268
EFFECT OF THERMAL PROCESSING ON THE PHENOLIC ANTIOXIDANTS OF COLORED POTATOES By BALUNKESWAR NAYAK A dissertation submitted in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY WASHINGTON STATE UNIVERSITY Department of Biological Systems Engineering MAY 2011

EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

EFFECT OF THERMAL PROCESSING ON THE PHENOLIC ANTIOXIDANTS OF

COLORED POTATOES

By

BALUNKESWAR NAYAK

A dissertation submitted in partial fulfillment of

the requirements for the degree of

DOCTOR OF PHILOSOPHY

WASHINGTON STATE UNIVERSITY

Department of Biological Systems Engineering

MAY 2011

Page 2: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

ii

To the Faculty of Washington State University:

The members of the Committee appointed to examine the dissertation of

BALUNKESWAR NAYAK find it satisfactory and recommend that it be accepted.

___________________________________

Juming Tang, Ph.D., Chair

___________________________________

Shyam S Sablani, Ph.D.

___________________________________

Jose De J Berrios, Ph.D.

___________________________________

Joseph R. Powers, Ph.D.

___________________________________

Jinwen Zhang, Ph.D.

Page 3: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

iii

ACKNOWLEDGMENTS

I express my greatest gratitude to Dr. Juming Tang for his advice, guidance and

encouragement throughout my study and research at Washington State University. I am at a loss

of words to express the importance and value of his helpful insights, understanding, and moral

support that enabled me to stay focused on my research. I also express hearty thanks to my

doctoral committee members: Drs. Shyam S Sablani, Jose De D Berrios, Joseph R. Powers, and

Jinwen Zhang for their valuable suggestions and feedback whenever I approached them with

problems.

I thank Dr. Rui Hai Liu, Associate Professor in the Department of Food Science, Cornell

University, Ithaca, NY for his guidance on experimental designs in microbiology during my stay

and research at Cornell University. I acknowledge the contributions of Dr. Sohan Birla for his

guidance in the data analysis and technical writing. I am indebted to Mr. Christopher Derito,

Cornell University for teaching me cellular antioxidant analysis.

I thank Galina Mikhaylenko for her valuable suggestions and support in conducting my

experiments smoothly in the lab. I gratefully acknowledge the assistance of James Pan and

Matthew Tom, USDA-ARS, Albany, CA for their assistance and feedback during the extrusion

experiment. I thank Frank Younce, Wayne Dewitt, and Vince Himsl for their technical support

and assistance during my instrumentation. I thank my friend Gopal and senior colleague Ram,

who were highly supportive and stood by me through the thick and thin of my doctoral study.

Last but not least, I want to express my heartfelt thanks to my parents, wife Pinky, brothers

Uma and Titu, sister Gayatri for their understanding, encouragement and patience during my

research.

Page 4: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

iv

EFFECT OF THERMAL PROCESSING ON THE PHENOLIC ANTIOXIDANTS OF

COLORED POTATOES

ABSTRACT

by Balunkeswar Nayak, Ph.D.

Washington State University

May 2011

Chair: Juming Tang

Foods with antioxidant capacity contribute health benefits and provide protection against

certain cancers, Alzheimer‘s dementia, and cardio-vascular diseases caused by oxidative

damage. Colored potatoes are a significant source of antioxidants from polyphenols, carotenoids,

and ascorbic acid. In this research, retention of total phenolics and antioxidant activity were

studied in fresh colored potatoes, and processed flakes were prepared as potential ingredients for

snack foods using freeze-drying, drum-drying and refractance window-drying. Extruded

products prepared from purple colored potatoes (‗Purple Majesty‘ cv) and yellow peas using a

twin-screw extruder were analyzed for the effect of extrusion cooking on their antioxidant

activity and anthocyanin level. Bioavailability of the phenolic antioxidants of extruded products

was analyzed using HepG2 liver cancer cells in a cellular antioxidant assay and compared with

unprocessed samples. Antioxidant potential of degradation products from purple potato

anthocyanins during high temperature processing and contribution to the total antioxidant

activity were measured using chemical assays (DPPH and ABTS). Thermal degradation kinetics

of anthocyanins prepared from ‗Purple Majesty‘ potatoes was conducted over a temperature

range of 100 – 150 °C. Dehydrated potato flakes showed no significant losses (P > 0.05) in total

antioxidant capacity (TAC) and total phenolics (TP) content, whereas 23 – 45% losses in total

Page 5: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

v

anthocyanins (TA) were observed during dehydration of potatoes. The quantity of TAC was

unchanged and the TP was largely retained (73 – 83%) in the extruded products prepared from

colored potatoes and yellow pea flours. Severe losses in TA (60 – 70%) of extruded products

were observed due to high temperature during extrusion cooking. The large bioavailability of the

extruded products was well supported by the ORAC antioxidant activity, total phenolics

(including free and bound fractions) and total flavonoids content irrespective of the degradation

of anthocyanins. The bioavailability in the extrudates could be attributed to the breakdown of the

conjugated phenolic antioxidants to their free forms, and formation of Maillard reaction products

with potential antioxidant activity. Thermal degradation in prepared anthocyanins followed a

first order reaction, but the degradation compounds had antioxidant potency contributing to the

TAC in the processed foods.

Page 6: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

vi

TABLE OF CONTENT

ACKNOWLEDGMENTS ............................................................................................................. iii

ABSTRACT ................................................................................................................................... iv

LIST OF TABLES ....................................................................................................................... xiii

LIST OF FIGURES ..................................................................................................................... xvi

CHAPTER ONE

INTRODUCTION .......................................................................................................................... 1

1. SUMMARY OF RELATED STUDIES AND PROBLEM STATEMENTS............................. 1

2. OBJECTIVES ............................................................................................................................. 4

3. DISSERTATION OUTLINES ................................................................................................... 5

REFERENCES ............................................................................................................................... 9

CHAPTER TWO

EFFECT OF PROCESSING ON THE PHENOLIC ANTIOXIDANTS OF FRUITS,

VEGETABLES AND GRAINS – A REVIEW ............................................................................ 14

1. INTRODUCTION .................................................................................................................... 14

1.1. Antioxidants and Mechanism ............................................................................................ 17

1.2. Measurement of antioxidant activity ................................................................................. 19

2. PHENOLIC ANTIOXIDANTS OF FRUITS, VEGETABLE AND GRAIN .......................... 22

2.1. Effect of processing on phytochemicals ............................................................................ 27

2.1.1. Extraction .................................................................................................................... 29

2.1.2. Blanching .................................................................................................................... 31

Page 7: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

vii

2.1.3. Cooking ....................................................................................................................... 33

2.1.4. Drying/Dehydration .................................................................................................... 37

2.1.5. Irradiation .................................................................................................................... 38

2.1.6. Extrusion ..................................................................................................................... 40

2.1.7. Non-thermal processing .............................................................................................. 41

2.1.8. Storage ........................................................................................................................ 43

2.1.9. Enzymatic/Chemical Oxidation .................................................................................. 47

3. ANTHOCYANINS ................................................................................................................... 49

3.1. Effect of processing on Anthocyanins ............................................................................... 56

3.1.1. Extraction .................................................................................................................... 56

3.1.2. Blanching .................................................................................................................... 57

3.1.3. Heating ........................................................................................................................ 59

3.1.4. Drying/Dehydration .................................................................................................... 61

3.1.5. CO2 treatment.............................................................................................................. 63

3.1.6. Addition of external ingredients ................................................................................. 63

3.1.7. Non-thermal processing .............................................................................................. 65

3.1.8. Storage ........................................................................................................................ 66

3.2. Degradation mechanism and products ............................................................................... 69

3.3. Degradation kinetics of anthocyanins ................................................................................ 72

4. SUMMARY .............................................................................................................................. 80

REFERENCES ............................................................................................................................. 82

Page 8: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

viii

CHAPTER THREE

COLORED POTATOES (SOLANUM TUBEROSUM L.) DRIED FOR ANTIOXIDANT-RICH

VALUE-ADDED FOODS.......................................................................................................... 116

1. INTRODUCTION .................................................................................................................. 116

2. MATERIALS AND METHODS ............................................................................................ 119

2.1. Raw Materials .................................................................................................................. 119

2.2. Production of potato flakes .............................................................................................. 119

2.2.1. Freeze-drying ............................................................................................................ 120

2.2.2. Drum-drying ............................................................................................................. 120

2.2.3. Refractance window-drying ..................................................................................... 120

2.3. Chemical Analysis ........................................................................................................... 121

2.3.1. Total antioxidant capacity ........................................................................................ 121

2.3.2. Total phenolics ......................................................................................................... 123

2.3.3. Total anthocyanins .................................................................................................... 123

2.4. Color determination ......................................................................................................... 125

2.5. Statistical analysis ............................................................................................................ 125

3. RESULTS AND DISCUSSION ............................................................................................. 126

3.1. Moisture content, total antioxidant capacity (TAC), total phenolics (TP) and total

anthocyanins (TA) in raw potatoes ......................................................................................... 126

3.2. Effects of blanching on TAC, TP and TA ....................................................................... 127

3.3 Total antioxidant capacity, total phenolics and total anthocyanins in dehydrated potato

flakes ....................................................................................................................................... 130

3.4. Correlation between TP and TAC.................................................................................... 132

Page 9: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

ix

3.5. Color in raw cultivars and dehydrated flakes .................................................................. 133

4. CONCLUSION ....................................................................................................................... 135

ACKNOWLEDGEMENTS ........................................................................................................ 136

REFERENCES ........................................................................................................................... 137

CHAPTER FOUR

EFFECT OF EXTRUSION ON THE ANTIOXIDANT CAPACITY AND COLOR

ATTRIBUTES OF EXPANDED EXTRUDATES PREPARED FROM PURPLE POTATO

AND YELLOW PEA FLOUR MIXES ...................................................................................... 142

1. INTRODUCTION .................................................................................................................. 143

2. MATERIALS AND METHODS ............................................................................................ 144

2.1. Materials .......................................................................................................................... 144

2.1.1. Production of potato flours ....................................................................................... 145

2.1.2. Sample preparation ................................................................................................... 145

2.1.3. Extrusion conditions ................................................................................................. 146

2.1.4. Experimental design.................................................................................................. 147

2.1.5. Determination of Expansion ratio ............................................................................. 147

2.1.6. Moisture content determination ................................................................................ 148

2.1.7. Pasting profile of potato flours ................................................................................. 148

2.1.8. Color evaluation ........................................................................................................ 148

2.2. Chemical Analyses........................................................................................................... 150

2.2.1. Total antioxidant capacity ......................................................................................... 150

2.2.2. Total phenolics ......................................................................................................... 151

Page 10: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

x

2.2.3. Total anthocyanins ................................................................................................... 152

2.2.4. Browning Index ....................................................................................................... 153

2.3. Statistical analysis ............................................................................................................ 153

3. RESULTS AND DISCUSSION ............................................................................................. 154

3.1. Expansion ratio ............................................................................................................... 154

3.2. Pasting behavior of potato flours ..................................................................................... 156

3.3. Color attributes................................................................................................................. 157

3.4. Total antioxidant capacity ................................................................................................ 159

3.5. Total phenolics ................................................................................................................. 162

3.6. Total anthocyanins ........................................................................................................... 164

3.7. Browning Index .............................................................................................................. 166

3.8. Correlation analyses ......................................................................................................... 167

4. CONCLUSIONS..................................................................................................................... 167

ACKNOWLEDGEMENTS ........................................................................................................ 168

REFERENCES ........................................................................................................................... 169

CHAPTER FIVE

BIOAVAILABILITY OF ANTIOXIDANTS IN EXTRUDED PRODUCTS PREPARED FROM

PURPLE POTATO AND DRY PEA FLOURS ......................................................................... 175

1. INTRODUCTION .................................................................................................................. 176

2. MATERIALS AND METHODS ............................................................................................ 178

2.1. Chemicals and Reagents .................................................................................................. 178

2.2. Preparation of Samples .................................................................................................... 178

Page 11: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xi

2.3 Extraction of Free Phenolic Compounds .......................................................................... 178

2.4. Extraction of Bound Phenolic Compounds...................................................................... 179

2.5. Determination of Total Phenolics .................................................................................... 179

2.6. Determination of Individual Phenolic Acids ................................................................... 180

2.7. Determination of Total Flavonoids .................................................................................. 181

2.8. Measurement of Antioxidant Activity (ORAC) .............................................................. 182

2.9. Cellular Antioxidant Activity assay ................................................................................. 182

2.10. Statistical Analyses ........................................................................................................ 184

3. RESULTS ............................................................................................................................... 184

3.1. Phenolic Contents ............................................................................................................ 184

3.3. Individual Phenolic Acids ................................................................................................ 187

3.4. Total Antioxidant Activity ............................................................................................... 189

3.5. Cellular Antioxidant Activity .......................................................................................... 192

3.6. Correlation Analyses ........................................................................................................ 191

4. DISCUSSION ......................................................................................................................... 193

ABBREVIATIONS USED ......................................................................................................... 201

ACKNOWLEDGEMENT .......................................................................................................... 201

REFERENCES ........................................................................................................................... 203

CHAPTER SIX

THERMAL DEGRADATION OF ANTHOCYANINS FROM PURPLE POTATOES

(‗PURPLE MAJESTY‘ CV) AND IMPACT ON ANTIOXIDANT CAPACITY. .................... 211

1. INTRODUCTION .................................................................................................................. 211

Page 12: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xii

2. MATERIALS AND METHODS ............................................................................................ 213

2.1 Chemicals .......................................................................................................................... 213

2.2. Materials .......................................................................................................................... 214

2.3. Extraction of anthocyanins .............................................................................................. 214

2.4. Purification of anthocyanins ............................................................................................ 215

2.4. Heat treatment of anthocyanins ....................................................................................... 215

2.5. HPLC Analyses ................................................................................................................ 217

2.7. Measurement of anthocyanins ......................................................................................... 218

2.8. Measurement of Antioxidant Capacity ............................................................................ 219

2.9. Determination of thermal kinetics parameters ................................................................. 220

3. RESULTS ............................................................................................................................... 222

3.1. Purification of anthocyanins ............................................................................................ 222

3.2. Degradation of anthocyanins ........................................................................................... 227

3.3. Antioxidant activity of the degradation compounds ........................................................ 229

4. DISCUSSION ......................................................................................................................... 230

5. CONCLUSIONS..................................................................................................................... 236

REFERENCES ........................................................................................................................... 239

CHAPTER SEVEN

CONCLUSIONS AND RECOMMENDATIONS ..................................................................... 246

1. CONCLUSIONS..................................................................................................................... 246

2. RECOMMENDATIONS ........................................................................................................ 248

APPENDIX ................................................................................................................................. 249

Page 13: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xiii

LIST OF TABLES

Table 2. 1. Major antioxidants in some common fruits, vegetables and grains. .......................... 21

Table 2. 2. In vitro assays used for measuring antioxidant activity of fruits and vegetables ...... 24

Table 2. 3. Comparison of different in vitro total antioxidant capacity assays. .......................... 25

Table 2. 4. Common anthocyanins present in fruits and vegetables. ........................................... 52

Table 2. 5. Degradation kinetic parameters of anthocyanins. ...................................................... 76

Table 3. 1. Total antioxidant capacity by DPPH assay, total phenolics content and total

anthocyanins of selected raw potato cultivars. ........................................................ 126

Table 3. 2. Total antioxidant capacity by DPPH assay, total phenolics and total anthocyanins of

raw, blanched and dried potato flakes from purple cultivars.. ................................. 129

Table 3. 3. Color atttributes of raw potato cultivars .................................................................. 133

Table 3. 4. Color atttributes of purple dehydrated flakes ......................................................... 135

Table 4. 1. Extrusion parameters for preparing extruded products using split yellow pea flours

and white potato flours. ........................................................................................... 146

Table 4. 2. Moisture contents and expansion ratios of the extrudates prepared from yellow pea

and purple potato flours ........................................................................................... 156

Table 4. 3. Flour pasting behaviors of white and purple potato flours. ..................................... 157

Table 4. 4. Color attributes of the extruded products prepared from yellow pea and purple potato

flours . ...................................................................................................................... 159

Page 14: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xiv

Table 5. 1. Percentage contributions of phytochemicals in free and bound extract of ingredients

(purple potato flour and dry pea flour), raw formulations and extruded products to

total phenolics, total antioxidant activity and total flavonoids ............................... 186

Table 5. 2. Quantity of free, bound and total individual phenolic acids present in extracts of

ingredients, raw formulations and extruded products ............................................. 188

Table 5. 3. Correlation analysis of phenolics, antioxidant activity (ORAC) and cellular

antioxidant activity of ingredients, raw formulations and extruded product. .......... 195

Table 6. 1. Experimental design of heat treatments with selected temperature and time

combinations. ........................................................................................................... 216

Table 6. 2. Concentrations of total anthocyanins from purple potato extract upon heating at 100

– 150 °C for 0 – 60 min. .......................................................................................... 226

Table 6. 3. Estimation of the order of anthocyanins degradation by examining r2 from plot of

zero-, half-, first and second order reactions. .......................................................... 227

Table 6. 4. Parameters for first-order kinetics and transition state equations for degradation of

anthocyanins from ‗Purple Majesty‘ potatoes after heat treatment over the

temperature range of 100 – 150 °C. ......................................................................... 227

Table 6. 5. Antioxidant Activity using DPPH radical scavenging assay of purified anthocyanins

from ‗Purple Majesty‘ potatoes upon heating at 100 – 150 °C for 0 – 60 min. ...... 234

Table 6. 6. Antioxidant Activity using ABTS radical scavenging assay of purified anthocyanins

from ‗Purple Majesty‘ potato upon heating at 100 – 150 °C for 0 – 60 min. .......... 235

Page 15: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xv

Table 6. 7. Progression of degradation compounds upon thermal exposure as measured by the

ratio of total antioxidant capacity and total anthocyanins in purified anthocyanins

samples over 100 – 150 °C. ..................................................................................... 237

Page 16: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xvi

LIST OF FIGURES

Figure 2. 1. Classification of Phytochemicals.. ........................................................................... 15

Figure 2. 2. Some of the plant phenols in food with antioxidant capacity. ................................. 16

Figure 2. 3. Generic structure of Flavonoid ................................................................................ 18

Figure 2. 4. Structures of common phenolic acids. ..................................................................... 18

Figure 2. 5. Classification of food antioxidants. ......................................................................... 20

Figure 2. 6. Mechanism of antioxidant activity involving free radicals. lipid radical (R·); peroxy

radical (ROO·); peroxide (ROOH); antioxidant (AH)............................................. 22

Figure 2. 7. Total phenols from daily consumption of fruits and vegetables in the American

diet............................................................................................................................ 26

Figure 2. 8. Heating/energy transfer medium in various processing methods. ........................... 28

Figure 2. 9. Changes in total antioxidant activity in vegetable matrix subjected to different

processing conditions.. ............................................................................................. 35

Figure 2. 10. Common anthocyanins found in flowers, fruits and vegetables. ............................. 50

Figure 2. 11. Spectral characteristics of potato anthocyanins, indication of glycosylation and

acylation patterns. .................................................................................................... 53

Figure 2. 12. Interrelationships between anthocyanin quantity and quality, and various factors

affecting the stability of anthocyanins.. ................................................................... 54

Figure 2. 13. Effect of pH on the possible degradation mechanism of anthocyanin (Malvidin-3-

glucoside).. ............................................................................................................... 55

Figure 2. 14. Enzymatic activity on the polyphenol.. .................................................................... 59

Page 17: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xvii

Figure 2. 15. Possible thermal degradation of non-acylated (Cyanidin-3-glucoside and

Pelargonidin-3-glucoside) and acylated (cyanidin glucosides with coumaric acid)

anthocyanins.. .......................................................................................................... 71

Figure 3. 1. Structure of anthocyanins. .................................................................................... 118

Figure 3. 2. Color of potato flakes; (A) without blanching; (B) with blanching ....................... 128

Figure 3. 3 Correlations between total antioxidant capacity by DPPH assay and total phenolics

in colored (purple, red, yellow) and white potato cultivars. .................................. 133

Figure. 4. 1. Effects of screw speed and feed moisture on the different expansion ratio of

extrudates prepared from white potato and yellow pea flours at 140 °C die

temperature ............................................................................................................ 155

Figure. 4. 2. RVA profile of white and purple potato flours. ..................................................... 158

Figure. 4. 3. Total antioxidant capacities by DPPH assay of raw formulations and extruded

products.. ................................................................................................................ 161

Figure. 4. 4. Total phenolic contents of raw formulations and extruded products.. ................... 163

Figure. 4. 5. Total anthocyanins contents and browning indices of raw formulations and extruded

products.. ................................................................................................................ 165

Figure 5. 1. Phenolic contents of ingredients, raw formulations and extruded products prepared

from raw formulations.. ......................................................................................... 185

Figure 5. 2. Flavonoid contents of ingredients, raw formulations and extruded products prepared

from raw formulations.. ......................................................................................... 187

Page 18: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xviii

Figure 5. 3 Antioxidant activity of ingredients, raw formulations and extruded products

prepared from raw formulations.. .......................................................................... 189

Figure 5. 4. Cellular antioxidant activity and EC50 value of ingredients, raw formulations and

extruded products prepared from raw formulations............................................... 191

Figure 6. 1. Specially designed thermal kinetics test (TKT) cells used for heat treatment of

purified anthocyanins (darker area) from ‗Purple Majesty‘ potatoes over a

temperature range of 100 -150 °C.. ........................................................................ 216

Figure 6. 2. Chromatogram of standards as detected using HPLC ........................................... 218

Figure 6. 3. HPLC-DAD profile of total polyphenols (in crude extract), phenolic acids (in ethyl

acetate portion) and purified anthocyanins (in HCl/Methanol portion) from ‗Purple

Majesty‘ potatoes. .................................................................................................. 223

Figure 6. 4 HPLC-DAD profile of control unheated purified anthocyanins from ‗Purple

Majesty‘ potatoes. .................................................................................................. 224

Figure 6. 5 MALDI mass spectra of the pigments from purified anthocyanins of ‗Purple

Majesty‘ potatoes;.. ................................................................................................ 225

Figure 6. 6. Chromatogram of thermal degradation compounds from purified anthocyanins

heated at 100 °C for 1 – 60 min... .......................................................................... 228

Figure 6. 7. First order plot for the degradation of anthocyanins during heating over a

temperature range of 100 -150 °C.. ........................................................................ 230

Figure 6. 8. Plot of ln (k) versus (1/T) for anthocyanin degradation during heating over the

temperature range of 100 -150 °C. ......................................................................... 231

Page 19: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

xix

Dedication

This dissertation is dedicated to my parents who provided

both emotional support and encouragement

Page 20: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

1

CHAPTER ONE

INTRODUCTION

1. SUMMARY OF RELATED STUDIES AND PROBLEM STATEMENTS

Phytochemicals derived from fruits and vegetables act effectively in preventing formation of

reactive oxygen, nitrogen, hydroxyl and lipid species either by scavenging free radicals,

repairing or removing damaged molecules (Velioglu et al., 1998). Consumption of antioxidant

rich foods maintains higher antioxidant levels in blood serum (Cao, Booth, Sadowski, & Prior,

1998; Cao, Russell, Lischner, & Prior, 1998). Phenolic antioxidants prevent many of the chronic

diseases associated with cancer, inflammation, atherosclerosis and ageing (Prior et al., 1998;

Zern et al., 2005). The relationship between consumption of fruits and vegetables to occurrence

of lung , colon, breast, cervical, esophageal, oral cavity, stomach, bladder, pancreatic and ovarian

cancers has been reviewed for almost 200 epidemiological studies (Block et al., 1992). The

same investigator reported that consumption of fruit and vegetable has significant protective

effect in reducing certain forms of cancer as demonstrated in 128 – 156 dietary studies.

Potatoes (Solanum tuberosum L.) have traditionally been perceived by consumers as a good

source of energy with additional benefits as they contain low fat and are cholesterol free.

Potatoes are also a good source of potassium, providing 21% of the recommended daily value.

Low sodium content in the diet reduces the risk of high blood pressure and heart attack for many

individuals. Potatoes provide 12% of the recommended daily value of dietary fiber beneficial for

a healthy digestive system and are an excellent source of vitamin C (45% of daily value), vitamin

B6 and iron (USDA National Nutrient Database, 2010). Colored potatoes, in particular, have

attracted the attention of investigators as well as consumers due to their taste, appearance, and

Page 21: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

2

high level of antioxidant activities. Antioxidant activities in colored potatoes are associated with

the presence of polyphenols-/- flavonoids, carotenoids, ascorbic acid, tocopherols, alpha-lipoic

acid and selenium (Lachman et al., 2000).

Food processing operations such as drying, cooking and extrusion are necessary steps in

production of convenience foods, but they may affect the retention of antioxidants in food

matrices (Clifford, 2000; Kader et al., 2002; Nicoli et al., 1999; Rossi et al., 2003; Wu et al.,

2004a; Wu et al., 2004b). For example, blanching and drying are two unit operations in

preparing shelf-stable potato flakes as ingredients for commercial production of a wide range of

foods products, including mashed potato and extruded snack foods. Bruising/wounding, boiling,

baking, freeze-drying and microwave cooking of white and colored potatoes have different

effects on the total phenolic content (Brown et al., 2003; Reyes and Cisneros-Zevallos, 2003). It

was reported that potato peels and patatin protein hydrolysate suppressed oxidation of beef

patties (Wang and Xiong, 2005) and peel waste retarded oxidation in radiation processed lamb

meat (Kanatt et al., 2005). Most of time, investigators assume vitamin C as an indicator for

processing effects (Burg and Fraile, 1995; Lathrop and Leung, 1980). However, Eberhardt, Lee,

& Liu (2000) observed that vitamin C contributed less than 0.4% to the total antioxidant activity

in apples indicating the importance of phytochemicals towards antioxidant activity. Other than

vitamin C, there is limited literature available on the effects of drying/thermal treatments on

phytochemicals of potatoes. Bioactive phytochemicals exist in free as well as soluble-conjugated

and bound forms (Adom and Liu, 2002). Bound phytochemicals, mostly in cell wall materials,

are difficult to digest in the upper gastrointestine and may digest in the colon which reduces risk

of colon cancer (Andreasen et al., 2001).

Page 22: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

3

According to Tiziani, Schwartz, & Vodovotz (2008), metabolism, absorption and

bioavailability of health-beneficial phytochemicals could be improved by combining the effects

of protein and individual phytochemicals. Liu (2004) reported the additive and synergistic

effects of biologically active compounds from fruits, vegetables and grains on health. Studies

have demonstrated that it is possible to puff pulse flours and pulse-based formulations into

potentially commercial nutritious snack and breakfast cereal-type products (Berrios et al., 2002;

Berrios et al., 2010; Berrios et al., 2004; Patil et al., 2007). Pulses (such as yellow peas) have

high levels of protein, dietary fibers, complex carbohydrates and folate, and are low in fat and

sodium (Madar and Stark, 2002). Therefore, it would be desirable to make commodities such as

colored potatoes and yellow peas into functional foods in the form of extruded snacks and

breakfast cereal-type food products. Understanding the profile and bioavailability of

phytochemicals in the unprocessed (raw formulations) and processed (extruded products) foods

is important for optimal use of these antioxidant compounds. Kinetic studies of natural color

pigments and analyses of degradation compounds will help design optimal processes to retain

maximal antioxidant activity in processed food.

Mishra, Dolan, & Yang (2008) studied the degradation of anthocyanins above 100°C and

hypothesized that anthocyanins could be used as food colorants in high temperature processes

such as for extruded snacks or baked cakes. Similarly, numerous studies have reported on the

degradation of anthocyanins in fruits and vegetables during processing and storage (Rossi et al.,

2003; Sadilova et al., 2006). The investigators either assumed or reported first-order kinetics for

anthocyanin degradation in selected fruits and vegetables (Garzon and Wrolstad, 2001; Kirca et

al., 2007; Reyes and Cisneros-Zevallos, 2007; Yue and Xu, 2008). However, those studies were

carried out either in whole pulp puree/extract. Puree/extract from whole pulp of fruits and

Page 23: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

4

vegetables or their powders (hence used as unpurified) contain anthocyanins with other

compounds such as salts, sugars and other colorless non-anthocyanin phenolics that could affect

the stability or degradation kinetics and antioxidant capacity of anthocyanins (Del Pozo-Insfran

et al., 2004). Higher antioxidant activity of anthocyanin degradation compounds than the

unheated anthocyanin was also reported by many investigators (Matsufuji et al., 2007; Sadilova

et al., 2007; Yue and Xu, 2008). However, limited investigations have reported on the effect of

processing on the phytochemicals of colored potatoes, on the bioavailability of products prepared

from potatoes and legumes using extrusion cooking technology, and on the thermal kinetics of

purified anthocyanins and antioxidant potencies of degradation compounds from the purple

potato.

2. OBJECTIVES

The overall objective of this dissertation is to discern a fundamental understanding of thermal

degradation of the phytochemicals present in colored potatoes and value added products prepared

from it. The specific aims were to:

Quantify the total antioxidant capacity, total phenolics, and total anthocyanins in selected

white and colored potato cultivars and to study the effect of blanching and consequent

drying on the retention of these quality parameters in purple potato cultivars;

Produce puffed extrudates from mixtures of colored potato and yellow pea flours, and

evaluate the effect of extrusion conditions on antioxidant capacities, color attributes, and

some physical characteristics of the extrudates;

Investigate the complete phytochemicals that exists in free and bound forms as well as

their contribution to bioactivity in the raw ingredients (purple potato flour and dry pea

Page 24: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

5

flour), raw formulations and processed products prepared from the above ingredients

using extrusion cooking; and

Evaluate the thermal degradation kinetics parameters of anthocyanins purified from

purple potatoes (‗Purple Majesty‘ cv) over the temperature range of 100 – 150 °C and

determine antioxidant potencies of degradation compounds from the anthocyanins.

3. DISSERTATION OUTLINES

This dissertation contains seven chapters to address the above objectives.

Chapter One: Introduction. This chapter summarizes related research conducted by various

investigators, problem statements of current research and proposed objectives to address the

problems in the study. It also provides an outline/structure of the dissertation.

Chapter Two: Effect of processing on the phenolic antioxidants of fruits, vegetables and

grains- a review. This chapter compiles and briefly discusses various phenolic antioxidants

present in fruits, vegetables and grains. Previous research results and findings on the effect of

various processing methods including extraction, dehydration/drying, cooking and storing on the

phenolic antioxidants are summarized.

Chapter Three: Colored potatoes (Solanum Tuberosum L.) dried for antioxidant-rich value-

added products. This chapter contains information on the selection of colored potatoes based on

the retention of antioxidant compounds in raw potatoes and dehydrated potato flakes prepared as

potential ingredients for snack foods. Purple, red, yellow and white potatoes are dehydrated

using drum-drying, freeze-drying and refractance window-drying to prepare flakes. Total

antioxidant capacity, phenolics and anthocyanins present in raw and dehydrated flakes are

measured using chemical assays. Color attributes of the raw and dehydrated flakes are also

measured to correlate with the presence of anthocyanins. Contributions of the total phenolics and

Page 25: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

6

anthocyanins to total antioxidants in potatoes are analyzed using correlation among the

parameters.

Chapter Four: Effect of extrusion on the antioxidant capacity and color attributes of

expanded extrudates prepared from purple potato and yellow pea flour mixes. In this chapter, the

total antioxidant capacity, total phenolics, total anthocyanins, and color attributes of unprocessed

formulations and processed products using a twin-screw extruder are evaluated. The effect of

extrusion cooking in producing expanded extrudates as well as retaining phenolic antioxidants is

also examined. Color attributes such as brightness, chroma, hue and browning index in extruded

products are evaluated for potential retention of natural color due to anthocyanins, as affected by

high temperature short time extrusion cooking.

Chapter Five: Bioavailability of antioxidants in extruded products prepared from purple

potato and dry pea flours. Bioavailability of phytochemicals in vivo in the extruded products

prepared from purple potatoes (‗Purple Majesty‘ cv) and dry peas are evaluated using a cellular

antioxidant assay using liver cancer HepG2 cells. Total phenolic antioxidants including that

present in free and bound fractions are considered for determining the antioxidant activity of the

processed products. Role of phenolic antioxidants evaluated using a cellular method is compared

with the commonly used chemical assay (oxygen radical absorption capacity) for a possible

correlation and better explanation of their health benefits.

Chapter Six: Thermal degradation of anthocyanins from purple potato (‗Purple Majesty‘ cv)

and impact on antioxidant activities. This chapter provides some explanations for possible

increase in the antioxidant activities of extruded products irrespective of reduction in its natural

color/anthocyanins. Anthocyanins are prepared from purple potatoes (‗Purple Majesty‘ cv) by

removing salts, sugars and colorless phenolics and heat treated (100 – 150 °C) and analyzed for

Page 26: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

7

potential antioxidant activity of the degradation compounds using chemical antioxidant assays.

Thermal degradation kinetics parameters including reaction order, activation energy, thermal

death time (D-values) and z-value of the purified anthocyanins are reported.

Chapter Seven: Conclusions and recommendations. Summary of the findings of the present

study and recommendations for future work are illustrated in this chapter.

Some chapters in the dissertation carry the styles of a particular journal, where it is published

or submitted. Full citations of these chapters included in this dissertation are as follows:

Chapter Two

Nayak, Balunkeswar, Berrios, J., Powers, J.R. and Tang, J. Effect of processing on phenolic

antioxidants of fruits and vegetables. Trends in Food Science and Technology (internal

review)

Chapter Three

Nayak, Balunkeswar, Powers, Joseph R., Berrios, J. and Tang, Juming. Potential of colored

potatoes for producing high antioxidant snack foods. Journal of Food Processing and

Preservation (in press).

Chapter Four

Nayak, Balunkeswar, Berrios, J., Powers, J.R., and Tang, J. Effect of extrusion on the

antioxidant capacity and color attributes of expanded products prepared from purple

potatoes and yellow peas. Food Chemistry (internal review)

Chapter Five

Nayak, Balunkeswar, Liu, R.H., Berrios, J., Tang, J., Derito, C. Bioavailability of

antioxidants in extruded products prepared from purple potato and dry pea flours. Journal

of Agricultural & Food Chemistry (internal review)

Page 27: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

8

Chapter Six

Nayak, Balunkeswar, Berrios, J., and Tang, J. Thermal degradation of anthocyanins in purple

potato and impact on antioxidant capacity. Journal of Food Engineering (internal review)

Page 28: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

9

REFERENCES

Adom, K. K., and Liu, R. H. (2002). Antioxidant activity of grains. Journal of Agricultural and

Food Chemistry. 50: 6182-6187.

Andreasen, M. F., Kroon, P. A., Williamson, G., and Garcia-Conesa, M. T. (2001). Intestinal

release and uptake of phenolic antioxidant diferulic acids. Free Radical Biology and

Medicine. 31: 304-314.

Berrios, J. D., Camara, M., Torija, M. E., and Alonso, M. (2002). Effect of extrusion cooking

and sodium bicarbonate addition on the carbohydrate composition of black bean flours.

Journal of Food Processing and Preservation. 26: 113-128.

Berrios, J. D., Morales, P., Camara, M., and Sanchez-Mata, M. C. (2010). Carbohydrate

composition of raw and extruded pulse flours. Food Research International. 43: 531-536.

Berrios, J. D., Wood, D. F., Whitehand, L., and Pan, J. (2004). Sodium bicarbonate and the

microstructure, expansion and color of extruded black beans. Journal of Food Processing

and Preservation. 28: 321-335.

Block, G., Patterson, B., and Subar, A. (1992). Fruit, vegetables, and cancer prevention - a

review of the epidemiologic evidence. Nutrition and Cancer-an International Journal. 18: 1-

29.

Brown, C. R., Wrolstad, R., Durst, R., Yang, C. P., and Clevidence, B. (2003). Breeding studies

in potatoes containing high concentrations of anthocyanins. American Journal of Potato

Research. 80: 241-249.

Burg, P., and Fraile, P. (1995). Vitamin C destruction during the cooking of a potato dish.

Lebensmittel-Wissenschaft und-Technologie. 28: 506-514.

Page 29: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

10

Clifford, M. N. (2000). Anthocyanins – nature, occurrence and dietary burden. Journal of the

Science of Food and Agriculture. 80: 1063-1072.

Del Pozo-Insfran, D., Brenes, C. H., and Talcottt, S. T. (2004). Phytochemical composition and

pigment stability of acai (Euterpe oleracea Mart.). Journal of Agricultural and Food

Chemistry. 52: 1539-1545.

Eberhardt, M. V., Lee, C. Y., and Liu, R. H. (2000). Nutrition - Antioxidant activity of fresh

apples. Nature. 405: 903-904.

Garzon, G. A., and Wrolstad, R. E. (2001). The stability of pelargonidin-based anthocyanins at

varying water activity. Food Chemistry. 75: 185-196.

Kader, F., Irmouli, M., Nicolas, J. P., and Metche, M. (2002). Involvement of blueberry

peroxidase in the mechanisms of anthocyanin degradation in blueberry juice. Journal of

Food Science. 67: 910-915.

Kanatt, S. R., Chander, R., Radhakrishna, P., and Sharma, A. (2005). Potato peel extract - a

natural antioxidant for retarding lipid peroxidation in radiation processed lamb meat. Journal

of Agricultural and Food Chemistry. 53: 1499-1504.

Kirca, A., Ozkan, M., and Cemeroglu, B. (2007). Effects of temperature, solid content and pH on

the stability of black carrot anthocyanins. Food Chemistry. 101: 212-218.

Lachman, J., Hamouz, K., Orsak, M., and Pivec, V. (2000). Potato tubers as a significant source

of antioxidants in human nutrition. Rostlinna Vyroba. 46: 231-236.

Lathrop, P. J., and Leung, H. K. (1980). Rates of ascorbic acid degradation during thermal

processing of canned peas. Journal of Food Science. 45: 152-153.

Liu, R. H. (2004). Potential synergy of phytochemicals in cancer prevention: mechanism of

action. J. Nutr. 134: 3479S-3485.

Page 30: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

11

Madar, Z., and Stark, A. H. (2002). New legume sources as therapeutic agents. British Journal of

Nutrition. 88: S287-S292.

Matsufuji, H., Kido, H., Misawa, H., Yaguchi, J., Otsuki, T., Chino, M., Takeda, M., and

Yamagata, K. (2007). Stability to light, heat, and hydrogen peroxide at different pH values

and DPPH radical scavenging activity of acylated anthocyanins from red radish extract.

Journal of Agricultural and Food Chemistry. 55: 3692-3701.

Mishra, D. K., Dolan, K. D., and Yang, L. (2008). Confidence intervals for modeling

anthocyanin retention in grape pomace during non-isothermal heating. Journal of Food

Science. 73: E9-E15.

Nicoli, M. C., Anese, M., and Parpinel, M. (1999). Influence of processing on the antioxidant

properties of fruit and vegetables. Trends in Food Science & Technology. 10: 94-100.

Patil, R. T., Berrios, J. D. J., Tang, J., and Swanson, B. G. (2007). Evaluation of methods for

expansion properties of legume extrudates. Applied Engineering in Agriculture. 23: 777-783.

Prior, R. L., Cao, G. H., Martin, A., Sofic, E., McEwen, J., O'Brien, C., Lischner, N., Ehlenfeldt,

M., Kalt, W., Krewer, G., and Mainland, C. M. (1998). Antioxidant capacity as influenced by

total phenolic and anthocyanin content, maturity, and variety of Vaccinium species. Journal

of Agricultural and Food Chemistry. 46: 2686-2693.

Reyes, L. F., and Cisneros-Zevallos, L. (2003). Wounding stress increases the phenolic content

and antioxidant capacity of purple-flesh potatoes (Solanum tuberosum L.). Journal of

Agricultural and Food Chemistry. 51: 5296-5300.

Reyes, L. F., and Cisneros-Zevallos, L. (2007). Degradation kinetics and colour of anthocyanins

in aqueous extracts of purple- and red-flesh potatoes (Solanum tuberosum L.). Food

Chemistry. 100: 885-894.

Page 31: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

12

Rossi, M., Giussani, E., Morelli, R., Lo Scalzo, R., Nani, R. C., and Torreggiani, D. (2003).

Effect of fruit blanching on phenolics and radical scavenging activity of highbush blueberry

juice. Food Research International. 36: 999-1005.

Sadilova, E., Carle, R., and Stintzing, F. C. (2007). Thermal degradation of anthocyanins and its

impact on color and in vitro antioxidant capacity. Molecular Nutrition & Food Research. 51:

1461-1471.

Sadilova, E., Stintzing, F. C., and Carle, R. (2006). Thermal degradation of acylated and

nonacylated anthocyanins. Journal of Food Science. 71: C504-C512.

Tiziani, S., Schwartz, S. J., and Vodovotz, Y. (2008). Intermolecular interactions in

phytochemical model systems studied by NMR diffusion measurements. Food Chemistry.

107: 962-969.

Velioglu, Y. S., Mazza, G., Gao, L., and Oomah, B. D. (1998). Antioxidant activity and total

phenolics in selected fruits, vegetables, and grain products. Journal of Agricultural and Food

Chemistry. 46: 4113-4117.

Wang, L. L., and Xiong, Y. L. L. (2005). Inhibition of lipid oxidation in cooked beef patties by

hydrolyzed potato protein is related to its reducing and radical scavenging ability. Journal of

Agricultural and Food Chemistry. 53: 9186-9192.

Wu, X., Beecher, G. R., Holden, J. M., Haytowitz, D. B., Gebhardt, S. E., and Prior, R. L.

(2004a). Lipophilic and hydrophilic antioxidant capacities of common foods in the United

States. Journal of Agricultural and Food Chemistry. 52: 4026-4037.

Wu, X. L., Gu, L. W., Holden, J., Haytowitz, D. B., Gebhardt, S. E., Beecher, G., and Prior, R.

L. (2004b). Development of a database for total antioxidant capacity in foods: a preliminary

study. Journal of Food Composition and Analysis. 17: 407-422.

Page 32: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

13

Yue, X., and Xu, Z. (2008). Changes of anthocyanins, anthocyanidins, and antioxidant activity in

bilberry extract during dry heating. Journal of Food Science. 73: C494-C499.

Zern, T. L., Wood, R. J., Greene, C., West, K. L., Liu, Y. Z., Aggarwal, D., Shachter, N. S., and

Fernandez, M. L. (2005). Grape polyphenols exert a cardioprotective effect in pre- and

postmenopausal women by lowering plasma lipids and reducing oxidative stress. Journal of

Nutrition. 135: 1911-1917.

http://www.nal.usda.gov/fnic/foodcomp/cgi-bin/list_nut_edit.pl. USDA National Nutrient

Database for Standard Reference, Release 23 (2010).

Page 33: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

14

CHAPTER TWO

EFFECT OF PROCESSING ON THE PHENOLIC ANTIOXIDANTS OF

FRUITS, VEGETABLES AND GRAINS – A REVIEW

1. INTRODUCTION

Phytochemicals are bioactive non-nutrient plant compounds in fruits, vegetables, grains and

other plant foods thought to promote health. These are broadly classified as carotenoids,

phenolics, alkaloids, nitrogen-containing, and organosulfur compounds (Figure 2.1).

Phytochemicals prevent many of the chronic diseases associated with cancer, inflammation,

atherosclerosis and ageing caused by free radicals and oxygen (Prior et al., 1998; van den Berg,

Haenen, van den Berg, & Bast, 1999; Zern et al., 2005). Consumption of antioxidant rich foods

maintains higher antioxidant levels in blood serum (Cao, Booth, Sadowski, & Prior, 1998; Cao,

Russell, Lischner, & Prior, 1998). The relationship between consumption of fruits and vegetables

to lung, colon, breast, cervical, esophageal, oral cavity, stomach, bladder, pancreatic and ovarian

cancers has been reviewed for almost 200 epidemiological studies (Block, Patterson, & Subar,

1992).

Phenolics are products of secondary metabolism in plants and have antioxidant activity

(Figure 2.2). These compounds are important in plants for protection against pathogens,

parasites and predators besides helping in reproduction and growth (Liu & Felice, 2007). In our

diet, 2/3 of all phenolics consist of flavonoids and 1/3 are derived from phenolic acids (Liu &

Felice, 2007). Flavonoids are found as conjugates in glycosylated or esterified forms in fruits and

vegetables. These types of compounds can also occur as aglycones as a result of food

Page 34: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

15

Phytochemicals

PhenolicsCarotenoids Alkaloids Nitrogen-

containingOrganosulfur

FlavonoidsPhenolic acids Stilbenes Coumarins Tannins

FlavonesFlavonols Isoflavonoids Flavanones Flavanols AnthocyanidinsHydroxy-

cinnamic acid

Hydroxy-

benzoic acid

Gallic acid

Protocatechuic

acid

Vannilic acid

Syringic acid

p-Hydroxy-

benzoic acid

P-Coumaric

acid

Caffeic acid

Ferulic acid

Sinapic acid

Apigenin

Chrysin

Luteolin

Quercetin

Kaempferol

Myricetin

Galangin

Fisetin

Genistein

Daidzein

Glycitein

formononetin

Catechin

Epicatechin

Epigallocatechin

Epicatechin-

gallate

Epigallocatechin

-gallate

Flavanones Cyanidin

Palargonidin

Delphidin

Peonidin

Petunidiin

Malvidin

Figure 2. 1. Classification of Phytochemicals. Adapted and modified from Liu & Felice (2007).

Page 35: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

16

OH

H3CO

CHO

Vanillin

OH

H3CO

Eugenol

OH

H3CO

O

O

OH OH

OHCOOH

Chlorogenic acid

OH

OH

O

O

COOH OH

OH

Rosmaric acid

OH

OH

O

OH

OH

OH

(+)- Catechin

OH

OH

O

OH

OH

OH

(-)- Epicatechin

OH

OH

O

OH

OH

O

O

rutinose

Rutin

OH

OH

O

OH

OH

Luteolin

Figure 2. 2. Some of the plant phenols in food with antioxidant capacity.

processing (Liu & Felice, 2007). The generic structure of flavonoids consists of 2 benzene rings

(A & B rings) linked by 3 carbons that are usually in an oxygenated heterocycle ring or C ring

(Figure 2.3). Hydroxybezoic acid and hydroxycinnamic acid derivatives are major phenolic

acids (Figure 2.4) present in bound form in plant cells. Hydroxybenzoic acid derivatives occur

as sugar derivatives and organic acids in plant foods besides present in lignins and hydrolyzable

tannins, whereas hydroxycinnamic acid derivatives are present in cellulose, lignin and proteins

through ester bonds. Food processing operations such as thermal processing, fermentation, and

freezing release these bound phenolic acids (Dewanto, Wu, Adom, & Liu, 2002).

Page 36: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

17

1.1. Antioxidants and Mechanism

Antioxidants are defined as the substances, when present at low concentrations compared to

those of an oxidizable substrate, which significantly delay or inhibit oxidation (process of losing

electrons) of that substrate (Halliwell & Gutteridge, 1990). A free radical is a species capable of

independent existence that contains one or more unpaired electrons (an unpaired electron being

one that is alone in an orbital) (Halliwell, Gutteridge, & Cross, 1992). For example, superoxide

(an oxygen-centered radical), thiyl (a sulfur-centered radical), trichloromethyl (a carbon-centered

radical), and nitric oxide are created as by-products of various metabolic reactions (Halliwell et

al., 1992). Antioxidants reduce localized oxygen concentration, prevent initiation of oxidation

(Halliwell & Gutteridge, 1990), inhibit radical oxygen species by directly scavenging free

radicals, singlet oxygen quenchers, peroxide decomposers, enzyme inhibitors or synergists such

as metal chelating agents or reducing agents (Namiki, 1990). In addition, both aqueous and lipid

phases exist in food systems with different reaction properties. For example, polar antioxidants,

such as ascorbic acid are dissolved in aqueous phase and react with hydrophilic hydroxyl or

peroxyl free radicals, whereas lipophilic antioxidants, such as tocopherols, dissolve in lipidic

phase reacting with liposoluble free radicals produced during lipid oxidation. Antioxidants,

according to their polarity, can accumulate at the water-oil interface forming an oriented mono-

molecular layer that protects the lipid phase against oxidation by oxygen dissolved in the

aqueous phase (Frankel, Huang, Kanner, & German, 1994). Such behaviors should be considered

while accessing the antioxidant functionality.

Food antioxidants are classified as (i) primary or chain-breaking antioxidants, (ii) synergist,

and (iii) secondary antioxidants (Figure 2.5). Major antioxidants in common fruits, vegetables

and grains are shown in Table 2.1. Primary antioxidants such as phenolic compounds terminate

Page 37: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

18

Figure 2. 3. Generic structure of Flavonoid

R1

R2

R3

COOH

Benzoic acid derivatives

R1

R2

R3

CH CH COOH

Cinnamic acid derivatives

Benzoic acid

Derivatives

Substitutions Cinnamic acid

derivatives

Substitutions

R1 R2 R3 R1 R2 R3

Benzoic acid H H H Cinnamic acid H H H

p-Hydroxybenzoic acid H OH H p-Coumaric acid H OH H

Protocatechuic acid H OH OH Caffeic acid OH OH H

Vanillic acid CH3O OH H Ferulic acid CH3O OH H

Syringic acid CH3O OH CH3O Sinapic acid CH3O OH CH3O

Gallic acid OH OH OH

Figure 2. 4. Structures of common phenolic acids.

the free radical chain reaction by donating hydrogen or electrons to free radicals and converting

them to more stable compounds. These antioxidants are effective in very low concentrations,

whereas they become prooxidants at higher concentration levels. Synergistic antioxidants, for

example ascorbic acid, act as hydrogen donators to the phenoxy radical to regenerate and provide

an acidic medium to stabilize the primary antioxidants. Secondary or preventive antioxidants

function by decomposing the lipid peroxides into stable end compounds (Rajalakshmi &

Narasimhan, 1996). The mechanism of antioxidant activity involving free radicals in lipid

oxidation includes three steps of initiation, propagation and termination (Figure 2.6). However,

O

5 4

3

2

6

7

2'

3'

4'

5'6'

8

A C

B

Page 38: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

19

an inhibition reaction is also considered that competes with the propagation step yielding more

stable products (Nawar, 1996).

1.2. Measurement of antioxidant activity

A number of methods have been used for measuring antioxidant activity, all of which

consider (i) measuring current state of oxidation in the model systems, and (ii) radical

scavenging assays. The methods include features such as a suitable substrate, an oxidation

initiator, and measurement of end point (Antolovich, Prenzler, Patsalides, McDonald, &

Robards, 2002). Shahidi & Zhong (2005) reviewed the methods for lipid oxidation.

Radical scavenging assays (Table 2.2) require no lipid substrate and directly involve either

measuring (i) hydrogen atom transfer (HAT) i.e. the ability of an antioxidant to quench free

radicals by hydrogen donation, or (ii) single electron transfer (SET) i.e. ability of a potential

antioxidant to transfer one electron to reduce any compound, including metal, carbonyl and

radicals in the assay. However, the end results in these assays remain the same, regardless of

mechanism, even if the kinetics and potential for side reactions vary (Prior, Wu, & Schaich,

2005). The concept of bond dissociation energy and ionization potential of antioxidants are

important to understand the mechanism and efficacy of their action (Wright, Johnson, &

DiLabio, 2001). HAT based methods (Table 2.2) use a free radical generator that generates

stable or short-lived radicals, an oxidizable molecular probe and an antioxidant. The added

antioxidant competes with probes for the radicals and thus, inhibits or retards the oxidation of the

probes. Most commonly used HAT based assays are oxygen radical absorbance capacity

(ORAC) and total radical-trapping antioxidant parameter (TRAP). SET based assay (Table 2.2)

such as ferric reducing antioxidant power (FRAP) involves an electron transfer from the

Page 39: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

20

Food

Antioxidants

Primary

Antioxidants

Secondary /Synergistic

Antioxidants

‗Hindred‘

PhenolsPhenols

Miscellaneous

Primary

Antioxidants

Oxygen

Scavengers

Chelating

Agents

Gallates

Hydroquinone

Trihydroxy-

butyrophenone

Nordihydroguairetic

acid

BHA

BHT

TBHQ

Tocopherols

Gum Guaiac

Ionox series

Sulfites

Ascorbic acid

Ascorbyl

palmitate

Erythorbic acid

Ethoxyquin

Anoxomer

Trolox-C

Polyphosphates

EDTA

Tartaric acid

Citric acid

Citrate esters

Phytic acid

Lecithin

Nitrites

Amino acids

Spice extracts

Flavonoids

Vitamin A

Β-Carotene

Tea extracts

Zinc

Selenium

Thiodipropionic

acid

Dilauryl,

Distearyl esters

Secondary

Antioxidants

Miscellaneous

Antioxidants

Figure 2. 5. Classification of food antioxidants. Adapted and modified from Rajalakshmi & Narasimhan (1996)

Page 40: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

21

Table 2. 1. Major antioxidants in some common fruits, vegetables and grains. Modified from

Sun & Powers (2007).

Source Major antioxidants

Fruits

Apple Benzoic acid, cinnamic acid, flavan-3-ols, anthocyanidin, flavonols, dihydrochalcones

Berry Anthocyanins, flavonols, flavanols, proanthocyanidins, ellagitannins, gallotannins,

stilbenoinds, phenolic acids

Grape Resveratol, catechin, anthocyanins, gallic acid

Grapefruit Narirulin, hesperetin, hesperidin, ascorbic acid

Vegetables

Asparagus Rutin, chlorogenic acid, ascorbic acid

Bean Quercetin, kaempferol glucoside, ascorbic acid

Beet Ferulic acid

Broccoli Caffeic acid, ferulic acid, quercetin, kaempferol, ascorbic acid

Cabbage Chlorogenic acid, caffeic acid, kaempferol, lutein, ascorbic acid

Carrot Chlorogenic acid, caffeic acid, β-carotene

Cauliflower Cinnamic acid, quercetin, ascorbic acid

Celery Chlorogenic acid, apigenin, apigenin glucoside, luteolin glucoside, α-tocopherol

Corn Caffeic acid, cinnamic acid, coumaric acid

Cucumber Ascorbic acid

Garlic Myricetin, apigenin, ascorbic acid

Lettuce Chlorogenic acid, caffeic acid, quercetin

Mushroom Ferulic acid, ascorbic acid

Onion Quercetin glucoside

Bell Pepper Caffeic acid, quercetin glucoside, luteolin glucoside, α-tocopherol, ascorbic acid

Potato Caffeic acid, cinnamon acid, p-hydroxybenzoic acid, ascorbic acid

Squash Β-carotene

Spinach Quercetin, ascorbic acid, patuletin glucoside, spinacetin glucoside, 5, 3‘-hydroxy-3-

methoxy-6,7-methylenedioxyflavone-4‘-glucronide methyl ester, 5-hydroxy-3,3‘-

dimethoxy-6,7- methylenedioxyflavone-4‘-glucronide methyl ester

Sweet potato Caffeic acid, cinnamic acid, α-tocopherol

Tomato Chlorogenic acid, caffeic acid, quercetin, ascorbic acid

Grains

Soybean Isoflavones, tocopherols, tocotrienols

Wheat Carotenoids, tocopherol, ferulic, vanillic, caffeic, coumaric and syringic acid, phytosterols

Rice bran Tocopherols, γ-oryzanol,

Oat Tocopherols, avenanthramides, p-hydroxybenzoic acid, vanillic acid, phytosterols

Corn Lutein, α- and β-carotene, β-cryptoxanthin, zeaxanthin, tocopherols, phytosterols

Page 41: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

22

Figure 2. 6. Mechanism of antioxidant activity involving free radicals. lipid radical (R·); peroxy

radical (ROO·); peroxide (ROOH); antioxidant (AH). Adapted from Shahidi & Zhong (2007)

antioxidants to the probe (oxidants), thus reducing the probe and oxidizing the antioxidants. The

oxidation reaction brings color change in the probe, which is proportional to the antioxidant

concentration and thus estimates the reducing capacity or antioxidant activity of the antioxidant

under investigation (Prior et al., 2005). Although trolox equivalent antioxidant capacity (TEAC)

and DPPH (2, 2-diphenyl-1-picrylhydrazyl) assays are also classified as SET based assays, these

two indicator radicals may be neutralized either by direct reduction via electron transfers or by

radical quenching via hydrogen atom transfer (Jimenez, Selga, Torres, & Julia, 2004). Table 2.3

provides a brief comparison of different antioxidant assays used to measure antioxidant activity

of fruits and vegetables.

2. PHENOLIC ANTIOXIDANTS OF FRUITS, VEGETABLE AND GRAIN

Consumption of foods such as grains, vegetables, and fruits containing antioxidant

compounds (Figure 2.7) may prevent many diseases and promote good health (Temple, 2000).

Vegetable consumption may provide protection against oxidative stress that is a pathogenic

mechanism of both carcinogenesis and atherosclerosis (Ames, Shigenaga, & Hagen, 1993). The

R H R

+ H

ROO

R

+ O2

ROO

+ R H

ROOH

+

R

ROO

+ A H

ROOH

+

A

Initiation

Propagation

Inhibition

Termination ROO

+ A

2 ROO

ROO

+R

R

+ R

Non-radical

products

Page 42: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

23

importance of various fruits and vegetables in the American diet has been compiled based on

their total phenolics, antioxidant capacity and daily consumption (Chun et al., 2005). The

mechanism for the protective effect of these phenolic antioxidants has been recently reviewed

(Kinsella, Frankel, German, & Kanner, 1993). Many of these phenols have been found to be

more powerful antioxidants than vitamins C, E, and β-carotene using an in vitro model for heart

disease, namely the oxidation of low density lipoproteins (Vinson, Dabbagh, Serry, & Jang,

1995). For example, antioxidant activities in colored potatoes are mainly due to the presence of

polyphenols / flavonoids, carotenoids, ascorbic acid, tocopherols, alpha-lipoic acid and selenium

(Lachman, Hamouz, & Orsak, 2005). The most common phenolics in potato are flavanols,

cinnamic acid, p-coumaric acid, caffeic acid, chlorogenic acid, ferulic acid, and quinic acid.

Studies on phenolic composition and antioxidant action of 23 vegetables showed highest total

phenol content and antioxidant activity (LDL oxidation) per fresh weight in beans (kidney and

pinto) (Vinson, Hao, Su, & Zubik, 1998). Similarly, Kahkonen et al. (1999) reported total

phenolic contents and antioxidant activities of a number of fruits, vegetables and cereal grains.

Alsaikhan, Howard, & Miller (1995) also reported that antioxidant activity of broccoli using β-

carotene/linoleic acid was highest followed by potato, carrot, onion and bell pepper. The main

antioxidative components in grain are classified as phenolic compounds such as anthocyanins,

tannins, and ferulic acid, and other substances (Martinez-Tome et al., 2004). White & Xing

(1997) reported the presence of p-hydroxybenzoic, protocatechuic, vanillic, trans-p-coumaric, (p-

hydroxyphenyl)- acetic, syringic, trans-sinapic, caffeic, and ferulic acids, with ferulic acid as the

most abundant phenolic acid in oat flour. According to Terao et al. (1993), caffeic acid showed

strong antioxidant activity followed by ferulic acid with moderate activity and p-coumaric acid

in a model solution. Chinese black-grained wheat has higher antioxidant activity compared to

Page 43: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

24

Table 2. 2. In vitro assays used for measuring antioxidant activity of fruits and vegetables.

Adapted from Huang, Ou, & Prior (2005)

Assays involving hydrogen atom transfer

reactions

ROO. + AH ROOH + A

.

ROO. + LH ROOH + L

.

1. ORAC (oxygen radical absorbance capacity)

2. TRAP (total radical-trapping antioxidant

parameter)

3. Crocin bleaching assay

4. IOU (inhibition oxygen uptake)

5. Inhibition of linoleic oxidation

6. Inhibition of LDL oxidation

Assays by electron transfer reaction

M (n) + e (from AH) AH.+ + M (n-1)

1. TEAC (Trolox Equivalent Antioxidant

Capacity)

2. FRAP (Ferric Reducing Antioxidant Power )

3. DPPH (2, 2-diphenyl-1-picrylhydrazyl)

4. Copper (II) reduction capacity

5. Total phenols assay by Folin-Ciocalteu reagents

Other assays 1. TOSC (total oxidant scavenging capacity)

2. Inhibition of Briggs-Rauscher oscillation

reaction

3. Chemiluminescence

4. Electrochemiluminescence

LH: substrate; AH: antioxidant; A.: antioxidant radical; ROO

.: peroxyl radical; M: stable radical;

e: electron.

white and blue wheat genotypes (Li, Shan, Sun, Corke, & Beta, 2005).

Genetics, environment (location), growing conditions (moisture, fertilization, pests and

disease burden etc.), processing methods, and storage affect the level of antioxidant activity of

phytochemicals in fruits and vegetables (Blessington, 2005; Dewanto, Wu, Adom et al., 2002;

Reyes, Miller, & Cisneros-Zevallos, 2004). For example, highest intensity of skin and flesh

color in potatoes has reported to be in those grown in sandy soils (Burton, 1989). A significant

interaction between the antioxidant capacity and year-wise genotype of blueberry cultivars is

reported (Connor, Luby, Tong, Finn, & Hancock, 2002). Alsaikhan et al. (1995) found that the

phenolic content and antioxidative activity of four potato cultivars was genotype-dependent and

not related to flesh color.

Page 44: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

25

Table 2. 3. Comparison of different in vitro total antioxidant capacity assays. Method ORAC TRAP FRAP TEAC TOSC Crocin DPPH

Reagents AAPH or

ABAP, Trolox,

Fluorescein

AAPH or ABAP,

Trolox,

R-phycoerythrin

TPTZ, FeCl3,

sodium acetate

ABTS, K2SO4,

Trolox

KMBA,

AAPH or

ABAP

Crocin, AAPH

or ABAP,

Trolox

DPPH

Instrument Fluorescent

plate reader

spectrometer

Fluorescent plate

reader

spectrometer

Spectro-

photometer

Spectro-

photometer

Head space

GC

Spectro-

photometer

Spectro-

photometer

Temperature 37 °C 37 °C 37 °C 37 °C 37 °C 37 °C 37 °C

Absorbance

(nm)

Excitation: 485

Emission: 520

Excitation: 495

Emission: 575

593 415 or 734 Depending on

detector type

443 515

pH 7.4 7.4 3.6 7.4 7.4 7.4

Endpoint End of

fluorescence

decay

Lag phase Fixed time

(4 – 10 min)

Fixed time

(4 – 6 min)

End of

production of

ethylene

Fixed time

(10 min)

% remaining

DPPH

Calibration Trolox solution Trolox solution Fe3+

standard

solution

Trolox solution - Trolox

solution

Trolox solution

Calculation of

Results

Area of the

fluorescence

curve decay

Length of the lag

phase

Absorbance of

final reading

minus blank

Absorbance

decrease in

presence of the

sample

Area under the

kinetic curve

Ka/Kc of

antioxidant

divided for

Ka/Kc of

Trolox

Absorbance

decrease in

presence of the

sample

Dimension of

results

µmol of Trolox

equivalent

µmol of Trolox

equivalent

Absorbance of

Fe2+

complex

produced by

antioxidant

reduction of

corresponding

tripyridyltriazine

Fe3+

complex

mM Trolox

equivalent to 1

mM test

substance

µmol of

Trolox

equivalent

% inhibition;

EC50; TEC50;

AE =

(1/ EC50)TEC50

ORAC (oxygen radical absorbance capacity); TRAP (total radical-trapping antioxidant parameter); TEAC (trolox equivalent antioxidant capacity);

FRAP (ferric reducing antioxidant power); DPPH (2, 2-diphenyl-1-picrylhydrazyl); TOSC (total oxidant scavenging capacity); ABTS (2,2'-azino-

bis(3-ethylbenzthiazoline-6-sulphonic acid)); AAPH or ABAP (Azo-bis(2-amidinopropane); TPTZ (2,4,6-tripyridyl-s-triazine); KMBA(α-keto-γ-

methiolbutyric acid); EC50 (concentration to decrease concentration of test free radical by 50%); TEC50 (time to decrease concentration of test free

radical by 50%);AE (antiradical efficiency). Adapted from Antolovich et al.(2002).

Page 45: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

26

Figure 2. 7. Total phenols from daily consumption of fruits and vegetables in the American diet.

Adapted from Chun et al. (2005).

Selected and relatively unstable antioxidants of nutritional interest (e.g. ascorbic acid) have

been commonly assessed as indicators of processing damage (Miller, Diplock, & Riceevans,

1995). However, it was reported that vitamin C provides less than 0.4% of total antioxidant

activity in apples (Dewanto, Wu, & Liu, 2002) . Antioxidant activity is correlated with the

occurrence of polyphenols including anthocyanins (Eberhardt, Lee, & Liu, 2000; Prior et al.,

1998). Complex mixtures of phytochemicals in whole foods are responsible for their health

Page 46: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

27

benefits and are better than single antioxidants due to a combination of additive and/or

synergistic effects (Eberhardt et al., 2000; Liu, 2002). Processed tomatoes and sweet corn exhibit

higher antioxidant activities than fresh ones due to the increased release of bound phenolic

compounds (Dewanto, Wu, Adom et al., 2002; Dewanto, Wu, & Liu, 2002) despite loss in

vitamin C content.

2.1. Effect of processing on phytochemicals

Processing of foods involve heating with different energy transfer media such as water, air,

oil and electromagnetic waves. In addition, storage can also be classified as passive processing

with no energy applied directly to foods (Figure 2.8). Polyphenolic compounds including

anthocyanins and proanthocyanidins are not completely stable during processing (Talcott,

Brenes, Pires, & Del Pozo-Insfran, 2003). Physical and biological factors such as temperature

increase and enzymatic activity may result in destruction of phenolics such as phenolic acids and

anthocyanins. After harvest these compounds can change during food processing and storage

(Kader, Irmouli, Nicolas, & Metche, 2002; Rossi et al., 2003), which may reduce related

bioactivity. During the processing of foods, various transformations of phenolics occur to

produce yellowish or brownish pigments (Clifford, 2000). Generally, food-processing procedures

are recognized as one of the major factors on the destruction or changes of natural

phytochemicals, which may affect the antioxidant capacity in foods (Nicoli, Anese, & Parpinel,

1999). But recent research has now established that food processing also has some positive

effects that improve the quality and health benefits of foods. These effects are mainly attributed

to the retained availability of some antioxidants. In addition, some compounds with potential

antioxidant activity after processing increases. For example, both hydrophilic and lipophilic-

Page 47: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

28

Figure 2. 8. Heating/energy transfer medium in various processing methods.

ORAC values in cooked tomatoes are significantly higher than in their raw forms (Wu, X. et al.,

2004; Wu, X. L. et al., 2004). Significant increases in hydrophilic-ORAC but decreases in

lipophilic-ORAC values are reported in baked russet potatoes compared to raw ones (Wu, X. et

al., 2004; Wu, X. L. et al., 2004). In contrast, lipophilic and hydrophilic-ORAC values of raw

broccoli and carrots are significantly higher than that of cooked forms (Wu, X. et al., 2004; Wu,

X. L. et al., 2004). Interestingly, cooking in boiling water decreases the radical scavenging

activity of peppers, whereas microwave heating without water increase the activity (Chuah,

Yamaguchi, & Matoba, 2005). Antioxidant properties of grains are affected by processing

temperature during thermal processes (Dewanto, Wu, & Liu, 2002). Antioxidant activity of

cereals during processing is important as phenolics are localized in outer layers (husk, pericarp,

Food Boiling

Blanching

Pasteurization

Sterilization

Evaporation

Extrusion cooking

Roasting

Baking

Drying

Microwave heating

Irradiation

Pulsed electric field

Ultrasound

sonication

Processed Food

Water

Air

Electromagnetic

Waves

Shallow and

Deep Fat Frying

Oil

Storage

(air, vacuum, inert gas,

refrigeration, freezing)

Page 48: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

29

testa, and aleurone cells) of the grain than in any other component. About 80% of the trans-

ferulic acid in rye and wheat grain was observed in the grain (White & Xing, 1997). Similarly,

freeze-dried fractions from durum wheat (Triticum durum) bran exhibit stronger antioxidant

activity than extracts from other milling fractions (Onyeneho & Hettiarachchy, 1992). Thus,

processing can alter antioxidant activity in both positive and negative ways. In the last decade,

excellent reviews by Nicoli et al. (1999), Klein & Kurilich (1993), and Kaur & Kapoor (2001)

provide brief information on the antioxidant activity as influenced by processing. Thus the

evaluation of processing factors influencing the antioxidant activity is imperative in optimizing

the conditions to increase or retain their activity and availability.

2.1.1. Extraction

Yields of phenolic contents from fruits and vegetables are affected by various extraction

solvents. Methanol has been reported to extract lower molecular weight phenolic compounds

compared to hexane and aqueous acetone in a fruit (Malus domestica) (Guyot, Marnet, Laraba,

Sanoner, & Drilleau, 1998). Extraction solvents have different effects on the peel, flesh and seed

of fruits. For example, phenolic content of pomegranate peel in methanol extract was more than

in ethanol extracts followed by water, whereas water extract of seeds contained high phenolic

content followed by methanol and ethanol extracts (Singh, Murthy, & Jayaprakasha, 2002). The

methanolic extract from pomegranate peel showed varied inhibition capacity to different

antioxidant assays such as 56, 58, and 93.7% in the thiobarbituric acid method, hydroxyl radical

scavenging activity and LDL oxidation, respectively (Singh et al., 2002). In another study,

acidified aqueous methanol effectively extracted more total phenolics and exhibited greater

antioxidant capacity from blueberries than acidified acetonitrile, or an acidified mixture of

methanol and acetone (Kalt et al., 2001). Among six extraction solvents (hexane, chloroform,

Page 49: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

30

ethyl acetate, acetone, methanol and water), the methanolic and aqueous extracts of Decalepis

hamiltonee, a tuber, has higher antioxidant activity using DPPH, superoxide and hydroxyl

radicals, and inhibits microsomal lipid peroxidation and exhibits strong reducing power and

metal chelating activity (Srivastava, Harish, & Shivanandappa, 2006). However, the phenolics

content in the methanolic extract of Decalepis hamiltonee was higher than that of the aqueous

extract. Whole oats extracted with 80% methanol resulted in substantially higher levels of total

phenolic compounds and exhibit higher antioxidant capacity than water extracts (Zielinski &

Kozlowska, 2000), while both 80% methanol or ethanol were found to be efficient at extracting

phenolic compounds from barley (Bonoli, Verardo, Marconi, & Caboni, 2004).

Among eight solvent combinations involving ether, diethyl ether, petroleum ether,

chloroform, dichloroethane and methanol it was found that methanol extracts had greatest

antioxidant activity in oat groats and hulls (Duve & White, 1991). Comparison of various

aqueous ethanol, methanol or acetone mixtures showed more phenolics were extracted from

wheat flour and bran using successive acidified methanol/water (50:50, v/v, pH 2) and acetone/

water (70:30, v/v) than 70:30 (v/v) of either ethanol: water or methanol: water (Perez-Jimenez &

Saura-Calixto, 2005). Better extraction efficiency of methanol: HCl and ethanol: HCl could be

attributed to the polarity difference of acidic extracting solvents compared to methanol solvent

(Li, Pickard, & Beta, 2007). In contrast, 95% ethanol: HCl (85:15 v/v) has found to be more

efficient in total phenolic extraction than methanol: HCl (99:1 v/v) followed by 100% methanol

in purple wheat bran (Li et al., 2007). Comparison of four solvent systems including 50%

acetone (v/v), 70% methanol (v/v), 70% ethanol (v/v), and 100% ethanol, 50% acetone (v/v)

showed as a better solvent for extracting phenolic antioxidants from wheat (Zhou & Yu, 2004).

Application of a second component in addition to the extraction solvents also enhanced

Page 50: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

31

extractability of phenolic antioxidants. Combination of supercritical carbon dioxide and ethanol

was found to extract greater antioxidant activity from marjoram than with either alone, or in

combination with hexane (Vagi et al., 2005). It was also observed that extraction of phenolic

compounds and antioxidant capacity from oat bran concentrate can be enhanced by microwave-

irradiating at 150 °C for 10 min in 50% ethanol (Stevenson et al., 2008).

2.1.2. Blanching

Blanching is an important processing step applied to soften the product as well as to

inactivate the enzymes that otherwise could cause browning or other possible reactions. The

blanching efficiency is determined by taking into account the complete inactivation of

peroxidase. The most commonly used method for thermal inactivation is heating by steam or hot

water, where the resistance to heat transfer at the surface is negligible compared to the internal

resistance to heat transfer (mechanism of heat penetration is by conduction). Thus, blanching

time depends on the dimension of food matrix. For example, blanching time for small size

products, such as peas, are 1-2 min, while for larger products, such as corn on the cob, is 11 min

(Feinberg, Winter, & Roth, 1968). The longer time required for the temperature rise at the cold

spot or slowest heating point, normally the geometric center, of a relatively large object such as

corn on the cob could damage the quality of the kernels, whereas shortening the length of

blanching time could reduce the degree of enzyme inactivation and thus result in shorter shelf-

life, nutritional and functional value. Thus, optimal blanching time is necessary for a particular

fruit or vegetable to preserve its overall nutritional and health promoting components.

A blanching time of 1 min in boiling water has been recommended for green leaves of sweet

potato to retain high antioxidant activity (Chu, Chang, & Hsu, 2000). Blanching opens up the

cell matrix and therefore, could increase the polyphenols yield during extraction that may either

Page 51: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

32

enhance or reduce the antioxidant activity. Blueberry juice has higher recovery of phenolic

compounds and strong radical-scavenging activity to DPPH and hydroxyl radicals after steam

blanching for 3 min than the non-blanched juice (Rossi et al., 2003). In contrast, blanching in

water at 98°C for 2 min diminishes the antioxidant capacity of purple carrots (Uyan, Baysal,

Yurdagel, & El, 2004). Amin & Lee (2005) observed that even 5 – 10 min of blanching in hot

water at 98°C reduced (p < 0.05) antioxidant activities and phenolics content of all vegetables

except for cabbage and mustard cabbage. After 15 min of blanching, the loss of antioxidant

activity (β-carotene bleaching assay) was highest in Chinese cabbage (40%) followed by Chinese

white cabbage (19%), mustard cabbage (9%) and red cabbage (4%). However, the total phenolic

content of Chinese cabbage increased (p < 0.05) after 15 min of blanching compared with other

vegetables.

Mizrahi (1996) observed that 2 min of ohmic blanching of large whole vegetables had similar

effects as 4 min of water blanching. The investigator also reported that the energy dissipated by

the electric current passing through the samples was capable of heating it uniformly and very

quickly regardless of its shape or size. In another study, ohmic blanching (25–40 V/cm) of

artichoke by-product was faster at inactivating the peroxidase enzyme, without producing

blanching waste water compared to hot water blanching at 85°C, thus retaining higher total

phenolic content (Icier, 2010). The same investigator reported that ohmic blanching (40 V/cm) at

85 °C had similar peroxidase inactivation times (310 ± 2 s) as water blanching at 100 °C (300 ±

2 s). Increase in the mass transfer and effective diffusion rate has been reported after ohmic

blanching of strawberries during osmotic dehydration (Allali, Marchal, & Vorobiev, 2010).

Peeling of fruits and vegetables followed by blanching affects phenolic antioxidants. For

example, peach puree containing periderm tissue blanched in boiling water (20 min) and

Page 52: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

33

pasteurized (boiling water for 30 min) was 7 – 11% higher in antioxidant activity (β-

carotene/linoleic acid assay) than peeled samples (Talcott, Howard, & Brenes, 2000b). Shorter

blanching of puree with periderm in boiling water (2 min) yielded the lowest initial antioxidant

activity, but the level of retention was greater during storage than that of peeled samples. The

same investigators also reported that total water-soluble phenolic compounds were higher in the

case of longer blanching time than that of shorter blanching time. That was due to increased

tissue softening and enhanced chemical extraction with the additional heat applied prior to

pasteurization.

2.1.3. Cooking

Cooking of vegetables, fruits and grains has a mixed effect on phenolic content and

antioxidant activity. Processing and heating during jam making (at 104 – 105 °C) decreases the

content of total phenolics of some varieties of cherries and plums, whereas no significant change

(p < 0.05) has observed in raspberries and in some varieties of cherries and plums (Kim &

Padilla-Zakour, 2004). However, the same study reported an increase or a decrease in antioxidant

activity (ABTS assay) of cherries, plums and raspberries depending on variety during jam

processing. Canning of raspberries and blueberries increases the phenolic content and antioxidant

activity by 50 and 53% respectively (Sablani et al., 2010). In contrast, the total phenolics has

decreased after dehydration of fresh plums, while antioxidant capacity in dried plums increased

compared to fresh plums (Piga, Del Caro, & Corda, 2003). During baking, the outer layer is

usually heated to over 120 °C, while the inner temperature remains lower than 95 – 100 °C. At

these temperatures, in addition to caramelization (between reducing sugars and ascorbic acid),

Maillard reaction (between reducing sugar and amino acids), Strecker degradation (dicarbonylic

compounds with amino acids), and hydrolysis of esters and glycosides of antioxidants, oxidation

Page 53: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

34

of phenolic antioxidants to quinones and their polymers occur. Chemical or enzymatic oxidation

of polyphenols is generally responsible for their loss of antioxidant activity. However, baking of

purple wheat bran at 177 °C for 20 min has not altered the total phenolic content in the processed

samples (Li et al., 2007). By contrast, the total phenolics and total antioxidant activity of sweet

corn has increased by 54 and 44%, respectively, after thermal processing at 100 – 121 °C for 10

– 50 min (Dewanto, Wu, & Liu, 2002). In other studies, antioxidant activities in processed

tomatoes (Dewanto, Wu, Adom et al., 2002; Re, Bramley, & Rice-Evans, 2002) and coffee

(Nicoli, Anese, Parpinel, Franceschi, & Lerici, 1997) were retained or higher than their fresh

equivalents. The increase or retention of antioxidant activities in processed foods is attributed to

the development of new compounds with potential antioxidant capacity (Figure 2.9), although

the content of naturally occurring antioxidants has significantly decreased due to the heat

processing (Anese, Manzocco, Nicoli, & Lerici, 1999; Nicoli et al., 1999; Nicoli et al., 1997).

Antioxidant compounds depletion in thermally treated fruits and vegetables may also be

attributed to consumption of ascorbic acid and polyphenols as reactants in the Maillard reaction

(Kaanane, Kane, & Labuza, 1988). Although a decrease in the antioxidant potential is found for

short heat treatments, a recovery of these properties has been reported during prolonged heat

treatment. For example, Jiratanan & Liu (2004) observed 12% reduction in phenolic content of

the beets at initial application of heat (115 °C for 15 – 30 min), but further processing raised its

content back to the equivalent of unprocessed beets and eventually increased by 14% after

processing at 115 °C for 45 min. Similar results in the antioxidant capacity of aged citrus juice

and orange juice has been reported as they became more discolored (Lee, 1992). The antioxidant

activity of Maillard reaction products can be mainly attributed to the high molecular weight

brown compounds, which are formed in the advanced stages of reaction.

Page 54: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

35

Figure 2. 9. Changes in total antioxidant activity in vegetable matrix subjected to different

processing conditions. Adapted and modified from Nicoli et al. (1999).

The initial reduction in the antioxidative activity can be attributed not only to thermal

degradation of naturally occurring antioxidants but also to formation of early MRPs with pro-

oxidant properties. The gain in antioxidant activity coincided with the formation of brown

MRPs.

Phenolic contents, including free and bound, and antioxidant activity during processing also

depend on the type of crop. Either increase or non-significant change in free, bound and total

phenolic content, total flavonoids, and total antioxidant activity of table beets were observed

during heat treatment at 105 – 125 °C for 15 – 45 min (Jiratanan & Liu, 2004). The same

investigators observed reductions in the antioxidant activity, phenolic contents and total

flavonoids (majority comes from free flavonoids) in green beans at similar processing conditions

100 – 121 °C for 10 – 40 min. The authors hypothesized that the processed beets and green beans

would contribute little to release of bound phytochemicals in the colon by intestinal microflora to

induce a site-specific reinforcement of antioxidants (Adom & Liu, 2002; Andreasen, Kroon,

Williamson, & Garcia-Conesa, 2001). After being heated at 90 °C for 147 h, the radical

Heating time

Anti

oxid

ant

acti

vit

y

Total antioxidant activity

Natural antioxidants

New antioxidants

Page 55: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

36

scavenging capacity (DPPH) of the roselle pigment model system increased from 18% to 43%,

and FRAP decreased from 980 to 640 mol/l, while TEAC (ABTS assay) remained around 2.2

Trolox (mmol/l) (Tsai & Huang, 2004).

Phenolic antioxidant activity in-vitro sometimes has a different effect on the bioavailability

of the compounds. For example, although heat treatments including moderate heating (IQF and

freeze-drying), high temperature heating (cooking at 100 °C for 5 min, canning and spray-

drying) and jam preparation retain most of the phenolic content and antioxidant activity (FRAP

and DPPH) of wild or cultivated blueberries found in unprocessed whole fruit, processing

diminished antiproliferation activity on heap-1c1c7 cells (Schmidt, Erdman, & Lila, 2005). The

changes in the antiproliferation activity on heap-1c1c7 cells are thought to be the breakdown of

the active proanthocyanidin oligomers ranging from monomeric catechin units to hexamers in

blueberries (Kayano et al., 2003).

Microwave cooking of vegetables did not resulted in a specific trend on phenolic

antioxidants. Microwave cooking (with a power of 800 Watt) either retains or decreases a small

quantity of total phenolic content in cauliflower, peas, spinach and Swiss chard, whereas

significant reduction occurs after boiling (6 – 13 min) (Natella, Belelli, Ramberti, & Scaccini,

2010). The same investigators observed significant increase or retention of total antioxidant

capacity in cauliflower, peas, spinach, Swiss chard, potatoes, tomatoes and carrots. In another

study, researchers investigated the effect of microwave cooking on different radical scavenging

capacity (lipoperoxyl, hydroxyl and ABTS radicals) of vegetables (Jiménez-Monreal, García-

Diz, Martínez-Tomé, Mariscal, & Murcia, 2009). The researchers observed 30 – 50% losses of

lipoperoxyl radical scavenging capacity in most of the vegetables, whereas artichoke, asparagus,

garlic, onion, and spinach retained and eggplant, maize, pepper, and Swiss chard increased their

Page 56: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

37

antioxidant activity after microwave cooking. Microwave cooked potatoes retained 55% of the

chlorogenic acids in the potatoes (cv NDA 1725) followed by boiled potatoes (35%), whereas

oven-baked potatoes lost all of them (Dao & Friedman, 1992). Commercially processed french-

fried potatoes, mashed potato flakes, and potato skins contained no chlorogenic acid in the same

study. The nature of the heat and their harshness in different cooking methods is attributed to

relative loss of chlorogenic acid in the potato. The total antioxidant activity of asparagus, using

ABTS assay, is significantly higher (P < 0.05) after refractance window and freeze-drying than

tray-drying, spouted bed, and combined microwave and spouted bed drying compared to raw

material (Nindo, Sun, Wang, Tang, & Powers, 2003). Similar results were observed in combined

microwave and hot air-drying of purple carrots (Uyan et al., 2004). By contrast, microwave

cooking of broccoli severely degrades phenolic content and antioxidant capacity in florets and

stems (Zhang & Hamauzu, 2004), with similar findings for potatoes (Tudela, Cantos, Espin,

Tomas-Barberan, & Gil, 2002). Studies on green vegetables and herbs found phenolic content

and antioxidant capacity increased or remained unchanged after microwave irradiation

(Turkmen, Sari, & Velioglu, 2005). Microwave heating of apple mash from 40 °C to 70 °C

results in an increase in phenolic and flavonoid compounds in the juice with increase in

temperature (Gerard & Roberts, 2004).

2.1.4. Drying/Dehydration

Saskatoon berries (‗Thiessen‘ and ‗Smoky‘ cv.) processed using freeze-drying, vacuum

microwave-drying, air-drying, and a combination of air-drying and vacuum microwave-drying

methods caused reduction (P < 0.05) in total phenolics, antioxidant activities and anthocyanin

contents as compared with fresh frozen berries (Kwok, Hu, Durance, & Kitts, 2004). However,

freeze-drying followed by vacuum microwave-drying render maximum antioxidant activity in

Page 57: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

38

the berries. Similar results are observed during drying of cranberries using vacuum microwave,

freeze-drying and air-drying method. Vacuum microwave-drying produce dehydrated product

with higher ORAC antioxidant activity followed by freeze-drying and air-drying (Leusink, Kitts,

Yaghmaee, & Durance, 2010). However, the antioxidant activity (ABTS) of freeze dried samples

was higher than vacuum microwave and air dried samples. One possible explanation for this

discrepancy could be that freeze-drying degraded cranberry antioxidants that contribute to the lag

phase more than vacuum microwave-drying. During microwave-drying, water molecules

transmit and absorb energy because of volumetric heat generation in the wet sample resulting in

higher interior temperature that helps water rapidly reached its boiling point compared to other

convective drying methods (Khraisheh, Cooper, & Magee, 1995). Microwave-drying in

combination with vacuum also reduce the drying temperature, oxygen exposure, and heating

time because of enhanced penetration of heat that provides a constant internal temperature and

lasts until the final stage of drying has been reached (Oliveira & Franca, 2002).

Steaming and flaking oat groats has been reported to decrease tocotrienols, caffeic acid and

some avenanthramides, but increase ferulic acid and vanillin, whereas autoclaving whole grains

increases contents of tocopherols, tocotrienols, and acids of vanillin, ferulic and p-coumaric, but

degrade avenanthramides (Bryngelsson, Dimberg, & Kamal-Eldin, 2002). Drum-drying of whole

meal or rolled oats results in decreases in all tocols and phenolic compounds but

avenanthramides are unaffected (Bryngelsson et al., 2002).

2.1.5. Irradiation

The antioxidant activity decreases in kiwi fruits after treatment with 2 and 3 kGy compared

to the non-irradiated and 1kGy during storage (1 – 3 weeks) (Kim & Yook, 2009). The same

investigators reported that electron-donating ability (DPPH) decreased slightly during storage for

Page 58: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

39

control samples; however, samples irradiated at1 and 2 kGy doses were not significantly

different throughout storage time. The total phenolic content and antioxidant capacity (FRAP) of

the irradiated (3 and 5 kGy) carrot juice are higher than that of the non-irradiated control,

whereas antioxidant activity of kale juice decreases after irradiation (5kGy) (Song et al., 2006).

Breitfellner, Solar, & Sontag (2002) studied the effect of gamma radiation (1 – 10 kGy) on

phenolic acids such as 4-hydroxybenzoic acid, gallic acid, cinnamic acid, p-coumaric acid, and

caffeic acid and their hydroxylation products in strawberry. The investigators observed a

significant increase in 4-hydroxybenzoic acid with the gamma radiation doses. Similarly, an

increase in the antioxidant activities due to irradiation was observed in fruit juice (2002) and

alfalfa sprouts (Fan & Thayer, 2001). Free radicals generated during irradiation act as stress

signals and may trigger stress-responses in lettuce (Fan, Toivonen, Rajkowski, & Sokorai, 2003)

resulting in increased antioxidant synthesis. Fan et al. (2003) hypothesized that irradiation

stimulates the synthesis of phenolics but not that of vitamin C, although both phenolics and

vitamin C are antioxidants. Gamma radiation of citrus fruit induces accumulation of 4-(3-methyl-

2-butenoxy) isonitrosoacetophenone that exhibits both antioxidant and antifungal activities

(Dubery, Louw, & van Heerden, 1999). In another study, seeds of kidney bean, cabbage and

beets exposed to ultraviolet (UV) irradiation (460-760 µW/cm2 for 30, 60 and 90 min)

stimulated synthesis of anthocyanins in leaves of kidney bean varieties (3-6%) and white beet

(14-21%) (Kacharava, Chanishvili, Badridze, Chkhubianishvili, & Janukashvili, 2009). In the

same study, low doses of irradiation were also effective (9-20%) in stimulating synthesis of

anthocyanins in cabbage and red beet. Free radicals produced by UV irradiation of seeds change

cell membrane permeability and electric potential, presumably initiating diverse metabolic

Page 59: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

40

responses including biosynthesis of antioxidants in addition to some stress factors (Barka,

Kalantari, Makhlouf, & Arul, 2000).

2.1.6. Extrusion

Extrusion cooking increases the phenolic content of oats (Zielinski, Kozlowska, & Lewczuk,

2001), cauliflower by-products (Stojceska, Ainsworth, Plunkett, Ibanoglu, & Ibanoglu, 2008),

and increases the antioxidant activity in sweet potatoes (Shih, Kuo, & Chiang, 2009), cauliflower

by-products (Stojceska et al., 2008) and retains in the fruit powders from blueberries,

cranberries, concord grapes and raspberries (Camire, Dougherty, & Briggs, 2007). The increase

or retention of phenolic compounds and antioxidant activity of the extrudates is the consequence

of the high temperature, water-stress and wounding, (Reyes, Villarreal, & Cisneros-Zevallos,

2007) and partly accounted for the presence of the high molecular weight Maillard reaction

products, which are formed at higher temperatures and act as antioxidants. In contrast, reduction

in phenolic content has also been reported for extruded bean (19–21%), oat cereals (24–46%)

and oat extrudates (50%) (Korus, Gumul, & Czechowska, 2007; Viscidi, Dougherty, Briggs, &

Camire, 2004; Zadernowski, Nowak-Polakowska, & Rashed, 1999). Antioxidant activities

(DPPH) and total phenolics in barley extrudate samples was reduced by 60–68% and 46–60%,

respectively, compared with that of the unprocessed barley flour (Altan, McCarthy, & Maskan,

2009). Similar losses in antioxidant activity during extrusion cooking were observed in sorghum

(Dlamini, Taylor, & Rooney, 2007) and grass peas (Grela, Jensen, & Jakobsen, 1999). The loss

of natural antioxidants during extrusion over 80 °C has been attributed to their low resistance to

heat (Zadernowskl, Nowak-Polakowska, & Rashed, 1999), evaporation and decomposition

(Hamama & Nawar, 1991) at elevated temperatures. It was also reported that high temperature

during extrusion can alter molecular structure of phenolic compounds and either reduce their

Page 60: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

41

chemical reactivity or decrease their extractability due to a certain degree of polymerization

(Alonso, Grant, Dewey, & Marzo, 2000) causing loss of antioxidant properties (Zadernowskl et

al., 1999). Interestingly, the effect of screw speed (co-rotating twin screw extruder) did not

follow a specific trend on the losses of antioxidant activity and total phenolics content of

extrudates from barley (Altan et al., 2009) and dry-mix of chick peas, corn, oats, carrot powder

and hazelnuts (Ozer, Herken, Guzel, Ainsworth, & Ibanoglu, 2006). The investigators

hypothesized that the increased shearing effects at increased speed were more dominant than the

effect of residence time on the destruction of antioxidant activity over the extrusion condition

even though increased screw speeds associates with decreased residence times.

2.1.7. Non-thermal processing

Effect of high pressure treatments on fruit and vegetable products are dependent on the food

matrix and processing parameters (pressure, time, and temperature), and antioxidant activity

assay methods (de Ancos, Gonzalez, & Cano, 2000; Fernandez-Garcia, Butz, Bognar, &

Tauscher, 2001; Fernandez-Garcia, Butz, & Tauscher, 2001; Sanchez-Moreno, De Ancos, Plaza,

Elez-Martinez, & Cano, 2009). Slight modification in the nutritional composition (vitamins C, A,

E, B1, B2, and folic acid), bioactive compounds and their antioxidant capacity (vitamins C, A

and E, carotenoid compounds, flavonoids) has been reported in purees (persimmons, tomatoes,

strawberries, kiwifruit), juices (lemon, orange, carrot, apple and broccoli) and a mix of vegetable

soup (gazpacho) (Donsi, Ferrari, & DiMatteo, 1996; Quaglia, Gravina, Paperi, & Paoletti, 1996;

Sanchez-Moreno et al., 2009).

Long time high pressurization (600 MPa/30 min) combined with thermal treatment (60 °C)

causes higher reduction in antioxidant capacity (linoleic acid/β-carotene assay) of freshly

squeezed apple juice than thermal treatment alone (60 °C/30 min) 25 and 10%, respectively,

Page 61: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

42

compared to untreated samples (Sanchez-Moreno et al., 2009). Similarly, the antioxidant

capacity of tomato puree (DPPH) is significantly reduced after high pressure treatments (400

MPa/25 °C/15 min), although no difference is observed between pressurized tomato puree and

low and high pasteurized products (70 °C/30 s and 90 °C/1 min). In contrast, the antioxidant

capacity (ABTS) of fruit juices (orange, apple, peach, and citrus) and vegetable purees (carrots

and tomato) has not affected after high pressure treatment (600 MPa/20 °C/60 min) (Butz et al.,

2003). Similar results have been reported on the antioxidant capacity (DPPH) of freshly

squeezed orange juice with combined treatments of high pressure/temperature (100 MPa/60

°C/5min, 350 MPa/30 °C/2.5 min, 400 MPa/40 °C/1 min) (Plaza, Sanchez-Moreno, Elez-

Martinez et al., 2006; Sanchez-Moreno, Plaza, de Ancos, & Cano, 2003).

Pulsed electric field (PEF) treatment (35 kV/cm in bipolar mode, 800 Hz pulse frequency, 4

μs pulse width, 750 μs total treatment time, temperature ≤ 50 °C) of freshly squeezed orange

juice has not altered the contents of naringenin, hesperetin or total flavanone, whereas

pasteurization using heat only (90 °C/1 min) reduces naringenin content of squeezed orange juice

(Sanchez-Moreno et al., 2005). Similarly, apple juice pasteurized using a PEF treatment (35

kV/cm, bipolar pulses of 4 μs 1,200 pulses per second) has a 14.5% reduction in the total

phenolic content, whereas thermal pasteurization (90 °C, 30 s) reduced it by 32.2% (Aguillar-

Rosas, Ballinas-Casarrubias, Nevarez-Moorillon, Martin-Belloso, & Ortega-Rivas, 2007). Plaza,

Sanchez-Moreno, De Ancos, & Cano (2006) reported that a low pasteurization treatment (70 °C,

30 s) or a PEF treatment of 35 kV/cm for 750 μs (4-μs bipolar pulses, 800 Hz) did not

significantly modify the antioxidant activity of treated compared to untreated orange juice.

Ultraviolet (UV) rays induce an increase in enzymes responsible for the biosynthesis of

secondary metabolites such as flavonoids, which act as UV screens preventing UV-induced

Page 62: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

43

damage in the genetic material of plant cells (Cantos, Garcia-Viguera, de Pascual-Teresa, &

Tomas-Berberan, 2000). Ultrasound (US) releases enzymes from cells for the secondary

metabolite biosynthesis due to mechanical stresses and microstreaming induced by acoustic

cavitations at low US intensity levels (Lin, Wu, Ho, & Qi, 2001). Prolonged US pre-treatment

combined with the same air-drying conditions decreases total phenolics, flavonoid contents and

the antioxidant capacity of dried apples (Opalic et al., 2009). The same investigator observed that

drying time had no significant effect on the contents of total phenolics, flavonoid or antioxidant

activity. However, the samples dried without the ultrasound pre-treatment was most sensory

acceptable. In contrast, ultraviolet (UV) rays and ultrasound increase the total phenolics, total

antioxidant capacity (TEAC and ORAC) of peanuts and US is more effective than UV in

increasing total antioxidants (Sales & Resurreccion, 2010).

2.1.8. Storage

The effects of storage on the polyphenols have been reported in apples (Price, Prosser,

Richetin, & Rhodes, 1999), broccoli (Price, Casuscelli, Colquhoun, & Rhodes, 1998), berries

(Hakkinen, Karenlampi, Mykkanen, & Torronen, 2000) and onions (Price, Bacon, & Rhodes,

1997). Hakkinen et al. (2000) observed that domestic processing and storage (–20 °C for 3 – 9

months) decreased myrecitin and kaempferol more than quercetin. The ORAC values of black

raspberries remained stable after IQF and throughout 3 months storage and increased by 18% at

6 months. The ORAC values of berries canned in syrup remained stable during storage, while

values for berries canned-in-water remained stable up to 1 month of storage and increased by

32% and 27% after 3 and 6 months storage. The ORAC values of puree, non-clarified and

clarified juices remained stable over the 6 months storage (Hager, Howard, Prior, &

Brownmiller, 2008). Similar results were observed in the storage of processed blackberries

Page 63: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

44

(Hager, Howard, & Prior, 2008). During the storage (at 10 °C for 1 – 3 days) of irradiated (3 and

5 kGy) kale juice, the antioxidant capacity decreases in spite of increase in total phenolic content

(Song et al., 2006). However, the antioxidant activity and total phenolic content of carrots

increase during the same storage period. The lack of decrease in the antioxidant activity in

carrots is attributed to the presence of β-carotene and its synergistic effect with other antioxidants

to protect against oxidation. In contrast, the same investigators hypothesized that presence of

vitamin C in kale juice contributes more to its antioxidant activity, which decreases during

irradiation, than the irradiation-induced phenolic compounds (Wrona, Korytowski, Rózanowska,

Sarna, & Truscott, 2003; Yun-Zhong, Sheng, & Guoyao, 2002). The total flavonoids content

remained constant during storage in both air dried and fresh spinach stored in modified

atmosphere (Gil, Ferreres, & Tomas-Barberan, 1999). The ORAC and percent polymeric color

values of canned cherries increase after 5 months storage at 22 °C (Chaovanalikit & Wrolstad,

2004a). This suggests that polymeric compounds form during storage compensated for the loss

of antioxidant capacity due to degradation of monomeric anthocyanins. Friedman (1997)

observed greater polyphenol content at lower temperature storage of potatoes and attributed to

less polyphenol oxidase (PPO) activity at lower temperatures. The investigator further

hypothesized that enzymatic activity is indirectly related to monomeric polyphenol content,

which means greater the enzymatic activity, the greater the transformation of monomeric

polyphenols to polymeric ones. The higher PPO activity at the higher storage temperature may

account for the greater discoloration due to transformation of polyphenols to polymeric pigments

at that temperature. In another study, the DPPH radical scavenging ability in new mulberry wine

was 71% and increased after a year‘s storage to 78%. At the same time, the FRAP reducing

power decreased from 5720 m mol/L to 4630 m mol/L (Tsai, Huang, & Huang, 2004). These

Page 64: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

45

changes support the hypotheses of Friedman (1997) on the conversion of monomeric

anthocyanins to the copigmented and polymeric forms during storage. Light has affect on the

phenolic acid content during storage. Chlorogenic acid levels of dark-stored potatoes increases,

but less compared to light-stored potatoes (Friedman, 1997; Griffiths, Bain, & Dale, 1995).

Gamma irradiation increases 0.21 units µg TE/g FW in antioxidant activity (DPPH) for each

day in storage, whereas 0.17 units in antioxidant activity (DPPH) with 1 unit increase of dosage

in potato (Atlantic cv). Increase in the phenolic content with storage has reported to be 1.98 units

per day and 0.16 units for 1 Gy increase in radiation (Blessington et al., 2007). The gain in

phenolic content may be due to dehydration, leading to concentration of solids at the end of the

storage period. Additionally, stimulation of synthesis of both antioxidants and polyphenols is

known to occur with stress, which may have increased at the end of the storage period due to

dehydration (Friedman, 1997; Kang & Saltveit, 2002). For example, the activity of PAL

(phenylalanine ammonia lyase), which produces a precursor to phenolic compounds, has been

reported to increase under stressful conditions, and this is associated with the accumulation and

synthesis of phenolic compounds (Kang & Saltveit, 2002).

Processing or storage can promote or enhance the stability and antioxidant properties of food

because of its complex matrix with lots of components mixed together in aqueous and lipid

phase. The antioxidant activity differs depending on measurement duration and mode,

temperature, oxygenation and medium in the model systems used. The behavior of antioxidants

also depends on the mediums such as bulk lipid, emulsified and aqueous, which is called the

polar paradox. According to the polar paradox, polar antioxidants are more effective in bulk

lipids, whereas non-polar antioxidants are effective in emulsified media (Cuvelier, Bondet, &

Berset, 2000; Porter, Levasseur, & Henick, 1977). The overall antioxidant activity would depend

Page 65: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

46

on the redox reactions occurring between different natural antioxidants and lipid oxidation

products (Halliwell, Murcia, Chirico, & Aruoma, 1995; Mortensen & Skibsted, 1997; Namiki,

1990). For example, when a small amount of olive oil is mixed with tomato puree, ascorbic acid

decreased after a few hours of storage. This is because of the ability of ascorbate compounds to

reduce the radical forms compared to α-tocopherol contained in the lipid matrix (Nicoli et al.,

1999). The investigators hypothesized that the above mechanism is due to the lower standard

redox potential value of the ascorbyl radical forms of the α-tocopherol radical. The interaction

between the vegetable matrix and the lipid fraction becomes more evident when heated. Thus, in

addition to study the antioxidant properties of complex foods, it must be taken into account that

water-soluble antioxidants could protect lipid soluble antioxidants, because of the polar paradox

(Porter et al., 1977).

There is not a particular trend in the effect of non-thermal treatment and storage on the

antioxidant capacity of fruits and vegetables. During storage at 4 °C of treated orange juice (100

MPa/60 °C/5min, 350 MPa/30 °C/2.5 min, 400 MPa/40 °C/1 min), the antioxidant capacity was

not significantly different from freshly squeezed orange juice after 10 days, whereas a 17 – 23 %

decrease was reported after 40 days (Plaza, Sanchez-Moreno, Elez-Martinez et al., 2006;

Sanchez-Moreno et al., 2003). In contrast, no significant differences in antioxidant capacity

(ABTS) of treated orange juice were observed at more extreme conditions of 500 MPa and 800

MPa for 5 min at 20 °C and stored for 21 days at 4 °C (Fernandez-Garcia, Butz, Bognar et al.,

2001). In another study, there was no difference (P > 0.05) in the antioxidant capacity of

combined treated apple juice (600 MPa/60 °C/30 min) and non treated samples after 1 month

storage at 4 °C (Sanchez-Moreno et al., 2009). Antioxidant capacity of carrot and tomato juices

was unaltered after treatment (250 MPa/35 °C/15 min) and storage for 30 days at 4 °C (Dede,

Page 66: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

47

Alpas, & Bayindirli, 2007). On the contrary, the antioxidant capacity of a tomato product known

as ―gazpacho‖ in Spain was not affected after high pressure treatments (150MPa/60◦C/15 min

and 350 MPa/60◦C/15 min) but decreased 39 – 46% after storage for 40 days at 4 °C (Plaza,

Sanchez-Moreno, De Ancos et al., 2006). However, reports on high pressure treatments for

whole fruits and vegetables are limited. Doblado, Frias, & Vidal-Valverde (2007) observed HPP

had no effect on antioxidant activity for broccoli, whereas a decrease in carrots and an increase

in green beans were observed when treated at 400 MPa/25 °C/2 min.

2.1.9. Enzymatic/Chemical Oxidation

Although chemical or enzymatic oxidations have been widely proven to cause a progressive

decrease in polyphenol antioxidant properties, processing or prolonged storage times can

promote or enhance the rate oxidation of phenolic compounds depending on intrinsic properties

of the food matrix as well as on processing conditions such as water activity, pH, time,

temperature, and oxygen availability (Nicoli et al., 1999). For example, thermal processing and

storage at high temperature (40 °C for 4 weeks) significantly increase the antioxidant activity of

carrot puree depending on the concentration and oxidation of the phenolic acids (Talcott,

Howard, & Brenes, 2000a). Similarly, the antioxidant properties of pasteurized (105 °C for 20

min) and air-bottled tea extracts increased during 30 day storage at 20 °C (Manzocco, Anese, &

Nicoli, 1998). The same investigators further observed that green tea extracts showed higher

phenol content and chain-breaking activity than those obtained from black tea leaves.

Additionally, reduction in the oxygen-scavenging ability of tea extracts after pasteurization and

storage is correlated with a progressive increase in the redox potential value for black tea extracts

indicating the gain in radical scavenging activity is associated with a corresponding decrease in

the reducing power of the tea extracts. The modifications in chain-breaking activity and oxygen

Page 67: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

48

uptake detected in tea extracts is ascribable to the progressive oxidation of polyphenols, a

process leading to the formation of macromolecular compounds with stronger radical scavenging

power. Catechin when subjected to a progressive enzymatic oxidation shows a remarkable

increase in its chain breaking activity prior to the formation of brown macromolecular

compounds, whereas a subsequent loss in the antioxidant properties has found for advanced

enzymatic oxidation steps (Cheigh, Um, & Lee, 1995). Enzymatic oxidation of the polyphenol

fraction causes a decrease in the chain-breaking activity, whereas chemical oxidation has found

to have the opposite effect (Manzocco et al., 1998; Nicoli et al., 1999). The above mechanism is

explained as different pathways of enzymatic and chemical oxidation to polyphenols, which

leads to the formation of compounds having markedly contrasting radical scavenging capacities.

The investigators hypothesized that chemical proceeds much slower than enzymatic oxidation

and the compounds formed during pasteurization and storage would have an intermediate

oxidation status when compared to those formed by enzymatic oxidation. Since the oxidation

products of phenolic compounds still retain antioxidant activity (Yen, Chen, & Peng, 1997), the

increased stability of partially oxidized polyphenols gained in chain-breaking efficiency during

the processing of the beverages. Thus, polyphenols with an intermediate oxidation state exhibit

higher radical scavenging efficiency than non-oxidized ones. The higher antioxidant properties

of partially oxidized polyphenols is attributed to their increased ability to donate a hydrogen

atom from the aromatic hydroxyl group to a free radical and/or to the capacity of their aromatic

structures to support the unpaired electron through delocalization around the π- electron system

(Nicoli et al., 1999).

Page 68: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

49

3. ANTHOCYANINS

Anthocyanins are water-soluble pigments responsible for the orange red through deep purple

colors produced by chemical combination of its C6-C3-C6 structure with glycosides, acyl

groups, and other molecules in flowers, fruits and vegetables (Figure 2.10). Mostly,

anthocyanins occur as glycosides such as 3-monoglycosides and 3, 5-diglycosides in nature.

Table 2.4 shows common anthocyanins present in fruits, vegetable and grains. In addition, many

anthocyanins have acylations (i.e. ester bonds between sugars and organic acids) with coumaric,

caffeic, ferulic, p-hydroxybenzoic, synapic, malonic, acetic, succinic, oxalic, and malic acids

(Francis, 1989). Substitution of hydroxyl and methoxyl groups influences the color of the

anthocyanins. For example, more hydroxyl groups provide bluish color, whereas methoxyl

groups increase the redness of the anthocyanins. Anthocyanins are present in the vacuole of the

plant cell and the molecules are protected by unique mechanisms such as self-association and

copigmentation in the cell. Self-association is to form helical stacks through the hydrophobic

attraction and hydrogen bonding between the flavylium nuclei. The stacking protects the

chromophores behind the sugar groups from the hydration reaction (Hoshino, Matsumoto, &

Goto, 1981). Copigmentation occurs through hydrogen bonding of the phenolic groups between

anthocyanin and flavone molecules (Francis, 1989). Flavonols, amino acids, benzoic acids,

coumaric and cinnamic acids also reacts with anthocyanins due to interaction known as

intermolecular copigmentation. Intramolecular copigmentation occurs in the anthocyanins

because of the acylation by organic acids. Acylation of the molecule is believed to improve

anthocyanin stability by protecting it from hydration (Patras, Brunton, O'Donnell, & Tiwari,

2010). A spectral characteristic of glycosylation and acylation in potato anthocyanin is shown in

Figure 2.11.

Page 69: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

50

O+

OH

OH

OH

O G

Pelargonidin-3-glucoside

O+

OH

OH

OH

O G

OH

Cyanidin-3-glucoside

O+

OH

OH

OH

O G

OCH3

OCH3

Malvidin-3-glucoside

O+

OH

OH

OH

O G

OH

OH

Delphidin-3-glucoside

O+

OH

OH

OH

O G

OCH3

Petunidin-3-glucoside

O+

OH

OH

OH

O G

OCH3

Peonidin-3-glucoside

OH

Figure 2. 10. Common anthocyanins found in flowers, fruits and vegetables.

Antioxidant activity of anthocyanins in-vitro and in-vivo has been well documented (Prior,

2003). Antioxidant activity of the anthocyanins depends on the number of free hydroxyl groups

attached to the structure. For example, petunidin glucoside (Figure 2. 10) in purple potatoes has

high antioxidant activity compared to others. However, total antioxidant activity depends on the

amount of phenolic acid and anthocyanins in the tuber (Hamouz, Lachman, Vokal, & Pivec,

1999; Lachman, Hamouz, Orsak, & Pivec, 2000). The same investigators also reported that

purple /blue coloration of the purple potato is due to acylation of cinnamic acid to anthocyanidin

giving high antioxidant activities, whereas glycosidic attachment at position 3 and 5 reduces

antioxidant activities. In another study, Wang, Cao, & Prior (1997) reported that cyanidin

glycosides tend to have higher antioxidant capacity than peonidin or malvidin glycosides because

of the free hydroxyl groups on the 3‘ and 4‘ positions.

Anthocyanins have many applications for regulatory authorities to check adulteration as well

Page 70: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

51

as in food industries. For example, prune juice adulterated with other fruit juices will show

increased levels of anthocyanins compared to prune juice alone (Van Gorsel, Li, Kerbel, Smits,

& Kader, 1992). Adulteration of blackberry jams with strawberries can be detected with the

profile analysis of pelargonidin and cyanidin-3-glucoside (Garcıia-Viguera, Zafrilla, & Tomás-

Barberán, 1997). In red raspberry juices, it was found that under-processed samples had higher

levels of polymeric color instead of the monomeric anthocyanin pigments, which is considered

as a poor quality product (Altamirano, Drdak, Simon, Smelik, & Simko, 1992). Other important

applications include fungicidal properties of anthocyanins that might block potato blight from

reaching tubers underground (Lachman et al., 2005). Cultivation of blueberries (Camire,

Chaovanalikit, Dougherty, & Briggs, 2002), grape (Camire et al., 2002), red and purple fleshed

potatoes give alternative sources of natural colorants from their rich anthocyanins (Lachman et

al., 2005; Rodriguez-Saona, Giusti, & Wrolstad, 1998). Potato peel powders have high

antioxidant activities (Singh & Rajini, 2004) and extracts from peels could be used with oils, fats

and other food products to suppress lipid oxidation (Rehman, Habib, & Shah, 2004). However,

lack of comprehensive knowledge on various factors affecting the stability of anthocyanins limits

their use.

The factors affecting the stability and interrelationship between the quality and quantity of

the anthocyanins have been reported in plants or food systems (Figure 2.12) (Jackman et al.,

1987). The stability of anthocyanins has also been related to the presence of number of hydroxyl

groups and methoxyl groups attributed to the flavylium structure besides glycosylation (Francis,

1989). The investigator also reported that diglucosides are more stable than monoglucosides.

However, presence of more sugar molecules in the diglucosides accelerates browning in the food

matrix. In another thermal degradation of anthocyanin study, Tanchev & Yoncheva (1974)

Page 71: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

52

Table 2. 4. Common anthocyanins present in fruits and vegetables. Fruit/veg/grain Major anthocyanin Minor anthocyanins Reference

Blackberry Cyanidin-3-glucoside Cyanidin-3-rutinoside; Cyanidin-3-dioxalylglucoside;

Cyanidin-3-xyloside Cyanidin-3-malonylglucoside

Fan-Chiang & Wrolstad (2005);

Rommel, Wrolstad, & Heatherbell

(1992)

Strawberries Pelargonidin-3-glucoside Cyanidin-3-glucoside Jackman et al. (1987)

Litchi pericarp Cyanidin-3-rutinoside Malvidin-3-acetylglucoside;

Quercetin-3-rutinoside; Cyanidin glucoside; Quercetin

glucoside

Lee & Wicker (1991); Sarni-

Manchado, Le Roux, Le Guerneve,

Lozano, & Cheynier (2000)

Blood orange Cyanidin-3-glucoside Delphinidin-3,5-diglucoside; Cyanidin-3,5-diglucoside;

Delphinidin-3-glucoside; Peonidin-3,5-diglucoside; Cyanidin-

3-(acetyl)-glucoside; Cyanidin-3-(feruloyl)-glucoside

Krifi et al. (2000)

Acai Cyanidin-3-glucoside Pelargonidin-3-glycoside Del Pozo-Insfran et al. (2004)

Bilberry Cyanidin galactoside Delphidin arabinoside; Cyanidin glucoside; Delphidin

galactoside; Delphidin glucoside; Petunidin glucoside;

Malvidin galactoside; Cyanidin arabinoside; Peonidin

glucoside; Malvidin arabinoside

Yue & Xu (2008)

Elderberry Cyanidin-3-sambubioside Cyanidin-3-glucoside;

Cyanidin 3,5-diglucoside;

Cyanidin-3-sambubioside-5-glucoside

Stintzing, Stintzing, Carle, Frei, &

Wrolstad (2002)

Red radish Pelargonidin-3-sophoroside-5-

glucoside

Matsufuji et al. (2007)

Black carrot Cyanidin-3-xylosylgalactoside Cyanidin-3-xylosylglucosyl-galactoside Stintzing et al. (2002)

Cherry Cyanidin-3- glucosyl-rutinoside Cyanindin-3-glucoside; Cyanidin-3-rutinoside; Peonidin-3-

rutinoside

Kim & Padilla-Zakour (2004)

Purple carrot Cyanidin 3-(sinapoylxylosyl-

glucosylgalactoside)

Malvidin 3-monoglucoside;

Peonidin 3-monoglucoside

Glassgen, Wray, Strack, Metzger,

& Seitz (1992)

Red cabbage Cyanidin-3-sophoroside-5-

glucoside

Cyanidin-3,5-diglucoside Giusti, Rodriguez-Saona, Griffin,

& Wrolstad (1999)

Purple-fleshed

potato

Petunidin-3-(p-coumaroyl-

rutinoside)-5-glucoside

Malvidin-3-(p-coumaroyl-rutinoside)-5-glucoside Lewis, Walker, Lancaster, &

Sutton (1998)

Red-fleshed

potato

Pelagonidin-3-(p-coumaroyl-

rutinoside)-5-glucoside

Peonidin-3-(p-coumaroyl-

rutinoside)-5-glucoside

Lewis et al. (1998)

Page 72: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

53

Figure 2. 11. Spectral characteristics of potato anthocyanins, indication of glycosylation and

acylation patterns. Modified from Rodriguez-Saona, Giusti, & Wrolstad (1998).

reported that the existence of three hydroxyl groups leads to more rapid degradation of

delphidin-3-rutinoside than malvidin-3-glucoside. Sadilova, Stintzing, & Carle (2006) reported

that methoxylation of the acyl moiety improves the structural integrity towards heat. The

equilibrium between the colored cationic form and colorless pseudobase and degradation is

directly influenced by pH (Figure 2.13). Anthocyanins are stable at pH (< 4) and become

colorless at high pH (Francis, 1989). In slightly alkaline solutions (pH 8 to 10) highly colored

ionized anhydro bases are formed. At pH 12, these ionized anhydro bases hydrolyze rapidly to

fully ionized chalcones (Bridle & Timberlake, 1997). Acylated anthocyanins retain more color at

Page 73: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

54

the higher pH than non-acylated anthocyanins in addition to their resistance to heat, light and

SO2. Giusti & Wrolstad (1996a) observed that the C-5 acylation of anthocyanins in red radish

provide high stability to the compounds. In another study on Vitis vinifera wines, substitution of

vitisins at C-4 position of anthocyanin structure provided resistance to color loss with SO2, at

higher pH values (Bakker & Timberlake, 1997). Acylation of anthocyanins with ferulic, sinapic

Figure 2. 12. Interrelationships between anthocyanin quantity and quality, and various factors

affecting the stability of anthocyanins. Adapted from Jackman et al.(1987).

and coumaric acids in black carrots prolongs their half-life compared with non-acylated

derivatives from strawberry and elderberry isolates (Sadilova, Carle, & Stintzing, 2007).

Acylation has not been shown to increase the stability of anthocyanins in some model systems.

For example, anthocyanins from grape, red cabbage and Ajuga reptans in pH 3.5 citrate buffer

solutions has not showed any difference in their degradation rate in spite of their different

Quantity of

anthocyanin

pigment

Quality of anthocyanin

B-ring structrure

Glycosidic substitution

Acylation

Hue

pKA

Rate of reactivity

Reaction with other

compounds

Leucoanthocyanins

Flavonols

Acetaldehyde

Acid

Polyphenoloxidase

Ascorbic acid

Proteins

Oxygen

Hydrogen peroxide

Metal ions

Sulfur dioxide

Browning reactions

Maillard

Enzymatic

Ascorbic acid

Page 74: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

55

+H+ OCH3

O+

OH

OH

OH

O G

OCH3

OCH3

OCH3

O+

OH

OH

OH

OH

OCH3

O

OH

OH

OH

OH

OH

OCH3

OCH3

O

OH

OH

O G

OCH3

O

OCH3

OCH3

O

O

OH

OH

O

OH

OH

OCH3

OCH3

OH

O

OH

OH

OHO

O

H2OH2O

OCH3

OH

HOOC

OCH3

OH

OH OH

CHO +

PhloroglucinaldehydeProtocatechuic acid

Malvidin-3-glucoside (blue)Malvidin-3-glucoside (red)

Quinoidal base or Anhydrobase

Deglycosylation

Malvidin

- Diketone

Figure 2. 13. Effect of pH on the possible degradation mechanism of anthocyanin (Malvidin-3-

glucoside). Adapted from Wong (1989).

Page 75: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

56

chemical configurations (Baublis, Spomer, & Berberjimenez, 1994). Grape anthocyanins include

mono and diglucosides of five different monoacylated aglycones (Lea, 1988), red cabbage

contains diacylated triglucosides of cyanidin (Mazza & Miniati, 1993), and Ajuga reptans

contains glucosylated cyanidin, acylated with p-hydroxycinnamic acid, ferulic acid, and malonic

acid (Callebaut, Hendrickx, Voets, & Motte, 1990). Similar findings have been reported on the

acylated aglycones of pelargonidin from strawberry juice or concentrate (Garzon & Wrolstad,

2002) and a model system (Garzón & Wrolstad, 2001).

3.1. Effect of processing on Anthocyanins

Processing can change the anthocyanin content and color in fruits and vegetables. The overall

color changes may be due to a number of factors such as pigment degradation, pigment

polymerization, reactions with other components of the formulation, nonenzymatic browning,

oxidation of tannins, and other reactions completely unrelated to the added colorant (Francis,

1989).

3.1.1. Extraction

Solvents to be used for extraction of anthocyanin pigments from fruits and vegetables

without disturbing the acylation, structure of anthocyanin and interfering with unwanted

compounds are reviewed in the literature (Jackman et al., 1987). Extraction procedures have

generally involved the use of acidic solvents, which denature the membranes of cell tissue and

simultaneously dissolve the pigments followed by ether precipitation. Acidified methanol

(HCl/methanol) has been the most commonly used extraction solvent, where HCl maintains low

pH for favoring the formation of flavylium chloride salt and methanol allows easier

concentration of extracted material. Lipid materials and unwanted polyphenols can be separated

Page 76: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

57

by using petroleum ether, diethyl ether or ethyl acetate. Many researchers have suggested using

alcohol with very low acid concentration or without acid to minimize the possible changes to the

acylated pigments during extraction (Jackman et al., 1987). In addition, use of neutral solvents

such as 60 % methanol, acetone/methanol/water mixtures, cold acetone, n-butanol or boiling

water has also been recommended (Jackman et al., 1987). However, application of external

technologies in addition to solvents for extraction of anthocyanin pigments has not been

discussed in detail. Optimal extraction of anthocyanins from grape skin was reported to be

carried out by formic acid/methanol solvent (5/95, v/v), with ultrasonication for 10 min,

extraction temperature at 25 °C, and extraction time for 1.5 hr (Li, Pan, Cui, & Duan, 2010). In

another study, yield of anthocyanins from red cabbage using high pressure CO2 (HPCD)

increased compared to conventional acidified water (CAW) (Xu et al., 2010). Fractionated high

pressure extractions using CO2 supercritical fluid extraction followed by CO2/ethanol-H2O

mixtures (1 – 100%, v/v) at 313 K and 20 MPa extracted higher anthocyanin contents from dry

and raw elderberry pomace. Presence of water both in the raw material and in the solvent

mixture accelerates higher extraction of anthocyanin from elderberry pomace (Seabra, Braga,

Batista, & de Sousa, 2010).

3.1.2. Blanching

Anthocyanins are degraded by a number of enzymes found in plant tissue such as

glycosidases, polyphenoloxidases (PPO), and peroxidases. Glycosidases produce anthocyanidins

and sugars, and anthocyanidins are very unstable and rapidly degraded. PPO catalyzes the

oxidation of o-dihydrophenols to o-quinones that further react to brown polymers (Figure 2.14).

Kader, Irmouli, Zitouni, Nicolas, & Metche (1999) proposed that Cyanidin 3-glucoside (o-

diphenolic) is degraded by a mechanism of coupled oxidation involving the enzymatically

Page 77: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

58

generated o-quinone with partial regeneration of the o-diphenolic co-substrate confirming the

role of PPO in anthocyanin degradation. In another study, pelargonidin-3-glucoside, were

degraded by a mechanism involving a reaction between the o-quinone and/or secondary products

of oxidation formed from the quinone and the anthocyanin pigment (Kader, Irmouli, Nicolas, &

Metche, 2001). The destruction of anthocyanins by enzymatic activity could be important in the

design of an extraction procedure and perhaps in the final formulation in a food (Francis, 1989;

Rossi et al., 2003). Blanching of fruits and vegetables inactivate enzymes and improves the color

by retaining anthocyanins in the processed foods. Addition of 30 ppm SO2 inhibits phenolase

activity in sour cherry juice (Cemeroglu, Velioglu, & Isik, 1994). Juice prepared from blanched

blueberry-pulp extract degrades anthocyanins, whereas unblanched extract cause 50% loss of

anthocyanins (Skrede, Wrolstad, & Durst, 2000). Similarly, blanching (steam for 3 min) of

blueberry fruits induced higher anthocyanin retention i.e. 23% instead of 12% (unblanched)

when processed into juice (Rossi et al., 2003). Anthocyanin content of blanched (98 °C for 2

min) purple carrots increase 27% compared with the fresh sample (Uyan et al., 2004). Increase in

the anthocyanin retention in fruits is attributed to two main factors; (1) reduction of enzyme

mediated anthocyanin degradation i.e. complete inactivation of native PPO, and (2) greater

extraction yield linked to the increase of fruit skin permeability caused by the heat treatment

(Kalt, McDonald, & Donner, 2000). Introduction of the blanching step has also a positive effect

on the recovery of individual anthocyanin. For example, in blueberry juice processing, the

percent recovery increase by 71 – 2672 % for monoglucosides of cyanidin, malvidin, petunidin,

peonidin and delphidin (Rossi et al., 2003). Heat inactivation by blanching (boiling water for 10

sec) retain or increase anthocyanin content of apple peels after oven-drying (Wolfe & Liu, 2003).

Similarly, steam blanching of purple- and red-fleshed potatoes reduces the peroxidase activity

Page 78: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

59

OH

OH

Enzyme

O2 + OH2 +

O

O

O-diphenol O-benzoquinone

O

O

O-benzoquinone

+ AnthocyaninOxidized

anthocyanin+

Degradation

Product

Figure 2. 14. Enzymatic activity on the polyphenol. Adapted from Wong (1989).

by 98 – 99% to retain the anthocyanins (Reyes & Cisneros-Zevallos, 2007). Interestingly, heat

and SO2 treatments of high bush blueberries increase the recovery of anthocyanins but not of

other polyphenols in the juices (Lee, Durst, & Wrolstad, 2002). Blanching (95 °C for 3 min) in

combination with pasteurization during preparation of blueberry purees reduce 43% total

monomeric anthocyanins compared to fresh fruit (Brownmiller, Howard, & Prior, 2008). Volden

et al. (2008) observed 59% loss in anthocyanin content of red cabbage after blanching.

3.1.3. Heating

At high temperatures, the structure of anthocyanin is opened to form chalcone, which is

degraded further to brown products (Francis, 1989). However, it has been observed that optimal

conditions permit the regaining of color on cooling if there is sufficient time (several hours) for

the reconversion. Processing of black raspberries in canned-in-water or syrup (87.8 – 93.3 °C for

4 min) losses 42 and 51% of total anthocyanin, respectively (Hager, A. et al., 2008), whereas

blackberry products losses 17.8 and 10.5%, respectively, in similar condition (Hager, T. J. et al.,

2008). In another study, only 50% of the total anthocyanin content of elderberry was lost after 3

Page 79: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

60

h of heating 95 °C (Sadilova et al., 2006). Degradation of anthocyanins occurs in strawberry

(Garzon & Wrolstad, 2002), raspberry (Kim & Padilla-Zakour, 2004) and sour cherry

(Cemeroglu et al., 1994) as the fruits are processed into juice, concentrate or jam and continues

during storage. The investigators also reported that the degradation of anthocyanins in

concentrates was greater compared to juices. Patras, Brunton, Da Pieve, & Butler (2009)

reported significant loss (P < 0.05) in anthocyanin content of blackberry (3%) and strawberry

(28%) puree processed at 70 °C for 2 min. Anthocyanin degradation in processed berry products

is attributed to indirect oxidation by phenolic quinones generated by polyphenol oxidase and

peroxidase (Kader, Rovel, Girardin, & Metche, 1997; Skrede et al., 2000). In contrast,

anthocyanins in Bing cherries increase slightly after canning (100 °C for 12 min) (Chaovanalikit

& Wrolstad, 2004b). The increase in the anthocyanin content is attributed to the increased

extraction efficiency in the softened fruits. Increase in membrane permeability at high

temperatures in the macerated peel tissue facilitate phenolic extraction (Spanos, Wrolstad, &

Heatherbell, 1990) and release of bound phenolic compounds by breakdown of the cellular

constituents was also reported in the literature (Dewanto, Wu, Adom et al., 2002).

Boiling and steaming cause 41% and 29% losses in anthocyanin content of red cabbage,

respectively (Volden et al., 2008). Similar losses were reported in blueberry products (Lee et al.,

2002). In contrast, Kirca, Ozkan, & Cemeroglu (2007) reported that anthocyanins from black

carrots were reasonably stable during heating at 70 – 80 °C due to di-acylation of anthocyanin

structure. According to the same investigators, the stability of monomeric anthocyanins in black

carrot juice and concentrates also depends on temperature, solids content and pH. Similarly,

cyanidin-3-rutinoside has the highest stability to the effect of thermal treatment at 95 °C in

blackcurrants (Rubinskiene, Viskelis, Jasutiene, Viskeliene, & Bobinas, 2005). Greater thermal

Page 80: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

61

stability (25 – 80 °C) of anthocyanins present in red cabbage compared to blackcurrants, grape

skins and elderberries in a soft drink model system is attributed to the protection of flavylium

system through co-pigmentation (Dyrby, Westergaard, & Stapelfeldt, 2001). In another study,

purple-fleshed potato and grape extracts showed lower color stability than red-fleshed potatoes

and purple carrots at 98 °C (Reyes & Cisneros-Zevallos, 2007). Maccarone, Maccarrone, &

Rapisarda (1985) studied the stabilization of anthocyanins in blood orange juice and found that

microwave pasteurization and addition of tartaric acid and glutathione improved the stability.

The investigators reported that complexation of anthocyanins with rutin and caffeic acid

provided the highest stability.

Extrusion cooking of corn meal with grape juice and blueberry concentrate at a die

temperature of 130 °C degrades 74% of anthocyanin compared to the unheated formulation

(Camire et al., 2002). Similarly, extrusion cooking of corn meal with dehydrated fruit powder

from blueberry, cranberry, concord grape and raspberry at the similar die temperature reduce up

to 90% of anthocyanin content for all the dehydrated fruits used except raspberries (Camire et

al., 2007). There is no difference (P > 0.05) observed in the total anthocyanin content of purple

wheat bran heated at 177 °C for 20 min compared to unheated bran (Li et al., 2007). However,

baking of muffins prepared from purple wheat bran with other ingredients at a similar

temperature for 7 – 12 min has not retained any anthocyanin.

3.1.4. Drying/Dehydration

Depending on the severity of heat application, dehydration reduces moisture content of fruits

and vegetables, which is important to maintaining the equilibrium media for stability of

anthocyanins. The rate of evaporation of water might have influence on the water soluble

anthocyanin pigments. Air-drying and freeze-drying increase the anthocyanin content of apple

Page 81: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

62

peels, whereas oven-drying has similar amount compared to fresh peels (Wolfe & Liu, 2003).

Interestingly, the blanched freeze dried peels contain more anthocyanins than the fresh apple

peels in the same study (Wolfe & Liu, 2003). The retention of total anthocyanins in Saskatoon

berries is higher (~ 30%) when dried by combining the air with microwave vacuum method than

with air-drying (15%) compare to fresh berries (Kwok et al., 2004). Better retention of

anthocyanin in berries is attributed to the reduced time of microwave vacuum-drying. The same

investigators reported that the combination of heat and atmospheric oxygen in air-drying favored

enzymatic browning activity of polyphenol oxidase, whereas less enzymatic browning was

observed for vacuum microwave-drying, because of reduced heat and oxygen exposure. Similar

results were observed during drying of cranberry using microwave vacuum, freeze-drying and

air-drying methods. Microwave vacuum-drying and freeze-drying produce dehydrated products

that contain a relatively higher level of total anthocyanins per gram dry solid compared to air-

drying (Leusink et al., 2010). Uyan et al. (2004) observed a 2-fold increase in anthocyanin

content during hot air dehydration of purple carrots, whereas combining microwave and hot air-

drying increased it 1.75-fold. However, jam processing (105 °C) reduces (P < 0.05) 21 – 89%

anthocyanin contents of cherries, plums and raspberries (Kim & Padilla-Zakour, 2004). Ersus &

Yurdagel (2007) studied microencapsulation of black carrot anthocyanins by spray-drying

(drying air inlet: 160 – 200 °C; outlet: 107 – 131 °C) using different maltodextrins as carrier and

coating agents. The investigators observed an increase in anthocyanin contents of powders (7.9 –

36.83%) with different maltodextrins at a constant air inlet with varying outlet temperatures,

whereas higher inlet temperatures (>160 – 180 °C) caused more anthocyanin losses.

Page 82: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

63

3.1.5. CO2 treatment

Reduction in red color intensity of the internal tissue in some fruits (Kader, 1986) and change

in external color from red to red-purple in strawberries with CO2 treatment (Ke, Goldstein,

Omahony, & Kader, 1991) has been reported. No significant difference (P > 0.05) in external

anthocyanin content of fresh strawberries stored under CO2 treated atmosphere compared to the

initial fruit, whereas internal anthocyanin content decreases significantly (P < 0.05)(Gil,

Holcroft, & Kader, 1997). The discrepancy of opposite behavior of anthocyanin in external and

internal tissues is attributed to higher concentrations of cyanidin 3-glucoside in external tissue

providing higher color stability, while pelargonidin glycosides are present in the internal tissue.

In addition, the accumulation of phenolic compounds was greater in external tissue which

provides more stability to the color.

3.1.6. Addition of external ingredients

Methods including addition of external ingredients have been proposed to retain

anthocyanins to reach higher color intensity. In frozen strawberries, it is shown that sugar

addition has a stabilizing effect on the total monomeric anthocyanins, enhancing the shelf life of

colored products (Wrolstad, Skrede, Lea, & Enersen, 1990). Similarly, litchi is a tropical fruit

and within 2 to 3 days after harvest its pericarp becomes desiccated and turns brown. To reduce

the color degradation, litchis are coated with chitosan (1.0 to 2.0%) and stored (4 ºC/90% relative

humidity). The use of chitosan delay changes of contents of anthocyanin, flavonoid, and total

phenolics. Interestingly, the activities of polyphenol oxidase and peroxidase, which have been

involved in anthocyanin degradation, are inhibited (Zhang & Quantick, 1997). The same

investigators also suggested that a plastic coating forms a protective barrier on the surface of the

Page 83: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

64

fruit and reduces the supply of oxygen for enzymatic oxidation of phenolics.

3.1.6.1. Effect of light

In general, it is considered that light has deleterious effects on anthocyanin stability and

exposure of natural colored beverages to light must be avoided. It has been reported that

presence of other flavonoids, flavone, isoflavone, and aurone sulfonates increase the

photostability of anthocyanins (Francis, 1989). On the other hand, light intensity has a profound

effect on apple color, because the light-exposed peel contains twice as much anthocyanin as a

shaded peel (Ju, Liu, & Yuan, 1995). Results similar to that of apple peel was observed in

anthocyanin extract heated at 43 °C for 160 min in a model system (Kearsley & Rodriguez,

1981). Interestingly, there is no difference (P > 0.05) in the color changes of raspberries, sweet

and sour cherries during light or dark storage, although the reaction rate of color degradation in

light is higher than that in dark storage (Ochoa, Kesseler, De Michelis, Mugridge, & Chaves,

2001). Light intensity has no affect (P > 0.05) on the stability of freeze dried elderberry

anthocyanin stored at different water activities (BrØnnum-Hansen & Flink, 1985).

3.1.6.2. Effect of oxygen and metal

Oxygen has a negative effect on anthocyanin stability (Markakis, 1982) depending on the

media conditions. Additionally, it has been suggested that flavonoids act as free radical

scavengers to protect anthocyanin molecules (Sarma, Sreelakshmi, & Sharma, 1997). The

presence of oxygen can accelerate the degradation of anthocyanins either through a direct

oxidative mechanism and/or through the action of oxidizing enzymes (Jackman et al., 1987). In

contrast, the presence or absence of oxygen has no influence on the stability of freeze dried

elderberry anthocyanins during storage at lower moisture content (BrØnnum-Hansen & Flink,

Page 84: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

65

1985). Anthocyanins are very reactive toward metals, and they form stable complexes with tin,

copper, and iron; it has been proposed that metal complexes could be used as colorants (Sarma et

al., 1997).

3.1.7. Non-thermal processing

Conventional heat processing of fruits and vegetables remains the most widely accepted

technology for food safety and shelf-life. However, in the last decade non-thermal technologies

such as high hydrostatic pressure, pulsed electric field and ultrasound have emerged as

alternative techniques for minimizing the degradation of anthocyanin content of fruit juices. For

example, more than 80% anthocyanin content of strawberry juice processed with high intensity

pulsed electric field (HIPEF) is retained (Odriozola-Serrano, Soliva-Fortuny, & Martin-Belloso,

2009a). However, Zhang et al. (2007) observed that processing of cyanidin-3-glucoside in

methanolic solution by pulsed electric field (at 1.2, 2.2 and 3.0 kV/cm, 300 numbers of pulses,

temperature ≤ 47 °C) degraded anthocyanin to colorless chalcones. It has been well documented

that electric field strength, pulse width, pulse frequency, pulse polarity, treatment time or pulse

shape are among the most important HIPEF processing parameters affecting microbial,

enzymatic inactivation, antioxidant activity and level of anthocyanins (Elez-Martínez & Martín-

Belloso, 2007; Marsellés-Fontanet & Martín-Belloso, 2007; Odriozola-Serrano et al., 2009a).

Odriozola-Serrano et al. (2009a) reported higher anthocyanin content of strawberry juice

submitted to bipolar pulses than in monopolar mode. Zabetakis, Leclerc, & Kajda (2000)

reported high hydrostatic pressure processing (200 – 800 MPa for 15 min) had minimal effect on

anthocyanin content of strawberry juice. Cano, Hernandez, & De Ancos (1997) reported that

peroxidase activity was decreased with high pressure up to 300 MPa for a treatment carried out

under 20 °C for 15 min, whereas above 300 MPa the activity increased slightly at this

Page 85: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

66

temperature. However, complete loss of enzymatic activity was achieved at 900 MPa. The same

investigators also observed strongly diminishing polyphenol oxidase activity with high-pressure

treatments up to 400 MPa. On the contrary, β-glucosidase has highest activity after a high-

pressure treatment at 400 MPa, and the activity decreased as the pressure treatment increased up

to 800 MPa (Zabetakis, Koulentianos, Orruño, & Boyes, 2000). Application of sonication using

ultrasound decreases anthocyanin content of blackberry juice by 5% (Tiwari, O'Donnell, &

Cullen, 2009). Degradation of compounds responsible for color and anthocyanins during

ultrasound processing is attributed to (i) oxidation reaction that is promoted by the interaction

with free radicals formed during sonication (Portenlanger & Heusinger, 1992); (ii) extreme

physical conditions occur during sonication (temperature up to 5000 K and pressure up to 500

MPa at micro-scale) (Suslick, 1988); and (iii) sonochemical reactions including generation of

free radicals, enhancement of polymerization/depolymerization reactions and other reactions

(Floros & Liang, 1994).

3.1.8. Storage

Proper storage of raw and/or processed foods is important for increasing shelf-life without

altering its color attributes, nutritional value and functionality. Formation of new anthocyanins

by the reaction of malvidin 3- monoglucoside and procyanidin B2 in the presence of

acetaldehyde (15 ºC for 4 months) has observed in a model system imitating wine. Three new

pigments have observed and their visible spectra show a bathochromic shift in relation to the

anthocyanin by the condensation of these compounds. Interestingly, these compounds have

improved stability (Francia-Aricha, Guerra, Rivas-Gonzalo, & Santos-Buelga, 1997).

Individual quick freezing (–70 °C/12 h) and storage (–20 °C/6 months) of black raspberries

increase the total anthocyanins (Hager, A. et al., 2008). Retention of anthocyanins in red

Page 86: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

67

raspberries has also observed during frozen storage (Mullen et al., 2002). The retention or

increase in the anthocyanin contents is attributed to moisture loss and enhanced extraction of

anthocyanins due to tissue softening. However, anthocyanin content of raspberries declines

linearly during storage (25 °C/6 months) with losses of 76% and 75% in canned-in water or

syrup, respectively (Hager, A. et al., 2008). Similarly, 65.8, 60.6, and 58.4% losses have

observed in blackberry canned-in-syrup products, canned-in-water products, and purees,

respectively, stored at 25°C for 6 months. In another study, total anthocyanins in strawberries

canned in 20 °Brix syrup declined 69% over 2 months at room temperature (Ngo, Wrolstad, &

Zhao, 2007). Residual polyphenoloxidase (PPO) activity is hypothesized to cause a decrease in

color characteristics in partially processed peach puree during storage (Bian, Gonzalez, &

Asleage, 1994), while nonenzymatic browning in pasteurized peach purees has attributed to

reducing the extent of sugar degradation and HMF formation (Garza, Ibarz, Pagan, & Giner,

1999). During storage of canned products, pelargonidin- 3-glucoside, the main anthocyanin in

strawberries, is hydrolyzed by acid to pelargonidin and further breaks into hydroxybenzoic acid

(Stintzing & Carle, 2004). Losses of anthocyanins and increase in percent polymeric color in

fruits are attributed to residual enzyme activity and/or condensation reactions of anthocyanins

with other phenolics (Brownmiller et al., 2008; Chaovanalikit & Wrolstad, 2004b; Hager, A. et

al., 2008; Hager, T. J. et al., 2008; Ngo et al., 2007). Ersus & Yurdagel (2007) observed 33 and

11% loss of anthocyanins in encapsulated black carrot anthocyanin powders with different

maltodextrins at 25 and 4 °C, respectively stored for 64 days. In contrast, phenolic acids such as

ferulic and syringic acid have also been shown to complex with anthocyanins in strawberry and

raspberry juices to improve the color (Rein & Heinonen, 2004). Condensation reactions have

been reported by the reaction of acetaldehyde, anthocyanins, and flavan-3-ols, producing the

Page 87: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

68

increment of color (up to seven times). It is believed that acetaldehyde forms a bridge between

the two flavonoids, and consequently condensation reactions could proceed and contribute to the

polymeric color (Francis, 1989).

Higher stability of anthocyanins can be achieved by using lower temperature and short-time

heating during processing and storage (Krifi et al., 2000; Rodriguez-Saona, Giusti, & Wrolstad,

1999). For example, cyanidin and delphinidin-rutinosides in blood orange juice are the most

stable anthocyanins found during storage for 12 months at -18 °C, whereas transformations at 4

°C under nitrogen storage induced a slow degradation process of anthocyanins (Krifi et al.,

2000). The investigators proposed that the condensation reactions in the presence of carbonyl

derivatives formed by sugar and ascorbic acid degradations under acidic conditions are

responsible for the transformation of anthocyanins into high molecular weight brown

compounds.

In model systems of red radish and red-fleshed potato anthocyanins, radish extracts have

higher stability during storage (2 °C and 25 °C for 65 weeks) than potato extracts (Rodriguez-

Saona et al., 1999). The presence of diacylation in red radish anthocyanin as compared to

monoacylated anthocyanins in red-potatoes is responsible for its enhanced stability. Diacylated

anthocyanins are stabilized by a sandwich-type stacking caused by hydrophobic interactions

between the planar aromatic residues of the acyl groups and the positively charged pyrylium

nucleus, which prevents the addition of nucleophiles such as water to the C2 and C4 position of

the anthocyanin, reducing the formation of pseudobase (Brouillard, 1981; Goto & Kondo, 1991).

In case of monoacylated anthocyanins, only one side of the pyrylium ring can be protected

against the nucleophilic attack (Brouillard, 1983).

Page 88: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

69

Water activity is another important factor influencing the stability of anthocyanins during

storage. Markaris, Livingston, & Fellers (1957) observed that strawberry anthocyanins are stable

when stored in dry crystalline form or on dry paper chromatograms. Bronnum-Hansen & Flink

(1985) observed that at water activity (aw) ≤ 0.31, water uptake had no effect on the anthocyanin

extracts from freeze dried elderberries, whereas at aw ≥ 0.5, a significant increase in anthocyanin

degradation rate during storage occurred. The investigators also reported that storage at high

temperature (50 °C) and high water activity (0.5 aw) had a most pronounced effect on the

stability of elderberry anthocyanins i.e. half-life was 2 months. Thakur & Arya (1989) found

that an increase in aw produced loss of grape anthocyanins adsorbed onto microcrystalline

cellulose. In contrast, degradation of strawberry anthocyanin increases with increasing aw in fruit

and in a model system (Erlandson & Wrolstad, 1972; Garzón & Wrolstad, 2001). In another

study, anthocyanin extract in a model system heated at 43 °C for 160 min was relatively stable at

different water activities (Kearsley & Rodriguez, 1981).

3.2. Degradation mechanism and products

Degradation of phenolic compounds is primarily caused by oxidation, cleavage of covalent

bonds or enhanced oxidation reactions due to thermal processing. There are two major pathways

for the degradation of anthocyanins (Markakis & Jurd, 1974). The first proceeds through the

carbinol pseudobase to the chalcone and coumarin glycoside, whereas second pathway involves

hydrolysis of the glycosidic bond as the first step in anthocyanin degradation to form the

anthocyanidin. The aglycon, which is more unstable than its glycosides, proceeds through a

highly unstable α-diketone intermediate to form aldehydes and benzoic acid derivatives. At first,

Markakis et al. (1957) hypothesized that opening of the heterocycle pyrylium ring and chalcone

formation is a first step of anthocyanin degradation and heating shifts the equilibrium toward the

Page 89: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

70

chalcone while the chalcone-flavylium reversion is very slow. Hrazdina (1971) reported that

heating decomposed anthocyanin into a chalcone structure, the latter being further transformed

into a coumarin glucoside derivative with a loss of the B-ring. Adams (1973) proposed

hydrolysis of sugar moiety and aglycone formation as initial degradation step possibly due to the

formation of cyclic- adducts. Tanchev & Ioncheva (1976) identified quercetin,

phloroglucinaldehyde and protocatechuic acid in addition to four other compounds by paper

chromatography following thermal degradation of anthocyanins.

Several studies have been reported on the thermal degradation of various anthocyanins and

their degradation products (Jackman & Smith, 1992; Patras et al., 2010; Sadilova et al., 2007;

Sadilova et al., 2006; Seeram, Bourquin, & Nair, 2001). Von Elbe & Schwartz (1996) suggested

that coumarin 3, 5-diglycosides are common thermal degradation products of anthocyanin 3, 5-

diglycosides. In another study on thermal degradation of strawberries, elderberries and black

carrots at pH 1, deglycosylation was proposed as the first step of anthocyanin degradation into

their respective aglycones yielding a phenolic acid (protocatechuic acid) and a phenolic aldehyde

(phloroglucinaldehyde) (Sadilova et al., 2006). However, the same investigator reported that

opening of the pyrylium ring and chalcone glycoside formation was the first step rather than

deglycosylation during the thermal degradation at pH 3.5 for the same anthocyanins (Sadilova et

al., 2007). A proposed mechanism of thermal degradation of acylated and non-acylated

anthocyanin is shown in Figure 2.15. High temperatures in combination with high pH caused

degradation of cherry anthocyanins resulting in three different benzoic acid derivatives (Seeram

et al., 2001). Sarni, Fulcrand, Souillol, Souquet, & Cheynier (1995) studied oxidative

degradation of o-diphenolic cyanidin-3-glucoside and non-o-diphenolic malvidin-3-glucoside in

the presence of caffeoyltartaric acid and grape polyphenoloxidase in model solutions. The same

Page 90: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

71

OH

OH

O+

O

OH

OH

glc

A C

B

Cyanidin-3-glucoside

Deglycosylation

OH

OH

O+

OH

OH

OH

A C

B

Cyanidin

COOH

OH

OH

OH

CHO

OHOH

ClevageC

leva

ge

OH

OH

O+

O

OH

glc

A C

B

Pelargonidin-3-glucoside

Deglycosylation

OH

OH

O+

OH

OH

Pelargonidin

Cle

vage

Cle

vage

A C

B

AB COOH

OH

OH

B

Cle

vage

4-Hydroxybenzoic acid

PhloroglucinaldehydeProtocatechuic acid

OH

OH

O+

O

OH

OH

gal glc coum

OH

OH

O+

OH

OH

OH

Cyanidin

Clevage

Cle

vage

Cle

vage

ClevageCoum-gal-glu

Deglycosylation

O

OH

OH

Cyanidin-3-gal-glc-coum

Coumaric acid

Figure 2. 15. Possible thermal degradation of non-acylated (Cyanidin-3-glucoside and

Pelargonidin-3-glucoside) and acylated (cyanidin glucosides with coumaric acid) anthocyanins.

Modified from Sadilova et al. (2007; 2006).

Page 91: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

72

investigators reported that both the anthocyanins reacted with the enzymatically generated

caffeoyltartaric acid o-quinone. They also indicated that cyanidin-3- glucoside was degraded

mostly by coupled oxidation, whereas malvidin-3-glucoside formed adducts with caffeoyltartaric

acid quinone. Jam making of cherries (Kim & Padilla-Zakour, 2004) and plums (Donovan,

Meyer, & Waterhouse, 1998) generated 5-(hydroxymethyl) furfural as a Maillard reaction

product.

3.3. Degradation kinetics of anthocyanins

Over the past few decades, there has been an increased concern for nutritional values of

foods in addition to preservation and microbiological safety. Application of kinetic models in

thermal processing of foods is important to assessing and predicting the influence of

operations/processing on critical quality parameters to minimize the undesirable changes and to

optimize quality of specific foods. Thermal degradation of anthocyanins has been studied for

various fruits such as strawberries (Garzón & Wrolstad, 2001; Markaris et al., 1957), raspberries

(Ochoa et al., 2001; Tanchev, 1972), Concord grapes (Calvi & Francis, 1978), plums (Ahmed,

Shivhare, & Raghavan, 2004), pomegranates (Marti, Perez-Vicente, & Garcia-Viguera, 2002),

sour cherries (Cemeroglu et al., 1994), radishes (Giusti & Wrolstad, 1996b), and red cabbage

(Dyrby et al., 2001). Table 2.5 provides detailed degradation kinetics and estimated parameters

of various fruits and vegetables. A varied number of possible inactivation/degradation reactions

in food matrices occur during processing, which may involve several reaction mechanisms. The

rate of inactivation/degradation is reflected by the numerical values of the kinetic parameter

estimates such as order of the reaction, rate constants, activation energy, D-values and z-value.

The order of the reaction in the thermal degradation of anthocyanins, which is dependent on the

concentration and time, could be predicted by the following relationship (Eq.1):

Page 92: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

73

nCkdt

dC)( (1)

where k is the rate constant, n is the reaction order, C is the concentration of the total

anthocyanins and t is the reaction time.

Order of reaction is determined by comparing the coefficient of regression, where exponent n

in eq. 1 was set to zero, half, one and two in zero-, half-, 1st and 2

nd order reactions, respectively.

The integrated form of zero-, half-, 1st and 2

nd order kinetic models is given in Eq. (2) – (5).

Zero-order: ktCCt 0 (2)

Half-order: ktCCt 02 (3)

First-order: ktC

C

o

t ln (4)

Second-order: ktCCt

0

11 (5)

Degradation of anthocyanins under isothermal conditions are reported to follow a first order

kinetics (Eq. 4) in strawberries, elderberries and black carrots at pH 3.5 (Sadilova et al., 2007),

purple-fleshed potatoes, red-fleshed potatoes, grapes and purple carrots (Reyes & Cisneros-

Zevallos, 2007) and delphidin-3-rutinoside isolates (from eggplant skin) and malvidin-3-

glucoside (from grapes) (Tanchev & Yoncheva, 1974). However, degradation of anthocyanins

from red radishes and red-fleshed potatoes follows a quadratic model during storage at 25 °C for

65 weeks (Rodriguez-Saona et al., 1999). For first order degradation kinetics, estimation of half-

lives is important. The half-lives (t1/2) of the anthocyanins were calculated as:

k

t2ln

21 (6)

Page 93: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

74

Acylation of anthocyanins in black carrots prolongs their half-life (t1/2 ~ 2.8 h) compared

with nonacylated derivatives from strawberry (t1/2 ~ 1.9 h) and elderberry (t1/2 ~ 1.9 h) isolates at

pH 3.5 (Sadilova et al., 2007). However, higher half-life values of 4.1 and 3.2 h are observed for

black carrots and strawberries, respectively, at similar temperature conditions (Sadilova et al.,

2006). In vegetables, diacylated anthocyanins have higher half-lives compared to monoacylated

anthocyanin. For example, red radish juice with diacylated anthocyanin has half-lives of 16

weeks compared to red-fleshed potato juice with a half-life of 10 weeks at pH 3.5 (Rodriguez-

Saona et al., 1999). The same researchers reported that the half-lives of prepared anthocyanin

extracts from red radishes (24 weeks) were higher than those of red-fleshed potatoes (11 weeks)

at pH 3.5.

Thermal processing (blanching, pasteurization, sterilization) and storage of foods involve

temperature as the major extrinsic factor for food safety and nutritional quality. Parameters to

estimate the influence of temperature on the quality of food such as activation energy (Ea), z-

value and Q10 (Eq. 7 – 10) which have been reported for acai (Del Pozo-Insfran et al., 2004),

blood orange juice (Cisse, Vaillant, Acosta, Dhuique-Mayer, & Dornier, 2009; Kirca &

Cemeroglu, 2003), blackberry juice, roselle extract (Cisse et al., 2009) and black carrot juice

(Kirca et al., 2007). The thermal death time method (D-z model) is used to estimate the decimal

reduction time (D-value) i.e. heating time required to reduce the anthocyanins concentration by

90% and z-value i.e. temperature change necessary to alter the thermal death time by one log

cycle (Holdsworth, 2000). Q10 is the factor by which the reaction rate is increased if the

temperature is raised by 10 °C.

kD

10ln (7)

Page 94: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

75

TTR

Ekk

ref

aref

11exp (8)

zTTDD refref /)()/log( (9)

T

T

k

kQ

)10(

10

(10)

)exp( ktCC

CC

eqo

eqt

(11)

where Dref is the D-value at temperature Tref, Ea is the activation energy (kJmol-1

), T is the

absolute temperature (K), R is the universal constant (8.135 Jmol-1

K-1

), kT is the reaction rate at

temperature T and k(T+10) is the reaction rate at temperature T+10, and Ceq is the final equilibrium

value of the concentration. Ochoa et al. (2001) used the concept of fractional conversion (Eq. 11)

to study the effect of light and room temperature on the kinetics of color change in raspberries,

sweet and sour cherries.

Calculation of activation energy using the two-step procedure has resulted in a relatively

large standard deviation and particularly with a large confidence interval caused by the small

number of degree of freedom (Arabshahi & Lund, 1985). Ochoa et al. (2001) used a non-linear

model (Eq. 12 & 13) to increase the degree of freedom, thus narrowing the confidence interval

for kinetics of color in fruits.

)exp( )((0),( ) ijTTTt tkCCiiiij

(12)

i

aijoTTt

RT

EtkCC

iiijexpexp

)(0),( (13)

where ),( iij TtC is the concentration of C at time tij and temperature Ti, and )(0 iTC is the initial

concentration at time zero for temperature Ti.

Page 95: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

76

Table 2. 5. Degradation kinetic parameters of anthocyanins.(d:days; h:hour) Fruit/vegetable Processing condition Degradation

kinetics

Kinetic parameter Reference

Acai 10 – 30 °C;

H2O2 (0- 30 mmol/L)

1st order k = (7.7 – 13.9) x 10

3 min

-1;

t1/2 = 90 – 50 min;

Q10 = 1.5 (10 – 20 °C);

Q10 = 1.2 (20 – 30 °C)

Del Pozo-Insfran et al.

(2004)

Grapes 25 – 98 °C;

pH 3.0

1st order k = 0.0006 – 0.2853 h

-1;

t1/2 = 47 d – 2.4 h;

D-value = 157 d – 8.1 h;

z-value = 28 °C; Q10 = 2.28;

Ea = 75.03 kJ/mol;

Reyes & Cisneros-Zevallos

(2007)

Grape skin

(in McIlvaine buffer

(B) and carbonated

soft drink (SD))

25 – 80 °C for 0.25 – 6 h;

pH 3.0

1st order k = 3.6 – 15 x 10

-3 h

-1 (B);

k = 2.4 – 320 x 10-3

h-1

(SD);

Ea = 58 kJ/mol (B);

Ea = 77 kJ/mol (SD)

Dyrby et al. (2001)

Concord grapes

pomace

Retort (126.7 °C) for >30 min Non-linear regression

(non-isothermal)

k110 °C = 0.0607 min-1

;

Ea = 65.32 kJ/mol

Mishra et al. (2008)

Raspberries 4 – 40 °C for 5 months Non-linear regression k = (2.00 – 7.10) x 103 d

-1;

Ea = 26 kJ/mol

Ochoa et al. (2001)

Sweet cherries 4 – 40 °C for 5 months Non-linear regression k = (1.28 – 6.95) x 103 d

-1;

Ea = 32.49 kJ/mol

Ochoa et al. (2001)

Sour cherries 4 – 40 °C for 5 months Non-linear regression k = (1.10 – 5.37) x 103 d

-1;

Ea = 34.8 kJ/mol

Ochoa et al. (2001)

Plum puree 50 – 90 °C for 0 – 20 min 1st order Ea = 37.48 kJ/mol Ahmed et al. (2004)

Blood orange juice 5 – 37 °C & 70 – 90 °C 1st order Ea = 73.2 – 89.5 kJ/mol (11.2 – 69 °Brix) Kirca & Cemeroglu (2003)

Blood orange juice 30 – 90 °C 1st order D-value = (13 – 158) x 10

3 s;

Z-value = 36 °C;

Ea = 66 kJ/mol;

∆H = 63 kJ/mol; ∆S = -149 J/mol-K

Cisse et al. (2009)

Blueberry juice 40 – 80 °C 1st order k = (0.064 – 2.25) x 10

3 min

-1;

t1/2 = 180.5 – 5.11 h;

Q10 = 4.27 (40 – 50 °C);

Q10 = 1.67 (50 – 60 °C);

Q10 = 2.95 (60 – 70 °C);

Q10 = 1.67 (70 – 80 °C);

∆H = 77.8 kJ/mol; ∆S = -43.07 J/mol-K

Kechinski et al. (2010)

Strawberries Stored at 25 °C 1st order t1/2 = 56 – 934 d Garzon & Wrolstad (2001)

Page 96: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

77

Strawberries (fresh-

cut)

5 – 20 °C with 80 kPa O2

flushing

Non-linear regression

(Weibull model)

kα = 4.4 x 10-3

– 4.4 x 10-2

d-1

Odriozola-Serrano et al.

(2009b)

Strawberry

concentrate

95 °C for 6 – 7 h 1st order t1/2 = 1.95 h (pH 3.5)

t1/2 = 3.2 h (pH 1.0)

Sadilova et al. (2007);

Sadilova et al. (2006)

Elderberry

concentrate

95 °C for 6 – 7 h 1st order t1/2 = 1.96 h (pH 3.5)

t1/2 = 1.9 h (pH 1.0)

Sadilova et al. (2007);

Sadilova et al. (2006)

Elderberries

(in McIlvaine buffer

(B) and carbonated

soft drink (SD))

25 – 80 °C for 0.25 – 6 h;

pH 3.0

1st order k = 31 – 90 x 10

-3 h

-1 (B);

k = 5.5 – 180 x 10-3

h-1

(SD);

Ea = 89 kJ/mol (B);

Ea = 56 kJ/mol (SD)

Dyrby et al. (2001)

Blackcurrant juice

(model system)

Heating

(4 – 100 °C)

1st order k = (0.16 – 310) x 10

-3 h

-1;

t1/2 = 180 d – 2.18 h;

Ea = 73 kJ/mol (21 – 100 °C)

Harbourne et al. (2008)

110 °C for 1 – 30 min Non-linear regression

(non-isothermal)

k110 °C = 1.02 h-1

;

Ea = 81.5 kJ/mol

Harbourne et al. (2008)

110 – 140 °C for 1 min Non-linear regression

(non-isothermal)

k = 9.954 h-1

;

Ea = 91.09 kJ/mol

Harbourne et al. (2008)

Blackcurrants

(in McIlvaine buffer

(B) and carbonated

soft drink (SD))

25 – 80 °C for 0.25 – 6 h;

pH 3.0

1st order k = 2.4 – 43 x 10

-3 h

-1 (B);

k = 4.5 – 87 x 10-3

h-1

(SD);

Ea = 69 kJ/mol (B);

Ea = 50 kJ/mol (SD)

Dyrby et al. (2001)

Blackberry juice &

concentrate

60 – 90 °C

(for 8.9 °Brix)

1st order k = 0.69 – 3.94 x 10

3 min

-1;

t1/2 = 16.7 h – 2.9 h;

Ea = 58.95 kJ/mol

Wang & Xu (2007)

Blackberry juice (65

°Brix)

5 – 37 °C 1st order k = 2.0 – 59.1 x 10

3 min

-1;

t1/2 = 330.1 h – 11.7 h;

Ea = 75.5 kJ/mol

Wang & Xu (2007)

Blackberry

concentrate

(65 °Brix)

5 – 37 °C 1st order k = 5.2 – 89.9 x 10

3 min

-1;

t1/2 = 133.3 h – 7.7 h;

Ea = 65.06 kJ/mol

Wang & Xu (2007)

Blackberry juice 100 – 180 °C Non-linear regression

(non-isothermal)

Ea = 92 ± 8 kJ/mol (100 – 140 °C);

Ea = 44 ± 40 kJ/mol (140 – 180 °C);

Jimenez et al. (2010)

Blackberry juice 100 – 180 °C 1st order (isothermal) Ea = 74 kJ/mol Jimenez et al. (2010)

Blackberry juice 30 – 90 °C 1st order D-value = (30 – 341) x 10

3 s

Z-value = 56 – 57 °C

Ea = 37 kJ/mol

∆H = 34 kJ/mol; ∆S = -232 to -233 J/mol-K

Cisse et al. (2009)

Page 97: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

78

Roselle extract 30 – 90 °C 1st order D-value = (30 – 2280) x 10

3 s

Z-value = 34 – 44 °C

Ea = 47 – 61 kJ/mol

∆H = 44 – 58 kJ/mol;

∆S = -205 to -165 J/mol-K

Cisse et al. (2009)

Black Carrot juice 70 – 90 °C;

pH 4.3 and pH 6.0;

11 – 64 °Brix

1st order Ea = 62.5 – 95.1 kJ/mol;

Q10 = 1.7 – 2.8 (70 – 80 °C);

Q10 = 2.0 – 2.2 (80 – 90 °C)

Kirca et al. (2007)

70 – 90 °C; pH 2.5 – 7.0 1st order Ea = 78.1 – 47.4 kJ/mol; Kirca et al. (2007)

Storage at 4 – 37 °C;

pH 4.3

1st order Ea = 62.1 – 86.2 kJ/mol;

Q10 = 2.3 – 3.1 (4 – 20 °C);

Q10 = 2.5 – 3.6 (20 – 37 °C)

Kirca et al. (2007)

Black carrot

concentrate

95 °C for 6 – 7 h 1st order t1/2 = 2.81 h (pH 3.5)

t1/2 = 4.1 h (pH 3.5)

Sadilova et al. (2007);

Sadilova et al. (2006)

Red radish juice

concentrate

Stored (2 or 25 °C for 65

weeks)

2nd

order t1/2 = 16 weeks (25 °C)

t1/2 > 65 weeks (2 °C)

Rodriguez-Saona et al.

(1999)

Red-fleshed potato

juice concentrate

Stored (2 or 25 °C for 65

weeks)

2nd

order t1/2 = 10 weeks (25 °C)

t1/2 = 60 weeks (2 °C)

Rodriguez-Saona et al.

(1999)

Purple-fleshed

potatoes

25 – 98 °C;

pH 3.0

1st order k = 0.0007 – 0.3259 h

-1;

t1/2 = 41 d – 2.1 h;

D-value = 137 d – 7.1 h;

z-value = 28.4 °C; Q10 = 2.25;

Ea = 72.49 kJ/mol;

Reyes & Cisneros-Zevallos

(2007)

Red-fleshed potatoes 25 – 98 °C;

pH 3.0

1st order k = 0.0003 – 0.0725 h

-1;

t1/2 = 89 d – 9.6 h;

D-value = 297 d – 32 h;

z-value = 31.5 °C; Q10 = 2.08;

Ea = 66.7 kJ/mol;

Reyes & Cisneros-Zevallos

(2007)

Purple carrots 25 – 98 °C;

pH 3.0

1st order k = 0.0001 – 0.1004 h

-1;

t1/2 = 216 d – 6.9 h;

D-value = 717 d – 23 h; z-value = 26 °C

Q10 = 2.44; Ea = 81.34 kJ/mol;

Reyes & Cisneros-Zevallos

(2007)

Red cabbage

(in McIlvaine buffer

(B) and carbonated

soft drink (SD))

80 °C;

pH 3.0

1st order k = 9.0 x 10

-3 h

-1 (B);

k = 3.6 x 10-3

h-1

(SD)

Dyrby et al. (2001)

Page 98: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

79

The isothermal methods work well for kinetic studies of samples with rapid heat transfer rate,

and where sufficient concentration remains when lag time ends. For example, for 1st order

isothermal reactions, ―sufficient concentration‖ for accurate estimation of k is C/C0 ≥ 25% (Back

& Arnold, 1977). Processing of solid or semi-solid foods enriched with anthocyanins at high

temperature may be non-isothermal because of their varying water content. In addition the come-

up-time i.e. time to reach ± 0.5 °C of the set temperature, will be different depending on the

sample matrix. For a non-isothermal process, temperature of an individual sample changes with

time. The non-isothermal model (Eq. 14 & 15) involves thermal history (β) combining both

individual time and temperature. The kinetic parameters such as reaction rate constant (k) and

activation energy (Ea) can be estimated by minimizing the sum of square errors (Dolan, 2003).

Recently, Mishra, Dolan, & Yang (2008) and Harbourne, Jacquier, Morgan, & Lyng (2008)

studied the non-isothermal kinetic degradation of anthocyanins of grape pomace and

blackcurrants, respectively, in model systems.

tk

o

t eC

C (14)

t

ref

a

TTR

E

0

11exp (15)

The Weibull model (Eq. 16) has the potential to describe chemical degradation kinetics in

addition to microbial and enzymatic kinetics. Recently, Oms-Oliu, Odriozola-Serrano, Soliva-

Fortuny, & Martín-Belloso (2009) used the Weibull model and accurately described the kinetics

of phenolics and antioxidant changes (DPPH) in fresh-cut watermelon stored at 5 – 20 °C for 14

days. In another study, Odriozola-Serrano, Soliva-Fortuny, & Martin-Belloso (2009b) reported

the accuracy of the Weibull model to describe the changes in anthocyanins and antioxidant

Page 99: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

80

activity (DPPH) of fresh-cut strawberries stored at similar conditions under a high oxygen

atmosphere (80 kPa).

).(exp[0 ktAOXAOX (16)

where AOX is the percentage relative antioxidant property, AOX0 is the intercept of the curve, t

(days) is the storage time, kα (per day) is the kinetic constant, which is the inverse of the scale

factor (α) and γ is the shape parameter. Mean storage time (tm) i.e. the time for 100% depletion of

the antioxidant property and reaction rate constant (k) is expressed as Eq. 17 & 18, respectively:

)1

1()1

(

k

tm (17)

m

cTTck )}](exp{1ln[ (18)

where Ґ is the gamma function, T (K) is the storage temperature, Tc (K) is the marker of the

temperature range where the changes accelerate, c (per K), and m are dimensional constants.

The enthalpy of activation (∆H) and entropy of activation (∆S) of anthocyanins has been

reported for blueberry juice (Kechinski et al., 2010) and blood oranges, blackberries, and roselle

(Cisse et al., 2009) using the Eyring-Polanyi model based on transition state theory:

RT

STH

b Teh

kk

(19)

where T is the absolute temperature (K), kb is the Boltzman constant (1.381 x 10-23

J/K), h is the

Planck constant (6.626 x 10-34

Js) and R is the gas constant (8.31 J/mol K).

4. SUMMARY

Most investigations in the literature on the phenolic antioxidants in fruits, vegetables, and

grains have shown that high temperature treatments can reduce the antioxidant activity, whereas

some studies reported to have increased in the antioxidant activity of processed foods. Some

Page 100: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

81

investigators also emphasized the intrinsic properties of the food matrix responsible for the

mixed behavior of the antioxidant compounds during processing studies of fruits and vegetables.

Thermal degradation of anthocyanins in fruits and vegetables has been extensively conducted

and reported in the model systems, juices and concentrates. Degradation kinetics of

anthocyanins, mostly studied under 100 °C, follows the first order reaction in these matrices.

However, the responses of antioxidant compounds depend on the specific fruit or vegetable.

Based on the available knowledge, it will be biased to predict the effect of thermal treatments on

phenolic antioxidants retention. There is a need to analyze these compounds on a case-by-case

basis. Therefore, the current study aimed to investigate the presence of phenolic antioxidants in

colored potatoes, their retention/degradation, bioavailability and kinetics in the processed (at >

100 °C) value-added products used in canning, baking and extrusion cooking. The results could

provide valuable information to farmers, researchers and industry.

Page 101: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

82

REFERENCES

Adams, J. B. (1973). Thermal degradation of anthocyanins with particular reference to the 3-

glycosides of cyanidin. I. In acidified aqueous solution at 100 °C. Journal of the Science of

Food and Agriculture, 24, 747-762.

Adom, K. K., & Liu, R. H. (2002). Antioxidant activity of grains. Journal of Agricultural and

Food Chemistry, 50, 6182-6187.

Aguillar-Rosas, S. F., Ballinas-Casarrubias, M. L., Nevarez-Moorillon, G. V., Martin-Belloso,

O., & Ortega-Rivas, E. (2007). Thermal and pulsed electric fields pasteurization of apple

juice: Effects on physicochemical properties and flavour compounds. Journal of Food

Engineering, 83, 41-46.

Ahmed, J., Shivhare, U. S., & Raghavan, G. S. V. (2004). Thermal degradation kinetics of

anthocyanin and visual colour of plum puree. European Food Research and Technology,

218, 525-528.

Allali, H., Marchal, L., & Vorobiev, E. (2010). Blanching of strawberries by ohmic heating:

effects on the kinetics of mass transfer during osmotic dehydration. Food and Bioprocess

Technology, 3, 406-414.

Alonso, R., Grant, G., Dewey, P., & Marzo, F. (2000). Nutritional assessment in vitro and in

vivo of raw and extruded peas (Pisum sativum L.). Journal of Agricultural and Food

Chemistry, 48, 2286-2290.

Alsaikhan, M. S., Howard, L. R., & Miller, J. C. (1995). Antioxidant activity and total phenolics

in different genotypes of potato (Solanum tuberosum, L). Journal of Food Science, 60, 341-

347.

Page 102: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

83

Altamirano, R. C., Drdak, M., Simon, P., Smelik, A., & Simko, P. (1992). Stability of red beet

pigment concentrate in maize starch. Journal of the Science of Food and Agriculture, 58,

595-596.

Altan, A., McCarthy, K. L., & Maskan, M. (2009). Effect of extrusion process on antioxidant

activity, total phenolics and beta-glucan content of extrudates developed from barley-fruit

and vegetable by-products. International Journal of Food Science and Technology, 44, 1263-

1271.

Ames, B. N., Shigenaga, M. K., & Hagen, T. M. (1993). Oxidants, antioxidants, and the

degenerative diseases of aging. Proceedings of the National Academy of Sciences of the

United States of America, 90, 7915-7922.

Amin, I., & Lee, W. Y. (2005). Effect of different blanching times on antioxidant properties in

selected cruciferous vegetables. Journal of the Science of Food and Agriculture, 85, 2314-

2320.

Andreasen, M. F., Kroon, P. A., Williamson, G., & Garcia-Conesa, M. T. (2001). Intestinal

release and uptake of phenolic antioxidant diferulic acids. Free Radical Biology and

Medicine, 31, 304-314.

Anese, M., Manzocco, L., Nicoli, M. C., & Lerici, C. R. (1999). Antioxidant properties of

tomato juice as affected by heating. Journal of the Science of Food and Agriculture, 79, 750-

754.

Antolovich, M., Prenzler, P. D., Patsalides, E., McDonald, S., & Robards, K. (2002). Methods

for testing antioxidant activity. Analyst, 127, 183-198.

Arabshahi, A., & Lund, D. B. (1985). Considerations in calculating kinetic parameters from

experimental data. Journal of Food Process Engineering, 7, 239-251.

Page 103: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

84

Back, J. V., & Arnold, K. J. (1977). Parameter estimation in engineering and science. New

York, NY: John Wiley & Sons.

Bakker, J., & Timberlake, C. F. (1997). Isolation, identification, and characterization of new

color-stable anthocyanins occurring in some red wines. Journal of Agricultural and Food

Chemistry, 45, 35-43.

Barka, E. A., Kalantari, S., Makhlouf, J., & Arul, J. (2000). Effects of UV-C irradiation on lipid

peroxidation markers during ripening of tomato (Lycopersicon esculentum L.) fruits.

Australian Journal of Plant Physiology, 27, 147-152.

Baublis, A., Spomer, A., & Berberjimenez, M. D. (1994). Anthocyanin pigments - comparison of

extract stability. Journal of Food Science, 59, 1219-&.

Bian, Y., Gonzalez, A. R., & Asleage, J. M. (1994). Effects of pasteurization or sodium benzoate

on long-term storage stability of peach puree. Journal of Food Quality, 17, 299-310.

Blessington, A., Miller, J. C., Nzaramba, M. N., Hale, A. L., Redivari, L., Scheuring, D. C., &

Hallman, G. J. (2007). The effects of low-dose gamma irradiation and storage time on

carotenoids, antioxidant activity, and phenolics in the potato cultivar Atlantic. American

Journal of Potato Research, 84, 125-131.

Blessington, T. (2005). The effects of cooking, storage and ionizing irradiation on carotenoids,

antioxidant activity and phenolics in potato. Texas A&M, College station.

Block, G., Patterson, B., & Subar, A. (1992). Fruit, vegetables, and cancer prevention - a review

of the epidemiologic evidence. Nutrition and Cancer-an International Journal, 18, 1-29.

Bonoli, M., Verardo, V., Marconi, E., & Caboni, M. F. (2004). Phenols in barley (Hordeum

vulgare L.) flour: Comparative spectrophotometric study among extraction methods of free

Page 104: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

85

and bound phenolic compounds. Journal of Agricultural and Food Chemistry, 52, 5195-

5200.

Breitfellner, F., Solar, S., & Sontag, G. (2002). Effect of gamma-irradiation on phenolic acids in

strawberries. Journal of Food Science, 67, 517-521.

Bridle, P., & Timberlake, C. F. (1997). Anthocyanins as natural food colours - Selected aspects.

Food Chemistry, 58, 103-109.

BrØnnum-Hansen, K., & Flink, J. M. (1985). Anthocyanin colourants from elderberry

(Sambucus nigra L.). 3. Storage stability of the freeze dried product. International Journal of

Food Science & Technology, 20, 725-733.

Brouillard, R. (1981). Origin of the exceptional colour stability of the Zebrina anthocyanin.

Phytochemistry, 20, 143-145.

Brouillard, R. (1983). The in vivo expression of anthocyanin colour in plants. Phytochemistry,

22, 1311-1323.

Brownmiller, C., Howard, L. R., & Prior, R. L. (2008). Processing and storage effects on

monomeric anthocyanins, percent polymeric color, and antioxidant capacity of processed

blueberry products. Journal of Food Science, 73, H72-H79.

Bryngelsson, S., Dimberg, L. H., & Kamal-Eldin, A. (2002). Effects of commercial processing

on levels of antioxidants in oats (Avena sativa L.). Journal of Agricultural and Food

Chemistry, 50, 1890-1896.

Burton, W. G. (1989). The Potato. New York, NY: Wiley & Sons.

Butz, P., FernandezGarcia, A., Lindauer, R., Dieterich, S., Bognar, A., & Tauscher, B. (2003).

Influence of ultra high pressure processing on fruit and vegetable products. Journal of Food

Engineering, 56, 233-236.

Page 105: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

86

Callebaut, A., Hendrickx, G., Voets, A. M., & Motte, J. C. (1990). Anthocyanins in cell-cultures

of Ajuga-Reptans. Phytochemistry, 29, 2153-2158.

Calvi, J., & Francis, F. (1978). Stability of concord grape (V. Labrusca) anthocyanins in model

systems. Journal of Food Science, 43, 1448-1456.

Camire, M. E., Chaovanalikit, A., Dougherty, M. P., & Briggs, J. (2002). Blueberry and grape

anthocyanins as breakfast cereal colorants. Journal of Food Science, 67, 438-441.

Camire, M. E., Dougherty, M. P., & Briggs, J. L. (2007). Functionality of fruit powders in

extruded corn breakfast cereals. Food Chemistry, 101, 765-770.

Cano, M. P., Hernandez, A., & De Ancos, B. (1997). High pressure and temperature effects on

enzyme inactivation in strawberry and orange products. Journal of Food Science, 62, 85-88.

Cantos, E., Garcia-Viguera, C., de Pascual-Teresa, S., & Tomas-Berberan, F. A. (2000). Effect

of postharvest ultraviolet irradiation on resveratrol and other phenolics of cv. Napoleon table

grapes. Journal of Agricultural and Food Chemistry, 48, 4606-4612.

Cao, G. H., Booth, S. L., Sadowski, J. A., & Prior, R. L. (1998). Increases in human plasma

antioxidant capacity after consumption of controlled diets high in fruit and vegetables.

American Journal of Clinical Nutrition, 68, 1081-1087.

Cao, G. H., Russell, R. M., Lischner, N., & Prior, R. L. (1998). Serum antioxidant capacity is

increased by consumption of strawberries, spinach, red wine or vitamin C in elderly women.

Journal of Nutrition, 128, 2383-2390.

Cemeroglu, B., Velioglu, S., & Isik, S. (1994). Degradation kinetics of anthocyanins in sour

cherry juice and concentrate. Journal of Food Science, 59, 1216-1218.

Chaovanalikit, A., & Wrolstad, R. E. (2004a). Anthocyanin and polyphenolic composition of

fresh and processed cherries. Journal of Food Science, 69, C73-C83.

Page 106: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

87

Chaovanalikit, A., & Wrolstad, R. E. (2004b). Total anthocyanins and total phenolics of fresh

and processed cherries and their antioxidant properties. Journal of Food Science, 69, C67-

C72.

Cheigh, H. S., Um, S. H., & Lee, C. Y. (1995). Antioxidant characteristics of melanin-related

products from enzymatic browning reaction of catechin in a model system. Enzymatic

Browning and Its Prevention, 600, 200-208.

Chu, Y. H., Chang, C. L., & Hsu, H. F. (2000). Flavonoid content of several vegetables and their

antioxidant activity. Journal of the Science of Food and Agriculture, 80, 561-566.

Chuah, A. M., Yamaguchi, H. T., & Matoba, T. (2005) 2nd International conference on

polyphenols and health (pp. 168). University of California, Davis, CA.

Chun, O. K., Kim, D. O., Smith, N., Schroeder, D., Han, J. T., & Lee, C. Y. (2005). Daily

consumption of phenolics and total antioxidant capacity from fruit and vegetables in the

American diet. Journal of the Science of Food and Agriculture, 85, 1715-1724.

Cisse, M., Vaillant, F., Acosta, O., Dhuique-Mayer, C., & Dornier, M. (2009). Thermal

degradation kinetics of anthocyanins from blood orange, blackberry, and roselle using the

Arrhenius, Eyring, and Ball models. Journal of Agricultural and Food Chemistry, 57, 6285-

6291.

Clifford, M. N. (2000). Anthocyanins – nature, occurrence and dietary burden. Journal of the

Science of Food and Agriculture, 80, 1063-1072.

Connor, A. M., Luby, J. J., Tong, C. B. S., Finn, C. E., & Hancock, J. F. (2002). Genotypic and

environmental variation in antioxidant activity, total phenolic content, and anthocyanin

content among blueberry cultivars. Journal of the American Society for Horticultural

Science, 127, 89-97.

Page 107: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

88

Cuvelier, M.-E., Bondet, V., & Berset, C. (2000). Behavior of phenolic antioxidants in a

partitioned medium: structure—Activity relationship. Journal of the American Oil Chemists'

Society, 77, 819-824.

Dao, L., & Friedman, M. (1992). Chlorogenic acid content of fresh and processed potatoes

determined by ultraviolet spectrophotometry. Journal of Agricultural and Food Chemistry,

40, 2152-2156.

de Ancos, B., Gonzalez, E., & Cano, M. P. (2000). Effect of high-pressure treatment on the

carotenoid composition and the radical scavenging activity of persimmon fruit purees.

Journal of Agricultural and Food Chemistry, 48, 3542-3548.

Dede, S., Alpas, H., & Bayindirli, A. (2007). High hydrostatic pressure treatment and storage of

carrot and tomato juices: Antioxidant activity and microbial safety. Journal of the Science of

Food and Agriculture, 87, 773-782.

Del Pozo-Insfran, D., Brenes, C. H., & Talcottt, S. T. (2004). Phytochemical composition and

pigment stability of acai (Euterpe oleracea Mart.). Journal of Agricultural and Food

Chemistry, 52, 1539-1545.

Dewanto, V., Wu, X. Z., Adom, K. K., & Liu, R. H. (2002). Thermal processing enhances the

nutritional value of tomatoes by increasing total antioxidant activity. Journal of Agricultural

and Food Chemistry, 50, 3010-3014.

Dewanto, V., Wu, X. Z., & Liu, R. H. (2002). Processed sweet corn has higher antioxidant

activity. Journal of Agricultural and Food Chemistry, 50, 4959-4964.

Dlamini, N. R., Taylor, J. R. N., & Rooney, L. W. (2007). The effect of sorghum type and

processing on the antioxidant properties of African sorghum-based foods. Food Chemistry,

105, 1412-1419.

Page 108: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

89

Doblado, R., Frias, J., & Vidal-Valverde, C. (2007). Changes in vitamin C content and

antioxidant capacity of raw and germinated cowpea (Vigna sinensis var. carilla) seeds

induced by high pressure treatment. Food Chemistry, 101, 918-923.

Dolan, K. D. (2003). Estimation of kinetic parameters for nonisothermal food processes. Journal

of Food Science, 68, 728-741.

Donovan, J. L., Meyer, A. S., & Waterhouse, A. L. (1998). Phenolic composition and antioxidant

activity of prunes and prune juice (Prunus domestica). Journal of Agricultural and Food

Chemistry, 46, 1247-1252.

Donsi, G., Ferrari, G., & DiMatteo, M. (1996). High pressure stabilization of orange juice:

Evaluation of the effects of process conditions. Italian Journal of Food Science, 8, 99-106.

Dubery, I. A., Louw, A. E., & van Heerden, F. R. (1999). Synthesis and evaluation of 4-(3-

methyl-2-butenoxy) isonitrosoacetophenone, a radiation-induced stress metabolite in Citrus.

Phytochemistry, 50, 983-989.

Duve, K. J., & White, P. J. (1991). Extraction and identification of antioxidants in oats. Journal

of the American Oil Chemists Society, 68, 365-370.

Dyrby, M., Westergaard, N., & Stapelfeldt, H. (2001). Light and heat sensitivity of red cabbage

extract in soft drink model systems. Food Chemistry, 72, 431-437.

Eberhardt, M. V., Lee, C. Y., & Liu, R. H. (2000). Nutrition - antioxidant activity of fresh

apples. Nature, 405, 903-904.

Elez-Martínez, P., & Martín-Belloso, O. (2007). Effects of high intensity pulsed electric field

processing conditions on vitamin C and antioxidant capacity of orange juice and gazpacho, a

cold vegetable soup. Food Chemistry, 102, 201-209.

Page 109: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

90

Erlandson, J. A., & Wrolstad, R. E. (1972). Degradation of anthocyanins at limited water

concentration. Journal of Food Science, 37, 592-595.

Ersus, S., & Yurdagel, U. (2007). Microencapsulation of anthocyanin pigments of black carrot

(Daucuscarota L.) by spray drier. Journal of Food Engineering, 80, 805-812.

Fan-Chiang, H. J., & Wrolstad, R. E. (2005). Anthocyanin pigment composition of blackberries.

Journal of Food Science, 70, C198-C202.

Fan, X. T., & Thayer, D. W. (2001). Quality of irradiated alfalfa sprouts. Journal of Food

Protection, 64, 1574-1578.

Fan, X. T., & Thayer, D. W. (2002). Gamma-radiation influences browning, antioxidant activity,

and malondialdehyde level of apple juice. Journal of Agricultural and Food Chemistry, 50,

710-715.

Fan, X. T., Toivonen, P. M. A., Rajkowski, K. T., & Sokorai, K. J. B. (2003). Warm water

treatment in combination with modified atmosphere packaging reduces undesirable effects of

irradiation on the quality of fresh-cut iceberg lettuce. Journal of Agricultural and Food

Chemistry, 51, 1231-1236.

Feinberg, B., Winter, F., & Roth, T. L. (1968). The preparation for freezing and freezing of

vegetables. In D. K. Tressler, W. B. Van-Arsdel & M. J. Copley (Eds.), The freezing

preservation of foods. Westport, Connecticut: Avi publishing Co.

Fernandez-Garcia, A., Butz, P., Bognar, A., & Tauscher, B. (2001). Antioxidative capacity,

nutrient content and sensory quality of orange juice and an orange-lemon-carrot juice product

after high pressure treatment and storage in different packaging. European Food Research

and Technology, 213, 290-296.

Page 110: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

91

Fernandez-Garcia, A., Butz, P., & Tauscher, B. (2001). Effects of high-pressure processing on

carotenoid extractability, antioxidant activity, glucose diffusion, and water binding of tomato

puree (Lycopersicon esculentum Mill.). Journal of Food Science, 66, 1033-1038.

Floros, J. D., & Liang, H. H. (1994). Acoustically assisted diffusion through membranes and

biomaterials. Food Technology, 48, 79-84.

Francia-Aricha, E. M., Guerra, M. T., Rivas-Gonzalo, J. C., & Santos-Buelga, C. (1997). New

anthocyanin pigments formed after condensation with flavanols. Journal of Agricultural and

Food Chemistry, 45, 2262-2266.

Francis, F. J. (1989). Food colorants - Anthocyanins. Critical Reviews in Food Science and

Nutrition, 28, 273-314.

Frankel, E. N., Huang, S. W., Kanner, J., & German, J. B. (1994). Interfacial phenomena in the

evaluation of antioxidants - bulk oils vs emulsions. Journal of Agricultural and Food

Chemistry, 42, 1054-1059.

Friedman, M. (1997). Chemistry, biochemistry, and dietary role of potato polyphenols. A review.

Journal of Agricultural and Food Chemistry, 45, 1523-1540.

Garcıia-Viguera, C., Zafrilla, P., & Tomás-Barberán, F. A. (1997). Determination of authenticity

of fruit jams by HPLC analysis of anthocyanins. Journal of the Science of Food and

Agriculture, 73, 207-213.

Garza, S., Ibarz, A., Pagan, J., & Giner, J. (1999). Non-enzymatic browning in peach puree

during heating. Food Research International, 32, 335-343.

Garzon, G. A., & Wrolstad, R. E. (2002). Comparison of the stability of pelargonidin-based

anthocyanins in strawberry juice and concentrate. Journal of Food Science, 67, 1288-1299.

Page 111: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

92

Garzón, G. A., & Wrolstad, R. E. (2001). The stability of pelargonidin-based anthocyanins at

varying water activity. Food Chemistry, 75, 185-196.

Gerard, K. A., & Roberts, J. S. (2004). Microwave heating of apple mash to improve Juice yield

and quality. Lebensmittel-Wissenschaft Und-Technologie-Food Science and Technology, 37,

551-557.

Gil, M. I., Ferreres, F., & Tomas-Barberan, F. A. (1999). Effect of postharvest storage and

processing on the antioxidant constituents (flavonoids and vitamin C) of fresh-cut spinach.

Journal of Agricultural and Food Chemistry, 47, 2213-2217.

Gil, M. I., Holcroft, D. M., & Kader, A. A. (1997). Changes in strawberry anthocyanins and

other polyphenols in response to carbon dioxide treatments. Journal of Agricultural and

Food Chemistry, 45, 1662-1667.

Giusti, M. M., Rodriguez-Saona, L. E., Griffin, D., & Wrolstad, R. E. (1999). Electrospray and

tandem mass spectroscopy as tools for anthocyanin characterization. Journal of Agricultural

and Food Chemistry, 47, 4657-4664.

Giusti, M. M., & Wrolstad, R. E. (1996a). Characterization of red radish anthocyanins. Journal

of Food Science, 61, 322-326.

Giusti, M. M., & Wrolstad, R. E. (1996b). Radish anthocyanin extract as a natural red colorant

for maraschino cherries. Journal of Food Science, 61, 688-694.

Glassgen, W. E., Wray, V., Strack, D., Metzger, J. W., & Seitz, H. U. (1992). Anthocyanins from

cell-suspension cultures of Daucus-Carota. Phytochemistry, 31, 1593-1601.

Goto, T., & Kondo, T. (1991). Structure and molecular stacking of anthocyanins - flower color

variation. Angewandte Chemie-International Edition in English, 30, 17-33.

Page 112: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

93

Grela, E. R., Jensen, S. K., & Jakobsen, K. (1999). Fatty acid composition and content of

tocopherols and carotenoids in raw and extruded grass pea (Lathyrus sativus L). Journal of

the Science of Food and Agriculture, 79, 2075-2078.

Griffiths, D. W., Bain, H., & Dale, M. F. B. (1995). Photoinduced changes in the total

chlorogenic acid content of potato (Solanum tuberosum) tubers. Journal of the Science of

Food and Agriculture, 68, 105-110.

Guyot, S., Marnet, N., Laraba, D., Sanoner, P., & Drilleau, J. F. (1998). Reversed-phase HPLC

following thiolysis for quantitative estimation and characterization of the four main classes of

phenolic compounds in different tissue zones of a French cider apple variety (Malus

domestica var. Kermerrien). Journal of Agricultural and Food Chemistry, 46, 1698-1705.

Hager, A., Howard, L. R., Prior, R. L., & Brownmiller, C. (2008). Processing and storage effects

on monomeric anthocyanins, percent polymeric color, and antioxidant capacity of processed

black raspberry products. Journal of Food Science, 73, H134-H140.

Hager, T. J., Howard, L. R., & Prior, R. L. (2008). Processing and storage effects on monomeric

anthocyanins, percent polymeric color, and antioxidant capacity of processed blackberry

products. Journal of Agricultural and Food Chemistry, 56, 689-695.

Hakkinen, S. H., Karenlampi, S. O., Mykkanen, H. M., & Torronen, A. R. (2000). Influence of

domestic processing and storage on flavonol contents in berries. Journal of Agricultural and

Food Chemistry, 48, 2960-2965.

Halliwell, B., & Gutteridge, J. M. C. (1990). The antioxidants of human extracellular fluids.

Archives of Biochemistry and Biophysics, 280, 1-8.

Halliwell, B., Gutteridge, J. M. C., & Cross, C. E. (1992). Free radicals, antioxidants, and human

disease - Where are we now. Journal of Laboratory and Clinical Medicine, 119, 598-620.

Page 113: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

94

Halliwell, B., Murcia, M. A., Chirico, S., & Aruoma, O. I. (1995). Free radicals and antioxidants

in food and in vivo - What they do and how they work. Critical Reviews in Food Science and

Nutrition, 35, 7-20.

Hamama, A. A., & Nawar, W. W. (1991). Thermal decomposition of some phenolic

antioxidants. Journal of Agricultural and Food Chemistry, 39, 1063-1069.

Hamouz, K., Lachman, J., Vokal, B., & Pivec, V. (1999). Influence of environmental conditions

and way of cultivation on the polyphenol and ascorbic acid content in potato tubers.

Rostlinna Vyroba, 45, 293-298.

Harbourne, N., Jacquier, J. C., Morgan, D. J., & Lyng, J. G. (2008). Determination of the

degradation kinetics of anthocyanins in a model juice system using isothermal and non-

isothermal methods. Food Chemistry, 111, 204-208.

Holdsworth, S. D. (2000). Thermal processing of packaged foods. New York: Blackie Academic

& Professional.

Hoshino, T., Matsumoto, U., & Goto, T. (1981). Self-association of some anthocyanins in neutral

aqueous solution. Phytochemistry, 20, 1971-1976.

Hrazdina, G. (1971). Reactions of the anthocyanidin-3,5-diglucosides: Formation of 3,5-di-(O-

[beta]--glucosyl)-7-hydroxy coumarin. Phytochemistry, 10, 1125-1130.

Huang, D. J., Ou, B. X., & Prior, R. L. (2005). The chemistry behind antioxidant capacity assays.

Journal of Agricultural and Food Chemistry, 53, 1841-1856.

Icier, F. (2010). Ohmic blanching effects on drying of vegetable byproduct. Journal of Food

Process Engineering, 33, 661-683.

Jackman, R. L., & Smith, J. L. (1992). Anthocyanins and betalains. In G. A. F. Hendry & J. D.

Houghton (Eds.), Natural food colorants (pp. 183 - 241). London: Blackie and Son Ltd.

Page 114: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

95

Jackman, R. L., Yada, R. Y., & Tung, M. A. (1987). A review - Separation and chemical

properties of anthocyanins used for their qualitative and quantitative analysis. Journal of

Food Biochemistry, 11, 279-308.

Jiménez-Monreal, A. M., García-Diz, L., Martínez-Tomé, M., Mariscal, M., & Murcia, M. A.

(2009). Influence of cooking methods on antioxidant activity of vegetables. Journal of Food

Science, 74, H97-H103.

Jimenez, A., Selga, A., Torres, J. U., & Julia, L. (2004). Reducing activity of polyphenols with

stable radicals of the TTM series. Electron transfer versus H-abstraction reactions in flavan-

3-ols. Organic Letters, 6, 4583-4586.

Jimenez, N., Bohuon, P., Lima, J., Dornier, M., Vaillant, F., & Perez, A. M. (2010). Kinetics of

anthocyanin degradation and browning in reconstituted blackberry juice treated at high

temperatures (100 - 180 °C). Journal of Agricultural and Food Chemistry, 58, 2314-2322.

Jiratanan, T., & Liu, R. H. (2004). Antioxidant activity of processed table beets (Beta vulgaris

var, conditiva) and green beans (Phaseolus vulgaris L.). Journal of Agricultural and Food

Chemistry, 52, 2659-2670.

Ju, Z. G., Liu, C. L., & Yuan, Y. B. (1995). Activities of chalcone synthase and udpgal -

Flavonoid-3-o-glycosyltransferase in relation to anthocyanin synthesis in apple. Scientia

Horticulturae, 63, 175-185.

Kaanane, A., Kane, D., & Labuza, T. (1988). Time and temperature effect on stability of

Moroccan processed orange juice during storage. Journal of Food Science, 53, 1470-1473.

Kacharava, N., Chanishvili, S., Badridze, G., Chkhubianishvili, E., & Janukashvili, N. (2009).

Effect of seed irradiation on the content of antioxidants in leaves of Kidney bean, Cabbage

and Beet cultivars. Australian Journal of Crop Science, 3, 137-145.

Page 115: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

96

Kader, A. A. (1986). Biochemical and physiological-basis for effects of controlled and modified

atmospheres on fruits and vegetables. Food Technology, 40, 99 -100 and 102 - 104.

Kader, F., Irmouli, M., Nicolas, J. P., & Metche, M. (2001). Proposed mechanism for the

degradation of pelargonidin 3-glucoside by caffeic acid o-quinone. Food Chemistry, 75, 139-

144.

Kader, F., Irmouli, M., Nicolas, J. P., & Metche, M. (2002). Involvement of blueberry

peroxidase in the mechanisms of anthocyanin degradation in blueberry juice. Journal of

Food Science, 67, 910-915.

Kader, F., Irmouli, M., Zitouni, N., Nicolas, J. P., & Metche, M. (1999). Degradation of cyanidin

3-glucoside by caffeic acid o-quinone. Determination of the stoichiometry and

characterization of the degradation products. Journal of Agricultural and Food Chemistry,

47, 4625-4630.

Kader, F., Rovel, B., Girardin, M., & Metche, M. (1997). Mechanism of browning in fresh

highbush blueberry fruit (Vaccinium corymbosum L). Role of blueberry polyphenol oxidase,

chlorogenic acid and anthocyanins. Journal of the Science of Food and Agriculture, 74, 31-

34.

Kahkonen, M. P., Hopia, A. I., Vuorela, H. J., Rauha, J. P., Pihlaja, K., Kujala, T. S., &

Heinonen, M. (1999). Antioxidant activity of plant extracts containing phenolic compounds.

J. Agric. Food Chem., 47, 3954-3962.

Kalt, W., McDonald, J. E., & Donner, H. (2000). Anthocyanins, phenolics, and antioxidant

capacity of processed lowbush blueberry products. Journal of Food Science, 65, 390-393.

Kalt, W., Ryan, D. A. J., Duy, J. C., Prior, R. L., Ehlenfeldt, M. K., & Vander Kloet, S. P.

(2001). Interspecific variation in anthocyanins, phenolics, and antioxidant capacity among

Page 116: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

97

genotypes of highbush and lowbush blueberries (Vaccinium section cyanococcus spp.).

Journal of Agricultural and Food Chemistry, 49, 4761-4767.

Kang, H. M., & Saltveit, M. E. (2002). Antioxidant capacity of lettuce leaf tissue increases after

wounding. Journal of Agricultural and Food Chemistry, 50, 7536-7541.

Kaur, C., & Kapoor, H. C. (2001). Antioxidants in fruits and vegetables - the millennium's

health. International Journal of Food Science & Technology, 36, 703-725.

Kayano, S., Yamada, N. F., Suzuki, T., Ikami, T., Shioaki, K., Kikuzaki, H., Mitani, T., &

Nakatani, N. (2003). Quantitative evaluation of antioxidant components in prunes (Prunus

domestica L.). Journal of Agricultural and Food Chemistry, 51, 1480-1485.

Ke, D. Y., Goldstein, L., Omahony, M., & Kader, A. A. (1991). Effects of short-term exposure to

low O2 and high CO2 atmospheres on quality attributes of strawberries. Journal of Food

Science, 56, 50-54.

Kearsley, M. W., & Rodriguez, N. (1981). The stability and use of natural colours in foods:

anthocyanin, β-carotene and riboflavin. International Journal of Food Science &

Technology, 16, 421-431.

Kechinski, C. P., Guimaraes, P. V. R., Norena, C. P. Z., Tessaro, I. C., & Marczak, L. D. F.

(2010). Degradation kinetics of anthocyanin in blueberry juice during thermal treatment.

Journal of Food Science, 75, C173-C176.

Khraisheh, M. A. M., Cooper, T. J. R., & Magee, T. R. A. (1995). Investigation and modeling of

combined microwave and air-drying. Food and Bioproducts Processing, 73, 121-126.

Kim, D. O., & Padilla-Zakour, O. I. (2004). Jam processing effect on phenolics and antioxidant

capacity in anthocyanin-rich fruits: Cherry, plum, and raspberry. Journal of Food Science,

69, S395-S400.

Page 117: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

98

Kim, K. H., & Yook, H. S. (2009). Effect of gamma irradiation on quality of kiwifruit (Actinidia

deliciosa var. deliciosa cv. Hayward). Radiation Physics and Chemistry, 78, 414-421.

Kinsella, J. E., Frankel, E., German, B., & Kanner, J. (1993). Possible mechanisms for the

protective role of antioxidants in wine and plant foods. Food Technology, 47, 85-89.

Kirca, A., & Cemeroglu, B. (2003). Degradation kinetics of anthocyanins in blood orange juice

and concentrate. Food Chemistry, 81, 583-587.

Kirca, A., Ozkan, M., & Cemeroglu, B. (2007). Effects of temperature, solid content and pH on

the stability of black carrot anthocyanins. Food Chemistry, 101, 212-218.

Korus, J., Gumul, D., & Czechowska, K. (2007). Effect of extrusion on the phenolic composition

and antioxidant activity of dry beans of Phaseolus vulgaris L. Food Technology and

Biotechnology, 45, 139-146.

Krifi, B., Chouteau, F., Boudrant, J., & Metche, M. (2000). Degradation of anthocyanins from

blood orange juices. International Journal of Food Science and Technology, 35, 275-283.

Kwok, B. H. L., Hu, C., Durance, T., & Kitts, D. D. (2004). Dehydration techniques affect

phytochemical contents and free radical scavenging activities of Saskatoon berries

(Amelanchier alnifolia Nutt.). Journal of Food Science, 69, S122-S126.

Lachman, J., Hamouz, K., & Orsak, M. (2005). Red and purple potatoes - A significant

antioxidant source in human nutrition. Chemicke Listy, 99, 474-482.

Lachman, J., Hamouz, K., Orsak, M., & Pivec, V. (2000). Potato tubers as a significant source of

antioxidants in human nutrition. Rostlinna Vyroba, 46, 231-236.

Lea, A. G. H. (1988). HPLC of natural pigments in foodstuffs. In R. Macrae (Ed.), HLPC in

Food Analysis (pp. 277-333). San Diego,

CA: Academic Press Inc.

Page 118: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

99

Lee, H. S. (1992). Antioxidative activity of browning reaction products isolated from storage-

aged orange juice. Journal of Agricultural and Food Chemistry, 40, 550-552.

Lee, H. S., & Wicker, L. (1991). Anthocyanin pigments in the skin of lychee fruit. Journal of

Food Science, 56, 466-&.

Lee, J., Durst, R. W., & Wrolstad, R. E. (2002). Impact of juice processing on blueberry

anthocyanins and polyphenolics: Comparison of two pretreatments. Journal of Food Science,

67, 1660-1667.

Leusink, G. J., Kitts, D. D., Yaghmaee, P., & Durance, T. (2010). Retention of antioxidant

capacity of vacuum microwave dried cranberry. Journal of Food Science, 75, C311-C316.

Lewis, C. E., Walker, J. R. L., Lancaster, J. E., & Sutton, K. H. (1998). Determination of

anthocyanins, flavonoids and phenolic acids in potatoes. I: Coloured cultivars of Solanum

tuberosum L. Journal of the Science of Food and Agriculture, 77, 45-57.

Li, W., Shan, F., Sun, S., Corke, H., & Beta, T. (2005). Free radical scavenging properties and

phenolic content of Chinese black-grained wheat. Journal of Agricultural and Food

Chemistry, 53, 8533-8536.

Li, W. D., Pickard, M. D., & Beta, T. (2007). Effect of thermal processing on antioxidant

properties of purple wheat bran. Food Chemistry, 104, 1080-1086.

Li, Z., Pan, Q. H., Cui, X. Y., & Duan, C. Q. (2010). Optimization on anthocyanins extraction

from wine grape skins using orthogonal test design. Food Science and Biotechnology, 19,

1047-1053.

Lin, L. D., Wu, J. Y., Ho, K. P., & Qi, S. Y. (2001). Ultrasound-induced physiological effects

and secondary metabolite (Saponin) production in Panax ginseng cell cultures. Ultrasound in

Medicine and Biology, 27, 1147-1152.

Page 119: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

100

Liu, R. H. (2002). Supplement quick fix fails to deliver. Food Technology International, 1, 71 -

72.

Liu, R. H., & Felice, D. L. (2007). Antioxidants and whole food phytochemicals for cancer

prevention. In F. Shahidi & C.-T. Ho (Eds.), Antioxidant Measurement and Application.

Washington DC: American Chemical Society.

Maccarone, E., Maccarrone, A., & Rapisarda, P. (1985). Stabilization of anthocyanins of blood

orange fruit juice. Journal of Food Science, 50, 901-904.

Manzocco, L., Anese, M., & Nicoli, M. C. (1998). Antioxidant properties of tea extracts as

affected by processing. Food Science and Technology-Lebensmittel-Wissenschaft &

Technologie, 31, 694-698.

Markakis, P. (1982). Stability of anthocyanins in foods. In P. Markakis (Ed.), Anthocyanins as

Food Colors (pp. 163 - 180 ). New York, NY: Academic Press.

Markakis, P., & Jurd, L. (1974). Anthocyanins and their stability in foods. CRC Critical Reviews

in Food Technology, 4, 437 - 456.

Markaris, P., Livingston, G. E., & Fellers, C. R. (1957). Quantitative aspects of strawberry

pigment degradation. Journal of Food Science, 22, 117-130.

Marsellés-Fontanet, Á. R., & Martín-Belloso, O. (2007). Optimization and validation of PEF

processing conditions to inactivate oxidative enzymes of grape juice. Journal of Food

Engineering, 83, 452-462.

Marti, N., Perez-Vicente, A., & Garcia-Viguera, C. (2002). Influence of storage temperature and

ascorbic acid addition on pomegranate juice. Journal of the Science of Food and Agriculture,

82, 217-221.

Page 120: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

101

Martinez-Tome, M., Murcia, M. A., Frega, N., Ruggieri, S., Jimenez, A. M., Roses, F., & Parras,

P. (2004). Evaluation of antioxidant capacity of cereal brans. Journal of Agricultural and

Food Chemistry, 52, 4690-4699.

Mazza, G., & Miniati, E. (1993). Anthocyanins in fruits, vegetables and grains. Boca Raton, FL:

CRC Press.

Miller, N. J., Diplock, A. T., & Riceevans, C. A. (1995). Evaluation of the total antioxidant

activity as a marker of the deterioration of apple juice oil storage. Journal of Agricultural and

Food Chemistry, 43, 1794-1801.

Mishra, D. K., Dolan, K. D., & Yang, L. (2008). Confidence intervals for modeling anthocyanin

retention in grape pomace during non-isothermal heating. Journal of Food Science, 73, E9-

E15.

Mizrahi, S. (1996). Leaching of soluble solids during blanching of vegetables by ohmic heating.

Journal of Food Engineering, 29, 153-166.

Mortensen, A., & Skibsted, L. H. (1997). Importance of carotenoid structure in radical-

scavenging reactions. Journal of Agricultural and Food Chemistry, 45, 2970-2977.

Mullen, W., Stewart, A. J., Lean, M. E. J., Gardner, P., Duthie, G. G., & Crozier, A. (2002).

Effect of freezing and storage on the phenolics, ellagitannins, flavonoids, and antioxidant

capacity of red raspberries. Journal of Agricultural and Food Chemistry, 50, 5197-5201.

Namiki, M. (1990). Antioxidants antimutagens in food. Critical Reviews in Food Science and

Nutrition, 29, 273-300.

Natella, F., Belelli, F., Ramberti, A., & Scaccini, C. (2010). Microwave and traditional cooking

methods: effect of cooking on antioxidant capacity and phenolic compounds content of seven

vegetables. Journal of Food Biochemistry, 34, 796-810.

Page 121: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

102

Nawar, W. W. (1996). Lipids. In R. O. Fennema (Ed.), Food Chemistry (pp. 225 - 320). New

York, NY: Marcel Dekker, Inc.

Ngo, T., Wrolstad, R. E., & Zhao, Y. (2007). Color quality of Oregon strawberries - Impact of

genotype, composition, and processing. Journal of Food Science, 72, C25-C32.

Nicoli, M. C., Anese, M., & Parpinel, M. (1999). Influence of processing on the antioxidant

properties of fruit and vegetables. Trends in Food Science & Technology, 10, 94-100.

Nicoli, M. C., Anese, M., Parpinel, M. T., Franceschi, S., & Lerici, C. R. (1997). Loss and/or

formation of antioxidants during food processing and storage. Cancer Letters, 114, 71-74.

Nindo, C. I., Sun, T., Wang, S. W., Tang, J., & Powers, J. R. (2003). Evaluation of drying

technologies for retention of physical quality and antioxidants in asparagus (Asparagus

officinalis, L.). Lebensmittel-Wissenschaft Und-Technologie-Food Science and Technology,

36, 507-516.

Ochoa, M. R., Kesseler, A. G., De Michelis, A., Mugridge, A., & Chaves, A. R. (2001). Kinetics

of colour change of raspberry, sweet (Prunus avium) and sour (Prunus cerasus) cherries

preserves packed in glass containers: light and room temperature effects. Journal of Food

Engineering, 49, 55-62.

Odriozola-Serrano, I., Soliva-Fortuny, R., & Martin-Belloso, O. (2009a). Impact of high-

intensity pulsed electric fields variables on vitamin C, anthocyanins and antioxidant capacity

of strawberry juice. Lwt-Food Science and Technology, 42, 93-100.

Odriozola-Serrano, I., Soliva-Fortuny, R., & Martin-Belloso, O. (2009b). Influence of storage

temperature on the kinetics of the changes in anthocyanins, vitamin C, and antioxidant

capacity in fresh-cut strawberries stored under high-oxygen atmospheres. Journal of Food

Science, 74, C184-C191.

Page 122: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

103

Oliveira, M. E. C., & Franca, A. S. (2002). Microwave heating of foodstuffs. Journal of Food

Engineering, 53, 347-359.

Oms-Oliu, G., Odriozola-Serrano, I., Soliva-Fortuny, R., & Martín-Belloso, O. (2009). Use of

Weibull distribution for describing kinetics of antioxidant potential changes in fresh-cut

watermelon. Journal of Food Engineering, 95, 99-105.

Onyeneho, S. N., & Hettiarachchy, N. S. (1992). Antioxidant activity of durum-wheat bran.

Journal of Agricultural and Food Chemistry, 40, 1496-1500.

Opalic, M., Domitran, Z., Komes, D., Belscak, A., Horzic, D., & Karlovic, D. (2009). The effect

of ultrasound pre-treatment and air-drying on the quality of dried apples. Czech Journal of

Food Sciences, 27, S297-S300.

Ozer, E. A., Herken, E. N., Guzel, S., Ainsworth, P., & Ibanoglu, S. (2006). Effect of extrusion

process on the antioxidant activity and total phenolics in a nutritious snack food.

International Journal of Food Science and Technology, 41, 289-293.

Patras, A., Brunton, N. P., Da Pieve, S., & Butler, F. (2009). Impact of high pressure processing

on total antioxidant activity, phenolic, ascorbic acid, anthocyanin content and colour of

strawberry and blackberry purees. Innovative Food Science & Emerging Technologies, 10,

308-313.

Patras, A., Brunton, N. P., O'Donnell, C., & Tiwari, B. K. (2010). Effect of thermal processing

on anthocyanin stability in foods; mechanisms and kinetics of degradation. Trends in Food

Science & Technology, 21, 3-11.

Perez-Jimenez, J., & Saura-Calixto, F. (2005). Literature data may underestimate the actual

antioxidant capacity of cereals. Journal of Agricultural and Food Chemistry, 53, 5036-5040.

Page 123: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

104

Piga, A., Del Caro, A., & Corda, G. (2003). From plums to prunes: Influence of drying

parameters on polyphenols and antioxidant activity. Journal of Agricultural and Food

Chemistry, 51, 3675-3681.

Plaza, L., Sanchez-Moreno, C., De Ancos, B., & Cano, M. P. (2006). Carotenoid content and

antioxidant capacity of Mediterranean vegetable soup (gazpacho) treated by high-

pressure/temperature during refrigerated storage. European Food Research and Technology,

223, 210-215.

Plaza, L., Sanchez-Moreno, C., Elez-Martinez, P., de Ancos, B., Martin-Belloso, O., & Cano, M.

P. (2006). Effect of refrigerated storage on vitamin C and antioxidant activity of orange juice

processed by high-pressure or pulsed electric fields with regard to low pasteurization.

European Food Research and Technology, 223, 487-493.

Portenlanger, G., & Heusinger, H. (1992). Chemical reactions induced by ultrasound and gamma

rays in aqueous solutions of L-ascorbic acid. Carbohydrate Research, 232, 291-301.

Porter, W. L., Levasseur, L. A., & Henick, A. S. (1977). Evaluation of some natural and

synthetic phenolic antioxidants in linoleic acid monolayers on silica. Journal of Food

Science, 42, 1533-1535.

Price, K. R., Bacon, J. R., & Rhodes, M. J. C. (1997). Effect of storage and domestic processing

on the content and composition of flavonol glucosides in onion (Allium cepa). Journal of

Agricultural and Food Chemistry, 45, 938-942.

Price, K. R., Casuscelli, F., Colquhoun, I. J., & Rhodes, M. J. C. (1998). Composition and

content of flavonol glycosides in broccoli florets (Brassica olearacea) and their fate during

cooking. Journal of the Science of Food and Agriculture, 77, 468-472.

Page 124: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

105

Price, K. R., Prosser, T., Richetin, A. M. F., & Rhodes, M. J. C. (1999). A comparison of the

flavonol content and composition in dessert, cooking and cider-making apples; distribution

within the fruit and effect of juicing. Food Chemistry, 66, 489-494.

Prior, R. L. (2003). Fruits and vegetables in the prevention of cellular oxidative damage. Am J

Clin Nutr, 78, 570S-578.

Prior, R. L., Cao, G. H., Martin, A., Sofic, E., McEwen, J., O'Brien, C., Lischner, N., Ehlenfeldt,

M., Kalt, W., Krewer, G., & Mainland, C. M. (1998). Antioxidant capacity as influenced by

total phenolic and anthocyanin content, maturity, and variety of Vaccinium species. Journal

of Agricultural and Food Chemistry, 46, 2686-2693.

Prior, R. L., Wu, X. L., & Schaich, K. (2005). Standardized methods for the determination of

antioxidant capacity and phenolics in foods and dietary supplements. Journal of Agricultural

and Food Chemistry, 53, 4290-4302.

Quaglia, G. B., Gravina, R., Paperi, R., & Paoletti, F. (1996). Effect of high pressure treatments

on peroxidase activity, ascorbic acid content and texture in green peas. Food Science and

Technology-Lebensmittel-Wissenschaft & Technologie, 29, 552-555.

Rajalakshmi, D., & Narasimhan, S. (1996). Food antioxidants: sources and methods of

evaluation. In Food Antioxidants: Technological, Toxicological, and Health Persepectives.

(pp. 65 - 157). New York, NY: Marcel Dekker, Inc.

Re, R., Bramley, P. M., & Rice-Evans, C. (2002). Effects of food processing on flavonoids and

lycopene status in a Mediterranean tomato variety. Free Radical Research, 36, 803-810.

Rehman, Z. U., Habib, F., & Shah, W. H. (2004). Utilization of potato peels extract as a natural

antioxidant in soy bean oil. Food Chemistry, 85, 215-220.

Page 125: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

106

Rein, M. J., & Heinonen, M. (2004). Stability and enhancement of berry juice color. Journal of

Agricultural and Food Chemistry, 52, 3106-3114.

Reyes, L. F., & Cisneros-Zevallos, L. (2007). Degradation kinetics and colour of anthocyanins in

aqueous extracts of purple- and red-flesh potatoes (Solanum tuberosum L.). Food Chemistry,

100, 885-894.

Reyes, L. F., Miller, J. C., & Cisneros-Zevallos, L. (2004). Environmental conditions influence

the content and yield of anthocyanins and total phenolics in purple- and red-flesh potatoes

during tuber development. American Journal of Potato Research, 81, 187-193.

Reyes, L. F., Villarreal, J. E., & Cisneros-Zevallos, L. (2007). The increase in antioxidant

capacity after wounding depends on the type of fruit or vegetable tissue. Food Chemistry,

101, 1254-1262.

Rodriguez-Saona, L. E., Giusti, M. M., & Wrolstad, R. E. (1998). Anthocyanin pigment

composition of red-fleshed potatoes. Journal of Food Science, 63, 458-465.

Rodriguez-Saona, L. E., Giusti, M. M., & Wrolstad, R. E. (1999). Color and pigment stability of

red radish and red-fleshed potato anthocyanins in juice model systems. Journal of Food

Science, 64, 451-456.

Rommel, A., Wrolstad, R. E., & Heatherbell, D. A. (1992). Blackberry juice and wine -

processing and storage effects on anthocyanin composition, color and appearance. Journal of

Food Science, 57, 385-391.

Rossi, M., Giussani, E., Morelli, R., Lo Scalzo, R., Nani, R. C., & Torreggiani, D. (2003). Effect

of fruit blanching on phenolics and radical scavenging activity of highbush blueberry juice.

Food Research International, 36, 999-1005.

Page 126: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

107

Rubinskiene, M., Viskelis, P., Jasutiene, I., Viskeliene, R., & Bobinas, C. (2005). Impact of

various factors on the composition and stability of black currant anthocyanins. Food

Research International, 38, 867-871.

Sablani, S. S., Andrews, P. K., Davies, N. M., Walters, T., Saez, H., Syamaladevi, R. M., &

Mohekar, P. R. (2010). Effect of thermal treatments on phytochemicals in conventionally and

organically grown berries. Journal of the Science of Food and Agriculture, 90, 769-778.

Sadilova, E., Carle, R., & Stintzing, F. C. (2007). Thermal degradation of anthocyanins and its

impact on color and in vitro antioxidant capacity. Molecular Nutrition & Food Research, 51,

1461-1471.

Sadilova, E., Stintzing, F. C., & Carle, R. (2006). Thermal degradation of acylated and

nonacylated anthocyanins. Journal of Food Science, 71, C504-C512.

Sales, J. M., & Resurreccion, A. V. A. (2010). Maximizing phenolics, antioxidants and sensory

acceptance of UV and ultrasound-treated peanuts. Lwt-Food Science and Technology, 43,

1058-1066.

Sanchez-Moreno, C., De Ancos, B., Plaza, L., Elez-Martinez, P., & Cano, M. P. (2009).

Nutritional approaches and health related properties of plant foods processed by high

pressure and pulsed electric fields. Critical Reviews in Food Science and Nutrition, 49, 552-

576.

Sanchez-Moreno, C., Plaza, L., de Ancos, B., & Cano, M. P. (2003). Effect of high-pressure

processing on health-promoting attributes of freshly squeezed orange juice (Citrus sinensis

L.) during chilled storage. European Food Research and Technology, 216, 18-22.

Sanchez-Moreno, C., Plaza, L., Elez-Martinez, P., De Ancos, B., Martin-Belloso, O., & Cano,

M. P. (2005). Impact of high pressure and pulsed electric fields on bioactive compounds and

Page 127: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

108

antioxidant activity of orange juice in comparison with traditional thermal processing.

Journal of Agricultural and Food Chemistry, 53, 4403-4409.

Sarma, A. D., Sreelakshmi, Y., & Sharma, R. (1997). Antioxidant ability of anthocyanins against

ascorbic acid oxidation. Phytochemistry, 45, 671-674.

Sarni-Manchado, P., Le Roux, E., Le Guerneve, C., Lozano, Y., & Cheynier, V. (2000). Phenolic

composition of litchi fruit pericarp. Journal of Agricultural and Food Chemistry, 48, 5995-

6002.

Sarni, P., Fulcrand, H., Souillol, V., Souquet, J. M., & Cheynier, V. (1995). Mechanisms of

anthocyanin degradation in grape must-like model solutions. Journal of the Science of Food

and Agriculture, 69, 385-391.

Schmidt, B. M., Erdman, J. W., & Lila, M. A. (2005). Effects of food processing on blueberry

antiproliferation and antioxidant activity. Journal of Food Science, 70, s389-s394.

Seabra, I. J., Braga, M. E. M., Batista, M. T. P., & de Sousa, H. C. (2010). Fractioned high

pressure extraction of anthocyanins from elderberry (Sambucus nigra L.) pomace. Food and

Bioprocess Technology, 3, 674-683.

Seeram, N. P., Bourquin, L. D., & Nair, M. G. (2001). Degradation products of cyanidin

glycosides from tart cherries and their bioactivities. Journal of Agricultural and Food

Chemistry, 49, 4924-4929.

Shahidi, F., & Zhong, Y. (2005)F. Shahidi (Ed.), Bailey's industrial oil and fat products (Vol. 1,

pp. 357 - 386). Hobiken, NY: John Wiley & Sons Ltd.

Shahidi, F., & Zhong, Y. (2007). Measurement of antioxidant activity in food and biological

systems. In Shahidi. F & Ho. C (Ed.), Antioxidant Measurement and Applications (pp. 36-

66): American Chemical Society, Wasington DC.

Page 128: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

109

Shih, M.-C., Kuo, C.-C., & Chiang, W. (2009). Effects of drying and extrusion on colour,

chemical composition, antioxidant activities and mitogenic response of spleen lymphocytes

of sweet potatoes. Food Chemistry, 117, 114-121.

Singh, N., & Rajini, P. S. (2004). Free radical scavenging activity of an aqueous extract of potato

peel. Food Chemistry, 85, 611-616.

Singh, R. P., Murthy, K. N. C., & Jayaprakasha, G. K. (2002). Studies on the antioxidant activity

of pomegranate (Punica granatum) peel and seed extracts using in vitro models. Journal of

Agricultural and Food Chemistry, 50, 81-86.

Skrede, G., Wrolstad, R. E., & Durst, R. W. (2000). Changes in anthocyanins and polyphenolics

during juice processing of highbush blueberries (Vaccinium corymbosum L.). Journal of

Food Science, 65, 357-364.

Song, H. P., Kim, D. H., Jo, C., Lee, C. H., Kim, K. S., & Byun, M. W. (2006). Effect of gamma

irradiation on the microbiological quality and antioxidant activity of fresh vegetable juice.

Food Microbiology, 23, 372-378.

Spanos, G. A., Wrolstad, R. E., & Heatherbell, D. A. (1990). Influence of processing and storage

on the phenolic composition of apple juice. Journal of Agricultural and Food Chemistry, 38,

1572-1579.

Srivastava, A., Harish, S. R., & Shivanandappa, T. (2006). Antioxidant activity of the roots of

Decalepis hamiltonii (Wight & Arn.). LWT - Food Science and Technology, 39, 1059-1065.

Stevenson, D. G., Inglett, G. E., Chen, D., Biswas, A., Eller, F. J., & Evangelista, R. L. (2008).

Phenolic content and antioxidant capacity of supercritical carbon dioxide-treated and air-

classified oat bran concentrate microwave-irradiated in water or ethanol at varying

temperatures. Food Chemistry, 108, 23-30.

Page 129: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

110

Stintzing, F. C., & Carle, R. (2004). Functional properties of anthocyanins and betalains in

plants, food, and in human nutrition. Trends in Food Science & Technology, 15, 19-38.

Stojceska, V., Ainsworth, P., Plunkett, A., Ibanoglu, E., & Ibanoglu, S. (2008). Cauliflower by-

products as a new source of dietary fibre, antioxidants and proteins in cereal based ready-to-

eat expanded snacks. Journal of Food Engineering, 87, 554-563.

Sun, T., & Powers, J. R. (2007). Antioxidants and antioxidant activities of vegetables. In F.

Shahidi. & C. T. Ho (Eds.), Antioxidant Measurement and Applicaitons (pp. 160 - 183).

Washington, DC: American Chemical Society.

Suslick, K. S. (1988). Ultrasounds: its chemical, physical and biological effects. New York, NY:

VHC Publishers.

Talcott, S. T., Brenes, C. H., Pires, D. M., & Del Pozo-Insfran, D. (2003). Phytochemical

stability and color retention of copigmented and processed muscadine grape juice. Journal of

Agricultural and Food Chemistry, 51, 957-963.

Talcott, S. T., Howard, L. R., & Brenes, C. H. (2000a). Antioxidant changes and sensory

properties of carrot puree processed with and without periderm tissue. Journal of

Agricultural and Food Chemistry, 48, 1315-1321.

Talcott, S. T., Howard, L. R., & Brenes, C. H. (2000b). Contribution of periderm material and

blanching time to the quality of pasteurized peach puree. Journal of Agricultural and Food

Chemistry, 48, 4590-4596.

Tanchev, S., & Ioncheva, N. (1976). Products of thermal degradation of the anthocyanins

cyanidin-3-glucoside, cyanidin-3-rutinoside and cyanidin-3-sophoroside. Food / Nahrung,

20, 889-893.

Page 130: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

111

Tanchev, S. S. (1972). Kinetics of the thermal degradation of anthocyanins of the raspberry.

Zeitschrift für Lebensmitteluntersuchung und -Forschung A, 150, 28-30.

Tanchev, S. S., & Yoncheva, N. (1974). Kinetics of thermal degradation of the anthocyanins

delphinidin-3-rutinoside and malvidin-3-glucoside. Food / Nahrung, 18, 747-752.

Temple, N. J. (2000). Antioxidants and disease: More questions than answers. Nutrition

Research, 20, 449-459.

Terao, J., Karasawa, H., Arai, H., Nagao, A., Suzuki, T., & Takama, K. (1993). Peroxyl radical

scavenging activity of caffeic acid and its related phenolic compounds in solution. Bioscience

Biotechnology and Biochemistry, 57, 1204-1205.

Thakur, B. R., & Arya, S. S. (1989). Studies on stability of blue grape anthocyanins.

International Journal of Food Science and Technology, 24, 321-326.

Tiwari, B. K., O'Donnell, C. P., & Cullen, P. J. (2009). Effect of sonication on retention of

anthocyanins in blackberry juice. Journal of Food Engineering, 93, 166-171.

Tsai, P. J., & Huang, H. P. (2004). Effect of polymerization on the antioxidant capacity of

anthocyanins in Roselle. Food Research International, 37, 313-318.

Tsai, P. J., Huang, H. P., & Huang, T. C. (2004). Relationship between anthocyanin patterns and

antioxidant capacity in mulberry wine during storage. Journal of Food Quality, 27, 497-505.

Tudela, J. A., Cantos, E., Espin, J. C., Tomas-Barberan, F. A., & Gil, M. I. (2002). Induction of

antioxidant flavonol biosynthesis in fresh-cut potatoes. Effect of domestic cooking. Journal

of Agricultural and Food Chemistry, 50, 5925-5931.

Turkmen, N., Sari, F., & Velioglu, Y. S. (2005). The effect of cooking methods on total

phenolics and antioxidant activity of selected green vegetables. Food Chemistry, 93, 713-

718.

Page 131: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

112

Uyan, S. E., Baysal, T., Yurdagel, O., & El, S. N. (2004). Effects of drying process on

antioxidant activity of purple carrots. Nahrung-Food, 48, 57-60.

Vagi, E., Rapavi, E., Hadolin, M., Peredi, K. V., Balazs, A., Blazovics, A., & Simandi, B.

(2005). Phenolic and triterpenoid antioxidants from Origanum majorana L. herb and extracts

obtained with different solvents. Journal of Agricultural and Food Chemistry, 53, 17-21.

van den Berg, R., Haenen, G. R. M. M., van den Berg, H., & Bast, A. (1999). Applicability of an

improved Trolox equivalent antioxidant capacity (TEAC) assay for evaluation of antioxidant

capacity measurements of mixtures. Food Chemistry, 66, 511-517.

Van Gorsel, H., Li, C., Kerbel, E. L., Smits, M., & Kader, A. A. (1992). Compositional

characterization of prune juice. Journal of Agricultural and Food Chemistry, 40, 784-789.

Vinson, J. A., Dabbagh, Y. A., Serry, M. M., & Jang, J. H. (1995). Plant flavonoids, especially

tea flavonols, are powerful antioxidants using an in vitro oxidation model for heart disease.

Journal of Agricultural and Food Chemistry, 43, 2800-2802.

Vinson, J. A., Hao, Y., Su, X., & Zubik, L. (1998). Phenol antioxidant quantity and quality in

foods: vegetables. Journal of Agricultural and Food Chemistry, 46, 3630-3634.

Viscidi, K. A., Dougherty, M. P., Briggs, J., & Camire, M. E. (2004). Complex phenolic

compounds reduce lipid oxidation in extruded oat cereals. Lebensmittel-Wissenschaft Und-

Technologie-Food Science and Technology, 37, 789-796.

Volden, J., Borge, G. I. A., Bengtsson, G. B., Hansen, M., Thygesen, I. E., & Wicklund, T.

(2008). Effect of thermal treatment on glucosinolates and antioxidant-related parameters in

red cabbage (Brassica oleracea L. ssp capitata f. rubra). Food Chemistry, 109, 595-605.

Von Elbe, J. H., & Schwartz, S. J. (1996). Colorants. In O. R. Fennema (Ed.), Food Chemistry

(Third ed., pp. 651 - 722). New York, NY: Marcel Dekker, Inc.

Page 132: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

113

Wang, H., Cao, G. H., & Prior, R. L. (1997). Oxygen radical absorbing capacity of anthocyanins.

Journal of Agricultural and Food Chemistry, 45, 304-309.

Wang, W. D., & Xu, S. Y. (2007). Degradation kinetics of anthocyanins in blackberry juice and

concentrate. Journal of Food Engineering, 82, 271-275.

White, P. J., & Xing, Y. (1997). Antioxidants from cereals and legumes. In F. Shahidi (Ed.),

Natural Antioxidants: Chemistry, Health Effects, and Applications. Champaign, IL: AOCS

Press.

Wolfe, K. L., & Liu, R. H. (2003). Apple peels as a value-added food ingredient. Journal of

Agricultural and Food Chemistry, 51, 1676-1683.

Wong, D. W. S. (1989). Mechanism and theory in food chemistry. New York: Van Nostrand

Reinhold.

Wright, J. S., Johnson, E. R., & DiLabio, G. A. (2001). Predicting the activity of phenolic

antioxidants: Theoretical method, analysis of substituent effects, and application to major

families of antioxidants. Journal of the American Chemical Society, 123, 1173-1183.

Wrolstad, R. E., Skrede, G., Lea, P., & Enersen, G. (1990). Influence of Sugar on Anthocyanin

Pigment Stability in Frozen Strawberries. Journal of Food Science, 55, 1064-&.

Wrona, M., Korytowski, W., Rózanowska, M., Sarna, T., & Truscott, T. G. (2003). Cooperation

of antioxidants in protection against photosensitized oxidation. Free Radical Biology and

Medicine, 35, 1319-1329.

Wu, X., Beecher, G. R., Holden, J. M., Haytowitz, D. B., Gebhardt, S. E., & Prior, R. L. (2004).

Lipophilic and hydrophilic antioxidant capacities of common foods in the United States.

Journal of Agricultural and Food Chemistry, 52, 4026-4037.

Page 133: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

114

Wu, X. L., Gu, L. W., Holden, J., Haytowitz, D. B., Gebhardt, S. E., Beecher, G., & Prior, R. L.

(2004). Development of a database for total antioxidant capacity in foods: a preliminary

study. Journal of Food Composition and Analysis, 17, 407-422.

Xu, Z., Wu, J., Zhang, Y., Hu, X., Liao, X., & Wang, Z. (2010). Extraction of anthocyanins from

red cabbage using high pressure CO2. Bioresource Technology, 101, 7151-7157.

Yen, G. C., Chen, H. Y., & Peng, H. H. (1997). Antioxidant and pro-oxidant effects of various

tea extracts. Journal of Agricultural and Food Chemistry, 45, 30-34.

Yue, X., & Xu, Z. (2008). Changes of anthocyanins, anthocyanidins, and antioxidant activity in

bilberry extract during dry heating. Journal of Food Science, 73, C494-C499.

Yun-Zhong, F., Sheng, Y., & Guoyao, W. (2002). Free radicals, antioxidants, and nutrition.

Nutrition (Burbank, Los Angeles County, Calif.), 18, 872-879.

Zabetakis, I., Koulentianos, A., Orruño, E., & Boyes, I. (2000). The effect of high hydrostatic

pressure on strawberry flavour compounds. Food Chemistry, 71, 51-55.

Zabetakis, I., Leclerc, N., & Kajda, P. (2000). The effect of high hydrostatic pressure on the

strawberry anthocyanins. Journal of Agricultural and Food Chemistry, 48, 2749-2754.

Zadernowski, R., Nowak-Polakowska, H., & Rashed, A. A. (1999). The influence of heat

treatment on the activity of lipo- and hydrophilic components of oat grain. Journal of Food

Processing and Preservation, 23, 177-191.

Zadernowskl, R., Nowak-Polakowska, H., & Rashed, A. A. (1999). The influence of heat

treatment on the activity of lipo-and hydrophilic components of oat grain. Journal of Food

Processing and Preservation, 23, 177-191.

Zern, T. L., Wood, R. J., Greene, C., West, K. L., Liu, Y. Z., Aggarwal, D., Shachter, N. S., &

Fernandez, M. L. (2005). Grape polyphenols exert a cardioprotective effect in pre- and

Page 134: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

115

postmenopausal women by lowering plasma lipids and reducing oxidative stress. Journal of

Nutrition, 135, 1911-1917.

Zhang, D. L., & Hamauzu, Y. (2004). Phenolics, ascorbic acid, carotenoids and antioxidant

activity of broccoli and their changes during conventional and microwave cooking. Food

Chemistry, 88, 503-509.

Zhang, D. L., & Quantick, P. C. (1997). Effects of chitosan coating on enzymatic browning and

decay during postharvest storage of litchi (Litchi chinensis Sonn.) fruit. Postharvest Biology

and Technology, 12, 195-202.

Zhang, Y., Liao, X. J., Ni, Y. Y., Wu, J. H., Hu, X. S., Wang, Z. F., & Chen, F. (2007). Kinetic

analysis of the degradation and its color change of cyanidin-3-glucoside exposed to pulsed

electric field. European Food Research and Technology, 224, 597-603.

Zhou, K. Q., & Yu, L. L. (2004). Effects of extraction solvent on wheat bran antioxidant activity

estimation. Lebensmittel-Wissenschaft Und-Technologie-Food Science and Technology, 37,

717-721.

Zielinski, H., & Kozlowska, H. (2000). Antioxidant activity and total phenolics in selected cereal

grains and their different morphological fractions. Journal of Agricultural and Food

Chemistry, 48, 2008-2016.

Zielinski, H., Kozlowska, H., & Lewczuk, B. (2001). Bioactive compounds in the cereal grains

before and after hydrothermal processing. Innovative Food Science & Emerging

Technologies, 2, 159-169.

Page 135: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

116

CHAPTER THREE

COLORED POTATOES (SOLANUM TUBEROSUM L.) DRIED FOR

ANTIOXIDANT-RICH VALUE-ADDED FOODS

ABSTRACT

Colored potatoes (Solanum tuberosum L.) are a significant source of antioxidants from

polyphenols, carotenoids, and ascorbic acid. In this study, retention of total antioxidants in fresh

colored potatoes and processed potato flakes prepared as potential ingredients for snack foods

was studied. Total antioxidant capacity, total phenolics and total anthocyanins were higher in

purple potato flesh compared to those from red, yellow and white potato cultivars. Peeled purple

potatoes were blanched and dehydrated by freeze-drying, drum-drying and refractance window-

drying to prepare potato flakes. Results showed no significant losses (P > 0.05) in total

antioxidant capacity and total phenolic content in flakes in all drying methods obtained under

study. However, 45, 41 and 23 % losses in total anthocyanins content were observed in potato

flakes after freeze-drying, drum-drying and refractive window-drying, respectively. Colored

potatoes could provide an excellent source of antioxidant-rich ingredient for the production of

nutritionally enhanced food products.

1. INTRODUCTION

Antioxidant phytochemicals in plants have recently attracted great attention from the

research community, food industry, and consumers. A large number of scientific papers report

the important role of phytochemicals in preventing many chronic diseases that are related to

oxidative stress caused by free radicals (van den Berg et al., 1999). Free radicals are associated

Page 136: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

117

with cancer, inflammation, atherosclerosis and ageing (Halliwell et al., 1992). Phytochemicals,

such as polyphenols in fruits and vegetables, possess high antioxidant activities that control the

generative oxidative reaction caused by reactive oxygen in living tissues (Kaur and Kapoor,

2001; Lachman et al., 2005). It has been reported that phenolic compounds including

anthocyanins (Figure 3.1) have potential to scavenge free radicals (Kalt et al., 1999).

Antioxidant capacity has been highly correlated to the amount of phenolic compounds and

anthocyanins present in dark colored fruits and vegetables (Brown et al., 2003; Prior et al.,

1998). Among common commercial fruits, blueberries have the largest antioxidant capacity

(Wang et al., 1996). It has also been reported that Andean purple corn and red-fleshed sweet

potatoes have even higher antioxidant capacity and antiradical activity than blueberries and

higher or similar content of phenolic compounds and anthocyanins (Cevallos-Casals and

Cisneros-Zevallos, 2003).

Potatoes (Solanum tuberosum L.) have traditionally been perceived by consumers as a

starchy food. Lewis, Walker, Lancaster & Sutton (1998) and more recently Jansen & Flamme

(2006) have reported phenolic, anthocyanin, and flavonoid contents of different varieties of

colored potato cultivars and breeding clones. Colored potatoes have attracted the attention of

investigators as well as consumers due to their antioxidant activities, taste, and appearance. The

antioxidant activity in colored potatoes are associated with the presence of polyphenols

anthocyanins, flavonoids, carotenoids, ascorbic acid, tocopherols, alpha-lipoic acid and selenium

(Lachman et al., 2005). Therefore, colored potatoes have the potential to be one of the richest

sources of antioxidants in the human diet. Brown (2005) reported 0.5 to 1 µg of carotenoids per

gram fresh weight (FW) in white, up to 20 µg per gram FW in deeply yellow to orange and 0.09

to 0.38 mg of total anthocyanins per gram FW in purple and red potato cultivars. The

Page 137: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

118

investigator also reported that an average potato contains 0.20 mg of vitamin C per gram FW that

contributes 13 % of total antioxidant activity in the tuber. Han et al., (2006a) reported that

purple potato extract prevented liver injury induced by D-galactosamine in rats. They also

reported that purple potato flakes have radical scavenging activities inhibiting linoleic acid

oxidation and improved antioxidant potential in rats by enhancing hepatic Mn-superoxide

dismutase (SOD), Cu/Zn-SOD and glutathione peroxidase (GSH-Px) mRNA expression.

Important food processing operations such as drying, cooking, and extrusion may affect the

retention of antioxidants in food matrices (Nicoli et al., 1999). However, other than vitamin C,

there is limited literature available on the effects of drying/thermal treatments on antioxidant

activities of potatoes.

Blanching and drying are the two most important unit operations in preparing shelf-stable

potato flakes as ingredients for commercial production of a wide range of foods products,

O+

OH

OH

OH

O G

Pelargonidin-3-glucoside

O+

OH

OH

OH

O G

OH

Cyanidin-3-glucoside

O+

OH

OH

OH

O G

OCH3

OCH3

Malvidin-3-glucoside

O+

OH

OH

OH

O G

OH

OH

Delphidin-3-glucoside

O+

OH

OH

OH

O G

OCH3

Petunidin-3-glucoside

O+

OH

OH

OH

O G

OCH3

Peonidin-3-glucoside

OH

Figure 3. 1. Structure of anthocyanins; G – Glucosides.

Page 138: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

119

including mashed potato and extruded snack foods. The objectives of this study were (i) to

quantify the total antioxidant capacity (TAC), total phenolics (TP), and total anthocyanins (TA)

in selected potato cultivars and (ii) to study the effect of blanching and consequent drying on the

retention of TAC, TP, and TA in purple potato cultivars.

2. MATERIALS AND METHODS

2.1. Raw Materials

Red (‗Red Rodeo‘) potatoes were purchased from an Oregon State (USA) potato grower.

Yellow (‗Yukon Gold‘) and white (‗Russet‘) potatoes were purchased from a local store in

Pullman, WA (USA). ‗Purple Majesty‘ potatoes were purchased in Fall of 2007 from Kiska

Farms, Burbank, WA (USA) and SLV Research Center, Colorado State University. The potatoes

were placed in a storage room at 4°C and 80 % relative humidity. Fresh purple potatoes procured

from Kiska Farms, were used for production of flakes by refractance window-drying comparison

with fresh red, yellow and white potatoes. Purple potatoes procured from Colorado State

University were used for the production of drum dried and freeze dried flakes.

2.2. Production of potato flakes

Based on the screening of potato varieties, flakes were prepared only from purple potatoes

for further investigation of the effects of blanching and dehydration on TAC, TP and TA. Stored

purple potatoes were peeled with an abrasive peeler for about 75 seconds. Peeled potatoes were

sliced with a mechanical slicer (Machine type RG-7, Ab Hallade Maskiner, Spanga, Sweden) to

6 mm thick before blanching in a steam blancher for 8 min to inactivate polyphenolic oxidase

(PPO) similar to the procedures reported for peroxidase inactivation (Reyes and Cisneros-

Zevallos, 2007). Blanched potato slices were cooled in ice water for 8 min and then pureed

Page 139: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

120

using a mixer (The Hobart Mfg Company, Troy, OH, USA). Additional water was added to the

puree to make it of uniform consistency, before placing the puree into the dryers.

2.2.1. Freeze-drying

Freeze-drying was selected as the reference method for drying. The potato puree was poured

on a tray to approximately 2 mm thick, frozen at -35°C for 1 h, placed into a freeze dryer and

dehydrated for 24 h at 3.33 Pa. The upper plate of the dryer was maintained at 20°C while the

condenser temperature was set at -64 °C. The dried samples were packed in polyethylene bags,

flushed with nitrogen, wrapped in aluminum sheets and stored at -30 °C for further analysis and

use.

2.2.2. Drum-drying

A 15.24cm × 20.32cm drum pilot scale counter rotating twin drum dryer (Blaw Knox Food

& Chemical Equipment Co., Buffalo, NY, USA) was used to dehydrate the puree. Pressurized

steam to the drums was maintained at 413 kPa corresponding to a saturation temperature of

water at 145 °C. The surface temperature of the drums was 135 - 138 °C. The gap between the

drums, rotating at 1.13 rpm, was set to 0.3 mm. The dried samples, packed in polyethylene bags,

were flushed with nitrogen wrapped in aluminum sheet and stored at -30 °C for further analysis

and use.

2.2.3. Refractance window-drying (RW)

A pilot scale refractance window dryer of an effective length and width of 1.83 × 0.60 m

developed by MCD Technologies, Inc. (Tacoma, Washington, USA) was used to dehydrate the

potato puree. The dryer consisted of a plastic conveyer belt rotating at speed of 1.04 m/min in

Page 140: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

121

contact with hot water circulated at 95 °C. Potato puree prepared using a Hobert mixer was

poured on a roller feeder, which deposited a thin layer (1 mm) of puree on the conveyor belt.

The conveyer belt transported the puree over a heating section for drying, then through a cooling

section, before the dried flakes were scraped from the end. Average air velocity over the

conveyer was 0.7 m/s and residence time of the puree on the belt was 1min 55 sec. Dried flakes

were packed in polyethylene bags, flushed with nitrogen, kept in aluminum bags, heat sealed and

stored at -30 °C for further analysis and use.

2.3. Chemical Analysis

Moisture content of the raw cultivars and dried flakes was determined by the vacuum oven

method (AOAC, 1995).

Folin-Ciocalteu reagent, potassium chloride, sodium acetate, Trolox (6-Hydroxy-2,5,7,8-

tetramethylchromane-2-carboxylic acid), DPPH (2, 2-diphenyl-1-picrylhydrazyl) and gallic acid

were purchased from Sigma-Aldrich. Laboratory grade methanol and ethanol were used in

extraction and preparation of samples.

2.3.1. Total antioxidant capacity (TAC)

Total antioxidant capacities of raw potatoes were quantified using the DPPH (2, 2-diphenyl-

1-picrylhydrazyl) assay (Brandwilliams et al., 1995). DPPH, a stable radical deep purple in

color, is reduced in the presence of antioxidants decolorizing the solution. The loss of color

results in a decrease in the absorbance intensity, thus providing a basis for measurement of

antioxidant activities in the extracts. Thirty gram of potatoes peeled by an abrasive peeler

(Model 15A, MJM Mfg Co., Culver city, CA, USA) were chopped and homogenized with 100

ml of HPLC grade methanol to a uniform consistency by a homogenizer (Omni Mix

Page 141: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

122

Homogenizer, Omni International, Waterbury, CT, USA). The samples were centrifuged

(Beckman J2-HS Centrifuge, USA) at 30000g at 4 °C for 20 min and the supernatants stored at -

20 °C for further analysis. A 6 × 10-5

M DPPH solution was prepared in methanol and stored at -

20 °C for analysis.

Stored supernatants were diluted two fold with methanol and 0.05 ml of diluted supernatants

was added to 1.95 ml of DPPH solution in a cuvette. Aliquots in cuvettes were covered with

Parafilm and vortexed with a mini vortexer (MV1, Wilmington, NC, USA) before taking

absorbance readings at selected times with a Ultraspec 4000 UV/visible spectrophotometer

(Pharmacia Biotech, Cambridge, England) until absorbance values reached a plateau. The

spectrophotometer was blanked with methanol. DPPH without sample was taken as control. For

each sample measured, the percentage of DPPH remaining was calculated as follows:

100)(0@

@

Tabsorbance

tTabsorbance

remainingDPPH

DPPHDPPH

where DPPH absorbance@T=t is the absorbance of DPPH at time t min and DPPH absorbance@T=0 is the

absorbance of DPPH at zero min. Trolox (6-Hydroxy-2,5,7,8-tetramethylchromane-2-carboxylic

acid) was used as a standard. The absorbance readings at 2h after which there was no additional

change were used to calculate TAC from the trolox standard curve. All the values of TAC were

expressed as micrograms of trolox equivalent per gram of dry weight sample (µg TE/g DW) ±

SD for three replications.

Total antioxidant capacity in purple dried flakes extracts were prepared from 10 g of dried

flakes initially rehydrated with 40 ml water and blended with 40 ml HPLC grade methanol. The

final volume of the mixture was made to 100 ml with aqueous methanol (50:50 v/v). The

mixture was centrifuged at 30,000g at 4 °C for 20 min and supernatants stored at -20 °C for

further analysis.

Page 142: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

123

2.3.2. Total phenolics (TP)

Total phenolics were determined from extracts prepared for TAC using an Folin-Ciocalteau

colorimetric method as described by Swain & Hillis (1959) and Singleton & Rossi (1965).

Supernatants were diluted four-fold with methanol. To 0.5 ml aliquots 8 ml of deionized water

was added followed by 0.5 ml of 0.25 N Folin-Ciocalteu (FC) reagent, the samples were

vortexed, kept for 3 min and 1 ml of 1 N sodium carbonate was added and the mixture vortexed.

The samples were kept at room temperature for 2 hours before taking readings at 725 nm in the

UV Spectrophotometer. One-half milliliter methanol was treated in the same way as the diluted

samples and used as a blank. The TP analyses were triplicated and means (± SD) reported as

micrograms gallic acid equivalent per gram of dry weight sample (µg GAE/g DW) from a

standard curve developed for gallic acid.

2.3.3. Total anthocyanins

Potatoes were extracted for anthocyanin analysis using the procedure of Fuleki & Francis

(1968) with modifications. Raw stored potatoes were peeled with an abrasive peeler for 75 sec

and the surface moisture removed with tissue paper. Peeled potatoes were chopped manually to

small pieces and 30 g homogenized with 150 ml of acidified ethanol (95% ethanol/1.5 N HCl,

85:15 v/v) to a uniform consistency. The samples were kept for 10 min at room temperature and

extracts were decanted into a beaker. The residues were washed with 100 ml of acidified ethanol

and extracts from first and second washings and residues were mixed, covered with Parafilm,

and stored for 90 min at 4 °C. The samples were centrifuged at 23000 x g at 4 °C for 15 min

and supernatants stored at -20 °C for further analysis.

Quantification of anthocyanins was carried out using the pH differential method described

Page 143: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

124

by Giusti & Wrolstad (2001). Diluted aliquots containing 0.2 ml of anthocyanin extract and 1.8

ml buffer were prepared in 2 ml cuvettes with KCl buffer (pH 1.0) or sodium acetate buffer (pH

4.5), respectively. The diluted aliquots were vortexed with a mini vortexer and equilibrated for

15 min at room temperature. Absorbance readings were taken at maximum wavelength (λ-max)

of 535nm for purple cultivars, 515nm for red and yellow cultivars and 700 nm for correcting for

turbidity (Reyes and Cisneros-Zevallos, 2003) in a UV/visible spectrophotometer, previously

blanked with distilled water. Total anthocyanins in purple cultivars were quantified by

considering malvidin-3-glucoside (Han et al., 2006b; Jansen and Flamme, 2006; Lewis et al.,

1998) as the major anthocyanin with MW =718.5 g/mol and molar extinction coefficient of

30200 l/cm per mol. Pelargonidin-3-glucosides having e = 27300 l/cm per mol and MW = 486.5

g/mol (Fuleki and Francis, 1968) were considered as the major anthocyanins in red cultivars,

whereas, cyanidin-3-glucosides were considered as the major anthocyanin in yellow cultivars.

Total anthocyanins content in potato extracts were calculated according to the formula (Giusti

and Wrolstad, 2001):

de

DFMWAlmgC

*

**)/(

where A = absorbance of the sample, MW = molecular weight of major anthocyanin, DF =

dilution factor, e = molar extinction coefficient of major anthocyanin and d = path length of the

cuvette (1 cm). The absorbances of the samples were calculated as

5.4700max0.1700max )()( pHpH AAAAA . Total anthocyanins were expressed as mg of major

anthocyanin per gram of DW sample mean ± SD for three replications.

For total anthocyanins determination in dehydrated purple potato flakes, 5 g of flakes were

rehydrated with 40 ml water and blended with 40 ml of acidified ethanol (95 % ethanol/1.5 N

Page 144: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

125

HCl (85:15 v/v)). The samples were thoroughly mixed by stirring manually and decanted to a

separate beaker. The residues were washed again, the extracts decanted and mixed with the first

decanted extracts. The total extracts were made to a total volume of 100 ml with 50:50 (v/v)

water: acidified ethanol and centrifuged at 23000g at 4 °C for 15 min. The supernatants were

stored at -20 °C for further analysis.

2.4. Color determination

Color of raw cultivars and dehydrated flakes was determined using a color meter (Minolta

Chroma CR200, Minolta Co., Osaka, Japan) with L*a*b* values depicting brightness,

greenness/redness and blue/yellowness, respectively. The color meter was calibrated with white

and black standards provided by the manufacturer. The hue angle h° [h°=arctan(b*/a*)] and

Chromaticity C* [C*=(a*2+b*

2)1/2

] were computed from a* and b*. Color differences between

dehydrated samples and raw samples were expressed as ∆E* where ∆E*= [(∆L*) 2

+ (∆a*) 2

+

(∆b*) 2]

1/2. Peeled potatoes were sliced to measure the color attributes in raw cultivars.

Dehydrated flakes were ground with a mortar and pestle and covered with an ultra-thin

transparent polyethylene sheet before measurements were taken.

2.5. Statistical analysis

Color, TAC, TP and TA data for raw and dehydrated potato samples were analyzed with

SAS software (version 9.1). Significant differences among treatments were determined using

ANOVA followed by Tukey‘s pair-wise comparisons at 95 % confidence level (P≤0.05).

Triplicate (n=3) data obtained from the different analyses were reported as arithmetic mean and

standard deviation.

Page 145: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

126

3. RESULTS AND DISCUSSION

3.1. Moisture content, total antioxidant capacity (TAC), total phenolics (TP) and total

anthocyanins (TA) in raw potatoes

Average moisture contents of the selected raw potato tubers were 77% (wet basis). There was

no significant difference (P > 0.05) in moisture contents between cultivars. Also, no significant

differences (P > 0.05) were observed in TAC, TP, and TA contents between purple potatoes

procured from Washington and Colorado state locations. However, it was observed that purple

potatoes contained more TAC, TP, and TA than red, yellow and white potatoes cultivars (Table

3.1). The TAC among the different sample cultivars ranged from 2403 to 9605 µg TE/g DW.

Samples from purple cultivars without skin had the highest TAC with 9605 µg TE/g DW.

White, red and yellow had similar TAC of ~2400–2500 µg TE/g DW. Reddivari, Hale & Miller

(2007) reported that antioxidant activity of whole potato cultivars varied from 157 to 832 µg

TE/g of FW using DPPH assay and 810 to 1622 µg TE/g of FW using ABTS assay. A higher

antioxidant capacity of whole purple and red-fleshed potatoes ranging from 513 to 1426 µg

TE/g of FW has also been reported (Reyes et al., 2005).

Table 3. 1. Total antioxidant capacity by DPPH assay, total phenolics content and total

anthocyanins of selected raw potato cultivars expressed as quantity per gram of dry weight

sample (n=3).

Cultivars Dry matter

(g/g of sample)

Total Antioxidant

Capacity

(µg Trolox/g DW)

Total Phenolics

(µg GAE /g DW)

Total

Anthocyanins

(mg/g DW)

Purple 0.23 9605 ± 404a 3347 ± 198

a 1.08 ± 0.09

a

Red 0.24 2542 ± 120 b 1457 ± 240

c 0.014 ± 0.004

b

Yellow 0.23 2403 ± 90 b 1425 ± 200

c 0.014 ± 0.004

b

White 0.23 2518 ± 471 b 2096 ± 482

b -

Significant differences within the values in the same column are indicated by different

superscript letters (P < 0.05).

Page 146: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

127

The amount of TP varied from 1425 to 3347 µg GAE/g DW among the cultivars under study.

The TP content in the flesh of purple potato cultivars was significantly higher (3347 µg GAE/g

DW) than those obtained in white (2096 µg GAE/g DW), red (1457 µg GAE/g DW), and yellow

(1425 µg GAE/g DW) cultivars. Andre et al.(2007) have previously reported TP content in white

and colored potato cultivars in the range of 2360 to 12370 μg GAE/g FW.

Total anthocyanins content in the flesh of purple potato cultivars (from Washington state)

was 1.08 mg MV-3-GLU /g DW, while the red and yellow cultivars contained smaller amount of

TA (Table 3.1). Similarly, Reyes et al. (2005) reported that the TA content in the flesh of purple

potatoes ranged from 0.11 to 1.74 mg CY-3-GLU / g FW and from 0.21 to 0.55 mg CY-3-GLU /

g FW in the flesh of red potatoes. Dark purple-black tubers having 2 to 5 mg of MV-3-GLU /g

FW and red-fleshed potato cultivars with 0.02 to 0.40 mg pelargonidin-3- glucoside/g FW was

reported by Rodriguez-Saona, Giusti & Wrolstad (1998). The differences in TAC, TP and TA

content for the studied cultivars could be due to environmental conditions, growing location,

harvesting date, maturity, sample preparation, and extraction methods used for their evaluation.

3.2. Effects of blanching on TAC, TP and TA

Preliminary refractance window-drying (RWD) of potato puree from purple cultivars without

blanching caused complete loss of purple color in the flakes (Figure 3.2). This loss in color may

be due to PPO activity (Reyes and Cisneros-Zevallos, 2007). Color loss in purple potato puree

was observed, once tubers were cut into slices and ground. The color of the puree turned brown

and became colorless after RW dehydration. Therefore, in an effort to prevent enzymatic

browning to occur, we steam blanched purple potato slices for 8 minutes prior to processing

them into puree for subsequent drying.

Page 147: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

128

Purple potato product obtained from steam blanched puree, prior to drum-drying and freeze-

drying, retained about 90 % of TA compared to raw puree (Table 3.2). Reyes et al. (2007)

reported that red and purple fleshed potatoes of ~5 mm thickness steam blanched for 3 min had

98 % reduction in peroxidase activity. Retention of 23 % of TA in blueberries was reported by

Rossi, Giussani, Morelli, Lo Scalzo, Nani & Torreggiani (2003) after blanching compared to

12% without blanching, before processing into juice. Using a reverse-phase HPLC separation

system for separating individual anthocyanins, Rossi et al. (2003) observed that recovery of

anthocyanin from blueberries after blanching was maximum for delphinidin-3-arabinoside

(1936%) followed by petunidin (586 %) and cyanidin (191 %) glucosides, whereas malvidin

glucosides showed the least recovery (143 %). The investigators also reported that the minor

recovery of malvidin glucosides could be due to the structure of the anthocyanin having a single

hydroxyl group on its phenolic ring that was least affected by PPO compared to the other

anthocyanins. In the present study, puree from purple potatoes without blanching prior to RWD

did not retain any color. The complete loss of color in the puree could be due to the activity of

PPO. On the other hand, retention of anthocyanins and color in potato puree after blanching

Figure 3. 2. Color of potato flakes; (A) without blanching; (B) with blanching

Page 148: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

129

could be due to inactivation of PPO and subsequent reduction in enzymatic anthocyanin

degradation.

Blanching of sliced potatoes also showed a significant increase (P < 0.05) in TAC (75 %)

and TP (108 %) in the puree compared to unblanched puree (Table 3.2). Wang (2002) reported

an increase in total antioxidant activities of 7 and 26 % after water and microwave blanching of

asparagus spears, respectively. The increases in the TAC and TP content after blanching could

be due to the greater extraction yield by opening up the cell and increasing cell permeability to

bound phenolics (Kalt et al., 2000). Release of lycopene in tomatoes (Dewanto et al., 2002a),

phenolics in apple peels (Wolfe and Liu, 2003) and ferulic acid in corn (Dewanto et al., 2002b)

have also been reported during heat processing. These reports agreed with the results of the

present study with regard to the beneficial effect of blanching of retention of TA activity in the

final product.

Table 3. 2. Total antioxidant capacity (TAC) by DPPH assay, total phenolics (TP) and total

anthocyanins (TA) of raw, blanched and dried potato flakes from purple cultivars. TAC, TP &

TA of blanched and dried flakes were compared with raw cultivars and expressed in dry weight

basis (n=3).

Total Antioxidant Capacity

(µg Trolox/g DW)

Total Phenolics

(µg GAE /g DW)

Total Anthocyanins

(mg/g DW)

Raw (CO) 8787 ± 630b 3835 ± 296

c 1.74 ± 0.16

a

Blanched 15358 ± 2948a 7993 ± 488

a 1.58 ± 0.08

a

Drum dried 7021 ± 911 b 4021 ± 136

c 1.03 ± 0.02

b

Freeze dried 8151 ± 37 b 3950 ± 124

c 0.96 ± 0.17

b

Raw* (WA) 9605 ± 405b* 3347 ± 198

c* 1.08 ± 0.09

a*

Refractance

window dried 7331 ± 246

b* 4680 ± 120

b* 0.83 ± 0.01

b*

* Cultivar from a different location and storage time was used for Refractance Window-drying,

CO: Colorado, WA: Washington. Significant differences within the values in the same column

are indicated by different superscript letters (P < 0.05).

Page 149: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

130

3.3 Total antioxidant capacity (TAC), total phenolics (TP) and total anthocyanins (TA) in

dehydrated potato flakes

All of the dehydrated flakes had similar moisture contents: 6.0 % (wet basis) for freeze dried

flakes; 5.9 % (wet basis) for RW dried flakes; and 5.2 % (wet basis) for drum dried flakes.

There was no significant difference (P > 0.05) among the moisture contents of the flakes. The

TAC content in freeze dried, refractance window dried, and drum dried flakes were 8151, 7331,

and 7021 µg TE/g DW, respectively. There was no significant change (P > 0.05) in TAC content

of dry flakes compared to raw sample for all drying methods used in this study. The TP content

in flakes prepared by drum-drying and freeze-drying were 4021 and 3950 µg GAE/g DW,

respectively. Also, significant changes were observed in TP, DD, and FD. However, there was a

significant increase (P < 0.05) of these components observed in the flakes prepared by

refractance window-drying (4680 µg GAE/g DW). Flakes prepared by drum-drying, freeze-

drying and refractance window-drying had TA contents of 1.03, 0.96 and 0.83 mg MV-3-GLY/g

DW, respectively. There were significant losses of 23-45 % TA contents observed in potato

flakes compared to raw samples (Table 3.2).

Many researchers have reported that processing causes major losses in the antioxidant

concentration of fruits and vegetables, mainly due to enzymatic and chemical oxidation of

antioxidants. However, most of the literature focuses on the degradation of vitamin C in fruits

and vegetables during blanching, dehydration or heating (Van den Broeck et al., 1998) and very

few data are available on polyphenols and other compounds that contribute to the antioxidant

activities in fruits and vegetables. In the present study, no significant change (P > 0.05) in TAC

was observed in the dehydrated potato flakes compared to raw samples, regardless of drying

method used. Blessington et al. (2007) reported an increase in the content of carotenoids and

Page 150: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

131

DPPH antioxidant activities during frying and microwave-drying of potatoes. Dewanto et al.

(2002b) also reported an increase of 44 % in TA activities in heat processed sweet corn (115 °C

for 25 min) despite a 25 % loss in vitamin C. There is no significant change (P > 0.05) observed

in antioxidant activities during freeze-drying of apple peels (Wolfe and Liu, 2003), baking (at

177 °C for 20 min) of wheat bran (Singh et al., 2007) and canning of chick pea proteins

processed at 121 °C for 20 min (Arcan and Yemenicioglu, 2007). Retention of antioxidant

activities in processed fruits and vegetables have been attributed to protein hydrolysis, Maillard

reaction and fermentation process (Nicoli et al., 1999). In the present study, the observed

retention of TAC might be due to a combination of the natural phytochemicals present in the

potato and Maillard Reaction Products (MRPs) that contribute to the overall antioxidant

activities of the potato flake. Reports by Nicoli et al., (1999) also supports that Maillard

products formed during processing contributed to the formation of antioxidants in roasted coffee

enhancing total antioxidant capacity of the product.

Phenolics were also retained after dehydration of purple potatoes by DD and FD and a

significant increase was observed due to RWD when compared to raw tubers. There is limited

literature available to compare the effect of drying and other thermal treatments on total

phenolics in potatoes. In a study on antioxidant values of phenolic acids in potatoes, Friedman

(1997) reported complete loss of phenolic acids by cooking. In contrast, Blessington (2005)

observed higher TP content in potatoes processed by microwaving, frying and baking than by

boiling. Wolfe et al., (2003) reported a significant increase of phenolic contents in freeze dried

apple peels compared to fresh peels. The retention and increase in TP during blanching and

dehydration may be attributed to opening of the cell matrix and release of bound phenolics.

Blanching followed by drying of peeled purple potatoes reduced the content of total

Page 151: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

132

anthocyanins by 20 to 40 % in the flakes. Singh et al., (2007) reported complete loss of

anthocyanins in purple wheat bran during baking. Other researchers reported either an increase

or no change in TA content during drying of carrots (Uyan et al., 2004). It has been well studied

that temperature, pH, oxygen and light affect the stability of anthocyanins. Reyes et al., (2007)

reported a first order thermal degradation of anthocyanin extracts from colored potatoes at

different pH. Anthocyanins lose the glucosides by deglycosylation to form chalcones that further

degrades to acids and aldehydes (Sadilova et al., 2006).

3.4. Correlation between TP and TAC

A small negative linear correlation was found between TP and TAC (r2= -0.119) in dried

potato flakes indicating little relationship between these parameters. Therefore, these data do not

suggest that TP are primarily responsible for the total antioxidant capacity in potato flakes

dehydrated by DD, FD and RW. However, in colored (purple, red, yellow) and white potato

cultivars a large positive correlation between TP and TAC (r2 = 0.886) was observed (Figure

3.3). Reyes et al. (2003) reported that wound-induced phenolic compounds in potatoes were

positively correlated to antioxidant activities. Many researchers also reported a high positive

acid equivalent) in whole raw potatoes (Reddivari et al., 2007; Reyes et al., 2005). In contrast,

Hale (2003) observed that concentration of phenolic acid contributed very little to antioxidant

correlation between antioxidant capacity and total phenolics content (in terms of chlorogenic

activities in raw potatoes (r2=0.18). There has been little information reported on the correlation

between TP and TAC in processed potatoes. The observed difference in correlation between raw

tubers and dehydrated potato flakes suggests that MRPs formed during processing may

contribute significantly to the total antioxidant capacity in the extracts from dehydrated flakes.

Page 152: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

133

3.5. Color in raw cultivars and dehydrated flakes

Visual color attributes of the raw cultivars expressed in terms of lightness, hue, and

chromaticity are shown in Table 3.3. Lightness (L*) of white potato was significantly higher (P

< 0.05) among the cultivars tested and purple potato had the smallest L* value (P < 0.05), while

the red and yellow cultivars had similar L* values. Chromaticity values followed the same

pattern observed for lightness that is white potatoes showed the highest and purple potatoes the

smallest color saturation. Red and yellow cultivars were not different in lightness or

Table 3. 3. Color attributes of raw potato cultivars (n=3)

Cultivars Lightness

(L*)

Chromaticity

(C*)

Hue angle

(h°)

Purple 19.3 ± 2.4c 8.2 ± 0.5

c 320.2 ± 0.9

a

Red 65.6 ± 2.1b 24.3 ± 1.7

a 271.7 ± 0.2

b

Yellow 63.8 ± 2.2 b

25.0 ± 0.5 a 89.6 ± 0.3

c

White 69.4 ± 1.6a 15.6 ± 0.6

b 271.6 ± 0.1

b

Significant differences within the values in the same column are indicated by different

superscript letters (P < 0.05)

R2 = 0.8858

1000

3000

5000

7000

9000

1000 2000 3000 4000 5000

TP (µgGAE/g DW)

TA

C (

µg

TE

/g D

W)

Figure 3. 3 Correlations between total antioxidant capacity by DPPH assay and total

phenolics in colored (purple, red, yellow) and white potato cultivars.

Page 153: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

134

chromaticity, but were significantly different in hue angle (P < 0.05). The purple potato cultivar

was significantly (P < 0.05) different in all color attributes from other cultivars.

Upon dehydration of white and colored potato samples, a significant increase in lightness

was observed for all samples under different dehydration processes (Table 3.4). Freeze dried

and RW dried flakes had similar L* values and were significantly brighter (P < 0.05) than drum

dried flakes. Similar results were reported on heating of black carrot, strawberry and elderberry

extracts (Sadilova et al., 2006). Color purity of the freeze dried flakes expressed as chroma was

similar to the raw potatoes whereas significantly higher values (P < 0.05) were observed for DD

and RW dried flakes. The higher saturation of color observed in the DD and RW dried samples

could be due to exposure of potato puree to air and light for shorter times during these drying

processes than samples processed by freeze-drying technology. Since, the residence time of

potato puree on DD and RW driers was around 53 sec and 115 sec, respectively. Whereas, puree

samples were kept for 24 h in the freeze dryer; this provided a higher exposure time under this

drying condition. Similarly, Sadilova et al. (2006) observed a significant decrease in chroma of

black carrot, strawberry and elderberry extracts after heating 7 hours depicting dullness/less

saturated color in the samples.

Hue values of all the potato flakes dehydrated under the different drying processes were

significantly different (P < 0.05) from each other (Table 3.4). Flakes from the DD were higher

in hue value depicting more reddish than RW and freeze dried flakes. Also, the color of raw

purple cultivars was shifted towards red with higher hue value than that of dehydrated flakes.

Reyes et al., (2007) reported an increase in hue-value in red and purple-fleshed potato extracts

when exposed to 98 °C. On the other hand, Sadilova et al. (2006) reported an initial decrease in

hue-values after 3h of heating of black carrot, strawberry and elderberry extracts at 98 °C,

Page 154: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

135

Table 3. 4. Color attributes of purple dehydrated flakes (n=3)

Drying process Lightness

(L*)

Chromaticity

(C*)

Hue angle

(h°)

Color difference

(∆E*)

Raw 19.3 ± 2.4c 8.2 ± 0.5

b 320.2 ± 0.9

a 0.0

Freeze dried 54.0 ± 2.a 7.2 ± 0.6

b 270.5 ± 0.1

d 35.3 ± 0.99

a

Drum dried 45.7 ± 1.2 b

12.2 ± 1.0 a 296.2 ± 0.3

b 27.0 ± 3.37

b

RW dried 55.3 ± 0.9a 12.2 ± 0.7

a 293.6 ± 0.2

c 36.5 ± 2.01

a

Significant differences within the values in the same column are indicated by different

superscript letters (P < 0.05).

followed by an increase in the hue value more than that of unheated extracts. The overall color

attributes of the dehydrated potato flakes were determined by color differences compared to the

raw potato sample. Lesser differences in color indicate better stability of pigments in the

samples. Color differences (∆E) in DD flakes were significantly less (P < 0.05) that those

obtained from FD and RW dried flakes, indicating more preservation of the original color in

flakes dried by DD technology. Despite the observed variation in color difference between DD

and FD flakes, the concentration of anthocyanins in those flakes was not significantly different

(P > 0.05).

4. CONCLUSION

Evaluation of antioxidant compounds present in purple, red, yellow, and white potatoes

showed that purple potato cultivars contain significantly higher total antioxidants, total phenolic

and total anthocyanins than other potato cultivars. Dry flakes prepared from steam-blanched

purple potatoes retained the purple color. However, those flakes obtained from purple potatoes

without previous blanching lost the purple color. Therefore, blanching was demonstrated to be an

important operation for treating purple potatoes before dehydration. Different drying

technologies (drum-drying, freeze-drying and refractance window-drying) used to prepare

dehydrated purple potato flakes did not significantly change (P > 0.05) total antioxidant content.

Page 155: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

136

Similar results were also observed in total phenolics content in the dehydrated purple potato

flakes prepared by freeze-drying and drum-drying. Conversely, a significantly higher (P < 0.05)

total phenolic content was obtained by refractance window-drying technology of potato flakes.

Losses of 23 to 45 % in total antioxidant content were observed in dehydrated potato flakes

processed under all drying methods. The results suggest that processing colored potatoes into

value-added antioxidant-rich ingredients may contribute to the production of healthy snacks and

other foods. A detailed kinetic study of antioxidants during processing of potatoes is needed to

better understand the behavior of antioxidants in food systems.

ACKNOWLEDGEMENTS

We acknowledge the financial support from the Washington State Potato Commission,

Moses Lake, WA, USA and partial support from the Washington State University Agricultural

Research Center.

Page 156: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

137

REFERENCES

Andre, C.M., Ghislain, M., Bertin, P., Oufir, M., Herrera, M.D., Hoffmann, L., Hausman, J.F.,

Larondelle, Y., Evers, D. 2007. Andean potato cultivars (Solanum tuberosum L.) as a source

of antioxidant and mineral micronutrients. Journal of Agricultural and Food Chemistry. 55,

366-378.

AOAC. 1995. Official methods of analysis (16th. ed.). Association of Official Analytical

Chemistry, Washington DC.

Arcan, I., Yemenicioglu, A. 2007. Antioxidant activity of protein extracts from heat-treated or

thermally processed chickpeas and white beans. Food Chem. 103, 301-312.

Blessington, A., Miller, J.C., Nzaramba, M.N., Hale, A.L., Redivari, L., Scheuring, D.C.,

Hallman, G.J. 2007. The effects of low-dose gamma irradiation and storage time on

carotenoids, antioxidant activity, and phenolics in the potato cultivar Atlantic. American

Journal of Potato Research. 84, 125-131.

Blessington, T. 2005. The effects of cooking, storage and ionizing irradiation on carotenoids,

antioxidant activity and phenolics in potato. Texas A&M, College station.

Brandwilliams, W., Cuvelier, M.E., Berset, C. 1995. Use of a Free-Radical Method to Evaluate

Antioxidant Activity. Food Science and Technology-Lebensmittel-Wissenschaft &

Technologie. 28, 25-30.

Brown, C.R. 2005. Antioxidants in potato. American Journal of Potato Research. 82, 163-172.

Brown, C.R., Wrolstad, R., Durst, R., Yang, C.P., Clevidence, B. 2003. Breeding studies in

potatoes containing high concentrations of anthocyanins. American Journal of Potato

Research. 80, 241-249.

Page 157: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

138

Cevallos-Casals, B.A., Cisneros-Zevallos, L. 2003. Stoichiometric and kinetic studies of

phenolic antioxidants from Andean purple corn and red-fleshed sweetpotato. Journal of

Agricultural and Food Chemistry. 51, 3313-3319.

Dewanto, V., Wu, X.Z., Adom, K.K., Liu, R.H. 2002a. Thermal processing enhances the

nutritional value of tomatoes by increasing total antioxidant activity. Journal of Agricultural

and Food Chemistry. 50, 3010-3014.

Dewanto, V., Wu, X.Z., Liu, R.H. 2002b. Processed sweet corn has higher antioxidant activity.

Journal of Agricultural and Food Chemistry. 50, 4959-4964.

Friedman, M. 1997. Chemistry, biochemistry, and dietary role of potato polyphenols. A review.

Journal of Agricultural and Food Chemistry. 45, 1523-1540.

Fuleki, T., Francis, F.J. 1968. Quantitative methods for anthocyanins-I. Extraction and

determination of total anthocyanin in cranberries. Journal of Food Science. 33, 72-77.

Giusti, M.M., Wrolstad, R. 2001. Characterization and measurement of anthocyanin by UV-

visible spectroscopy In: Current Protocols in Food Analytical Chemistry., (R.E. Wrolstad,

ed.) pp. pp F1.2.1-F1.2.13, John Wiley & Sons, New York.

Hale, A.L. 2003. Screening potato genotypes for antioxidant activity, identification of the

responsible compounds and differentiating Russet Norkotah strains using AFLP and

microsatellite marker analysis., Texas A&M University, College Station, Texas.

Halliwell, B., Gutteridge, J.M.C., Cross, C.E. 1992. Free-Radicals, Antioxidants, and Human-

Disease - Where Are We Now. Journal of Laboratory and Clinical Medicine. 119, 598-620.

Han, K.H., Hashimoto, N., Shimada, K., Sekikawa, M., Noda, T., Yamauchi, H., Hashimoto, M.,

Chiji, H., Topping, D.L., Fukushima, M. 2006a. Hepatoprotective effects of purple potato

Page 158: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

139

extract against D-galactosamine-induced liver injury in rats. Bioscience Biotechnology and

Biochemistry. 70, 1432-1437.

Han, K.H., Sekikawa, M., Shimada, K., Hashimoto, M., Hashimoto, N., Noda, T., Tanaka, H.,

Fukushima, M. 2006b. Anthocyanin-rich purple potato flake extract has antioxidant capacity

and improves antioxidant potential in rats. British Journal of Nutrition. 96, 1125-1133.

Jansen, G., Flamme, W. 2006. Coloured potatoes (Solanum tuberosum L.) - anthocyanin content

and tuber quality. Genetic Resources and Crop Evolution. 53, 1321-1331.

Kalt, W., Forney, C.F., Martin, A., Prior, R.L. 1999. Antioxidant capacity, vitamin C, phenolics,

and anthocyanins after fresh storage of small fruits. Journal of Agricultural and Food

Chemistry. 47, 4638-4644.

Kalt, W., McDonald, J.E., Donner, H. 2000. Anthocyanins, phenolics, and antioxidant capacity

of processed lowbush blueberry products. Journal of Food Science. 65, 390-393.

Kaur, C., Kapoor, H.C. 2001. Antioxidants in fruits and vegetables - the millennium's health.

International Journal of Food Science and Technology. 36, 703-725.

Lachman, J., Hamouz, K., Orsak, M. 2005. Red and purple potatoes - A significant antioxidant

source in human nutrition. Chemicke Listy. 99, 474-482.

Lewis, C.E., Walker, J.R.L., Lancaster, J.E., Sutton, K.H. 1998. Determination of anthocyanins,

flavonoids and phenolic acids in potatoes. I: Coloured cultivars of Solanum tuberosum L.

Journal of the Science of Food and Agriculture. 77, 45-57.

Nicoli, M.C., Anese, M., Parpinel, M. 1999. Influence of processing on the antioxidant

properties of fruit and vegetables. Trends in Food Science & Technology. 10, 94-100.

Prior, R.L., Cao, G.H., Martin, A., Sofic, E., McEwen, J., O'Brien, C., Lischner, N., Ehlenfeldt,

M., Kalt, W., Krewer, G., Mainland, C.M. 1998. Antioxidant capacity as influenced by total

Page 159: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

140

phenolic and anthocyanin content, maturity, and variety of Vaccinium species. Journal of

Agricultural and Food Chemistry. 46, 2686-2693.

Reddivari, L., Hale, A.L., Miller, J.C. 2007. Determination of phenolic content, composition and

their contribution to antioxidant activity in specialty potato selections. American Journal of

Potato Research. 84, 275-282.

Reyes, L.F., Cisneros-Zevallos, L. 2003. Wounding stress increases the phenolic content and

antioxidant capacity of purple-flesh potatoes (Solanum tuberosum L.). Journal of

Agricultural and Food Chemistry. 51, 5296-5300.

Reyes, L.F., Cisneros-Zevallos, L. 2007. Degradation kinetics and colour of anthocyanins in

aqueous extracts of purple- and red-flesh potatoes (Solanum tuberosum L.). Food Chemistry.

100, 885-894.

Reyes, L.F., Miller, J.C., Cisneros-Zevallos, L. 2005. Antioxidant capacity, anthocyanins and

total phenolics in purple- and red-fleshed potato (Solanum tuberosum L.) genotypes.

American Journal of Potato Research. 82, 271-277.

Rodriguez-Saona, L.E., Giusti, M.M., Wrolstad, R.E. 1998. Anthocyanin pigment composition

of red-fleshed potatoes. Journal of Food Science. 63, 458-465.

Rossi, M., Giussani, E., Morelli, R., Lo Scalzo, R., Nani, R.C., Torreggiani, D. 2003. Effect of

fruit blanching on phenolics and radical scavenging activity of highbush blueberry juice.

Food Research International. 36, 999-1005.

Sadilova, E., Stintzing, F.C., Carle, R. 2006. Thermal degradation of acylated and nonacylated

anthocyanins. Journal of Food Science. 71, C504-C512.

Singh, S., Gamlath, S., Wakeling, L. 2007. Nutritional aspects of food extrusion: a review.

International Journal of Food Science & Technology. 42, 916-929.

Page 160: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

141

Singleton, V.L., Rossi, J.A., Jr. 1965. Colorimetry of Total Phenolics with Phosphomolybdic-

Phosphotungstic Acid Reagents. Am. J. Enol. Vitic. 16, 144-158.

Swain, T., Hillis, W.E. 1959. The phenolic constituents of Prunus domestica-I. The quantitative

abalysis of phenolic constituents. Journal of Food Science and Agriculture. 10, 63-68.

Uyan, S.E., Baysal, T., Yurdagel, O., El, S.N. 2004. Effects of drying process on antioxidant

activity of purple carrots. Nahrung-Food. 48, 57-60.

van den Berg, R., Haenen, G.R.M.M., van den Berg, H., Bast, A. 1999. Applicability of an

improved Trolox equivalent antioxidant capacity (TEAC) assay for evaluation of antioxidant

capacity measurements of mixtures. Food Chemistry. 66, 511-517.

Van den Broeck, I., Ludikhuyze, L., Weemaes, C., Van Loey, A., Hendrickx, M. 1998. Kinetics

for Isobaric-Isothermal Degradation of L-Ascorbic Acid. J. Agric. Food Chem. 46, 2001-

2006.

Wang, H., Cao, G.H., Prior, R.L. 1996. Total antioxidant capacity of fruits. Journal of

Agricultural and Food Chemistry. 44, 701-705.

Wang, S.W. 2002. The influence of drying and thermal treatments on antioxidant activity in

asparagus. In: Department of Food Science and Human Nutrition, Washington State

University, Pullman.

Wolfe, K.L., Liu, R.H. 2003. Apple peels as a value-added food ingredient. Journal of

Agricultural and Food Chemistry. 51, 1676-1683.

Page 161: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

142

CHAPTER FOUR

EFFECT OF EXTRUSION ON THE ANTIOXIDANT CAPACITY AND

COLOR ATTRIBUTES OF EXPANDED EXTRUDATES PREPARED

FROM PURPLE POTATO AND YELLOW PEA FLOUR MIXES

ABSTRACT

Foods with antioxidant capacity contribute health benefits and provide protection against

certain cancers, Alzheimer‘s dementia, and cardio-vascular diseases caused by oxidative

damage. The effect of extrusion cooking on the antioxidant capacity and color attributes of

extruded products prepared from three selected formulations of purple potato and yellow pea

flours using a co-rotating twin screw extruder were studied. Expansion ratios of the extruded

products prepared varied from 3.93 to 4.75. The total antioxidant capacities (TAC) of the

extruded products, using DPPH assay, were 3769 – 4116 µg trolox equivalent/g dry weight

sample and were not significantly different (p > 0.05) from their respective raw formulations.

The total phenolic contents (TP) of the extruded products varied from 2088 to 3766 µg of gallic

acid equivalent/g dry weight sample and retained 73 – 83 % of TP from the raw formulations

after extrusion. The total anthocyanins contents (TA) in the extrudates were 0.116 – 0.228 mg of

malvidin-3-glucosides/g dry weight sample. Compared with their raw formulations, significant

losses (60–70%) of TA in the extruded products occurred during extrusion cooking. Browning

indices and color attributes such as brightness, chroma and hue angles agreed with degradation

of anthocyanins in the extruded products. However, extrusion cooking retained antioxidant

capacities of the raw formulations in the extruded products either in their natural forms or

Page 162: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

143

degraded products with radical scavenging activity. This study demonstrated the potential for the

production of puffed extruded food products with improved antioxidant content from colored

potatoes and pulse formulations.

1. INTRODUCTION

Colored potatoes are rich in anthocyanins, which are known for providing natural color to

fruits and vegetables and for exhibiting antioxidant properties (Cevallos-Casals & Cisneros-

Zevallos, 2003; Nayak, Berrios, Powers, Tang & Ji, 2010). Antioxidants play an important role

to protect against diseases by reacting with and quenching oxidative free radicals, reducing

peroxides, chelating transition metals, and stimulating anti-oxidative defense enzyme activities

(Velioglu, Mazza, Gao & Oomah, 1998). Rice-Evans, Miller, Bolwell, Bramley, & Pridham

(1995) reported that anthocyanins are more effective antioxidants in vitro than ascorbic acid and

vitamin E. Pulses (such as yellow peas) are packed with a high content of nutritional ingredients

such as protein, dietary fibers, complex carbohydrates and folate, and are low in fat and sodium

(Madar & Stark, 2002). Most of these nutritional ingredients are also associated with health

benefits including hypochlesterolemic effects (Pusztai, Grant, Buchan, Bardocz, de Carvalho &

Ewen, 1998), prevention of osteoporosis (Messina, 1999), and certain cancers (Lamartiniere,

2000). Therefore, it would be desirable to make commodities such as colored potatoes and

yellow peas into functional foods in the form of extruded snacks and breakfast cereal-type food

products.

Extrusion cooking is a high temperature, short time process in which food materials are

plasticized and cooked by the combination of moisture, pressure, temperature and mechanical

shear, resulting in molecular transformation and chemical reactions. This technology is

preferable to other processing technologies because of it being a continuous process with high

Page 163: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

144

productivity, significant retention of nutritional quality (Singh, Gamlath & Wakeling, 2007),

natural color and flavor of foods (Bhandari, D'Arcy & Young, 2001).

Expansion ratio is one of the important characteristics of puffed extruded products. Extrusion

cooking of legumes restricts the expansion ratio of the extruded product, but addition of starch-

containing ingredients, such as potato flours, could improve puffing of the extrudates for the

development of expanded-type foods.

Some studies have demonstrated that it is possible to puff pulse flours and pulse-based

formulations into potentially commercial nutritious snack and breakfast cereal-type products

(Berrios, Camara, Torija & Alonso, 2002; Berrios, Morales, Camara & Sanchez-Mata, 2010;

Berrios, Wood, Whitehand & Pan, 2004; Patil, Berrios, Tang & Swanson, 2007). The goals of

this study were to (i) produce puffed extrudates from mixes of colored potatoes and yellow pea

flours, and (ii) evaluate the effect of extrusion conditions on antioxidant capacities, color

attributes, and some physical characteristics of the extrudates.

2. MATERIALS AND METHODS

2.1. Materials

Fresh ‗Purple Majesty‘ cultivar potatoes were purchased from SLV Research Center,

Colorado State University, Colorado, USA. White (‗Russet‘ cultivar) potatoes were purchased

from a local store. The potatoes were stored at 4 °C and 80 % relative humidity for a couple of

weeks, before processing. Split yellow peas were purchased from Giusto‘s specialty food, South

San Francisco, CA. The peas were pin milled into fine flours and stored at room temperature

until use.

Page 164: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

145

2.1.1. Production of potato flours

Purple and white potatoes were peeled using an abrasive peeler and sliced to 6 mm thickness.

Potato slices were blanched in a steam blancher for 8 min to ensure peroxidase inactivation,

since 98 % reduction in peroxidase activity was previously achieved in ~5 mm thickness red-

and purple-fleshed potatoes slices blanched for 3 min. (Reyes and Cisneros-Zevallos, 2007).

Blanched potato slices were immediately cooled in ice water for 8 min, to reduce thermal shock,

drained, and then pureed using a mixer (The Hobart Mfg Company, Troy, OH, USA). Water

(half of the weight of the blanched potatoes) was added to the puree to make it of uniform

consistency, before applying it to a drum dryer. A 15.24 cm × 20.32 cm pilot-scale, counter-

rotating twin-drum dryer (Blaw Knox Food & Chemical Equipment Co., Buffalo, NY, USA) was

used to dehydrate the puree. The surface temperature of the drums was maintained at 135 – 138

°C. The gap between the drums, rotating at 1.13 rpm, was maintained at 0.3 mm. The dehydrated

flakes prepared by drum-drying were pin milled into flour and stored at -30 °C until further use.

The mean particle sizes of pin milled purple potato flour (PPF) and white potato flour (WPF)

were 225 and 220 μm, respectively, as analyzed on a laser scattering particle size distribution

analyzer (Horiba LA-900, Horiba Instruments Incorporated, Irvine, CA).

2.1.2. Sample preparation

A mix of 35/65 (w/w) WPF and split yellow pea flours (SYPF) were prepared in a mixing

bowl under continuous mixing for 10 min. Three formulations with ratios of 35/65, 50/50, 65/35

(w/w), purple potato flour (PPF) and SYPF, respectively, were prepared in the same way. These

formulations will henceforth be referred to as 35, 50 and 65% PPF. All the formulations were

kept in polyethylene bags and stored at room temperature overnight before extrusion processing.

Page 165: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

146

2.1.3. Extrusion conditions

A co-rotating twin-screw extruder (Micro 18, American Leistritz Extruder Corp., NJ, USA)

was used to process the different extrudates under study. The extruder was equipped with five

independently-controlled heating zones that were electrically heated and water cooled by means

of a water cooling system. The temperature of the feeding zone was maintained constant at 80

°C. The detailed temperature profile in the heating zones, feed moisture and screw speeds is

given in Table 4.1. Barrel wall and die temperatures were monitored by respective

thermocouples attached to the top and bottom of the five heating/cooling zones and the die. A

Table 4. 1. Extrusion parameters for preparing extruded products using split yellow pea flours

and white potato flours.

Parameters

Feed rate (g/min) : 45

Feed moisture (% wb) : 17, 21, 25

Screw speed (rpm) : 200, 250, 300

Temperature profiles (°C)

(i) die temperature at 120 °C

(ii) die temperature at 130 °C

(iii) die temperature at 140 °C

: 80, 90, 100, 110, 120

: 80, 100, 110, 120, 130

: 80, 100, 115, 130, 140

pressure transducer was also attached at the die to monitor the operating pressure. Raw

formulations were fed at a constant rate of 45 g/min using a volumetric twin-screw feeder (K-

Tron Process Group, Pitman, NJ, USA). A Bran Luebee metering pump (Pumps & Process

Equipment, Inc., IL, USA) connected to the feeding zone was used to add water to the feed

during processing. The feed moisture content was maintained by changing the water flow rate

while keeping the feed rate constant. Extrusion parameters were displayed on an in-built monitor

to the extruder and the data were auto-saved on a personal computer. Extruded samples were

collected once the operation reached a steady state condition, i.e., the drift in torque was minimal

Page 166: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

147

for at least 5 min. The samples were collected for 3 min, cooled at room temperature under

natural convection conditions, double packed in polyethylene bags, flushed with nitrogen, and

stored at -30 °C for further analysis.

2.1.4. Experimental design

Extrusion cooking of the WPF and SYPF formulations was designed following the

procedures of Box and Behnken (1960). Three independent extrusion parameters, namely feed

moisture content (% wb), screw speed (rpm), and die temperature (°C) were considered for a

three level (+1, 0, -1) design. The effects of these parameters on the expansion ratio of the

extrudates were investigated using response surface methodology (RSM). A total 15 experiments

were designed according to N = k2 + k + cp, where N is the total number of experiments, k is the

factor number, and cp is the number of replicates at the central point. Three factors with three

levels and three replicates at the central point were considered for the experimental design. The

optimum extrusion conditions to obtain acceptable expansion ratios were considered for

producing extruded products from PPF and SYPF.

2.1.5. Determination of Expansion ratio

Five randomly chosen extruded rods per extrusion run were considered for measurement.

Five readings at the nodes and space between the nodes of the rods were taken with a caliper for

calculating the mean diameter of the extrudates. The expansion ratio was calculated as the ratio

of mean cross-sectional diameter of an extrudate to diameter of the die (Camire, Dougherty &

Briggs, 2007).

Page 167: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

148

2.1.6. Moisture content determination

Moisture contents of the raw PPF, SYPF and extruded samples were determined using the

standard procedures of AACC moisture-air-oven method number 44-45A (AACC, 2000).

2.1.7. Pasting profile of potato flours

Pasting profile of WPF and PPF was evaluated using a Rapid Visco Analyzer (RVA)

following the procedures of Batey, Cutin & Moore (1997). Three grams (dry basis) of pin milled

WPF or PPF was put into an RVA canister followed by adding deionized water to a final net

weight of 28.0 g and analyzed with continuous stirring at 160 rpm. The mixture was initially held

at 60 °C for 2 min, followed by 4 minutes of heating to 95 °C at 5.83 °C/min and held there for 4

min. The mixture was then cooled to 50 °C in four minutes at 11.25 °C/min and held for 4 min

for a total run time of 20 minutes The parameters such as time-to-peak viscosity, peak viscosity

(the maximum hot paste viscosity), trough viscosity (the trough at the minimum hot paste

viscosity), and final viscosity (the viscosity at the end of the test) were measured. Breakdown

and total set back associated with the degree of collapse of swollen starch granules

corresponding to release of solubilized starch capable of re-association during cooling were

calculated as following:

Breakdown = Peak viscosity – Trough viscosity

Total setback = Final viscosity – Trough viscosity

2.1.8. Color evaluation

Extruded samples were ground into flour using a food processor to pass through a US # 35

sieve (0.5 mm), whereas raw flour formulations required no further preparation for evaluation of

color attributes. Color attributes were determined using a computer vision system (CVS)

Page 168: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

149

following the procedures of Pandit, Tang, Liu & Mikhaylenko (2007). Briefly, the CVS included

a digital camera (Nikon D70 model) with 18 to 70 mm zoom lens providing 6.1 megapixel

resolutions, a lighting system, and a personal computer. Flour samples were kept in a cylindrical

container and placed on a white plate inside a shooting tent. Images of the samples were taken

with the digital camera mounted downwards on the top of the tent at 50 cm above the sample

plate. The images were downloaded to a PC and analyzed using Adobe Photoshop CS2 software

(version 8.0, Adobe Systems Inc, San Jose, CA, USA) to determine the color parameters (CIE L,

a and b). Since color values in Photoshop software are encoded from 0 to 255, standard scaling

values were determined using the method of Briones & Aguilera (2005):

5.2*

LL (1)

120255

240*

aa (2)

120255

240*

bb (3)

where L, a and b values are from Photoshop and L*,a* and b* values are standardized values

depicting brightness, greenness/redness and blue/yellowness, respectively. The hue angle h° and

Chromaticity C* were computed from a* and b* using the following equations:

)*

*(tan 1

abh (4)

22 *)(*)(* baC (5)

Hue angle (h°), the angular representation of color, is often described as "red," "blue," etc.,

whereas chromaticity describes the purity (saturation) of color. The reference values for h° at

0/360°, 90°, 180°, 270° are magenta red, yellow, bluish-green, and blue, respectively. Color

Page 169: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

150

differences between extruded samples and their respective raw formulations were expressed as

∆E* where,

222 *)(*)(*)(* baLE (6)

2.2. Chemical Analyses

Folin-Ciocalteu reagents, potassium chloride, sodium acetate, trolox (6-Hydroxy-2,5,7,8-

tetramethylchromane-2-carboxylic acid), DPPH (2, 2-diphenyl-1-picrylhydrazyl) and gallic acid

were purchased from Sigma-Aldrich (St. Louis, MO, USA). Laboratory grade methanol and

ethanol were used in extraction and preparation of samples.

2.2.1. Total antioxidant capacity (TAC)

Ten grams of flour from the raw or extruded samples were blended with aqueous methanol

(50:50 v/v) under constant stirring for 2 min at room temperature. The final volume of the

mixture was brought to 100 mL and kept for 90 min at 4 °C, before centrifuged at 30,000 g at 4

°C for 20 min. The supernatants were collected and stored at -20 °C for further analysis.

TAC on the raw and extruded samples was determined using DPPH assay following the

procedures of Brandwilliams, Cuvelier, & Berset (1995). DPPH (a stable, deep purple color

radical) is reduced in the presence of antioxidants decolorizing the solution. Loss of color results

in a decrease in the absorbance intensity, which can be monitored spectrophotometrically at 515

nm, provides the basis for measurement of the antioxidant capacity of the extracts. Stored

supernatants were brought to room temperature and DPPH assay was prepared by adding 0.05

mL of supernatants to 1.95 mL of freshly prepared DPPH solution (6 × 10-5

M) in a cuvette. The

cuvettes were covered with parafilms and thoroughly mixed before taking readings at 515 nm

using a UV/visible spectrophotometer (Ultraspec 4000, Pharmacia Biotech, Cambridge,

Page 170: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

151

England) until a plateau was reached (Brandwilliams et al., 1995). The absorbance at 2 h was

considered optimum for determining the TAC of samples as there was little change in their

absorbance reading toward the end of time. Methanol was used as a blank, whereas DPPH

without sample was taken as control. For each measured sample, the percentage of DPPH

remaining was calculated as:

100)(0@

@

Tabsorbance

tTabsorbance

remainingDPPH

DPPHDPPH (7)

where DPPH absorbance@T=t is the absorbance of DPPH at time t min and DPPH absorbance@T=0 is the

absorbance of DPPH measured at zero min. The TAC was quantified from a trolox standard

curve and expressed as micrograms of trolox equivalent per gram of dry weight sample (µg TE/g

DW). Three replicates were considered for determination of TAC.

2.2.2. Total phenolics (TP)

Extracts from the total antioxidant assay were used for determining TP on the raw and

extruded samples using Folin-Ciocalteu colorimetric method, following the procedures of Swain

and Hillis (1959) and Singleton and Rossi (1965). Briefly, in presence of phenolates, the Folin-

Ciocalteu reagents reduce and produce molybdenum-tungsten blue, which can be measured with

a spectrophotometer. Extracted supernatants were brought to room temperature before diluting

them four times with 50% aqueous methanol. To 0.5 mL supernatants, 8 mL of deionized water

was added followed by 0.5 mL of 0.25 N FCR. The samples were mixed thoroughly and

equilibrated for 3 min at room temperature. After 3 to 4 min, 1 mL of 1 N sodium carbonate was

added to the mixture and mixed. Sodium carbonate raises the pH of the phenols to be oxidized

rapidly in an alkaline medium to form phenolates. The aliquots were kept at room temperature

for 2 h before taking readings 725 nm. 0.5 mL methanol was treated in the same way as the

Page 171: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

152

diluted samples and used as blank. TP was quantified from a gallic acid standard curve and

expressed as micrograms gallic acid equivalent per gram of dry weight sample (µg GAE/g DW).

Three replicates were considered for determination of the TP.

2.2.3. Total anthocyanins (TA)

Determination of TA on the raw and extruded samples was carried out following the

procedures of Fuleki & Francis (1968) with modifications. Five grams of flour from the raw or

extruded samples were blended with aqueous acidified ethanol (50/50 water/acidified ethanol,

v/v) and thoroughly stirred for 2 min. Acidified ethanol was prepared from 85:15 95%

ethanol/1.5 N HCl. The final volume of the mixture was brought to 100 mL with aqueous

acidified ethanol. The mixture was covered with Parafilm and kept for 90 min at 4 °C to

equilibrate, before centrifuging at 23,000 g at 4 °C for 15 min. The supernatants were collected

and stored at -20 °C for further analysis.

TA contents were quantified using the pH differential method, following the procedure of

Giusti & Wrolstad (2001). Anthocyanin assays were prepared by adding 0.2 mL of supernatant

to 1.8 mL of KCl buffer (pH 1.0) or sodium acetate buffer (pH 4.5). The cuvettes containing

aliquots were covered with Parafilm, thoroughly mixed and equilibrated for 15 min at room

temperature, before reading absorbances. Malvidin-3-glucoside was considered the major

anthocyanin in purple potato flour (Han et al., 2006; Jansen & Flamme, 2006; Lewis, Walker,

Lancaster & Sutton, 1998) at the maximum wavelength (λmax) of 535 nm with molecular weight

(MW) of 718.5 g/mol and molar extinction coefficient of 30200 L-1

cm-1

mol-1

. Absorbance

readings at 535 nm (λmax) and 700 nm (for correcting turbidity) (Reyes & Cisneros-Zevallos,

2003) were taken in a UV/visible spectrophotometer, previously blanked with distilled water.

Page 172: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

153

The TA of the raw formulations and the extrudates were calculated according to the following

formula (Giusti et al., 2001):

de

DFMWAlmgC

*

**)/( (8)

where A = Absorbance of the sample given by 5.4700max0.1700max )()( pHpH AAAAA MW =

molecular weight of malvidin-3-glucoside, DF = dilution factor, e = molar extinction coefficient

and d= path length of the cuvette (1 cm). TA contents were expressed as mg of malvidin-3-

glucosides per gram of DW sample (mg mv-3-glu/g DW). Individual anthocyanin were not

identified or calculated.

2.2.4. Browning Index (BI)

Degradation of anthocyanins and formation of brown Maillard Reactions Products (MRP)

was assessed based on the BI of the products. Extracts from the total anthocyanins were used to

determine BI in the raw and the extruded samples, following the procedures of Jackman, Yada &

Tung (1987). Absorbance readings of 2 mL of supernatants were taken at 535 (λmax), 420, and

700 nm for calculating Browning Index of the samples as:

700535

700420

AA

AABI

(9)

where A420, A535 and A700 were absorbances at 420, 535, and 700 nm, respectively.

2.3. Statistical analysis

All physical and chemical data obtained from the raw formulations and extruded samples

were collected and analyzed with SAS (version 9.1, SAS Institute Inc, Cary, NC, USA) using

analysis of variance (ANOVA). Tukey‘s pair-wise comparison at 95 % confidence level was

Page 173: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

154

used to identify statistical significant differences (p < 0.05). All the data were expressed as mean

± standard deviation.

3. RESULTS AND DISCUSSION

3.1. Expansion ratio (ER)

The use of RSM allowed determining how the ER of the extrudates produced from WPF and

SYPF formulations varied under the influence of the selected extrusion parameters of screw

speed (200, 250, and 300 rpm) and die temperature (120, 130, and 140 oC). At the selected

extrusion parameters, the ER of the extrudates produced from the WPF and SYPF formulation

ranged from 1.78 to 5.18 (Figure. 4.1). The amount of PPF was limited compared to that of

WPF. Therefore, experimental designs for the extrusion of the PPF and SYPF formulations were

based on the optimized conditions for ER of the extrudates produced using WPF and SYPF

formulations. Expansion ratio of about 5.0 was selected from the optimized extrusion conditions

with 17% (wb) feed moisture, 250 rpm screw speed and 140 °C temperature. However, actual

extrusion processing of PPF and SYPF formulations were conducted at a screw speed of 300 rpm

and temperature at 130 °C in addition to 140 °C, to reduce the residence time and exposure to

heat of the mix in the extruder barrel, to achieve maximum color retention in the final extrudates.

The expansion ratios of the extrudates prepared from the PPF and SYPF formulations varied

from 3.93 to 4.75 (Table 4.2). Camire et al., (2007) reported diametric expansions of 1.90 to

1.93 in extruded products, processed at extrusion conditions of 175 rpm screw speed, 163 °C die

temperature, and a feed rate of 255 g/min. The formulation was prepared from 84.3% cornmeal

and other food ingredients (sucrose and dehydrated blueberry, cranberry, raspberry, and concord

grape powders). The ER of the extrudates produced from formulations containing 50 and 65%

Page 174: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

155

PPF were not significantly different (p > 0.05) at 130 and 140 °C. This tended to indicate that a

difference in processing temperatures of 10 °C might not be sufficient to promote a significant

increase (p < 0.05) in ER of the extrudates. However, the ER of extrudates produced from

formulations containing 35% PPF were significantly smaller (p < 0.05) than those produced from

formulations containing 50 and 65% PPF, under the two die temperatures under study of 130 and

140 oC. It is known that starch has a positive effect on increasing expansion, while fiber and/or

protein have a negative and lowering effect on expansion of extrudates (Conway, 1971a;

Conway, 1971b; Guy & Horn, 1988; Kim & Maga, 1987). These findings support the results on

ER reported in this study, as high potato flour (source of starch) and low dry pea flour content

(source of protein and fiber) in the extrudates produced from formulations containing 50 and

3002

250

3

4

5

18 20 20022 24

Expansion

Ratio

% Feed Moisture

Screw Speed

Figure. 4. 1. Effects of screw speed and feed moisture on the different expansion ratio of

extrudates prepared from white potato and yellow pea flours at 140 °C die temperature

Page 175: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

156

65% PPF, presented extrudates with the highest values of ER. Conversely, low concentration of

potato flour and high concentration of dry pea flour, in the extrudates produced from

formulations containing 35% PPF, resulted in extrudates with significantly (p < 0.05) lower

values of ER. Additionally, extrudates produced from formulations containing 35% PPF showed

large variability of diameter (used for calculation of ER) among the different extrudates.

Therefore, the difference (p < 0.05) of ER observed between extrudates produced at 130 and 140

oC is attributed to this indicated variability.

Table 4. 2. Moisture content and expansion ratios of the extrudates prepared from yellow pea

and purple potato flours (n=3); Feed moisture content: 17% (wet basis).

Formulations

(Potato/Pea)

Treatment Moisture content

(% wet basis)

Expansion Ratio

35/65 w/w

Raw Formulation 9.32 0

Extruded @130 °C 7.03 4.28 ± 0.11b

Extruded @140 °C 6.97 3.93 ± 0.22c

50/50 w/w

Raw Formulation 9.01 0

Extruded @130 °C 7.26 4.74 ± 0.10a

Extruded @140 °C 7.39 4.48 ± 0.12a

65/35 w/w

Raw Formulation 8.69 0

Extruded @130 °C 7.43 4.75 ± 0.27a

Extruded @140 °C 7.19 4.53 ± 0.26a

Significant differences within the values in the same column are indicated by different

superscript letters (p < 0.05, Tukey‘s pairwise comparison test)

3.2. Pasting behavior of potato flours

Differences and/or similarities in the pasting behavior of white and purple potato flours can

be explained based on their RVA‘s pasting curves. The pasting behaviors of both the WPF and

PPF followed similar patterns with regard to times to attain their peak, trough and final

viscosities (Figure. 4.2). Breakdown and total setback were 433 and 601 cP, for WPF, and 314

and 629 cP for PPF (Table 4.3). The similar proximate composition presented by the white and

Page 176: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

157

Table 4. 3. Pasting behaviors of white and purple potato flours.

Flours Peak

viscosity (cP)

Trough

viscosity (cP)

Final

viscosity (cP)

Breakdown*

(cP)

Total

Setback# (cP)

White potato 1380 947 1548 433 601

Purple potato 1285 971 1600 314 629

*Breakdown = Peak viscosity – Trough viscosity;

#Total setback = Final viscosity – Trough viscosity.

purple potato flours may be responsible for the similar pasting profile patterns displayed by the

two flours. Similar observations were previously reported for pasting profiles of white and

colored potato flours (Hoover, 2001). Use of PPF was justified and used with SYPF for further

analyses with the optimized extrusion process parameters.

3.3. Color attributes

The brightness (L*) or color lightness among the different raw formulations prepared from

PPF and SYPF varied significantly (p < 0.05) from 77 to 88 on a scale of 0–100. The raw

formulation containing 35% PPF was significantly (p < 0.05) brighter than formulation

containing 50% PPF and this one was significantly (p < 0.05) brighter than formulation

containing the brightness color parameter. That is, a decrease in chroma and and hue values with

an increase of PPF in the formulations (Table 4.4). The raw SYPF flours had a whitish color

while potato flour has a purplish color, due to their anthocyanins content. Therefore, as the

amount of PPF increased in the formulations, the color parameters of brightness, chroma and hue

significantly (p < 0.05) decreased.

When comparing the brightness values (L*) of the different raw formulations with those of

their extrudates it was observed that extrusion processing, at die temperatures at 130 and 140 °C,

caused a significant (p < 0.05) decrease in brightness, chroma and hue, at all levels of PPF

addition. It is known that reducing sugars and proteins (amino acids) in foods can react under

Page 177: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

158

hue values with and 65% PPF. The values of chroma and hue followed the same trend as

high processing temperatures to promote nonenzymatic browning (Maillard reaction), which

result in darkening of the final product. Potatoes are high in sugars and dry peas are high in

protein (amino acids). Therefore, the observed decrease in brightness is attributed to the Maillard

reaction, as a consequence of extrusion processing. Similarly, previous researchers have

indicated that extrusion of the whey protein concentrate and corn starch gave higher color

differences with increasing amylose content (Matthey & Hanna, 1997). Additionally, the

degradation of purple anthocyanins due to extrusion temperatures could have generated Maillard

reaction products that promoted the changes in color parameters of brightness, chroma and hue

values, observed between the unprocessed (raw) and extruded products.

Color difference (∆E*) was used to represent the color change between the unprocessed and

Time (Sec)

0 200 400 600 800 1000 1200 1400

Vis

cosi

ty (

cP)

0

200

400

600

800

1000

1200

1400

1600

1800

Tem

per

atu

re (

°C)

40

50

60

70

80

90

100

Purple flour

White flour

Temp (°C)

Figure. 4. 2. RVA profile of white and purple potato flours.

Page 178: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

159

processed flours (effect of processing). In this study, the values of ∆E* for all flours increased

significantly (p < 0.05) as the processing temperature increased from 130 to 140 oC (Table 4.4).

The values of ∆E* for the extruded flours containing 35 and 50% PPF were similar, at both

indicated processing temperatures. However, those from extruded flours containing 65% PPF

were significantly (p < 0.05) greater. These results indicate that the color in the PPF had, along

with processing temperatures, a direct effect on the values of ∆E*. Additionally, they indicated

that the results obtained for ∆E* support those obtained for the color parameters of brightness,

chroma and hue values. Berrios et al., (2004) reported that there are no established threshold or

cut-off values for color development of an acceptable legume-based snack, due to the lack of this

type of product in the market place. Therefore, the color data in this study may have important

value for future product development of legume pulse-based snack type products.

Table 4. 4. Color attributes of the extruded products prepared from yellow pea and purple potato

flours (n = 3).

Formulations

(Potato/Pea) Treatment

Brightness

(L*)

Chroma

(C*)

Hue

(h°)

Color difference

(∆E*)

35/65 w/w

Raw Formulation 88.10 ± 0.4a 8.60 ± 0.3

f 89.31 ± 0.4

a 0.00

Extruded @130 °C 73.00 ± 1.6d 26.48 ± 0.8

d 87.85 ± 1.2

a 23.41 ± 0.8

c

Extruded @140 °C 73.82 ± 1.5d 29.87 ± 0.4

b 87.61 ± 1.5

a 25.62 ± 0.7

b

50/50 w/w

Raw Formulation 82.10 ± 1.3b 5.20 ± 0.3

g 67.83 ± 0.6

c 0.00

Extruded @130 °C 70.57 ± 0.7e 22.73 ± 0.8

e 87.71 ± 1.4

a 21.31 ± 1.3

c

Extruded @140 °C 72.72 ± 1.4d 28.33 ± 0.4

c 87.47 ± 1.3

a 25.29 ± 0.5

b

65/35 w/w

Raw Formulation 77.03 ± 1.0c 4.04 ± 0.2

h 21.35 ± 2.4

d 0.00

Extruded @130 °C 69.46 ± 1.7e 27.28 ± 0.7

d 83.64 ± 1.6

b 26.75 ± 0.6

b

Extruded @140 °C 73.80 ± 0.4d 33.19 ± 0.8

a 84.61 ± 0.3

b 31.74 ± 0.8

a

a Significant differences within the values in the same column are indicated by different

superscript letters (p < 0.05, Tukey‘s pairwise comparison test).

3.4. Total antioxidant capacity

The TAC of the raw formulations and extrudates were determined using DPPH assay and

expressed as µg TE/g DW (Figure. 4.3). The TAC of the raw formulations with 35, 50 and 65%

Page 179: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

160

PPF were 3936 ± 71, 4025 ± 35 and 4083 ± 37 µg TE/g DW, respectively. These values were not

significantly different (p > 0.05). However, data clearly showed that TAC increased with

increasing PPF in the formulations. Those increases represented 2.21 and 1.42%, from 35 to 50

and from 50 to 65% PPF in the formulations, respectively. On the other hand, 0 – 35% PPF

addition resulted in a 60.5% increase in TAC (data not shown in Figure. 4.3). This significant (p

< 0.05) increase in TAC can be attributed to the purple color in the PPF. Since, studies on the

TAC of colorful fruits and vegetables, at the Jean Mayer USDA Human Nutrition Research

Center on Aging at Tuffs University, revealed that a large group of color compounds are

flavonoids (including anthocyanins) with potent antioxidant protection against peroxyl radicals

(McBride, 1996; Wang, Cao & Prior, 1997). When comparing the TAC values of the different

raw formulations with those of their extrudates it was observed that, even though the TAC values

of the extrudates were lower, they were not significantly (p > 0.05) different. Similar TAC

content in the raw formulations and extruded products could be attributable to the effect of

extrusion on (i) breaking complex polyphenols into low molecular weight phenolic compounds

with scavenging activity, (ii) interaction of the phenolics with protein under heat treatment, and

(iii) formation of Maillard Reaction Products. On the other hand, with the exception of

extrudates formulated with 50% PPF, the TAC values of those formulations extruded at die

temperatures of 140 °C showed significantly (p < 0.05) lower TAC values than their raw

counterparts. High temperature extrusion promotes the Maillard reaction and the formation of

brown compounds that may have had an effect on the TAC values of the extrudates (Anese,

Manzocco, Nicoli & Lerici, 1999). Additionally, the potential binding of phenolic compounds to

the protein matrix may account for the decrease in TAC values observed in those formulations

extruded at die temperatures at 140 °C compared to the raw samples. At high protein

Page 180: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

161

concentration complex interactions and cross-linking of different protein molecules with

phenolic compounds forms a hydrophobic surface (Mcmanus, Davis, Beart, Gaffney, Lilley &

Haslam, 1985). No significant effect (p > 0.05) as a result of extrusion temperatures of 130 and

140 °C was observed on the TAC of extruded products formulated with 35, 50, and 65% PPF,

respectively. This result may indicate that an in order to see an effect on TAC, an increase in die

temperature greater than 10 oC needs to be used. Camire et al., (2007) reported no significant

change (p > 0.05) in antioxidant activity in control and cranberry extruded products prepared

with corn at a substitution level of 1% and extrusion temperature of 165 °C.

3500

3600

3700

3800

3900

4000

4100

4200

4300

35% PPF 50% PPF 65% PPF

Formulations

To

tal

an

tio

xid

an

t ca

pa

city

(μg

TE

/g D

W)

Non-extruded

Extruded @130 °C

Extruded @140 °C

ab

bc

c

aa

a

a

ab

b

Figure. 4. 3. Total antioxidant capacities by DPPH assay of raw formulations and extruded

products. The formulations contained x% of purple potato flour and (100-x)% of dry pea

flour.aValues in each bar with no letters in common are significantly different (p < 0.05).

Page 181: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

162

3.5. Total phenolics

The TP of the raw formulations and extruded products were determined using the FC reagent

method and expressed as µg of GAE/g DW (Figure. 4.4). Comparing the TP content of different

raw formulations it was determined that the TP content of the 65% PPF formulation (4548 ± 117

µg of GAE/g DW) was significantly higher (p < 0.05) than that of 50% PPF (3838 ± 286 µg of

GAE/g DW) formulation; and this one was significantly higher (p < 0.05) than the TP of the

35% PPF (2818 ± 46 µg of GAE/g DW) formulation. Previous researchers have demonstrated

that purple-fleshed potato cultivars had higher phenolic contents than white-fleshed cultivars

(Nayak et al., 2010; Stushnoff et al., 2008) and yellow peas (Xu & Chang, 2008). These reports

support the results obtained in the present study that purple-fleshed potatoes had higher content

of TP than yellow peas. Therefore, those raw formulations containing the highest proportion of

PPF had also the highest content of TP. A similar pattern on TP content was observed in the

extruded products. However, significant losses (p < 0.05) in TP were determined in products

processed at die temperature of 130 °C prepared from formulations containing 50 and 65% PPF,

when compared to TP of their raw formulations. One exception was observed for extruded

products prepared from formulations containing 35% PPF, whose TP contents were not

significantly different (p > 0.05) from the raw formulations. This could be attributed to high

standard deviation values determined for those extruded samples. Similarly, products processed

at a die temperature of 140 °C prepared from formulations containing 35, 50 and 65% PPF had

significantly (p < 0.05) less TP content than their respective raw formulations. Viscidi,

Dougherty, Briggs & Camire (2004) reported significant loss of total phenolics during extrusion

of oat cereals. Additionally, Zadernowski, Nowak-Polakowska & Rashed (1999) reported losses

of up to ~60% of phenolic compounds in extruded oat samples, compared to their respective raw

Page 182: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

163

0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

35% PPF 50% PPF 65% PPF

Formulations

To

tal

ph

eno

lics

g G

AE

/g D

W)

Non-extruded

Extruded @130 °C

Extruded @140 °C

c

cd

d

b

c

c

a

b b

Figure. 4. 4. Total phenolic contents of raw formulations and extruded products. The

formulations contained x% of purple potato flour and (100-x)% of dry pea flour. aValues in each

bar with no letters in common are significantly different (p < 0.05).

samples. These reports corroborate well with the findings of the present study. Contrary to the

previous reports, Camire et al.(2007) reported a higher content of soluble phenolics, as ferulic

acid equivalents, in Concord grape and raspberry extrudates compared to their control samples.

Phenolic compounds are heat-liable and can break upon exposure to high temperatures.

Therefore, losses in the TP content of formulations under extrusion are expected to occur, due to

break down of complex polyphenols into other phenolic or non-phenolic compounds, as a

consequence of high temperatures conditions. However, extrusion die temperatures of 130 and

140 °C had no significant effect (p > 0.05) on the TP content of the extrudates (Figure. 4.4).

This indicated that, under the extrusion possessing conditions of the study, a die temperature

differential of 10 °C had not detrimental effect on TP.

Page 183: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

164

3.6. Total anthocyanins

The TA content in the raw formulations and extrudates was determined using the pH

differential method and expressed as mg of mal-3-glu/g DW (Figure. 4.5). Results obtained from

the different raw formulations demonstrated that the TA content in the 65% PPF formulation

(0.729 ± 0.04 mg of mal-3-glu/g DW) was significantly higher (p < 0.05) than TA in formulation

containing 50% PPF (0.584 ± 0.05 mg of mal-3-glu/g DW) followed by formulation with 35%

PPF (0.363 ± 0.01 mg of mal-3-glu/g DW). In general, the TA content in the extruded products

prepared from formulations containing 65, 50 and 35% PPF and processed at die temperatures of

130 and 140°C followed same trend observed for the raw formulations. This is a progressive and

significant decrease in TA content as the percentage of PPF in the formulations decreased from

65 to 35%. Higher concentrations of PPF in the formulations contributed to higher content of

TA. Since the purple color in potato flour is mainly due to the presence of the anthocyanins

petunidin and malvidin glucosides present in the potato flesh and skin (Stushnoff et al., 2008);

whereas, preliminary TA results in yellow pea flour showed negligible values (data not shown).

Compared to their raw counter parts, extruded samples showed a significant loss in TA at all the

different levels of PPF in the formulations. The losses in TA were more evident in extrudates

processed at the highest die temperature of 140 °C. Different from the results obtained

previously, where a die temperature differential of 10 °C had no detrimental effect on TP, the

extrudates containing 35 and 50% PPF processed at a die temperature of 140 °C reflected a

significant loss in TA compared to those processed at 130 °C. Extrudates containing the highest

percentage of PPF (65%) processed at die temperatures of 130 and 140 °C presented similar TA

losses. Stability of anthocyanins is affected by a number of factors such as temperature, pH,

light, oxygen, enzymes, ascorbic acid, sulfur dioxide, sugars, metal ions, etc., (Francis, 1989).

Page 184: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

165

Degradation of anthocyanins in the extruded products could be attributed to the breaking of

anthocyanin structures at high temperatures. High temperature at the initial step could open

either the pyrylium ring of the anthocyanins and form chalcone (Sadilova, Carle & Stintzing,

2007), or hydrolyze the glycosidic moiety and form aglycon (Sadilova, Stintzing & Carle, 2006),

0

0.2

0.4

0.6

0.8

1

35% PPF 50% PPF 65% PPF

Formulations

To

tal

an

tho

cya

nin

s

(mg

MV

-3-G

ly/g

DW

)

0.00

0.10

0.20

0.30

0.40

0.50

0.60

Bro

wn

ing

in

dex

Non-extruded Extruded @130 °C Extruded @140 °C

Non-extruded Extruded @130 °C Extruded @140 °C

c

fg

b

d

e

a

dd

Total anthocyanins

Browning index

Figure. 4. 5. Total anthocyanins contents and browning indices of raw formulations and

extruded products. The formulations contained x% of purple potato flour and (100-x)% of dry

pea flour. aValues in each bar with no letters in common are significantly different (p < 0.05).

providing degradation products as quercetin, phloroglucinaldehyde and protocatechuic acid.

Degradation of anthocyanins in the extruded products might be also due to the formation of

browning compounds caused by the Maillard reaction at high temperatures (Nicoli, Anese &

Parpinel, 1999). A study on the extrusion cooking of blueberry and grape anthocyanins, used as

breakfast cereal colorants, by Camire, Chaovanalikit, Dougherty & Briggs (2002) reported losses

of 90% of blueberry anthocyanins and 74% of grape anthocyanins, induced by extrusion

Page 185: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

166

processing. Similarly, Camire et al., (2007) reported almost 90% loss in anthocyanins content in

extruded corn products containing fruit powders. These reports support the results of the present

study and indicate that when processing food materials that are a good source of anthocyanins,

special attention should be giving to the processing conditions to avoid significant losses.

3.7. Browning Index (BI)

Results of BI determined in raw formulations (Figure. 4.5) showed that the BI of

formulations with 35 and 50% PPF (0.19 ± 0.0, 0.20 ± 0.0, respectively) were not different from

(p > 0.05) each other. However, the BI of formulation with 65% PPF (0.15 ± 0.0) was lower (p <

0.05) than the former formulations. Since BI values are calculated based on the ratio of

absorbances at 420, 535 nm subtracting haziness at 700 nm, a higher absorbance value of the

65% PPF formulation at 535 nm (because of higher concentration of PPF than other

formulations) contributed to the observed lower BI values. The BI of the extrudates processed at

130 and 140 °C were significantly higher (p < 0.05) than their respective raw formulations.

Additionally, the BI of extrudates prepared at 140 °C (0.34 ± 0.04, 0.34 ± 0.00 and 0.49 ± 0.04

for 35, 50 and 65% PPF, respectively) were significantly higher (p < 0.05) than those processed

at 130 °C (0.29 ± 0.01, 0.29 ± 0.01 and 0.33 ± 0.03 for 35, 50 and 65% PPF, respectively).

Higher BI in the extrudates might be due to the formation of browning compounds by the

Maillard reaction of amino acids present in peas and reducing sugar in potato flours at high

temperature. Degradation of anthocyanins in the extrudates with change in color attributes also

agrees with higher BI in the extrudates. Karel and Labuza (1968) reported hydrolysis of sucrose

giving reducing sugars, which has potential for browning in a model system containing sucrose.

Browning of extruded products could have related to the feed moisture content in the

formulation, concentration of ingredients and other extrusion parameters. Dominance influence

Page 186: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

167

of water on the rate of browning in systems containing carbonyl compounds have been reported

in the literature (Erlandson & Wrolstad, 1972).

3.8. Correlation analyses

Correlations (r2) among the TP, TAC, TA, BI and color attributes were analyzed using

Pearson‘s correlation coefficients method. The TAC was not strongly correlated with the TP

(r2 = 0.53, p > 0.05) or TA (r

2 = 0.46, p > 0.05). The correlation coefficients showed that the

phenolic compounds including anthocyanins were not solely responsible for antioxidant capacity

in the formulations and extruded products. Presence of other secondary metabolites such as

volatile oils, carotenoids and vitamins also might have contributed to the total antioxidant

capacity in the raw formulations and extrudates. Strong correlation of the TA with the TP (r2 =

0.76, p < 0.05) agrees the contribution of anthocyanins to the phenolic compounds. Contents of

the TA in the formulations and extruded products were negatively correlated with the BI (r2 = -

0.71, p < 0.05), chroma (r2 = -0.89, p < 0.05), and hue (r

2 = -0.86, p < 0.05) but positively

correlated with brightness (r2 = 0.52, p < 0.05). Anthocyanins are responsible for the purple color

of PPF. The negative correlations of TA with the BI, chroma and hue were attributable to the

concentrations of PPF in the formulations and degradation of anthocyanins in the extruded

products.

4. CONCLUSIONS

Naturally colored extruded puffed food products rich in antioxidants can be produced from

yellow pea and purple potato flours using extrusion cooking technology. Although degradation

in total anthocyanins content was determined, some purple color was retained in the final

extruded products. This indicated that high temperature-short time extrusion processing is a

Page 187: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

168

suitable process for the fabrication of products from antioxidant-rich colored ingredients. The

total antioxidant capacities in the extruded products were retained due to the preservation of

phenolics during processing. Addition of the purple potato flour to yellow pea flour provided an

acceptable expansion ratio to the extruded products. Presence of natural color in the final

extrudates could play a major role in consumer attraction and acceptability, as well as

marketability of the developed extruded food products. More research on the use of extrusion

parameters and their effect on the kinetics of anthocyanins are necessary to study the stability of

natural color in the final extrudates. Naturally colored food ingredients, such as purple potato

flour, has potential for substituting in place of artificial colors, which are generally added as

coatings during the downstream processes.

ACKNOWLEDGEMENTS

We gratefully acknowledge the assistance of James Pan and Matthew Tom, Processed Foods

Research Unit, WRRC, USDA-ARS, Albany, CA for their assistance and feedback during the

extrusion experiment. We also acknowledge the US Dry Pea and Lentil Council for their

financial support for this project.

Page 188: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

169

REFERENCES

AACC (2000). Approved methods of the American Association of Cereal Chemists (10th ed.). St.

Paul, MN, USA: American Association of Cereal Chemists.

Anese, M., Manzocco, L., Nicoli, M. C., & Lerici, C. R. (1999). Antioxidant properties of

tomato juice as affected by heating. Journal of the Science of Food and Agriculture, 79(5),

750-754.

Batey, I. L., Curtin, B. M., & Moore, S. A. (1997). Optimization of rapid-visco analyser test

conditions for predicting Asian noodle quality. Cereal Chemistry, 74(4), 497-501.

Berrios, J. D., Camara, M., Torija, M. E., & Alonso, M. (2002). Effect of extrusion cooking and

sodium bicarbonate addition on the carbohydrate composition of black bean flours. Journal

of Food Processing and Preservation, 26(2), 113-128.

Berrios, J. D., Morales, P., Camara, M., & Sanchez-Mata, M. C. (2010). Carbohydrate

composition of raw and extruded pulse flours. Food Research International, 43(2), 531-536.

Berrios, J. D., Wood, D. F., Whitehand, L., & Pan, J. (2004). Sodium bicarbonate and the

microstructure, expansion and color of extruded black beans. Journal of Food Processing

and Preservation, 28(5), 321-335.

Bhandari, B., D'Arcy, B., & Young, G. (2001). Flavour retention during high temperature short

time extrusion cooking process: a review. International Journal of Food Science and

Technology, 36(5), 453-461.

Box, G. E. P., Behnken, D.W., (1960). Some new three level designs for the study of quantitative

variables. Technometrics, 2, 455-475.

Page 189: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

170

Brandwilliams, W., Cuvelier, M. E., & Berset, C. (1995). Use of a free radical method to

evaluate antioxidant activity. Food Science and Technology-Lebensmittel-Wissenschaft &

Technologie, 28(1), 25-30.

Briones, V., & Aguilera, J. M. (2005). Image analysis of changes in surface color of chocolate.

Food Research International, 38(1), 87-94.

Camire, M. E., Chaovanalikit, A., Dougherty, M. P., & Briggs, J. (2002). Blueberry and grape

anthocyanins as breakfast cereal colorants. Journal of Food Science, 67(1), 438-441.

Camire, M. E., Dougherty, M. P., & Briggs, J. L. (2007). Functionality of fruit powders in

extruded corn breakfast cereals. Food Chemistry, 101(2), 765-770.

Cevallos-Casals, B. A., & Cisneros-Zevallos, L. (2003). Stoichiometric and kinetic studies of

phenolic antioxidants from Andean purple corn and red-fleshed sweetpotato. Journal of

Agricultural and Food Chemistry, 51(11), 3313-3319.

Conway, H. F. (1971a). Extrusion cooking of cereals and soybeans-I. Food Product

Development, 5(2), 27 - 29.

Conway, H. F. (1971b). Extrusion cooking of cereals and soybeans-II. Food Product

Development, 5(3), 14 - 17.

Erlandson, J. A., & Wrolstad, R. E. (1972). Degradation of anthocyanins at limited water

concentration. Journal of Food Science, 37(4), 592-595.

Francis, F. J. (1989). Food colorants - Anthocyanins. Critical Reviews in Food Science and

Nutrition, 28(4), 273-314.

Fuleki, T., & Francis, F. J. (1968). Quantitative methods for anthocyanins-I. Extraction and

determination of total anthocyanin in cranberries. Journal of Food Science, 33, 72-77.

Page 190: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

171

Giusti, M. M., & Wrolstad, R. (2001). Characterization and measurement of anthocyanin by UV-

visible spectroscopy In: R. E. Wrolstad, Current Protocols in Food Analytical Chemistry.,

vol. F 1.2 (pp. pp F1.2.1-F1.2.13): John Wiley & Sons, New York.

Guy, R. C. E., & Horn, A. W. (1988). Extrusion and co-extrusion of cereals. In: J. M. V.

Blanshard, & J. V. Michelle, Food structure - Its creation and evaluation (pp. 331 - 349).

Butterworths, London.

Han, K. H., Sekikawa, M., Shimada, K., Hashimoto, M., Hashimoto, N., Noda, T., Tanaka, H., &

Fukushima, M. (2006). Anthocyanin-rich purple potato flake extract has antioxidant capacity

and improves antioxidant potential in rats. British Journal of Nutrition, 96(6), 1125-1133.

Hoover, R. (2001). Composition, molecular structure, and physicochemical properties of tuber

and root starches: a review. Carbohydrate Polymers, 45(3), 253-267.

Jackman, R. L., Yada, R. Y., & Tung, M. A. (1987). A review - Separation and chemical

properties of anthocyanins used for their qualitative and quantitative analysis. Journal of

Food Biochemistry, 11(4), 279-308.

Jansen, G., & Flamme, W. (2006). Coloured potatoes (Solanum tuberosum L.) - anthocyanin

content and tuber quality. Genetic Resources and Crop Evolution, 53(7), 1321-1331.

Karel, M., & Labuza, T. P. (1968). Nonenzymic browning in model systems containing sucrose.

Journal of Agricultural and Food Chemistry, 16(5), 717-719.

Kim, C. H., & Maga, J. A. (1987). Properties of extruded whey-protein concentrate and cereal

flour blends. Lebensmittel-Wissenschaft & Technologie, 20(6), 311-318.

Lamartiniere, C. A. (2000). Protection against breast cancer with genistein: a component of soy.

American Journal of Clinical Nutrition, 71(6), 1705s-1707s.

Page 191: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

172

Lewis, C. E., Walker, J. R. L., Lancaster, J. E., & Sutton, K. H. (1998). Determination of

anthocyanins, flavonoids and phenolic acids in potatoes. I: Coloured cultivars of Solanum

tuberosum L. Journal of the Science of Food and Agriculture, 77(1), 45-57.

Madar, Z., & Stark, A. H. (2002). New legume sources as therapeutic agents. British Journal of

Nutrition, 88, S287-S292.

Matthey, F. P., & Hanna, M. A. (1997). Physical and functional properties of twin-screw

extruded whey protein concentrate corn starch blends. Food Science and Technology-

Lebensmittel-Wissenschaft & Technologie, 30(4), 359-366.

McBride, J. (1996). Plant pigments - Paint a rainbow of antioxidants. Agricultural Research, 44,

4 - 8.

Mcmanus, J. P., Davis, K. G., Beart, J. E., Gaffney, S. H., Lilley, T. H., & Haslam, E. (1985).

Polyphenol interactions .1. Introduction - Some observations on the reversible complexation

of polyphenols with proteins and polysaccharides. Journal of the Chemical Society-Perkin

Transactions 2(9), 1429-1438.

Messina, M. J. (1999). Legumes and soybeans: overview of their nutritional profiles and health

effects. American Journal of Clinical Nutrition, 70(3), 439s-450s.

Nayak, B., Berrios, J. J., Powers, J. R., Tang, J., & Ji, Y. (2010). Colored potatoes (Solanum

tuberosum L.) dried for antioxidant-rich value-added foods. Journal of Food Processing and

Preservation (in Press).

Nicoli, M. C., Anese, M., & Parpinel, M. (1999). Influence of processing on the antioxidant

properties of fruit and vegetables. Trends in Food Science & Technology, 10(3), 94-100.

Page 192: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

173

Pandit, R. B., Tang, J., Liu, F., & Mikhaylenko, G. (2007). A computer vision method to locate

cold spots in foods in microwave sterilization processes. Pattern Recognition, 40(12), 3667-

3676.

Patil, R. T., Berrios, J. D. J., Tang, J., & Swanson, B. G. (2007). Evaluation of methods for

expansion properties of legume extrudates. Applied Engineering in Agriculture, 23(6), 777-

783.

Pusztai, A., Grant, G., Buchan, W. C., Bardocz, S., de Carvalho, A. F. F. U., & Ewen, S. W. B.

(1998). Lipid accumulation in obese Zucker rats is reduced by inclusion of raw kidney bean

(Phaseolus vulgaris) in the diet. British Journal of Nutrition, 79(2), 213-221.

Reyes, L. F., & Cisneros-Zevallos, L. (2003). Wounding stress increases the phenolic content

and antioxidant capacity of purple-flesh potatoes (Solanum tuberosum L.). Journal of

Agricultural and Food Chemistry, 51(18), 5296-5300.

Rice-Evans, C. A., Miller, N. J., Bolwell, P. G., Bramley, P. M., & Pridham, J. B. (1995). The

relative antioxidant activities of plant-derived polyphenolic flavonoids. Free Radical

Research, 22(4), 375-383.

Sadilova, E., Carle, R., & Stintzing, F. C. (2007). Thermal degradation of anthocyanins and its

impact on color and in vitro antioxidant capacity. Molecular Nutrition & Food Research,

51(12), 1461-1471.

Sadilova, E., Stintzing, F. C., & Carle, R. (2006). Thermal degradation of acylated and

nonacylated anthocyanins. Journal of Food Science, 71(8), C504-C512.

Singh, S., Gamlath, S., & Wakeling, L. (2007). Nutritional aspects of food extrusion: a review.

International Journal of Food Science & Technology, 42(8), 916-929.

Page 193: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

174

Singleton, V. L., & Rossi, J. A., Jr. (1965). Colorimetry of total phenolics with

phosphomolybdic-phosphotungstic acid reagents. Am. J. Enol. Vitic., 16(3), 144-158.

Stushnoff, C., Holm, D., Thompson, M. D., Jiang, W., Thompson, H. J., Joyce, N. I., & Wilson,

P. (2008). Antioxidant properties of cultivars and selections from the Colorado potato

breeding program. American Journal of Potato Research, 85(4), 267-276.

Swain, T., & Hillis, W. E. (1959). The phenolic constituents of Prunus domestica-I. The

quantitative abalysis of phenolic constituents. Journal of Food Science and Agriculture, 10,

63-68.

Velioglu, Y. S., Mazza, G., Gao, L., & Oomah, B. D. (1998). Antioxidant activity and total

phenolics in selected fruits, vegetables, and grain products. Journal of Agricultural and Food

Chemistry, 46(10), 4113-4117.

Viscidi, K. A., Dougherty, M. P., Briggs, J., & Camire, M. E. (2004). Complex phenolic

compounds reduce lipid oxidation in extruded oat cereals. Lebensmittel-Wissenschaft Und-

Technologie-Food Science and Technology, 37(7), 789-796.

Wang, H., Cao, G. H., & Prior, R. L. (1997). Oxygen radical absorbing capacity of anthocyanins.

Journal of Agricultural and Food Chemistry, 45(2), 304-309.

Xu, B., & Chang, S. K. C. (2008). Effect of soaking, boiling, and steaming on total phenolic

content and antioxidant activities of cool season food legumes. Food Chemistry, 110(1), 1-

13.

Zadernowski, R., Nowak-Polakowska, H., & Rashed, A. A. (1999). The influence of heat

treatment on the activity of lipo- and hydrophilic components of oat grain. Journal of Food

Processing and Preservation, 23(3), 177-191.

Page 194: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

175

CHAPTER FIVE

BIOAVAILABILITY OF ANTIOXIDANTS IN EXTRUDED PRODUCTS

PREPARED FROM PURPLE POTATO AND DRY PEA FLOURS

ABSTRACT

Measuring antioxidant activity using a biologically relevant assay is more important for

understanding the role of phytochemicals in-vivo than the use of chemical assays. A cellular

antioxidant activity assay using HepG2 liver cancer cells could provide more biologically

relevant information on bioactive compounds in raw as well as processed food products. The

objective of this study was to investigate the complete phytochemical profiles, antioxidant

activity, cellular antioxidant activity, and their contribution to bioactivity in purple potato flour,

dry pea flour, raw formulations and processed products prepared from the above ingredients

using extrusion cooking. Free phenolics in purple potato flour and dry pea flour contributed 68

and 87%, respectively, to the total phenolics and 64 and 86%, respectively, to the total

antioxidant activity (ORAC value). Caffeic, p-coumaric and ferulic acids were mostly observed

in the bound extracts of raw formulations as well as in extrudates whereas chlorogenic acid was

predominant in free extracts. Extruded products prepared from raw formulations using extrusion

cooking of the above ingredients had either retained or increased the total phenolics, ORAC

antioxidant activity and flavonoids compared to the raw formulations. Extrusion processing

increased the cellular antioxidant activity of extrudates prepared from 50:50 (w/w) of ingredients

than control raw formulations in a dose-dependent manner.

Page 195: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

176

1. INTRODUCTION

Processed fruits and vegetables are generally believed to have fewer naturally occurring

antioxidants than fresh produce resulting in reduced health benefits. One of the major reasons

for measuring vitamin C levels in foods is its perception as an indicator for health benefits of

processed products (1-4). However, Eberhardt et al. (5) observed that vitamin C contributed less

than 0.4% to the total antioxidant activity in apples indicating the importance of phytochemicals

towards total antioxidant activity. Antioxidant compounds derived from fruits and vegetables act

in preventing formation of reactive oxygen, nitrogen, hydroxyl and lipid species either by

scavenging free radicals, or by repairing or removing damaged molecules (6). These bioactive

phytochemicals are proposed to prevent chronic diseases such as cardiovascular, diabetes and

certain forms of cancers (7-10). Bioactive phytochemicals exist in free as well as soluble-

conjugated and bound forms (11). Bound phytochemicals, mostly in cell wall materials, are

difficult to digest in the upper gastrointestine and may be digested by bacteria in the colon to

provide health benefits and reduce the risk of colon cancer (12). Only a small portion of

flavonoids absorbed across the intestinal membrane are partly transformed to glucuronides and

sulfates; however, the majority of the flavonoids are degraded by intestinal microflora. Several

phenolic acids are produced by bacterial enzymes utilizing hydrolysis, dehydroxylation, cleavage

of the heterocyclic oxygen-containing ring and decarboxylation of flavonoids. Reabsorption of

these phenolic acids increases the antioxidant protection with their radical scavenging ability

(13).

Many researchers have reviewed and documented the contributions of phenolic compounds

including anthocyanins, individual phenolic acids and carotenoids present in potato skin and

flesh to antioxidant activity (14-16). Wounding, boiling, baking, drum-drying, freeze-drying and

Page 196: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

177

microwave cooking of white and colored potatoes had different effects on the total phenolic

content (17-19). Potato protein hydrolysate inhibited lipid oxidation of beef patties by

scavenging free radicals (20) and potato peel extracts retarded oxidation in radiation-processed

lamb meat (21).

Legumes are rich sources of protein and dietary fiber. Consumption of legumes help prevent

osteoporosis (22), certain cancers (23), and reduces body lipid accumulation (24). Various

effects of thermal processed cool season legumes on antioxidant activity, phenolic content and

potential health benefits were reported (25, 26). Combining both purple potatoes and dry peas

may produce healthy snack foods incorporating natural color along with positive effects on

human health. Limited investigations have been done on the effect of processing on formulations

from potatoes and legumes on characterizing the phytochemicals profile and contributions to

antioxidant activity. Metabolism, absorption, and bioavailability of health-beneficial

phytochemicals could be improved by combining protein and individual phytochemicals. Liu (9)

reported the additive and synergistic effects of biologically active compounds from fruits,

vegetables, and whole grains are responsible for their health benefits.

The objective of this study was to investigate the complete phytochemical profiles that exist

in free and bound forms as well as their contribution to the total antioxidant activity in the raw

ingredients (purple potato flour and dry pea flour), raw formulations and processed products

prepared from the above ingredients using extrusion cooking. The bioactivity of phytochemicals

in the raw and processed samples was also determined using the cellular antioxidant activity

assay.

Page 197: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

178

2. MATERIALS AND METHODS

2.1. Chemicals and Reagents

Folin-Ciocalteu reagent, sodium nitrite, catechin, sodium borohydride, chloranil, and vanillin

and gallic acid were purchased from Sigma (St. Louis, MO). Sodium hydroxide, hexane,

aluminum chloride, and acetonitrile were obtained from Fisher Scientific (Pittsburgh, PA), while

ethyl acetate, trifluoroacetic acid, methanol, hydrochloric acid, acetic acid, acetone and ethanol

were of analytical grade and were purchased from Mallinckrodt Chemicals (Phillipsburg, NJ).

Tetrahydrofuran and aluminum chloride were purchased from Fisher Scientific (Fair Lawn, NJ).

2.2. Preparation of Samples

‗Purple Majesty‘ potatoes were peeled, sliced, blanched and drum dried in the school of Food

Science pilot plant, Washington State University. Split dry peas and drum dried potato flakes

were pin milled to produce flours at the Processed Foods Research Unit, Western Regional

Research Center, USDA, Albany, CA. Raw formulations were prepared from purple potato flour

(PPF) and dry pea flour (DPF) at selected concentrations (35/65, 50/50 and 65/35 PPF/DPF

w/w). Extruded products were prepared from the raw formulations using a Clextral twin screw

extruder at 130 °C die temperature, 300 rpm and 17% feed moisture. Detailed extrusion

conditions were elaborated in chapter 4. The extrudates were milled to flour using a coffee

grinder and stored at -80 °C until further use. The flours for routine analysis were stored at -20

C.

2.3 Extraction of Free Phenolic Compounds

Free phenolic compounds in the ingredients, raw formulations, or extruded flours were

Page 198: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

179

extracted using the method previously reported in our laboratory (27, 28). Briefly, 1-2 g of flour

sample was homogenized with 50 mL of 80% chilled acetone for 10 min using a high speed

homogenizer. After centrifugation of homogenized flour at 2500g for 5 min, the supernatant was

removed and extraction was repeated one more time. Supernatants were pooled, evaporated at 45

C to dryness and reconstituted with water to a final volume of 10 mL. The extracts were stored

at -75 °C until use.

2.4. Extraction of Bound Phenolic Compounds

Bound phenolics of the ingredients, raw formulations, or extruded flours were extracted

using the method previously reported in our laboratory (11, 28, 29). Briefly, 1-2 grams of flour

samples were extracted twice with 80% chilled acetone with centrifugation at 2500g for 5 min,

and the supernatant was discarded after each extraction. The residues were digested with 20 mL

of 2 M sodium hydroxide at room temperature for 1 h with shaking under nitrogen gas. The

mixture was neutralized with an appropriate amount of hydrochloric acid and extracted with

hexane to remove lipids. The final solution was extracted five times with ethyl acetate. The ethyl

acetate fraction was evaporated to dryness. The resulting residues were reconstituted in 10 mL of

water and stored at -75 °C until use.

2.5. Determination of Total Phenolics

The total phenolics of each extract was determined using methods previously described by

Singleton et al. (30) and modified in our laboratory (31, 32). Briefly, 400 µL of deionized water

and 100 µL of a known dilution of the extract or standard solution were added to a test tube.

Folin-Ciocalteu reagent (100 µL) was added to the solution and allowed to react for 6 min. Then,

1 mL of 7% sodium carbonate solution and 800 µL of deionized water was added into the test

Page 199: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

180

tubes, and the mixture mixed well. The color developed for 90 min at room temperature, and

absorbance was read at 760 nm using a MRX II DYNEX spectrophotometer (DYNEX

Technologies, Inc., Chantilly, VA). The measurements for free, bound and total phenolics were

compared to a standard curve of prepared gallic acid solutions and expressed as micrograms of

gallic acid equivalents (GAE) per gram dry weight (DW) sample ± SD for triplicate extracts.

2.6. Determination of Individual Phenolic Acids

Chlorogenic, caffeic, p-coumaric and ferulic acids in sample extracts were quantified using a

reverse phase HPLC procedure employing a Supelcosil LC-18-DB, 150 mm x 4.6 mm, 3 mm

column as reported previously (33). Briefly, isocratic elution was conducted with 20%

acetonitrile in water adjusted to pH 2 with trifluoroacetic acid, at a flow rate of 1.0 mL/min,

delivered using a Waters 515 HPLC pump (Waters Corp., Milford, MA). A Waters 2487 dual

wavelength absorbance detector (Waters Corp.) was used for UV detection of analytes at 280

nm. Data signals were acquired and processed on a PC running the Waters Millennium software,

version 3.2 (1999) (Waters Corp.). The retention times of the standards, chlorogenic, caffeic, p-

coumaric and ferulic acids, were 3.3, 4.7, 7.4 and 8.4 min, respectively. The phenolic acid

concentrations of sample extracts were extrapolated from the pure phenolic acid standard curves

(chlorogenic acid, r2=0.99; caffeic acid, r

2=0.99; p-coumaric acid, r

2=0.99; and ferulic acid,

r2=0.99). Twenty microliter injections were made in each run, and areas were used for all

calculations. The selected individual peaks were identified by the retention times and co-

injection of the pure standards. The method was validated by the recovery of chlorogenic acid,

and the percentage recovery for chlorogenic acid was 96.5 ± 6.22 (n=3).

Page 200: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

181

2.7. Determination of Total Flavonoids

The total flavonoid contents of the samples were determined using the method of Sodium

Borohydride/Chloronil (SBC) total flavonoid assay with modifications (34). Briefly, stored

sample extracts for total phenolic analysis were thawed and added into test tubes (15 × 150 mm),

then dried at 45 °C under nitrogen gas, and reconstituted in 1 mL of THF/EtOH (1:1, v/v).

Catechin standards (0.3-10.0 mM) were prepared fresh each day before use in 1 mL of

THF/EtOH (1:1, v/v). Each test tube with 1 mL of sample solution or 1 mL of catechin standard

solution had 0.5 mL of 50 mM NaBH4 solution and 0.5 mL of 74.56 mM AlCl3 solution added

and was then shaken in an orbital shaker (Laboratory-Line Instruments, Inc., Melrose Park, IL)

for 30 min at room temperature. Then an additional 0.5 mL of NaBH4 solution was added into

each test tube with continual shaking for another 30 min at room temperature. Cold acetic acid

solution (2.0 mL of 0.8 M, 4 °C) was added into each test tube, and the solutions were shaken in

the orbital shaker in the dark for 15 min after thorough mixing. Then, 1 mL of 20 mM chloranil

was added into each tube, which was heated at 95 °C with shaking for 60 min in a reciprocal

shaking bath (Precision Scientific Inc., Chicago, IL). The temperature in the reciprocal shaking

bath was maintained using glycerin. The reaction solutions were cooled using tap water, and the

final volume was brought to 4 mL using methanol. Then, 1 mL of 1052 mM vanillin was added

into each tube, followed by mixing. Concentrated HCl (2 mL of 12 M) was added into each tube,

and the reaction solutions were kept in the dark for 15 min after thorough mixing. Aliquots of the

final reaction solutions (200 μL) were added into each well of a 96-well plate after centrifuging

for 3 min at 2500g, and absorbances were measured at 490 nm using a MRX Microplate Reader

with Revelation work station (Dynex Technologies, Inc., Chantilly, VA). Total flavonoids were

expressed as micrograms of catechin equivalents per gram of dry weight sample. Data were

Page 201: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

182

reported as mean ±SD for three replicates.

2.8. Measurement of Antioxidant Activity (ORAC)

The antioxidant activity of the samples (ingredients, raw formulations and extruded products)

were determined using the oxygen radical absorbance capacity(ORAC) assay described by Prior

et al. (35) and modified in our laboratory (36). Briefly, 20 µL of blank, trolox standard, or

sample extracts in 75 mM potassium phosphate buffer, pH 7.4 (working buffer), was added to

triplicate wells in a black, clear-bottom, 96-well microplate. The triplicate samples were

distributed throughout the microplate and were not placed side-by-side, to avoid any effect on

readings due to location. In addition, no outside wells were used, as use of those wells results in

greater variation. A volume of 200 µL of 0.96 µM fluorescein in working buffer was added to

each well and incubated at 37 C for 20 min, with intermittent shaking, before the addition of 20

µL of freshly prepared 119 mM ABAP in working buffer using a 12-channel pipetter. The

microplate was immediately inserted into a Fluoroskan Ascent FL plate reader

(ThermoLabsystems) at 37 C. The decay of fluorescence at 538 nm was measured with

excitation at 485 nm every 4.5 min for 2.5 h. The areas under the fluorescence versus time curve

for the samples minus the area under the curve for the blank were calculated and compared to a

standard curve of the areas under the curve for 6.25, 12.5, 25, and 50 µM trolox standards minus

the area under the curve for blank. ORAC values were expressed as mean micromoles of trolox

equivalents (TE) per gram of dry weight sample ± SD for triplicate data.

2.9. Cellular Antioxidant Activity (CAA) assay

The assay for CAA was performed following the procedures of Wolfe and Liu (37). Briefly,

HepG2 liver cancer cells from the passages of 5 and 7 were seeded at a density of 6 x 104 in 100

Page 202: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

183

µL complete growth medium per well on a 96-well microplate in a humidified 5 % CO2

incubator at 37 °C. The wells on the boundary of the microplate were filled with 200 µL of PBS.

Twenty-four hours after seeding, the growth medium was removed and the wells were washed

with PBS. The wells were treated in triplicate with 100 µL of solutions containing different

concentrations of antioxidant extracts plus 25 µM DCFH-DA dissolved in antioxidant treatment

media for 1 h at 37 °C. Then treatment mediums were removed and wells were washed with 100

µL of PBS to remove extracellular residues. One hundred microliters of 600 µM ABAP in

oxidant treatment medium (HBSS) was applied to all the cells and the microplate was

immediately placed into a Fluoroskan Ascent FL plate reader (ThermoLabsystems, Franklin,

MA) at 37 °C. Emission was measured for 1 h at 538 nm after excitement at 485 nm in every 5

min. The blank wells contained cells treated with DCFH-DA, HBSS and antioxidant extracts

without ABAP whereas the control wells contained cells treated with DCFH-DA, HBSS and

ABAP without antioxidant extracts.

The area under curve for fluorescence (after subtraction of blank and initial fluorescence

values) versus time was integrated to calculate the CAA value at each concentration of the

sample extracts as follows:

CAA unit = )/(1 CASA

where SA is the integrated area under the sample fluorescence vs time curve and CA is the

integrated area from the control curve. The median effective dose (EC50), i.e. dose required to

give a 50% inhibition for sample extract, was determined from the median effective plot of

log(fa/fu) vs log(dose), where fa is the fraction affected (CAA unit) and fu is the fraction

unaffected (1 – CAA unit) by the treatment. In the experiments, quercetin was used as a

standard.

Page 203: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

184

2.10. Statistical Analyses

All results were reported as mean ± SD for at least three analyses for each type of extraction

and parameter. Results were subjected to ANOVA, and significance of differences between

means were determined using Tukey‘s multiple comparison test run on SAS (version 9.1, SAS

Institute Inc., Cary, NC). Correlations among various parameters were also investigated using

Pearson‘s correlation coefficient.

3. RESULTS

3.1. Phenolic Contents

The phenolic contents of the ingredients, raw formulations, and extruded products were

determined using the Folin-Ciocalteu method and expressed as µg GAE/g DW sample (Figure

5.1). Free, bound, and total phenolics in purple potato flour (PPF) were 2008 ± 118, 932 ± 66,

and 2940 ± 183 µg GAE/g DW sample, respectively. Dry pea flour (DPF) had 366 ± 20, 55 ± 2,

and 421 ± 19 µg GAE/g DW sample of free, bound, and total phenolics, respectively. Extruded

products prepared from 35% PPF had significantly higher (p < 0.05) free phenolic contents

(1947 ± 60 µg GAE/g DW sample) than its raw formulation (919 ± 28 µg GAE/g DW sample).

Similarly, free phenolic contents in the extruded products prepared from 50% PPF (2692 ± 42 µg

GAE/g DW sample) and 65% PPF (3977 ± 36 µg GAE/g DW sample) were significantly

higher(p < 0.05) than their raw formulations (1552 ± 53 and 2290 ± 92 µg GAE/g DW sample,

respectively). The bound phenolic contents in the extruded products prepared from 35, 50 and

65% PPF (175 ± 8, 378 ± 28, and 331 ± 35 µg GAE/g DW sample, respectively) were

significantly lower (p < 0.05) than their raw formulations (415 ± 26, 478 ± 21 and 451 ± 31 µg

GAE/g DW sample, respectively). The total phenolic contents of the extruded products prepared

Page 204: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

185

0

1000

2000

3000

4000

5000

PPF DPF Raw Extruded Raw Extruded Raw Extruded

Ingredients 35% Purple potato

flour formulation

50% Purple potato

flour formulation

65% Purple potato

flour formulation

Ph

eno

lic

con

ten

ts

(µg

ga

llic

aci

d e

q./

g s

am

ple

, D

W)

Free Bound Total

a

b

c

d

cd

d

eef ff

f

g

h

ii

jj j j j j j

kl

Figure 5. 1. Phenolic contents of ingredients, raw formulations, and extruded products prepared

from raw formulations. The formulation contains x% of purple potato flour and (100-x)% of dry

pea flour (DPF). Bars with different letters are significantly different (p < 0.05).

from 35, 50 and 65% PPF (2122 ± 53, 3070 ± 62 and 4308 ± 64 µg GAE/g DW sample,

respectively) were significantly higher (p < 0.05) than their raw formulations (1334 ± 51, 2030 ±

39 and 2741 ± 120 µg GAE/g DW sample, respectively). Table 5.1 shows the percentage

contributions of free and bound phenolics to the total phenolic content of raw ingredients and

extruded products.

3.2. Flavonoid Contents

The flavonoid contents of the raw ingredients, formulations and extruded products were

determined using the SBC assay and were expressed as µg catechin eq/g DW sample (Figure

Page 205: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

186

5.2). The free, bound and total flavonoid contents of potato flour were 5495 ± 42, 724 ± 23, and

6219 ± 417 µg catechin eq/g DW sample, respectively. The free, bound, and total flavonoid

contents of DPF were 1305 ± 36, 731 ± 71, and 2036 ± 72 µg catechin eq/g DW sample,

respectively. The free flavonoid contents of the extruded products prepared from 35, 50 and 65%

PPF were 2606 ± 220, 2929 ± 153, and 3468 ± 64 µg catechin eq/g DW sample, respectively,

and were significantly higher (p < 0.05) than their raw formulations (1944 ± 100, 2196 ± 68, and

2890 ± 103 µg catechin eq/g DW sample, respectively). Extruded products prepared from 35 and

50% PPF had significantly higher (p < 0.05) bound flavonoid contents (592 ± 52 and 470 ± 32

µg catechin eq/g DW sample, respectively) than their raw samples (273 ± 24 and 289 ± 26 µg

catechin eq/g DW sample, respectively). However, the bound flavonoid content in the extruded

products prepared from 65% PPF (290 ± 27 µg catechin eq/g DW sample) was not significantly

different (p > 0.05) from its raw formulation (230 ± 19 µg catechin eq/g DW sample). The total

flavonoid content of the extruded products prepared from 35% PPF (3198 ± 220 µg catechin eq/g

Table 5. 1. Percentage contributions of phytochemicals in free and bound extract of ingredients

(purple potato flour and dry pea flour), raw formulations, and extruded products to total

phenolics, total antioxidant activity and total flavonoids (n=3). Total Phenolics Total Antioxidant Activity

(ORAC value)

Total Flavonoids

Free

(%)

Bound

(%)

Free (%) Bound (%) Free

(%)

Bound

(%)

Purple potato flour (PPF) 68 32 64 36 88 12

Dry pea flour (DPF) 87 13 86 14 64 36

35% PPF* Raw 69 31 63 37 88 12

Extruded 92 8 88 12 82 18

50% PPF Raw 76 24 74 26 88 12

Extruded 88 12 88 12 86 14

65% PPF Raw 84 16 83 17 93 7

Extruded 92 8 93 7 92 8

* The formulation contains x% of purple potato flour and (100-x)% of dry pea flour. Extruded

products were prepared from the raw formulations using extrusion cooking technology.

Page 206: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

187

0

1000

2000

3000

4000

5000

6000

7000

PPF DPF Raw Extruded Raw Extruded Raw Extruded

Ingredients 35% Purple potato

flour formulation

50% Purple potato

flour formulation

65% Purple potato

flour formulation

Fla

vo

no

id c

on

ten

ts

g c

ate

chin

/g s

am

ple

, D

W)

Free Bound Total

a

b

ccdcd

de dee

efe

ffg

fgfg

g

ji

ii

kkkl

h

Figure 5. 2. Flavonoid contents of ingredients, raw formulations and extruded products prepared

from raw formulations. The formulation contains x% of purple potato flour and (100-x)% of dry

pea flour (DPF). Bars with different letters are significantly different (p < 0.05).

DW sample) was significantly higher (p < 0.05) than its raw formulation (2217 ± 124 µg

catechin eq/g DW sample). Similarly, extruded products prepared from 50 and 65% PPF (3400 ±

149 and 3758 ± 89 µg catechin eq/g DW sample, respectively) were significantly higher (p <

0.05) than their raw formulations (2486 ± 53 and 3120 ± 111 µg catechin eq/g DW sample,

respectively).

3.3. Individual Phenolic Acids

Detected individual phenolic acids present in raw formulations and extruded products are

Page 207: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

188

Table 5. 2. Quantity of free, bound, and total individual phenolic acids present in extracts of ingredients, raw formulations and

extruded products (Mean ± SD, n=3)

Free

(µg/g sample, DW)

Bound

(µg/g sample, DW)

Total

(µg/g sample, DW)

Chloro Caffeic p-Coum Ferul Chloro Caffeic p-Coum Ferul Chloro Caffeic p-Coum Ferul

Purple potato

flour 714 ±55

d 14±1

b 52 ±0.3

b n.d. 202 ±8

c 801±55

a 1019 ±34

a 109 ±11

b 916 ±55

e 815 ±56

a 1071 ± 35

a 109 ±11

b

Dry pea flour 2.0 ±1g n.d. n.d. n.d. 236 ±20

b 15 ±3

f n.d. 22 ±3

d 260 ±25

g 15 ±3

f n.d. 22 ±3

d

35%

PPF*

Raw 298 ± 24f 10±1

c 6±1

d n.d. 253±27

b 306±45

b 334±10

e 59±1

c 551±46

f 316±46

b 340±10

e 59±1

c

Exd 984±10c n.d. n.d. n.d. 581±60

a 77±6

e 362±11

d 188±12

a 1565±66

c 77±7

e 362±11

e 188±12

a

50%

PPF

Raw 540±77e 13±1

b 37±8

c n.d. 253±33

b 339±39

b 387±54

cd 57±7

c 794±107

e 353±39

b 425±47

d 57±7

c

Exd 2011±211b n.d. n.d. n.d. 609±60

a 168±17

c 970±113

ab 179±27

a 2620±170

b 168±17

c 970±113

ab 179±27

a

65%

PPF

Raw 1011±42c 23±4

a 117±18

a n.d. 243±21

b 309±52

b 463±35

c 58±4

c 1254±63

d 333±55

b 581±51

c 58±4

c

Exd 2967±267a n.d. n.d. n.d. 574±81

a 93±7

d 857±73

b 159±21

a 3541±348

a 93±7

d 857±73

b 159±21

a

* The formulation contains x% of purple potato flour and (100-x)% of dry pea flour in the formulation. Extruded products were

prepared from the raw formulations using extrusion cooking technology. aValues in each column with no letters in common are

significantly different (p < 0.05). Chloro – Chlorogenic acid; p-Coum – Coumaric acid; n.d. – not detected

Page 208: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

189

0

10

20

30

40

50

60

70

80

90

100

110

120

130

140

PPF DPF Raw Extruded Raw Extruded Raw Extruded

Ingredients 35% Purple potato

flour formulation

50% Purple potato

flour formulation

65% Purple potato

flour formulation

An

tio

xid

an

t A

ctiv

ity

(O

RA

C)

(µm

ol

Tro

lox

eq

./g

sa

mp

le, D

W)

Free Bound Total

a

a

b

bcc c

de

dee

ee

f

f f

g hh h

hhi j i

k

Figure 5. 3 Antioxidant activity of ingredients, raw formulations and extruded products prepared

from raw formulations. The formulation contains x% of purple potato flour and (100-x)% of dry

pea flour (DPF) . Bars with different letters are significantly different (p < 0.05).

presented in Table 5.2, expressed as µg/g DW sample. Chlorogenic acid was the prominent

phenolic acid in the free phytochemicals of the samples whereas caffeic and p-coumaric acids

were prominent in the bound phytochemicals of the samples followed by chlorogenic acid.

Quantity of caffeic, p-coumaric and ferulic acids were either less or not detected in the free

phytochemicals of the raw formulations and extruded products. Chlorogenic acids in the free and

bound phytochemicals of the extruded products were 984 – 2967 µg/g DW sample and 574 –

609 µg/g DW sample, respectively, and were significantly higher (p < 0.05) than their raw

formulations (297 – 1011 µg/g DW sample and 243 – 253 µg/g DW sample, respectively).

Page 209: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

190

Extrusion of raw formulations significantly lowered (p < 0.05) the caffeic acid content (306 –

339 to 77 – 168 µg/g DW sample) whereas processing increased the quantity of p-coumaric acid

(334 – 463 to 362 – 970 µg/g DW sample) in the bound phytochemicals of all raw formulations

except 35% PPF. Quantities of ferulic acid in the extruded products (159 – 188 µg/g DW

sample) were also significantly higher (p < 0.05) than their raw formulations (57 – 59 µg/g DW

sample). Individual phenolic acids except caffeic acid in the total phytochemicals were

significantly increased (p < 0.05) in the extruded products compared to their raw formulations.

However, no significant difference (p > 0.05) was observed in the content of total p-coumaric

acid in the extruded product and raw formulation.

3.4. Total Antioxidant Activity

The total antioxidant activities of the ingredients, raw formulations and extruded products

were determined using the ORAC assay and expressed as µmol TE/g DW sample (Figure 5.3)

The free, bound, and total ORAC values of PPF were 47.82 ± 1.66, 26.80 ± 0.89, and 74.62 ±

2.53 µmol TE/g DW sample, respectively. The free, bound and total ORAC values of DPF

samples were 8.86 ± 0.88, 1.43 ±0.07, and 10.29 ± 0.92 µmol TE/g DW sample, respectively.

The free ORAC values of extruded products prepared from 35, 50 and 65% PPF (45.48 ± 2.34,

75.29 ± 5.32 and 113.83 ± 6.67 µmol TE/g DW sample, respectively) were significantly higher

(p < 0.05) than these of their raw formulations (17.69 ± 1.56, 33.36 ± 4.27 and 57.80 ± 1.44

µmol TE/g DW sample, respectively). The bound ORAC values of extruded products prepared

from 35 and 65% PPF (6.08 ± 0.23, and 8.60 ± 1.42 µmol TE/g DW sample, respectively) were

significantly lower (p < 0.05) than these of their raw formulations (10.28 ± 0.52, and 12.25 ±

0.64 /g DW sample, respectively). However, there was no significant difference (p > 0.05) in the

Page 210: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

191

ORAC values of the bound phytochemicals in the extruded products prepared from 50% PPF

(9.93 ± 0.54 µmol TE /g DW sample) formulation compared to its raw formulation (11.58 ± 1.74

0.000

0.050

0.100

0.150

0.200

PPF DPF Raw Extruded Raw Extruded Raw Extruded

Ingredients 35% Purple potato

flour formulation

50% Purple potato

flour formulation

65% Purple potato

flour formulation

Cel

lula

r A

nti

ox

ida

nt

Act

ivit

y

(µm

ol

qu

erce

tin

eq

./g

sa

mp

le, D

W)

0

100

200

300

400

500

EC

50

va

lue

(mg

/mL

)

CAA

EC50

a

b

d

d

c

c c

D

B C

D

A

CC

Figure 5. 4. Cellular antioxidant activity and EC50 value of ingredients, raw formulations, and

extruded products prepared from raw formulations. The formulation contains x% of purple

potato flour (DPF) and (100-x)% of dry pea flour. Bars with different letters are significantly

different (p < 0.05). Capital and small letters are for EC50 and cellular antioxidant activity,

respectively.

µmol TE /g DW sample). Extruded products prepared from 35, 50 and 65% PPF had

significantly higher (p < 0.05) ORAC values (51.57 ± 2.57, 85.22 ± 5.81 and 122.43 ± 7.52 µmol

TE/g DW sample, respectively) than these of their raw formulations (27.98 ± 1.87, 44.95 ± 5.92

and 70.05 ± 1.00 µmol TE/g DW sample, respectively).

Page 211: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

192

3.5. Cellular Antioxidant Activity

The CAA of raw formulations and extruded products were quantified using the protocol of

the CAA method and expressed as EC50 in mg/mL and CAA values in µmol quercetin equivalent

(QE)/g DW sample (Figure 5.4). The EC50 value of CAA of PPF sample was 52.1 ± 4.0 mg/mL

(CAA = 0.084 µmol QE/g DW sample) whereas no CAA activity was detected in the DPF

sample. The EC50 values of CAA of the extruded product prepared from 35% PPF (64.7 ± 9.0

mg/mL) and 50% PPF (38.9 ± 7.4 mg/mL) were significantly lower (p < 0.05) than these of their

raw formulations (424.2 ± 344 and 195.8 ± 10 mg/mL, respectively). However, no significant

difference in the EC50 values of CAA was observed between the raw (77.8 ± 19 mg/mL) and

extruded products (70.3 ± 4.4 mg/mL) prepared from 65% PPF. Similarly, the CAA values of the

35, 50 and 65% PPF raw formulations (0.017 ± 0.018, 0.025 ± 0.001, and 0.06 ± 0.015 µmol

QE/g DW sample, respectively) and extruded products (0.56 ± 0.008, 0.128 ±0.023, and 0.063 ±

0.004 µmol QE/g DW sample, respectively) followed the same trend as their EC50 values. The

EC50 value of quercetin standard in all the replicates were in the range of 3.35 – 4.5 mg/mL,

which were similar as reported by Wolfe & Liu (37).

3.6. Correlation Analyses

Relationships among total phenolics, total ORAC values and CAA with individual phenolic

acids and flavonoids in free and bound forms were determined using Pearson‘s correlation

coefficient (Table 5.3). Free phenolic content was significantly correlated (p < 0.05) to total

ORAC values (r2 = 0.98) and CAA values (r

2 = 0.66). Similar correlation patterns were observed

between total phenolics and total ORAC values and CAA values. Free and total flavonoids were

significantly correlated (p < 0.05) to total phenolics, ORAC and CAA values. Among individual

phenolic acids, free and total chlorogenic acid contents were significantly correlated (p < 0.05) to

Page 212: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

193

the total phenolics, ORAC and CAA values. p-Coumaric and ferulic acids in the bound and total

extracts significantly correlated (p < 0.05) to the total phenolics, ORAC and CAA values.

However, caffeic acid did not show any positive correlation with phenolics, ORAC or CAA

values. Total ORAC values were significantly correlated (p < 0.05) with CAA values (r2 = 0.69).

4. DISCUSSION

Phytochemicals in food samples are present in both free and bound form (11, 12). Without

accounting for bound phytochemicals, the total phytochemical content will be underestimated

(11). Therefore, studying free and bound phytochemicals provides accurate information on the

total phytochemicals and the contributions of free, soluble conjugated and bound to the total

phenolics and their total antioxidant activity. The cellular antioxidant activity assay, which is

used in this study, is a biologically relevant way to quantify the bioactivity of antioxidants in

food products as it takes into consideration the cellular uptake, metabolism, and distribution of

bioactive compounds in the cell (38). The values of cellular antioxidant activities of samples

were also compared with a chemical assay, ORAC. This study was designed to determine the

phytochemicals, their relationships and contributions to total antioxidant activity in potato and

dry pea flours, raw formulations, and processed snack foods applying extrusion technology on

the raw formulations.

Our study showed that majority of the phenolics in purple potato and dry pea flours were present

in free rather than in bound form (Table 5.1). The quantity of total phenolics in potato flour

prepared from the flesh was within the range (1120 – 12370 µg GAE/g DW sample with skin) of

potato cultivars grown in the Andes mountains of South America (39). Results from a Colorado

potato breeding program have shown similar total phenolic contents in processed (microwaved

and freeze-dried) ‗Purple Majesty‘ potatoes with skin (19). The percentage contribution of free

Page 213: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

194

phenolics to the total was 68% whereas bound phenolics contributed 32% in potato flour (Table

5.1). This is similar to the results of Chu et al., (40) who reported the contributions of 60 and

40% by free and bound phenolics, respectively, to the total phenolics in white potatoes with skin

(40). Therefore, the total phenolic contents reported in the previous literature citations were

underestimated by not including the bound phenolics in potatoes. Bound phenolics might be

associated with the plant cell walls that survive upper gastrointestinal digestion and reach the

colon where most of it is released by intestinal microflora (11, 12). Total phenolic content in dry

pea flour in the present study (Figure 5.1) was 3 – 5 fold lower than in the reported literature

(25, 26). This could be due to (i) use of previously milled flours compared to whole raw legumes

by other researchers; (ii) different solvents for extraction of phenolic compounds (41). Free

phenolics in dry pea flour contributed more (87%) to the total phenolics than the bound ones

(13%). Raw formulations from potato and dry pea flours in selected concentrations had 69 – 84%

free phenolics. Extrusion cooking of raw formulations increased the percentage contributions of

free phenolics to the total and decreased the contributions from bound phenolics (Table 5.1).

Increase in the total phenolicsin the extruded food products (50 – 60%) rather than in raw

formulations could be due to (i) breaking of conjugated phenolics into free phenolics, and (ii)

leaching of soluble fibers, proteins and other non-phenolic soluble components such as mono-,

di- and oligosaccharides (26). Han and Baik (26) reported a similar increase in total phenolics in

cooked and soaked chickpeas, yellow peas, green peas and soybeans. However, in another report,

total phenolics in cooked cool season food legumes were significantly reduced (p < 0.05) when

compared to uncooked samples (25).

Free and bound phytochemicals in potato flour contributed 64 and 36%, respectively, to the

total ORAC antioxidant activity, similar to their contributions to the total phenolics. The total

Page 214: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

195

Table 5. 3. Correlation analysis of phenolics, antioxidant activity (ORAC) and cellular antioxidant activity of ingredients, raw

formulations and extruded product.

Free extract Bound extract Total

Phenolics

Total

ORAC

value

CAA Phenolics

Total

ORAC

value

CAA Total

Phenolics

Total

ORAC

value

CAA

Phenolics 0.9780

(<.0001)

0.6565

(0.0005)

0.3295

(0.1159)

0.3936

(0.0570)

0.9847

(<.0001)

0.6988

(0.0001)

Flavonoids 0.6636

(0.0004)

0.6288

(0.0010)

0.6008

(0.0019)

-0.3210

(0.1262)

-0.2752

(0.1931)

0.0677

(0.7534)

0.5814

(0.0029)

0.5555

(0.0048)

0.5832

(0.0028)

Chlorogenic

acid

0.8843

(<.0001)

0.9176

(<.0001)

0.6047

(0.0017)

0.5117

(0.0106)

0.5583

(0.0046)

0.5605

(0.0044)

0.8529

(<.0001)

0.8897

(<.0001)

0.6181

(0.0013)

Caffeic acid 0.0240

(0.9113)

-0.0655

(0.7610)

-0.0708

(0.7423)

0.1912

(0.3707)

0.1121

(0.6019)

0.2377

(0.2634)

0.1878

(0.3796)

0.1073

(0.6178)

0.2299

(0.2799)

p-Coumaric

acid

0.1575

(0.4622)

0.0926

(0.6668)

0.0880

(0.6827)

0.8402

(<.0001)

0.8329

(<.0001)

0.8551

(<.0001)

0.8465

(<.0001)

0.8318

(<.0001)

0.8532

(<.0001)

Ferulic acid n.d. n.d. n.d. 0.6419

(0.0007)

0.6583

(0.0005)

0.7197

(<.0001)

0.6419

(0.0007)

0.6583

(0.0005)

0.7197

(<.0001)

ORAC

value

0.8762

(<.0001)

0.3514

(0.0923)

0.3188

(0.1290)

0.2749

(0.1936)

0.9847

(<.0001)

0.6984

(0.0001)

Correlation coefficients were calculated as r2. Significant levels are given in the parenthesis.

Page 215: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

196

ORAC value in potato flour was in the range of hydrophilic ORAC values (28.25 – 250.67 µmol

TE/g DW) of Andean potato cultivars (39). Individual ORAC values of free and bound

phytochemicals could not be compared due to limited information in the literature. The total

ORAC value in dry pea flour was comparable to that reported in the literature (42) and had

similar contributions from free (86%) and bound (14%) phytochemicals as was the case in total

phenolics (Table 5.1). Han and Baik (26) reported free and bound phytochemicals contributed

68 and 32% respectively, to the total TEAC antioxidant activity of yellow peas. Processing of all

raw formulations using extrusion cooking significantly increased (p < 0.05) the total ORAC

values of extruded products. Increase in the total ORAC values followed the same pattern as in

total phenolics in the extruded products. It was also observed that ORAC values of bound

phytochemicals significantly decreased (p < 0.05) and ORAC values of free phytochemicals

increased in the extruded products. This phenomenon could be attributed to (i) breaking of

conjugated phytochemicals to release free phytochemicals (31), (ii) prevention of enzymatic

oxidation, and (iii) darker colors of the extruded products indicating formation of Maillard

reaction products having antioxidant properties (43). ORAC values of processed green peas,

yellow peas and chickpeas were significantly increased (p < 0.05) (27 – 114%, 12 – 67% and 25

– 40%, respectively) after pressure boiling as compared to the raw legume (42). However, the

FRAP values and DPPH antioxidant activity of legumes were decreased significantly (p < 0.05)

by conventional and pressure boiling (25, 42). An increase in antioxidant activities of sweet corn,

teas and tomatoes with thermal processing were also reported (4, 31, 44).

Quantities of free flavonoids in potato and dry pea flours were significantly higher (p < 0.05)

when compared to their bound flavonoids. The total flavonoids of both flours followed similar

pattern as free flavonoids, because of the larger contribution from free flavonoids (Table 5.1).

Page 216: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

197

There is limited literature available to compare the free, bound and total flavonoids in colored

potatoes (45). The total flavonoids of dry pea flour in our study was higher when compared to

the reported total flavonoid contents of whole (41, 46) and dehulled yellow peas (47). Quantities

of free flavonoids increased after extrusion cooking of all the raw formulations. Interestingly, the

contents of bound flavonoids were also significantly increased (p < 0.05) in the extruded

products when compared to their raw formulations. Choi et al. (48) reported a significant

increase of free flavonoids in Shiitake mushrooms after heat treatment at 100 and 121 °C when

compare to raw mushrooms. However, the investigators observed decreases in the bound

flavonoids after heat treatment. Increase in flavonoid contents in the extrudates could be

attributed to (i) disruption of plant cell walls providing better extractability, (ii) breaking of

chemical bonds of higher molecular weight polyphenols and forming soluble low molecular

weight polyphenol compounds, and (iii) inter-conversion of flavonoids in different forms (49).

Variyar et al. (50) reported an increase in antioxidant potential of soybeans with the dose of γ-

irradiation due to increased levels of genistin (an isoflavone) and degradation products of

diadzein.

Phenolic acids are sources of dietary phenols, which are linked to cellulose, lignin and

proteins though ester bonds. For example, chlorogenic acid has strong antioxidant activity and is

already demonstrated to have several desirable effects on biochemical processes involved in

carcinogenesis (51). p-Coumaric acid was the major phenolic acid observed in potato flour,

followed by chlorogenic, caffeic and ferulic acid (Table 5.2). Most of the p-coumaric and caffeic

acids were observed in the bound phenolic fraction whereas chlorogenic acid was higher in the

free phenolics. Ferulic acid was observed only in the bound phenolics. Adom and Liu reported

that more than 93% ferulic acids were present in the bound form in corn, wheat, oats and rice

Page 217: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

198

(11). The content of free chlorogenic acid we reported in this study was similar to the total

chlorogenic content in ‗Purple Majesty‘ potatoes found in a Colorado Potato Breeding Program

(19). Interestingly, our results demonstrated the bound phenolics contributed additional ~20% of

total chlorogenic acid, which was commonly underestimated in the published literature. Lewis et

al. (52) and Lachman et al. (53) reported in detail the presence of chlorogenic, caffeic, p-

coumaric, ferulic and other phenolic acids in colored potatoes. Chlorogenic followed by ferulic

and caffeic acids were detected mostly in the bound phenolics in dry pea flour whereas p-

coumaric acid was not detected. Dry pea flour contained only chlorogenic acid in the free

phenolic fraction. The content of total chlorogenic acid in this study was higher when compared

to the reported content of whole yellow peas (25). The investigators, however, reported having

p-coumaric acid along with chlorogenic and gallic acids in yellow peas. A multifold increase in

chlorogenic acid in free and bound phenolic fractions of formulations rather than raw ingredients

could be due to the binding of o-hydroxy phenolic group in chlorogenic acid to protein via the

bidentate hydrogen bond (54). The mechanism for the binding of chlorogenic acid to sunflower

proteins involved both hydrogen bonding and covalent linkages between oxidized phenolics and

nucleophilic amino acid side chains, such as lysine and cysteine (55). Behavior of caffeic acid in

the raw formulations could also be due to the same mechanism of protein-bound phenolics

interactions as chlorogenic acid. Quantities of chlorogenic acids in the free and bound phenolics

in all extruded products were significantly higher (p < 0.05) when compared to their raw

formulations (Table 5.2). A similar pattern was observed in total chlorogenic acid contents in the

extruded products. Xu and Chang (25) reported significant increases in gallic, chlorogenic and

total phenolic acids in yellow peas after pressure boiling. Significant decreases in caffeic acid in

bound fraction and total phenolics were observed in extruded products when compared to their

Page 218: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

199

raw formulations whereas those decreases were not detected in free phenolics. A decrease in

caffeic acid contents with exposure of potato strips to home processing conditions was reported

(56). Total p-coumaric and ferulic acid contents were increased in extruded products compared

to their raw formulations, mostly due to increases in their bound phenolics. However, caffeic, p-

coumaric and ferulic acids were not detected in free phenolic extractions after extrusion of raw

formulations. The undetected phenolic acids in extrudates may be due to alkaline hydrolysis that

partly or completely broke down original phenolic acids. Increases in total chlorogenic, p-

coumaric and ferulic contents in the extruded products could be attributed to the alkaline

hydrolysis of caffeic acid present in conjugated phenolics. The reasons for change in the

individual phenolic acids during extrusion cooking could be explained according to Fleuriet and

Macheix (57) as (i) oxidative degradation of phenolic acids; (ii) break down of conjugated

phenolics and release of free phenolic acids; and (iii) formation of complex phenolic acids.

Increases in the contents of chlorogenic, p-coumaric and ferulic acids and low molecular

phenolic acids from the breakdown of caffeic acid could have increased the total phenolics in the

extruded products. Thermal decomposition of caffeic acid generated compounds such as

tetraoxygenated phenylinadan isomers with higher antioxidant activity than caffeic acid (58).

Analysis of individual phenolic acids suggested that the breakdown of higher polyphenols and

formation of novel compounds after processing could have increased the total phenolics as well

as antioxidant activity. Significant positive correlations among individual phenolic acids except

caffeic acid to the total phenolics and total ORAC antioxidant activity justify these assumptions.

Increases in the total phenolics after processing and a strong correlation (r2 = 0.9847) with

ORAC antioxidant activity showed that phenolics were mostly responsible for the chemical

antioxidant activity in the extruded products.

Page 219: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

200

The bioavailability of phytochemicals in free phytochemical extracts in the raw ingredients,

raw formulations and extruded products were studied using the cellular antioxidant activity

assay. Potato flour showed potent cellular antioxidant activity with an EC50 of 52 mg/mL. The

lower EC50 values indicate higher cellular antioxidant activity. Dry pea flour had no

measureable cellular antioxidant activity. Cellular antioxidant activities were observed in all the

raw formulations and their extruded products (Figure 5.4). Extruded products prepared from

50% PPF had higher cellular antioxidant activity and lower EC50 value among the extrudates (p <

0.05), and also had much higher cellular antioxidant activity (p < 0.05) and lower EC50 value (p <

0.05) than that of its raw formulation. The effect of extrusion on the products prepared from 65%

PPF was not significantly different (p > 0.05) in the EC50 or CAA values. Similar results were

observed in the 35% PPF raw formulation; CAA values were increased in the extrudates when

compared to its raw formulation (Figure 5.4). There was no specific pattern observed in EC50

values of extruded products. It could be possible that products extruded from a 50% PPF raw

formulation had a better composition of potato and dry pea flour for possible additive and

synergistic effects of phytochemicals responsible for the higher cellular antioxidant activity as

suggested by Liu (9). The CAA values of extruded product from 50% PPF formulation was

comparable with those of plums and red grapes and higher than those of cherries, kiwifruit,

mangoes, peaches and pears (36). Extruded products had higher cellular antioxidant activity

when compared to the raw materials, which may be due to extrusion processing increasing the

amount of bioaccessible phytochemicals as well as increased cellular uptake of these

phytochemicals.

In summary, a complete study on the contributions of free and bound phytochemicals to the

total phenolics, flavonoids, ORAC antioxidant activity, and cellular antioxidant activity in

Page 220: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

201

‗Purple Majesty‘ potato and dry pea flours was made. Raw formulations of these ingredients and

extruded products prepared were analyzed in detail with regard to changes in individual phenolic

acids, bound and free phytochemicals, and their contributions to ORAC antioxidant activities

after extrusion. In most of the extruded products, the total phenolics, flavonoids and ORAC

values increased after extrusion processing. Cellular antioxidants were observed in potato flour,

raw formulations, and extruded products. Extrusion processing improved the cellular antioxidant

activity of 50% PPF raw formulation. Measuring the antioxidant activity using the CAA assay in

extruded food products is important for assessing the bioavailability of processed foods because

it is more biologically relevant than chemical antioxidant assays (37). Further studies on the

antioxidant activities of individual phytochemicals and changes during processing are needed to

understand the detailed mechanism of breaking and formation of total phytochemicals and their

contribution to bioavailability.

ABBREVIATIONS USED

PPF, purple potato flour; DPF, dry pea flour; PBS, phosphate-buffer saline; ABAP, 2,2′-

azobis (2-amidinopropane) dihydrochloride; CAA, cellular antioxidant activity; DCFH, 2′,7′-

dichlorofluorescin; DCFH-DA, 2′,7′-dichlorofluorescin diacetate; HBBS, Hank‘s Balanced Salt

Solution; FRAP, ferric reducing/antioxidant power; ORAC, oxygen radical absorbance capacity;

GAE, gallic acid equivalent; QE, quercetin equivalents; TE, Trolox equivalents; TEAC, Trolox

equivalent antioxidant capacity; TRAP, total radical-scavenging antioxidant parameter.

ACKNOWLEDGEMENT

We gratefully acknowledge the assistance of Jams Pan and Matt, Processed Foods Research

Unit, WRRC, USDA-ARS, Albany, CA for their assistance and feedback during the extrusion

Page 221: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

202

experiment. We also acknowledge the US Dry Pea and Lentil Council for their financial support

for this study.

Page 222: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

203

REFERENCES

1. Burg, P.; Fraile, P., Vitamin C destruction during the cooking of a potato dish.

Lebensmittel-Wissenschaft und-Technologie 1995, 28, (5), 506-514.

2. Lathrop, P. J.; Leung, H. K., Rates of ascorbic acid degradation during thermal

processing of canned peas. Journal of Food Science 1980, 45, (1), 152-153.

3. Murcia, M A.; López-Ayerra, B.; Martinez-Tomé, M.; Vera, A. M.; García-Carmona, F.,

Evolution of ascorbic acid and peroxidase during industrial processing of broccoli.

Journal of the Science of Food and Agriculture 2000, 80, (13), 1882-1886.

4. Dewanto, V.; Wu, X. Z.; Adom, K. K.; Liu, R. H., Thermal processing enhances the

nutritional value of tomatoes by increasing total antioxidant activity. Journal of

Agricultural and Food Chemistry 2002, 50, (10), 3010-3014.

5. Eberhardt, M. V.; Lee, C. Y.; Liu, R. H., Nutrition - antioxidant activity of fresh apples.

Nature 2000, 405, (6789), 903-904.

6. Liu, R. H.; Hotchkiss, J. H., Potential genotoxicity of chronically elevated nitric oxide - a

review. Mutation Research-Reviews in Genetic Toxicology 1995, 339, (2), 73-89.

7. Cohen, J. H.; Kristal, A. R.; Stanford, J. L., Fruit and vegetable intakes and prostate

cancer risk. Journal of the National Cancer Institute 2000, 92, (1), 61-68.

8. Lunet, N.; Lacerda-Vieira, A.; Barros, H., Fruit and vegetables consumption and gastric

cancer: A systematic review and meta-analysis of cohort studies. Nutrition and Cancer-

an International Journal 2005, 53, (1), 1-10.

9. Liu, R. H., Potential synergy of phytochemicals in cancer prevention: mechanism of

action. J. Nutr. 2004, 134, (12), 3479S-3485.

Page 223: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

204

10. Liu, R. H., Health benefits of fruit and vegetables are from additive and synergistic

combinations of phytochemicals. Am J Clin Nutr 2003, 78, (3), 517S-520.

11. Adom, K. K.; Liu, R. H., Antioxidant activity of grains. Journal of Agricultural and Food

Chemistry 2002, 50, (21), 6182-6187.

12. Liu, R. H., Whole grain phytochemicals and health. Journal of Cereal Science 2007, 46,

(3), 207-219.

13. Pietta, P.-G., Flavonoids as antioxidants. Journal of Natural Products 2000, 63, (7),

1035-1042.

14. Friedman, M., Chemistry, biochemistry, and dietary role of potato polyphenols. A

review. Journal of Agricultural and Food Chemistry 1997, 45, (5), 1523-1540.

15. Alsaikhan, M. S.; Howard, L. R.; Miller, J. C., Antioxidant activity and total phenolics in

different genotypes of potato (Solanum tuberosum, L). Journal of Food Science 1995, 60,

(2), 341-347.

16. Brown, C. R.; Wrolstad, R.; Durst, R.; Yang, C. P.; Clevidence, B., Breeding studies in

potatoes containing high concentrations of anthocyanins. American Journal of Potato

Research 2003, 80, (4), 241-249.

17. Nayak, B.; Berrios, J. J.; Powers, J. R.; Tang, J.; Ji, Y., Colored potatoes (Solanum

tuberosum L.) dried for antioxidant-rich value-added foods. Journal of Food Processing

and Preservation (in Press) 2010.

18. Reyes, L. F.; Cisneros-Zevallos, L., Wounding stress increases the phenolic content and

antioxidant capacity of purple-flesh potatoes (Solanum tuberosum L.). Journal of

Agricultural and Food Chemistry 2003, 51, (18), 5296-5300.

Page 224: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

205

19. Stushnoff, C.; Holm, D.; Thompson, M. D.; Jiang, W.; Thompson, H. J.; Joyce, N. I.;

Wilson, P., Antioxidant properties of cultivars and selections from the Colorado potato

breeding program. American Journal of Potato Research 2008, 85, (4), 267-276.

20. Wang, L. L.; Xiong, Y. L. L., Inhibition of lipid oxidation in cooked beef patties by

hydrolyzed potato protein is related to its reducing and radical scavenging ability.

Journal of Agricultural and Food Chemistry 2005, 53, (23), 9186-9192.

21. Kanatt, S. R.; Chander, R.; Radhakrishna, P.; Sharma, A., Potato peel extract - a natural

antioxidant for retarding lipid peroxidation in radiation processed lamb meat. Journal of

Agricultural and Food Chemistry 2005, 53, (5), 1499-1504.

22. Messina, M. J., Legumes and soybeans: overview of their nutritional profiles and health

effects. American Journal of Clinical Nutrition 1999, 70, (3), 439s-450s.

23. Lamartiniere, C. A., Protection against breast cancer with genistein: a component of soy.

American Journal of Clinical Nutrition 2000, 71, (6), 1705s-1707s.

24. Pusztai, A.; Grant, G.; Buchan, W. C.; Bardocz, S.; de Carvalho, A. F. F. U.; Ewen, S. W.

B., Lipid accumulation in obese Zucker rats is reduced by inclusion of raw kidney bean

(Phaseolus vulgaris) in the diet. British Journal of Nutrition 1998, 79, (2), 213-221.

25. Xu, B.; Chang, S. K. C., Phytochemical profiles and health-promoting effects of cool-

season food legumes as influenced by thermal processing. Journal of Agricultural and

Food Chemistry 2009, 57, (22), 10718-10731.

26. Han, H.; Baik, B. K., Antioxidant activity and phenolic content of lentils (Lens culinaris),

chickpeas (Cicer arietinum L.), peas (Pisum sativum L.) and soybeans (Glycine max),

and their quantitative changes during processing. International Journal of Food Science

and Technology 2008, 43, (11), 1971-1978.

Page 225: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

206

27. Sun, J.; Chu, Y. F.; Wu, X. Z.; Liu, R. H., Antioxidant and anti proliferative activities of

common fruits. Journal of Agricultural and Food Chemistry 2002, 50, (25), 7449-7454.

28. de la Parra, C.; Serna Saldivar, S. O.; Liu, R. H., Effect of processing on the

phytochemical profiles and antioxidant activity of corn for production of masa, tortillas,

and tortilla chips. Journal of Agricultural and Food Chemistry 2007, 55, (10), 4177-

4183.

29. Adom, K. K.; Sorrells, M. E.; Liu, R. H., Phytochemical profiles and antioxidant activity

of wheat varieties. Journal of Agricultural and Food Chemistry 2003, 51, (26), 7825-

7834.

30. Singleton, V. L.; Orthofer, R.; Lamuela-Raventos, R. M., Analysis of total phenols and

other oxidation substrates and antioxidants by means of Folin-Ciocalteu reagent.

Oxidants and Antioxidants, Pt A 1999, 299, 152-178.

31. Dewanto, V.; Wu, X. Z.; Liu, R. H., Processed sweet corn has higher antioxidant activity.

Journal of Agricultural and Food Chemistry 2002, 50, (17), 4959-4964.

32. Adom, K. K.; Sorrells, M. E.; Liu, R. H., Phytochemicals and antioxidant activity of

milled fractions of different wheat varieties. Journal of Agricultural and Food Chemistry

2005, 53, (6), 2297-2306.

33. Okarter, N.; Liu, R. H., Health benefits of whole grain phytochemicals. Critical Reviews

in Food Science and Nutrition 2010, 50, (3), 193-208.

34. He, X.; Liu, D.; Liu, R. H., Sodium borohydride/chloranil-based assay for quantifying

total flavonoids. Journal of Agricultural and Food Chemistry 2008, 56, (20), 9337-9344.

35. Prior, R. L.; Hoang, H.; Gu, L. W.; Wu, X. L.; Bacchiocca, M.; Howard, L.; Hampsch-

Woodill, M.; Huang, D. J.; Ou, B. X.; Jacob, R., Assays for hydrophilic and lipophilic

Page 226: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

207

antioxidant capacity (oxygen radical absorbance capacity (ORAC(FL))) of plasma and

other biological and food samples. Journal of Agricultural and Food Chemistry 2003, 51,

(11), 3273-3279.

36. Wolfe, K. L.; Kang, X.; He, X.; Dong, M.; Zhang, Q.; Liu, R. H., Cellular antioxidant

activity of common fruits. Journal of Agricultural and Food Chemistry 2008, 56, (18),

8418-8426.

37. Wolfe, K. L.; Liu, R. H., Cellular antioxidant activity (CAA) assay for assessing

antioxidants, foods, and dietary supplements. Journal of Agricultural and Food

Chemistry 2007, 55, (22), 8896-8907.

38. Spencer, J. P. E.; El Mohsen, M. M. A.; Rice-Evans, C., Cellular uptake and metabolism

of flavonoids and their metabolites: implications for their bioactivity. Archives of

Biochemistry and Biophysics 2004, 423, (1), 148-161.

39. Andre, C. M.; Ghislain, M.; Bertin, P.; Oufir, M.; Herrera, M. D.; Hoffmann, L.;

Hausman, J. F.; Larondelle, Y.; Evers, D., Andean potato cultivars (Solanum tuberosum

L.) as a source of antioxidant and mineral micronutrients. Journal of Agricultural and

Food Chemistry 2007, 55, (2), 366-378.

40. Chu, Y. F.; Sun, J.; Wu, X. Z.; Liu, R. H., Antioxidant and anti proliferative activities of

common vegetables. Journal of Agricultural and Food Chemistry 2002, 50, (23), 6910-

6916.

41. Xu, B. J.; Chang, S. K. C., A comparative study on phenolic profiles and antioxidant

activities of legumes as affected by extraction solvents. Journal of Food Science 2007,

72, (2), S159-S166.

Page 227: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

208

42. Xu, B.; Chang, S. K. C., Effect of soaking, boiling, and steaming on total phenolic

contentand antioxidant activities of cool season food legumes. Food Chemistry 2008,

110, (1), 1-13.

43. Nicoli, M. C.; Anese, M.; Parpinel, M., Influence of processing on the antioxidant

properties of fruit and vegetables. Trends in Food Science & Technology 1999, 10, (3),

94-100.

44. Manzocco, L.; Anese, M.; Nicoli, M. C., Antioxidant properties of tea extracts as affected

by processing. Food Science and Technology-Lebensmittel-Wissenschaft & Technologie

1998, 31, (7-8), 694-698.

45. Christine, E. L.; John, R. L. W.; Jane, E. L., Changes in anthocyanin, flavonoid and

phenolic acid concentrations during development and storage of coloured potato

(Solanum tuberosum L) tubers. Journal of the Science of Food and Agriculture 1999, 79,

(2), 311-316.

46. Xu, B. J.; Yuan, S. H.; Chang, S. K. C., Comparative analyses of phenolic composition,

antioxidant capacity, and color of cool season legumes and other selected food legumes.

Journal of Food Science 2007, 72, (2), S167-S177.

47. Aharon, S.; Hana, B.; Yoram, K.; Ilan, S.; Michal, O.-S.; Shmuel, G., Determination of

polyphenols, flavonoids, and antioxidant capacity in colored chickpea (Cicer arietinum

L.). Journal of Food Science 2010, 75, (2), S115-S119.

48. Choi, Y.; Lee, S. M.; Chun, J.; Lee, H. B.; Lee, J., Influence of heat treatment on the

antioxidant activities and polyphenolic compounds of Shiitake (Lentinus edodes)

mushroom. Food Chemistry 2006, 99, (2), 381-387.

Page 228: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

209

49. Coward, L.; Smith, M.; Kirk, M.; Barnes, S., Chemical modification of isoflavones in

soyfoods during cooking and processing. American Journal of Clinical Nutrition 1998,

68, (6), 1486s-1491s.

50. Variyar, P. S.; Limaye, A.; Sharma, A., Radiation-induced enhancement of antioxidant

contents of soybean (Glycine max Merrill). Journal of Agricultural and Food Chemistry

2004, 52, (11), 3385-3388.

51. Feng, R. T.; Lu, Y. J.; Bowman, L. L.; Qian, Y.; Castranova, V.; Ding, M., Inhibition of

activator protein-1, NF-kappa B, and MAPKs and induction of phase 2 detoxifying

enzyme activity by chlorogenic acid. Journal of Biological Chemistry 2005, 280, (30),

27888-27895.

52. Lewis, C. E.; Walker, J. R. L.; Lancaster, J. E.; Sutton, K. H., Determination of

anthocyanins, flavonoids and phenolic acids in potatoes. I: Coloured cultivars of Solanum

tuberosum L. Journal of the Science of Food and Agriculture 1998, 77, (1), 45-57.

53. Lachman, J.; Hamouz, K.; Orsak, M., Red and purple potatoes - A significant antioxidant

source in human nutrition. Chemicke Listy 2005, 99, (7), 474-482.

54. Mcmanus, J. P.; Davis, K. G.; Beart, J. E.; Gaffney, S. H.; Lilley, T. H.; Haslam, E.,

Polyphenol interactions .1. Introduction - Some observations on the reversible

complexation of polyphenols with proteins and polysaccharides. Journal of the Chemical

Society-Perkin Transactions 2 1985, (9), 1429-1438.

55. Sastry, M. C. S.; Rao, M. S. N., Binding of chlorogenic acid by the isolated polyphenol-

free 11s protein of sunflower (Helianthusa annuus) seed. Journal of Agricultural and

Food Chemistry 1990, 38, (12), 2103-2110.

Page 229: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

210

56. Im, H. W.; Suh, B.-S.; Lee, S.-U.; Kozukue, N.; Ohnisi-Kameyama, M.; Levin, C. E.;

Friedman, M., Analysis of phenolic compounds by high-performance liquid

chromatography and liquid chromatography/mass spectrometry in potato plant flowers,

leaves, stems, and tubers and in home-processed potatoes. J. Agric. Food Chem. 2008,

56, (9), 3341-3349.

57. Fleuriet, A., Macheix, J. J., Phenolics in fruits and vegetables. In Flavonoids in Health

and Disease, Rice-Evans, C., Packer, L, Ed. CRC Press: Boca Raton, FL, 2003; pp 1 - 42.

58. Chen, J. H.; Ho, C. T., Antioxidant activities of caffeic acid and its related

hydroxycinnamic acid compounds. Journal of Agricultural and Food Chemistry 1997,

45, (7), 2374-2378.

Page 230: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

211

CHAPTER SIX

THERMAL DEGRADATION OF ANTHOCYANINS FROM PURPLE

POTATOES (‘PURPLE MAJESTY’ CV) AND IMPACT ON

ANTIOXIDANT CAPACITY.

ABSTRACT

Thermal degradation kinetics of purified anthocyanins from purple-fleshed potatoes (‗Purple

Majesty‘ cv.) was studied over a temperature range of 100 and 150 °C. Anthocyanin was

prepared by removing salts, sugars and colorless non-anthocyanin phenolics from the crude

extract were quantified using HPLC and spectrophotometry for heat induced degradation

compounds. Colorless phenolics from the anthocyanins were separated using HPLC and

monitored at two wavelengths, i.e., 280 and 520 nm. The thermal degradation of purified

anthocyanins followed a first order reaction with reaction rate constants (k-values) of 0.0262 –

0.2855 min-1

, activation energy of 72.89 kJ/mol, thermal death times (D-values) of 8.06 – 8789

min and z-value of 47.84 °C. The enthalpy and entropy of activation was 59.97 kJ/mol and –

116.46 J/mol K, respectively. But the total antioxidant capacity in the thermally treated samples,

measured using DPPH and ABTS assay, were retained indicating antioxidant activities of

degradation compounds in the samples.

1. INTRODUCTION

Anthocyanins are one of the most important groups of water soluble plant pigments largely

responsible for the cyanic colors of flowers, fruits, vegetables and grains accumulating in the

vacuoles of epidermal or subepidermal cells (1). Recently, anthocyanins have gained increasing

Page 231: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

212

attention as functional compounds for natural color as a replacement for synthetic dyes in food

products (2). In addition, these compounds have health related beneficial antioxidative (3-5) and

anticarcinogenic properties (6, 7). Mishra et al. (8) studied the degradation of anthocyanins

above 100 °C and suggested that anthocyanins could be used as food colorants in high

temperature processes such as for extruded snacks or baked cakes. However, successful use of

anthocyanins either as natural colors or as nutraceuticals depends on their physical and chemical

stability during different processing conditions.

Commercial production of colored potatoes such as the ‗Purple Majesty‘ cultivar has

attractive taste and appearance. An added benefit of colored potatoes is their high antioxidant

capacity mainly due to the presence of polyphenols including anthocyanins (9, 10). Stushnoff et

al. (10) reported five petunidin glycosides and single glycoside of each of malvidin, peonidin and

delphidin aglycones in ‗Purple Majesty‘ potatoes. Various mechanisms of stability of

anthocyanins such as association between pigments and cofactors (polyphenols, metal ions, other

anthocyanins) have been proposed in numerous reports (11-13). Acylation of aromatic acids to

the structure (14, 15) and side chain double bond (16) has also been attributed to the stability of

anthocyanins.

Besides the structure, stability of anthocyanins depends on temperature, light, enzymes,

metal ions, sugars, ascorbic acid, oxygen, and presence of other phenolic compounds, (17).

Several studies have reported on the degradation of anthocyanins in fruits and vegetables during

processing and storage (18, 19). The investigators either assumed or reported first order kinetics

for thermal degradation of anthocyanin in selected fruits and vegetables (20-23). Those studies

were carried out either at temperatures below 100 °C or in whole pulp puree/extract. In general,

extracts from whole pulp of fruits and vegetables or fruit and vegetable contain anthocyanins

Page 232: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

213

along with other compounds such as salts, sugars and other colorless non-anthocyanin phenolics

that could affect the stability or degradation kinetics and antioxidant capacity of anthocyanins

(12, 24). Little information is available regarding the thermal kinetics of purified anthocyanins

and antioxidant potencies of degradation compounds. Such information is desirable as reference

starting point to study the interaction of anthocyanin with other compounds in more complicated

food matrix. Stintzing et al., (4) reported color, visual detection thresholds, hydration constants

and ORAC antioxidant activities of purified cyanidin-based anthocyanins. Stability and

antioxidant activities of several purified anthocyanins have been affected by the B-ring structure

and glycosylation (14, 19, 25). In our previous study, extrusion cooking of products prepared

from ‗Purple Majesty‘ potatoes and dry peas retained total antioxidant activity brought by the

raw ingredient materials used to produce extrudates in spite of 60 – 70% decrease in anthocyanin

content. We hypothesized that at high temperatures, such as in extrusion cooking; the

degradation compounds from anthocyanin could be responsible for the overall antioxidant

activity in the extrudates. In the present study, total anthocyanins purified from ‗Purple Majesty‘

potatoes was used to evaluate the thermal kinetics parameters over a temperature range of 100 –

150 °C and antioxidant potencies of degradation compounds from the anthocyanins.

2. MATERIALS AND METHODS

2.1 Chemicals

Folin-Ciocalteu reagents, potassium chloride, sodium acetate, gallic acid, trolox (6-Hydroxy-

2,5,7,8-tetramethylchromane-2-carboxylic acid), DPPH (2,2-diphenyl-1-picrylhydrazyl),

acetonitrile, ABTS (2,2'-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid)), CHCA (α-Cyano-4-

hydroxycinnamic acid), ethyl acetate, and potassium persulfate were purchased from Sigma-

Page 233: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

214

Aldrich (St. Louis, MO, USA). Laboratory grade methanol was used in extraction and

preparation of samples.

2.2. Materials

Fresh purple potatoes (‗Purple Majesty‘ cv) were purchased from Colorado State University

in Fall 2008. Potatoes were washed with tap water and stored at 4 °C in the Washington State

University Food Processing Pilot Plant. To prepare flakes, potatoes were peeled using a

mechanical abrasive peeler for about 75 seconds, sliced to 6mm thickness pieces and steam

blanched for 8 min to inactivate polyphenolic oxidase (PPO) similar to the procedures reported

for peroxidase inactivation (22). The puree was then dehydrated to approximately 5.2% (wet

basis) using a 15.24cm × 20.32 cm pilot scale counter rotating twin drum dryer (Blaw Knox

Food & Chemical Equipment Co., Buffalo, NY, USA). Pressurized steam to the drums was

maintained at 413 kPa corresponding to a saturation temperature of water at 145 °C. The surface

temperature of the drums was 135 - 138 °C. The gap between the drums, rotating at 1.13 rpm,

was set to 0.3 mm. Potato flakes were cooled at room temperature and stored at -20 °C for

further analysis. Moisture content of the flakes were determined using the standard procedures of

AACC moisture-air-oven method number 44-45A (26) and the samples were expressed in dry

weight (DW) basis.

2.3. Extraction of anthocyanins

Purple potato flakes were ground using a food processer and passed through a sieve US # 35

(0.5 mm). Fifty g of flakes were homogenized in 500 ml of extraction solvent (1.5 N HCl:

Methanol: water; 10/70/20 v/v/v) using a high speed homogenizer (Omni Mix Homogenizer,

Omni International, Waterbury, CT, USA). The homogenate was filtered through a double layer

Page 234: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

215

of cheese cloth after keeping the mixture at 4 °C for 90 min. The residue was again homogenized

twice in 250 ml of extraction solvent at room temperature and filtered. All filtrates were pooled

and stored at -20 °C for further analyses.

2.4. Purification of anthocyanins

Potato extracts were centrifuged at 23000g at 4 °C for 15 min using a centrifuge (Beckman

J2-HS Centrifuge, USA) and 10 ml of clear supernatants were loaded to a Sep-Pak C18 column

(part no: WAT023635, Waters, Milford, MA) previously activated with HPLC grade methanol

followed by 0.01% aqueous HCl. Anthocyanins and polyphenolics were adsorbed onto the

column (silica-based bonded phase with strong hydrophobicity) while sugars, acids, and other

water-soluble compounds were removed by washing the column with 20 ml 0.01% aqueous HCl.

The column was further washed with 20 ml ethyl acetate to remove colorless non-anthocyanin

phenolics. Anthocyanins from the extract were collected by washing the column with 40 ml

0.01% HCl in methanol. The anthocyanin-rich extract was dried with a rotary evaporator at 30

°C to dryness and re-suspended in 10 ml deionized water. The pH of the aqueous anthocyanin

extracts ranged from of 5.9 – 6.0. These samples were stored at -20 °C for thermal kinetics study.

Purity of anthocyanins was checked with HPLC at a wavelength of 280 nm.

2.5. Heat treatment of anthocyanins

Thermal kinetics test (TKT) cells (Figure 6.1) were used in an oil bath containing silicon oil

for heat treatment. The test cell was custom designed to reduce the come-up time (CUT). Stored

sample was thawed to room temperature before transferring 1 ml of sample into the TKT cells.

Based on preliminary results of total anthocyanin contents during heat treatment, duplicate

Page 235: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

216

Figure 6. 1. (A) Specially designed thermal kinetics test (TKT) cells used for heat treatment of

purified anthocyanins (darker area) from ‗Purple Majesty‘ potatoes over a temperature range of

100 -150 °C. (B) TKT cells showing the detail dimensions.

Table 6. 1. Experimental design of heat treatments with selected temperature and time

combinations. Time = 0 min at come-up-time.

Temperature (°C) Heating time (min)

100 0 5 15 30 60

110 0 5 15 30 60

120 0 5 15 30 60

130 0 5 15 30 45

140 0 5 10 15 30

150 0 3 5 10 20

samples were considered. Total anthocyanin content in the purified extracts was measured over a

temperature range of 100 – 150 °C. During the treatment, sample temperatures were measured

using a pre-calibrated type T copper-constantan thermocouple with a diameter of 0.1 mm and

recorded at 0.5 sec intervals with a temperature data logger (TracerDAQ Pro, Measurement

Computing, Norton, MA, USA). The come-up-time (CUT e.g., time to reach 0.5 °C below the

set temperature) was 60 – 135 sec. The treatment time zero was considered to be the end of the

CUT interval. Details for the experimental design of thermal treatment are given in Table 6.1.

The samples in the test cells were cooled immediately in ice water after treatments to reduce the

A B

Page 236: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

217

thermal shock. Samples were collected for the cell and stored in 2 ml centrifuge tubes at -20 °C

for further quantification, HPLC and mass spectrometry analyses.

2.6. HPLC Analyses

The analytical reverse phase HPLC technique was used for identification and separation of

anthocyanins. Stored samples (control and heated) were thawed at room temperature, each

sample passed through a Whatman 0.45 μm NYL filter to vials before been applied to HPLC.

The experiments were carried out using a Varian Star HPLC solvent delivery and control system

with automatic sample injector and a variable Varian UV/Vis detector. A Varian Microsorb-MV

100-5 C18 250 x 4.6 mm column fitted with a 10 x 3 mm Varian Chromo guard column was

maintained at room temperature. Twenty μL samples were injected for each test. Two elution

solvents used were 10% (v/v) acetic acid (A) and 50% (v/v) aqueous acetonitrile (B). Flow rate

of the elution solvents was maintained at 1 mL/min with a linear 30 min gradient from 0 to 30%

B followed by 5 min hold at 30% B. The column was then washed with 50% B for 5 min and

then returned to 0% B. The column was re-equilibrated for 10 min before the next analysis. The

eluted compounds were monitored at 280 nm for phenolics and 520 nm for anthocyanins (27).

Standards of gallic, protocatechuic, chlorogenic, caffeic, p-coumaric and ferulic acid with

residence times of 3.336, 4.699, 8.083, 9.664, 15.893 and 20.643 min, respectively, (Figure 6.2)

were used to determine the presence of these colorless phenolic acids in the anthocyanin extracts.

2.6. Matrix-assisted laser desorption ionization (MALDI) mass spectroscopy

MALDI-TOF-MS is a rapid technique to identify a group of anthocyanins with different

masses. The mass spectrometry analyses were performed using a 4800 Plus MALDI TOF/TOF

Analyzer (Applied Biosystems, Framington , MA, USA). The positive ion reflector mode was

used. The instrument was equipped with a pulsed nitrogen laser (337 nm, 3 ns pulse duration,

Page 237: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

218

5 10 15 20 25 30 35

Minutes

-5

0

10

20

30

40

Ab

sorb

an

ce (

mA

U)

20.6436. Ferulic acid

15.8935. p-coumaric acid

9.6644. Caffeic acid

8.0833. Chlorogenic acid

4.6992. Protocatechuic acid

3.3361. Gallic acid

Retention

time (min)

Standard

12

3

4

5

6

Figure 6. 2. Chromatogram of standards as detected using HPLC

3Hz frequency). All spectra were obtained by averaging 25 laser shots. The matrix used in this

study was α-cyano-4-hydroxycinnamic acid (CHCA). Five mg of CHCA were solubilized in

1mL acetonitrile/water mixture (1:1 v/v/) for preparing the matrix stock solution. The sample to

be analyzed was prepared by mixing purified anthocyanin extracts (control or heat treated) with

the matrix (1:1 v/v). The sample mixtures were applied to the plate and analyzed.

2.7. Measurement of anthocyanins

Determination of total anthocyanin was based on the pH differential method following the

procedures of Giusti & Wrolstad (27). Absorbance readings were taken at a maximum

wavelength (λ-max) of 535nm and 700 nm for correcting for turbidity (28) in an UV/visible

Page 238: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

219

spectrophotometer, previously blanked with distilled water. Total anthocyanins were quantified

and expressed in mg equivalent malvidin-3-glucoside (29-31) according to the following

formula:

de

DFMWALmgC

)/( (1)

where A = absorbance of the sample, MW = molecular weight of malvidin-3-glycoside (718.5

g/mol), DF = dilution factor, d = path length of the cuvette (1 cm), and e = molar extinction

coefficient of malvidin-3-glucoside (30200 l/cm per mol). The absorbances of the samples were

calculated as 5.4700max0.1700max )()( pHpH AAAAA .

2.8. Measurement of Antioxidant Capacity

Antioxidant capacities of the samples were measured using DPPH and ABTS assay.

DPPH assay: DPPH, a stable radical, deep purple in color, is reduced in the presence of

antioxidants. The loss of color results in a decrease in the absorbance intensity, thus providing a

basis for measurement of antioxidant activities in the extracts. The assay was based on the

procedures of Brandwilliams et al. (32). Control or heat treated sample (0.05 ml) was added to

1.95 ml of 6 × 10-5

M DPPH solution in a cuvette and the absorbance readings at 515 nm were

taken after 2 h of equilibration time using an Ultraspec 4000 UV/visible spectrophotometer

(Pharmacia Biotech, Cambridge, England).

ABTS assay: Total antioxidant capacity of the samples was also quantified using ABTS

radical cation decolorization assay (33). Working solution for the assay was prepared by mixing

stock solutions of 7 mM ABTS and 2.4 mM potassium persulfate in equal volumes and allowing

them to react for 12 h at room temperature in the dark. One ml of ABTS.+

solution was diluted

with 60 ml methanol to obtain an absorbance of 0.70 ± 0.01 units at 734 nm. Fresh ABTS.+

Page 239: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

220

solution was prepared for each assay. One ml of sample was added to 1 ml of diluted ABTS.+

solution and allowed to equilibrate for 7 min before measuring the absorbance at 734 nm.

In both antioxidant assays, the spectrophotometer was blanked with methanol. Percentage

inhibition of DPPH and ABTS was calculated as:

100)1((%) control

sample

Abs

AbsInhibition (2)

where Abscontrol is the absorbance of DPPH or ABTS radical with methanol; Abssample is the

absorbance of DPPH or ABTS radical with sample extract or standard. The total antioxidant

capacity was quantified from a trolox standard curve and expressed as trolox equivalent per gram

of dry weight sample (µg TE/g DW) ± SD for duplicate samples.

2.9. Determination of thermal kinetics parameters

The order of the reaction in the thermal degradation of anthocyanins was predicted by

differential method using the following model:

nCkdt

dC)( (3)

where k is the rate constant, n is the reaction order, C is the concentration of the total

anthocyanins and t is the reaction time.

Order of reaction was determined by graphical analysis, where exponent n in eq. 3 was set to

zero, half, one and two to compare the coefficients of determination among zero-, half-, 1st and

2nd

order reactions, respectively. The integrated form of zero-, half-, 1st and 2

nd order kinetic

models is listed below:

Zero-order: ktCCt 0 (4)

Half-order: ktCCt 02 (5)

Page 240: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

221

First-order: ktC

C

o

t ln (6)

Second-order: ktCCt

0

11 (7)

Using the experimental anthocyanins data, the coefficient of determination was observed to be

minimum for n = 1, predicting a 1st order reaction. According to the activated complex theory for

chemical reaction rates, for 1st order, Arrhenius equation relates the reaction rate constants to the

absolute temperature (34):

RT

EAk a lnln (8)

where Ea is the activation energy (kJmol-1

), A is pre-exponential factor/frequency factor (s-1

), T is

the absolute temperature (K) and R is the universal constant (8.135 Jmol-1

K-1

). The reaction rate

constant k and the activation energy, Ea were determined graphically from a plot of ln (C/Co) vs

time and ln k vs 1/T, respectively.

Thermal death time method (D-z model) was used to estimate the decimal reduction time (D-

value) i.e. heating time required to reduce the anthocyanins concentration by 90% and z-value

i.e. temperature change necessary to alter the thermal death time by one log cycle (34) with the

following relationships:

kD

10ln (9)

zTTDD refref /)()/log( (10)

where Dref is the D-value at temperature Tref. The half-lives (t1/2) of the anthocyanins were

calculated as:

k

t2ln

21 (11)

Page 241: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

222

The enthalpy of activation (∆H) and entropy of activation (∆S) was estimated using the

Eyring-Polanyi model based on the transition state theory (35):

RT

STH

b eTh

kk

. (12)

Where T is the absolute temperature (K), kb is the Boltzman constant (1.381 x 10-23 J/K), h is

the Planck constant (6.626x10-34 Js) and R is the gas constant (8.31 J/mol K).

3. RESULTS

3.1. Purification of anthocyanins

Aqueous anthocyanin extracts analyzed using HPLC showed the presence of chlorogenic,

caffeic, p-coumaric, ferulic and protocatechuic acids along with other unknown phenolic

compounds at 280 nm (Figure 6.3). After separation of salts, sugars and colorless phenolics

from the aqueous anthocyanin extracts, purified anthocyanins as detected at 280 nm did not show

any peak coincident with phenolic acid standards other than ferulic acid. When the purified

anthocyanins were detected at 520 nm (Figure 6.4), it was observed that the peak for the major

anthocyanin was eluted at the same time as the ferulic acid peak. Absence of individual ferulic

acid mass (m/z: 176.2) in MALDI data (Figure 6.5A) showed no contamination of phenolic

acids in the purified anthocyanin preparations. Comparison of spectra of the purified

anthocyanins from ‗Purple Majesty‘ potatoes with the pure anthocyanin (36) revealed acylation

of a phenolic acid to the major anthocyanin. The MALDI data (Figure 6.5A) showed the

presence of petunidin glucosides (m/z: 479), petunidin glucosides acylated with ferulic acid

(m/z: 963) and also petunidin glucosides acylated with coumaric acid (m/z: 933) in the purified

anthocyanins (10, 37-39). The latter mass (m/z: 933) agreed with the observations of Stushnoff et

Page 242: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

223

5 10 15 20 25 30 35

Minutes

0

25

50

75

100

125

Ab

sorb

an

ce (

mA

U)

Purified Anthocyanins

(acidified methanol portion)

Total Polyphenols

(in crude extract)

Phenolic acids

(ethyl acetate portion)

Wavelength = 280 nm

2

3

4

5

6

Figure 6. 3. HPLC-DAD profile of total polyphenols (in crude extract), phenolic acids (in ethyl

acetate portion) and purified anthocyanins (in HCl/Methanol portion) from ‗Purple Majesty‘

potatoes purified using a Sep-Pak C18 column and as detected at 280 nm in HPLC.

Peak 2: protocatechuic acid; peak 3: chlorogenic acid; peak 4: caffeic acid; peak 5: p-coumaric

acid, and peak 6: ferulic acid. anthocyanin eluted at the same time as the ferulic acid peak.

al. (10) on the presence of petunidin-3-rutinoside-5-glucoside acylated with coumaric acid as the

major anthocyanin in ‗Purple Majesty‘ potatoes. The investigator also reported having petunidin-

3-rutinoside-5-glucoside acylated with ferulic acid in ‗Purple Majesty‘ potatoes along with

delphidin, malvidin and peonidin aglycones with single glucosides. Although 1 – 2 smaller and

broader peaks were also observed in the HPLC chromatogram (Figure 6.4), no other individual

anthocyanin was characterized.

Page 243: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

224

5 10 15 20 25 30 35

Minutes

0

25

50

75

Ab

sorb

an

ce (

mA

U)

280 nm

520 nm

Figure 6. 4 HPLC-DAD profile of control unheated purified anthocyanins from ‗Purple

Majesty‘ potatoes as detected at 280 and 520 nm.

99 300 501 702 903 1104

Mass (m/z)

5.0E+4

0

10

20

30

40

50

60

70

80

90

100

933.2020

479.1046 852.4272

317.0674 868.4166

850.4072293.0998

866.3948172.0401 379.0930

190.049

854.4277481.1025

963.1570766.6755318.6223147.6709

464.1113

3

4

2

1

A

% I

nte

nsit

y

99 300 501 702 903 1104

Mass (m/z)

5.0E+4

0

10

20

30

40

50

60

70

80

90

100

933.2020

479.1046 852.4272

317.0674 868.4166

850.4072293.0998

866.3948172.0401 379.0930

190.049

854.4277481.1025

963.1570766.6755318.6223147.6709

464.1113

3

4

2

1

A

99 300 501 702 903 1104

Mass (m/z)

5.0E+4

0

10

20

30

40

50

60

70

80

90

100

933.2020

479.1046 852.4272

317.0674 868.4166

850.4072293.0998

866.3948172.0401 379.0930

190.049

854.4277481.1025

963.1570766.6755318.6223147.6709

464.1113

3

4

2

1

A

% I

nte

nsit

y

Page 244: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

225

99 300 501 702 903 1104

Mass (m/z)

7.1E+4

0

10

20

30

40

50

60

70

80

90

100

852.2131

317.0797 868.1868

932.8741379.0927

172.0307479.0848

190.0452

770.9835164.0598

147.0283

303.1453 772.9853 948.8604481.0866130.1362

% I

nte

nsit

y

C

99 300 501 702 903 1104

Mass (m/z)

7.1E+4

0

10

20

30

40

50

60

70

80

90

100

852.2131

317.0797 868.1868

932.8741379.0927

172.0307479.0848

190.0452

770.9835164.0598

147.0283

303.1453 772.9853 948.8604481.0866130.1362

% I

nte

nsit

y

C

Figure 6. 5 MALDI mass spectra of the pigments from purified anthocyanins of ‗Purple

Majesty‘ potatoes; (A) spectra of control purified anthocyanins samples; peak 1: petunidin; peak

2: petunidin monoglucoside; peak 3: petunidin-3-rutinoside-5glucoside acyalated with coumaric

acid; peak 4: petunidin-3-rutinoside-5glucoside acyalated with ferulic acid; (B) spectra of heat

treated samples at 100 °C for 30 min; (C) spectra of heat treated samples at 100 °C for 60 min.

Note: Scales of spectra A, B and C are different.

99 300 501 702 903 1104

Mass (m/z)

2.1E+4

0

10

20

30

40

50

60

70

80

90

100

852.5448

868.5381

317.0477

171.9863332.0712 786.2531

% I

nte

nsit

yB

99 300 501 702 903 1104

Mass (m/z)

2.1E+4

0

10

20

30

40

50

60

70

80

90

100

852.5448

868.5381

317.0477

171.9863332.0712 786.2531

% I

nte

nsit

yB

Page 245: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

226

Table 6. 2. Concentrations of total anthocyanins from purple potato extract upon heating at 100 – 150 °C for 0 – 60 min.

(mean ± SD, n = 2; time = 0 min at come-up-time)

Total Anthocyanins (mg mal-3-glu/g dry weight sample)

Heating time (min)

Heating temperature (° C)

100 110 120 130 140 150

Control 0.487 ± 0.030a 0.487 ± 0.030a 0.487 ± 0.030a 0.487 ± 0.030a 0.487 ± 0.030a 0.487 ± 0.030a

0 0.467 ± 0.009a 0.420 ± 0.005b 0.436 ± 0.002b 0.402 ± 0.018b 0.368 ± 0.017b 0.364 ± 0.007b

3 - - - - - 0.098 ± 0.012c

5 0.410 ± 0.013b 0.344 ± 0.017c 0.285 ± 0.020c 0.202 ± 0.004c 0.084 ± 0.008c 0.044 ± 0.005cd

10 - - - - 0.030 ± 0.005cd 0.011 ± 0.008e

15 0.339 ± 0.007c 0.216 ± 0.009d 0.138 ± 0.017d 0.059 ± 0.005d 0.007 ± 0.002e -

20 - - - - - 0.001 ± 0.000f

30 0.186 ± 0.005d 0.082 ± 0.014e 0.042 ± 0.006e 0.010 ± 0.005e 0.001 ± 0.001f -

45 - - - 0.003 ± 0.004f - -

60 0.101 ± 0.003e 0.034 ± 0.007f 0.006 ± 0.003f - - -

Significant differences within the values in the same column are indicated by different superscript letters (p < 0.05, Tukey‘s

pairwise comparison test)

Page 246: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

227

3.2. Degradation of anthocyanins

Total anthocyanin contents (TA) in the purified anthocyanin preparation determined before

and after heat treatments are summarized in Table 6.2. The TA of the control unheated extract

was 0.487 ± 0.03 mg mav-3-glu/g of DW sample. Thermal degradation of anthocyanins with

heating at 100 °C is clearly reflected by the reduction in the peaks detected at 520 nm (Figure

6.6A). A similar trend was observed when using 280 nm for detection with a major peak with

elution time of 20.643 min (Figure 6.6B).

The order of thermal degradation was estimated by examining the coefficient of

determination (r2) from plots of TA versus treatment time over the temperature range of

Table 6. 3. Estimation of the order of anthocyanins degradation by examining r2 from plot of

zero-, half-, first and second order reactions. (n = 2) Temperature (°C) Zero order Half-order First-order Second-order

100 0.9316 0.9644 0.9849 0.9792

110 0.8451 0.9207 0.9759 0.8428

120 0.7404 0.8958 0.9978 0.8632

130 0.7041 0.867 0.9918 0.8428

140 0.4993 0.704 0.9655 0.8511

150 0.4886 0.7194 0.9828 0.8428

Mean 0.7015 0.8452 0.9831 0.8703

Table 6. 4. Parameters for first-order kinetics and transition state equations for degradation of

anthocyanins from ‗Purple Majesty‘ potatoes after heat treatment over the temperature range of

100 – 150 °C. (n = 2) Temp

(°C)

k (min-1

)

x 103

r2

t1/2

(min)

D-value

(min)

Z-value

(°C)

Ea

(kJ/mol)

∆H

(kJ/mol)

∆S

(J/mol K)

100 26.2 0.9849 26.456 87.885 47.85 72.89 59.97 – 116.46

110 43.2 0.9759 16.045 53.301

120 71.4 0.9978 9.708 32.249

130 110 0.9918 6.301 20.933

140 194.1 0.9655 3.571 11.863

150 285.5 0.9828 2.428 8.065

Page 247: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

228

5 10 15 20 25 30 35

Minutes

0

10

20

30

40A

bso

rba

nce

(m

AU

) 1 min

Wavelength = 520 nm

15 min

30 min

60 min

A

5 10 15 20 25 30 35

Minutes

0

25

50

75

Ab

sorb

an

ce (

mA

U)

1min

15 min

30 min

60 min

Channel 1 = 280 nm

23

5

6

x

y

z

B

Figure 6. 6. Chromatogram of thermal degradation compounds from purified anthocyanins

heated at 100 °C for 1 – 60 min. (A) Peaks detected at 520 nm; (B) Peaks detected at 280 nm.

Peaks 2, 3, 5 and 6 were protocatechuic, caffeic, p-coumaric and ferulic acids, respectively. x, y

and z were unknown peaks.

Page 248: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

229

100 – 150 °C (Table 6.3). Based on the mean r2, thermal degradation of anthocyanins appeared

to follow 1st order kinetics (r

2 = 0.98). The temperature dependent rate constants, k-values, from

1st order over 100 – 150 °C as calculated from a plot of ln (C/Co) versus treatment time (Figure

6.7) were 0.0262 – 0.2855 min-1

. As expected, the reaction rate increased almost 10 times and

the D-values decreased 11 times (8.06 – 87.89 min) as the heating temperatures increased from

100 to 150 °C (Table 6.4). The activation energy, Ea and z-value of the degradation reaction

calculated from the Arrhenius plot (Figure 6.8) were 72.89 KJ/mol and 47.84 °C, respectively.

The activation enthalpy (∆H) and entropy (∆S) estimated from transition state theory were 59.97

kJ/mol and –116.46 J/mol K, respectively (Table 6.4).

3.3. Antioxidant activity of the degradation compounds

DPPH and ABTS antioxidant assays were performed to evaluate the total antioxidant

capacity of the anthocyanin preparation before and after heating. The TAC of the unheated

sample using DPPH and ABTS assays were 1237 ± 14 and 1546 ± 5 µg TE/g DW sample,

respectively. For each heating temperature, most of the TAC of heated samples either increased

or remained unchanged (p < 0.05) compared to unheated control samples. However, ABTS

antioxidant assay of degradation compounds for samples treated at 110 and 140 °C showed some

decrease in their TAC values. The TAC of the degradation compounds from anthocyanins ranged

from 1243 ± 97 to 1860 ± 21 µg TE/g DW sample using the DPPH assay (Table 6.5) and 1302 ±

82 to 1715 ± 21 µg TE/g DW sample using the ABTS assay (Table 6.6). Thermally induced

changes for the production of compounds from anthocyanins degradation did not follow a

particular trend. The ratio of TAC and TA in the control unheated purified anthocyanins was

2.54 and 3.17 using DPPH and ABTS assay, respectively, and increased to a maximum of 1316

(using DPPH assay) and 1581 (using ABTS assay) after heating at 150 °C (Table 6.7).

Page 249: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

230

Treatment time (min)

0 10 20 30 40 50 60 70

ln (

C/C

o)

-7

-6

-5

-4

-3

-2

-1

0100 °C

110 °C

120 °C

130 °C

140 °C

150 °C

1st order

model

Figure 6. 7. First order plot for the degradation of anthocyanins during heating over a

temperature range of 100 -150 °C. Data were means of duplicate samples.

4. DISCUSSION

Protocatechuic, chlorogenic, caffeic and p-coumaric acids (peaks 2, 3, 4 & 5 in Figure 6.3)

in the crude anthocyanins extract were not observed in the purified anthocyanins when examined

at 280 nm. During purification of anthocyanins, washing of the Sep-Pak C18 column with ethyl

acetate removed the colorless non-anthocyanin phenolics (peaks 2, 3, 4 and 5 in Figure 6.3).

Prior to the removal of colorless phenolics, salts and sugars were also removed from the crude

extract with acidified water. A HPLC chromatogram of the purified anthocyanins showed one

large peak eluting at the same time as ferulic acid and smaller peaks were detected for other

anthocyanins at 520 nm. Purified anthocyanins from elderberries, strawberries and black carrots

have shown similar chromatograms when detected at 280 and 520 nm (14, 19). From the HPLC

Page 250: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

231

1000/T in absolute temperature (K)

2.30 2.35 2.40 2.45 2.50 2.55 2.60 2.65 2.70

ln k

-4

-3

-2

-1

0

Figure 6. 8. Plot of ln (k) versus (1/T) for anthocyanin degradation during heating over the

temperature range of 100 -150 °C.

and MS information, the presence of petunidin glucosides acylated with a phenolic acid in

purified anthocyanins was indicated. Stushnoff et al. (10) reported the presence of petunidin-3-

rutinoside-5-glucoside acylated with coumaric and ferulic acid in addition to other types of

anthocyanins in ‗Purple Majesty‘ potatoes.

Degradation of anthocyanins in the extract is associated with reduction in its color. A similar

observation was noted while heating purple potatoes at high temperature (40). In this study,

anthocyanins in the extracts were reduced to negligible levels after 15 min of heating at 140 °C

and 10 min of heating at 150 °C. Yue & Xu (23) observed undetectable levels of anthocyanins in

the bilberry extracts after 10 min of heating at 150 °C. Reduction to negligible amount of

anthocyanins in heated samples could be due to chalcone formation from anthocyanins in the

Page 251: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

232

process of thermal exposure or loss of glycosyl moieties (41). The breakdown of anthocyanin

compounds to other colorless small molecular weight compounds cause the sample to lose its

original color (22). Over the range of 100 – 150 °C, degradation kinetics followed a 1st order

reaction, which is in agreement with previous reports (22, 24, 42-44). The reaction rate constants

and half-lives over the range of 100 – 150 °C confirmed the influence of temperature on

anthocyanins and agreed with a previous report by Yue et al. (23) on dry heating of bilberry

extract. However, the k-value for purified anthocyanins in the present study was 10-fold higher

than the reported values for crude anthocyanins in a blackcurrant juice (pH 3.4) model at 100 °C

(24) and purple-fleshed potatoes (pH 3.0) at 98 °C (22). This shows that purified anthocyanins

degrade at a faster rate compared to anthocyanins in crude or unpurified extracts obtained

directly from food material, which could be due to inter- and intramolecular co-pigmentation

reactions in the pigments and cofactors such as colorless non-anthocyanin phenolic compounds

in the extract (11, 12). The values of half-lives over 100 – 125 °C in the present study at pH 5.95

± 0.05 (Table 6.4) were comparable with the half-lives of bilberry extracts during dry heating

(23) but lower than reported values at 100 °C i.e. 2.18 h in a model blackcurrant juice (24) and

98 °C (1.6, 5.6, 2 and 4.9 h) for purple-flesh potatoes, red-flesh potatoes, grapes and purple

carrots, respectively (22). Sadilova et al. (14, 19) reported that the half-lives of purified

anthocyanins at 95 °C were 1.95, 1.96 and 2.81 h at pH 3.5 and 3.2, 1.9 and 4.1 h at pH 1.0 for

strawberries, elderberries and black carrots, respectively. This indicates that purified

anthocyanins from ‗Purple Majesty‘ potatoes (pH 6.0) are less heat stable than those of

strawberries, elderberries and black carrots. One possible explanation is the difference in the

chemical structure of anthocyanin (45), intramolecular stacking of acylated anthocyanins (46),

types of acylation of anthocyanins (14) and types of sugar moieties (47). For example, cyanidin-

Page 252: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

233

3-galactoside-xyloside-glucoside-sinapic acid, cyanidin-3-galactoside-xyloside-glucoside-ferulic

acid, and cyanidin-3-galactoside-xyloside-glucoside-coumaric acid in black carrots have half-

lives of 2.94, 3.43 and 3.1 h, respectively at pH 3.5 and 2.57, 2.39 and 2.16 h at pH 1.0 when

heated at 95 °C (14, 19). The same investigators also reported that different anthocyanins such as

pelargonidin-3-glucoside in strawberries and cyanindin-3-glucoside in elderberries have half-

lives of 2.12 and 1.82 at pH 3.5 and 2.12 and 1.95 at pH 1.0 upon heating at 95 °C. In ‗Purple

Majesty‘ potatoes, anthocyanins such as petunidin/malvidin/delphidin/peonidin-3-rutinoside-5-

glucoside are acylated with coumaric or ferulic acids (10), which might have different stability

when exposed to heat. Matsufuji et al., (48) reported anthocynins in red radish extract acylated

with p-coumaric acid or ferulic acid had a little more stability than anthocyanins acylated with

caffeic acids. The z-value of Purple Majesty potato anthocyanins was higher than the reported

values for all blue and CO94165-3P/P potato varieties (22) but comparable to roselle

anthocyanin extract (35).

Activation energy for the degradation of purified anthocyanins (72.89 KJ/mol ) in the present

study was in agreement with reported values for purple-flesh potatoes (72.49 kJ/mol) over 25 –

98 °C (22), blood orange juice and concentrate (73.2 – 89.5 kJ/mol) over 5 – 90 °C (44), sour

cherry concentrate (73.06 kJ/mol) over –18 to 80 °C (43), and blueberry extract (60 – 80 kJ/mol)

over 80 – 150 °C (23). Higher activation energy implies that a smaller temperature change could

degrade a compound more rapidly. The enthalpy of activation in the degradation of purified

anthocyanins from ‗Purple Majesty‘ potatoes was higher than blackberry (34 kJ/mol) and roselle

extracts (except Thai variety) (44 – 48 kJ/mol) but less than blood orange (63 kJ/mol) over 30 –

100 °C (35). This indicates that the degradation rate of purified anthocyanins from ‗Purple

Majesty‘ potatoes was less affected by temperature, even over a temperature range of 100 – 150

Page 253: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

234

Table 6. 5. Antioxidant Activity (µg trolox eq./g DW sample) using DPPH radical scavenging

assay of purified anthocyanins from ‗Purple Majesty‘ potatoes upon heating at 100 – 150 °C for

0 – 60 min. (mean ± SD, n=2; come-up-time = 0 min)

Total Antioxidant Activity (µg trolox eq./g dry weight sample)

Heating

time

(min)

Heating temperature (° C)

100 110 120 130 140 150

Control 1237 ± 14e 1237 ± 14b 1237 ± 14b 1237 ± 14c 1237 ± 14c 1237 ± 14b

0 1352 ±21d 1404 ± 83a 1336 ± 7a 1639 ± 56a 1500 ± 34a 1326 ± 34a

3 - - - - - 1364 ± 89a

5 1511 ± 21b 1355 ± 41b 1341 ± 28a 1525 ± 7b 1500 ± 7a 1355 ± 21a

10 - - - - 1476 ± 41a 1277 ± 117a

15 1860 ±21a 1574 ± 117a 1356 ± 49a 1719 ± 28a 1340 ± 27b -

20 - - - - - 1316 ± 21a

30 1431 ± 7c 1277 ± 124b 1252 ± 14b 1525 ± 63b 1359 ± 14b -

45 - - - 1510 ± 98b - -

60 1431 ± 21c 1389 ± 255a 1243 ± 97a - - -

Significant differences within the values in the same column are indicated by different

superscript letters (p < 0.05, Tukey‘s pairwise comparison test)

°C than those from blood orange juice. Enthalpy of activation measures the energy barrier, which

must be overcome by reacting molecules and is related to the strength of bonds that are broken

and made in the formation of the transition state from the reactants (49). Entropy of activation

relates to the number of molecules with the appropriate energy which actually react and the

values provide an insight into the roles of the reactants in anthocyanins degradation by including

steric and orientation requirements (49). The estimated entropy of activation in ‗Purple Majesty‘

potatoes (–116.46 J/mol K ) was less negative than that of blood orange juice (–149 J/mol.K) ,

blackberry juice (–233 J/mol.K) and roselle extracts (–165 to –205 J/mol.K) (35). The more

negative entropy of activation indicates a smaller number of species in the transition state (49).

Page 254: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

235

Antioxidant capacity and color of control unheated purified anthocyanins extracts is due to

high resonance of the fully conjugated 10-electron A-C ring-system in the structure, with some

contribution by the B-ring as well. The resonance of the structure leads to low reactivity and the

groups attached to the structure such as hydroxyl, methoxy, and glycosyl add stability (3, 17). After

thermal treatment of the purified extracts, degradation of original acylated anthocyanins

occurred. The MS data shows spectra of pigments in the samples heat treated at 100 °C for 30

and 60 min and the formation of new compounds at different heating times (Figure 6.5B &

6.5C). Even if the mechanism is not clear, possible explanations include chalcone formation,

deglycosylation and formation of compounds such as coumarin derivatives (50), benzoic

derivatives (51) and trihydroxybenzaldehyde (14). Formation of five different types of

Table 6. 6. Antioxidant Activity (µg trolox eq./g DW sample) using ABTS radical scavenging

assay of purified anthocyanins from ‗Purple Majesty‘ potato upon heating at 100 – 150 °C for 0

– 60 min. (mean ± SD, n=2; come-up-time = 0 min)

Total Antioxidant Activity (µg trolox eq./g dry weight sample)

Heating

time

(min)

Heating temperature (° C)

100 110 120 130 140 150

Control 1546 ± 5c 1546 ± 5a 1546 ± 5a 1546 ± 5b 1546 ± 5a 1546 ± 5b

0 1653 ± 22b 1491 ± 32a 1538 ± 5ab 1663 ± 22a 1539 ± 32a 1715 ± 21a

3 - - - - - 1622 ± 55b

5 1710 ± 19a 1433 ± 34b 1501 ± 21b 1573 ± 39a 1421 ± 14b 1619 ± 64b

10 - - - - 1400 ± 20b 1617 ± 12b

15 1687 ±38a 1433 ± 3b 1488 ± 8b 1576 ± 28a 1386 ± 6c -

20 - - - - - 1581 ± 40b

30 1649 ± 22b 1370 ±

119b 1482 ± 11b 1537 ± 45ab 1365 ± 23c -

45 - - - 1460 ± 64c - -

60 1559 ± 41c 1302 ± 82b 1469 ± 13c - - -

Significant differences within the values in the same column are indicated by different

superscript letters (p < 0.05, Tukey‘s pairwise comparison test)

Page 255: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

236

anthocyanidins from heated bilberry extract after cleavage of conjugated sugars from

anthocyanins has been reported (23). The retention or increase in the TAC of heated anthocyanin

preparation compared to unheated purified anthocyanins is emphasized by noting the increasing

ratio of TAC and TA in the samples (Table 6.7) and formation of new compounds as observed in

increasing peak areas detected at 280 nm in chromatograms from the degradation of

anthocyanins when heated at 100 °C for 0 – 60 min (Figure 6.6). The ratio of TAC to TA using

DPPH assay varied from 2.54 – 1359.46, whereas using ABTS assay the ratio was 3.81 – 1581

over the temperature range of 100 – 150 °C at 0 – 60 min. This shows that the anthocyanin

degradation compounds exhibit higher antioxidant activity than unheated anthocyanin, which

agrees with previous reports (14, 23, 48). It could be assumed that acylated anthocyanins are

cleaved into their corresponding acyl-glycosides, then into intermediate chalcones and finally

into colorless phenolics such as phenolic acids and aldehydes that contribute to the TAC in the

heated samples. Sadilova et al., (14) observed retentions of 74.4, 86.5 and 84.6 % antioxidant

activity in the anthocyanin degradation compounds after heating 4 h at 95 °C in strawberry,

black carrot and elderberry anthocyanins isolates, respectively, because of formation of

compounds such as protocatechuic acid, phologlucinaldehyde, and 4-hydroxybenzoic acid.

Sreeram et al. (51) reported that the antioxidant activity of degradation compounds of cyanidin

glycosides from tart cherries were comparable to commercial antioxidants such as butylated

hydroxytoulene (BHT).

5. CONCLUSIONS

The present study evaluated the thermal degradation of purified anthocyanins from ‗Purple

Majesty‘ potatoes after removal of salts, sugars and colorless non-anthocyanin phenolics. The

thermal degradation of purified anthocyanins followed first order kinetics in an Arrhenius type

Page 256: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

237

Table 6. 7. Progression of degradation compounds upon thermal exposure as measured by the ratio of total antioxidant capacity

(TAC) and total anthocyanins (TA) in purified anthocyanins samples over 100 – 150 °C. (A) TAC measured using DPPH assay; (B)

TAC measure using ABTS assay.

Ratio of Total Antioxidant Capacity versus Total Anthocyanins

Heating

time

(min)

Heating temperature (° C)

100 110 120 130 140 150

DPPH ABTS DPPH ABTS DPPH ABTS DPPH ABTS DPPH ABTS DPPH ABTS

Control 2.54 3.17 2.54 3.17 2.54 3.17 2.54 3.17 2.54 3.17 2.54 3.17

0 2.89 3.54 3.34 3.55 3.06 3.53 4.08 4.14 4.08 4.18 3.64 4.71

3 - - - - - - - - - - 13.92 16.55

5 3.69 4.17 3.94 4.17 4.71 5.27 7.55 7.79 17.86 16.92 30.79 36.80

10 - - - - - - - - 49.19 46.67 116.10 147

15 5.49 4.98 7.29 6.63 9.82 10.78 29.13 26.71 191.44 198.57 - -

20 - - - - - - - - - - 1315.84 1581

30 7.68 8.85 15.57 16.71 29.82 35.29 152.51 153.70 1359.46 1365 - -

45 - - - - - - 503.40 486.67 - - - -

60 14.14 15.40 40.85 38.29 207.09 244.83 - - - - - -

Page 257: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

238

relationship. The total antioxidant capacity of purified anthocyanins after heat treatment was

either increased or retained compared to the unheated samples because of compensation in

antioxidant activity of the degradation compounds. Characterization of degradation compounds

from heat treatments based on their structure and functionality might be needed to understand the

mechanism responsible for the measured antioxidant activity.

Page 258: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

239

REFERENCES

1. Wagner, G. J., Cellular and Subcellular Localization in Plant Metabolism. In Recent

Advances in Phytochemistry, Creasy, L. L.; Hrazdina, G., Eds. Plenum Press: New York,

NY, 1982; Vol. 16, pp 1 - 45.

2. Downham, A.; Collins, P., Colouring our foods in the last and next millennium.

International Journal of Food Science and Technology 2000, 35, (1), 5-22.

3. Rice-Evans, C. A.; Miller, N. J.; Paganga, G., Structure-antioxidant activity relationships

of flavonoids and phenolic acids. Free Radical Biology and Medicine 1996, 20, (7), 933-

956.

4. Stintzing, F. C.; Stintzing, A. S.; Carle, R.; Frei, B.; Wrolstad, R. E., Color and

antioxidant properties of cyanidin-based anthocyanin pigments. Journal of Agricultural

and Food Chemistry 2002, 50, (21), 6172-6181.

5. Kahkonen, M. P.; Hopia, A. I.; Vuorela, H. J.; Rauha, J. P.; Pihlaja, K.; Kujala, T. S.;

Heinonen, M., Antioxidant Activity of Plant Extracts Containing Phenolic Compounds. J.

Agric. Food Chem. 1999, 47, (10), 3954-3962.

6. Hou, D. X., Potential mechanisms of cancer chemoprevention by anthocyanins. Current

Molecular Medicine 2003, 3, (2), 149-159.

7. Han, K. H.; Hashimoto, N.; Shimada, K.; Sekikawa, M.; Noda, T.; Yamauchi, H.;

Hashimoto, M.; Chiji, H.; Topping, D. L.; Fukushima, M., Hepatoprotective effects of

purple potato extract against D-galactosamine-induced liver injury in rats. Bioscience

Biotechnology and Biochemistry 2006, 70, (6), 1432-1437.

Page 259: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

240

8. Mishra, D. K.; Dolan, K. D.; Yang, L., Confidence Intervals for Modeling Anthocyanin

Retention in Grape Pomace during Nonisothermal Heating. Journal of Food Science

2008, 73, (1), E9-E15.

9. Lachman, J.; Hamouz, K.; Orsak, M., Red and purple potatoes - A significant antioxidant

source in human nutrition. Chemicke Listy 2005, 99, (7), 474-482.

10. Stushnoff, C.; Holm, D.; Thompson, M. D.; Jiang, W.; Thompson, H. J.; Joyce, N. I.;

Wilson, P., Antioxidant properties of cultivars and selections from the Colorado potato

breeding program. American Journal of Potato Research 2008, 85, (4), 267-276.

11. Malien-Aubert, C.; Dangles, O.; Amiot, M. J., Color stability of commercial

anthocyanin-based extracts in relation to the phenolic composition. Protective effects by

intra and intermolecular copigmentation. Journal of Agricultural and Food Chemistry

2001, 49, (1), 170-176.

12. Del Pozo-Insfran, D.; Brenes, C. H.; Talcottt, S. T., Phytochemical composition and

pigment stability of acai (Euterpe oleracea Mart.). Journal of Agricultural and Food

Chemistry 2004, 52, (6), 1539-1545.

13. Mazza, G.; Brouillard, R., The Mechanism of Copigmentation of Anthocyanins in

Aqueous-Solutions. Phytochemistry 1990, 29, (4), 1097-1102.

14. Sadilova, E.; Carle, R.; Stintzing, F. C., Thermal degradation of anthocyanins and its

impact on color and in vitro antioxidant capacity. Molecular Nutrition & Food Research

2007, 51, (12), 1461-1471.

15. Giusti, M. M.; Wrolstad, R. E., Acylated anthocyanins from edible sources and their

applications in food systems. Biochemical Engineering Journal 2003, 14, (3), 217-225.

Page 260: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

241

16. Redus, M.; Baker, D. C.; Dougall, D. K., Rate and equilibrium constants for the

dehydration and deprotonation reactions of some monoacylated and glycosylated

cyanidin derivatives. Journal of Agricultural and Food Chemistry 1999, 47, (8), 3449-

3454.

17. Delgado-Vargas, F.; Paredes-López, O., Anthocyanins and betalains. In Natural

Colorants for Food and Nutraceutical Uses, CRC Press Boca Raton, 2003; pp 167 - 219

18. Rossi, M.; Giussani, E.; Morelli, R.; Lo Scalzo, R.; Nani, R. C.; Torreggiani, D., Effect of

fruit blanching on phenolics and radical scavenging activity of highbush blueberry juice.

Food Research International 2003, 36, (9-10), 999-1005.

19. Sadilova, E.; Stintzing, F. C.; Carle, R., Thermal degradation of acylated and nonacylated

anthocyanins. Journal of Food Science 2006, 71, (8), C504-C512.

20. Garzon, G. A.; Wrolstad, R. E., The stability of pelargonidin-based anthocyanins at

varying water activity. Food Chemistry 2001, 75, (2), 185-196.

21. Kirca, A.; Ozkan, M.; Cemeroglu, B., Effects of temperature, solid content and pH on the

stability of black carrot anthocyanins. Food Chemistry 2007, 101, (1), 212-218.

22. Reyes, L. F.; Cisneros-Zevallos, L., Degradation kinetics and colour of anthocyanins in

aqueous extracts of purple- and red-flesh potatoes (Solanum tuberosum L.). Food

Chemistry 2007, 100, (3), 885-894.

23. Yue, X.; Xu, Z., Changes of Anthocyanins, Anthocyanidins, and Antioxidant Activity in

Bilberry Extract during Dry Heating. Journal of Food Science 2008, 73, (6), C494-C499.

24. Harbourne, N.; Jacquier, J. C.; Morgan, D. J.; Lyng, J. G., Determination of the

degradation kinetics of anthocyanins in a model juice system using isothermal and non-

isothermal methods. Food Chemistry 2008, 111, (1), 204-208.

Page 261: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

242

25. Wang, H.; Cao, G. H.; Prior, R. L., Oxygen radical absorbing capacity of anthocyanins.

Journal of Agricultural and Food Chemistry 1997, 45, (2), 304-309.

26. AACC, Approved methods of the American Association of Cereal Chemists (10th ed.).

American Association of Cereal Chemists: St. Paul, MN, USA, 2000; p Method 44-15A.

27. Giusti, M. M.; Wrolstad, R., Characterization and measurement of anthocyanin by UV-

visible spectroscopy In Current Protocols in Food Analytical Chemistry., John Wiley &

Sons ed.; Wrolstad, R. E., Ed. John Wiley & Sons, New York: 2001; Vol. F 1.2, pp pp

F1.2.1-F1.2.13.

28. Reyes, L. F.; Cisneros-Zevallos, L., Wounding stress increases the phenolic content and

antioxidant capacity of purple-flesh potatoes (Solanum tuberosum L.). Journal of

Agricultural and Food Chemistry 2003, 51, (18), 5296-5300.

29. Han, K. H.; Sekikawa, M.; Shimada, K.; Hashimoto, M.; Hashimoto, N.; Noda, T.;

Tanaka, H.; Fukushima, M., Anthocyanin-rich purple potato flake extract has antioxidant

capacity and improves antioxidant potential in rats. British Journal of Nutrition 2006, 96,

(6), 1125-1133.

30. Jansen, G.; Flamme, W., Coloured potatoes (Solanum tuberosum L.) - anthocyanin

content and tuber quality. Genetic Resources and Crop Evolution 2006, 53, (7), 1321-

1331.

31. Lewis, C. E.; Walker, J. R. L.; Lancaster, J. E.; Sutton, K. H., Determination of

anthocyanins, flavonoids and phenolic acids in potatoes. I: Coloured cultivars of Solanum

tuberosum L. Journal of the Science of Food and Agriculture 1998, 77, (1), 45-57.

Page 262: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

243

32. Brandwilliams, W.; Cuvelier, M. E.; Berset, C., Use of a Free-Radical Method to

Evaluate Antioxidant Activity. Food Science and Technology-Lebensmittel-Wissenschaft

& Technologie 1995, 28, (1), 25-30.

33. Re, R.; Pellegrini, N.; Proteggente, A.; Pannala, A.; Yang, M.; Rice-Evans, C.,

Antioxidant activity applying an improved ABTS radical cation decolorization assay.

Free Radical Biology and Medicine 1999, 26, (9-10), 1231-1237.

34. Holdsworth, S. D., Thermal Processing of Packaged Foods. Blackie Academic &

Professional: New York, 2000; p 70 - 93.

35. Cisse, M.; Vaillant, F.; Acosta, O.; Dhuique-Mayer, C.; Dornier, M., Thermal

Degradation Kinetics of Anthocyanins from Blood Orange, Blackberry, and Roselle

Using the Arrhenius, Eyring, and Ball Models. Journal of Agricultural and Food

Chemistry 2009, 57, (14), 6285-6291.

36. Hong, V.; Wrolstad, R. E., Use of Hplc Separation Photodiode Array Detection for

Characterization of Anthocyanins. Journal of Agricultural and Food Chemistry 1990, 38,

(3), 708-715.

37. Wang, J.; Kalt, W.; Sporns, P., Comparison between HPLC and MALDI-TOF MS

analysis of anthocyanins in highbush blueberries. Journal of Agricultural and Food

Chemistry 2000, 48, (8), 3330-3335.

38. Giusti, M. M.; Rodriguez-Saona, L. E.; Griffin, D.; Wrolstad, R. E., Electrospray and

tandem mass spectroscopy as tools for anthocyanin characterization. Journal of

Agricultural and Food Chemistry 1999, 47, (11), 4657-4664.

39. Jones, C. M.; Mes, P.; Myers, J. R., Characterization and inheritance of the Anthocyanin

fruit (Aft) tomato. Journal of Heredity 2003, 94, (6), 449-456.

Page 263: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

244

40. Nayak, B.; Berrios, J. J.; Powers, J. R.; Tang, J.; Ji, Y., Colored Potatoes (Solanum

Tuberosum L.) Dried for Antioxidant-Rich Value-Added Foods. Journal of Food

Processing and Preservation (in Press, DOI: 10.1111/j.1745-4549.2010.00502.x) 2010.

41. Adams, J. B., Thermal degradation of anthocyanins with particular reference to the 3-

glycosides of cyanidin. I. In acidified aqueous solution at 100 °C. Journal of the Science

of Food and Agriculture 1973, 24, (7), 747-762.

42. Wang, W. D.; Xu, S. Y., Degradation kinetics of anthocyanins in blackberry juice and

concentrate. Journal of Food Engineering 2007, 82, (3), 271-275.

43. Cemeroglu, B.; Velioglu, S.; Isik, S., Degradation Kinetics of Anthocyanins in Sour

Cherry Juice and Concentrate. Journal of Food Science 1994, 59, (6), 1216-1218.

44. Kirca, A.; Cemeroglu, B., Degradation kinetics of anthocyanins in blood orange juice and

concentrate. Food Chemistry 2003, 81, (4), 583-587.

45. Von Elbe, J. H.; Schwartz, S. J., Colorants. In Food Chemistry, Third ed.; Fennema, O.

R., Ed. Marcel Dekker, Inc: New York, NY, 1996; pp 651 - 722.

46. Dangles, O.; Saito, N.; Brouillard, R., Anthocyanin Intramolecular Copigment Effect.

Phytochemistry 1993, 34, (1), 119-124.

47. Attoe, E. L.; Von Elbe, J. H., Photochemial Degradation of Betanine and Selected

Anthocyanins. Journal of Food Science 1981, 46, (6), 1934-1937.

48. Matsufuji, H.; Kido, H.; Misawa, H.; Yaguchi, J.; Otsuki, T.; Chino, M.; Takeda, M.;

Yamagata, K., Stability to light, heat, and hydrogen peroxide at different pH values and

DPPH radical scavenging activity of acylated anthocyanins from red radish extract.

Journal of Agricultural and Food Chemistry 2007, 55, (9), 3692-3701.

Page 264: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

245

49. Ariahu, C. C.; Ogunsua, A. O., Thermal degradation kinetics of thiamine in periwinkle

based formulated low acidity foods. International Journal of Food Science & Technology

2000, 35, (3), 315-321.

50. Jackman, R. L.; Smith, J. L., Anthocyanins and betalains. In Natural food colorants,

Hendry, G. A. F.; Houghton, J. D., Eds. Blackie and Son Ltd.: London, 1992; pp 183 -

241.

51. Seeram, N. P.; Bourquin, L. D.; Nair, M. G., Degradation Products of Cyanidin

Glycosides from Tart Cherries and Their Bioactivities. Journal of Agricultural and Food

Chemistry 2001, 49, (10), 4924-4929.

Page 265: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

246

CHAPTER SEVEN

CONCLUSIONS AND RECOMMENDATIONS

1. CONCLUSIONS

In this study, the profile of the phenolic antioxidants and their bioactivity in colored potatoes,

value added products prepared from colored potatoes, and thermal kinetics of potato anthocyanin

was investigated. White, red, yellow and purple potato cultivars were investigated for total

phenolics, antioxidant activity and total anthocyanin content. The effects of steam blanching

following drying technologies (freeze-drying, drum-drying and refractance window-drying) on

the phenolic antioxidants were determined. Extruded products prepared from purple potato and

yellow pea flour using extrusion cooking technology were analyzed for their bioavailability as

well as chemical antioxidant activity by determining total phenolics and flavonoids including

existence in free and bound forms. A flow chart showing the detail recommended process for

producing puffed extrudates from purple potato (‗Purple Majesty‘ cv) and yellow pea flours is

given in Appendix 1. The thermal kinetics of purified anthocyanin from ‗Purple Majesty‘

potatoes was studied using specially designed thermal kinetic test cells in an oil bath over a

temperature range of 100 – 150 °C and the degradation products were analyzed for potential

antioxidant activity. From the above studies, it can be concluded that:

Purple potatoes (‗Purple Majesty‘ cv) contains significantly higher total phenolics,

antioxidant activity, and total anthocyanins than red (‗Red Cliff‘ cv), yellow (‗Yukon Gold‘ cv),

and white potato (‗Russet‘ cv) cultivars and can be used as a source for producing antioxidant-

rich value-added products.

Page 266: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

247

Blanching is an important step for processing purple potatoes to retain natural color

(anthocyanins) in dehydrated flakes.

Different drying technologies (drum-drying, freeze-drying and refractance window-

drying) used to prepare dehydrated purple potato flakes either retained or increased (P < 0.05)

the total phenolic content and antioxidant capacity.

Losses of 23 to 45 % in total anthocyanin content were observed in dehydrated potato

flakes processed under all drying methods.

Natural colored extruded puffed food products with an acceptable expansion ratio can be

produced from yellow pea and purple potato flours using extrusion cooking technology.

Although major degradation in the quantity of the total anthocyanin content was noticed,

some color in the final product survived high temperature extrusion.

The total antioxidant capacities (DPPH and ORAC) of the extruded products were

retained or increased because of retention of some natural antioxidant compounds and formation

of new compounds with potent antioxidant activity.

The total phenolics and flavonoids in most of the extruded products also increased 33 –

37% and 17 – 30%, respectively, after extrusion cooking.

Positive dose-response inhibitory effects on the HepG2 cells were observed in potato

flour, raw formulations, and extruded products.

Extrusion cooking improved the dose-response inhibitory effect of 50% PPF raw

formulation on the HepG2 cells, thus increasing the CAA value of extruded products.

The thermal degradation of purified anthocyanins followed first order kinetics in an

Arrhenius type relationship.

Page 267: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

248

The total antioxidant capacity of purified anthocyanins after heat treatment was either

increased or retained compared to the unheated samples because of compensation in antioxidant

activity of the degradation products.

2. RECOMMENDATIONS

The present study was focused on the phytochemicals present in colored potatoes and their

response to different processing conditions. The following future studies are recommended to

help reveal the effect of processing on phytochemicals in fruits and vegetables more completely:

Although positive results of antioxidant activities of processed products were observed,

the effect of storage on those phytochemicals was not studied.

Colored potatoes contain different types of antioxidant compounds at varying

concentrations, availability and activity depending on their intrinsic properties.

Fundamental research on the effect processing has on individual compounds is needed for

interpreting proper food and health relationships.

Value-added products from potatoes and peas further complicate the existence and

bioavailability of phytochemicals in free and conjugated states. Interactions between

natural and heat-induced antioxidants need to be investigated for possible synergistic

effects using advanced technologies such as Nuclear magnetic resonance (NMR) and

Matrix-assisted laser desorption/ionization (MALDI) is important to help understand the

overall antioxidant activity of the processed products.

In this study, bioavailability of antioxidant compounds was analyzed using HepG2 liver

cancer cells, which should be extended to other types of cells to develop a more complete

understanding of bioavailability.

Page 268: EFFECT OF THERMAL PROCESSING ON THE PHENOLIC

249

APPENDIX

Flow chart showing recommended processes for producing extrudates from purple potato

(‗Purple Majesty‘ cv) and yellow pea flours

Purple Majesty Potatoes Peeling Slicing

( 6 mm thick)

Blanching

(Steam; 8 min) Pureeing

Dehydration/Drying

(Drum drying; drum surface temp:135 – 138 °C;

speed: 1.13 rpm; gap: 0.3 mm)

Pin Milling

Potato flour

(225 µm) Pea flour

Mixing/Formulations

(8.7 – 9.3% moisture content, wet basis)

Extrusion Cooking

(feed moisture: 17% wb; screw speed:

300 rpm; die temp: 130 & 140 °C)

Extrudates

(Cooling at room temperature)

Cooling

(Ice water; 8 min)