μL of each sample was dissolved in 0.9 ml methanol and

then was passed through a 0.45 μm filter. A 20 μl aliquot of this sample solution was

separated using a Shimadzu HPLC system equipped with a Shimadzu diode array

detector (DAD), a Waters 474 Scanning Fluorescent Detector and a 250 mm × 4.6 mm

i.d., 5 μm, Inertsil C18 analytical column. The mobile phase consisted of 4 % purified

water with 3 mM phosphoric acid at pH 2.6 (solvent A) and 96 % methanol (solvent B)

using isocratic gradient at a flow rate of 1 ml/min. Column temperature was set at 38 °C.

Trolox was detected using DAD at a wavelength of 280 nm. Detection of α-tocopherol

was conducted using both DAD at 295 nm and fluorescence at an excitation wavelength

of 290 nm and an emission wavelength of 330 nm. Trolox and α-tocopherol in the

samples were identified by comparing their relative retention times and UV spectra with

authentic compounds and were quantitated using an external standard method.

Depletion of antioxidants was determined using At/A0×100%

Where Ao and At were the peak area at time zero and t, respectively.

Page 112: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

92

5.2.5 Determination the concentration of ferric iron in MCT

Ferric iron content was determined by UV-visible spectrophotometric method using

calcein (CA) dye method developed by Thomas et al (219). Briefly, calcein solution was

prepared by adding 0.75 g of calcein in 400 g MCT and stirred overnight to dissolve the

dye. The solution was centrifuged at 10,000 rpm for 10 min and upper clear orange color

solution was collected and defined as the dye solution. Equal volume of dye solution was

added directly to bulk oil samples containing ferric iron and the absorbance was

measured at 352 nm using a Shimadzu UV-2101 PC UV−vis scanning spectrophotometer

(Shimadzu, Kyoto, Japan).

Remaining ferric iron in MCT was determined using Abt/Ab0×100%

Where Abo and Abt were the absorbance at time zero and t, respectively.

5.2.6 EPR spectroscopy

EPR measurements were carried out at room temperature using a Bruker EPR

Elexsys-500 spectrometer operating at the X-band. The samples were placed in 707-SQ-

250 M, thin wall quartz EPR sample tubes (Wilmad Glass Co., Buena, NJ, USA) and

inserted into the ER 4122-SHQE high sensitivity TE102 cylindrical mode single cavity

(Q ~ 3000) optical window of the EPR system. Instrument settings were: center field,

3470 G; scan range, 100 G; gain, 60; time constant, 128 ms; modulation amplitude, 1 G;

phase 0°; microwave power, 20 mW. Data collection was performed using the

computerized program Xepr.

Page 113: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

93

5.2.6 Statistical analysis

Duplicate experiments were performed with freshly prepared samples. All data

shown represents the mean values ± standard deviation of triplicate measurements.

Statistical analysis of lipid oxidation kinetics was performed using a one-way analysis of

variance. A significance level of p<0.05 between groups was accepted as being

statistically different. In all cases, comparisons of the means of the individual groups

were performed using Duncan’s multiple range tests. All calculations were performed

using SPSS17 (http://www.spss.com; SPSS Inc., Chicago, IL).

5.3 Results

5.3.1 Effects of the metal chelator, Deferoxamine (DFO) on the oxidative stability of

stripped soybean oil

Transition metals are implicated in many lipid oxidation pathways as they are able to

generate free radicals that initiate and propagate lipid oxidation in food systems

(220,221). The different redox states of iron have different pathways by which they can

promote oxidation (222). For example, low-valence transition metals can promote the

decomposition of lipid hydroperoxides into free radicals such as the alkoxyl radical

(reaction 1). The high-valence state of transition metals can also decompose lipid

hydroperoxide (reaction 2) but this pathway is very slow and may not be very important

in foods. However, compounds such as antioxidants can reduce metals resulting in the

formation of the highly prooxidative low-valence state thus promoting oxidation

(reaction 3) (207,223).

Fe2+ + LOOH → Fe3+ + LO•+-OH (reaction 1)

Page 114: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

94

Fe3+ + LOOH → Fe2+ + LOO•+H+ (reaction 2)

Fe3+ + AH→ Fe2++A•+H+ (reaction 3)

Bulk oil naturally contains transition metals. Therefore, gaining a better

understanding of the chemical role of transition metals in bulk oil oxidation is necessary

to develop effective prevention strategies to extend the shelf-life of oils. To gain a better

understanding of the role of iron on lipid oxidation in bulk oils, we studied the effect of

deferoxamine (DFO), a specific Fe(III) chelator, on the oxidative stability of stripped

soybean oil (SSO) in the absence and presence of phospholipid reverse micelles.

The formation of lipid hydroperoxides (LH) and headspace hexanal in SSO

containing different levels of phospholipids (0 and 1000 μM DOPC) and DFO (0 and 2

mM) was measured during incubation at 55 °C (Figures 5.1 A and B). The lag phase for

both lipid hydroperoxides and headspace hexanal for SSO lasted 4 days. The addition of

DOPC reduced the lag phase by 2 days, which is in agreement with our previous study

indicating that DOPC reverse micelles act as a prooxidant (209). The incorporation of

DFO had no significant effect on the lag phase of SSO (p>0.05), but extended the lag

phase of SSO containing DOPC from 2 to 7 days (p<0.05).

Page 115: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

95

Figure 5.1. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 2 mM deferoxamine (DFO) and/or 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) during storage at 55 °C. Some of the error bars are within data points.

In order to understand if iron played a role in lipid oxidation in the presence of free

radical scavenging antioxidants, different combinations of DFO, α-tocopherol and Trolox

were added to the SSO in the presence and absence of phospholipids reverse micelles

0 2 4 6 8 10 12 140

200

400

600

800

Lip

id H

ydro

per

oxid

es (

mm

ol/

Kg

oil)

Storage time (day)

SSO SSO + DFO SSO + DOPC SSO + DOPC +DFO

A

0 2 4 6 8 10 12 14

0

200

400

600

800

1000

B

Hex

an

al (

mol

/Kg o

il)

Storage time (day)

SSO SSO + DFO SSO + DOPC SSO + DOPC + DFO

Page 116: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

96

during incubation at 55 °C (Figures 5.2 and 5.3). The lag phase of lipid hydroperoxides

(LH) and hexanal formation was 25 days for SSO containing 100 μM α-tocopherol

(Figures 5.2 A and B). In the presence of DOPC reverse micelles and -tocopherol, the

lag phase decreased to 16 days for both LH and hexanal formation, again showing that -

tocopherol was less effective in the presence of DOPC reverse micelles (188). DFO

increased the lag phase of LH and hexanal formation in the presence of -tocopherol

both in the presence and absence of DOPC reverse micelles. In the absence of DOPC

reverse micelles, the lag phase was increased by 5 days while in the presence of DOPC

reverse micelles the lag phase was increased by 10-11 days. Unlike the previous

experiment, DFO was able to inhibit lipid oxidation in the absence of DOPC micelles

suggesting that -tocopherol was somehow increasing the prooxidant activity of iron.

0 5 10 15 20 25 30 35

0

100

200

300

400

500

600

700

Lip

id H

ydro

per

oxid

es (

mm

ol/

Kg

oil)

Storage time (day)

SSO+Toc SSO+Toc+DFO SSO+DOPC+Toc SSO+DOPC+Toc+DFO

A

Page 117: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

97

Figure 5.2. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 100 M α-tocopherol, 2 mM deferoxamine (DFO) and/or 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) during storage at 55 °C. Some of the error bars are within data points.

The impact of DFO on the antioxidant activity of Trolox was also tested in the

presence and absence of DOPC reverse micelles (Figure 5.3 A and B). The lag phase of

lipid hydroperoxides (LH) and hexanal formation was 40 days for the SSO containing

100 μM Trolox showing that the more polar Trolox was a more effective antioxidant than

-tocopherol as predicted by the antioxidant polar paradox (224). However, unlike -

tocopherol, the incorporation of DFO did not increase the lag phase of LH and hexanal

formation in the absence of DOPC micelles. As observed previously, DOPC again

diminished the effectiveness of Trolox decreasing the lag phase to 21 days for both LH

and hexanal formation. Even though DOPC decreased the activity of both -tocopherol

and Trolox, the more hydrophilic Trolox was still the most effective of the two. When

DOPC micelles, Trolox and DFO were present in combination, DFO helped partially

offset the prooxidative effect of DOPC by increasing the lag phase for LH and hexanal

0 5 10 15 20 25 30 35

0

100

200

300

400

500

600

700

B

Hex

anal

(

mol

/Kg

oil

)

Storage time (day)

SSO+Toc SSO+Toc+DFO SSO+DOPC+Toc

SSO+DOPC+Toc+DFO

Page 118: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

98

formation to 29 days.

Figure 5.3. Formation of lipid hydroperoxides (A) and hexanal (B) in stripped soybean oil containing 100 M Trolox, 2 mM deferoxamine (DFO) and/or 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) during storage at 55 °C. Some of the error bars are within data points.

DFO, a powerful iron chelator, has been successfully used in clinical settings to treat

patients with acute and chronic iron overload syndromes (225). DFO has a remarkably

high affinity for ferric ion allowing it to inhibit the initiation and propagation of lipid

0 5 10 15 20 25 30 35 40 450

100

200

300

400

500

600

700

Lip

id H

ydro

per

oxi

des

(m

mol

/ K

g o

il)

Storage time (day)

SSO+Trolox SSO+Trolox+DFO SSO+DOPC+Trolox SSO+DOPC+Trolox+DFO

A

0 5 10 15 20 25 30 35 40 450

100

200

300

400

500

600

700

B

Hex

anal

(

mol

/Kg

oil)

Storage time (day)

SSO+Trolox SSO+Trolox+DFO SSO+DOPC+Trolox SSO+DOPC+Trolox+DFO

Page 119: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

99

oxidation (226). Therefore, it was anticipated that oxidative stability of bulk oil would be

improved by DFO if iron was an important prooxidant. In this study, DFO was not an

effective antioxidant in soybean oil stripped of its minor components including

endogenous antioxidants and phospholipids. However, DFO was effective in the

presence of DOPC reverse micelle indicating that the iron was an important prooxidant in

the micelles. This could occur since phospholipids can bind iron (227) and thus could

concentrate iron at the lipid-water interface of reverse micelles. Lipid hydroperoxides are

also surface active and could accumulate at the water-oil interface of the reverse micelles

(14). Thus, the prooxidant activity of DOPC could be due to their ability to concentrate

both endogenous iron and lipid hydroperoxides at the water-lipid interface of reverse

micelles thereby increasing the ability of iron to decompose lipid hydroperoxides.

The combination of the free radical scavengers, -tocopherol and Trolox, and DOPC

reverse micelles also impacted the ability of DFO to inhibit lipid oxidation. In the

absence of DOPC reverse micelles, DFO was a weak antioxidant in the presence of -

tocopherol (increased lag phase approximately 5 days) and did not inhibit oxidation in the

presence of Trolox. However, in the presence of DOPC reverse micelles and -

tocopherol or Trolox, DFO was a strong antioxidant increasing the lag phase of oxidation

to 11 and 13 days for -tocopherol and Trolox, respectively. This observation could be

due to the ability of DFO to inhibit iron-promoted generation of free radical that would

deplete the antioxidants in the DOPC reverse micelles system. However, it is also

possible that DFO could improve the ability of -tocopherol or Trolox by inhibiting

direct degradation by iron.

Page 120: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

100

5.3.2 Effects of Fe(III) and Fe(II) on the depletion of α-tocopherol and Trolox in

MCT

In order to better understand interactions between high and low-valence iron with -

tocopherol and Trolox, a medium chain triacylglycerides (MCT) model bulk oil system

was used. MCT was used in these experiments since is does not contain unsaturated fatty

acids and thus there would be no fatty acid oxidation that could alter α-tocopherol or

Trolox concentrations. Figure 5.4A shows the depletion of α-tocopherol and Trolox

during storage at 55 oC in the presence of high-valence Fe(III). The concentration of both

α-tocopherol and Trolox were relatively constant in the absence of Fe(III). Incubation of

α-tocopherol or Trolox with 150 M Fe(III) resulted in a similar trend in antioxidant

depletion with 15 and 12 % of α-tocopherol and Trolox consumed, respectively, after one

hour storage. α-Tocopherol and Trolox concentrations in the presence of 150 M Fe(III)

reached equilibrium (~50% consumed) after 12 h. When 600 M of Fe(III) was added,

the pattern of antioxidants consumption was different with α-tocopherol being completely

consumed after 8 h of storage. Increasing the Fe(III) concentration to 600 M had much

less effect on Trolox consumption with a maximum of approximately 60% depletion of

Trolox after 24 h. The observed consumption of -tocopherol and Trolox in the presence

of Fe(III) could be due to the ability of these antioxidants donate an electron to iron thus

converting ferric to ferrous ions. Since Fe(III) was more reactive with -tocopherol than

Trolox, this could help explain why DFO was able to improve the antioxidant activity of

-tocopherol but not Trolox in the absence of DOPC reverse micelles (Figure 5.2 and

5.3) as DFO could decrease -tocopherol’s ability to reduce ferric ions to the more

prooxidative ferrous ions and could also decrease Fe(III) promoted -tocopherol

Page 121: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

101

consumption which would help maintain a higher concentrations of -tocopherol thus

more effectively inhibiting lipid oxidation.

Figure 5.4. Depletion of α-tocopherol and Trolox (100 M) in medium chain triacylglycerols (MCT) in the presence of (A) Fe(III); and (B)Fe(II)

Since -tocopherol and Trolox consumption reached an equilibrium and did not

continue to decrease in concentrations, this suggests that Fe(II) did not cause antioxidant

consumption. To test this hypothesis, Fe(II) was also added to the MCT model system

0 10 20 30 40 50

0

20

40

60

80

100D

eple

tion

of

anti

oxid

an

t (%

)

Storage time(hrs)

100M Trolox+MCT 100M Toc+MCT

150M Fe3+

+100M Trolox

150MFe3+

+100M Toc

600M Fe3+

+100M Trolox

600M Fe3+

+100M Toc

A

0 10 20 30 40 50

0

20

40

60

80

100

B

Dep

leti

on o

f an

tiox

ida

nt

(%)

Storage time(hrs)

100M Trolox+MCT

100M Toc+MCT

150M Fe2+

+100M Trolox

150M Fe2+

+100M Toc

600M Fe2+

+100M Trolox

600M Fe2+

+100M Toc

Page 122: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

102

(Figure 5.4B). Initial experiments were performed with reagent grade ferrous chloride

powder but rapid -tocopherol and Trolox consumption was observed (data not shown).

It was thought that this could be due to contaminating Fe(III) so experiments were

repeated using a high purity form of anhydrous ferrous supplied under vacuum. In these

experiments, 150 M Fe(II) did not significantly decrease -tocopherol and Trolox for

the first 12 h of storage compared to a 42 and 50% decrease in -tocopherol and Trolox,

respectively, during the same time period in the presence of 150 M Fe(III) (Figure

5.4A). In the presence of 600 M Fe(II), Trolox consumption was more rapid than -

tocopherol. In all cases, -tocopherol and Trolox consumption increased during

prolonged storage but the level of consumption was always significantly lower than

Fe(III) (p>0.05, Figure 5.4A). The depletion of antioxidants during the later stages of

incubation may be due to the oxidation of ferrous to ferric upon reaction with oxygen.

5.3.3 Effects of Fe(III) and Fe(II) on the depletion of α-tocopherol or Trolox in the

presence of DOPC reverse micelles

Previous results showed that DFO more effectively increased the antioxidant activity

of α-tocopherol and Trolox in the presence of DOPC micelles (Figure 5.2 and 5.3). This

could be due to DFO’s ability to decrease iron-promoted decomposition of lipid

hydroperoxides into free radicals, a factor that would decrease -tocopherol and Trolox

consumption and thus increase their ability to scavenge free radicals. However, it is also

possible that the improved activity of DFO in DOPC reverse micelles could be due to its

ability to inhibit the direct consumption of -tocopherol and Trolox by iron which again

would increase antioxidant concentrations and would also decrease the reduction of ferric

Page 123: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

103

into the more prooxidative ferrous ions. To test this possibility, the depletion of α-

tocopherol and Trolox by Fe(III) or Fe(II) in MCT in the presence of DOPC reverse

micelles was determined (Figure 5.5 and 5.6). These experiments were also conducted

with DC4PC, a phosphatidylcholine with butyric acid. DC4PC does not form reverse

micelles at the concentrations tested (188), so experiments could be conducted with the

same concentration of the choline head group in the absence of association colloids.

Figure 5.5. Depletion of α-tocopherol and Trolox (100 M) in medium chain

0 10 20 30 40 500

20

40

60

80

100

Dep

leti

on o

f a

nti

oxid

an

t (%

)

Storage time(hrs)

150M Fe3+

+100M Trolox+DOPC

150M Fe3+

+100M Toc+DOPC

150M Fe3+

+100M Trolox+DC4PC

150M Fe3+

+100M Toc+DC4PC

A

0 10 20 30 40 50

0

20

40

60

80

100

Dep

leti

on o

f an

tiox

idan

t (%

)

Storage time(hrs)

600M Fe3+

+100M Trolox+DOPC

600M Fe3+

+100M Toc+DOPC

600M Fe3+

+100M Trolox+DC4PC

600M Fe3+

+100M Toc+DC4PC

B

Page 124: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

104

triacylglycerols (MCT) containing 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) in the presence of (A) 40 ppm Fe(III); and (B) 100 ppm Fe(III)

In the absence of Fe(III), neither DOPC nor DC4PC promoted the consumption of -

tocopherol and Trolox and (Figure 5.4A and Band -tocopherol and Trolox were

stable in the absence of added iron (data not shown). In the presence of 150 M Fe(III)

and DOPC reverse micelles, the consumption of both -tocopherol and Trolox was less

than in the absence of DOPC (Figure 5.5A). For example, after 24 h of incubation, -

tocopherol concentrations were 10% lower than time 0 in the presence of DOPC reverse

micelles compared to 40% lower in the absence of DOPC reverse micelles. In the

presence of 600 M Fe(III), Trolox was depleted more rapidly than -tocopherol in the

presence of DOPC reverse micelles while the opposite was true in the absence of DOPC.

Trolox has been found to more highly associated than -tocopherol with water in DOPC

micelles (212). This means that it could have greater contact with phosphatidylcholine

bound iron at the water-lipid interface and thus could be more rapidly depleted.

In the presence of DC4PC which does not form reverse micelles, Fe(III) was rapidly

consumed by both α-tocopherol and Trolox. At low (150 M) and high (600 M) Fe(III)

concentrations, DC4PC slightly increased both α-tocopherol and Trolox consumption

compared to the absence of phospholipid (Figure 5.6A and B). DC4PC would be

expected to bind iron as would any phosphatidylcholine. The ability of DC4PC to slightly

promote the consumption of α-tocopherol and Trolox could be due to its ability to

increase the solubility of iron in the oil by acting as a lipid soluble metal chelator and

thus making the iron more reactive if the iron chelated to DC4PC would still be redox

Page 125: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

105

active. Since DC4PC does not form reverse micelles at the concentrations tested, this

suggests that the ability of DOPC to inhibit both α-tocopherol and Trolox depletion was

not due to the phospholipid head group but instead was due to the presence of reverse

micelles.

Figure 5.6. Depletion of α-tocopherol and Trolox (100 M) in medium chain triacylglycerols (MCT) containing 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) in the presence of (A) 40 ppm Fe(II); and (B)100 ppm Fe(II)

5.3.4 Measurement of Fe(III) loss by α-tocopherol and Trolox

0 10 20 30 40 50

0

20

40

60

80

100

Dep

leti

on

of

an

tioxid

an

t (%

)

Storage time(hrs)

150M Fe2+

+100M Trolox+DOPC

150M Fe2+

+100M Toc+DOPC

150M Fe2+

+100M Trolox+DC4PC

150M Fe2+

+100M Toc+DC4PC

A

0 10 20 30 40 500

20

40

60

80

100

Dep

leti

on o

f an

tiox

idan

t (%

)

Storage time(hrs)

600M Fe2+

+100M Trolox+DOPC

600M Fe2+

+100M Toc+DOPC

600M Fe2+

+100M Trolox+DC4PC

600M Fe2+

+100M Toc+DC4PC

B

Page 126: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

106

Many researchers have determined iron concentration in bulk oil (228-230). For

example, Mendil and coworkers measured iron content in bulk oil with a flame atomic

absorption spectrometer (FAAS) and total iron was found to be 139, 127, 105, and 52

ppm (i.e., 200-800 M) in olive oil, hazelnut oil, sunflower oil and corn oil, respectively

(230). While this experiment is helpful to verify that iron exist in commercially refined

oils, it does not provide any information on the redox state of the iron. This is not

uncommon since analytic techniques to measure iron redox state in bulk oil are limited.

Verifying the existence of iron recycling pathway by measuring the redox form of iron

has been extremely helpful in understanding the role of iron redox form in lipid oxidation

in both in vivo systems and oil-in-water emulsion (231,232). Therefore, in this study we

initially attempted to track the concentration of Fe(III) in MCT with calcein

acetoxymethyl ester (calcein-AM), an Fe(III) specific probe used successfully in cells and

biological fluids (233). However, calcein-AM solubility in MCT was quite limited so

calcein was successfully substituted due to its higher solubility in hydrophobic solvents

(219). Preliminary studies showed that calcein had strong absorbance at 365nm in MCT

which was quenched by ferric chloride. It also showed a good linear relationship (r>0.99)

with Fe(III) in MCT over the range of 0-600 M Fe(III) (data not shown). Room

temperature was chosen in these experiments since the 55 °C used in the previous

experiments caused too rapid of a reaction between ferric ions and the antioxidants.

The ability of -tocopherol and Trolox to reduce Fe(III) was determined at 600 M

Fe(III) and varying concentrations of the antioxidants (Figure 5.7A). After 24 h of

incubation, Fe(III) concentrations only decreased slightly (<5%) in the absence of added

antioxidants. -Tocopherol significantly decreased Fe(III) at concentrations as low as 20

Page 127: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

107

M. Maximal decrease in Fe(III) concentrations was observed at -tocopherol

concentrations of 60 M. Trolox exhibited a similar trend on Fe(III) concentrations

although its maximal decrease in Fe(III) concentrations was 40 M. At 100 M, the

decreased in Fe(III) concentrations was slightly greater for Trolox than -tocopherol

(p<0.05).

DOPC reverse micelles by themselves caused a greater decrease in Fe(III) (Figure

5.7B) than in the absence of DOPC (Figure 5.7A) suggesting that the DOPC by itself

could reduce Fe(III) (Figure 5.7B). DOPC reverse micelles decreased the ability of both

-tocopherol and Trolox to reduce iron as significant decreases in Fe(III) were not seen

until -tocopherol and Trolox concentrations were greater than 40M. As in the absence

of DOPC reverse micelles, at 100 M, the decreased in Fe(III) concentrations was

slightly greater for Trolox than -tocopherol (p<0.05).

Overall, DC4PC (Figure 5.7C) changed the concentration of Fe(III) in a similar

manner to the absence of phospholipid (Figure 5.7A). For example, DC4PC had very

little impact on Fe(III) concentrations in the absence of antioxidants. In addition, both -

tocopherol and Trolox were able to significantly decreased Fe(III) at concentrations as

low as 20 M as seen in the absence of phospholipid. However, DC4PC decreased the

maximal decrease in Fe(III) concentrations by -tocopherol from 60 to 40 M.

Page 128: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

108

0

20

40

60

80

100 600M Fe

3+

600M Fe3+

+Toc

600M Fe3+

+Trolox

Rem

ain

ing o

f F

e3+(%

)

Antioxidant concentration(M)

0 20 40 60 80 100

aA

bb

c

c c

B

CC

C

C

A

0

20

40

60

80

100 600M Fe3+

+DOPC

600M Fe3+

+Toc+1000M DOPC

600M Fe3+

+Trolox+1000M DOPC

Rem

ain

ing

of

Fe3+

(%)

Antioxidant concentration(M)0 20 40 60 80 100

B

aAa

a

bb b

A A

B B B

Page 129: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

109

Figure 5.7. Fe(III) development in medium chain triacylglycerols (MCT) containing (A) α-tocopherol or Trolox; (B) 1000 M of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC); (C) 1,2-dibutyryl-sn-glycero-3-phosphocholine (DC4PC) after 24 h storage at room temperature. For each group, different letters on the top of columns represent significant differences (p < 0.05)

5.3.5 Effects of Fe(III) and Fe(II) on the formation of PBN spin adducts in MCT

Since both α-tocopherol and Trolox can scavenge free radicals generated in oils, one

may argue that the depletion of antioxidants in this research could be due to the presence

of free radicals. In order to clarify this potential pitfall and provide more evidence for the

different role of Fe(III) and Fe(II) in bulk oil, an EPR study was employed in an effort to

detect the formation of free radicals in our MCT model system under different conditions.

A comparison of the EPR signals of SSO (Figure 5.8A) and MCT (Figure 5.8B)

after 24 h storage at 55 oC showed that PBN trapped free radicals were only observed in

SSO indicating that there were no detectable free radicals in our MCT model system.

When antioxidants and iron were added to MCT, the formation of α-tocopheroxyl

(Figure 5.8C) or Trolox (Figure 5.8D) radicals were not detected. This was true in the

0

20

40

60

80

100 600M Fe3+

+DC4PC

600M Fe3+

+Toc+DC4PC

600M Fe3+

+Trolox+DC4PC

0

Rem

ain

ing

of F

e3+(%

)

Antioxidant concentration(M)20 40 60 80 100

aA

bB b

B

c C cC c C

C

Page 130: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

110

presence or absence of DOPC or DC4PC (data not shown). To determine if iron was

reactive in the MCT model, it was added in combination with cumene hydroperoxide. In

this system, Fe(II) was observed to form PBN trapped radicals (Figure 5.8E) but Fe(III)

was not (Figure 5.8F). This confirms that Fe(II) could be an active prooxidant in bulk

oils and that Fe(III) is much less reactive as has been observed oil-in-water emulsions.

Figure 5.8. EPR detection of radical formations in the reaction consisting of (A) 0.1 M PBN in stripped soybean oil; (B) 0.1 M PBN in MCT; (C) 0.1 M PBN+100 μM α-tocopherol in MCT; (D) 0.1 M PBN+100 μM Trolox in MCT; (E) 0.1 M PBN+10 mM cumene hydroperoxides+40 ppm Fe(II) in MCT; (F) 0.1 M PBN+cumene hydroperoxides+40 ppm Fe(III) in MCT after 24 h storage at 55 oC

5.4 Discussion

Cumene Peroxides+Fe3+

at 24 hrsF

C

PBN in Oxidized SSO (24 hrs)

Cumene Peroxides+Fe2+

at 24 hrs

B

A

E

PBN in MCT

PBN in MCT containing Fe3+

+TOC

D PBN in MCT containing Fe3+

+TROLOX

Field (G)

Page 131: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

111

DFO was able to inhibit lipid oxidation in SSO in the presence of DOPC reverse

micelles indicating that iron was an active prooxidant in the presence of association

colloids. Since refined oils contain association colloids due to the presence of surface

active minor components (e.g. phospholipids, free fatty acids, lipid hydroperoxides, etc.)

and water (121), this could help explain why the metal chelator citric acid is an effective

antioxidant in commercial oils. DFO was even more effective at increasing the oxidative

stability of oils with DOPC reverse micelles in the presence of -tocopherol and Trolox.

This could be due to the decreased iron-promoted generation of free radicals or direct

iron-promoted decomposition of antioxidants.

Many phenolic antioxidants, such as the chlorogenic, caffeic, sinapic and ferulic acid, can

reduce Fe(III) to Fe(II) in aqueous environments resulting in the consumption of the

antioxidant. For example, 1 molecule of caffeic acid was reported to reduce 9 atoms of

Fe(III) (234,235). Our result showed that -tocopherol and Trolox can reduce Fe(III) in

bulk oil as seen by both the ability of Fe (III) to decrease antioxidant concentrations

(Figure 5.4) and the ability of antioxidants to decrease Fe(III) concentrations (Figure

5.7B). In this study, DOPC reverse micelles decreased interactions between iron and -

tocopherol or Trolox as seen by both a decrease in iron-promoted -tocopherol and

Trolox decomposition and a decrease in the ability of -tocopherol and Trolox to

decrease Fe(III) concentrations. This suggests that DFO did not decrease the iron-

promoted decomposition of -tocopherol and Trolox. Therefore, it is more likely that the

ability of DFO to enhance the activity of -tocopherol and Trolox in DOPC reverse

micelles was due to its ability to inhibit iron-promoted generatation of free radicals

through pathways such as hydroperoxides decomposition. This could occur if iron

Page 132: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

112

concentrated at the water-lipid interface of the DOPC reverse micelles and increased its

ability to decompose surface active lipid hydroperoxides. Decrease production of free

radicals from decomposing lipid hydroperoxides would decrease the consumption of the

antioxidants thus increase their effectiveness.

CHAPTER 6

Page 133: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

113

CONCLUSION

This research has shown that 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) can

self-assemble in stripped soybean oil (SSO) to form thermodynamically stable

association colloid structure, most likely reverse micelles. Interfacial tension and

fluorescence spectrometry results showed the critical micelle concentration (CMC) of

DOPC in stripped soybean oil was 650 and 950 μM, respectively which are well within

the phospholipid concentrations found in refined vegetable oils. Light scattering

attenuation results showed water concentrations was a critical factor in DOPC reverse

micelle structure with a 500 ppm water threshold above which association colloids

altered the optical properties of stripped soybean oil. Small-angle X-ray scattering (SAXS)

and fluorescence probes again showed that water had a very strong impact on the

properties of the association colloids formed by DOPC.

Measurement of primary and secondary lipid oxidation products revealed that

reverse micelles formed by DOPC reverse micelles had a pro-oxidant effect, shortening

the lag phase of SSO at 55 °C. This research also showed that 1,2-dibutyl-sn-glycero-3-

phosphocholine (DC4PC) did not form association colloid in stripped soybean oil, nor did

it change lipid oxidation kinetics of stripped soybean oil indicating that it was not the

choline head group that responsible for the observed accelerated lipid oxidation but

instead was likely the reverse micelle itself. The addition of ferric ion chelator,

deferoxamine (DFO) in stripped soybean oil significantly prevented the lipid oxidation

caused by DOPC reverse micelles extending lag phase from 2 to 7 days. This indicated

that iron was a strong prooxidant in the DOPC reverse micelles. Lipid hydroperoxides

are also surface active and therefore likely concentrate in the DOPC reverse micelles.

Page 134: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

114

Therefore the proposed mechanism of oxidation in DOPC reverse micelles is iron

promoted lipid hydroperoxide decomposition into free radicals.

DOPC reverse micelles can also impact the physical location of antioxidants such as

non-polar α-tocopherol and its polar analog Trolox. The steady state fluorescence

intensity of the probe, N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)-1,2-dihexadecanoyl-sn-

glycero-3-phosphoethanolamine, triethylammonium salt (NBD-PE), revealed that both α-

tocopherol and Trolox were associated with DOPC reverse micelle in bulk oil. Trolox

primarily concentrated in the water pool of reverse micelle since it quenched the NBD-

PE intensity with increasing concentrations. A portion of α-tocopherol was also

associated with the aqueous phase as evidence by the fact that it did not impact NBD-PE

intensity when its concentration was increased from 50 to 100 μM. The shortened life

time of NBD-PE fluorescence by Trolox but not -tocopherol was provided evidence that

Trolox was more highly associated with the water pool of the DOPC reverse micelles.

Increasing the concentration of both -tocopherol and Trolox increased the stability of

bulk stripped soybean oil containing DOPC reverse micelles. Hydrophilic Trolox had

better antioxidant activity than hydrophobic α-tocopherol. This could be due to Trolox’s

ability to more highly associate with the water phase of the reverse micelles where lipid

oxidation is prevalent.

Since -tocopherol and Trolox both associate with the DOPC reverse micelles, it is

possible that they could also interact with iron at the oil-water interface as proposed in

Figure 6.1. High-valence (FeIII) and low-valence-state (FeII) iron could interact with

Trolox and α-tocopherol) causing their depletion or increasing the reactivity of iron if the

antioxidant caused iron reduction. However, DOPC reverse micelles minimized iron-

Page 135: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

115

antioxidant interactions as seen by their ability to decrease iron-promoted -tocopherol

and Trolox degradation as well as decrease the ability of the antioxidants to reduce Fe(III)

to Fe(II). It is unclear how the DOPC reverse micelles decrease interactions between iron

and -tocopherol or Trolox. It does not seem to be due to physical hindrance since the

reverse micelles protected both hydrophilic Trolox which would concentrate in the

aqueous phase and hydrophobic -tocopherol which would concentrate in the lipid phase.

Therefore, protection could be due to the phosphate group that would bind iron and

decrease its ability to interact with the antioxidants.

Figure 6.1. Proposed mechanism of bulk oil oxidation that contains reverse micelles

Overall, this research identified several important new insights into potential lipid

Page 136: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

116

oxidation mechanisms in bulk oil. First, it showed that DOPC can form physical

structures such as reverse micelles in vegetable oils at the concentrations of

phospholipids and water found in commercially refined oils. These physical structures

increase lipid oxidation rates especially oxidation promoted by iron. Antioxidants were

also shown to associate with the DOPC reverse micelles with the more hydrophilic

Trolox associating with the association colloids than -tocopherol. The evidence of

acceleration of oxidation by DOPC reverse micelles suggests that these physical

structures could be a major site for oxidation reactions which could help explain by

hydrophilic antioxidants are often more effective than hydrophobic antioxidants in bulk

oils.

The results of this research also identified a number of areas where future

experiments would be beneficial in developing knowledge of how association colloids

impact bulk oil oxidation mechanisms and could present new opportunities for inhibiting

lipid oxidation:

(i) Investigation the impact of free fatty acids on bulk oil oxidation containing

association colloids;

(ii) Detection dissolved oxygen in oil and its role on iron redox recycling;

(iii) Identification the reaction products between α-tocopherol or Trolx and Fe(III)

and their role on bulk oil oxidation;

(iv) Investigation the relationship between lipid hydroperoxides generation,

decomposition and the iron recycling in bulk oil system.

(v) Investigation the impact of other phospholipids, such as phosphatidylserine (PS)

and phosphatidylethanolamine (PE) on bulk oil oxidation, specifically on the physical

Page 137: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

117

structure formation and chemical stability.

The knowledge gained from this study will improve our understanding of causes and

prevention of lipid oxidation in bulk oils. It suggests that DOPC reverse micelles can

alter the microenvironment where chemical degradation reactions occur, such as lipid

oxidation. In addition, lipid oxidation in bulk oils is not only influenced by the traditional

chemical factors, such as fatty acid composition, transition metals and antioxidants, but is

also related to the existence of physical structures. Understanding how these physical

structures impact lipid oxidation and antioxidant mechanisms could lead to the

development of new antioxidant technologies and/or methods to use existing antioxidants

more effectively.

BIBLIOGRAPHY

1. Salas, J. J.; Sánchez, J.; Ramli, U. S.; Manaf, A. M.; Williams, M.; Harwood, J. L.,

Page 138: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

118

Biochemistry of lipid metabolism in olive and other oil fruits. Progress in Lipid Research 2000, 39, (2), 151-180.

2. Laguerre, M.; Lecomte, J.; Villeneuve, P., Evaluation of the ability of antioxidants to counteract lipid oxidation: Existing methods, new trends and challenges. Progress in Lipid Research 2007, 46, (5), 244-282.

3. Frankel, E. N., Lipid oxidation. Progress in Lipid Research 1980, 19, (1-2), 1-22.

4. Evans, M. V.; Turton, H. E.; Grant, C. M.; Dawes, I. W., Toxicity of linoleic acid hydroperoxide to Saccharomyces cerevisiae: involvement of a respiration-related process for maximal sensitivity and adaptive response. The Journal of Bacteriology 1998, 180, (3), 483-490.

5. Ebong, P. E.; Owu, D. U.; Isong, E. U., Influence of palm oil (Elaesis guineensis) on health. Plant Foods for Human Nutrition 1999, 53, (3), 209-222.

6. Zidkova, J.; Sajdok, J.; Kontrova, K.; Kotrbova-Kozak, A.; Hanis, T.; Zidek, V.; Fucikova, A., Effects of oxidised dietary cod liver oil on the reproductive functions of Wistar rat. Czechoslovak Journal of Food Sciences 2004, 22, (3), 108-120.

7. Ichinose, T.; Nobuyuki, S.; Takano, H.; Abe, M.; Sadakane, K.; Yanagisawa, R.; Ochi, H.; Fujioka, K.; Lee, K. G.; Shibamoto, T., Liver carcinogenesis and formation of 8-hydroxy-deoxyguanosine in C3H/HeN mice by oxidized dietary oils containing carcinogenic dicarbonyl compounds. Food and Chemical Toxicology 2004, 42, (11), 1795-1803.

8. Shahidi, F.; Zhong, Y., Revisiting the polar paradox theory: A critical overview. Journal of Agricultural and Food Chemistry 2011, dio:10.1021/jf104750m.

9. Branen, A. L., Toxicology and biochemistry of butylated hydroxyanisole and butylated hydroxytoluene. Journal of the American Oil Chemists Society 1975, 52, (2), 59-63.

10. Shahidi, F.; Zhong, Y., Novel antioxidants in food quality preservation and health promotion. European Journal of Lipid Science and Technology 2010, 112, (9), 930-940.

11. FDA, Electronic Code of Federal Regulations. In CFR 182.10. Spices and other natural seasonings and flavorings: 2011.

12. Decker, E., Oxidation in foods and beverages and antioxidant applications: understanding mechanisms of oxidation and antioxidant activity (volume 1). Woodhead Publishing: 2010.

13. McClements, D. J.; Decker, E. A., Lipid oxidation in oil-in-water emulsions: Impact of molecular environment on chemical reactions in heterogeneous food systems. Journal of Food Science 2000, 65, (8), 1270-1282.

Page 139: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

119

14. Chaiyasit, W.; Elias, R. J.; McClements, D. J.; Decker, E. A., Role of physical structures in bulk oils on lipid oxidation. Critical Reviews in Food Science and Nutrition 2007, 47, (3), 299-317.

15. Bhattacharyya, A.; Bhattacharyya, D., Deacidification of high FFA rice bran oil by reesterification and alkali neutralization. Journal of the American Oil Chemists Society 1987, 64, (1), 128-131.

16. Smiles, A.; Kakuda, Y.; MacDonald, B., Effect of degumming reagents on the recovery and nature of lecithins from crude canola, soybean and sunflower oils. Journal of the American Oil Chemists Society 1988, 65, (7), 1151-1155.

17. Ohlson, J. S. R., Processing effects on oil quality. Journal of the American Oil Chemists Society 1976, 53, (6), 299-301.

18. Lanser, A. C.; List, G. R.; Holloway, R. K.; Mounts, T. L., FTIR estimation of free fatty acid content in crude oils extracted from damaged soybeans. Journal of the American Oil Chemists Society 1991, 68, (6), 448-449.

19. de Alencar, E. R.; Faroni, L. R. D.; Peternelli, L. A.; da Silva, M. T. C.; Costa, A. R., Influence of soybean storage conditions on crude oil quality. Revista Brasileira de Engenharia Agrícola e Ambiental 2010, 14, (3), 303-308.

20. Gunstone, F. D.; Padley, F. B., Lipid technologies and applications. CRC: 1997.

21. Sarkadi, D. S., Hydrolysis during deodorization of fatty oils. Catalytic action of fatty acids. Journal of the American Oil Chemists Society 1959, 36, (4), 143-145.

22. O'Brien, R. D., Fats and oils: formulating and processing for applications. CRC: 2008.

23. Alimentarius, C., Codex standard for named vegetable oils. Codex Stan 1999, 210.

24. Maza, A.; Ormsbee, R. A.; Strecker, L. R., Effects of deodorization and steam-refining parameters on finished oil quality Journal of the American Oil Chemists Society 1992, 69, (10), 1003-1008.

25. Farhoosh, R.; Einafshar, S.; Sharayei, P., The effect of commercial refining steps on the rancidity measures of soybean and canola oils. Food Chemistry 2009, 115, (3), 933-938.

26. Ceriani, R.; Meirelles, A. J. A., Simulation of batch physical refining and deodorization processes. Journal of the American Oil Chemists Society 2004, 81, (3), 305-312.

27. Fritsch, C., Measurements of frying fat deterioration: A brief review. Journal of the American Oil Chemists Society 1981, 58, (3), 272-274.

Page 140: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

120

28. Goñi, F. M.; Alonso, A., Structure and functional properties of diacylglycerols in membranes. Progress in Lipid Research 1999, 38, (1), 1-48.

29. Coleman, R. A.; Lee, D. P., Enzymes of triacylglycerol synthesis and their regulation. Progress in Lipid Research 2004, 43, (2), 134-176.

30. Gee, P. T., Analytical characteristics of crude and refined palm oil and fractions. European Journal of Lipid Science and Technology 2007, 109, (4), 373-379.

31. Sleeter, R., Effects of processing on quality of soybean oil. Journal of the American Oil Chemists Society 1981, 58, (3), 239-247.

32. Ruiz-Méndez, M. V.; Márquez-Ruiz, G.; Dobarganes, M. C., Relationships between quality of crude and refined edible oils based on quantitation of minor glyceridic compounds. Food Chemistry 1997, 60, (4), 549-554.

33. Huang, A. H. C., Oil bodies and oleosins in seeds. Annual Review of Plant Biology 1992, 43, (1), 177-200.

34. Johnson, L. A., Recovery of fats and oils from plant and animal sources. Introduction to fats and oils. AOCS Press, Champaign 2000, 108-135.

35. Verleyen, T.; Sosinska, U.; Ioannidou, S.; Verh, R.; Dewettinck, K.; Huyghebaert, A.; De Greyt, W., Influence of the vegetable oil refining process on free and esterified sterols. Journal of the American Oil Chemists Society 2002, 79, (10), 947-953.

36. Koprivnjak, O.; Skevin, D.; Valic, S.; Majetic, V.; Petricevic, S.; Ljubenkov, I., The antioxidant capacity and oxidative stability of virgin olive oil enriched with phospholipids. Food Chemistry 2008, 111, (1), 121-126.

37. Gogolewski, M.; Nogala-Kalucka, M.; Szeliga, M., Changes of the tocopherol and fatty acid contents in rapeseed oil during refining. European Journal of Lipid Science and Technology 2000, 102, (10), 618-623.

38. Rossi, M.; Gianazza, M.; Alamprese, C.; Stanga, F., The effect of bleaching and physical refining on color and minor components of palm oil. Journal of the American Oil Chemists Society 2001, 78, (10), 1051-1055.

39. Kim, H.; Kim, S. G.; Choi, Y.; Jeong, H. S.; Lee, J., Changes in tocopherols, tocotrienols, and fatty acid contents in grape seed oils during oxidation. Journal of the American Oil Chemists Society 2008, 85, (5), 487-489.

40. Latif, S.; Anwar, F., Effect of aqueous enzymatic processes on sunflower oil quality. Journal of the American Oil Chemists Society 2009, 86, (4), 393-400.

41. Boon, C. S.; McClements, D. J.; Weiss, J.; Decker, E. A., Factors influencing the chemical stability of carotenoids in foods. Critical Reviews in Food Science and Nutrition 2010, 50, (6), 515-532.

Page 141: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

121

42. Jung, M. Y.; Yoon, S. H.; Min, D. B., Effects of processing steps on the contents of minor compounds and oxidation of soybean oil. Journal of the American Oil Chemists Society 1989, 66, (1), 118-120.

43. Usuki, R.; Suzuki, T.; Endo, Y.; Kaneda, T., Residual amounts of chlorophylls and pheophytins in refined edible oils. Journal of the American Oil Chemists Society 1984, 61, (4), 785-788.

44. Guillén, M. D.; Goicoechea, E.; Palencia, G.; Cosmes, N., Evidence of the formation of light polycyclic aromatic hydrocarbons during the oxidation of edible oils in closed containers at room temperature. Journal of Agricultural and Food Chemistry 2008, 56, (6), 2028-2033.

45. Kmiecik, D.; Korczak, J.; Rudzińska, M.; Michałowska, A. G.; Hęś, M., Stabilization of phytosterols in rapeseed oil by natural antioxidants during heating. European Journal of Lipid Science and Technology 2009, 111, (11), 1124-1132.

46. Moret, S.; Conte, L. S., Polycyclic aromatic hydrocarbons in edible fats and oils: occurrence and analytical methods. Journal of Chromatography A 2000, 882, (1-2), 245-253.

47. Sekeroglu, G.; Gogus, F.; Fadiloglu, S., Determination of benzo(a)pyrene in vegetable oils by high performance liquid chromatography. Journal of Food Quality 2007, 30, (3), 300-308.

48. Speer, K.; Steeg, E.; Horstmann, P.; Kuhn, T.; Montag, A., Determination and distribution of polycyclic aromatic-hydrocarbons in native vegetable-oils, smoked fish products, mussels and oysters, and bream from the river elbe. Journal of High Resolution Chromatography 1990, 13, (2), 104-111.

49. Guillén, M. D.; Goicoechea, E.; Palencia, G.; Cosmes, N., Evidence of the formation of light polycyclic aromatic hydrocarbons during the oxidation of edible oft in closed containers at room temperature. Journal of Agricultural and Food Chemistry 2008, 56, (6), 2028-2033.

50. Guillén, M. D.; Goicoechea, E., Formation of oxygenated alpha,beta-unsaturated aldehydes and other toxic compounds in sunflower oil oxidation at room temperature in closed receptacles. Food Chemistry 2008, 111, (1), 157-164.

51. Bondioli, P.; Mariani, C.; Lanzani, A.; Fedeli, E.; Muller, A., Squalene recovery from olive oil deodorizer distillates. Journal of the American Oil Chemists Society 1993, 70, (8), 763-766.

52. Moreda, W.; Perez-Camino, M. C.; Cert, A., Gas and liquid chromatography of hydrocarbons in edible vegetable oils. Journal of Chromatography A 2001, 936, (1-2), 159-171.

Page 142: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

122

53. Gomes, T.; Caponio, F.; Delcuratolo, D., Fate of oxidized triglycerides during refining of seed oils. Journal of Agricultural and Food Chemistry 2003, 51, (16), 4647-4651.

54. Ferrari, R. A.; Schulte, E.; Esteves, W.; Brül, L.; Mukherjee, K. D., Minor constituents of vegetable oils during industrial processing. Journal of the American Oil Chemists Society 1996, 73, (5), 587-592.

55. deGreyt, W. F.; Kellens, M. J.; Huyghebaert, A. D., Polymeric and oxidized triglyceride content of crude and refined vegetable oils - An overview. Fett-Lipid 1997, 99, (8), 287-290.

56. Gomes, T., Oligopolymer, diglyceride and oxidized triglyceride contents as measures of olive oil qualit. Journal of the American Oil Chemists Society 1992, 69, (12), 1219-1223.

57. de Greyt, W. F.; Kellens, M. J.; Huyghebaert, A. D., Polymeric and oxidized triglyceride content of crude and refined vegetable oils:an overview. Lipid/Fett 1997, 99, (8), 287-290.

58. Hopia, A., Analysis of high-molecular-weight autoxidation products using high-performance size-exclusion chromatography .2. changes during processing. Food Science and Technology-Lebensmittel-Wissenschaft & Technologie 1993, 26, (6), 568-571.

59. Bell, J. A., Chemistry: a project of the American Chemical Society. WH Freeman: 2004.

60. Vertucci, C. W.; Leopold, A. C., Bound water in soybean seed and its relation to respiration and imbibitional damage. Plant Physiology 1984, 75, (1), 114-117.

61. O'Brien, R. D., Fats and oils: formulating and processing for applications. CRC: 2004.

62. Garcia-Torres, R.; Ponagandla, N. R.; Rouseff, R. L.; Goodrich-Schneider, R. M.; Reyes-De-Corcuera, J. I., Effects of dissolved oxygen in fruit juices and methods of removal. Critical Reviews in Food Science and Nutrition 2009, 8, (4), 409-423.

63. Montgomery, H. A. C.; Thom, N. S.; Cockburn, A., Determination of dissolved oxygen by the Winkler method and the solubility of oxygen in pure water and sea water. Journal of Applied Chemistry 1964, 14, (7), 280-296.

64. Windrem, D. A.; Plachy, W. Z., The diffusion-solubility of oxygen in lipid bilayers. Biochimica et Biophysica Acta (BBA) - Biomembranes 1980, 600, (3), 655-665.

65. Schrader, U.; Becker, K.; Heiss, R., Der Einfluß von Diffusion und Löslichkeit auf die Reaktion von Sauerstoff mit kompakten Lebensmitteln. Mçchen: Universitç, Fachbereich Maschinenwesen, Diss 1979, 31, 33-40.

Page 143: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

123

66. Battino, R.; Rettich, T. R.; Tominaga, T., The solubility of oxygen and ozone in liquids. Journal of Physical and Chemical Reference Data 1983, 12, (2), 164-177.

67. Ke, P.; Ackman, R., Bunsen coefficient for oxygen in marine oils at various temperatures determined by an exponential dilution method with a polarographic oxygen electrode. Journal of the American Oil Chemists Society 1973, 50, (11), 429-435.

68. Schaich, K. M., Lipid oxidation: Theoretical aspects in Bailey’s Industrial Oil and Fat Products. John Wiley and Sons Inc.,: Hoboken, NJ, 2005.

69. Schafer, F. Q.; Qian, S. Y.; Buettner, G. R., Iron and free radical oxidations in cell membranes. Cell and Molecular Biology 2000, 46, (3), 657-662.

70. Frankel, E. N., Lipid oxidation. Oily Press Bridgewater: 2005.

71. Parenti, A.; Spugnoli, P.; Masella, P.; Calamai, L., Influence of the extraction process on dissolved oxygen in olive oil. European Journal of Lipid Science and Technology 2007, 109, (12), 1180-1185.

72. Masella, P.; Parenti, A.; Spugnoli, P.; Calamai, L., Nitrogen stripping to remove dissolved oxygen from extra virgin olive oil. European Journal of Lipid Science and Technology 2010, 112, (12), 1389-1392.

73. Maloba, F. W.; Rooney, M. L.; Wormell, P.; Nguyen, M., Improved oxidative stability of sunflower oil in the presence of an oxygen-scavenging film. Journal of the American Oil Chemists Society 1996, 73, (2), 181-185.

74. Tawfik, M. S.; Huyghebaert, A., Interaction of packaging materials and vegetable oils: oil stability. Food Chemistry 1999, 64, (4), 451-459.

75. Miyashita, K.; Takagi, T., Study on the oxidative rate and prooxidant activity of free fatty acids. Journal of the American Oil Chemists Society 1986, 63, (10), 1380-1384.

76. Yoshida, H., Influence of fatty acids of different unsaturation in the oxidation of purified vegetable oils during microwave irradiation. Journal of the Science of Food and Agriculture 1993, 62, 41-47.

77. Yoshida, H.; Kondo, I.; Kajimoto, G., Participation of free fatty acids in the oxidation of purified soybean oil during microwave heating. Journal of the American Oil Chemists Society 1992, 69, (11), 1136-1140.

78. Aubourg, S. P., Fluorescence study of the prooxidant effect of free fatty acids on marine lipids. Journal of the Science of Food and Agriculture 2001, 81, (4), 385-390.

79. Paradiso, V. M.; Gomes, T.; Nasti, R.; Caponio, F.; Summo, C., Effects of free fatty acids on the oxidative processes in purified olive oil. Food Research International 2010, 43, (5), 1389-1394.

Page 144: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

124

80. Mistry, B. S.; Min, D. B., Isolation of Sn-a-monolinolein from soybean oil and Its effect on oil oxidative stability. Journal of Food Science 1987, 52, (3), 786-790.

81. Mistry, B. S.; Min, D. B., Prooxidant effects of monoglycerides and diglycerides in soybean oil. Journal of Food Science 1988, 53, (6), 1896-1897.

82. Colakoglu, A. S., Oxidation kinetics of soybean oil in the presence of monoolein, stearic acid and iron. Food Chemistry 2007, 101, (2), 724-728.

83. Wang, T.; Jiang, Y.; Hammond, E., Effect of randomization on the oxidative stability of corn oil. Journal of the American Oil Chemists Society 2005, 82, (2), 111-117.

84. Gomes, T.; Caponio, F.; Bruno, G.; Summo, C.; Paradiso, V. M., Effects of monoacylglycerols on the oxidative stability of olive oil. Journal of the Science of Food and Agriculture 2010, 90, (13), 2228-2232.

85. Aubourg, S. P., Effect of partially hydrolysed lipids on inhibition of oxidation of marine lipids. European Food Research and Technology 2001, 212, (5), 540-545.

86. Husain, S. R.; Terao, J.; Matsushita, S., Effect of browning reaction products of phospholipids on autoxidation of methyl linoleate. Journal of the American Oil Chemists Society 1986, 63, (11), 1457-1460.

87. Takenaka, A.; Hosokawa, M.; Miyashita, K., Unsaturated phosphatidylethanolamine as effective synergist in combination with a-tocopherol. Journal of Oleo Science 2007, 56, (10), 511-516.

88. Bandarra, N.; Campos, R.; Batista, I.; Nunes, M.; Empis, J., Antioxidant synergy of α-tocopherol and phospholipids. Journal of the American Oil Chemists Society 1999, 76, (8), 905-913.

89. Hahnel, D.; Beyer, K.; Engelmann, B., Inhibition of peroxyl radical-mediated lipid oxidation by plasmalogen phospholipids and [alpha]-tocopherol. Free Radical Biology and Medicine 1999, 27, (9-10), 1087-1094.

90. Hildebrand, D. H.; Terao, J.; Kito, M., Phospholipids plus tocopherols increase soybean oil stability. Journal of the American Oil Chemists Society 1984, 61, (3), 552-555.

91. Sugino, H.; Ishikawa, M.; Nitoda, T.; Koketsu, M.; Juneja, L. R.; Kim, M.; Yamamoto, T., Antioxidative activity of egg yolk phospholipids. Journal of Agricultural and Food Chemistry 1997, 45, (3), 551-554.

92. Koga, T.; Terao, J., Phospholipids increase radical-scavenging activity of vitamin E in a bulk oil model system. Journal of Agricultural and Food Chemistry 1995, 43, (6), 1450-1454.

93. Hahnel, D.; Huber, T.; Kurze, V.; Beyer, K.; Engelmann, B., Contribution of

Page 145: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

125

copper binding to the inhibition of lipid oxidation by plasmalogen phospholipids. Biochemical Journal 1999, 340, (Pt 2), 377-383.

94. Khan, M. A.; Shahidi, F., Tocopherols and phospholipids enhance the oxidative stability of borage and evening primrose triacylglycerols. Journal of Food Lipids 2000, 7, (3), 143-150.

95. Koga, T.; Terao, J., Antioxidant activity of a novel phosphatidyl derivative of vitamin E in lard and its model system. Journal of Agricultural and Food Chemistry 1994, 42, (6), 1291-1294.

96. Seppanen, C. M.; Song, Q. H.; Csallany, A. S., The antioxidant functions of tocopherol and tocotrienol homologues in oils, fats, and food systems. Journal of the American Oil Chemists Society 2010, 87, (5), 469-481.

97. Kiokias, S.; Varzakas, T.; Oreopoulou, V., In vitro activity of vitamins, flavonoids, and natural phenolic antioxidants against the oxidative deterioration of oil-based systems. Critical Reviews in Food Science and Nutrition 2008, 48, (1), 78-93.

98. Yoshida, Y.; Niki, E.; Noguchi, N., Comparative study on the action of tocopherols and tocotrienols as antioxidant: chemical and physical effects. Chemistry and Physics of Lipids 2003, 123, (1), 63-75.

99. Yoshida, Y.; Tsuchiya, J.; Niki, E., Interaction of -tocopherol with copper and its effect on lipid peroxidation. Biochimica et Biophysica Acta (BBA) - General Subjects 1994, 1200, (2), 85-92.

100. Decker, A., Phenolics: prooxidants or antioxidants? Nutrition Reviews 1997, 55, (11), 396-398.

101. Kim, H. J.; Lee, H. O.; Min, D. B., Effects and prooxidant mechanisms of oxidized tocopherol on the oxidative stability of soybean oil. Journal of Food Science 2007, 72, (4), C223-C230.

102. Jung, M. Y.; Min, D. B., Effects of oxidized -,-and-tocopherols on the oxidative stability of purified soybean oil. Food Chemistry 1992, 45, (3), 183-187.

103. Choe, E.; Min, D. B., Mechanisms of antioxidants in the oxidation of foods. Critical Reviews in Food Science and Nutrition 2009, 8, (4), 345-358.

104. Lee, S. H.; Min, D. B., Effects, quenching mechanisms, and kinetics of carotenoids in chlorophyll-sensitized photooxidation of soybean oil. Journal of Agricultural and Food Chemistry 1990, 38, (8), 1630-1634.

105. Carlsson, D.; Suprunchuk, T.; Wiles, D., Photooxidation of unsaturated oils: effects of singlet oxygen quenchers. Journal of the American Oil Chemists Society 1976, 53, (10), 656-660.

Page 146: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

126

106. Endo, Y.; Usuki, R.; Kaneda, T., Antioxidant effects of chlorophyll and pheophytin on the autoxidation of oils in the dark. I. Comparison of the inhibitory effects. Journal of the American Oil Chemists Society 1985, 62, (9), 1375-1378.

107. Jung, M. Y.; Min, D. B., Effects of quenching mechanisms of carotenoids on the photosensitized oxidation of soybean oil. Journal of the American Oil Chemists Society 1991, 68, (9), 653-658.

108. Burton, G. W., Antioxidant action of carotenoids. J Nutr 1989, 119, (1), 109-111.

109. Burton, G. W.; Ingold, K. U., Beta-carotene: an unusual type of lipid antioxidant. Science 1984, 224, (4649), 569-573.

110. Steenson, D.; Min, D., Effects of β-carotene and lycopene thermal degradation products on the oxidative stability of soybean oil. Journal of the American Oil Chemists Society 2000, 77, (11), 1153-1160.

111. Psomiadou, E.; Tsimidou, M., On the role of squalene in olive oil stability. Journal of Agricultural and Food Chemistry 1999, 47, (10), 4025-4032.

112. Malecka, M., The effect of squalene on the heat stability of rapeseed oil and model lipids. Food/Nahrung 1991, 35, (5), 541-542.

113. Mateos, R.; Dominguez, M. M.; Espartero, J. L.; Cert, A., Antioxidant effect of phenolic compounds, alpha-tocopherol, and other minor components in virgin olive oil. Journal of Agricultural and Food Chemistry 2003, 51, (24), 7170-7175.

114. Mateos, R.; Domíguez, M. M.; Espartero, J. L.; Cert, A., Antioxidant effect of phenolic compounds, a-tocopherol, and other minor components in virgin olive oil. Journal of Agricultural and Food Chemistry 2003, 51, (24), 7170-7175.

115. Gordon, M. H.; Magos, P., The effect of sterols on the oxidation of edible oils. Food Chemistry 1983, 10, (2), 141-147.

116. Winkler, J. K.; Warner, K., The effect of phytosterol concentration on oxidative stability and thermal polymerization of heated oils. European Journal of Lipid Science and Technology 2008, 110, (5), 455-464.

117. Wang, T.; Hicks, K.; Moreau, R., Antioxidant activity of phytosterols, oryzanol, and other phytosterol conjugates. Journal of the American Oil Chemists Society 2002, 79, (12), 1201-1206.

118. Nyströ, L.; Achrenius, T.; Lampi, A. M.; Moreau, R. A.; Piironen, V., A comparison of the antioxidant properties of steryl ferulates with tocopherol at high temperatures. Food Chemistry 2007, 101, (3), 947-954.

119. Yoon, S. H.; Jung, M. Y.; Min, D. B., Effects of thermally oxidized triglycerides on the oxidative stability of soybean oil. Journal of the American Oil Chemists Society

Page 147: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

127

1988, 65, (10), 1652-1656.

120. Gomes, T.; Delcuratolo, D.; Paradiso, V. M., Pro-oxidant action of polar triglyceride oligopolymers in edible vegetable oils. European Food Research and Technology 2008, 226, (6), 1409-1414.

121. Xenakis, A.; Papadimitriou, V.; Sotiroudis, T. G., Colloidal structures in natural oils. Current Opinion in Colloid and Interface Science 2010, 15, (1-2), 55-60.

122. Lutton, E. S., Phase behavior of aqueous systems of monoglycerides. Journal of the American Oil Chemists Society 1965, 42, (12), 1068-1070.

123. Srisiri, W.; Benedicto, A.; O'Brien, D. F.; Trouard, T. P.; OrO'd, G.; Persson, S.; Lindblom, G., Stabilization of a bicontinuous cubic phase from polymerizable monoacylglycerol and diacylglycerol. Langmuir 1998, 14, (7), 1921-1926.

124. Monduzzi, M.; Ljusberg-Wahren, H.; Larsson, K., A 13C NMR study of aqueous dispersions of reversed lipid phases. Langmuir 2000, 16, (19), 7355-7358.

125. Chen, C. H.; Van Damme, I.; Terentjev, E. M., Phase behavior of C18 monoglyceride in hydrophobic solutions. soft matter 2008, 5, (2), 432-439.

126. Ichikawa, S.; Sugiura, S.; Nakajima, M.; Sano, Y.; Seki, M.; Furusaki, S., Formation of biocompatible reversed micellar systems using phospholipids. Biochemical Engineering Journal 2000, 6, (3), 193-199.

127. Abraham, S.; Narine, S. S., Self-Assembled nanostructures of oleic acid and their capacity for encapsulation and controlled delivery of nutrients. Journal of Nanoscience and Nanotechnology 2009, 9, (11), 6326-6334.

128. Chen, C. H.; Terentjev, E. M., Aging and metastability of monoglycerides in hydrophobic solutions. Langmuir 2009, 25, (12), 6717-6724.

129. Chen, C. H.; Terentjev, E. M., Effects of water on aggregation and stability of monoglycerides in hydrophobic solutions. Langmuir 2010, 26, (5), 3095-3105.

130. Kasaikina, O. T.; Kortenska, V. D.; Kartasheva, Z. S.; Kuznetsova, G. M.; Maximova, T. V.; Sirota, T. V.; Yanishlieva, N. V., Hydrocarbon and lipid oxidation in micro heterogeneous systems formed by surfactants or nanodispersed Al2O3, SiO2 and TiO2. Colloids and Surfaces A: Physicochemical and Engineering Aspects 1999, 149, (1-3), 29-38.

131. Subramanian, R.; Nakajima, M., Membrane degumming of crude soybean and rapeseed oils. Journal of the American Oil Chemists Society 1997, 74, (8), 971-975.

132. Subramanian, R.; Nakajima, M.; Kimura, T.; Maekawa, T., Membrane process for premium quality expeller-pressed vegetable oils. Food Research International 1998, 31, (8), 587-593.

Page 148: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

128

133. Manjula, S.; Subramanian, R., Membrane technology in degumming, dewaxing, deacidifying, and decolorizing edible oils. Critical Reviews in Food Science and Nutrition 2006, 46, (7), 569-592.

134. Bottino, A.; Capannelli, G.; Comite, A.; Ferrari, F.; Marotta, F.; Mattei, A.; Turchini, A., Application of membrane processes for the filtration of extra virgin olive oil. Journal of Food Engineering 2004, 65, (2), 303-309.

135. de Morais Coutinho, C.; Chiu, M. C.; Basso, R. C.; Ribeiro, A. P. B.; Gonçlves, L. A. G.; Viotto, L. A., State of art of the application of membrane technology to vegetable oils: A review. Food Research International 2009, 42, (5-6), 536-550.

136. Badan Ribeiro, A. P.; Bei, N.; Guaraldo Gonlves, L. A.; Cunha Petrus, J. C.; Viotto, L. A., The optimisation of soybean oil degumming on a pilot plant scale using a ceramic membrane. Journal of Food Engineering 2008, 87, (4), 514-521.

137. Waraho, T.; McClements, D. J.; Decker, E. A., Mechanisms of lipid oxidation in food dispersions. Trends in Food Science and Technology 2011, 21, (1), 3-13.

138. Laranjinha, J., Redox cycles of caffeic acid with -tocopherol and ascorbate. Methods in Enzymology 2001, 335, 282-295.

139. Chaiyasit, W.; Stanley, C.; Strey, H.; McClements, D.; Decker, E., Impact of surface active compounds on iron catalyzed oxidation of methyl linolenate in AOT–water–hexadecane systems. Food Biophysics 2007, 2, (2), 57-66.

140. Chaiyasit, W.; McClements, D. J.; Weiss, J.; Decker, E. A., Impact of surface-active compounds on physicochemical and oxidative properties of edible oil. Journal of Agricultural and Food Chemistry 2007, 56, (2), 550-556.

141. Trunova, N. A.; Kartasheva, Z. S.; Maksimova, T. V.; Bogdanova, Y. G.; Kasaikina, O. T., Decomposition of cumene hydroperoxide in the systems of normal and reverse micelles formed by cationic surfactants. Colloid Journal 2007, 69, (5), 655-659.

142. Kasaikina, O. T.; Golyavin, A. A.; Krugovov, D. A.; Kartasheva, Z. S.; Pisarenko, L. M., Micellar catalysis in the oxidation of lipids. Moscow University Chemistry Bulletin 2010, 65, (3), 206-209.

143. Kasaikina, O. T.; Kancheva, V. D.; Maximova, T. V.; Kartasheva, Z. S.; Yanishlieva, N. V.; Kondratovich, V. G.; Totseva, I. R., Catalytic effect of amphiphilic components on the lipid oxidation and lipid hydroperoxide decomposition. Oxidation Communication 2006, 29, (3), 574-584.

144. Kasaikina, O. T.; Kartasheva, Z. S.; Pisarenko, L. M., Effect of surfactants on liquid-phase oxidation of hydrocarbons and lipids. Russian Journal of General Chemistry 2008, 78, (8), 1533-1544.

Page 149: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

129

145. Kortenska, V. D.; Yanishlieva, N. V.; Kasaikina, O. T.; Totzeva, I. T.; Boneva, M. I.; Russina, I. F., Phenol antioxidant efficiency in various lipid substrates containing hydroxy compounds. European Journal of Lipid Science and Technology 2002, 104, (8), 513-519.

146. Alamed, J.; Chaiyasit, W.; McClements, D. J.; Decker, E. A., Relationships between free radical scavenging and antioxidant activity in foods. Journal of Agricultural and Food Chemistry 2009, 57, (7), 2969-2976.

147. Waraho, T.; Cardenia, V.; Rodriguez-Estrada, M. T.; McClements, D. J.; Decker, E. A., Prooxidant mechanisms of free fatty acids in stripped soybean Oil-in-Water emulsions. Journal of Agricultural and Food Chemistry 2009, 57, (15), 7112-7117.

148. Laity, P. R.; Taylor, J. E.; Wong, S. S.; Khunkamchoo, P.; Norris, K.; Cable, M.; Andrews, G. T.; Johnson, A. F.; Cameron, R. E., A review of small-angle scattering models for random segmented poly(ether-urethane) copolymers. Polymer 2004, 45, (21), 7273-7291.

149. Walker, L. M., Scattering from polymer-like micelles. Current Opinion in Colloid & Interface Science 2009, 14, (6), 451-454.

150. Albani, J. R.; Ebooks, C., Principles and applications of fluorescence spectroscopy. Wiley Online Library: 2007.

151. Goodson, T.; Varnavski, O.; Wang, Y., Optical properties and applications of dendrimer-metal nanocomposites. International Reviews in Physical Chemistry 2004, 23, (1), 109-150.

152. Rist, M. J.; Marino, J. P., Fluorescent nucleotide base analogs as probes of nucleic acid structure, dynamics and interactions. Current Organic Chemistry 2002, 6, (9), 775-793.

153. Chaiyasit, W.; Elias, R.; McClements, D. J.; Decker, E., Role of physical structures in bulk oils on lipid oxidation. Critical Reviews in Food Science and Nutrition 2007, 47, (3), 299-317.

154. Chaiyasit, W.; Stanley, C. B.; Strey, H. H.; McClements, D. J.; Decker, E. A., Impact of surface active compounds on iron catalyzed oxidation of methyl linolenate in AOT–Water–Hexadecane systems. Food Biophysics 2007, 2, (2), 57-66.

155. Frankel, E. N., Lipid oxidation. Oily Press Dundee: 1998.

156. Huang, S. W.; Hopia, A.; Schwarz, K.; Frankel, E. N.; German, J. B., Antioxidant Activity of a-Tocopherol and Trolox in Different Lipid Substrates: Bulk Oils vs Oil-in-Water Emulsions. Journal of Agricultural and Food Chemistry 1996, 44, (2), 444-452.

157. Boon, C. S.; Xu, Z.; Yue, X.; McClements, D. J.; Weiss, J.; Decker, E. A., Factors affecting lycopene oxidation in Oil-in-Water emulsions. Journal of Agricultural and

Page 150: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

130

Food Chemistry 2008, 56, (4), 1408-1414.

158. Wang, Y. H.; Mai, Q. Y.; Qin, X. L.; Yang, B.; Wang, Z. L.; Chen, H. T., Establishment of an evaluation model for Human milk fat substitutes. Journal of Agricultural and Food Chemistry 58, (1), 642-649.

159. Felgner, A.; Schlink, R.; Kirschenbuhler, P.; Faas, B.; Isengard, H. D., Automated Karl Fischer titration for liquid samples water determination in edible oils. Food Chemistry 2008, 106, (4), 1379-1384.

160. Kanamoto, R.; Wada, Y.; Miyajima, G.; Kito, M., Phospholipid-phospholipid interaction in soybean oil. Journal of the American Oil Chemists Society 1981, 58, (12), 1050-1053.

161. Semenyuk, A. V.; Svergun, D. I., GNOM-a program package for small-angle scattering data processing. Journal of Applied Crystallography 1991, 24, (5), 537-540.

162. Svergun, D. I., Mathematical methods in small-angle scattering data analysis. Journal of Applied Crystallography 1991, 24, (5), 485-492.

163. Svergun, D. I., Determination of the regularization parameter in indirect-transform methods using perceptual criteria. Journal of Applied Crystallography 1992, 25, (4), 495-503.

164. Chattopadhyay, A.; Mukherjee, S.; Raghuraman, H., Reverse micellar organization and dynamics: A wavelength-selective fluorescence approach. Journal of Physical Chemistry B 2002, 106, (50), 13002-13009.

165. Shantha, N. C.; Decker, E. A., Rapid, sensitive, iron-based spectrophotometric methods for determination of peroxide values of food lipids. Journal of AOAC International 1994, 77, (2), 421-424.

166. Tarver, T., Food Nanotechnology. Food Technology 2006, 60, (11), 22-26.

167. Papadimitriou, V.; Pispas, S.; Syriou, S.; Pournara, A.; Zournpanioti, M.; Sotiroudis, T. G.; Xenakis, A., Biocompatible microemulsions based on limonene: Formulation, structure, and applications. Langmuir 2008, 24, (7), 3380-3386.

168. Zhang, S., Fabrication of novel biomaterials through molecular self-assembly. Nature biotechnology 2003, 21, (10), 1171-1178.

169. King, M. D.; Marsh, D., Head group and chain length dependence of phospholipid self-assembly studied by spin-label electron spin resonance. Biochemistry 1987, 26, (5), 1224.

170. Srisiri, W.; Benedicto, A.; O'Brien, D. F.; Trouard, T. P.; Oradd, G.; Persson, S.; Lindblom, G., Stabilization of a bicontinuous cubic phase from polymerizable monoacylglycerol and diacylglycerol. Langmuir 1998, 14, (7), 1921-1926.

Page 151: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

131

171. Shiao, S. Y.; Chhabra, V.; Patist, A.; Free, M. L.; Huibers, P. D. T.; Gregory, A.; Patel, S.; Shah, D. O., Chain length compatibility effects in mixed surfactant systems for technological applications. Advances in Colloid and Interface Science 1998, 74, 1-29.

172. Ali, N.; Sul, W. H.; Lee, D. Y.; Kim, D. H.; Park, S. Y., Structures of the gylindrical and vesicular micelles of an P4VP-longer asymmetric PS-b-P4VP. Macromolecular Research 2009, 17, (8), 553-556.

173. Chaiyasit, W.; McClements, D. J.; Weiss, J.; Decker, E. A., Impact of surface-active compounds on physicochemical and oxidative properties of edible oil. Journal of Agricultural and Food Chemistry 2008, 56, (2), 550-556.

174. Eicke, H. F.; Christen, H., Is water critical to the formation of micelles in apolar media? Helvetica Chimica Acta 1978, 61, (6), 2258-2263.

175. Tung, S. H.; Huang, Y. E.; Raghavan, S. R., A new reverse wormlike micellar system: Mixtures of bile salt and lecithin in organic liquids. Journal of the American Chemical Society 2006, 128, (17), 5751-5756.

176. Langevin, D., Micelles and microemulsions. Annual Review of Physical Chemistry 1992, 43, (1), 341-369.

177. Ohshima, A.; Narita, H.; Kito, M., Phospholipid reverse micelles as a milieu of an enzyme reaction in an apolar system. Journal of biochemistry 1983, 93, (5), 1421.

178. Gochman-Hecht, H.; Bianco-Peled, H., Structure of AOT reverse micelles under shear. Journal of colloid and interface science 2005, 288, (1), 230-237.

179. Mezzasalma, S. A.; Koper, G. J. M.; Shchipunov, Y. A., Lecithin organogel as a binary blend of monodisperse polymer-like micelles. Langmuir 2000, 16, (26), 10564-10565.

180. Gupta, R.; Muralidhara, H. S.; Davis, H. T., Structure and phase behavior of phospholipid-based micelles in nonaqueous media. Langmuir 2001, 17, (17), 5176-5183.

181. Koga, T.; Terao, J., Phospholipids increase radical-scavenging activity of vitamin-e in a bulk oil model system. Journal of Agricultural and Food Chemistry 1995, 43, (6), 1450-1454.

182. Min, D. B.; Boff, J. M., Lipid oxidation of edible oil. Food lipid chemistry, nutrition, and biotechnology 2002, 2, 335-363.

183. Choe, E.; Min, D. B., Mechanisms and factors for edible oil oxidation. Critical Reviews in Food Science and Nutrition 2006, 5, (4), 169-186.

184. Decker, E. A.; Warner, K.; Richards, M. P.; Shahidi, F., Measuring antioxidant effectiveness in food. Journal of Agricultural and Food Chemistry 2005, 53, (10), 4303-4310.

Page 152: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

132

185. Chaiyasit, W.; Elias, R. J.; McClements, D. J.; Decker, E. A., Role of physical structures in bulk oils on lipid oxidation. Critical Reviews in Food Science and Nutrition 2007, 47, (3), 299-317.

186. Frankel, E. N., Antioxidants in lipid foods and their impact on food quality. Food Chemistry 1996, 57, (1), 51-55.

187. Hudson, B. J. F.; Ghavami, M., Phospholipids as antioxidant synergists for tocopherols in the autoxidation of edible oils. Lebensmittel-Wissenschaft & Technologie 1984, 17, (4), 191-194.

188. Chen, B.; Han, A.; McClements, D. J.; Decker, E. A., Physical Structures in Soybean Oil and Their Impact on Lipid Oxidation. Journal of Agricultural and Food Chemistry 2010, 58, (22), 11993–11999.

189. Nwosu, C. V.; Boyd, L. C.; Sheldon, B., Effect of fatty acid composition of phospholipids on their antioxidant properties and activity index. Journal of the American Oil Chemists Society 1997, 74, (3), 293-297.

190. Saito, H.; Ishihara, K., Antioxidant activity and active sites of phospholipids as antioxidants. Journal of the American Oil Chemists Society 1997, 74, (12), 1531-1536.

191. Bandarra, N. M.; Campos, R. M.; Batista, I.; Nunes, M. L.; Empis, J. M., Antioxidant synergy of alpha-tocopherol and phospholipids. Journal of the American Oil Chemists Society 1999, 76, (8), 905-913.

192. Takenaka, A.; Hosokawa, M.; Miyashita, K., Unsaturated phosphatidylethanolamine as effective synergist in combination with -tocopherol. Journal of Oleo Science 2007, 56, (10), 511-516.

193. Le Grandois, J.; Marchioni, E.; Zhao, M.; Giuffrida, F.; Ennahar, S.; Bindler, F., Oxidative Stability at High Temperatures of Oleyol and Linoleoyl Residues in the Forms of Phosphatidylcholines and Triacylglycerols. Journal of Agricultural and Food Chemistry 58, (5), 2973-2979.

194. Frankel, E. N.; Huang, S. W.; Kanner, J.; German, J. B., Interfacial phenomena in the evaluation of antioxidants: bulk oils vs emulsions. Journal of Agricultural and Food Chemistry 1994, 42, (5), 1054-1059.

195. Porter, W. L.; Black, E. D.; Drolet, A. M., Use of polyamide oxidative fluorescence test on lipid emulsions: contrast in relative effectiveness of antioxidants in bulk versus dispersed systems. Journal of Agricultural and Food Chemistry 1989, 37, (3), 615-624.

196. Mazaletskaya, L. I.; Sheludchenko, N. I.; Shishkina, L. N., Lecithin influence on the effectiveness of the antioxidant effect of flavonoids and α-tocopherol. Applied Microbiology and Microbiology 2010, 46, (2), 135-139.

Page 153: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

133

197. Ramadan, M. F., Quercetin increases antioxidant activity of soy lecithin in a triolein model system. LWT - Food Science and Technology 2008, 41, (4), 581-587.

198. Svenson, S., Controlling surfactant self-assembly. Current Opinion In Colloid & Interface Science 2004, 9, (3-4), 201-212.

199. Nagarajan, R., Solubilization in aqueous solutions of amphiphiles. Current Opinion In Colloid & Interface Science 1996, 1, (3), 391-401.

200. Fanun, M., Microstructure of Mixed Nonionic Surfactants Microemulsions Studied By SAXS and DLS. Journal Of Dispersion Science and Technology 2009, 30, (1), 115-123.

201. Chaiyasit, W.; McClements, D. J.; Decker, E. A., The relationship between the physicochemical properties of antioxidants and their ability to inhibit lipid oxidation in bulk oil and oil-in-water emulsions. Journal of Agricultural and Food Chemistry 2005, 53, (12), 4982-4988.

202. Evans, J. C.; Kodali, D. R.; Addis, P. B., Optimal tocopherol concentrations to inhibit soybean oil oxidation. Journal of the American Oil Chemists Society 2002, 79, (1), 47-51.

203. Shao, B.; Heinecke, J. W. HDL, lipid peroxidation, and atherosclerosis. Journal of Lipid Research 2009, 50, 599-601.

204. Waraho, T.; McClements, D. J.; Decker, E. A. Mechanisms of lipid oxidation in food dispersions. Trends in Food Science and Technology 2010, 22, 3-13.

205. Singh, M.; Nam, D. T.; Arseneault, M.; Ramassamy, C. Role of by-products of lipid oxidation in Alzheimer's disease brain: a focus on acrolein. Journal of Alzheimer's disease 2010, 21, 741-756.

206. Sun, Y. E.; Wang, W. D.; Chen, H. W.; Li, C. Autoxidation of unsaturated lipids in food emulsion. Critical Reviews in Food Science and Nutrition 2011, 51, 453-466.

207. Chen, B.; McClements, D. J.; Decker, E. A. Minor components in food oils: a critical review of their roles on lipid oxidation chemistry in bulk oils and emulsions. Critical Reviews in Food Science and Nutrition 2011, 51, 901-916.

208. Laguerre, M.; Chen, B.; Lecomte, j.; Villeneuve, P.; McClements, D. J.; Decker, E. A. Antioxidant properties of chlorogenic acid and its alkyl esters in stripped corn oil in combination with phospholipids and/or water. Journal of Agricultural and Food Chemistry 2011, 59, 10361-10366.

209. Chen, B.; Han, A.; Laguerre, M.; McClements, D. J.; Decker, E. A. Role of reverse micelles on lipid oxidation in bulk oils: impact of phospholipids on antioxidant activity of a-tocopherol and Trolox. Food and Function 2011, 2, 302-309.

Page 154: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

134

210. Benedet, J. A.; Shibamoto, T. Role of transition metals, Fe (II), Cr (II), Pb (II), and Cd (II) in lipid peroxidation. Food Chemistry 2008, 107, 165-168.

211. Laguerre, M.; Lecomte, J.; Villeneuve, P. Evaluation of the ability of antioxidants to counteract lipid oxidation: existing methods, new trends and challenges. Progress in Lipid Research 2007, 46, 244-282.

212. Shahidi, F.; Zhong, Y. Lipid oxidation and improving the oxidative stability. Chemcal Society Review 2010, 39, 4067-4079.

213. Keceli, T.; Gordon, M. H. Ferric ions reduce the antioxidant activity of the phenolic fraction of virgin olive oil. Journal of Food Science 2002, 67, 943-947.

214. Sorensen, A. D. M.; Haahr, A. M.; Becker, E. M.; Skibsted, L. H.; Bergenstahl, B.; Nilsson, L.; Jacobsen, C. Interactions between iron, phenolic compounds, emulsifiers, and pH in omega-3-enriched oil-in-water emulsions. Journal of Agricultural and Food Chemistry 2008, 56, 1740-1750.

215. Mei, L.; McClements, D. J.; Decker, E. A. Lipid oxidation in emulsions as affected by charge status of antioxidants and emulsion droplets. Journal of Agricultural and Food Chemistry 1999, 47, 2267-2273.

216. Decker, E. A.; Hultin, H. O. Nonenzymic catalysts of lipid oxidation in mackerel ordinary muscle. Journal of Food Science 1990, 55, 951-953.

217. Ouchi, A.; Ishikura, M.; Konishi, K.; Nagaoka, S.-i.; Mukai, K. Kinetic study of the prooxidant effect of a-Tocopherol. Hydrogen abstraction from lipids by a-Tocopheroxyl radical. Lipids 2009, 44, 935-943.

218. Terao, J.; Matsushita, S. The peroxidizing effect of a-tocopherol on autoxidation of methyl linoleate in bulk phase. Lipids 1986, 21, 255-260.

219. Thomas, F.; Serratrice, G.; Béguin, C.; Aman, E. S.; Pierre, J. L.; Fontecave, M.; Laulhère, J. P. Calcein as a fluorescent probe for ferric iron. Journal of Biological Chemistry 1999, 274, 13375-13382.

220. Boon, C. S.; McClements, D. J.; Weiss, J.; Decker, E. A. Role of iron and hydroperoxides in the degradation of lycopene in oil-in-water emulsions. Journal of Agricultural and Food Chemistry 2009, 57, 2993-2998.

221. Decker, E. A.; McClements, D. J. Transition metal and hydroperoxide interactions: an important determinant in the oxidative stability of lipid dispersions. Inform 2001, 12, 251-255.

222. Fuster, M. D.; Lampi, A. M.; Hopia, A.; Kamal-Eldin, A. Effects of a-and g-tocopherols on the autooxidation of purified sunflower triacylglycerols. Lipids 1998, 33, 715-722.

Page 155: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

135

223. Schaich, K. M. Metals and lipid oxidation. Contemporary issues. Lipids 1992, 27, 209-218.

224. Frankel, E. N.; Huang, S. W.; Aeschbach, R.; Prior, E. Antioxidant activity of a rosemary extract and its constituents, carnosic acid, carnosol, and rosmarinic acid, in bulk oil and oil-in-water emulsion. Journal of Agricultural and Food Chemistry 1996, 44, 131-135.

225. Suzer, T.; Coskun, E.; Demir, S.; Tahta, K. Lipid peroxidation and glutathione levels after cortical injection of ferric chloride in rats: effect of trimetazidine and deferoxamine. Research in Experimental Medicine 2000, 199, 223-229.

226. Min, B.; Cordray, J. C.; Ahn, D. U. Antioxidant effect of fractions from chicken breast and beef loin homogenates in phospholipid liposome systems. Food Chemistry 2011, 128, 299-307.

227. Dacaranhe, C.; Terao, J. A unique antioxidant activity of phosphatidylserine on iron-induced lipid peroxidation of phospholipid bilayers. Lipids 2001, 36, 1105-1110.

228. Ansari, R.; Kazi, T. G.; Jamali, M. K.; Arain, M. B.; Sherazi, S. T.; Jalbani, N.; Afridi, H. I. Improved extraction method for the determination of iron, copper, and nickel in new varieties of sunflower oil by atomic absorption spectroscopy. Journal of AOAC International 2008, 91, 400-407.

229. Nunes, L. S.; Barbosa, J. T. P.; Fernandes, A. P.; Lemos, V. A.; Santos, W. N. L. d.; Korn, M. G. A.; Teixeira, L. S. G. Multi-element determination of Cu, Fe, Ni and Zn content in vegetable oils samples by high-resolution continuum source atomic absorption spectrometry and microemulsion sample preparation. Food Chemistry 2011, 127, 780-783.

230. Mendil, D.; Uluözlü, ÖD,; Tüzen, M.; Soylak, M. Investigation of the levels of some element in edible oil samples produced in Turkey by atomic absorption spectrometry. Journal of Hazardous Material 2009, 165, 724-728.

231. Mei, L.; McClements, D. J.; Wu, J.; Decker, E. A. Iron-catalyzed lipid oxidation in emulsion as affected by surfactant, pH and NaCl. Food Chemistry 1998, 61, 307-312.

232. Mei, L.; Decker, E. A.; McClements, D. J. Evidence of iron association with emulsion droplets and its impact on lipid oxidation. Journal of Agricultural and Food Chemistry 1998, 46, 5072-5077.

233. Espósito, B. P.; Epsztejn, S.; Breuer, W.; Cabantchik, Z. I. A review of fluorescence methods for assessing labile iron in cells and biological fluids. Analitical Biochemistry 2002, 304, 1-18.

234. Hynes, M. J.; O'Coinceanainn, M. The kinetics and mechanisms of reactions of iron (III) with caffeic acid, chlorogenic acid, sinapic acid, ferulic acid and naringin.

Page 156: Minor Components and Their Roles on Lipid Oxidation in Bulk Oil That Contains Association Colloids

136

Journal of Inorganic Biochemistry 2004, 98, 1457-1464.

235. Deiana, S.; Gessa, C.; Marchetti, M.; Usai, M. Phenolic acid redox properties: pH influence on iron (III) reduction by caffeic acid. Soil Science Society of America Journal 1995, 59, 1301-1307.