phytochemistry and pharmacology of plants from the ginger family

261
Southern Cross University ePublications@SCU eses 2008 Phytochemistry and pharmacology of plants from the ginger family, Zingiberaceae Hans Wohlmuth Southern Cross University ePublications@SCU is an electronic repository administered by Southern Cross University Library. Its goal is to capture and preserve the intellectual output of Southern Cross University authors and researchers, and to increase visibility and impact through open access to researchers around the world. For further information please contact [email protected]. Publication details Wohlmuth, H 2008, 'Phytochemistry and pharmacology of plants from the ginger family, Zingiberaceae', PhD thesis, Southern Cross University, Lismore, NSW. Copyright H Wohlmuth 2008

Upload: others

Post on 04-Feb-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Phytochemistry and pharmacology of plants from the ginger family

Southern Cross UniversityePublications@SCU

Theses

2008

Phytochemistry and pharmacology of plants fromthe ginger family, ZingiberaceaeHans WohlmuthSouthern Cross University

ePublications@SCU is an electronic repository administered by Southern Cross University Library. Its goal is to capture and preserve the intellectualoutput of Southern Cross University authors and researchers, and to increase visibility and impact through open access to researchers around theworld. For further information please contact [email protected].

Publication detailsWohlmuth, H 2008, 'Phytochemistry and pharmacology of plants from the ginger family, Zingiberaceae', PhD thesis, Southern CrossUniversity, Lismore, NSW.Copyright H Wohlmuth 2008

Page 2: Phytochemistry and pharmacology of plants from the ginger family

PHYTOCHEMISTRY AND PHARMACOLOGY OF PLANTS FROM THE GINGER FAMILY, ZINGIBERACEAE

Hans Wohlmuth, BSc

Submitted for the Degree of Doctor of Philosophy

Department of Natural and Complementary Medicine Southern Cross University

Lismore, Australia

April 2008

Page 3: Phytochemistry and pharmacology of plants from the ginger family

i

Flowering Curcuma australasica, an endemic Australian Zingiberaceae with potential anti-inflammatory activity, cultivated at Southern Cross University, Lismore, Australia.

Page 4: Phytochemistry and pharmacology of plants from the ginger family

ii

Declaration

certify that the work presented in this thesis is, to the best of my knowledge and belief,

original, except as acknowledged in the text, and that the material has not been

submitted, either in whole or in part, for a degree at this or any other university.

I acknowledge that I have read and understood the University's rules, requirements,

procedures and policy relating to my higher degree research award and to my thesis. I certify

that I have complied with the rules, requirements, procedures and policy of the University (as

they may be from time to time).

Name: Hans Wohlmuth

Signature:…………………………………………………………………..

Date: ………………………………………………………………………..

I

Page 5: Phytochemistry and pharmacology of plants from the ginger family

iii

SYNOPSIS This thesis reports on a series of investigations into the phytochemistry and pharmacology of

plants belonging to the ginger family, Zingiberaceae (incl. Costaceae). The work falls into

two main parts. The first part examines the pungent compounds and essential oil in 17

clones of ginger (Zingiber officinale) with a view to identify one or more with unique

chemistry and consequent particular therapeutic (or flavouring) prospects. The second part

comprises the screening of 41 taxa for inhibition of PGE2 and other biological activities, with

the primary aim of identifying species with potential anti-inflammatory activity. This part

tested the hypothesis that the combination of ethnobotanical and taxonomic information is a

productive strategy to identify previously unrecognised plant species with therapeutic

potential.

Chapter 1 provides a general introduction to plants as medicines and the field of

ethnopharmacology. It also provides an overview of the process of inflammation, in

particular arachidonic acid metabolism.

The main literature review is presented in Chapter 2. It reviews the literature relating to the

chemistry and pharmacology of 15 genera in the Zingiberaceae, with the focus on species

included in the experimental work. The Zingiberaceae is rich in species used as traditional

medicines or spices, but extensive information about their chemistry and pharmacology is

available only for a few species, most notably ginger and turmeric (Curcuma longa). These

attributes make the family an ideal target for a screening project, since phylogenetically

related plant species usually display a significant degree of similarity in the kinds of

secondary metabolites they produce.

Chapter 3 describes the preliminary experimental work with ginger. This work aimed at

determining a suitable extraction solvent and method, guided by the activity of the extracts in

a cyclooxygenase-1 (COX-1) bioassay. It also established an HPLC method suitable for the

quantification of gingerols and shogaols in the extracts.

Seventeen ginger clones, including commercial cultivars and 12 experimental clones, were

analysed by HPLC for their content of pungent compounds. The result of this work is

reported in Chapter 4. Because ginger is a sterile cultigen, there is an increased likelihood

that chemically distinct and genetically stable clones may exist. This work identified one

cultivar that when compared with other clones contained a significantly higher level of

Page 6: Phytochemistry and pharmacology of plants from the ginger family

iv

pungent gingerols. The analysis included 12 tetraploid clones, but these did not display

elevated gingerol production compared with their diploid parent cultivar.

The essential oils obtained from the same 17 ginger clones by steam distillation were

analysed by GC-MS. These results are presented in Chapter 5. The oil from one particular

clone was distinctly different from the others; this was the same clone that had a very high

gingerol content. The essential oil of this clone differed from the others by having a lower

citral content and higher levels of sesquiterpene hydrocarbons. The unique chemistry of this

clone in terms of aroma, pungency and flavour should make it of interest to the flavour,

fragrance and pharmaceutical industries.

Chapter 6 presents the results of the screening of 41 taxa in an in vitro cell-based bioassay

for inhibition of PGE2 production. A number of the samples were also tested for antioxidant

activity in the oxygen radical absorbance capacity (ORAC) assay, for inhibition of nitric

oxide production and for modulation of natural killer cell activity. Known medicinal plants,

in particular ginger and turmeric, emerged as the most active in these assays. Included in the

work were seven native Australian species not previously investigated for pharmacological

activity. Two of these species showed good activity in the PGE2 assay and were selected for

further investigations.

Chapter 7 reports on the bioactivity-guided fractionation of these two native Australian

species, Curcuma australasica and Pleuranthodium racemigerum. Inhibition of PGE2 was

used as the primary bioassay in this process, but fractions with high activity in that assay

were also tested for cytotoxic properties. This work succeeded in isolating and structurally

characterising a novel curcuminoid compound with potent PGE2 inhibitory activity from P.

racemigerum as well as two known compounds from C. australasica.

The final chapter (Chapter 8) provides a short summary and concluding remarks, and

identifies areas for future research arising from the present work.

Page 7: Phytochemistry and pharmacology of plants from the ginger family

v

PUBLICATIONS ARISING FROM THE WORK IN THIS THESIS

Wohlmuth, H, Leach, DN, Smith, MK, Myers, SP (2005). Gingerol content of diploid and

tetraploid clones of ginger (Zingiber officinale Roscoe). Journal of Agricultural and

Food Chemistry 53: 5772-5778.

Wohlmuth, H, Smith, MK, Brooks, LO, Myers, SP, Leach, DN (2006). Essential oil

composition of diploid and tetraploid clones of ginger (Zingiber officinale Roscoe)

grown in Australia. Journal of Agricultural and Food Chemistry 54: 1414-1419.

Wohlmuth, H, Deseo, MA, Brushett, DJ, MacFarlane, G, Waterman, PG, Stevenson, LM,

Leach, DN (2007). Biological activity and novel cytotoxic curcuminoid from

Pleuranthodium racemigerum - an Australian Zingiberaceae. Planta Medica 73: 947

[Poster abstract].

Page 8: Phytochemistry and pharmacology of plants from the ginger family

vi

ACKNOWLEDGMENTS Many individuals have supported the work presented in this thesis in a variety of ways, and

without their assistance, guidance, help and support the project would not have been

possible. I am grateful to them all.

First and foremost, my warm thanks and appreciation go to my principal supervisor,

Associate Professor David Leach, for his guidance, support, patience and friendship. I also

thank my associate supervisor, Dr Lesley Stevenson, whose valuable input was most

appreciated, even with she moved across the Tasman. My thanks also go to Professor

Stephen Myers, who inspired and encouraged me to do a PhD, whose idea it was to work on

ginger, and who was my supervisor in the early days.

Most of the work presented in this thesis was carried out in the laboratories of the Centre for

Phytochemistry and Pharmacology at Southern Cross University, and in addition to my

supervisors I wish to thank staff and fellow students of the Centre for their invaluable

assistance during the project. No one deserves more thanks and credit than Don Brushett,

who tirelessly and patiently introduced me to the intricacies of analytical chemistry, and I am

also deeply grateful to Dr Myrna Deseo (NMR), Paul Connellan (flow cytometry), Ashley

Dowell (analytical chemistry) and Dion Thompson (pharmacology) for their generous

assistance. I also thank Linda Banbury, Karen Beattie, Bill Eickhoff, Yen Lin Ho, Dr Denise

Hunter, Dr Gloria Karagianis, Kate MacFarlane, Aaron Pollack, Kelly Shepherd, Emeritus

Professor Peter Waterman and Kelly Winter for their support.

The plant materials I worked on in this project came from a variety of sources. I am

indebted to Mike Smith, who so generously made available ginger cultivars and

experimental clones arising from his work at the Maroochy Research Station (Queensland

Department of Primary Industries and Fisheries). I am equally grateful to Dr Greg Leach of

the Northern Territory’s Department of Natural Resources, Environment and the Arts, who

facilitated access to the living collection of Zingiberaceae in the George Brown Darwin

Botanic Gardens. Similarly, through the generosity of Curator David Warmington, I was

able to access the extensive living collection of Zingiberaceae in the Flecker Botanic

Gardens in Cairns. I wish to thank Rigel Jensen of Malanda, whose impressive botanical and

local knowledge of the North Queensland Wet Tropics made the task of collecting native

Australian species in the wild a surprisingly easy one.

Page 9: Phytochemistry and pharmacology of plants from the ginger family

vii

Many of the plants were grown for several years at Southern Cross University, and I owe the

Curator of the Medicinal Plant Garden, Geoff Callan, and his staff much thanks for the

construction of purpose build beds and for general support and friendship. I also gratefully

acknowledge the support of the university’s Ground Staff who let me have a large number of

plants growing in pots in their nursery for several years.

I acknowledge and appreciate the support of my employer, Southern Cross University, in

particular that of Professor Iain Graham, Head of the School of Health and Human Sciences.

I also thank my colleagues in the Department of Natural and Complementary Medicine, in

particular Sue Evans and Holly Muggleston, who shouldered most of the burden during the

times I was away working on my research and thesis.

I also thank Dr Myrna Deseo and Associate Professor John Stevens who provided valuable

comments on the final draft.

Finally, thanks to Catherine for her love and support, and for never resenting me being busy

with my gingers.

Page 10: Phytochemistry and pharmacology of plants from the ginger family

viii

ABBREVIATIONS

5-HETE 5-hydroxyeicosatetraenoic acid

12-HETE 12-hydroxyeicosatetraenoic acid

12-HTT 12-hydroxy-5,8,10-heptadecatrienoic acid

ADP Adenosine diphosphate

ATCC American Type Culture Collection

AP-1 Activator protein-1

BCE Before current era (= B.C.)

B.P. British Pharmacopoeia

CE Current era (= A.D.)

CONSORT Consolidated Standards of Reporting Trials

COX Cyclooxygenase

ED50 Median effective dose

ELAM-1 Endothelial leukocyte adhesion molecule-1

EtOH Ethanol

g Gram

GC Gas chromatography

GC-MS Gas chromatography – mass spectrometry

GM-CSF Granulocyte-macrophage colony-stimulating factor

HCl Hydrochloric acid

HDL High density lipoprotein

HETEs Hydroxyeicosatetraenoic acids

HPETEs Hydroperoxyeicosatetraenoic acids

HPLC High-performance liquid chromatography

IC50 Median inhibition concentration

ICAM-1 Intracellular adhesion molecule-1

IFN-β Interferon-beta

IFN-γ Interferon-gamma

IL-1 Interleukin-1

IL-1β Interleukin-1-beta

IL-2 Interleukin-2

Page 11: Phytochemistry and pharmacology of plants from the ginger family

ix

IL-6 Interleukin-6

IL-9 Interleukin-9

iNOS Inducible nitric oxide synthase

JNK c-Jun N-terminal kinase

LC-MS Liquid chromatography – mass spectrometry

LDL Low density lipoprotein

IP-10 Interferon-γ activated protein

L Litre

LD50 Median lethal dose

LOX Lipoxygenase

5-LOX 5-Lipoxygenase

LPS Lipopolysaccharide

LTC4 Leukotriene C4

LTD4 Leukotriene D4

LTE4 Leukotriene E4

M Molar

MAPK, MAP kinase Mitogen-activated protein kinase

MIC Minimum inhibitory concentration

MCP-1 Macrophage (monocyte) chemotactic protein-1

MCSF Macrophage colony-stimulating factor

mg Milligram

MIP-1α Monocyte inflammatory protein-1-alpha

mL Millilitre

MMPs Metalloproteinases

MS Mass spectrometry

NFκB Nuclear factor kappa B

NK cells Natural killer cells

NMR Nuclear magnetic resonance

NO Nitric oxide

NSAIDs Non-steroidal anti-inflammatory drugs

ORAC Oxygen radical absorption capacity

PAF Platelet-activating factor

PBS Phosphate-buffered saline

PG Prostaglandin

Page 12: Phytochemistry and pharmacology of plants from the ginger family

x

PGD2 Prostaglandin D2

PGE2 Prostaglandin E2

PGF2α Prostaglandin F2-alpha

PGH2 Prostaglandin H2

PGI2 Prostaglandin I2 (prostacyclin)

PLA2 Phospholipase A2

PMNs Polymorphonuclear neutrophils

q.i.d. quarter in die (four times a day)

RCF Relative centrifugal force

SD Standard deviation

TFA Trifluoroacetic acid

t.i.d. ter in die (three times a day)

TLC Thin layer chromatography

TNF Tumour necrosis factor

TNF-α Tumour necrosis factor-alpha

TxA2 Thromboxane A2

TxB2 Thromboxane B2

UV Ultraviolet

UV-VIS Ultraviolet-visible

UVB Ultraviolet B

VAS Visual analogue scale

VCAM-1 Vascular cell adhesion molecule-1

VLDL Very low density lipoprotein

VR1 Vanilloid receptor-1

WOMAC Western Ontario and McMaster Osteoarthritis Index

μg Microgram

μL Microlitre

μM Micromolar

Page 13: Phytochemistry and pharmacology of plants from the ginger family

xi

TABLE OF CONTENTS PHYTOCHEMISTRY AND PHARMACOLOGY OF PLANTS FROM THE GINGER

FAMILY, ZINGIBERACEAE I

SYNOPSIS III

SYNOPSIS III

PUBLICATIONS ARISING FROM THE WORK IN THIS THESIS V

ACKNOWLEDGMENTS VI

ABBREVIATIONS VIII

TABLE OF CONTENTS XI

LIST OF TABLES XIX

LIST OF FIGURES XXI

1. GENERAL INTRODUCTION 1

1.1 Plants as medicines 1

1.2 Ethnobotany and ethnopharmacology 3 1.2.1 Ethnobotany 3 1.2.2 Ethnopharmacology 3 1.2.3 Ethnopharmacology and bioprospecting 4 1.2.4 Ethnobotany and ethnopharmacology as strategies in bioprospecting 5 1.2.5 Ethnopharmacology and phytotherapy 5

1.3 Inflammation 6 1.3.1 Overview over the inflammatory process 6 1.3.2 Cellular products as inflammatory mediators 7 1.3.3 Arachidonic acid metabolism 8

1.3.3.1 Prostaglandins 11 1.3.3.2 Thromboxanes 12

Page 14: Phytochemistry and pharmacology of plants from the ginger family

xii

1.3.3.3 Leukotrienes 12

1.4 Plants as anti-inflammatory agents 13

2. LITERATURE REVIEW 16

2.1 Introduction 16

2.2 The ginger family, Zingiberaceae Martinov 16 2.2.1 Zingiberaceae in Australia 17

2.3 Genus Zingiber Boehmer 18 2.3.1 Ginger (Zingiber officinale Roscoe) 19

2.3.1.1 Ginger chemistry 20 2.3.1.1.1 Non-volatile compounds 20

2.3.1.1.1.1 Historical background 21 2.3.1.1.1.2 Gingerols 22 2.3.1.1.1.3 Gingerols in fresh ginger 23 2.3.1.1.1.4 Gingerol degradation products 23 2.3.1.1.1.5 Other non-volatile compounds 26

2.3.1.1.2 Volatile oil 27 2.3.1.2 Pharmacokinetics of ginger 30 2.3.1.3 Pharmacodynamics of ginger 32

2.3.1.3.1 Anti-oxidant activity 32 2.3.1.3.2 Effects on arachidonic acid metabolism and eicosanoids 34

2.3.1.3.2.1 Assays using isolated enzymes 34 2.3.1.3.2.2 Cell-based assays 35 2.3.1.3.2.3 Animal studies 37 2.3.1.3.2.4 Summary: Effects on arachidonic acid metabolism and eicosanoids 39

2.3.1.3.3 Effects on cytokines and chemokines 40 2.3.1.3.4 Anti-cancer activity 41 2.3.1.3.5 Anti-microbial activity 41 2.3.1.3.6 Immunomodulatory effects 42 2.3.1.3.7 Gastrointestinal effects 42 2.3.1.3.8 Metabolic effects 43 2.3.1.3.9 Anti-atherosclerotic effects 43 2.3.1.3.10 Other cardiovascular effects 44 2.3.1.3.11 Analgesic effects 44 2.3.1.3.12 Antipyretic activity 45

2.3.1.4 Safety data 45 2.3.1.4.1 Acute toxicity 45

Page 15: Phytochemistry and pharmacology of plants from the ginger family

xiii

2.3.1.4.2 Teratogenicity and embryotoxicity 45 2.3.1.4.3 Mutagenicity 46

2.3.1.5 Clinical studies of ginger 46 2.3.1.5.1 Clinical studies of ginger in arthritis 46

2.3.2 Other Zingiber species 49 2.3.2.1 Thai ginger (Zingiber montanum) 49 2.3.2.2 Zingiber ottensii 51 2.3.2.3 Beehive ginger (Zingiber spectabile) 51

2.4 Genus Curcuma L. 52 2.4.1 Curcuma longa 52

2.4.1.1 Chemistry 52 2.4.1.2 Pharmacology 53

2.4.1.2.1 Pharmacokinetics of curcumin 54 2.4.1.2.2 Pharmacodynamics of curcumin 54

2.4.1.2.2.1 Anti-oxidant activity 54 2.4.1.2.2.2 Anti-inflammatory activity 57 2.4.1.2.2.3 Anti-carcinogenic and anti-tumour activity 59 2.4.1.2.2.4 Other pharmacological activities 61

2.4.2 Curcuma parviflora 62 2.4.3 Other Curcuma species 62

2.5 Genus Boesenbergia Kuntze 63 2.5.1 Chemistry 63 2.5.2 Pharmacology 64

2.6 Genus Hedychium Koenig 65 2.6.1 Chemistry 66 2.6.2 Pharmacology 66

2.7 Genus Kaempferia L. 67 2.7.1 Chemistry 67 2.7.2 Pharmacology 67

2.8 Genus Scaphochlamys Baker 68

2.9 Genus Alpinia Roxburgh 68 2.9.1 Chemistry 69 2.9.2 Pharmacology 70

2.10 Genus Pleuranthodium (K. Schum.) R. M. Smith 72

2.11 Genus Etlingera Giseke 72

Page 16: Phytochemistry and pharmacology of plants from the ginger family

xiv

2.12 Genus Elettaria Maton 73 2.12.1 Chemistry 73 2.12.2 Pharmacology 74

2.13 Genus Hornstedtia Retz. 75

2.14 Genus Renealmia L.f. 75

2.15 Genus Costus L. 75 2.15.1 Chemistry 76 2.15.2 Pharmacology 77

2.16 Genus Tapeinochilos Miquel 77

2.17 Summary 77

2.18 Aims and research questions 78 2.18.1 Part 1: Phytochemical investigations of 17 ginger clones 78

2.18.1.1 Research questions 78 2.18.1.2 Specific aims 79 2.18.1.3 Research methods 80

2.18.2 Part 2: Screening Zingiberaceae for pharmacological activity 80 2.18.2.1 Research questions 81 2.18.2.2 Specific aims 81 2.18.2.3 Research methods 82

3. INHIBITION OF CYCLOOXYGENASE-1 BY GINGER RHIZOME EXTRACTS 83

3.1 Introduction 83

3.2 Materials and Methods 83 3.2.1 Plant material 83 3.2.2 HPLC analysis 84 3.2.3 Cyclooxygenase-1 assay 84

3.2.3.1 Enzyme reaction 85 3.2.3.2 Extraction of arachidonic acid and metabolites 85 3.2.3.3 Separation of arachidonic acid and metabolites 85

3.2.4 Statistical analysis 86

3.3 Results and Discussion 87 3.3.1 Experiment 1 87 3.3.2 Experiment 2 89

Page 17: Phytochemistry and pharmacology of plants from the ginger family

xv

4. GINGEROL CONTENT OF SEVENTEEN GINGER (ZINGIBER OFFICINALE) CLONES 92

4.1 Introduction 92

4.2 Materials and Methods 93 4.2.1 Plant materials 93 4.2.2 Sample preparation 95 4.2.3 HPLC methods 95 4.2.4 Statistical analysis 96

4.3 Results 97 4.3.1 Measurement reproducibility 97 4.3.2 Quantification of gingerols in test samples 97 4.3.3 Correlation between gingerols 100 4.3.4 Gingerol ratios 102 4.3.5 Stability of gingerols 103 4.3.6 Gingerol content of tetraploid clones 103

4.4 Discussion 104 4.4.1 Gingerols 104 4.4.2 Shogaols 106 4.4.3 Ploidy 106

5. ESSENTIAL OIL COMPOSITION OF SEVENTEEN GINGER (ZINGIBER OFFICINALE) CLONES 108

5.1 Introduction 108

5.2 Materials and Methods 109 5.2.1 Plant materials and distillation 109 5.2.2 Chemical analysis 109 5.2.3 Statistical analysis 110

5.3 Results 110 5.3.1 Oil composition 110 5.3.2 Principal components analysis 113 5.3.3 Cluster analysis 115 5.3.4 Citral content 116 5.3.5 Neral to geranial ratio 117

5.4 Discussion 118

Page 18: Phytochemistry and pharmacology of plants from the ginger family

xvi

5.4.1 Oil composition 118 5.4.2 Citral content 120 5.4.3 Neral to geranial ratio 121

6. PHARMACOLOGICAL ACTIVITY OF ZINGIBERACEAE SPECIES 123

6.1 Introduction 123

6.2 Materials and Methods 125 6.2.1 Plant materials 125 6.2.2 Extraction methods 128

6.2.2.1 Extraction method 1 (ethanolic) 128 6.2.2.2 Extraction method 2 (ethanolic) 129 6.2.2.3 Extraction method 3 (ethanolic) 129 6.2.2.4 Extraction method 4 (ethanolic) 129 6.2.2.5 Extraction method 5 (aqueous) 130 6.2.2.6 Extraction method 6 (aqueous) 130

6.2.3 Bioassay methods 130 6.2.3.1 Inhibition of PGE2 production 130 6.2.3.2 Oxygen Radical Absorbance Capacity (ORAC) 132 6.2.3.3 Inhibition of nitric oxide production 132 6.2.3.4 Modulation of natural killer cell activity 133

6.3 Results 136 6.3.1 Inhibition of PGE2 production 136 6.3.2 Oxygen Radical Absorbance Capacity (ORAC) 138 6.3.3 Inhibition of nitric oxide production 139 6.3.4 Modulation of natural killer cell activity 140

6.4 Discussion 142 6.4.1 Inhibition of PGE2 production 142 6.4.2 Oxygen Radical Absorbance Capacity (ORAC) 144 6.4.3 Inhibition of nitric oxide production 145 6.4.4 Modulation of natural killer cell activity 146 6.4.5 Conclusion 147

7. BIOACTIVITY-GUIDED FRACTIONATION OF TWO NATIVE AUSTRALIAN ZINGIBERACEAE 149

7.1 Introduction 149

Page 19: Phytochemistry and pharmacology of plants from the ginger family

xvii

7.2 Materials and Methods 150 7.2.1 Plant material 150 7.2.2 Extraction methods 150

7.2.2.1 Curcuma australasica 150 7.2.2.2 Pleuranthodium racemigerum 150

7.2.3 Fractionation by preparative HPLC 151 7.2.3.1 Curcuma australasica 151 7.2.3.2 Pleuranthodium racemigerum 151

7.2.4 Nuclear magnetic resonance spectroscopy 152 7.2.4.1 One-dimensional NMR spectroscopy 152

7.2.4.1.1 1H NMR 152 7.2.4.1.2 J-modulated 13C NMR 152

7.2.4.2 Two-dimensional homonuclear correlation NMR spectroscopy 153 7.2.4.2.1 1H-1H Correlation spectroscopy (COSY) 153

7.2.4.3 Two-dimensional heteronuclear correlation spectroscopy 153 7.2.4.3.1 Heteronuclear single quantum correlation (HSQC) spectroscopy 153 7.2.4.3.2 Heteronuclear multiple-bond correlation (HMBC) spectroscopy 153

7.2.5 Determination of accurate mass 154 7.2.6 UV spectroscopy 154 7.2.7 Cytotoxicity assays 154 7.2.8 PGE2 assay 155

7.3 Results 155 7.3.1 Curcuma australasica 155

7.3.1.1 Activity-guided fractionation 155 7.3.1.2 Cytotoxic activity of Compound 1 160 7.3.1.3 Structural elucidation of Compound 1 nd 2 160

7.3.2 Pleuranthodium racemigerum 167 7.3.2.1 Activity-guided fractionation 167 7.3.2.2 Cytotoxic activity of Compound 3 170 7.3.2.3 Structural elucidation of Compound 3 172 7.3.2.4 Accurate Mass 176 7.3.2.5 UV spectrophotometric data and extinction coefficient 176

7.4 Discussion 178 7.4.1 Curcuma australasica 178 7.4.2 Pleuranthodium racemigerum 179

8. CONCLUDING REMARKS 180

8.1 Phytochemical investigations of ginger (Zingiber officinale) 180

Page 20: Phytochemistry and pharmacology of plants from the ginger family

xviii

8.2 Screening Zingiberaceae for pharmacological activity 182

8.3 Direction of future research 183

APPENDIX A: CLINICAL EFFICACY TRIALS OF GINGER PREPARATIONS 185

APPENDIX B: QUALITY ASSESSMENT OF CLINICAL TRIALS OF GINGER IN OSTEOATHRITIS 190

APPENDIX C: INHIBITION OF PGE2 IN 3T3 CELLS 192

REFERENCES 196

Page 21: Phytochemistry and pharmacology of plants from the ginger family

xix

LIST OF TABLES

Table 1-1. Cytokines involved in inflammation. 8

Table 1-2. Examples of plant compounds with in vitro anti-inflammatory activity. 15

Table 2-1. Taxonomy of the family Zingiberaceae. 17

Table 2-2. Subfamilies, tribes and representative genera of the Zingiberaceae. 17

Table 2-3. Zingiberaceae species occurring in Australia. 18

Table 3-1. Effect of extraction method: cyclooxygenase-1 inhibitory activity and content of major pungent compounds of ginger extracts.

87

Table 3-2. Effect of extraction solvent: cyclooxygenase-1 inhibitory activity and content of major pungent compounds of ginger extracts.

89

Table 3-3. Gingerol ratios in ginger extracts compared with values from the literature. 91

Table 4-1. Ginger clones studied, their genotype and origin. 94

Table 4-2. Results of pairwise analyses of 17 ginger clones in terms of [6]-, [8]- and [10]-gingerol content.

100

Table 4-3. Pearson Product-Moment correlations between the concentration of gingerols in fresh rhizomes of seventeen ginger clones assayed by HPLC at zero and five months.

102

Table 4-4. Mean±SE (range) concentrations of gingerols in two commercial ‘Queensland’ clones and 12 tetraploid clones (µg/g).

103

Table 4-5. Literature data on gingerol content of fresh ginger rhizomes. 105

Table 5-1. Composition of essential oils of 17 clones of ginger analysed by GC-MS on a BPX-5 column.

112

Table 5-2. Content of 14 constituents in essential oils of 16 ‘typical’ clones of ginger and one ‘atypical’ clone, ‘Jamaican’ (Z46).

113

Table 5-3. Varimax rotated component matrix for two component solution for the ten most abundant ginger essential oil constituents.

114

Table 6-1. Zingiberaceae samples screened for biological activity. 126

Table 6-2. Inhibition of PGE2 production in 3T3 murine fibroblasts. 137

Table 6-3. Oxygen radical absorbance capacity (ORAC). 138

Table 6-4. Inhibition of nitric oxide production in LPS-stimulated RAW264 macrophages.

139

Table 6-5. Effect of ethanolic extracts on natural killer cell activity against K562 leukaemia cells.

140

Table 6-6. Effect of aqueous extracts on natural killer cell activity against K562 leukaemia cells.

141

Table 7-1. Inhibition of PGE2 production in 3T3 murine fibroblast cells by fractions of Curcuma australasica extract.

157

Table 7-2. Cytotoxicity of Compound 1 from Curcuma australasica in P388D1 murine lymphoma cells.

160

Page 22: Phytochemistry and pharmacology of plants from the ginger family

xx

Table 7-3. 1H and 13C NMR spectral data of Compound 1 (zederone) from Curcuma australasica.

162

Table 7-4. 1H and 13C NMR spectral data of Compound 2 (1(10)E,4E-furanodien-6-one) from Curcuma australasica.

164

Table 7-5. 1H NMR spectral data from the literature for the isomers isofuranodienone and 1(10)Z,4Z-furanodien-6-one.

166

Table 7-6. Inhibition of PGE2 production in 3T3 murine fibroblast cells by combined fractions of Pleuranthodium racemigera extract.

168

Table 7-7. LD50 (µM) values for cytotoxic activity of Compound 3 and curcumin against five cancer cell lines.

172

Table 7-8. 1H and 13C NMR data from one-dimensional (1H and J-modulated 13C NMR) and two-dimensional correlation (COSY, HSQC and HMBC) NMR spectroscopy experiments on Compound 3 from Pleuranthodium racemigerum.

173

Table 7-9. Accurate mass determination for Compound 3. 176

Table 7-10. Peak absorption (λmax) and extinction coefficients (ε) of Compound 3 at 203 nm, 225 nm and 278 nm.

177

Page 23: Phytochemistry and pharmacology of plants from the ginger family

xxi

LIST OF FIGURES

Fig. 1-1. Schematic representation of the metabolism of arachidonic acid catalysed by cyclo-oxygenase (COX) and lipoxygenase (LOX) enzymes.

10

Fig. 2-1. Structure of [6]-gingerol, the most abundant gingerol in ginger rhizome. 21

Fig. 2-2. Structure of zingerone. 21

Fig. 2-3. Structure of [6]-shogaol. 21

Fig. 2-4. Structures of the major gingerols in ginger. 22

Fig. 2-5. Dehydration of gingerols to shogaols. 24

Fig. 2-6. Conversion of gingerols to zingerone and aliphatic aldehydes at high temperature.

25

Fig. 2-7. Paradols, gingerdiols and gingerdiones from Zingiber officinale. 27

Fig. 2-8. Volatile sesquiterpenes from Zingiber officinale. 28

Fig. 2-9. Volatile monoterpenoids from Zingiber officinale. 29

Fig. 2-10. Curcuminoids from Zingiber montanum. 50

Fig. 2-11. Structural diagrams of the major curcuminoids in turmeric rhizome: curcumin, demethoxycurcumin and bisdemethoxycurcumin.

53

Fig. 2-12. Parviflorene A and F from Curcuma parviflora. 62

Fig. 2-13. Boesenbergin A and B from Boesenbergia rotunda. 64

Fig. 2-14. 1’-acetoxychavicol acetate from Alpinia galanga. 70

Fig. 2-15. Terpinyl acetate and 1,8-cineole from Elettaria cardamomum. 74

Fig. 2-16. Diosgenin from Costus malortieanus. 76

Fig. 3-1. Gingerol content (mg/mL extract) of ginger extracts. 90

Fig. 4-1. Typical HPLC chromatogram of ethanolic extract of fresh ginger rhizome (Z44) obtained with HPLC Method 2.

98

Fig. 4-2. Concentrations of [6]-, [8]- and [10]-gingerol in fresh rhizomes of 17 ginger clones.

99

Fig. 4-3. Scatter plots illustrating the correlation between concentrations (µg/g) of [6]- and [8]-gingerol (A), [6]- and [10]-gingerol (B), and [8]- and [10]-gingerol (C) in fresh ginger rhizomes.

101

Fig. 5-1. Volatile constituents from Zingiber officinale essential oil. 111

Fig. 5-2. Component plot in rotated space showing Varimax rotated data on two components based on the ten most abundant constituents of ginger essential oil.

115

Fig. 5-3. Dendrogram of hierarchical cluster analysis of 17 essential oils of ginger. 116

Fig. 5-4. Scatter plot showing the relationship between the percentage content of the stereoisomers neral and geranial in essential oils from 17 clones of ginger.

117

Fig. 7-1. LC-MS chromatogram of Curcuma australasica extract redissolved in methanol.

156

Page 24: Phytochemistry and pharmacology of plants from the ginger family

xxii

Fig. 7-2. LC-MS chromatogram for Compound 1 (Fraction 17) from Curcuma australasica.

158

Fig. 7-3. Mass spectrum (M+1; left) and UV spectrum (right) for Compound 1 (Fraction 17) from Curcuma australasica.

158

Fig. 7-4. LC-MS chromatogram for Compound 2 (Fraction 20) from Curcuma australasica.

159

Fig. 7-5. Mass spectrum (M+1; left) and UV spectrum (right) for Compound 2 (Fraction 20) from Curcuma australasica.

159

Fig 7-6. Structure of zederone (Compound 1). 161

Fig. 7-7. Structure of 1(10)E,4E-furanodien-6-one (Compound 2) from Curcuma australasica.

165

Fig. 7-8. LC-MS chromatogram of Pleuranthodium racemigerum extract redissolved in 90% aqueous methanol.

167

Fig. 7-9. LC-MS chromatogram for Compound 3 from Pleuranthodium racemigerum. 169

Fig. 7-10. Mass spectrum (M+1; top) and UV spectrum (bottom) for Compound 3 from Pleuranthodium racemigerum.

169

Fig. 7-11. Cytotoxic effect of fraction Compound 3 from Pleuranthodium racemigerum on 3T3 murine fibroblasts.

170

Fig. 7-12. Dose-response curves for cytotoxic activity of Compound 3 from Pleuranthodium racemigerum against five cell lines.

171

Fig. 7-13. Heteronuclear correlations in Ring A. 174

Fig. 7-14. Heteronuclear correlations in Ring B. 175

Fig. 7-15. Chemical structure of Compound 3 from Pleuranthodium racemigerum, 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-2E-heptene.

175

Fig. 7-16. UV spectrum of 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-2E-heptene (Compound 3).

177

Page 25: Phytochemistry and pharmacology of plants from the ginger family

1

1. GENERAL INTRODUCTION This thesis reports on research into the chemistry and pharmacology of plants from the

Zingiberaceae family, in particular in terms of their potential use as anti-inflammatory

agents.

The experimental work presented in this thesis is divided into two main parts. The first part

focuses on the common ginger (Zingiber officinale), which in recent years has attracted

attention as a potential anti-inflammatory agent. Seventeen clones of ginger were analysed

in terms of their content of pungent compounds with a view to identify any with a profile

that might be of particular interest from a pharmacological point of view. Because ginger is

also an important flavouring commodity, profiling of the essential oil was also carried out.

The second part of the work comprises the screening of some 43 Zingiberaceae taxa,

representing 14 genera, for potential anti-inflammatory activity in a whole cell assay. Simple

chemical profiling of these extracts was carried out, and two were subjected to activity-

guided fractionation.

This chapter provides a brief introduction to the use of plants as medicines and the process of

inflammation. Chapter 2 provides a review of the plant species investigated during the

course of this work.

1.1 Plants as medicines

Plants have provided humans with medicines since time immemorial. The oldest known

document concerning medicinal plants and their uses is the Chinese Pen Ts’ao, which was

written 4800 years ago and describes no less than 360 plants, suggesting that herbal medicine

was already at an advanced stage in China at this time (Mann, 1992). In Mesopotamia (part

of present-day Iraq), 4600-year old clay tablets inscribed with cuneiform characters have

been found that contain references to familiar medicinal plants such as myrrh, licorice and

the opium poppy (Cragg & Newman, 2002). Another famous early document detailing the

use of plants as medicines is the Ebers papyrus from Egypt, which was written about 3500

years ago (Mann, 1992).

Page 26: Phytochemistry and pharmacology of plants from the ginger family

2

Much more ancient, albeit less conclusive, evidence suggests that humans might have

employed the pharmacological properties of plants much earlier. At the famous burial site in

the Shanidar Cave in the northern part of Iraq, a Neanderthal (Homo neanderthalensis) was

laid to rest with bunches of flowers about 60,000 years ago (Solecki, 1975). Of the eight

plants identified in the grave from preserved pollen, seven are considered medicinal plants

today. There is of course no way of knowing with certainty whether they were placed in the

grave because of their medicinal properties, to serve the dead man on his final journey, or

whether they simply were used for decorative purposes.

The more recent discovery of the ‘Iceman’ on the Italian-Austrian border in the Alps

provides intriguing evidence of early use of medicinal fungi in Europe. This hunter, who

had been lying well preserved in the ice for about 5300 years, was found to be in possession

of a fungus, the birch polypore (Piptoporus betulinus), which is known to have purgative and

antibiotic properties, and which he might well have been using to treat the whipworm

infestation of his intestines (Heinrich et al., 2004).

Plants play a key role in sophisticated ancient traditional medical systems such as traditional

Chinese medicine and Ayurveda of India, and have also been central in the Greco-Roman

medical tradition, which developed into modern biomedicine. Hippocrates (468-377 BCE),

used more than 400 plant species for therapeutic purposes, and it was a Roman army surgeon

by the name of Dioscorides (c. 40-80 CE) who wrote the most influential early European

manual of medicinal plants, De Materia Medica, in the first century CE (Griggs, 1997). This

comprehensive work included illustrations and descriptions of about 600 plant species, along

with text detailing their uses, doses and potential toxic effects. The writings of Galen (c.

129-199 CE), who classified herbs according to their humoral properties, had a profound and

almost unimaginable impact on medical thought in Europe for about 1500 years (Griggs,

1997). The English apothecary Nicholas Culpeper (1616-1654) wrote many herbal books,

the most famous being The English Physician (1653), in which he presented herbal medicine

in an astrological framework (Griggs, 1997).

Although modern biomedicine to a significant degree employs synthetic drugs as therapeutic

agents, plants still occupy a prominent place in contemporary pharmacy, either as sources of

pharmaceutical drugs in the form of isolated plant compounds, as sources of precursors to

drugs, or as sources of compounds that have served as models for synthetic or semisynthetic

drugs. It has been estimated that about one-half of all drugs in current use are natural

Page 27: Phytochemistry and pharmacology of plants from the ginger family

3

compounds or derivatives thereof (Iwu, 2002). It is however important to realise that despite

the many advances of biomedicine, the progress afforded residents of first world countries is

beyond the reach of the majority of the world’s population. For the majority of people, many

of whom live in miserable poverty, crude plants preparations are still the main form of

medicine. In acknowledgment of this situation, the World Health Organization (WHO) is

actively promoting the development of traditional medicine (Anonymous, 2002).

1.2 Ethnobotany and ethnopharmacology

1.2.1 Ethnobotany

The term ethnobotany lacks a singular, uniformly agreed definition. The term was coined by

J. W. Harshberger, who defined it as “…the use of plants by aboriginal peoples”

(Harshberger, 1896). Since then, ethnobotany has been redefined and reinterpreted by many

scholars in the area (for an overview, see Cotton, 1996). One of the broadest definitions of

ethnobotany is that provided by Martin, who described it as the subdiscipline of

ethnoecology that is concerned with local people’s interaction with plants (Martin, 1995).

Early ethnobotany was focused on plants of economic significance or potential, while

contemporary ethnobotany tends to have a far broader scope and include, for example,

traditional agricultural knowledge and traditional vegetation management (Cotton, 1996).

Throughout the history of formal ethnobotany, medicinal plants have been an area of keen

interest to many ethnobotanists.

1.2.2 Ethnopharmacology

Ethnopharmacology is a multidisciplinary field devoted to the study of pharmacologically

active agents traditionally used by humans. The term ethnopharmacology was coined as

recently as 1967 by Efron, who used the term in the context of hallucinogenic substances

(Heinrich & Gibbons, 2001).

Page 28: Phytochemistry and pharmacology of plants from the ginger family

4

More recently, ethnopharmacology has been defined as “… the interdisciplinary scientific

exploration of biologically active agents traditionally employed or observed by man”

(Bruhn & Holmstedt, 1981).

Ethnopharmacology applies conventional chemical and pharmacological analysis to

traditional medicines and in doing so differs from two related disciplines: medical

anthropology, which examines health and disease from a cultural perspective, and medical

ethnobotany, which is concerned with the use of plants within traditional medical systems

(Cotton, 1996).

Although ethnopharmacology is not exclusively concerned with plants or plant products, the

plant kingdom is the major focus of the discipline, because this kingdom has provided

humans with the greatest number of pharmacologically active substances throughout history.

The multidisciplinary nature of ethnopharmacology is evidenced by the important roles

played by fields such as botany, pharmacognosy, natural product chemistry, pharmacology,

toxicology, anthropology and others (Heinrich & Gibbons, 2001; Houghton, 2002).

1.2.3 Ethnopharmacology and bioprospecting

Heinrich and Gibbons (2001) have noted the differences between ethnopharmacology and

bioprospecting, while acknowledging that the two approaches are not mutually exclusive. In

brief, ethnopharmacology aims to develop (through increased knowledge and understanding)

the use of crude plant preparations in local communities, whereas the goal of bioprospecting

is the identification and development of compounds from nature as pharmaceutical drugs in

the international market place (Heinrich & Gibbons, 2001). Although the aims of these two

approaches to natural products research are vastly different from a socio-economic

viewpoint, many of the methodologies will often be the same, and work focussed on one

approach might yield results that are relevant to the other. For example, phytochemical and

pharmacological investigations of a traditional medicine might lead to the identification of a

compound that can be developed into a pharmaceutical drug. Well known examples of this

include ephedrine (from Ephedra spp.), atropine (from Atropa bella-donna and other

Solanaceae) and more recently the development of the anti-malarial drug artemether from the

Page 29: Phytochemistry and pharmacology of plants from the ginger family

5

lead compound artemisinin in the traditional Chinese herb Artemisia annua (Evans, 2002;

Wright, 2002).

1.2.4 Ethnobotany and ethnopharmacology as strategies in bioprospecting

Different approaches can be employed in the process of bioprospecting. Cotton (1996)

outlined three main approaches to the collection of plants for screening: the random method,

where every species in a given area is included; phylogenetic targeting, where a particular

taxon (such as a family) is targeted, because it is already known to be good source of

pharmacologically active metabolites; and the ethno-directed sampling, which is guided by

traditional plant use. The latter approach is based on the notion that initial screening and

selection has already been conducted effectively by the owners of the traditional knowledge.

The ethno-directed approach to identifying plants with biological activity has been shown in

a number of studies to be more efficient than the random method at identifying plants with

promising pharmacological activity. Such studies include one aimed at identifying plants

with anti-HIV activity from Central America (Balick, 1990) and another that showed that

plants used ethnomedically to treat viral infections were more than 100 times more likely to

yield compounds with anti-viral activity than randomly collected plants (Carlson, 2002).

However, it has been argued that the successful development of drugs from traditional

medicines is most likely for conditions such as inflammation, gastrointestinal or nervous

system disorders, because these pathologies are widely recognised and treated in indigenous

systems of medicine (Cox, 1994).

1.2.5 Ethnopharmacology and phytotherapy

The term phytotherapy is used to describe the use of plant-based, chemically complex

therapeutic agents in contemporary, mostly industrialised societies. Phytotherapy is usually

based on a history of traditional use, but it differs from traditional indigenous herbal

medicine by employing industrialised extraction and manufacturing methods and by being

cosmopolitan in scope. Hence phytomedicines made from plants from around the globe are

Page 30: Phytochemistry and pharmacology of plants from the ginger family

6

available in most industrialised countries. Ethnopharmacology has the potential to increase

our knowledge and understanding of traditional herbal medicines, how they work, how they

are best prepared, and how they can be applied in a safe and efficacious manner. Due to the

chemical complexity of both traditional herbal medicines and modern phytomedicines, the

task of elucidating their pharmacology is a complex one indeed. A full understanding of

how complex mixtures of plant compounds interact with the human body and with each

other is probably not achievable, and pharmacological investigations of plant extract almost

always focus on one or a few ‘active’ constituents, i.e. compounds with profound biological

activity. It should always be borne in mind that many other compounds present in a plant

extract could potentially play a role in the overall activity of that extract, for example by

modulating the pharmacokinetic and/or pharmacodynamic properties of the ‘actives’.

Despite this caveat, ethnopharmacological investigations clearly have much to offer modern

phytotherapy, and the long-term success of the ‘herbal renaissance’ currently experienced in

most of the industrialised world undoubtedly depends on the scientific underpinning of

traditional or anecdotal uses.

1.3 Inflammation

Inflammation is being implicated in the pathophysiology of an increasing number of

diseases. In addition to conditions traditionally considered to be inflammatory in nature,

inflammation is now considered to have a role in a wide range of pathologies, including

cardiovascular disease (Hansson, 2005; Kaperonis et al., 2006), cancer (Zhang & Rigas,

2006), diabetes (Deans & Sattar, 2006; Duncan & Schmidt, 2006), age-related macular

degeneration (Rodrigues, 2007), Parkinson’s disease (Hald et al., 2007), Alzheimer’s disease

(Eikelenboom et al., 2006), and possibly depression (Kulmatycki & Jamali, 2006).

1.3.1 Overview over the inflammatory process

Inflammation is a rapid and non-specific response to cellular injury in vascularised tissues.

The inflammatory response is produced and controlled by complex interactions between

Page 31: Phytochemistry and pharmacology of plants from the ginger family

7

cellular and plasma protein components. The cellular component involves intercellular

communication effected by a range of cytokines.

The inflammatory response commences with a brief constriction of arterioles followed by

vasodilation and exudation of protein-containing plasma and blood cells into the injured

tissue. This causes swelling and oedema. Meanwhile leukocytes adhere to vessel walls and

cause the endothelial cells to contract, creating enough space between these cells for the

leukocytes to enter the extravascular tissue. Increased vascular permeability is maintained

until the inflammatory state is resolved, and it is the interplay between blood cells and

plasma proteins in the affected tissue that controls the inflammatory response and interacts

with part of the immune response.

The principal cell types involved in inflammation are mast cells, endothelial cells,

phagocytic leukocytes (polymorphonuclear neutrophils, macrophages, and eosinophils), and

platelets.

Mast cells play a key role in the initiation of the inflammatory response. Degranulation leads

to the release of stored chemicals such as histamine, which causes increased vascular

permeability, and mast cells also synthesise pro-inflammatory mediators such as

prostaglandins, leukotrienes and platelet-activating factor (PAF).

Endothelial cells express adhesion molecules (selectins) for leukocytes and platelets, and

also produce nitric oxide (NO), which causes vasodilation but also may play a regulatory

role by suppressing mast cell and platelet function. Endothelial cells also produce two

prostaglandin derivatives with opposite action: the vasoconstrictor thromboxane A2 (TxA2)

and the vasodilator prostacyclin (PGI2), and it is the interplay between these two regulatory

compounds that allows for platelet aggregation to occur only at the site of injury (McCance

& Huether, 2002).

1.3.2 Cellular products as inflammatory mediators

The various cells involved in the inflammatory response produce a range of compounds that

act as inflammatory mediators, including cytokines and products of arachidonic acid

metabolism, i.e. prostaglandins and leukotrienes.

Page 32: Phytochemistry and pharmacology of plants from the ginger family

8

Cytokines are proteins produced by a range of different cell types. The major types of

cytokines are the interleukins and the interferons, but the class also includes tumour necrosis

factors, colony-stimulating factors, transforming growth factor, and others (McCance &

Huether, 2002).

Table 1-1 lists cytokines that play an important role in inflammation.

Table 1-1. Cytokines involved in inflammation.

(After McCance et al. 2002).

Cytokine Source Main actions relevant to inflammation

IL-1 Mainly macrophages Inflammatory mediator; increases prostaglandin production

IL-6 Monocytes, macrophages, T and helper T cells

Stimulates inflammatory response

IL-9 Helper T cells T cell and mast cell growth factor IFN-γ (IL-18) T and helper T cells, NK cells Activates macrophages TNF-α Macrophages, mast cells,

lymphocytes Increases other cytokines; increases inflammatory and immune responses

Macrophage colony-stimulating factor

Inflamed endothelial cells, monocytes, lymphocytes, fibroblasts

Macrophage growth factor

Transforming growth factor-β Macrophages, platelets, lymphocytes

Chemotactic for macrophages; increases IL-1 production

1.3.3 Arachidonic acid metabolism

Products of arachidonic acid metabolism such as prostaglandins and leukotrienes play key

roles in inflammation, and their syntheses are well established targets in the pharmacological

treatment of inflammation.

Arachidonic acid (5Z,8Z,11Z,14Z-eicosatetraenoic acid) is an unsaturated, 20-carbon,

omega-6 fatty acid found in cell membranes. It can be obtained from the diet or be derived

Page 33: Phytochemistry and pharmacology of plants from the ginger family

9

from linoleic acid. In the cell membrane, arachidonic acid is esterified to phospholipid, and

arachidonate must be liberated from phospholipid before it can act as a substrate for

enzymatic modification. These modifications, catalysed by various enzymes, are known as

arachidonic acid metabolism and can lead to the formation of inflammatory mediators

collectively known as eicosanoids, i.e. prostaglandins (PGs), thromboxanes (TXs),

hydroxyeicosatetraenoic acids (HETEs) and leukotrienes (LTs) (Calder, 2005; Eberhart &

Dubois, 1995).

Arachidonic acid is released from the cell membrane by phospholipase enzymes, in

particular phospholipase A2, and the free acid can be metabolised to eicosanoids by cyclo-

oxygenase (COX) and lipoxygenase (LOX) enzymes. Metabolism catalysed by COX

enzymes gives rise to prostaglandins of the 2-series as well as thromboxanes, while LOX

metabolism leads to the formation of leukotrienes (Fig. 1-1).

COX is also known as prostaglandin endoperoxide synthase. This protein possesses two

discrete activities: the cyclo-oxygenase activity first inserts two oxygen molecules into

arachidonic acid resulting in PGG2; this is followed by the reduction of PGG2 to PGH2, a

result of the protein’s peroxidase activity (Needleman et al., 1986). Since prostaglandins of

the 2-series, including PGE2, are involved in many inflammatory processes, the inhibition of

the COX pathway of arachidonic acid metabolism is a prime pharmacological target and one

that has been exploited long before the pathway was known. Non-steroidal anti-

inflammatory drugs (NSAIDs) inhibit the cyclo-oxygenase activity (but not the peroxidase

activity) of prostaglandin endoperoxide synthase (Marnett et al., 1999). Acetylsalicylic acid

(aspirin) does so in an irreversible fashion, by acetylating the enzyme (Foster, 1996;

Needleman et al., 1986).

Two isoforms of the COX enzyme are known; these are known as COX-1 and COX-2 and

their respective protein products show differential distribution (Marnett et al., 1999). COX-1

is constitutively expressed in most tissues, where it catalyses the biosynthesis of eicosanoids

(prostaglandins and thromboxanes) that regulate numerous cellular processes. In contrast,

COX-2 activity is generally undetectable in most tissues, but the expression of COX-2 can

be rapidly induced in inflammatory cells in response to stimulation by pro-inflammatory

cytokines or by growth factors (Calder, 2005). A new class of anti-inflammatory and

analgesic agents that are selective COX-2 inhibitors were developed in the 1990s (e.g.

celecoxib, refecoxib). The rationale for this drug development was that selective COX-2

Page 34: Phytochemistry and pharmacology of plants from the ginger family

10

inhibitors were not expected to interfere with homoeostatic physiological processes and

should therefore be less likely than non-selective COX inhibitors to cause the unwanted side-

effects typical of traditional NSAIDs (Lipsky, 1999; Sautebin, 2000; Simon et al., 1999).

Fig. 1-1. Schematic representation of the metabolism of arachidonic acid catalysed by cyclo-oxygenase (COX) and lipoxygenase (LOX) enzymes.

Already in the late 1990s there was growing evidence, however, that viewing COX-2 as

solely pro-inflammatory and COX-1 as essentially benign was too simplistic. COX-2 has

now been shown to be involved in normal physiology such as the regulation of vascular and

renal blood flow (Brater et al., 2001), and Sautebin reviewed work suggesting that COX-2

could have anti-inflammatory action, while COX-1 could contribute to inflammation under

certain circumstances (Sautebin, 2000). Nevertheless, COX-2 inhibitory drugs (coxibs) were

aggressively marketed and became very widely prescribed for inflammatory conditions in

industrialised countries in the early 2000s, only for some of them to be withdrawn from the

5-HPETE

PGH2 15-HETE 12-HETE LTA4

15-HPETE 12-HPETE

5-HETE

PGE2 PGD2 PGF2α PGI2 TX2

PGG2

Lipoxin A4 LTB4 LTC4

LTD4

LTE4

Phospholipase A2

15-LOX 12-LOX 5-LOX COX

PGJ4

Arachidonic acid in cell membrane

Free arachidonic acid

PEROXIDASE

Page 35: Phytochemistry and pharmacology of plants from the ginger family

11

market a few years later (2004) due to the increased cardiovascular risks associated with

their use (rofecoxib [Vioxx®] was withdrawn; celecoxib [Celebrex®] is still used, but in a far

more discriminate manner) (Fitzgerald, 2004; Yoon & Baek, 2005).

A so-called splice variant of COX-1, known as COX-3, COX-1b or COX-1v, is also known,

but after initial speculation that it might be a target for paracetamol, medical interest in it has

waned (Hersh et al., 2005).

The alternative metabolic pathway of arachidonic acid is controlled by the lipoxygenase

(LOX) enzymes (Fig. 1-1). They catalyse the insertion of molecular oxygen into

polyunsaturated fatty acids with a 1Z,4Z-pentadiene system. Depending of the location of

the oxygen insertion into arachidonic acid, mammalian lipoxygenase enzymes are classified

as 5-, 12- or 15-lipoxygenases. The primary products are 5-, 12- and 15-

hydroperoxyeicosatetraenoic acids (HPETEs), which are subsequently reduced to the

corresponding hydroxyeicosatetraenoic acids (HETEs) or, in case of the 5-LOX pathway,

converted to the eicosanoids known as leukotrienes (Calder, 2005; Needleman et al., 1986;

Schneider & Bucar, 2005).

1.3.3.1 Prostaglandins

The prostaglandins are a group of cyclic, 20-C unsaturated fatty acids, which all share a

double-bond at C13-C14. Based on structural differences, the prostaglandins (PGs) are

divided into 9 groups, named A-I. A subscript numeral indicates the number of unsaturated

carbon bonds in the compound, and the subscripts α- and β- denotes the orientation of a

hydroxyl-group on C9 below or above the molecular plane (e.g. PGF2α) (Foster, 1996).

Originally discovered in seminal fluid from the prostate (giving rise to their name),

prostaglandins are known to be synthesised and released in virtually all body tissues. No

tissue (except seminal fluid) appears to have the capacity to store prostaglandins, so rate of

release reflects the rate of biosynthesis. Prostaglandins are produced in response to a variety

of stimuli, including inflammation, allergic responses and trauma, and they usually exert

their action locally, close to their site of release. Several prostaglandins are known to be

metabolised rapidly in the liver, kidneys and lungs (Foster, 1996).

Page 36: Phytochemistry and pharmacology of plants from the ginger family

12

Five prostanoid receptors named DP, EP, FP, IP and TP have been identified that show some

degree of selectivity for PGD2, PGE2, PGF2α, PGI2 and thromboxane A2 (TxA2),

respectively. The existence of several subtypes of the EP receptor has been proposed and it

is believed that PGE2 exerts different actions at these (Foster, 1996).

The formation of different PGs from the unstable metabolite PGH2 is cell-specific. For

example, PGI2 (prostacyclin) is the predominant prostaglandin in the vascular endothelium,

where it inhibits platelet aggregation and causes vasodilation. In mast cells, PGD2 is the

major COX product, while PGE2 is the prevailing prostaglandin in the kidney (Needleman et

al., 1986). PGE2 has several pro-inflammatory activities, including vasodilation and

increasing vascular permeability, inducing fever, and enhancing pain and oedema caused by

other mediators such as bradykinin and histamine. Large amounts of PGE2 and PGF2 can be

produced by stimulated monocytes and macrophages, whereas stimulated neutrophils

produce moderate amounts of PGE2 (Calder, 2005).

1.3.3.2 Thromboxanes

The thromboxanes are eicosanoids arising from the COX metabolic pathway like the

prostaglandins. Thromboxane A2 (TxA2) is the major metabolite of PGH2 in platelets.

Unlike its metabolite thromboxane B2 (TxB2), which is inactive, TxA2 contracts vascular

smooth muscle, induces platelet aggregation and causes serotonin release (Needleman et al.,

1986).

1.3.3.3 Leukotrienes

Leukotrienes (LTs) are potent biologically active compounds produced by the 5-LOX

pathway via 5-HPETE. LTC4, LTD4, and LTE4 are produced mainly by mast cells,

basophils, eosinophils and endothelial cells from the precursor LTA4, while LTB4 is

produced in neutrophils, monocytes and macrophages (Calder, 2005; Jala & Haribabu,

2004).

LTB4 is a potent chemotactic agent for leucocytes, it increases vascular permeability, causes

the release of lysosomal enzymes, and promotes the generation of reactive oxygen species

Page 37: Phytochemistry and pharmacology of plants from the ginger family

13

and the production of inflammatory cytokines including tumour necrosis factor α (TNFα),

interleukin-1 (IL-1) and interleukin-6 (IL-6). LTC4, LTD4, and LTE4 increase vascular

permeability, but also cause bronchoconstriction and promote hypersensitivity reactions

(Calder, 2005).

Leukotrienes have been associated with a range of inflammatory conditions including

asthma, colitis, dermatitis, rheumatoid arthritis and septic peritonitis, and recent work has

suggested an important role for leukotrienes in the development and progression of

atherosclerosis (Jala & Haribabu, 2004). The modification of leukotrienes as a

pharmacological target has only been achieved relatively recently, but several drugs

(inhibitors of leukotrienes biosynthesis and receptor antagonists) are now approved for the

treatment of asthma and allergic rhinitis (Jala & Haribabu, 2004; Valli, 2003).

1.4 Plants as anti-inflammatory agents

When you examine a man with an irregular wound…and that wound is

inflamed…[there is] a concentration of heat; the lips of that wound are reddened

and that man is hot in consequence…Then you must make cooling substances for

him to draw the heat out…leaves of willow.

(Translated entry from the Ebers papyrus, as quoted in Mann 1992)

The bark and leaves of willow (Salix spp.) are among the most famous plants with anti-

inflammatory properties. A source of a range of salicylates, mostly in glycoside form,

willow has been used medicinally for millennia. The parent compound of these salicylates,

salicylic acid, was shown to be highly effective in the treatment of rheumatic fever in one of

the earliest clinical trials in history. Its corrosive effects in the gastrointestinal tract lead to

the development of a derivative, acetylsalicylic acid, which was marketed by the German

pharmaceutical company Bayer in 1899 under the name Aspirin (Mann, 1992). The most

successful pharmaceutical drug of all time, acetylsalicylic acid became the prototype for the

class of pharmaceuticals known as non-steroidal anti-inflammatory drugs (NSAIDs).

However, it was not until 1971 that the main mechanism of action for Aspirin was

elucidated. In groundbreaking work led by John Vane, who was later rewarded with a Nobel

Page 38: Phytochemistry and pharmacology of plants from the ginger family

14

Prize, the key to Aspirin’s anti-inflammatory action was shown to be its potent ability to

inhibit cyclooxygenase (Mann, 1992).

Over the past two decades, numerous plant extracts and plant compounds have been

investigated for their ability to modulate inflammation. Most of these investigations have

been conducted in vitro or in animal models, while only a relatively small number of human

trials have been conducted in this area. Plant compounds with anti-inflammatory activity

have been reviewed and arachidonic acid metabolism, nitric oxide and nuclear factor kappa

B (NFκB) identified as major targets (Bremner & Heinrich, 2002; Calixto et al., 2003). In

many cases such anti-inflammatory activity appears to be the result of the ability of a

compound to inhibit the action and/or biosynthesis of pro-inflammatory cytokines,

chemokines or adhesion molecules involved in the inflammatory process, for example by

activating transcription factors (incl. NFκB) and protein kinases (Calixto et al., 2004). Table

1-2 provides examples of anti-inflammatory plant compounds and their targets.

Page 39: Phytochemistry and pharmacology of plants from the ginger family

15

Table 1-2. Examples of plant compounds with in vitro anti-inflammatory activity.

Compound(s) Plant source Molecular target Reference

Isocoumarins Hydrangea dulcis Mast cells (Matsuda et al., 1999)

Resveratrol Vitis vinifera Mast cells (Baolin et al., 2004)

Terminoside A Terminalia arjuna Nitric oxide (Ali et al., 2003)

Orixalone A Orixa japonica Nitric oxide (Ito et al., 2004)

Eugenol Ocimum sanctum COX-1 (Kelm et al., 2000)

Ursolic acid Plantago major COX-2 (Ringbom et al., 1998)

Wogonin Scutellaria baicalensis

COX-2 expression (Chen et al., 2001)

Rhamnetin Guiera senegalensis 5-lipoxygenase (Bucar et al., 1998)

Maesanin Maesa lanceolata 5-lipoxygenase (Abourashed et al., 2001)

Ugandensidial Warburgia ugandensis

5-lipoxygenase 12-lipoxygenase

(Wube et al., 2006)

In terms of arachidonic acid metabolism, it is noteworthy that a number of plant compounds

and extracts have been shown to be dual inhibitors of cyclo-oxygenase and 5-lipoxygenase

enzymes in vitro (Liu et al., 1998; Resch et al., 1998).

Page 40: Phytochemistry and pharmacology of plants from the ginger family

16

2. LITERATURE REVIEW 2.1 Introduction

Many plants belonging to the ginger family, Zingiberaceae, have a history of medicinal use

in systems of traditional medicine. Best known are ginger (Zingiber officinale) and turmeric

(Curcuma longa), both of which has been the subject of substantial pharmacological and

clinical investigations over the last three decades, but many lesser known species are also

used, mostly in tropical Asia, where the majority are native. Several species in the family are

also important spices.

This chapter provides background information on the genera included in the present work,

with emphasis on chemistry and pharmacology, in particular as it relates to potential anti-

inflammatory activity. Ginger has been given special attention in this review. Zingiberaceae

genera not included in the present study have not been reviewed.

2.2 The ginger family, Zingiberaceae Martinov

The ginger family, Zingiberaceae, is a monocotyledenous family in the order Zingiberales.

The family comprises some 52 genera with a total about 1100 species. The family is

essentially tropical in distribution, with few species occurring in temperate climates, and is

particularly richly represented in the Indomalesian flora, i.e. from India to New Guinea.

Zingiberaceae species typically have thickened rhizomes with secretory cells producing

essential oil (Mabberley, 1997). The taxonomy of the Zingiberaceae according to K.

Kubitzki’s The Families and Genera of Vascular Plants as presented by Mabberley (1997) is

outlined in Table 2-1.

Page 41: Phytochemistry and pharmacology of plants from the ginger family

17

Table 2-1. Taxonomy of the family Zingiberaceae. (After Mabberley 1997).

Phylum Angiospermae Class Monocotyledonae (Liliopsida) Subclass Zingiberidae Order Zingiberales Family Zingiberaceae

The Zingiberaceae is divided into two subfamilies (Table 2-2). The subfamily

Zingiberoideae comprises four tribes. The subfamily Costoideae is commonly treated as a

separate family, Costaceae (Meissner) Nakai (Mabberley, 1997).

Table 2-2. Subfamilies, tribes and representative genera of the Zingiberaceae. (After Mabberley 1997.) Genera in bold were included in the present work.

Subfamily Tribe Representative genera

Zingiberoideae Hedychieae Boesenbergia, Curcuma, Hedychium, Kaempferia, Scaphochlamys

Zingibereae Zingiber Alpinieae Aframomum, Alpinia, Amomum, Elettaria,

Etlingera, Hornstedtia

Globbeae Globba Costoideae Costus, Tapeinochilus

2.2.1 Zingiberaceae in Australia

In Australia, the Zingiberaceae is represented by 14 native species from 8 genera. In

addition, 7 introduced species from 5 genera also occur (Smith, 1987) (Table 2-3).

Page 42: Phytochemistry and pharmacology of plants from the ginger family

18

Table 2-3. Zingiberaceae species occurring in Australia.

Native species in bold were included in the present study. (After Smith 1987).

Genus Native species Introduced and naturalised species

Alpinia A. arctiflora (F. Muell.) Benth. A. arundelliana (Bailey) Schumann A. caerulea (R. Br.) Benth. A. hylandii R. M. Smith A. modesta F. Muell. ex Schumann Amomum A. dallachyi F. Muell. A. queenslandicum R. M. Smith Costus C. potierae F. Muell. C. dubius (Afzel.) Schumann Curcuma C. australasica J. D. Hook C. longa L. Etlingera E. australasica (R. M. Smith) R. M. Smith Globba G. marantina L. Hedychium H. coronarium Koenig H. gardnerianum Sheppard ex

Ker Gawler Hornstedtia H. scottiana (F. Muell.) Schumann Kaempferia Kaempferia sp. Pleuranthodium P. racemigerum (F. Muell.) R. M. Smith Tapeinochilos T. ananassae (Hassk.) Schumann Zingiber Z. officinale Roscoe

Z. zerumbet (L.) Smith

2.3 Genus Zingiber Boehmer

The genus Zingiber comprises approximately 60 species ranging from India through tropical

Asia (Mabberley, 1997; Smith, 1987). There are no native Australian representatives of this

genus (despite the erroneous statement to the contrary by Mabberley, 1997), but two species,

Z. officinale Roscoe and Z. zerumbet (L.) Smith, have been reported as naturalised in tropical

north Queensland (Hnatiuk, 1990; Smith, 1987).

Page 43: Phytochemistry and pharmacology of plants from the ginger family

19

The generic name comes from the Greek Zingiberi, which in turn is derived from an Indian

word meaning ‘root’ (Smith, 1987).

Several species of Zingiber have a history of traditional medicinal, culinary or other

ethnobotanical uses. These include Japanese or mioga ginger (Z. mioga (Thunb.) Roscoe),

used against malaria and as a vermifuge in China (Mabberley, 1997); Z. montanum (J.

König) Theilade (syn. Z. cassumunar, Z. purpureum), used for a wide range of complaints in

India and South-East Asia (Johnson, 1999); and Z. zerumbet (L.) Smith, which has also been

employed for a range of conditions in Asia and the Pacific (Johnson, 1999).

The common ginger (Z. officinale Roscoe), which has been a major focus of the present

work, is treated in detail below. In addition, other Zingiber species included in this work are

briefly reviewed.

2.3.1 Ginger (Zingiber officinale Roscoe)

Ginger (Zingiber officinale Roscoe) is a sterile, reed-like plant with a pungent and aromatic

rhizome on which it relies for vegetative propagation. The plant is a cultigen, that is, it is

only known from cultivation. Its wild origins are not known with certainty but are believed

to be India or South-East Asia (Mabberley, 1997; Vaughan & Geissler, 1997). Ginger has a

very long history of use, both as a spice and as a medicinal plant, and is mentioned in ancient

Sanskrit texts and in classical Buddhist, Arabic, Greek and Roman literature (Govindarajan,

1982a). It was used widely in Europe by the tenth century (Vaughan & Geissler, 1997) and

was first exported from Jamaica, where it became a significant agricultural crop, in 1547

(Mabberley, 1997). It is now cultivated in many tropical and subtropical regions including

India, Africa, China, the West Indies and Australia, with the annual world production

estimated at 100,000 tons in 2000 (Bartley & Jacobs, 2000; Evans, 2002).

Ginger rhizome is valued as a spice for its combination of pungent and aromatic qualities,

which arise from its content of phenolic compounds and essential oil, respectively. Ginger is

used as flavouring in a vast array of foods, including savoury dishes such as curries, and

sweets such as cakes and biscuits, and also in beverages such as ginger ale, ginger beer and

ginger wine.

Page 44: Phytochemistry and pharmacology of plants from the ginger family

20

Ginger rhizome (known as Rhizoma Zingiberis in pharmacy) is used in several traditional

systems of medicine, including Traditional Chinese Medicine, Ayurveda and Western herbal

medicine (WHO, 1989; Williamson, 2002). Its traditional uses cover a great variety of

complaints including dyspepsia, flatulence and colic, nausea and vomiting, colds and flu,

migraine, as well as muscular and rheumatic disorders (WHO, 1999).

2.3.1.1 Ginger chemistry

The secondary metabolites found in the rhizome of ginger that are of primary interest can

broadly be divided into volatile compounds (extractable by steam distillation) and non-

volatile phenolic compounds, the major ones of which have pungent properties. It is

generally considered that the pharmacological activity of ginger rhizome resides with

compounds from these classes, in particular the non-volatile pungent phenolic compounds.

The term oleoresin, when applied to ginger, refers to the volatile oil, the pungent compounds

and other compounds extracted by means of solvents (ethanol or acetone) (Connell, 1969;

Govindarajan, 1982a).

2.3.1.1.1 Non-volatile compounds

Ginger owes its pungency to phenolic compounds. In the fresh rhizome the major type

comprises a series of homologous phenolic alkanones known as gingerols and derivatives

thereof such as gingerdiols. The principal of these compounds is [6]-gingerol with 8- and 10-

gingerol occurring in lower concentrations (Connell & Sutherland, 1969; Denniff et al.,

1981). When subjected to heat or alkali treatment, however, gingerols are converted to a

corresponding series of homologous shogaols by dehydration and/or to the compound

zingerone (Connell, 1969; Connell & Sutherland, 1969). The shogaols possess greater

pungency than the corresponding gingerols (Denniff et al., 1981).

Page 45: Phytochemistry and pharmacology of plants from the ginger family

21

2.3.1.1.1.1 Historical background

In 1879, Tresh isolated an oily pungent concentrate from ginger oleoresin and called it

gingerol (Tresh, 1879). In 1917, English and Japanese researchers independently isolated

two pungent ginger compounds, gingerol (Fig. 2-1) and zingerone (Fig. 2-2) (Lapworth et

al., 1917).

In 1927, the Japanese group published the structural characterisation of another pungent

ginger compound, shogaol (Fig. 2-3) (after shoga, the Japanese word for ginger) (Connell,

1969; Govindarajan, 1982a).

H3CO

HO

CH3

O

Fig. 2-2. Structure of zingerone.

H3CO

HO

(CH2)4

O OH

CH3

Fig. 2-1. Structure of [6]-gingerol, the most abundant gingerol in ginger rhizome.

Fig. 2-3. Structure of [6]-shogaol.

H3CO

HO

(CH2)4

O

CH3

Page 46: Phytochemistry and pharmacology of plants from the ginger family

22

Little progress on the chemistry of pungent ginger compounds was made until the 1960s,

when D. W. Connell of the Queensland Department of Primary Industries commenced his

work in the area.

2.3.1.1.1.2 Gingerols

Comparing the chemical composition of commercially prepared ginger oleoresin and an

oleoresin extracted with cold solvent, Connell's group found surprising differences. The

major pungent compound identified in the commercial oleoresin was shogaol and despite

repeated attempts, gingerol could not be isolated from this sample. The freshly prepared

oleoresin extracted with cold solvent, however, had a different major constituent, which was

isolated and identified as gingerol (Connell, 1969; Connell & Sutherland, 1969).

Extensive work on ginger oleoresin led Connell and Sutherland to suggest that although both

shogaol and zingerone had been isolated from oleoresins, they were in fact both artefacts, or

at the most very minor constituents of ginger rhizomes. The presence of shogaol and

zingerone in easily detectable quantities, according to Connell and Sutherland, were

indicative of the oleoresin having been exposed to excessive heat in the course of extraction

(Connell & Sutherland, 1969).

Although gingerol is normally an oily substance, Connell and Sutherland were able to obtain

a crystalline solid when storing the gingerol in hexane at –30°C. This solid was shown to

consist of a mixture of homologous phenolic ketones, identified as [6]-, [8]- and [10]-

gingerol (Fig. 2-4). [4]- and [12]-gingerols were not identified, but their presence in trace

amounts could not be excluded.

H3CO

HO

(CH2)n

O OH

CH3

[6]-gingerol: n = 4[8]-gingerol: n = 6[10]-gingerol: n = 8

Fig. 2-4. Structures of the major gingerols in ginger.

Page 47: Phytochemistry and pharmacology of plants from the ginger family

23

He and coworkers also identified [6]-, [8]- and [10]-gingerol in a methanol extract of fresh

ginger rhizome analysed by HPLC-electrospray mass spectrometry (He et al., 1998).

2.3.1.1.1.3 Gingerols in fresh ginger

Several studies have reported on the concentration of gingerols in fresh ginger rhizomes.

Zhang and colleagues analysed by reversed-phase high-performance liquid chromatography

(HPLC) freshly harvested rhizomes from Hawaii extracted with methanol, and found the

concentration of [6]-gingerol to be 2100 µg per gram fresh rhizome. The corresponding

concentrations of [8]- and [10]-gingerol were 288 and 533 µg/g, respectively (Zhang et al.,

1994). Fresh rhizomes purchased locally in the United States, extracted by methylene

chloride and analysed by HPLC, yielded 880, 93 and 120 µg per gram fresh rhizome of [6]-,

[8]- and [10]-gingerol, respectively (Hiserodt et al., 1998). Fresh rhizome from Taiwan was

reported to have a [6]-gingerol concentration of 806 µg/g by HPLC (Young et al., 2002). A

supercritical CO2 extract of ginger grown on the Australian east coast and analysed by

negative ion electrospray-MS contained 120, 19 and 24 µg per gram fresh rhizome of [6]-,

[8]- and [10]-gingerol, respectively (as well as shogaols that may have formed from

gingerols during processing) (Bartley, 1995). The considerable variation in gingerol

concentrations across these studies may reflect genetic or environmental differences, as well

as the variable methodological approaches, or a combination of these. Because ginger is a

sterile cultigen with a very long history of cultivation in different parts of the world, genetic

differences between clones are likely to be an important determinant of variation in

secondary metabolites.

2.3.1.1.1.4 Gingerol degradation products

Connell (1969) commented on the remarkable extent to which chemical changes occur in

ginger. It is interesting to note that these changes are reflected in the different therapeutic

applications of fresh and processed ginger in Oriental medicine (Hikino, 1985).

Page 48: Phytochemistry and pharmacology of plants from the ginger family

24

The main pungent compounds in fresh ginger, [6]-, [8]- and [10]-gingerol, are thermally

unstable and can undergo at least two reactions (Connell, 1969; Connell & Sutherland,

1969).

Firstly, [6]-, [8]- and [10]-gingerol can undergo dehydration and convert to [6]-, [8]- and

[10]-shogaol, respectively, when exposed to high temperature or subjected to prolonged

storage (Fig. 2-5) (He et al., 1998; Zhang et al., 1994).

Secondly, a retro-aldol reaction can give rise to zingerone and a series of aliphatic aldehydes,

which can cause undesirable flavours in the oleoresin (Fig. 2-6) (Connell, 1969).

Experiments showed that aldehyde formation occurred when oleoresin was heated to

temperatures above 200°C.

+ H2O H3CO

HO

(CH2)n

O

CH3

[6]-shogaol: n = 4[8]-shogaol: n = 6[10]-shogaol: n = 8

H3CO

HO

(CH2)n

O OH

CH3

Fig. 2-5. Dehydration of gingerols to shogaols.

Heat

Page 49: Phytochemistry and pharmacology of plants from the ginger family

25

In studies of the fate of pure gingerol at pH 3-5, which is the typical pH-range for oleoresin

samples, the conversion to shogaol was found to proceed at a very slow rate at room

temperature. At higher temperature, however, the reaction occurred more rapidly. Under

alkaline conditions, the dehydration reaction proceeded readily at room temperature,

following first order reaction kinetics. Studying this dehydration reaction in the oleoresin at a

constant temperature of 120°C, the gingerol content was found to decrease, with the rate of

loss increasing with decreasing pH. This reaction also followed first order kinetics. The

half-life of gingerol in ginger oleoresin at 120°C was found to range from approximately 3

hours at pH 2.4 to 15 hours at pH 7.2 (Connell, 1969; Connell & Sutherland, 1969).

Connell also extracted an oleoresin from freshly harvested ginger rhizome, using cold

acetone, which was subsequently removed from the extract at or below room temperature

(Connell, 1969). This extract showed a high gingerol content with little or no shogaol

present. From the same batch of ginger rhizome, two other oleoresins were prepared, one

from the whole dried rhizome and another from the sliced dried rhizome. Whole rhizomes

required longer heating time to dry than did sliced rhizomes. As expected, the oleoresin

from the whole dried rhizome had a higher shogaol content.

One of these oleoresins, with a natural pH of 3.5, was stored at ambient temperature for a

period of 18 months. During this time, a large proportion of the gingerol was converted to

H3CO

HO

(CH2)n

O OH

CH3

H3CO

HO

CH3

O

+ CH3(CH2) nCHO

200°C+

Fig. 2-6. Conversion of gingerols to zingerone and aliphatic aldehydes at high temperature.

zingerone

Page 50: Phytochemistry and pharmacology of plants from the ginger family

26

shogaol, and whereas gingerol was the major pungent compound in the original oleoresin,

shogaol was predominant in the stored product (Connell, 1969).

Whether shogaols occur in fresh ginger rhizome or form as artefacts of various extraction

and manufacturing processes has been a matter of some contention. Govindarajan, in his

comprehensive review of ginger from 1982, stated that several thin-layer chromatography

(TLC) studies had shown that shogaol is ‘definitely’ present in the fresh rhizome as a minor

component (Govindarajan, 1982a). This appeared to have been confirmed by liquid

chromatography – mass spectrometry (LC-MS) by He and co-workers who reported a small

amount of [6]-shogaol (He et al., 1998), but their sample was a methanolic extract prepared

by reflux, a process that would likely have provided the necessary heat for [6]-shogaol to

form from [6]-gingerol. [6]-Shogaol (and in one case also 8- and 10-shogaol) have also been

reported from supercritical carbon dioxide extracts of fresh ginger by GC-MS (Bartley,

1995; Bartley & Jacobs, 2000), but the high temperatures involved in this analytical method

(up to 290ºC) readily account for the formation of shogaols from gingerols. Chen and co-

workers found only a trace amount of [6]-shogaol in a supercritical CO2 extract prepared

from freeze-dried fresh rhizomes and analysed by HPLC, which does not involve high

temperatures (Chen et al., 1986b), and TLC of extracts of fresh Australian ginger did not

demonstrate the presence of shogaol (Connell & Sutherland, 1969).

More recent studies have indicated that in aqueous solution [6]-gingerol and [6]-shogaol

exist in a pH-dependent equilibrium. One study found that [6]-gingerol was most stable at

pH 4, while at pH 1 and 100ºC conversion of [6]-gingerol to [6]-shogaol was relatively fast

and reached equilibrium within two hours (Bhattarai et al., 2001). Another study found that

in simulated gastric fluid (pH 1 at 37ºC) [6]-gingerol and [6]-shogaol underwent fist order

reversible dehydration and hydration reactions to form [6]-shogaol and [6]-gingerol,

respectively (Bhattarai et al., 2007). The same study found that in simulated intestinal fluid

(pH 7.4 at 37ºC) insignificant interconversion of the two compounds took place.

2.3.1.1.1.5 Other non-volatile compounds

A range of other compounds related to gingerols as well as several diarylheptanoids have

been isolated from ginger rhizome. These include [4]-, [6]-, [8]- and [10]-gingerdiol, [6]-

methylgingerdiol, [6]-methylgingerdiacetate, [6]-paradol, gingerdiones, [6]-gingesulfonic

Page 51: Phytochemistry and pharmacology of plants from the ginger family

27

acid, [6]-hydroxyshogaol, [6]-, [8]- and [10]-dehydroshogaol and hexahydrocurcumin

(Kikuzaki, 2000 and references therein).

More recently, workers at the University of Arizona have published numerous compounds

from ginger, including many novel compounds (Jolad et al., 2004; Jolad et al., 2005). These

include paradols, dihydroparadols, acetyl derivatives of gingerols, 3-dihydroshogaols,

gingerdiols, mono- and diacetyl derivatives of gingerdiols, 1-dehydrogingerdiones,

diarylheptanoids, methyl [8]-paradol, methyl [6]-isogingerol, methyl [4]-shogaol, [6]-

isoshogaol, 6-hydroxy-[8]-shogaol, 6-hydroxy-[10]-shogaol, 6-dehydro-[6]-gingerol, three 5-

methoxy-[n]-gingerols (n = 4, 8 and 10), 3-acetoxy-[4]-gingerdiol, 5-acetoxy-[6]-gingerdiol,

diacetoxy-[8]-gingerdiol, methyl diacetoxy-[8]-gingerdiol, 5-acetoxy-3-deoxy-[6]-gingerol,

1-hydroxy-[6]-paradol and others. Some of these compounds could potentially be artifacts.

Some of the compounds reported from ginger are shown in Fig. 2-7 below.

(CH2)n

HO

On=6 [6]-paradoln=8 [8]-paradoln=10 [10]-paradol

HO

OH

(CH2)n

OHn=2 [4]-gingerdioln=4 [6]-gingerdioln=6 [8]-gingerdioln=8 [10]-gingerdiol

HO

O

(CH2)n

On=4 [6]-gingerdionen=6 [8]-gingerdionen=8 [10]-gingerdionen=10 [12]-gingerdione

CH3

CH3

CH3

MeO

MeO

MeO

Fig. 2-7. Paradols, gingerdiols and gingerdiones from Zingiber officinale.

2.3.1.1.2 Volatile oil

Commercial ginger oil is obtained by steam distillation of coarsely ground dried ginger

rhizome, and most published compositional analyses refer to oil prepared from dried raw

Page 52: Phytochemistry and pharmacology of plants from the ginger family

28

material. Ginger oil distilled from dry material is characterised by a high proportion of

sesquiterpene hydrocarbons and relatively small amounts of monoterpene hydrocarbons and

oxygenated compounds (Govindarajan, 1982a). The major sesquiterpene hydrocarbons are

zingiberene, ar-curcumene, β-bisabolene, (-)-β-sesquiphellandrene and (E,E)-α-farnesene

(Lawrence, 1995b), although published data indicate the relative abundance of these

compounds varies greatly (Fig. 2-8).

H3C

ar-curcumene

zingiberene

H

β-sesquiphellandrene

β-bisabolene

(E,E)-α-farnesene

α-copaene

Fig. 2-8. Volatile sesquiterpenes from Zingiber officinale.

Both zingiberene and (-)-β-sesquiphellandrene can be oxidised to ar-curcumene in oil stored

under unfavourable conditions (Connell & Jordan, 1971; Govindarajan, 1982a). High levels

of ar-curcumene may therefore be an indicator of degraded oil, but could also possibly

reflect distillation conditions (Govindarajan, 1982a).

Page 53: Phytochemistry and pharmacology of plants from the ginger family

29

Other constituents of ginger essential oil widely reported include α-pinene, camphene, 6-

methyl-5-hepten-2-one, myrcene, α- and β-phellandrene, limonene, 1,8-cineole, linalool,

borneol, α-terpineol, citronellol, neral, geraniol, geranial, bornyl acetate, 2-undecanone,

citronellyl acetate, α-copaene and geranyl acetate (Lawrence, 1995b; Lawrence, 1997;

Lawrence, 2000). Some of the monoterpenoids in the oil are shown in Fig. 2-9.

myrcene

α-pinene

α-phellandrene β-phellandrene

OO

bornyl acetate

O

O

citronellyl acetate

O

Ogeranyl acetate

Fig. 2-9. Volatile monoterpenoids from Zingiber officinale.

Australian ginger grown in Queensland yields a steam-distilled oil known for its citrus-like

aroma. Connell and Jordan compared 35 samples of commercially and laboratory prepared

Australian ginger oils with ten samples of commercial oils from other major ginger-

producing areas (Jamaica, Nigeria, Sierra Leone, China and India) (Connell & Jordan, 1971).

The compounds chiefly responsible for this citrus-like aroma were identified as the

monoterpene aldehydes geranial and neral (the mixture of which is known as citral). The 35

Australian oil samples analysed by Connell and Jordan contained 3-20% geranial and 1-10%

neral, whereas the oils from other parts of the world contained these compounds from trace

amounts to 3 and 1%, respectively (Connell & Jordan, 1971). The same study found that

Page 54: Phytochemistry and pharmacology of plants from the ginger family

30

apart from their high citral content, Australian ginger oils were similar to other oils in terms

of their major constituents, these being (-)-zingiberene (20-28%), (-)-β-sesquiphellandrene

(7-11%), ar-curcumene (6-10%) and β-bisabolene (5-9%). These sesquiterpenes, however,

are only weakly odorous and are believed to make only minor contributions to the aroma and

flavour of the oil (Connell & Jordan, 1971).

It is clear from more recent studies, however, that ginger oils from regions other than

Australia can also have a high citral content. The highest citral content of ginger oil reported

in the literature was 66% in one of five samples (ranging from 28% to 66%) prepared from

fresh rhizomes from Fiji (Smith & Robinson, 1981). Rhizomes (probably fresh) from the

Central African Republic steam-distilled for five hours yielded an oil containing 34.6% citral

(Menut et al., 1994), and a Mauritian ginger oil distilled from freshly harvested rhizomes

contained 26.6% citral (Gurib-Fakim et al., 2002). Ekundayo and colleagues compared oils

distilled for three hours from fresh and sun-dried Nigerian ginger (Ekundayo et al., 1988).

The oil made from fresh and dried plant material contained 23.9% and 14.3% citral,

respectively. Some Asian ginger oils have been reported to have low levels of citral

(Macleod & Pieris, 1984; Vernin & Parkanyi, 1994), but an Indian oil distilled (4 hours)

from fresh rhizomes contained 25% (Sharma et al., 2002), and an oil from Taiwan reportedly

had a citral content of 25% (Lawrence, 1995a).

Indian investigators have reported that the citrus-like aroma of ginger oil (which is due to

citral) is characteristic of oil distilled from fresh or quickly dried ginger, whereas commercial

oils and oils made from sun-dried ginger lack the lemony note (Govindarajan, 1982a). Thus

it is presently unclear whether the reported variability in citral content is due to genetic,

environmental or ontogenic factors, to factors relating to post-harvest handling or processing,

or to a combination of these.

2.3.1.2 Pharmacokinetics of ginger

No information is available regarding the pharmacokinetics of crude ginger preparations.

Several in vitro and animal studies and one in vitro study employing human hepatic and

intestinal microsomes have investigated the metabolism of the major pungent compound in

Page 55: Phytochemistry and pharmacology of plants from the ginger family

31

ginger, [6]-gingerol. One in vitro study has explored the hepatic metabolism of [6]-shogoal.

These studies are outlined below.

Experiments in rats with acute renal or hepatic failure showed that the elimination half-life

and plasma concentration of intravenously administered [6]-gingerol increased significantly

after hepatic failure was induced with carbon tetrachloride. The induction of renal failure,

however, did not affect clearance of [6]-gingerol from plasma. These results suggest that the

liver but not the kidney is involved in the elimination of [6]-gingerol (Naora et al., 1992).

The same study found more than 90% of [6]-gingerol present in plasma was bound to serum

protein. [6]-Gingerol incubated with phenobarbital-induced rat liver supernatant containing

the NADPH-generating system was converted to two diastereomers of [6]-gingerdiol (Surh

& Lee, 1994a).

Another study characterised a number of metabolites following the oral administration of

[6]-gingerol to rats (Nakazawa & Ohsawa, 2002). The major metabolite in the bile was (S)-

[6]-gingerol-4’-O-β-glucuronide, while the β-glucuronidase-treated urine contained vanillic

and ferulic acids, 9-hydroxy-[6]-gingerol, (S)-(+)-[6]-gingerol and two substituted organic

acids. These results demonstrated that orally administered [6]-gingerol undergoes

conjugation, and ω-1 oxidation and β-oxidation of a phenolic side chain. About half of the

oral dose of [6]-gingerol was excreted in the bile as the glucuronide. Incubation of [6]-

gingerol with rat liver produced only (S)-[6]-gingerol-4’-O-β-glucuronide and 9-hydroxy-

[6]-gingerol (glucuronidation and ω-1 oxidation), suggesting that metabolism must also

occur outside the liver. The contribution of gut microflora was demonstrated by the fact that

gut sterilisation decreased the amount of urinary metabolites.

In a recent study, [6]-gingerol incubated with human hepatic microsomes fortified with

uridine diphosphoglucuronic acid (UDPGA; a source of glucuronic acid) was glucuronidated

primarily at the phenolic hydroxyl group with a small amount of glucuronidation also

occurring at the aliphatic hydroxyl group, while human intestinal microsomes formed the

phenolic glucuronide only (Pfeiffer et al., 2006).

[6]-shogaol incubated with rat liver supernatant containing the NADPH-generating system

was converted to saturated ketone and reduced alcohol metabolites as well as an allyl

alcohol, 1-(4'-hydroxy-3'-methoxyphenyl)-deca-4-ene-3-ol (Surh & Lee, 1994b).

Page 56: Phytochemistry and pharmacology of plants from the ginger family

32

It is clear from the above that the current state of knowledge of the pharmacokinetics of

ginger compounds in humans is embryonic. Expanding this knowledge and including

information about oral bioavailability of compounds with known pharmacological activity

should be a priority and ought to precede further clinical trials of ginger for inflammatory

conditions.

2.3.1.3 Pharmacodynamics of ginger

The pharmacology and therapeutic use of ginger have been the subject of several recent

reviews (Afzal et al., 2001; Chrubasik et al., 2005; Grzanna et al., 2005).

Studies conducted in vitro have shown gingerols to act as agonists of the vanilloid receptor

(VR1) (Dedov et al., 2002). This receptor is also activated by capsaicin, the major pungent

principle in cayenne and chilli pepper, which shares structural features with the gingerols

(vanillin-derived compounds with phenolic 3-methoxy-, 4-hydroxy-groups). Earlier work

suggested that [6]-shogaol might share its site of action with capsaicin (Onogi et al., 1992);

if this were true, [6]-shogaol might also be a VR1 agonist.

Ginger extracts and various ginger constituents have demonstrated a wide range of

pharmacological activities in vitro and in vitro. Here the emphasis will be on the activities

relevant to inflammation, while other pharmacological activities will be briefly summarised.

2.3.1.3.1 Anti-oxidant activity

Ginger extracts have demonstrated anti-oxidant activity in a number of in vitro assay

systems, including 1,1-diphenyl-2-picrylhydrazyl (DPPH) radical scavenging activity,

inhibitory effect on oxidation of methyl linoleate under aeration and heating by the Oil

Stability Index (OSI) method, inhibitory effect on oxidation of liposome induced by 2,2'-

azobis(2-amidinopropane)dihydrochloride (AAPH) (Masuda et al., 2004), oxygen radical

absorbance capacity (ORAC) (Ninfali et al., 2005), Fe2+-induced lipid peroxidation, 1,1-

diphenyl-2-picrylhydrazyl (DPPH) radical quencher assay (Kuo et al., 2005), β-carotene

bleaching method (Kaur & Kapoor, 2002) and ferric reducing ability of plasma (FRAP)

(Halvorsen et al., 2002).

Page 57: Phytochemistry and pharmacology of plants from the ginger family

33

[6]-Gingerol, [6]-shogaol, [6]-gingerdiol, two other gingerol-related compounds as well as 8

arylheptanoids delayed the oxidation of linoleic acid in the ferric thiocyanate assay

(Kikuzaki & Nakatani, 1993). More recently, Masuda and coworkers have identified more

than 50 compounds with anti-oxidant activity from ginger rhizome (Masuda et al., 2004).

They were either related to gingerols or diarylheptanoids. Structure-activity studies of the

gingerol-type compounds suggested that both the substituents on and the length of the alkyl

chain might contribute to the anti-oxidant activity. A glucoside of [6]-gingerdiol has also

been found to possess strong anti-oxidant activity in vitro (Sekiwa et al., 2000).

Several animal studies have demonstrated the anti-oxidant activity of ginger in vivo.

Lipid peroxidation was significantly lowered in rats fed ginger (1%); the activities of the

antioxidant enzymes superoxide dismutase (SOD), catalase (CAT) and glutathione

peroxidase were maintained, and the blood glutathione content was significantly increased.

Similar effects were achieved with ascorbic acid (100 mg/kg) treatment, suggesting that at

least in this assay ginger was comparable to ascorbic acid as an antioxidant (Ahmed et al.,

2000a).

Rats on a high fat diet supplemented with ginger (35 mg/kg and 70 mg/kg) had lowered

levels of tissue thiobarbituric acid reactive substances (TBARS) and hydroperoxides, raised

activities of SOD and CAT, and raised levels of reduced glutathione in the aorta, liver,

kidney and intestine, compared with unsupplemented controls (Jeyakumar et al., 1999).

In another study, rats fed dried ginger (1%) showed significantly attenuated oxidative stress

and lipid peroxidation in response to exposure to the organophosphorous pesticide malathion

for 4 weeks (20 ppm) (Ahmed et al., 2000b).

Atherosclerotic, apolipoprotein E-deficient mice fed ginger extract showed a significant

reduction in the LDL basal oxidative state, with reduced susceptibility to oxidation and

aggregation (Fuhrman et al., 2000).

Pre-treatment with a hydroethanolic extract of ginger reduced the severity of radiation

sickness and the mortality in mice exposed to gamma irradiation (Jagetia et al., 2003).

Elevated lipid peroxidation and depletion of glutathione following irradiation were both

reduced by the ginger treatment. Treatment with ginger after irradiation was not effective

(Jagetia et al., 2004).

Page 58: Phytochemistry and pharmacology of plants from the ginger family

34

No studies have as yet directly assessed the anti-oxidant potential of ginger in humans, but

based on the substantial amount of evidence from both in vitro and in vivo studies, there are

sound reasons to predict that ginger might be a potent anti-oxidant in humans, although any

systemic effects could be limited if the compounds responsible have poor oral

bioavailability.

2.3.1.3.2 Effects on arachidonic acid metabolism and eicosanoids

Arachidonic acid metabolism is a major inflammatory pathway, and most of the work

relevant to the anti-inflammatory activity of ginger has focused on its effects on this

pathway. This substantial body of work is reviewed below under the headings (1) Assays

using isolated enzymes, (2) Cell-based assays, and (3) Animal studies.

2.3.1.3.2.1 Assays using isolated enzymes

Kiuchi and coworkers were the first group to publish details of the ability of ginger

compounds to modulate arachidonic acid metabolism (Kiuchi et al., 1982). They worked on

fresh ginger rhizomes obtained from a Japanese market and isolated [6]-gingerol, [6]-

gingerdione, [10]-gingerdione, [6]-dehydrogingerdione and [10]-dehydrogingerdione. All of

these compounds were shown to be potent inhibitors of prostaglandin-synthetase

(cyclooxygenase) in vitro (although details of the assay method were not provided), with

IC50 values comparable to that of indomethacin. Kiuchi's group subsequently reported the

potent inhibitory activity on prostaglandin-synthetase of [6]-gingerol, [10]-gingerol, [6]-

shogaol and numerous gingerol derivatives (Kiuchi et al., 1992). Very strong inhibitory

activity was found for [10]-gingerol, [6]-acetylgingerol, [6]-shogaol, [6]-

dehydrogingerdione, [10]-dehydrogingerdione, [6]-gingerdione and [10]-gingerdione, all of

which had IC50 values <2.5 µM.

[8]-Paradol has also been shown to be a potent inhibitor of ovine cyclooxygenase-1 (COX-1)

with an IC50 value of 4±1 μM (Nurtjahja-Tjendraputra et al., 2003).

Of particular interest was the finding by Kiuchi’s group that the homologous series of

synthetic gingerols (of which [6]-, [8]- and [10]-gingerol occur in ginger) were dual

Page 59: Phytochemistry and pharmacology of plants from the ginger family

35

inhibitors of arachidonic acid metabolism, i.e. inhibitors of both the cyclooxygenase and 5-

lipoxygenase pathways (Kiuchi et al., 1992). This finding appears to confirm an earlier

finding suggesting that several ginger compounds inhibit both pathways (Flynn et al., 1986),

while being at odds with the finding by Srivastava (see below), that fractionated ginger

preparations increased the formation of lipoxygenase products (Srivastava, 1984b;

Srivastava, 1986). It is possible, however, that these seemingly incongruous results are in

fact all accurate and reflect chemical differences between the test preparations.

Dual inhibitors of arachidonic acid metabolism are of significant interest as potential

therapeutic substances, because their inhibitory effect on the formation of eicosanoids is

likely to some extent to resemble the therapeutic effects of steroidal drugs, which inhibit

phospholipase A2 and therefore inhibit the release from phospholipid membranes of

arachidonic acid itself, resulting in potent anti-inflammatory activity.

2.3.1.3.2.2 Cell-based assays

Srivastava published two studies describing the effects of a ginger extract and its fractions on

arachidonic acid metabolism and platelet aggregation in washed human platelets (Srivastava,

1984a; Srivastava, 1984b). The preparation used in these studies was described as an

aqueous extract, but it is unclear whether it was prepared from fresh or dried plant material.

Regardless, thromboxane B2 (TxB2) formation was reduced by 73%, while formation of

prostaglandin F2α (PGF2α), prostaglandin E2 (PGE2) and prostaglandin D2 (PGD2) and the

prostaglandin endoperoxides, prostaglandin G2 (PGG2) and prostaglandin H2 (PGH2) was

also significantly reduced.

Additional data published by Srivastava and colleagues confirmed the previously observed

effects on the production of TxB2 and also found inhibition of lipoxygenase products by

ginger components at higher concentrations (Bordia et al., 1997; Srivastava, 1986).

An 80% ethanolic ginger extract (uncertain whether from fresh or dried material) was found

to significantly and dose-dependently inhibit prostaglandin release (expressed in terms of

PGE2) by stimulated rat peritoneal leukocytes (Mascolo et al., 1989).

Further evidence of ginger’s inhibitory activity on thromboxane formation was provided by

Bordia and coworkers who showed that an n-hexane extract of ginger powder dose-

Page 60: Phytochemistry and pharmacology of plants from the ginger family

36

dependently inhibited thromboxane formation in washed platelets challenged with (1-14C)-

labelled arachidonic acid (IC50 =16±8µg/mL, n=3). Using autoradiography, the

investigators found that 12-hydroxy-5,8,10-heptadecatrienoic acid (12-HHT) was reduced

with a concomitant increase in 12-hydroxyeicosatetraenoic acid (12-HETE) (Bordia et al.,

1997).

Inhibition of LPS-induced PGE2 production in human leukaemia HL-60 cells with IC50

values comparable to that of indomethacin was demonstrated for fractions of fresh ginger

(Jolad et al., 2004) and commercially processed dry ginger (Jolad et al., 2005; Lantz et al.,

2007).

Flynn and colleagues were the first to study the effects of pure synthetic ginger compounds

on arachidonic acid metabolism (Flynn et al., 1986). This study used intact human

leukocytes and found gingerdione to be a potent inhibitor of both cyclooxygenase and 5-

lipoxygenase. [6]-Shogaol and a related compound preferentially inhibited 5-

hydroxyeicosatetraenoic acid (5-HETE) formation (i.e. the 5-lipoxygenase pathway), while

[6]-gingerol and dehydroparadol appeared to be more potent inhibitors of PGE2 formation

(i.e. the cyclooxygenase pathway). All compounds studied possessed a 4-hydroxy-3-

methoxyphenyl moiety (which is also a major determinant of flavour). The same study also

found curcumin from turmeric (Curcuma longa) to be a dual inhibitor of arachidonic acid

metabolism.

Guh et al. investigated the anti-platelet mechanism of ‘gingerol’ isolated from ginger in

washed rabbit platelets (Guh et al., 1995). It is not clear from the report whether the plant

material was fresh or dried, and the method of isolation is not described. Neither is it clear

what is meant by ‘gingerol’. The report includes a diagram of [8]-gingerol that carries the

caption ‘Chemical structure of gingerol’. However, since [8]-gingerol is not the major

gingerol present in ginger, some uncertainty as to the exact identity of ‘gingerol’ remains.

‘Gingerol’ and indomethacin (a COX inhibitor) were found to significantly inhibit the

arachidonic acid-induced formation of TxB2 and PGD2 in a concentration-dependent manner.

In contrast, imidazole, which is a thromboxane synthase inhibitor, reduced TxB2 formation

while causing a highly significant rise in PGD2 production. These findings are important,

because they demonstrate that ‘gingerol’, like indomethacin, inhibits cyclooxygenase rather

than thromboxane synthase. It was also demonstrated that the anti-platelet activity of

Page 61: Phytochemistry and pharmacology of plants from the ginger family

37

‘gingerol’ was due to the subsequent reduced production of TxA2, a potent platelet

aggregating agent.

Venkateshwarlu was the first to study the effects of ginger compounds on eicosanoid

metabolism in whole blood (Venkateshwarlu, 1997). He showed that [6]- and [10]-gingerol

isolated from fresh ginger dose-dependently inhibited the formation of TxB2 in horse blood.

Curcumin and another compound isolated from dried rhizomes of turmeric (Curcuma longa)

as well as piperine isolated from dried pepper (Piper nigrum) fruits were also found to be

active.

[8]-gingerol and [8]-gingerdiol were potent inhibitors of COX activity in RBL-2H3 cells

with IC50 values of 1.54 and 3.3 μM, respectively, which was similar to that of indomethacin

(0.76 μM) (Koo et al., 2001). The inhibitory activity of the major pungent phenolic in fresh

ginger, [6]-gingerol, was an order of magnitude weaker (50 μM) in this study.

Tjendraputra and co-workers showed that a number of naturally occurring gingerol-type

compounds ([8]- and [10]-gingerol, [6]-shogaol, [8]-gingerdiol and [8]-paradol, as well as

several synthetic analogues) inhibited COX-2 activity in cultured human airways epithelial

A549 cells, with IC50 values in the low micromolar range (Tjendraputra et al., 2001).

Interestingly, [6]-gingerol did not inhibit COX-2 in this assay.

2.3.1.3.2.3 Animal studies

Sharma and coworkers examined the effects of a steam-distilled ginger oil on chronic

adjuvant arthritis in rats (Sharma et al., 1994). The oil, which was obtained from a company

in Germany, contained camphene, limonene, 1,8-cineole, linalool, citral and borneol. It is

noticeable that three of the major constituents of most ginger oils, zingiberene, ar-

curcumene, and β-sesquiphellandrene, were not listed as constituents of this oil. Rats in the

verum group (n=8) were given ginger oil (33 mg/kg) orally once daily for 26 days,

commencing the day prior to the induction of adjuvant arthritis. Ginger oil treatment

produced significant suppression of knee joint swelling on day 2 after induction of oedema

(p<0.05) and an even greater suppression of joint swelling on day 26 (p<0.001). In terms of

paw swelling, a significant inhibitory effect was evident at 18-26 days (p<0.001). It is

noteworthy that this study used a steam-distilled oil, which should lack gingerols and

Page 62: Phytochemistry and pharmacology of plants from the ginger family

38

shogaols. If the ginger preparation indeed was devoid of gingerols and shogaols, these

findings indicate that steam-distilled ginger oil contains other compounds with potential anti-

inflammatory activity.

In 1997, Sharma et al. re-published the findings of the study first published in 1994 (Sharma

et al., 1997). Included were new data on plasma and tissue kallikrein (also known as

kallidinogenase) levels in synovial tissue. The kallidinogenases are enzymes responsible for

the formation of kinins from their precursors, kininogens and kallidinogens. Kinins have

been implicated in many physiological functions, including the regulation of blood flow and

blood pressure, allergy and inflammation. Increases in kinin formation or kinin release have

been shown to occur in many inflammatory conditions. The presence of high levels of

plasma kallidinogenase-like enzyme has previously been demonstrated in blood-free paw

tissue from rats with adjuvant inflammation. The study found that raised levels of plasma

kallidinogenase fell significantly in rats treated with ginger oil for 26 days (p<0.01). Ginger

oil treatment produced no change in synovial tissue kallidinogenase levels. According to the

authors, these findings suggest that ginger oil might possess anti-rheumatic and anti-

inflammatory properties by inhibiting the kinin-forming kallidinogenase enzyme.

An aqueous preparation made from fresh ginger was given to rats in a study examining the

ex vivo effects on TxB2 and PGE2 synthesis (Thomson et al., 2002). Administered orally or

intraperitoneally at two doses (50 mg/kg and 500 mg/kg), the low dose significantly reduced

serum PGE2 levels only when given orally, whereas the higher dose was effective with either

route of administration. Serum TxB2 levels were reduced significantly only with oral

administration of the higher dose.

[6]-Gingerol has been tested for anti-inflammatory activity against carrageenan-induced paw

oedema in rats (Young et al., 2005). At doses of 50 mg/kg, 100 mg/kg and 250 mg/kg, [6]-

gingerol was effective in reducing oedema; an ED50 value of 85 mg/kg suggested a

considerable anti-inflammatory effect in this assay.

Penna and co-workers found that rat paw oedema induced by carrageenan, compound 48/80

(a synthetic mast cell-degranulating compound) and serotonin was significantly inhibited by

intraperitoneal administration of a 70% aqueous ethanolic extract made from fresh ginger

rhizomes (Penna et al., 2003). Local application of the extract also inhibited 48/80-induced

(but not Substance P- and bradykinin-induced) skin oedema, and intraperitoneal

Page 63: Phytochemistry and pharmacology of plants from the ginger family

39

administration one hour prior to serotonin injection significantly reduced serotonin-induced

skin oedema. It was concluded that the anti-oedematogenic activity of the ginger extract was

at least in part due to serotonin receptor antagonism.

A study investigating the anti-tumour promoting effects of [6]-gingerol found that when

applied topically to mouse skin stimulated with a tumour promoter, the compound inhibited

COX-2 expression in vivo (Kim et al., 2005b). The suggested mechanism of this effect was

blocking of the p38 mitogen-activated protein (MAP) kinase–nuclear factor kappa B (NFκB)

signalling pathway (NFκB regulates COX-2 induction).

2.3.1.3.2.4 Summary: Effects on arachidonic acid metabolism and eicosanoids

There is a substantial body of evidence from in vitro studies suggesting that ginger extracts

and a range of ginger constituents (including [6]-, [8]- and [10]-gingerol, and [6]-shogaol)

inhibit cyclooxygenase (COX). This causes a reduction in the formation of prostanoids of

the 2-series, including PGE2, which has potent pro-inflammatory properties in certain

circumstances, and TxA2, which causes platelet aggregation and serotonin release, and

contracts smooths vascular muscle. Some phenolic ginger constituents, including [8]- and

[10]-gingerol and [6]-shogaol (but not [6]-gingerol), were relatively potent inhibitors of

COX-2 in vitro.

Animal studies have confirmed the ability of ginger extracts to inhibit PGE2, thromboxane

and inflammation, and studies using essential oil suggest that compounds other than those of

gingerol type possess anti-inflammatory activity. Unlike the finding from the in vitro study

of COX-2 inhibition, [6]-gingerol was found to inhibit COX-2 when applied topically to

mouse skin. In this instance COX-2 inhibition appeared to be caused by the blocking of a

signalling pathway involving NFκB. Serotonin antagonism has also been suggested as a

possible mechanism of anti-inflammatory action.

Thus the emerging picture of ginger as an anti-inflammatory agent is consistent with that of a

typical phytomedicine: it contains multiple active compounds and as such it appears to have

multiple pharmacological targets.

Page 64: Phytochemistry and pharmacology of plants from the ginger family

40

2.3.1.3.3 Effects on cytokines and chemokines

Cytokines are immunoregulatory proteins (such as interleukins, tumour necrosis factor, and

interferons) secreted by cells, especially of the immune system. They provide

communication between macrophages and various subsets of lymphocytes; all of them act in

a non-specific manner to enhance the inflammatory response (McCance & Huether, 2002).

A hot extract of dried ginger in 50% aqueous ethanol was tested for its effects on cytokine

secretion by human peripheral blood mononuclear cells (Chang et al., 1995). The secretion

of IL-1β, IL-6 and granulocyte-macrophage colony-stimulating factor (GM-CSF) was

increased in the presence of the ginger extract at low concentrations (10-30 mg/mL), but

only after prolonged reaction time (18 or 24 hours); higher concentrations were less

effective. There was no significant effect on tumour necrosis factor-alpha (TNF-α).

Chemokines are chemotactic molecules, at least 50 of which are known to play a role in

inflammation. Joint chondrocytes and synoviocytes produce a range of chemokines, which

cause migration of leukocytes into the inflamed tissue and mediate cartilage degradation by

decreasing proteoglycan synthesis and activating matrix metalloproteinases (MMPs), which

catalyse the breakdown of cartilage (Phan et al., 2005).

An extract prepared from dried ginger rhizomes, described as having a high (sic) content of

[6]-, [8]- and [10]-gingerol, gingerdione and gingerdiol, was tested in vitro for its effects on

the expression of two chemokines, macrophage chemotactic factor (MCP-1) and interferon-γ

activated protein (IP-10), in human synoviocytes (Phan et al., 2005). These two chemokines

have been shown to be involved in the pathogenesis of both osteoarthritis and rheumatoid

arthritis by inducing the migration of leukocytes and monocytes from the blood to areas of

inflammation. MCP-1 also has been found to induce MMPs. IP-10 is a chemoattractant for

monocytes, natural killer cells and activated T-cell, in particular type 1 helper cells, causing

upregulation of interferon-γ. The experiment used two cultivated synoviocyte cell lines

obtained from patients with osteoarthritis during knee replacement surgery. Incubation with

ginger extract inhibited the secretion of both MCP-1 and IP-10 (by 40% and 54%,

respectively), but only in one of the two cell lines tested. Incidentally, the study

demonstrated a synergistic effect when extracts of Zingiber officinale and Alpinia galanga

(galangal) were combined.

Page 65: Phytochemistry and pharmacology of plants from the ginger family

41

These findings suggest that some aspects of ginger’s anti-inflammatory activity may be

mediated by effects on cyto- and chemokines.

2.3.1.3.4 Anti-cancer activity

An ethanolic ginger extract applied topically to mouse skin provided a highly significant

protective effect against the development of skin tumours, and this was associated with the

inhibition of 12-O- tetradecanoylphorbol-13-acetate (TPA)-caused induction of epidermal

ornithine decarboxylase, cyclooxygenase and lipoxygenase activities (Katiyar et al., 1996).

A subsequent study showed [6]-gingerol to have similar activity (Park et al., 1998).

A more recent study showed that topical application of [6]-gingerol inhibited COX-2

expression in mouse skin stimulated with the tumour promoter TPA (Kim et al., 2005b).

Results from this study suggested that the inhibition of COX-2 expression was the result of

the blocking of the p38 MAP kinase- NFκB signalling pathway.

A cytotoxic or cytostatic effect mediated by apoptosis was found for [6]-gingerol and [6]-

paradol in human promyelocytic leukaemia HL-60 cells (Lee & Surh, 1998), and also for

four diarylheptanoids and two shogaols (Wei et al., 2005).

2.3.1.3.5 Anti-microbial activity

Ginger extracts have demonstrated antimicrobial activity against a wide range of pathogenic

micro-organisms; these include both Gram-positive and Gram-negative bacteria and the

yeast Candida albicans, but the antimicrobial activity of ginger extracts appears to be

moderate in comparison with highly antimicrobial plant extracts.

Of particular interest is an in vitro study showing that a crude methanolic extract (MIC 6-50

µg/mL) and a gingerol-containing fraction (MIC 0.8-12.5 µg/mL) significantly inhibited the

growth of 19 strains of Helicobacter pylori, the micro-organism associated with peptic ulcer

disease as well as gastric and colon cancer (Mahady et al., 2003).

Page 66: Phytochemistry and pharmacology of plants from the ginger family

42

2.3.1.3.6 Immunomodulatory effects

Few studies have examined the potential immunomodulatory activity of ginger.

Non-specific immunity was increased in rainbow trout eating a diet containing 1% of a dried

aqueous ginger extract for three weeks (Dugenci et al., 2003). Mice fed a 50% ethanolic

ginger extract (25 mg/kg) for seven days had higher haemagglutinating antibody titre and

plaque-forming cell counts, consistent with improved humoral immunity (Puri et al., 2000).

One in vitro study found that ginger suppressed lymphocyte proliferation; this was mediated

by decreases in IL-2 and IL-10 production (Wilasrusmee et al., 2002). Another study found

an aqueous ginger extract significantly increased the production of IL-1β, IL-6 and TNF-α in

activated peritoneal mouse macrophages (Ryu & Kim, 2004). The same study found that

splenocyte proliferation and cytokine production were stimulated in a dose-dependant

manner when mice were given an aqueous ginger extract for 4 weeks (50-500 mg/kg).

2.3.1.3.7 Gastrointestinal effects

An orally administered acetone extract of ginger (75 mg/kg) as well as [6]-shogaol (2.5

mg/kg) and [6]-, [8]- and [10]-gingerol (5 mg/kg) enhanced gastric motility (charcoal meal)

in mice (Yamahara et al., 1990). Both acetone and 50% ethanolic extracts (100-500 mg/kg)

and ginger juice (2-4 mL/kg) reversed cisplatin-induced delay in gastric emptying in rats

when given orally (Sharma & Gupta, 1998). The effect on gastric motility may at least in

part explain the anti-emetic properties of ginger.

Both ginger extracts and several ginger compounds including 6-gingerol have demonstrated

ulcer-protective activity in rats (Agrawal et al., 2000; Sertie et al., 1992; Yamahara et al.,

1988; Yoshikawa et al., 1992). Ginger stimulated bile secretion in rats, with [6]- and [10]-

gingerol being chiefly responsible for this activity (Yamahara et al., 1985), and ginger has

also been shown to stimulate intestinal lipase, trypsin, chymotrypsin, amylase, sucrase and

maltase activity when fed to rats for an 8-week period (Platel & Srinivasan, 1996; Platel &

Srinivasan, 2000). These findings support the traditional use of ginger as a digestive

stimulant (Mills & Bone, 2000). Ginger may also increase the conversion of cholesterol to

Page 67: Phytochemistry and pharmacology of plants from the ginger family

43

bile acids by increasing the activity of hepatic cholesterol-7-α-hydroxylase, the rate-limiting

enzyme of bile acid biosynthesis (Srinivasan & Sambaiah, 1991).

2.3.1.3.8 Metabolic effects

Significant hypoglycaemic activity in rabbits was produced by an ethanolic Soxhlet extract

of ginger (Mascolo et al., 1989), and two animal studies have suggested that ginger may

improve insulin sensitivity (Goyal & Kadnur, 2006; Kadnur & Goyal, 2005).

2.3.1.3.9 Anti-atherosclerotic effects

A number of animal studies have demonstrated hypocholesterolemic action of ginger and

ginger extracts. These studies have shown decreased levels of total cholesterol, low-density

lipoprotein (LDL)-, very low-density lipoprotein (VLDL)-cholesterol and triglycerides, and

increases in high-density lipoprotein (HDL)-cholesterol (Ahmed & Sharma, 1997; Bhandari

et al., 1998; Fuhrman et al., 2000; Giri et al., 1984; Murugaiah et al., 1999).

In a more recent study, air-dried ginger powder (100 mg/kg orally daily) fed to rabbits with

experimentally induced atherosclerosis for 75 days inhibited atherosclerotic changes in the

aorta and coronary arteries by about 50% (Verma et al., 2004). In this study the ginger

treatment did not cause any significant lowering of serum lipids, but lipid peroxidation was

decreased and fibrinolytic activity increased.

The effects of an aqueous ginger extract, administered orally or intraperitoneally to rats for 4

weeks, on serum cholesterol and triglyceride levels and on PGE2 and TxB2 have been

reported (Thomson et al., 2002). At 500 mg/kg both modes of administration led to a

significant reduction in serum cholesterol levels, whereas only intraperitoneal administration

led to a reduction with a daily dose of 50 mg/kg. Neither dose caused significant changes to

serum triglyceride levels. The higher dose administered orally significantly lowered serum

PGE2 and TxB2 levels, and the lower dose also reduced PGE2 levels.

Page 68: Phytochemistry and pharmacology of plants from the ginger family

44

It is evident from this that ginger has demonstrated considerable potential as an anti-

atherosclerotic agent in animal studies, but as yet this promise has not been confirmed in

human trials.

2.3.1.3.10 Other cardiovascular effects

Ginger extracts as well as [6]- and [8]-gingerol have been shown to modulate eicosanoid

responses in smooth vascular muscles ex vivo (Hata et al., 1998; Kimura et al., 1989; Pancho

et al., 1989).

[6]- and [8]-gingerol and related analogues inhibited arachidonic acid-induced serotonin

release by human platelets in a dose range similar to the effective dose aspirin (IC50 values

between 45 and 83 μM), and the same compounds were also effective inhibitors of

arachidonic acid-induced platelet aggregation (Koo et al., 2001). COX inhibition was

demonstrated in RBL-2H3 cells, suggesting this might be the underlying mechanism.

An early study found a dose-dependent positive inotropic action of [6]-, [8]- and [10]-

gingerol on isolated guinea pig left atria (Shoji et al., 1982), and ‘gingerol’ stimulated the

Ca2+-pumping ATPase activity of fragmented sarcoplasmic reticulum prepared from

mammalian skeletal and cardiac muscle (Kobayashi et al., 1987).

[6]-Gingerol and [6]-shogaol lowered systemic blood pressure in anaesthetised rats at doses

of 10-100 μg/kg and caused bradycardia when administered intravenously (Suekawa et al.,

1984; Suekawa et al., 1986).

In a recent study a crude extract (70% aqueous methanol) of fresh ginger induced a dose-

dependent fall in arterial blood pressure of anaesthetised rats; this effect was shown to be

mediated through blockade of voltage-dependent calcium channels (Ghayur & Gilani, 2005).

2.3.1.3.11 Analgesic effects

[6]-gingerol had analgesic effects in mice in both the acid-induced writhing test and the

formalin test, suggesting the analgesic activity resulted from peripheral and possible anti-

inflammatory action (Young et al., 2005). [6]-shogaol has also been shown to inhibit acetic

Page 69: Phytochemistry and pharmacology of plants from the ginger family

45

acid-induced writhing in mice and to elevate the nociceptive threshold of the yeast-inflamed

paw (Suekawa et al., 1984). Experiments carried out by Onogi and co-workers suggested

that [6]-shogaol inhibits the release of Substance P by stimulation of the primary afferents

from their central terminal and hence shares this site of action with capsaicin (Onogi et al.,

1992).

2.3.1.3.12 Antipyretic activity

A Soxhlet extract of ginger in 80% ethanol reduced yeast-induced fever in rats by 38% when

administered orally (100 mg/kg) (Mascolo et al., 1989). This was comparable to the anti-

pyretic effect of acetylsalicylic acid at the same dose. The ginger extract did not affect the

temperature of normothermic rats. This anti-pyretic activity may be mediated by COX

inhibition.

2.3.1.4 Safety data

2.3.1.4.1 Acute toxicity

A concentrated Soxhlet extract of ginger in 80% ethanol was well tolerated orally in mice at

doses up to 2.5 g/kg, but doses of 3.0 and 3.5 g/kg caused 10-30% mortality. Death was

caused by involuntary contractions of skeletal muscle; other symptoms included

gastrointestinal spasm, hypothermia, diarrhoea and anorexia (Mascolo et al., 1989).

Another study found that a hydroethanolic ginger extract was nontoxic in mice when

administered by oral gavage up to a dose of 1.5 g/kg body weight (Jagetia et al., 2004).

2.3.1.4.2 Teratogenicity and embryotoxicity

Pregnant rats administered 20 g/L or 50 g/L ginger tea via their drinking water from

gestation day 6 to 15 showed no gross malformation on fetuses (Wilkinson, 2000).

However, embryonic loss in the ginger-treated group was twice that of controls. Surviving

Page 70: Phytochemistry and pharmacology of plants from the ginger family

46

fetuses in the ginger group were significantly heavier and showed more advanced skeletal

development compared with controls.

A patented ginger extract was tested for teratogenic potential in pregnant rats (Weidner &

Sigwart, 2001). The extract caused neither maternal nor developmental toxicity at daily

doses of up to 1 g/kg body weight.

2.3.1.4.3 Mutagenicity

[6]-gingerol and to a far lesser extent [6]-shogaol were shown to have mutagenic properties

in an assay using Escherichia coli Hs30 as an indicator strain of mutagenesis (Nakamura &

Yamamoto, 1983). Despite this finding, ginger is not considered a mutagenic substance,

presumably due to its long history of safe use.

2.3.1.5 Clinical studies of ginger

Data from clinical trials support the use of ginger and ginger preparations in motion sickness

and seasickness, morning sickness, hyperemesis gravidarum, postoperative nausea and

osteoarthritis, although not all clinical trials have produced positive outcomes. The clinical

efficacy trials on ginger published prior to 2008 are summarised in Appendix A.

2.3.1.5.1 Clinical studies of ginger in arthritis

Few clinical studies have examined the efficacy of ginger in osteoarthritis. Preceding proper

clinical trials, Srivatava and Mustafa published in 1989 a case series of 7 patients suffering

from rheumatoid arthritis who had reported improvements in their condition following the

consumption of ginger (Srivastava & Mustafa, 1989). The same investigators followed up

these observations with a questionnaire-based survey of people who were self-medicating

ginger for their condition (Srivastava & Mustafa, 1992). This survey provided data from a

total of 56 patients, including the 7 people whose cases had been published in the first report.

Of the 56 patients, 28 suffered from rheumatoid arthritis, 18 from osteoarthritis and 10 from

Page 71: Phytochemistry and pharmacology of plants from the ginger family

47

muscular discomfort. The majority of patients reported marked relief in symptoms of pain

and swelling as a result of ginger consumption.

The results obtained by Srivastava and Mustafa in their survey were promising. However,

the fact that the data came from an uncontrolled study with self-selected subjects meant that

these results offered little in terms of proper evidence for the efficacy of ginger in arthritic

and rheumatic conditions.

The first proper clinical trial to examine the usefulness of ginger in osteoarthritis was

conducted by a Danish group and published in 2000 (Bliddal et al., 2000a). This

randomised, placebo-controlled, cross-over study compared the effects of a standardised

ginger extract with those of ibuprofen in osteoarthritis. Fifty-six out-patients (41 women, 15

men) completed the study. Of these, 36 had osteoarthritis of the knee and 20 osteoarthritis of

the hip. Mean duration of osteoarthritis was 7.7 years (range 1-30 years). One week prior to

randomisation, treatment with analgesics and NSAIDs was discontinued, but acetaminophen

(paracetamol) was allowed as a rescue drug for pain relief throughout the study, up to a

maximum dose of 3 g daily. Subjects were randomised to three treatment periods of three

weeks each, during which they received (in different order) capsules containing either 170

mg of a standardised ginger extract t.i.d., ibuprofen 400 mg t.i.d., or placebo t.i.d. The ginger

extract had a standardised content of hydroxy-methoxy-phenyl compounds, but no further

details were provided, making it impossible to estimate the dose on a fresh or dried

equivalent basis. There were no wash-out periods between the three treatments. The primary

outcome variable was pain assessed using a visual analogue scale (VAS); secondary

outcomes were the Lequesne-index (a questionnaire-based disability score) for either hip or

knee, and range of motion.

The results produced a ranking in terms of efficacy on pain level and function with ibuprofen

being more effective than the ginger and placebo. The same ranking applied to the

consumption of the rescue medication, with ibuprofen treatment being associated with the

lowest consumption of the rescue medication. No difference between the ginger extract and

placebo was found in a test for multiple comparisons. However, explorative statistical testing

of the first period of treatment (before cross-over) found a significantly better effect of both

ginger and ibuprofen compared with placebo (p<0.05). This finding is particularly

interesting, because the design of this cross-over study was marred by the lack of wash-out

periods between treatments. The lack of wash-out periods and the short duration (three

Page 72: Phytochemistry and pharmacology of plants from the ginger family

48

weeks) of each treatment means that the trial carried out by Bliddal's group failed to

convincingly establish whether or not ginger might be useful in the treatment of

osteoarthritis.

In 2003, Wigler and co-workers published the results of small clinical trial of ginger in

osteoarthritis of the knee (Wigler et al., 2003). This was a randomised, double-blind,

placebo-controlled study of 6 months’ duration. Twenty-nine subjects were divided into

verum and placebo groups, which were crossed over after 12 weeks. The intervention was a

ginger extract produced by supercritical carbon dioxide extraction and formulated into an

enteric coated capsule containing 250 mg extract including 10 mg ‘gingerol’. The capsule

was designed to release 20% of its content under the acidic conditions of the stomach and the

remainder in the intestine. Subjects took one capsule (or identical placebo) four times daily

(equivalent to 40 mg ‘gingerol’ daily). Primary outcome measures were pain on movement

and handicap as assessed by subjects on a visual analogue scale based on the Western

Ontario and McMaster Osteoarthritis Index (WOMAC). Both groups showed a significant

decrease in both outcome measures after 12 weeks, but there was no significant difference

between the groups. After 24 weeks, however, the difference between the ginger and

placebo groups was highly significant for both outcomes in favour of the ginger intervention

(p>0.01).

Although the study conducted by Wigler and colleagues also did not include a wash-out

phase between treatments, the duration (12 weeks) of each treatment means that this was less

likely to invalidate the findings. In fact, the group that started on the active treatment

continued to improve for 2 weeks after being switched to placebo, suggesting the effects of

the ginger treatment continued after the treatment was withdrawn.

A third larger trial (n=247), rather misleadingly published in 2001 under the title ‘Effects of

a ginger extract on knee pain in patients with osteoarthritis’, is often cited as evidence for the

efficacy of ginger in osteoarthritis of the knee (Altman & Marcussen, 2001). This assertion

is not strictly valid, however, as the intervention employed in this trial was a patented extract

containing not only Zingiber officinale but also Alpinia galanga (galangal). This

randomised, double-blind, placebo-controlled, parallel-group, 6-week study found a

statistically significant effect of the extract in reducing symptoms of osteoarthritis of the

knee.

Page 73: Phytochemistry and pharmacology of plants from the ginger family

49

The reports of the only two randomised controlled trials of ginger extracts as the sole active

in the treatment of osteoarthritis (Bliddal et al., 2000b; Wigler et al., 2003) were scored

against the CONSORT (Consolidated Standards of Reporting Trials) (Moher et al., 2001)

and the elaborated CONSORT for herbal interventions (Gagnier et al., 2006) (see Appendix

B). Each item was given a score of 0, 0.5 or 1, depending on whether the information was

provided not at all or to an unsatisfactory extent (0), to some extent (0.5), or to a satisfactory

or mostly satisfactory extent (1) in the trial report. With total scores of 17.5 (Bliddal) and

17.0 (Wigler) out of a possible 27, the reporting of the two trials was of equal and reasonable

quality, although neither provided qualitative chemical information (such as an HPLC

chromatogram) about the intervention used, and the study by Bliddal and colleagues did not

use the binomial for ginger to unambiguously define the material taxonomically.

The results of the only two randomised controlled trials of ginger in osteoarthritis are

encouraging, if not conclusive. The results of the most recent trial suggest that treatment for

up to 24 weeks is required for the full benefits to manifest. A large, rigorous trial with a

chemically well-defined intervention is now warranted to more conclusively establish the

role (if any) of ginger in the treatment of osteoarthritis.

2.3.2 Other Zingiber species

Other Zingiber species included in this work are reviewed below.

2.3.2.1 Thai ginger (Zingiber montanum)

Zingiber montanum (J. König) Theilade is more widely cited in the literature under the

synonyms Z. cassumunar Roxb. and Z. purpureum Roscoe.

This species is used in the treatment of asthma in traditional Thai medicine and has been

shown to have an anti-histamine effect in asthmatic individuals (500 mg orally) (Piromrat et

al., 1986). A methanolic extract inhibited PGE2 production by in human promonocytic U937

cells (IC50 = 7.7 µg/mL) (Jiang et al., 2006b). The rhizome contains a number of

phenylbutanoid compounds (Han et al., 2004; Jitoe et al., 1993; Lu et al., 2005; Masuda &

Jitoe, 1995), some of which have been shown to have potent anti-inflammatory activity both

Page 74: Phytochemistry and pharmacology of plants from the ginger family

50

in vitro and in vivo through inhibition of COX and lipoxygenase (LOX) pathways

(Jeenapongsa et al., 2003; Panthong et al., 1997), including inhibition of COX-2 in

lipopolysaccharide (LPS)-stimulated RAW 264.7 cells (Han et al., 2005). One

phenylbutenoid dimer had anti-proliferative activity in several human cancer cell lines (Lee

et al., 2007). Also present are curcuminoids (cassumunin A and B, cassumunarin A, B and

C; Fig. 2-10) with potent anti-oxidant activity (Masuda & Jitoe, 1994; Nagano et al., 1997).

The cassumunarins were also shown to have anti-inflammatory activity in vivo (Masuda &

Jitoe, 1994). The sesquiterpene zerumbone has antifungal activity (Kishore & Dwivedi,

1992).

Extracts of Z. montanum have demonstrated in vitro antioxidant activity (Chirangini et al.,

2004; Habsah et al., 2000a; Jitoe et al., 1992), anti-allergic activity (Tewtrakul &

Subhadhirasakul, 2007) and anti-tumour promoter activity (Vimala et al., 1999). Extracts

also inhibited P-glycoprotein in human uterine sarcoma cells (Go et al., 2004) and

significantly inhibited CYP3A4 in human liver microsomes (Subehan et al., 2006). In

animals, extracts have shown anti-inflammatory and analgesic effects (Ozaki et al., 1991;

Pongprayoon et al., 1997).

H3CO

RH3CO HO

OCH3

O O

OCH3

OH

cassumunin A: R = Hcassumunin B: R = -OCH3

H3CO

OCH3

OCH3

OH

O O

OH

OCH3

cassumunarin A

Fig. 2-10. Curcuminoids from Zingiber montanum.

Page 75: Phytochemistry and pharmacology of plants from the ginger family

51

The essential oil, which primarily consists of monoterpenoids and phenylbutanoids (Taroeno

et al., 1991), was active against a wide range of Gram-positive and Gram-negative bacteria,

dermatophytes and yeasts (Pithayanukul et al., 2007).

2.3.2.2 Zingiber ottensii

Zingiber ottensii Valeton is another species native to South-east Asia. The rhizome is

reportedly used to treat lumbago and is an ingredient in a sedative lotion used in the

treatment of convulsions (Sirirugsa, 1999).

The rhizome contains a number of terpenoids and diarylheptanoids (Akiyama et al., 2006;

Sirat, 1994). Zerumbone is the major constituent of the essential oil (Sirat & Nordin, 1994),

which had moderate cytotoxicity in the brine shrimp assay (Thubthimthed et al., 2005).

Extracts of the rhizome have been found to possess antimicrobial (Azmi Muda et al., 2002;

Mohtar et al., 1998) and anti-oxidant (Habsah et al., 2000b) activities. No other information

about the pharmacological activity of this species was located.

2.3.2.3 Beehive ginger (Zingiber spectabile)

Zingiber spectabile Griff. is native to Thailand and the Malayan peninsula and is widely used

as an ornamental in tropical and subtropical areas. An infusion of the leaves has been

reported to have been used to treat infected eyelids (Sirirugsa, 1999).

The rhizome contains an essential oil reported to contain terpinen-4-ol (23.7%), labda-8

(17),12-diene-15,16-dial (24.3%), α-terpineol (13.1%), and β-pinene (10.3%) as major

constituents (Sirat & Leh, 2001). A methanolic extract inhibited PGE2 production in human

promonocytic U937 cells (IC50 = 1.2 µg/mL) (Jiang et al., 2006b), and a solvent extract

demonstrated antimicrobial activity (Ghosh et al., 2000). No other information pertaining to

the chemistry and pharmacology of this species was located.

Page 76: Phytochemistry and pharmacology of plants from the ginger family

52

2.4 Genus Curcuma L.

The genus Curcuma contains approximately 40 mostly tropical Asian species. Best known is

turmeric (C. longa L.), a cultigen of likely Indian origin, which is widely used as a spice and

as an orange and yellow dye (Mabberley, 1997). Several other species are also used for

culinary purposes including mango ginger (C. amada Roxb.), Bombay or Indian arrowroot

(C. angustifolia Roxb.) and zedoary (C. zedoaria (Christm.) Roscoe, syn. C. zerumbet)

(Mabberley, 1997).

In Australia, the genus is represented by a single native species, C. australasica J. D. Hook,

which is sometimes, rather misleadingly, called Cape York lily, and finds use as an

ornamental plant. Its natural habitat is shady rainforest margins in tropical parts of

Queensland and the Northern Territory, and it also occurs in New Guinea (Smith, 1987).

More than a dozen species of Curcuma have been used in traditional systems of medicine

(Johnson, 1999).

2.4.1 Curcuma longa

Turmeric (C. longa L., syn. C. domestica Val.) and its major active constituent curcumin

have been the subject of hundreds of scientific studies, the majority of which have been

published since 2000. The interest has focused on curcumin and its anti-oxidant, anti-

carcinogenic and anti-inflammatory actions.

2.4.1.1 Chemistry

The major constituents of interest in turmeric are the coloured diarylheptanoids known as

curcuminoids, which constitute around 5% of the dried rhizome. The major of these is the

unsaturated β-diketone curcumin (diferuloylmethane), which together with

desmethoxycurcumin and bisdesmethoxycurcumin make up 50-60% of the curcuminoids

present in the rhizome (Fig. 2-11). Dihydrocurcumin has also been reported (Evans, 2002;

WHO, 1999).

Page 77: Phytochemistry and pharmacology of plants from the ginger family

53

Turmeric also contains 5-6% volatile oil made up of mono- and sesquiterpenes including

zingiberene, curcumene, α- and β-turmerone (Evans, 2002; WHO, 1999).

O O

HO OH

H3CO OCH3

Curcumin

O O

HO OH

H3CO OCH3

H

bis-keto form(pH 3-7)

enolate form(pH >8)

O O

HO OH

OCH3

Demethoxycurcumin

O O

HO OHBisdemethoxycurcumin

Fig. 2-11. Structural diagrams of the major curcuminoids in turmeric rhizome: curcumin, demethoxycurcumin and bisdemethoxycurcumin. Curcumin exists in a pH-dependent equilibrium between its bis-keto and enolate forms. (After Higdon, 2007; Sharma et al., 2005).

2.4.1.2 Pharmacology

The biological and pharmacological activities of curcumin have been the subject of many

reviews. These have included general reviews (Bengmark, 2006; Maheshwari et al., 2006;

Sharma et al., 2005) and reviews focussing on anti-carcinogenic and chemopreventive

Page 78: Phytochemistry and pharmacology of plants from the ginger family

54

activities (Duvoix et al., 2005; Karunagaran et al., 2005; Leu & Maa, 2002; Lin, 2004;

Narayan, 2004; Thangapazham et al., 2006), mechanism of action (Joe et al., 2004), and the

potential role of curcumin in the treatment of Alzheimer’s disease (Ringman et al., 2005).

2.4.1.2.1 Pharmacokinetics of curcumin

In rodents, curcumin has low oral bioavailability and may undergo partial intestinal

metabolism. Following absorption, curcumin undergoes extensive first-pass metabolism and

excretion with the bile (Sharma et al., 2005). The major curcumin metabolites in

suspensions of both human and rat hepatocytes were hexahydrocurcumin and

hexahydrocurcuminol and these were less potent inhibitors of PGE2 production in human

colonic epithelial cells than curcumin itself (Ireson et al., 2001).

In humans, curcumin also has low oral bioavailability. In one study, a dose of up to 200 mg

failed to produce serum levels detectable at the level of detection (0.63 ng/mL) (Ruffin et al.,

2003). Administration of 2 g turmeric powder to fasting volunteers resulted in curcumin

plasma concentrations below 10 ng/mL, but this was increased 20-fold when piperine (the

pungent alkaloid from Piper species) was co-administered (Shoba et al., 1998). Curcumin

sulfate and curcumin glucuronide have been identified as metabolites in human urine

(Sharma et al., 2004) and in intestinal tissues of patients with colorectal cancer (Garcea et

al., 2005).

2.4.1.2.2 Pharmacodynamics of curcumin

2.4.1.2.2.1 Anti-oxidant activity

Numerous studies have shown curcumin to possess potent anti-oxidant activity both in vitro

and in vivo. Mechanistic studies have found that curcumin is a phenolic chain-breaking anti-

oxidant, which donates hydrogen atoms from its phenolic groups (Barclay et al., 2000;

Priyadarsini et al., 2003; Sun et al., 2002). The resulting phenoxyl radical can be repaired by

water-soluble anti-oxidants such as vitamin C (Jovanovic et al., 2001).

Page 79: Phytochemistry and pharmacology of plants from the ginger family

55

Curcumin can act directly as a scavenger of free radicals such as singlet oxygen (Das & Das,

2002), superoxide (Biswas et al., 2005; Mishra et al., 2004) and hydroxyl radical (Biswas et

al., 2005), but also exerts its anti-oxidant activity by enhancing endogenous defenses against

oxidative damage. Cultured alveolar epithelial cells (A549) showed increased glutathione

expression compared with controls (Biswas et al., 2005), and mice receiving dietary

curcumin (2%) had significantly increased activities of glutathione peroxidase, glutathione

reductase, glucose 6-phosphate dehydrogenase and catalase in the kidneys compared with

controls (Iqbal et al., 2003). This induction of detoxification enzymes may play a role in the

chemopreventive effects of curcumin.

The anti-oxidant activity of turmeric and curcumin has been demonstrated in a variety of in

vitro assay systems, including oxygen radical absorbance capacity (ORAC), radical

scavenging assays using 1,1-diphenyl-2-picryl hydrazyl (DPPH) and 2,2'-azobis-3-

ethylbenzthiazoline-6-sulfonic acid (ABTS), the ferric reducing antioxidant power (FRAP)

assay (Tilak et al., 2004), and the Trolox equivalent antioxidant capacity (TEAC) assay

(Betancor-Fernandez et al., 2003) .

Radical scavenging and anti-oxidant activities have been confirmed in vivo in a variety of

animal models. Some of these studies are briefly summarised here.

Curcumin was shown to protect rat brain homogenate against lipid peroxidation caused by

lead and cadmium (Daniel et al., 2004), and pre-treatment with curcumin significantly

reduced or abolished cadmium-induced hepatic oxidative damage in rodents (Eybl et al.,

2004; Eybl et al., 2006). Curcumin protected the rat forebrain against cerebral

ischaemia/reperfusion oxidative injury (Ghoneim et al., 2002).

A hydroethanolic turmeric extract fed to rabbits (1.66 mg/kg) receiving a high-fat,

atherogenic diet for 30 days significantly reduced the oxidative damage to erythrocyte and

liver microsome membranes (Mesa et al., 2003), and dietary curcumin (0.2%) reduced lipid

peroxidation and the depletion of intracelluluar anti-oxidants (thiols and glutathione) in

erythrocytes of rats fed a high-fat diet for 8 weeks (Kempaiah & Srinivasan, 2004). Mice fed

a turmeric extract (1% of diet) for just one week had reduced levels of erythrocyte

phospholipid hydroperoxides and liver triacylglycerol levels about half those of controls

(Asai et al., 1999).

Page 80: Phytochemistry and pharmacology of plants from the ginger family

56

Curcumin (200 mg/kg) significantly reduced lipid peroxidation in chronic streptozotocin-

diabetic rats, but had no significant effect on superoxide dismutase, catalase, or reduced

glutathione levels (Majithiya & Balaraman, 2005). The same curcumin dose administered

for 8 weeks to diabetic rats had a preventive effects on the cross-linking of collagen and the

formation of advanced glycation end products, a process linked to oxidative stress (Sajithlal

et al., 1998). Low dietary levels of curcumin (0.002% and 0.01%) delayed the progression

of diabetic cataract in the lens of diabetic rats (Kumar et al., 2005).

Rats subjected to oxidative stress caused by dietary ethanol or oxidised polyunsaturated fatty

acids had elevated liver marker enzymes and lipid peroxidative indices, but curcumin

abrogated this effect, suggesting it potentially could provide some protection against

alcoholic liver disease (Rukkumani et al., 2004). Lipid peroxidation was significantly

modulated and anti-oxidant status improved in rats suffering lung toxicity caused by

subcutaneous injection of nicotine (Kalpana & Menon, 2004).

Curcumin has also been shown to protect rodents against drug-induced oxidative organ

injury. Gentamycin-induced renal oxidative damage in rats (Farombi & Ekor, 2006),

nephro- and cardiotoxicity in rats caused by adriamycin (Venkatesan, 1998; Venkatesan et

al., 2000), and inflammatory and oxidative lung injury caused by bleomycin in rats

(Venkatesan et al., 1997) were all reduced by curcumin. Curcumin also reduced cisplatin-

induced clastogenesis in rats through it radical scavenging activity (Antunes et al., 2000), but

did not protect against cisplatin-induced lipid peroxidation and nephrotoxicity (Antunes et

al., 2001).

Several studies have found that curcumin can also promote the generation of reactive oxygen

species (ROS) under some circumstances (Ahsan et al., 1999; Atsumi et al., 2005a; Atsumi

et al., 2005b). One such study found that in human myeloid leukemia (HL-60) cells, low

curcumin concentration reduced ROS generation, while high concentrations promoted it

(Chen et al., 2005). A recent study using myelomonocytic U937 cells found that the anti-

oxidative action of curcumin was preceded by a time and dose dependent oxidative stimulus

and suggested that excessive curcumin concentrations may be harmful to cells (Strasser et

al., 2005).

Page 81: Phytochemistry and pharmacology of plants from the ginger family

57

In a placebo-controlled study involving 20 subjects with tropical pancreatitis, 500 mg of

curcumin (with 5 mg piperine to enhance the bioavailability of curcumin) reduced markers

of erythrocyte lipid peroxidation, but did not improve pain (Durgaprasad et al., 2005).

2.4.1.2.2.2 Anti-inflammatory activity

Numerous studies have indicated that the key to the anti-inflammatory activity of curcumin

is its ability to inhibit the activation of the transcription factor NFκB (Brennan & O'Neill,

1998; Funk et al., 2006; Jobin et al., 1999; Jung et al., 2006; Kang et al., 2004; Kumar et al.,

1998; Moon et al., 2005; Singh & Aggarwal, 1995; Strasser et al., 2005). NFκB plays a

critical role in the transcriptional regulation of pro-inflammatory gene expression in various

cells. Jobin and colleagues demonstrated that curcumin blocks a signal upstream of NFκB-

inducing kinase and IκB kinase, both of which are involved in the activation of NFκB itself

(Jobin et al., 1999).

Curcumin has also been found to inhibit the production of inflammatory cytokines involved

in NFκB activation. In human peripheral blood monocytes and alveolar macrophages

stimulated with phorbol ester or lipopolysaccharide (LPS), curcumin inhibited IL-8,

monocyte inflammatory protein-1 (MIP-1α), monocyte chemotactic protein-1 (MCP-1), IL-

1β, and TNF-α (Abe et al., 1999). At a concentration of 5 μM, curcumin inhibited LPS-

induced production of TNF and IL-1 by a human monocytic macrophage cell line (Chan,

1995), and in rats with adjuvant inflammation curcumin reduced levels of IL-1β and C-

reactive protein (Banerjee et al., 2003). Curcumin also inhibited IL-1β-mediated

intracellular adhesion molecule-1 (ICAM-1) and IL-8 gene expression in several epithelial

intestinal cell lines, resulting in the inhibition of NFκB activation (Jobin et al., 1999).

Crude organic turmeric extracts inhibited LPS-induced TNF-α and PGE2 production in HL-

60 cells with IC50 values of 15.2 μg/mL and 0.92 μg/mL, respectively (Lantz et al., 2005).

An important aspect of the inflammatory response is the migration of endothelial cells from

the vascular system into the tissues, following their recruitment of leukocytes. Treatment of

endothelial cells with TNF caused monocytes to adhere as a result of NFκB-dependent

expression of ICAM-1, vascular cell adhesion molecule-1 (VCAM-1), and endothelial

leukocyte adhesion molecule-1 (ELAM-1); pre-treatment with curcumin completely blocked

Page 82: Phytochemistry and pharmacology of plants from the ginger family

58

adhesion (Kumar et al., 1998). These findings suggest that inhibition of leukocyte

recruitment by endothelial cells (as a consequence of NFκB inhibition) may contribute to the

anti-inflammatory activity of curcumin.

The inhibition of NFκB activation also leads to subsequent reduction in the expression of

COX-2 and other inflammatory mediators whose genes are regulated by NFκB, and several

studies have shown curcumin and turmeric to reduce PGE2 levels (Funk et al., 2006; Hong et

al., 2004; Huang et al., 1991; Ireson et al., 2001; Lantz et al., 2005). Although one study

reported the direct inhibition of prostaglandin E synthase-1 and COX-2 (Moon et al., 2005),

other studies have suggested that PGE2 levels are reduced not through inhibition of COX

catalytic activity, but as a result of upstream inhibition of COX-2 expression (Funk et al.,

2006; Hong et al., 2004). Curcumin strongly inhibited COX-2 mRNA and protein

expression in UVB-irradiated keratinocytes; the mechanism appeared to involve the

suppression of p38 mitogen-activated protein kinase (MAPK) and c-Jun N-terminal kinase

(JNK) activities (Cho et al., 2005).

In addition to inhibiting COX-2 expression, curcumin also affects arachidonic acid

metabolism by blocking the phosphorylation and thus the activation of PLA2, the enzyme

controlling the release of arachidonic acid from membranes, and by inhibiting the activity of

5-LOX (Hong et al., 2004).

Curcumin inhibits DNA binding of another transcription factor, activator protein-1 (AP-1),

which together with NFκB controls a key mediator of coagulation, endothelial tissue factor

(Bierhaus et al., 1997). Curcumin also inhibited DNA binding of AP-1 in microglial cells

(Kang et al., 2004), but did not reduce activation of AP-1 in mice suffering non-alcoholic

steatohepatitis, although NFκB activation was prevented (Leclercq et al., 2004).

Curcumin also inhibits the expression of nitric oxide (NO), which along with its derivatives

acts as an inflammatory mediator. The derivatives nitrite and peroxynitrite are capable of

damaging DNA and are therefore carcinogenic. The inhibition of nitric oxide production by

stimulated RAW 264.7 macrophages when cells were treated with curcumin (10 μM) was

shown to result from reduced expression of inducible nitric oxide synthase (iNOS) (Brouet &

Ohshima, 1995). Similar inhibition of nitric oxide production has been observed in mouse

peritoneal cells (Chan et al., 1995), while inhibition of both NO production and expression

of iNOS was demonstrated in rat primary microglial cells (Jung et al., 2006). In the latter

Page 83: Phytochemistry and pharmacology of plants from the ginger family

59

study, several signaling pathways (incl. NFκB, p38 MAPK and JNK) were found to be

involved in curcumin’s inhibition of NO, suggesting that the compound has multiple

molecular targets.

In an in vivo study with two oral treatments of curcumin (92 ng/g body weight), the

expression of hepatic iNOS RNA was reduced by 50-70% in mice injected with LPS (Chan

et al., 1998). Interestingly, inhibition was seen only when the curcumin was administered

after fasting, suggesting that concurrent food intake interfered with the bioavailability of

curcumin. This study is also important because is showed that curcumin can exert important

pharmacological activity when administered orally in very low doses, at least in mice.

In rodent brain microglial cells, curcumin reduced the inflammatory response (induction of

COX-2 and iNOS in response to stimuli) by inhibiting Janus kinase-STAT signalling (Kim et

al., 2003).

2.4.1.2.2.3 Anti-carcinogenic and anti-tumour activity

The chemopreventive and anti-tumour properties of curcumin (and turmeric) have been the

subject of numerous studies and, as previously mentioned, a number of reviews (Duvoix et

al., 2005; Karunagaran et al., 2005; Leu & Maa, 2002; Lin, 2004; Narayan, 2004;

Thangapazham et al., 2006).

Anti-tumour activity of curcumin has been demonstrated in many human cancer cells lines,

including colorectal (Goel et al., 2001; Lev-Ari et al., 2006), renal (Jung et al., 2005), lung

(Lee et al., 2005a; Shishodia et al., 2003), oral (Sharma et al., 2006), breast epithelial (Lee et

al., 2005b), pancreatic (Li et al., 2005) and bladder cells (Park et al., 2006). This anti-

tumour activity has been confirmed in several animal studies (Pal et al., 2005; Rao et al.,

1995; Volate et al., 2005).

Few human studies have been conducted in this area; those that have been carried out have

mostly been small and explorative in nature. Colorectal cancer patients who ingested 3.6 g

curcumin daily for 7 days had significantly fewer nucleotide adducts in malignant tissue

(p<0.05) but unchanged levels of COX-2 (Garcea et al., 2005). Both an ethanolic turmeric

extract and a curcumin-containing ointment was reported to have provided remarkable

symptomatic relief (smell, itch) in 62 patients with external cancers (Kuttan et al., 1987).

Page 84: Phytochemistry and pharmacology of plants from the ginger family

60

Numerous studies on curcumin suggest that it exerts its anti-cancer activities through a

variety of mechanisms at the initiation, promotion and progression stages of carcinogenesis

(Chauhan, 2002; Kawamori et al., 1999).

A major mechanism appears to be the induction of apoptosis (Anto et al., 2002; Atsumi et

al., 2005b; Divya & Pillai, 2006; Li et al., 2005). This occurs via the mitochondrial

pathway, where curcumin reduces mitochondrial membrane potential (Banjerdpongchai &

Wilairat, 2005; Karunagaran et al., 2005; Volate et al., 2005). Curcumin-induced apoptosis

in human renal cancer cells was found to be associated with upregulation of death receptor 5

expression (Jung et al., 2005), and in human colon cancer cell lines the induction of

apoptosis by curcumin was accompanied by down-regulation of COX-2 and reduced PGE2

synthesis (Lev-Ari et al., 2006).

Elevated levels of COX-2 are commonly seen in various types of malignant cells, and COX-

2 is therefore recognized as a molecular target in chemoprevention (Lee et al., 2005b).

COX-2 has also been implicated in tumour growth and invasiveness (Li et al., 2005). In

particular, COX-2 is selectively over-expressed in colon tumours and believed to play an

important role in colon carcinogenesis (Plummer et al., 1999). Many studies have shown

that curcumin down-regulates COX-2 (and subsequent PGE2 production) in cancer cells

(Atsumi et al., 2005b; Divya & Pillai, 2006; Goel et al., 2001; Lee et al., 2005a; Lee et al.,

2005b; Lev-Ari et al., 2006; Plummer et al., 1999; Sharma et al., 2006; Tunstall et al., 2006;

Yoysungnoen et al., 2006; Zhang et al., 1999). As outlined in the review of curcumin’s anti-

inflammatory activity above, it appears that the inhibition of the transcription factor NFκB is

the central mechanism by which curcumin inhibits the expression of COX-2 in cancer cells

(Divya & Pillai, 2006; Lee et al., 2005a; Lee et al., 2005b; Li et al., 2005; Plummer et al.,

1999; Sharma et al., 2006; Shishodia et al., 2003).

Curcumin has also demonstrated other anti-cancer mechanisms. These include protection of

DNA against genotoxic agents (Polasa et al., 2004), cytotoxicity (Chen et al., 1999; Divya &

Pillai, 2006; Su et al., 2006), causing cell cycle arrest (at G2/M) (Chauhan, 2002; Chen et

al., 1999; Park et al., 2006), inhibition of viral oncogenes (Divya & Pillai, 2006), inhibition

of tumour angiogenesis (Li et al., 2005; Yoysungnoen et al., 2006), production of reactive

oxygen species (Atsumi et al., 2005a; Banjerdpongchai & Wilairat, 2005), and the induction

of detoxification enzymes (glutathione S-transferases) (Piper et al., 1998; Singhal et al.,

1999).

Page 85: Phytochemistry and pharmacology of plants from the ginger family

61

2.4.1.2.2.4 Other pharmacological activities

Curcumin has demonstrated a range of other pharmacological activities. These include

immunological effects (Kang et al., 1999; Kobayashi et al., 1997; Ranjan et al., 1998; Youn

et al., 2006), anti-diabetic effects (Mahesh et al., 2005; Sharma et al., 2006), and

hepatoprotective action (Kang et al., 2002; Nanji et al., 2003; Park et al., 2000; Sreepriya &

Bali, 2006; Sugiyama et al., 2006; Xu et al., 2003).

Histamine release by leukemia cells was markedly decreased by both curcumin and

tetrahydrocurcumin (Suzuki et al., 2005). Curcumin is a potent inhibitor of Kv1.4 K+

channels (Liu et al., 2006) and inhibits arachidonic acid- and PAF-induced platelet

aggregation and platelet TxA2 synthesis (Shah et al., 1999). Curcumin reversed endothelial

dysfunction caused by homocysteine (Ramaswami et al., 2004) and had anti-atherogenic

effects in animals (Olszanecki et al., 2005; Ramirez-Tortosa et al., 1999). Curcumin had

anti-fibrotic action in rats with glomerular disease (Gaedeke et al., 2005) and inhibited or

reduced a range of inflammatory molecules and signals (incl. NFκB, AP-1, IL-6, TNF-α and

iNOS) in rat models of both ethanol and non-ethanol pancreatitis (Gukovsky et al., 2003).

Dietary curcumin (0.002%) protected rats against galactose-induced cataract formation

(Suryanarayana et al., 2003), and curcumin had potent anti-ulcer effects in a rat model by

preventing glutathione depletion, lipid peroxidation, and protein oxidation (Swarnakar et al.,

2005).

Curcumin improved wound healing following gamma-irradiation (Jagetia & Rajanikant,

2004) and also facilitated wound healing in diabetic rodents (Sidhu et al., 1999).

Curcumin has produced positive results in animal models of inflammatory bowel disease;

this effect is associated with the inhibition of NFκB (Jian et al., 2005; Jiang et al., 2006a;

Salh et al., 2003; Ukil et al., 2003; Zhang et al., 2006). A small pilot study with patients

with ulcerative proctitis or Crohn’s disease has yielded promising results (Holt et al., 2005).

Several in vitro and in vivo studies have identified curcumin as a promising agent for the

prevention and treatment of Alzheimer’s disease (Baum & Ng, 2004; Lim et al., 2001; Ono

et al., 2004; Yang et al., 2005). Curcumin has also been proposed as an agent worth

investigating for the treatment of cystic fibrosis (Davis & Drumm, 2004), but the interest in

pursuing this appears to have waned following further studies (Mall & Kunzelmann, 2005;

Song et al., 2004).

Page 86: Phytochemistry and pharmacology of plants from the ginger family

62

2.4.2 Curcuma parviflora

Curcuma parviflora Wallich is native to Thailand, where is has been used in the topical

treatment of cuts (Sirirugsa, 1999) and for the detoxification of scorpion bites (Takahashi et

al., 2003). It has been shown to contain a series of dimeric sesquiterpenes, parviflorenes A-

H with cytotoxic properties (Takahashi et al., 2003; Toume et al., 2004; Toume et al., 2005).

Parviflorene F (Fig. 2-12) induced apoptosis in HeLa cells (Ohtsuki et al., 2008).

OH

OH

OH

OH

parviflorene A

parviflorene F

OH

Fig. 2-12. Parviflorene A and F from Curcuma parviflora.

2.4.3 Other Curcuma species

No information regarding the chemistry or pharmacology was identified for the other two

Curcuma species included in this work, C. australasica J. D. Hook and C. cordata Wallich.

Several other species of Curcuma are used for medicinal purposes, particularly in Asia.

These include Javanese turmeric C. xanthorrhiza Roxb. (included in the British

Pharmacopoeia (Anonymous, 2007), C. kwangsiensis S. G. Lee & C. F. Liang, C.

Page 87: Phytochemistry and pharmacology of plants from the ginger family

63

phaeocaulis Valeton, C. wenyujin Y. H. Chen & C. Ling (all three are included in the

Pharmacopoeia of the People’s Republic of China (Anonymous, 2005), and C. zedoaria

Roscoe (included in the Japanese Pharmacopoeia (Anonymous, 2001)).

However, since these species were included in the present work, they are not reviewed here.

2.5 Genus Boesenbergia Kuntze

The genus Boesenbergia comprises about 30 species and is distributed from India to New

Guinea (Mabberley, 1997). There are no native or naturalised members of this genus in

Australia.

Fingerroot (Boesenbergia rotunda (L.) Mansf. (syn. Boesenbergia pandurata Schltr.,

Kaempferia pandurata Roxb., Gastrochilus panduratus (Roxb.) Ridl.) is used as a culinary

and medicinal herb in Southeast Asia. Fresh rhizomes are slightly pungent and have a

characteristic aroma. They are used for the treatment of colic, fungal infections, dry cough,

rheumatism and muscular pains, and for their reputed aphrodisiac properties (bin Jantan et

al., 2001; Murakami et al., 1993; Trakoontivakorn et al., 2001).

Boesenbergia rotunda was included in the present work.

2.5.1 Chemistry

A range of flavonoids have been identified in the rhizome of B. rotunda, notably chalcones

(including boesenbergin A and B [Fig. 2-13], panduratin A, cardamomin and

dihydromethoxychalcone) and flavanones (including pinostrobin, pinocembrin, alpinetin and

5-hydroxy-7-methoxyflavanone) (Murakami et al., 1993; Trakoontivakorn et al., 2001). A

chalcone derivative, 4-hydroxypanduratin A, has also been identified (Trakoontivakorn et

al., 2001).

Steam distilled rhizome oils from B. rotunda rhizomes from Malaysia, Indonesia and

Thailand showed considerable qualitative and quantitative differences in composition (bin

Page 88: Phytochemistry and pharmacology of plants from the ginger family

64

Jantan et al., 2001). The major volatile oil constituents identified were camphor (16.1-

32.1%), geraniol (16.2-26.0%), (E)-β-ocimene (19.0-23.7%) and 1,8-cineole (7.5-13.9%).

O

O

OMe

HO

boesenbergin A

O

OH

MeO

O

boesenbergin B

Fig. 2-13. Boesenbergin A and B from Boesenbergia rotunda.

2.5.2 Pharmacology

Fingerroot has attracted attention as a potential chemopreventive agent. Both chalcones and

flavanones from this species have demonstrated anti-mutagenic activity against mutagenic

heterocyclic amines in Salmonella typhimurium TA98 assays (Trakoontivakorn et al., 2001).

An ethanolic extract of rhizomes had potent inhibitory activity (similar to that of turmeric,

Curcuma longa) against human HT-29 colon cancer and MCF-7 breast cancer cell lines

(Kirana et al., 2003).

A methanolic extract of fresh B. rotunda rhizomes showed strong inhibition of tumour

promoter-induced Epstein-Barr virus activation (Murakami et al., 1993). This in vitro assay

is used to screen for agents with possible anti-tumour promoting properties. The chalcone

cardamonin was isolated and identified as having potent inhibitory activity in this assay (IC50

Page 89: Phytochemistry and pharmacology of plants from the ginger family

65

= 3.1 µM). Panduratin A has also been shown to induce apoptosis in human colon cancer

HT-29 cells (Yun et al., 2005).

These in vitro findings are at odds with a medium-term bioassay of hepatocarcinogenesis in

rats (Tiwawech et al., 2000). This study found that the administration of B. rotunda dried

rhizome, combined with a hepatocarcinogenic heterocyclic amine found in cooked meats,

resulted in an increase in the genesis of preneoplastic liver cell foci, suggestive of a possible

cancer promoting effect.

Panduratin A isolated from B. rotunda potently and dose-dependently inhibited nitric oxide

(IC50 = 0.175 µM) and PGE2 (IC50 = 0.0195 µM) production in RAW264.7 cells, suppressed

the expression of iNOS and COX-2 and reduced NFκB transcriptional activity without

significant cytotoxicity (Yun et al., 2003).

2.6 Genus Hedychium Koenig

The genus Hedychium comprises approximately 50 species distributed from the Himalayas

to New Guinea. The genus is also possibly native to Madagascar. Many species, including

H. coronarium Koenig, are cultivated as ornamental plants and often called ginger lilies

(Mabberley, 1997). There are no Hedychium species native to Australia, but two species, H.

coronarium and H. gardnerianum Sheppard ex Ker-Gawler, have become naturalised

(Smith, 1987).

Several species of Hedychium have been reported to have ethnobotanical uses including H.

coronarium, which has been employed for rhinitis, tonsillitis, halitosis, swelling and tumours

(Johnson, 1999). In Mauritius, the rhizomes of H. coronarium and H. flavescens are both

used in traditional medicine as topical applications for rheumatism (Gurib-Fakim et al.,

1997). Hedychium species are not used for culinary purposes (Omata et al., 1991).

The ground rhizome of H. spicatum Koenig is the major component of abir, a scented

powder used in Hindu ceremonies (Mabberley, 1997). This species has also been used

medicinally. The rhizome has been employed as a stomachic, carminative and stimulant,

whereas the root and stalk are reported to have analgesic and anti-inflammatory actions,

having been used in liver, urinary and respiratory disorders (Tandan et al., 1997).

Page 90: Phytochemistry and pharmacology of plants from the ginger family

66

Only Hedychium coronarium was included in the present work.

2.6.1 Chemistry

The steam-distilled volatile oil from the rhizome of H. gardnerianum was found to contain

67% monoterpenes and 31% sesquiterpenes, including 18% cadinane derivatives

(Weyerstahl et al., 1998). The major constituents of the oil were β-pinene (24.5%), α-pinene

(21.5%), p-cymene (6.2%), α-cadinol (4.9%), 10-epi-α-cadinol (2.4%), ar-curcumene

(2.3%) and linalool (2.1%).

Rhizome oil of H. coronarium from India was found to contain α- and β-pinene, limonene,

δ-3-carene, β-phellandrene, p-cymene, 1,8-cineole, β-caryophyllene, β-caryophyllene oxide,

linalool and elemol in varying quantities (Dixit et al., 1990). In addition to these

constituents, the following compounds also occur in the oil: myrcene, α-phellandrene,

borneol, methyl salicylate and eugenol (Omata et al., 1991).

Steam-distilled essential oil of fresh H. coronarium rhizome from Tahiti, to where the

species has been introduced from Indo-China, contained 1,8-cineole (40.2%), β-pinene

(24.8%), α-pinene (7.8%) and α-terpineol (5.4%) as major constituents (Lechat-Vahirua et

al., 1993).

In a study related to the present work steam-distilled oil from rhizomes of H. coronarium

grown in Mauritius was characterised by α-muurolol (17%), α-terpineol (16%) and 1,8-

cineole (11%) (Gurib-Fakim et al., 2002).

Seven labdane diterpenes have also been isolated from the rhizome of H. coronarium

(Nakatani et al., 1994).

2.6.2 Pharmacology

Anthelmintic properties in vitro (Dixit & Varma, 1975) and fungitoxicity against Aspergillus

flavus (Singh et al., 1984) have been reported for the essential oil of H. coronarium. The

ethanolic extract of the rhizome of this species does not appear to have been studied

Page 91: Phytochemistry and pharmacology of plants from the ginger family

67

previously, but in vivo analgesic and anti-inflammatory effects have been reported for an

ethanolic extract of H. spicatum (Tandan et al., 1997).

2.7 Genus Kaempferia L.

The genus Kaempferia contains approximately 50 species ranging from India to Southern

China and Malaysia. Among these, K. rotunda L. and in particular K. galanga L. are used

for medicinal and culinary purposes (Johnson, 1999; Mabberley, 1997).

The rhizome of lesser galangal (K. galanga) (also known as East Indian galangal, kacholam,

kentjur, chekur, and sometimes, misleadingly, simply ‘galangal’) is used medicinally and

cultivated in several countries, including India, China, Vietnam, Indonesia and the Sudan

(Luger et al., 1996; Vimala et al., 1999). It is considered to be warming, promoting vital

energy circulation, and is used in the treatment of complaints caused by cold, including

headache, dyspepsia, vomiting, diarrhoea, toothache, and abdominal and pectoral pain, while

a liniment applied topically through massage is used in rheumatism (Luger et al., 1996).

It should be noted that the binomial Kaempferia pandurata is a synonym for Boesenbergia

rotunda (refer to Section 2.4 above).

Kaempferia galanga and K. rotunda were included in the present work.

2.7.1 Chemistry

The rhizome of K. galanga contains a volatile oil with ethyl-p-methoxy-E-cinnamate as a

constituent (Luger et al., 1996). Other constituents reported from the rhizome of this species

are ethyl cinnamate, cinnamaldehyde, camphene, l-∆3-carene, borneol, p-methoxystyrene

and pentadecane (Noro et al., 1983).

2.7.2 Pharmacology

Several studies have investigated the biological activity of K. galanga. An ethanolic extract

exhibited moderate activity in a cell assay of anti-tumour promoter activity (inhibition of 12-

O-tetradecanoyl phorbol-13-acetate (TPA)-induced Epstein-Barr virus early antigen (EBV-

Page 92: Phytochemistry and pharmacology of plants from the ginger family

68

EA) in Raji cells) (Vimala et al., 1999). Another study showed moderate inhibition of

tumour promoter-induced Epstein-Barr virus activation (Murakami et al., 1993). This is a

similar in vitro assay, which like the first is used to screen for agents with possible anti-

tumour promoting properties.

A methanolic extract of K. galanga rhizomes showed in vitro cytotoxic activity against HeLa

cells, with an IC50 value of 80 µg/mL (Kosuge et al., 1985). Ethyl-p-methoxy-E-cinnamate

was isolated from this extract and identified as a cytotoxic compound, with an IC50 value of

35 µg/mL.

Ethyl-p-methoxy-trans-cinnamate isolated from K. galanga rhizomes has also been shown to

inhibit monoamine oxidase in vitro, with an IC50 value of 6.8 x 10−5 M (Noro et al., 1983).

2.8 Genus Scaphochlamys Baker

The genus Scaphochlamys includes 30 species, which occur in India, Southeast Asia and

New Guinea (Mabberley, 1997).

Scaphochlamys biloba and S. kunstleri were included in the present work.

No information about the ethnobotany, chemistry or pharmacology of this genus was located.

2.9 Genus Alpinia Roxburgh

The genus Alpinia comprises some 200 species distributed mainly in Asia and the Pacific

(Mabberley, 1997). Five species of Alpinia are indigenous to Australia: A. arctiflora (F.

Muell.) Benth., A. arundelliana (F. M. Bailey) Schumann, A. caerulea (R. Br.) Benth. (syn.

A. coerulea Benth.), A. hylandii R. M. Smith and A. modesta Schumann (Hnatiuk, 1990).

The species formerly known as A. racemigera F. Muell. is now named Pleuranthodium

racemigerum (F. Muell.) R. M. Smith (refer to Section 2.10). Both A. caerulea (commonly

known as ‘native ginger’ in eastern Australia) and A. arundelliana grow in rainforest and wet

sclerophyll forest in coastal New South Wales, north from the Central Coast, and in

Page 93: Phytochemistry and pharmacology of plants from the ginger family

69

Queensland, while the other three species are restricted to north-eastern Queensland

(Hnatiuk, 1990; Wilson & Hastings, 1993).

Many members of the genus are used as spices and/or as medicinal plants in indigenous

systems of traditional medicine around the world. This is particularly the case in China,

India and South-East Asia, but medicinal use of Alpinia species has also been reported from

Saudi Arabia and Brazil (Tewari et al., 1999). Galangal (greater galangal), A. galanga (L.)

Swartz is widely cultivated in Southeast Asia, where it is used as a spice and a traditional

medicine (Kubota et al., 1998; Someya et al., 2001).

The following Alpinia taxa were included in the present work: A. arctiflora (F. Muell.)

Benth., A. caerulea (R. Br.) Benth., A. calcarata Rosc., A. galanga (L.) Swartz, A.

luteocarpa Elmer, A. malaccensis Rosc., A. modesta Schumann, A. mutica Roxb., A.

purpurea ‘Eileen McDonald’, A. spectabile ‘Giant Orange’ and A. zerumbet (Pers.) B.L.

Burtt & R.M. Sm.

2.9.1 Chemistry

The chemistry of a range of Alpinia species has been reviewed by Tewari and co-workers

(Tewari et al., 1999), although it is evident that this review was not comprehensive.

Essential oil from Alpinia species, mostly extracted from rhizomes but also in some cases

from leaves, stems and seeds, contain terpenoids and phenylpropanoids typical of essential

oils. This includes monoterpenoids such as α- and β-pinene, geraniol, borneol, citronellol,

linalool, 1,8-cineole and camphor, sesquiterpenoids including eudesmol, β-

sesquiphellandrene and curcumene, and phenylpropanoids such as methyl eugenol (Tewari et

al., 1999). Four acetoxy-cineole isomers, viz. the (trans and cis)-2- and 3-acetoxy-1,8-

cineoles, have been identified as the major aroma constituents in the rhizome of A. galanga

(Kubota et al., 1998; Kubota et al., 1999). The major pungent compound in A. galanga is

1’-acetoxychavicol acetate (Fig. 2-14), which is at least ten times less pungent and produces

a more short-lived pungent sensation than capsaicin (Gautschi et al., 1999). Several pungent

diarylheptanoids have also been isolated from Alpinia species (Tewari et al., 1999).

Page 94: Phytochemistry and pharmacology of plants from the ginger family

70

O

O

O

O

Fig. 2-14. 1’-acetoxychavicol acetate from Alpinia galanga.

Two kava-pyrones, 5,6-dehydrokawain and dihydro-5,6-dehydrokawain, have been isolated

from the leaves of A. zerumbet (Pers) B. L. Burtt & R. M. Smith (syn. A. speciosa K.

Schum.), a species used as a diuretic and hypotensive agent in Brazil (Kuster et al., 1999).

Computerised searches of the databases Chemical Abstracts, MEDLINE, CAB Abstracts,

Current Contents Connect and Web of Science using the search terms ‘Alpinia caerulea’ and

‘Alpinia coerulea’ yielded no results relating the chemistry or pharmacology of this

Australian species.

2.9.2 Pharmacology

Several species of Alpinia are used in traditional medicine. Greater galangal (A. galanga) is

used for the treatment of stomach ache in China and Thailand (Yang & Eilerman, 1999), and

a number of species are used in traditional Ayurvedic medicine in India for inflammation,

cancer, urinary calculi, and as a digestive, liver and spleen tonic for the treatment of various

digestive complaints (Tewari et al., 1999). A. zerumbet is used medicinally in Brazil (Kuster

et al., 1999).

Alpinia galanga has demonstrated gastric anti-secretory and anti-ulcer effects in

experimental animals (Tewari et al., 1999) and is reported to possess anti-tumour (Itokawa et

al., 1987), anti-fungal (Haraguchi et al., 1996; Scheffer et al., 1981) and anti-oxidant activity

(Cheah & Abu, 2000). An extract of A. galanga in combination with a ginger (Zingiber

officinale) rhizome extract is marketed as an anti-inflammatory agent for use in arthritis

under the name Zinaxin (http://www.zinaxinrapid.com/). This product was the study

Page 95: Phytochemistry and pharmacology of plants from the ginger family

71

medication in one clinical trial in patients with osteoarthritis (Altman & Marcussen, 2001)

(refer to Section 2.2.1.5.1) (Altman & Marcussen, 2001).

A methanolic extract of A. galanga rhizome showed strong inhibitory effect in an in vitro

assay using inhibition of Epstein-Barr virus activation as a means of identifying potential

anti-tumour promoters (Murakami et al., 1994). This study identified 1’-acetoxychavicol

acetate, the major pungent principle in A. galanga, as the active principle with an IC50 value

of 1.5 µM.

Alpinia malaccensis and A. mutica had in vitro anti-bacterial and potent anti-oxidant activity

(Habsah et al., 2000b).

In Brazil, A. zerumbet is used as a diuretic and anti-hypertensive agent (Kuster et al., 1999).

The anti-hypertensive effect is likely due to the presence of several flavonoids with calcium

channel blocking activity. The kava pyrones present (see above) also have anti-platelet, anti-

convulsant, analgesic, sedative and anxiolytic actions. An aqueous extract of the leaves was

found to promote central nervous system excitation, followed by depression and

hypokinesia.

The major pungent principle of A. galanga, 1’-acetoxychavicol acetate, possesses

considerable biological activity, with anti-ulcer, anti-tumour and anti-microbial properties, as

well as inhibition of xanthine oxidase having been demonstrated (Kubota et al., 1998). Most

recently, the compound has been shown to inhibit the replication of human

immunodeficiency virus type 1 (Ye & Li, 2006).

Acute (24 h) and chronic (90 days) oral toxicity studies on the ethanolic extract of A.

galanga rhizome were carried out in mice (Qureshi et al., 1992). Acute extract dosages were

0.5, 1.0, and 3 g per kg body weight, while the chronic dosage was 100 mg/kg/day. No

significant mortality was observed compared with controls, but galangal treated mice

experienced a significant weight gain and an increased level of red blood cells. In male

mice, a highly significant increase in sperm count, sperm motility and weight of sexual

organs was recorded.

Diarylheptanoids possess potent anti-inflammatory properties. An early study found that

diarylheptanoids including yakuchinone A isolated from A. officinarum were dual inhibitors

of arachidonic acid metabolism, i.e. they inhibited both prostaglandin synthetase (COX) and

Page 96: Phytochemistry and pharmacology of plants from the ginger family

72

5-lipoxygenase (Kiuchi et al., 1992). The anti-inflammatory activity of diarylheptanoids

may in part result from their suppressive effect on the surface expression of inducible

adhesion molecules in endothelial cells and subsequent leukocyte adhesion (Yamazaki et al.,

2000), but other mechanisms may also be involved. Two diarylheptanoids isolated from A.

oxyphylla Miquel, yakuchinone A and B, were found to down-regulate the expression of

COX-2 and iNOS through suppression of NFκB in mouse skin treated with a tumour

promoter (Chun et al., 2002).

2.10 Genus Pleuranthodium (K. Schum.) R. M. Smith

The genus Pleuranthodium comprises some 23 species distributed in New Guinea, Australia

and the Pacific (Mabberley, 1997). In Australia, Pleuranthodium racemigerum (F. Muell.)

R. M. Smith (syn. Alpinia racemigera F. Muell., Psychanthus racemiger (F. Muell.) R. M.

Smith) occurs in the Wet Tropics region of Far North Queensland and was included in the

present work. No information about the chemistry or pharmacology of this species was

located in the literature.

2.11 Genus Etlingera Giseke

The genus Etlingera is made up of around 60 species distributed from India to New Guinea.

The genus is closely allied to, and sometimes included in, the genus Amomum (Mabberley,

1997).

One species is native to Australia, E. australasica R. M. Smith, which is confined to

rainforest in the north-eastern Cape York Peninsula in North Queensland (Smith, 1987).

Torch ginger (E. elatior (Jack) R. M. Smith) from Malesia is cultivated as an ornamental,

and its flowers are used as a condiment in cooking (Mabberley, 1997). The leaves of a

Tahitian species, E. cevuga Smith (syn. Amomum cevuga, Geanthus cevuga), are used as a

yellow dyeing material for dying vegetable fibres for traditional dresses (Lechat-Vahirua et

al., 1993).

Page 97: Phytochemistry and pharmacology of plants from the ginger family

73

The present work included Etlingera australasica R.M. Sm. and E. elatior ‘Burma Torch’.

No information about the chemistry and pharmacology of these taxa was identified, but the

essential oil of E. cevuga has been reported to contain methyl eugenol (47%) and (E)-methyl

isoeugenol (18%), as principal constituents (Lechat-Vahirua et al., 1993). No

pharmacological data pertaining to the genus Etlingera was located.

2.12 Genus Elettaria Maton

The genus Elettaria comprises seven species and is distributed from India to western

Malesia. Cardamom (E. cardamomum Maton), a native of India, produces aromatic seeds,

which are used in cooking and medicine (Mabberley, 1997). Today cardamom seeds are

obtained from the cultivated variety E. cardamomum var. minuscula, and the main producing

countries are Sri Lanka, India and Guatemala (Evans, 2002).

The medicinal use of cardamom seeds dates back several thousand years in the Indian system

of Ayurveda, and they were known to the Greeks in the 4th century BCE. In India,

cardamom is used for a variety of complaints including digestive problems, asthma,

bronchitis, kidney stones, anorexia and debility (Chevallier, 2001).

Cardamom fruit is included in the British Pharmacopoeia (Anonymous, 2007), and the seed

was approved by the German Commission E for the treatment of dyspepsia (Blumenthal,

1998).

2.12.1 Chemistry

Cardamom seeds yield 2.8-6.2% volatile oil with a high content of terpinyl acetate and 1,8-

cineole (Fig. 2-15) and smaller amounts of other monoterpenes including alcohols and esters

(Evans, 2002; Lawrence, 1998).

Page 98: Phytochemistry and pharmacology of plants from the ginger family

74

O

O

terpinyl acetate

O

1,8-cineole

Fig. 2-15. Terpinyl acetate and 1,8-cineole from Elettaria cardamomum.

No information on the chemical constituents of the underground parts of cardamom was

identified.

2.12.2 Pharmacology

Cardamom seed oil had anti-inflammatory activity against acute carrageenan-induced paw

oedema in the rat and analgesic activity in mice, producing 50% protection against writhing

induced by intraperitoneal administration of 0.02% β-benzoquinone (Al-Zuhair et al., 1996).

The same authors also reported anti-spasmodic activity in rabbit intestine in vitro resulting

from muscarinic receptor blockage. The seed oil administered intravenously (5-20 µL per

kg) caused dose-dependent decrease in arterial blood pressure in the rat, and large doses

relaxed spontaneously contracting rabbit jejunum in vitro (El Tahir et al., 1997).

The volatile oil of cardamom has been shown to enhance transdermal absorption of

indomethacin, as a result of its monoterpene content (Huang et al., 1993; YawBin et al.,

1999).

An aqueous cardamom seed extract (10% w/v) was shown to increase gastric acid secretion

almost three-fold in anaesthetised rats (Vasudevan et al., 2000).

Ethanolic extracts of combinations of cardamom leaves and root and also of roots and seeds

inhibited the growth of Staphylococcus aureus (Daswani & Bohra, 2000). No other

pharmacological information pertaining to root or rhizome preparations was located.

Page 99: Phytochemistry and pharmacology of plants from the ginger family

75

2.13 Genus Hornstedtia Retz.

The 24 species in the genus Hornstedtia occur in South-east Asia, Papua New Guinea and

Australia (Mabberley, 1997).

The species included in the present work, Hornstedtia scottiana (F. Muell.) Schum., is found

in New Guinea and Far North Queensland (Anonymous, 2008). It is reportedly edible (Dick,

1992), but no information about its chemical composition or biological activity was located.

A labdane diterpene, labda-8(17),12-diene-15,16-dial, has been reported from H. scyphifera

(Koenig) Steud. (Sirat et al., 1994), but no other information about the chemistry or

pharmacology of the genus was located.

2.14 Genus Renealmia L.f.

The genus Renealmia consists of some 60 species in tropical parts of the Americas and

Africa (Mabberley, 1997). Several species are reported to be used locally as medicinal

plants (Dos Santos & Sant'Ana, 2000; Lans et al., 2001; Otero et al., 2000; Zhou, 1997).

Diterpenoids (Boukouvalas et al., 2006; Sekiguchi, 2001; Yang, 1999), dihydrochalcones

(Gu, 2002), diarylheptanoids (Sekiguchi, 2002) and sesquiterpenes (Kaplan et al., 2000)

have been reported from the genus.

The Central American species Renealmia cernua (Sw. ex Roem. & Schult.) J.F. Macbr. was

included in the present work. No information about the chemistry or pharmacology of this

species was identified.

2.15 Genus Costus L.

The genus Costus is a tropical genus with 42 species and is placed in the family Costaceae

by some authorities (Mabberley, 1997). There is one endemic species in Australia, C.

potierae (Smith, 1987). Various species have been used in traditional medicine in Europe

and the Americas for a wide range of ailments (Johnson, 1999). C. speciosus (Koenig) Sm.

from Indomalesia is a commercial source of diosgenin, which occurs in the rhizome in a

Page 100: Phytochemistry and pharmacology of plants from the ginger family

76

concentration up to 2.6% on a dry weight basis (Gupta et al., 1981; Kaphai et al., 1977;

Mabberley, 1997).

It should be noted that the Asian drug ‘costus root’ is not derived from the genus Costus but

from Saussurea lappa (Asteraceae) (de Kraker et al., 2001; Moeslinger et al., 2000).

The following species of Costus were included in the present work: C. barbatus Suess., C.

leucanthus Maas, C. malortieanus H. Wendland, C. productus Gleason ex Maas, C.

pulverulentus C. Presl and C. tappenbeckianus J. Braun et K. Schum.

2.15.1 Chemistry

Diosgenin (Fig. 2-16) has been reported from C. malortieanus (Prasad & Ammal, 1983), and

other steroidal glycosides (including furostanol and spirostanol glycosides) have been

identified in the rhizomes of several other species (Agrawal et al., 1984; Inoue et al., 1995;

Pereira et al., 1999; Rui et al., 1997; Silva et al., 1998; Willuhn & Pretzsch, 1985).

O

HO

H

H

O

H

Fig. 2-16. Diosgenin from Costus malortieanus.

No further information was located pertaining to the chemistry of the species included in the

present work.

Page 101: Phytochemistry and pharmacology of plants from the ginger family

77

2.15.2 Pharmacology

No information about the pharmacology of any of the species included in the present work

was located, but C. spiralis, which is used in Brazilian folk medicine for urinary tract

complaints and calculi, reduced the growth of foreign body-induced calculi in the bladder of

rats (Viel et al., 1999), and C. lucanusianus, a traditional anti-abortion drug in the Ivory

Coast, produced complete inhibition of oxytocin-induced contractions in isolated rat uterus

(Foungbe et al., 1987).

2.16 Genus Tapeinochilos Miquel

The genus Tapeinochilos comprises approximately twelve species distributed in Indonesia,

New Guinea and Australia (Mabberley, 1997; Smith, 1987). The genus is closely related to

Costus, and is placed in the family Costaceae by some taxonomists (Mabberley, 1997). One

species, T. ananassae (Hassk.) K. Schum. (syn. T. queenslandiae K. Schum) occurring in

Queensland and New Guinea (Smith, 1987), was included in the present work.

No information about the chemistry or pharmacological activity of T. ananassae or any other

members of this genus was located.

2.17 Summary

The Zingiberaceae contains many species that have been used as medicinal plants or spices,

in particular in Southeast Asia. Extensive information about the chemistry and

pharmacology of these plants is available only for a few species, in particular ginger and

turmeric. For the majority of species, including many species traditionally used as

medicines, limited or no information is available, and this is also true for whole genera in

some cases.

Since the family clearly produces compounds with interesting pharmacological action, the

investigation of little known species should prove rewarding.

Page 102: Phytochemistry and pharmacology of plants from the ginger family

78

2.18 Aims and research questions

The overall aim of this work has been to contribute to the existing body of knowledge about

members of the Zingiberaceae as medicinal plants, in particular in terms of their potential

use as anti-inflammatory agents. The work comprised two distinct parts. The first part

focused on the phytochemistry of a number of ginger (Zingiber officinale) clones with the

aim of identifying one or more with unique chemistry and consequent particular therapeutic

prospects. The second part involved the screening of species from various genera in a

number of bioassays relevant to potential anti-inflammatory activity, with the aim of

increasing the understanding of established medicinal species and the potential identification

of species with novel therapeutic prospects.

2.18.1 Part 1: Phytochemical investigations of 17 ginger clones

That pharmacologically active constituents in medicinal plants can vary both qualitatively

and quantitatively is well established, and a range of genetic, environmental, ontogenic and

biologic factors can affect the production of secondary metabolites in plants (Evans, 2002;

Wijesekera, 1991).

Ginger is particularly interesting in this context, because the species is sterile and can only

propagate by vegetative means. This, combined with the fact that the plant has been widely

cultivated in tropical regions around the world for several hundreds of years, makes it

probable that genetically stable and potentially chemically distinct clones exist.

Access to 17 ginger clones (including 12 of tetraploid genotype) through the Queensland

Department of Primary Industries represented an excellent opportunity to investigate

phytochemical variability resulting from genotypic differences.

2.18.1.1 Research questions

A number of research questions arose from the literature review in relation to the

phytochemical variability of ginger clones:

Page 103: Phytochemistry and pharmacology of plants from the ginger family

79

1. Are there significant qualitative and/or quantitative phytochemical differences

between different clones of ginger that are genetically determined and independent of

environmental, ontogenic or biologic factors?

2. If such genetically determined phytochemical differences exist, would it be possible

to select one or more clones with characteristics that would make them particularly

suitable for pharmaceutical or other specific uses?

3. Does fresh ginger contain shogaols, or do they only form from gingerols post

harvest?

4. Does the essential oil composition vary significantly between clones and if so, how?

2.18.1.2 Specific aims

Based on the research questions above, a number of specific aims were formulated for this

part of the project. They were:

1. To assess ginger extracts, in a pilot study, for in vitro COX-1 inhibitory activity,

determine their content of gingerols and shogaols, and examine whether a correlation

between activity and levels of gingerols/shogaols exist.

2. To compare the 17 ginger clones in terms of their yield of pungent gingerols and

shogaols, both of which have demonstrated anti-inflammatory activity in vitro and

which are responsible for the pungency of ginger.

3. To identify one or more clones that yield particularly high amounts of these

compounds and therefore would be suitable for pharmaceutical use.

4. To determine whether fresh ginger contains shogaols.

5. To compare the 17 clones in terms of their essential oil composition.

6. To determine if all clones have high citral content (Australian ginger is said to

possess a particular ‘lemony’ aroma).

Page 104: Phytochemistry and pharmacology of plants from the ginger family

80

2.18.1.3 Research methods

The following methods were employed in order to meet the specific aims of this part of the

project:

1. Cultivation of ginger clones under uniform conditions.

2. Standardised extraction of ginger rhizomes.

3. Assay of COX-1 inhibition of in vitro.

4. Quantification of gingerols and shogaols by HPLC.

5. Extraction of essential oil from ginger rhizomes by steam distillation.

6. Analysis of essential oils by GC-MS.

2.18.2 Part 2: Screening Zingiberaceae for pharmacological activity

The second part of this work involved the screening of a number of Zingiberaceae species in

a range of bioassays. This work adds to the existing knowledge base regarding the

pharmacology of several traditional medicinal plants and provides novel information about

species that have not previously been investigated in this way.

Also, the work was designed to test the hypothesis that the combination of ethnobotanical

and taxonomic information is a productive strategy to identify previously unrecognised plant

species with therapeutic potential, either in the form of phytomedicines or by providing lead

compounds for subsequent drug development.

As evidenced by the literature review above, the Zingiberaceae contains many species that

have been used as traditional medicines. Among the most well known medicinal species

from the family are ginger (Zingiber officinale), turmeric (Curcuma longa), galangal

(Alpinia galanga) and fingerroot (Boesenbergia rotunda). Thus ethnobotanical information

has been used to identify a plant family and specific genera with significant pharmacological

and therapeutic activity. Subsequently, taxonomic information (based on phylogenetic

relationships) has been applied to identify species related to established medicinal species at

Page 105: Phytochemistry and pharmacology of plants from the ginger family

81

the generic or family level. This approach is based on the well established concept that

phylogenetically related plant species usually display a significant degree of similarity in the

kinds of secondary metabolites they produce (Evans, 2002).

2.18.2.1 Research questions

A number of research questions arose in relation to the second part of this work:

1. Can the mode of action of well known medicinal species from the Zingiberaceae be

further elucidated through bioassays?

2. Can Zingiberaceae species with previously unrecognised pharmacological activity be

identified through screening in bioassays?

3. Can species with previously unrecognised pharmacological activity be identified

from genera with recognised medicinal species (such as Zingiber, Curcuma and

Alpinia)?

4. If species with previously unrecognised pharmacological activity are identified, do

they contain active compounds that are similar to active compounds in related

medicinal species?

2.18.2.2 Specific aims

In accordance with the research questions, the specific aims of this part of the work were:

1. To screen a number of Zingiberaceae species in a range of bioassays (inhibition of

PGE2, antioxidant activity, inhibition of nitric oxide, stimulation of natural killer cell

activity, cytotoxicity).

2. Through bioassay work, to contribute to the understanding of the mode of action of

established medicinal plants.

3. To identify previously unrecognised species with pharmacological activity that

makes them candidates for novel phytomedicines or drug leads.

Page 106: Phytochemistry and pharmacology of plants from the ginger family

82

4. To isolate and determine the structure of major active compounds in previously

unrecognised medicinal species.

2.18.2.3 Research methods

To address the specific aims of this part of the project, the following methods were

employed:

1. Inhibition of PGE2 in 3T3 murine fibroblasts.

2. Oxygen radical absorption capacity (ORAC) assay.

3. Nitric oxide production by RAW 264 murine leukemic monocytes.

4. Cytotoxic activity of human natural killer (NK) cells against K562 human myeloid

leukemic cells by flow cytometry.

5. Cytotoxicity in murine and human cell lines.

6. Chemical profiling of extracts by liquid chromatography – mass spectrometry (LC-

MS).

7. Bioactivity-guided fractionation by semi-preparative high-performance liquid

chromatography (HPLC).

8. Structural elucidation by nuclear magnetic resonance (NMR) spectroscopy.

Page 107: Phytochemistry and pharmacology of plants from the ginger family

83

3. INHIBITION OF CYCLOOXYGENASE-1 BY GINGER RHIZOME EXTRACTS

3.1 Introduction

The aims of this study were: (i) to determine the most suitable method for extraction of

ginger rhizome; (ii) to determine the most suitable extraction medium for ginger rhizome;

(iii) to establish a protocol for the quantitative determination of major pungent ginger

compounds by high-performance liquid chromatography (HPLC); and (iv) to establish a

protocol for the determination of the biological activity of ginger rhizome extracts in a

cyclooxygenase-1 (COX-1) bioassay. The determination of the most suitable extraction

method and medium was guided by the biological activity observed in the bioassay.

3.2 Materials and Methods

3.2.1 Plant material

A single rhizome of Zingiber officinale 'Queensland', grown outdoors under sprinklers at

Southern Cross University over the previous 7 months, provided the raw material for all

ginger extracts used in this study. The plant was grown in a black 300 mm plastic pot in a

sand-peat 1:1 growth medium containing the following nutrients (g per m3): dolomite 3600,

ammonium sulfate 544, superphosphate 184, potassium sulfate 248, magnesium sulfate 472,

copper sulfate 7.2, zinc sulfate 9.6, iron sulfate 7.2 (Smith & Hamill, 1996).

The freshly harvested rhizome was peeled and chopped finely with a knife. For hot

extractions, 10 g of chopped rhizome was added to 150 ml of extraction medium (water or

ethanol), ground with a tissue tearer (Heidolph Diax 600) and extracted in a Soxhlet

apparatus for 5 hours. Extracts were reduced to near dryness under vacuum on a rotary

evaporator, then reconstituted in 100 ml fresh extraction medium. In Experiment 1, residues

from ethanolic extractions were dissolved in 75% ethanol and 25% water, whereas residues

from aqueous extractions were dissolved in 75% water and 25% ethanol. Laboratory grade

potable alcohol (96%) was used in the preparation of ginger extracts.

Page 108: Phytochemistry and pharmacology of plants from the ginger family

84

For extractions at ambient temperature (approximately 23°C), 1 g of chopped rhizome was

added to 10 ml of extraction medium, ground with a tissue tearer (Heidolph Diax 600), then

placed in a sonicator and macerated at ambient temperature for 20 minutes. The macerate

was then centrifuged (1800 RCF for 5 minutes), the supernatant removed, and the pellet

extracted again with 10 ml fresh extraction medium following the same procedure. The

supernatant fluids from both extractions were combined, reduced to near dryness under

vacuum on a rotary evaporator, then reconstituted in 10 ml fresh extraction medium. In

Experiment 1, residues from ethanolic extractions were dissolved in 75% ethanol and 25%

water, whereas residues from aqueous extractions were dissolved in 75% water and 25%

ethanol.

A preliminary experiment showed that two subsequent extractions with ethanol yielded 96%

of the substances extracted by 4 subsequent extractions (data not shown).

3.2.2 HPLC analysis

Reversed-phase HPLC analysis of the plant extracts was performed on a Hewlett Packard

1100 HPLC fitted with a HP LiChrospher 100 RP-18e (5µm) column. Mobile phase was

pumped at 1mL/min (90% H2O/10% acetonitrile to 10% H2O/90% acetonitrile over 20 min).

Injection volume was 10 µL. Data were collected using a UV/visible diode array detector

scanning from 195 to 330 nm.

Identification of [6]-gingerol, [8]-gingerol, [10]-gingerol, [6]-shogaol and [8]-shogaol was

based on comparisons of retention time and UV-spectra with pure synthesised standards

obtained from the Department of Pharmacy at the University of Sydney. Absolute ethanol

(99%) analytical grade was used in the preparation of standards. Quantification of the

compounds mentioned above was based on standard curves prepared with pure standards and

eugenol as an external standard. Duplicate injections were analysed of all samples.

3.2.3 Cyclooxygenase-1 assay

The COX-1 assay was carried out using 14C-labelled arachidonic acid as a substrate. Ginger

extracts were incubated with labeled arachidonic acid and cyclooxygenase-1 for 15 min at

Page 109: Phytochemistry and pharmacology of plants from the ginger family

85

37° C, after which the reaction was stopped. Arachidonic acid metabolites were extracted

and separated by thin layer chromatography and quantified using a liquid scintillation

counter. Samples were assayed in triplicates. Details of the assay are given below.

3.2.3.1 Enzyme reaction

40 µl Cox-buffer (0.1 M TRIS at pH 8.0 containing 0.5mM sodium EDTA and 0.5 mM

phenol), 10 µl co-factor-1 (aqueous solution of 10 µM hematin [Sigma] and 0.02 M NaOH)

and 10 µl co-factor-2 (0.1 mM TRIS buffer containing 10 mM reduced glutathione [Sigma]

and 10 mM (–)-epinephrine-(+)-bitartrate [Sigma] adjusted to pH 8.0) was added to

Eppendorf tubes and incubated for 5 minutes at 37°C. 10 µl dilute ovine cyclooxygenase-1

solution (Cayman Chemicals; 20 µl in 190 µl Cox-buffer) was added to each tube followed

by 10 µl sample or control solvent. 40 µl 14C-arachidonic acid (Amersham, Australia) was

diluted with 360 µl Cox-buffer and 20 µl of this solution added to each tube. Tubes were

vortexed gently after each addition of material. Tubes were incubated for 15 minutes at 37°C

with gentle shaking after which the reaction was stopped by the addition of 40 µl formic acid

solution to each tube.

3.2.3.2 Extraction of arachidonic acid and metabolites

After the enzymatic reaction was stopped, 200 µl chloroform (CHCl3) was added to each

tube, which was then vortexed at high speed for approximately 10 seconds. The chloroform

(bottom) layer was then transferred to a clean Eppendorf tube using a micropipette, and the

chloroform evaporated under a nitrogen gas stream.

3.2.3.3 Separation of arachidonic acid and metabolites

This separation was carried out by means of thin-layer chromatography (TLC). Two mobile

phases were prepared as follows: Mobile Phase-1 (hexane 70 ml, diethyl ether 30 ml, glacial

acetic acid 1 ml) and Mobile Phase-2 (ethyl acetate 40 ml, methanol 10 ml, distilled water 25

Page 110: Phytochemistry and pharmacology of plants from the ginger family

86

ml, combined in a separation funnel). Extracted substances were reconstituted in 20 µl

CHCl3 and applied to a 200 x 200 mm TLC plate (0.2mm Silicon F254; Merck, Germany).

Two plates were used at a time. Standards (arachidonic acid 2 µl and prostaglandin E2/D2 10

µl [Sigma]) were applied to one lane on each plate. After the application of samples and

standards, the plates were allowed to dry, then run in a TLC tank (CAMAG) containing

Mobile Phase-1 until the solvent front reached a groove cut in the gel at 10 cm. Plates were

removed from the tank, allowed to dry for 15 minutes, then inserted into another TLC tank

containing Mobile Phase-2 and run until the solvent front reached a pencil mark at 3 cm.

Plates were again removed from the tank, allowed to dry for 15 minutes and then placed in

an iodine chamber for approximately 45 minutes. After this development, the arachidonic

acid standard (Rf 65-70) and the prostaglandin standard (just below the 3 cm line) were

marked with a pencil on both plates. Two parallel lines, 10-13 mm apart and equidistant

from the marked prostaglandin standard, were drawn horizontally across the plate and the

number of each lane written between the lines. The plate was then cut up along the parallel

lines and the grooves separating the lanes, resulting in 9 pieces of silica gel containing the

prostaglandin metabolites of arachidonic acid from each plate. The silica gel from each

numbered piece was then scraped off the aluminium plate and transferred to a scintillation

tube to which 200 µl distilled water and 200 µl Soluene-350 (Packard, Australia) was added.

After thorough vortexing, tubes were left to stand for 30 minutes, after which 2 ml Insta-Gel

(Packard, Australia) and 2 ml Hionic-Fluor (Packard, Australia) was added to each tube.

Tubes were vortexed for 10 seconds, then placed in a liquid scintillation counter (Minaxiβ

Tri-carb® 4000; United Technologies Packard) running Program 4. Results were recorded as

counts per minute (CPMA/K).

3.2.4 Statistical analysis

Data were analysed using Excel 97-SR-1 (Microsoft Corporation, Redmond, WA). Statistics

calculated were Student's 2-tailed t-test and the correlation coefficient.

Page 111: Phytochemistry and pharmacology of plants from the ginger family

87

3.3 Results and Discussion

3.3.1 Experiment 1

This experiment aimed to determine the most efficient extraction method for gingerols. Hot

Soxhlet extraction was compared with maceration/sonication at room temperature (23°C)

using two different solvents, water and ethanol. The biological activity in the

cyclooxygenase-1 assay and the content of assayed pungent compounds of these extracts are

shown in Table 3-1.

Table 3-1. Effect of extraction method: cyclooxygenase-1 inhibitory activity and content of major pungent compounds of ginger extracts.

All extracts were prepared from fresh rhizome; the rhizome:solvent ratio was 1:4 (w/v). a mean±s.d of 6 injections; b mean±s.d. of 2 injections, c mean±s.d of 4 samples; d mean±s.d of 3 samples; e mean±s.d. of 2 samples; f 75% ethanol-25% water; g 75% water-25% ethanol. CPMA/K: counts per minute. * Two-tailed t-test for COX-inhibitory effect, sample v control.

Cold EtOH Hot EtOH EtOH Controlf

Cold Water Hot Water Water Controlg

6-gingerol (mg/mL)

0.2922 ±0.0005a

0.2841 ±0.0005b

0.1913 ±0.012b

0.2398 ±0.0001b

8-gingerol (mg/mL)

0.0842 ±0.0005a

0.0829 ±0.0006b

0.0491 ±0.0013b

0.0356 ±0.0001b

10-gingerol (mg/mL)

0.0868 ±0.0036a

0.0810 ±0.0003b

0.0402 ±0.0016b

0b

6-shogaol (mg/mL)

0a 0.0282 ±0.0000b

0b 0.0476 ±0.0001b

CPMA/K 12791 ±2199c

12225 ±942c

58496 ±9886c

51878 ±2595d

41476 ±1832d

78417 ±4502e

Inhibition of Cox-1 (%)

78.13±17.19 79.10±7.71 0±16.90 33.84±5.00 47.11±4.42 0±5.74

p-value* 0.002 0.002 0.038 0.036

All extracts demonstrated statistically significant inhibition of cyclooxygenase-1 in the

bioassay (two-tailed t-test; Table 3-1). The inhibitory effect ranged from approximately 34%

for the cold aqueous extract to almost 80% for both ethanolic extracts, compared with

controls. The data show that the choice of solvent was a stronger determinant of biological

Page 112: Phytochemistry and pharmacology of plants from the ginger family

88

activity than was the extraction method. The cold ethanolic extract showed significantly

more potent inhibitory activity than did the cold aqueous extract (p= 0.00003), while the hot

Soxhlet ethanolic extract was also significantly more potent than the hot Soxhlet aqueous

extract (p= 0.00022). Overall, the ethanolic extracts were significantly more potent in terms

of cyclooxygenase-1 inhibition than were the aqueous extracts (p= 0.00002).

It is interesting to note that while the extraction method appeared to have little effect on the

inhibitory activity when ethanol was used as the solvent, the aqueous hot Soxhlet extract was

significantly more active than the cold aqueous extract (47% vs 34% inhibition, p= 0.00653).

The [6]-gingerol content of the extracts showed a very high correlation to the inhibitory

activity in the cyclooxygenase-1 assay (r= 0.976). The [8]-gingerol content also showed

considerable correlation to the inhibitory activity (r= 0.891).

The amount of [10]-gingerol in the ethanolic extracts was approximately twice the amount

measured in the cold aqueous extract, while no [10]-gingerol was identified in the hot

Soxhlet extract. Being less polar than [6]- and [8]-gingerol, smaller amounts of [10]-gingerol

would be expected to be extracted in an aqueous solvent. It is possible that what small

amount of [10]-gingerol may have been extracted was degraded by the heat in the hot

Soxhlet extraction process.

It is noteworthy that [6]-shogaol was identified only in the hot Soxhlet extracts. This is in

agreement with the notion that [6]-shogaol is a degradation product of [6]-gingerol that

forms when [6]-gingerol is exposed to excessive heat (Connell & Sutherland, 1969; He et al.,

1998; Zhang et al., 1994). The present data also suggest that 6-shogaol is in fact absent from

fresh ginger rhizome.

Based on the findings of this experiment, it was decided to employ maceration with

sonication at ambient temperature in subsequent experiments, since hot Soxhlet extraction is

considerably more time consuming and was not found to provide additional extraction

efficiency.

Page 113: Phytochemistry and pharmacology of plants from the ginger family

89

3.3.2 Experiment 2

This experiment aimed to determine what extraction medium would yield the most active

extract in terms of inhibitory activity in the cyclooxygenase-1 bioassay. Water and ethanol

were identified as potential solvents because of their widespread use and acceptance as

solvents for herbal medicines. Five different extraction media were compared: ethanol

(96%), 75% ethanol-25% water, 50% ethanol-50% water, 25% ethanol-75% water, and

water (100%). All extractions were carried out as macerations with sonication at room

temperature (23°C) as described earlier.

The biological activity in the cyclooxygenase-1 assay and the content of assayed pungent

compounds of these extracts is shown in Table 3-2.

Table 3-2. Effect of extraction solvent: cyclooxygenase-1 inhibitory activity and content of major pungent compounds of ginger extracts.

All extracts were prepared from fresh rhizome; the rhizome:solvent ratio was 1:4 (w/v). a mean±s.d of 2 injections; b mean±s.d of 3 samples; c mean±s.d of 2 samples; d control values calculated on the basis of values for ethanol and water; CPMA/K: counts per minute. * Two-tailed t-test for COX-inhibitory effect, sample v control.

EtOH (96%) 75%EtOH+

25%Water 50%EtOH+ 50%Water

25%EtOH+ 75%Water

100% Water mm

6-gingerola (mg/mL)

0.3701±0.0005 0.3211±0.0028 0.3003±0.0008 0.2744±0.0016 0.1772±0.0007

8-gingerola (mg/mL)

0.0648±0.0008 0.0577±0.0006 0.0482±0.0008 0.0440±0.0010 0.0155±0.0002

10-gingerola (mg/mL)

0.1076±0.0009 0.0941±0.0003 0.0754±0.0005 0.0570±0.0001 0.0136±0.0002

CPMA/Kb 3396±390 6609±993 12507±2270 21353±5213 23030±5348

Control value 41197±6165b 40303d 39410d 38516d 37622±1951c

Inhibition (%) 91.76±11.49 83.60±15.02 68.26±18.16 44.56±24.41 38.78±23.22

p-value* 0.009 0.009 0.008 0.014 0.029

All extracts demonstrated statistically significant inhibition of COX-1 in the bioassay (two-

tailed t-test; Table 3-2). The ethanol content of the solvent was highly correlated to

inhibitory activity (r = 0.983) as well as [6]-, [8]- and [10]-gingerol content (r= 0.956, 0.939

Page 114: Phytochemistry and pharmacology of plants from the ginger family

90

and 0.971, respectively). This clearly shows that pure ethanol is a superior solvent to hydro-

ethanolic mixtures with respect to COX-1 inhibitory activity of the resulting extracts, as well

as in terms of gingerol content of these extracts.

Both [6]-, [8]- and [10]-gingerol content of the extracts showed high correlation to inhibitory

activity in the bioassay (r = 0.904, 0.889 and 0.941, respectively). It is not possible from this

experiment to determine to what extent each of these compounds contributes to the

inhibitory activity. Kiuchi and colleagues found the IC50 values of [6]-, [8]- and [10]-

gingerol to be 4.6, 5.0 and 2.5µM, respectively, against prostaglandin synthase

(cyclooxygenase), suggesting that [10]-gingerol may be the most potent of the three

gingerols (Kiuchi et al., 1992). If this is the case, [10]-gingerol may contribute significantly

to the inhibitory activity of the extracts, despite its relative low concentration compared with

the major gingerol, [6]-gingerol.

The inhibitory activity of more than 90% recorded for the ethanolic extract was very

promising and encouraging for future studies.

The gingerol content of the extracts is shown in Fig. 3-1. [6]-Shogaol was not detected in any

of the extracts, a further confirmation that this compound does not occur in fresh ginger (see

above).

Fig. 3-1. Gingerol content (mg/mL extract) of ginger extracts.

Page 115: Phytochemistry and pharmacology of plants from the ginger family

91

In Table 3-3, the ratios of [6]-gingerol : [8]-gingerol : [10]-gingerol in the extracts are given

and compared to values from the literature.

Table 3-3. Gingerol ratios in ginger extracts compared with values from the literature.

a (Zhang et al., 1994); b (Chen et al., 1986b); c (Connell & Sutherland, 1969)

[6]-gingerol:[8]-gingerol:[10]-gingerol

Ethanol (96%) 5.7 : 1 : 1.7

75% Ethanol + 25% Water 5.6 : 1 : 1.6

50% Ethanol + 50% Water 6.2 : 1 : 1.6

25% Ethanol + 75% Water 6.2 : 1 : 1.3

100% Water 11.4 : 1 : 0.9

Fresh ginger, Hawaiia 7 : 1 : 2

Freeze-dried ginger, Taiwanb 7 : 1 : 1.4

Ginger oleoresinc 4.3 : 1 : 2.4

From Table 3-3 it can be seen that the ratio between the three gingerols is relatively stable

when the extraction medium contains at least 25% ethanol. In pure water, however, the ratio

is shifted significantly in favor of the most polar of the compounds, [6]-gingerol.

The results obtained in this experiment clearly show that of the extraction media tested,

ethanol (96%) was the most effective. The ethanol extract showed the greatest inhibitory

activity in the COX-1 assay and had the highest concentration of the putative active

compounds.

In conclusion, this study found that fresh ginger rhizome extracted by maceration with

sonication at room temperature, using ethanol (96%) as the extraction medium, resulted in

extracts with a high concentration of pungent gingerols and potent inhibitory activity in the

cyclooxygenase-1 bioassay.

Page 116: Phytochemistry and pharmacology of plants from the ginger family

92

4. GINGEROL CONTENT OF SEVENTEEN GINGER (ZINGIBER OFFICINALE) CLONES

4.1 Introduction

As outlined in Chapter 2, the main pungent compounds in fresh ginger are a series of

homologous phenolic ketones known as gingerols. The major gingerol is [6]-gingerol, while

[8]- and [10]-gingerol occur in smaller quantities. The gingerols are thermally unstable and

are converted under high temperature to [6]-, [8]- and [10]-shogaol (after shoga, the

Japanese word for ginger) (He et al., 1998). Shogaols, which are more pungent than

gingerols, are the major pungent compounds in dried ginger rhizome. Both gingerols and

shogoals have pharmacological activity including inhibiting COX in vitro (refer to Chapter

2).

In Australia, the most widely grown ginger cultivar is ‘Queensland’, and it is estimated that

40% of the world’s confectionary ginger products are prepared from this cultivar (Smith et

al., 2004b). The origin of this cultivar remains uncertain. Various authors have suggested

that it arrived in Australia from the Cochin coast of India (Connell, 1986), from Fiji

(Leverington, 1975), or from China in the early 1900s (Miles, 1980).

This chapter reports on the analysis of 17 ginger clones grown in Eastern Australia,

including commercial cultivars and 12 experimental tetraploid clones derived from

‘Queensland’, and the quantification by HPLC of the major pungent phenolic compounds,

viz. gingerols and shogaols. The objectives of this study were to explore the variability of

Australian ginger clones in terms of their content of pungent phenolic compounds with a

view to identify one or more high-yielding clones as candidates for commercial cultivation

for flavour or pharmaceutical use.

Page 117: Phytochemistry and pharmacology of plants from the ginger family

93

4.2 Materials and Methods

4.2.1 Plant materials

Seventeen clones of ginger (Table 4-1) were obtained from the Queensland Department of

Primary Industries & Fisheries, Maroochy Research Station at Nambour, Queensland. They

included two selections of the cultivar ‘Queensland’, which is grown commercially in

Queensland, the cultivars ‘Jamaican’, ‘Brazilian’ and ‘Canton’, which were introduced to

Queensland between 1970 and 1972 for cultivar evaluation studies, and twelve experimental

clones developed at the Maroochy Research Station at Nambour, including the newly

released cultivar ‘Buderim Gold’ (Smith & Hamill, 2002). The experimental clones were

obtained after in vitro colchicine treatment of shoots of diploid (2n=22) ‘Queensland’ parent

material provided by J. Roscoe. They were confirmed as solid tetraploids except for one

(Z30), which proved to be a periclinal chimera with both diploid and tetraploid tissues

(Smith et al., 2004b).

Page 118: Phytochemistry and pharmacology of plants from the ginger family

94

Table 4-1. Ginger clones studied, their genotype and origin.

Ploidy: *: confirmed as solid tetraploids by flow cytometry; $: chimera with both diploid and tetraploid tissue sectors; ^: presumed to be tetraploid from stomatal measurements; #: unknown but presumed to be diploid. BGL = Buderim Ginger Ltd.

ID Genotype Cultivar name Origin

Z22 Tetraploid^ (Unnamed) Derived from ‘Queensland’ (Selection 1) by colchicine treatment

Z23 Tetraploid^ (Unnamed)

Z24 Tetraploid* (Unnamed)

Z25 Tetraploid^ (Unnamed)

Z26 Tetraploid* ‘Buderim Gold’

Z27 Tetraploid* (Unnamed)

Z28 Tetraploid^ (Unnamed)

Z29 Tetraploid^ (Unnamed)

Z30 Tetraploid$ (Unnamed)

Z31 Tetraploid* (Unnamed)

Z32 Tetraploid^ (Unnamed)

Z33 Tetraploid* (Unnamed)

Z44 Diploid ‘Queensland’ (Selection 1)

Selected by J. Roscoe, BGL

Z45 Diploid ‘Queensland’ (Selection 2)

Selected by L. Palmer, BGL

Z46 Diploid ‘Jamaican’ Imported from Jamaica

Z47 Diploid# ‘Brazilian’ Imported from Brazil

Z58 Diploid ‘Canton’ Imported from China

The clones were grown from rhizome stock in raised, outdoor beds under uniform, irrigated

conditions at Southern Cross University, Lismore, New South Wales (28° 49’ S, 153° 18’ E)

for approximately eight months.

Page 119: Phytochemistry and pharmacology of plants from the ginger family

95

4.2.2 Sample preparation

Fresh rhizomes were washed and a cylindrical sample was taken from the thickest part of the

rhizome using an apple corer. The epidermis was removed and the sample cut into cubes

approximately 1.5 × 1.5 × 1.5 mm. A five-gram sample was placed in a large centrifuge tube

to which twice the sample mass of 99% ethanol was added. The preparation was sonicated

for 20 minutes (Ultrasonic Cleaner 50 Hz, Unisonics Pty Ltd, Manly Vale, Australia) and

subsequently centrifuged for 5 minutes at 4000 rotations per minute (Hettich Universal 16A

Centrifuge, Tuttlingen, Germany). The supernatant was transferred to a brown glass vial

using a transfer pipette, stored at 4° C and filtered through a Zymark/Millipore Automation

Certified Filter (Glassfiber APFB 1.0 µm prefilter, hydrophilic PTFE membrane 0.45 µm)

before being injected onto the HPLC column.

In order to take account of variation between individual plants, three samples from different

rhizomes were prepared for each clone.

4.2.3 HPLC methods

The same extracts were analyzed by HPLC on two occasions, approximately 5 months apart,

during which period they were stored at 4° C. Measurement reproducibility was determined

by repeated analyses of a mixed standard solution (see below).

Method 1. The first (baseline) reversed-phase HPLC analysis of the extracts was performed

on an Agilent (Palo Alto, California) 1100 HPLC fitted with a HP LiChrospher 100 RP-18e

(5µm) column. Mobile phase A consisted of HPLC-grade water obtained from an in-house

Milli-Q system (Waters, Milford, MA), mobile phase B consisted of HPLC-grade

acetonitrile (EM Science, Gibbstown, NJ); both contained 0.05% trifluoroacetic acid (Sigma-

Aldrich, St. Louis, MO). The gradient eluting mobile phase was A:B (70:30, v/v) to A:B

(10:90, v/v) over 20 min followed by A:B (10:90, v/v) for 10 min. Mobile phase was

pumped at 1.00 mL/min, the column temperature was 40 ºC, and the injection volume was

10.0 µL. Data were collected using a UV/visible diode array detector collecting absorption

spectra from 200 to 400 nm with quantification performed at 228 nm.

Page 120: Phytochemistry and pharmacology of plants from the ginger family

96

Method 2. The second analysis (at 5 months) was carried out on an Agilent 1100 Series

LC/MSD system fitted with a Phenomenex (Torrance, California) Luna® 5µ (150 x 4.6mm)

C-18 column. Mobile phase A and mobile phase B were as described above. The gradient

eluting mobile phase was A:B (70:30, v/v) to A:B (10:90, v/v) over 20 min followed by A:B

(10:90, v/v) for 5 min. Mobile phase was pumped at 0.40 mL/min, the column temperature

was 40 ºC, and the injection volume was 10.0 µL. Data were collected using a UV/visible

diode array detector collecting absorption spectra from 200 to 400 nm with quantification

performed at 228 nm.

Standards. Synthetic standards of [6]-gingerol (99%), [8]-gingerol (99%), [10]-gingerol

(98%), [6]-shogaol (99%) and [8]-shogaol (99%) were obtained from the Department of

Pharmacy, University of Sydney (Sydney, NSW). A known quantity of each standard was

dissolved in 99% ethanol and made up to 10 mL in a volumetric flask. A mixed standard

solution was prepared by combining 100 µL of each of the individual standard solutions of

[6]-, [8]-, and [10]-gingerol, and [6]- and [8]-shogaol. A ten-fold dilution of this mixed

standard was employed in the analyses. It contained the standard compounds in

concentrations ranging from 15 to 25 µg/mL. Identification of gingerols and shogaols in

samples was based on comparisons of retention time and UV-spectra with the standards.

Quantification was based on standard curves prepared with pure standards.

4.2.4 Statistical analysis

Statistical analyses were performed using SPSS (Chicago, Illinois) for Windows Release

11.0.0. The mean, standard deviation and coefficient of variance were calculated for

repeated measurement of the standard solution. The mean and standard error were calculated

for the analyses of the samples. Two-factor repeated measure analysis of variance was used

to compare the content of the three gingerols (‘gingerol’ = the within factor) from the 17

clones (‘clone’ = the between factor). The analysis was based on three replicates for each

clone. The assumption of compound symmetry was not met, so tests of effects were adjusted

using the Greenhouse-Geisser method. Pairwise comparisons were also employed to

compare the gingerol content across the clones. The correlation between the different

gingerols was examined by way of scatter plots and Pearson Product-Moment correlations.

The stability over time of the gingerols was explored by repeated measure t-tests. The mean,

Page 121: Phytochemistry and pharmacology of plants from the ginger family

97

standard error and range were calculated for the gingerol content of diploid and tetraploid

clones, respectively.

4.3 Results

4.3.1 Measurement reproducibility

In order to determine the reproducibility of the HPLC measurements, repeated analyses of

the mixed standard solution were carried out by both methods. The analysis by Method 1

showed a high degree of reproducibility when the same sample was injected 9 times over a

5-day period. The coefficient of variance ranged from 2.2% to 6.4% for the five standard

compounds. The reproducibility of the measurements by Method 2 was assessed by

comparing eleven injections of the mixed standard over a 37-hour period. This analysis

showed a very high degree of reproducibility with the coefficient of variance being less than

2% for all compounds.

4.3.2 Quantification of gingerols in test samples

Rhizome extracts of seventeen clones of ginger were analyzed twice, approximately five

months apart. For each clone, three extracts were prepared from different rhizomes. The

mean concentrations of [6]-gingerol, [8]-gingerol and [10]-gingerol in three rhizomes of each

clone were calculated from the peak areas obtained at 228 nm. A typical HPLC trace is

shown in Fig. 4-1.

Page 122: Phytochemistry and pharmacology of plants from the ginger family

98

min16 18 20 22 24 26 28 30 32

mAU

0

50

100

150

200

250

300

[6]-g

inge

rol (

17.3

4)

[8]-g

inge

rol (

21.2

4)

[10]

-gin

gero

l (24

.80)

Fig. 4-1. Typical HPLC chromatogram of ethanolic extract of fresh ginger rhizome (Z44) obtained with HPLC Method 2.

The concentrations of gingerols in freshly extracted ginger are shown in Fig. 4-2. The most

abundant gingerol in all clones was [6]-gingerol, which occurred at concentrations ranging

from 132 to 339 µg per gram fresh rhizome (mean 215±54 µg/g). [8]-Gingerol and [10]-

gingerol occurred in lower concentrations; [8]-gingerol ranged from 40 to 177 µg/g (mean

75±31), while [10]-gingerol ranged from 37 to 148 µg/g (mean 73±26). The cultivar

‘Jamaican’ (Z46) had the highest total gingerol content of any clone (664 µg/g). Neither [6]-

nor [8]-shogaol was identified in any of the samples (freshly prepared or stored 5 months).

Page 123: Phytochemistry and pharmacology of plants from the ginger family

99

0

100

200

300

400

500

600

700

22 23 24 25 26 27 28 29 30 31 32 33 44 45 46 47 58

Clone ID (Z no.)

mic

rogr

am/g

ram

6-gingerol 8-gingerol 10-gingerol

Fig. 4-2. Concentrations of [6]-, [8]- and [10]-gingerol in fresh rhizomes of 17 ginger clones. Values are means of three separate rhizomes. Error bars indicate standard deviation (only positive half shown).

Two-factor repeated measure analysis of variance (Greenhouse-Geisser test) of the clone by

gingerol interaction effect showed that the mean values of [6]-, [8]- and [10]-gingerol varied

significantly across the 17 clones (F=2.335, p=0.01). When the mean total gingerol content

of each clone was compared with the mean value for all clones, only two clones showed a

statistically significant difference from the overall mean. These were the cultivar ‘Jamaican’

(Z46), which contained a significantly higher concentration of gingerols (p=0.002), and one

of the experimental tetraploid clones (Z25), which had a gingerol content that was

significantly lower than the overall mean (p=0.028).

The difference in gingerol content between the clones was also explored by way of pairwise

comparisons on each of the three gingerols. This analysis is summarised in Table 4-2, which

shows the number of significantly (p<0.05) different cases between each clone and the other

16 clones for each of the three gingerol compounds. The pairwise analysis confirms that the

cultivar ‘Jamaican’ (Z46) is the most outstanding clone in terms of gingerol content. This is

Page 124: Phytochemistry and pharmacology of plants from the ginger family

100

particularly true in terms of [10]-gingerol concentration, which is significantly higher in

‘Jamaican’ than in all but one other clone (Z31).

Table 4-2. Results of pairwise analyses of 17 ginger clones in terms of [6]-, [8]- and [10]-gingerol content.

Each clone is compared with all other clones and the number of significantly (p<0.05) different cases (max. 16) is shown for each gingerol. Note that Clone Z46 (‘Jamaican’) showed 32 significantly different comparisons, distinguishing it from all other clones.

Clone (Z no.) 22 23 24 25 26 27 28 29 30 31 32 33 44 45 46 47 58

[6]-gingerol 2 2 - 7 2 - 4 2 1 1 2 2 1 8 8 2 2

[8]-gingerol 1 5 1 3 2 - 1 - 1 6 - 2 - 2 9 2 2

[10]-gingerol 1 4 1 2 1 2 1 1 1 2 1 1 1 1 15 1 1

Total number 4 11 2 12 5 2 6 3 3 9 1 5 2 11 32 5 5

4.3.3 Correlation between gingerols

The correlations between the mean concentrations of [6]-, [8]- and [10]-gingerol in the

freshly prepared extracts of 17 clones are illustrated by way of scatter plots in Fig. 4-3. A

linear relationship between the concentrations of [6]-, [8]- and [10]-gingerol is apparent. The

cultivar ‘Jamaican’ (Z46) stands out from the others by containing higher concentrations of

[8]- and [10]-gingerol relative to [6]-gingerol (Fig. 4-3A-B).

Page 125: Phytochemistry and pharmacology of plants from the ginger family

101

6-gingerol

400300200100

8-gi

nger

ol

180

160

140

120

100

80

60

40

20

'Jamaican'

A

6-gingerol

400300200100

10-g

inge

rol

160

140

120

100

80

60

40

20

B 'Jamaican'

8-gingerol

18016014012010080604020

10-g

inge

rol

160

140

120

100

80

60

40

20

C 'Jamaican'

Fig. 4-3. Scatter plots illustrating the correlation between concentrations (µg/g) of [6]- and [8]-gingerol (A), [6]- and [10]-gingerol (B), and [8]- and [10]-gingerol (C) in fresh ginger rhizomes. Based on HPLC data from 17 different clones.

Page 126: Phytochemistry and pharmacology of plants from the ginger family

102

The Pearson Product-Moment correlations between the concentrations of [6]-, [8]- and [10]-

gingerol in the rhizomes at zero and five months were calculated (Table 4-3). Since the

scatter plots clearly show a positive correlation between the different gingerols a one-tailed

test for significance was chosen. The highly significant correlations confirm the strong

positive correlation between the three gingerols.

Table 4-3. Pearson Product-Moment correlations between the concentration of gingerols in fresh rhizomes of seventeen ginger clones assayed by HPLC at zero and five months.

*: p<0.0005 (one-tailed)

[6]-gingerol [8]-gingerol

0 months 5 months 0 months 5 months

[8]-gingerol 0.882* 0.850*

[10]-gingerol 0.880* 0.803* 0.959* 0.914*

4.3.4 Gingerol ratios

The ratio of [6]-gingerol:[8]-gingerol:[10]-gingerol was calculated for the 17 clones based on

the HPLC data. The mean ratio of [6]-gingerol:[8]-gingerol:[10]-gingerol was 3:1:1 across

the clones, but some clones deviated considerably from this ratio, for example Z29

(4.4:1:1.2) and Z46 (‘Jamaican’) (1.9:1:0.8). Since both [8]- and [10]-gingerol possess

considerable pungency (albeit less than [6]-gingerol) (Govindarajan, 1982b), the relatively

high levels of [8]- and [10]-gingerol present in ‘Jamaican’ in addition to the high

concentration of [6]-gingerol make this clone by far the most pungent of the 17 clones

assayed. This was confirmed organoleptically.

Page 127: Phytochemistry and pharmacology of plants from the ginger family

103

4.3.5 Stability of gingerols

Ethanolic extracts were assayed twice approximately 5 months apart. Between analyses the

extracts were refrigerated at 4º C. The concentrations of [6]-, [8]- and [10]-gingerol at zero

and five months were compared by way of repeated measure (paired samples) t-tests. The

mean concentrations of [6]- and [8]-gingerol did not change significantly in the 17 clones

over the 5-month period, and these compounds therefore appear to be stable in ethanolic

solution at 4º C for this period of time. The concentration of [10]-gingerol, however, showed

a small (5%) but statistically significant (p=0.01) decrease over the same period.

4.3.6 Gingerol content of tetraploid clones

The mean, standard error and ranges for gingerol concentrations in two commercial

selections of the diploid ‘Queensland’ cultivar and twelve experimental tetraploid clones are

shown in Table 4-4. ‘Queensland’ (Selection 1) is the parent clone from which the

tetraploids were generated by colchicine treatment. Both diploid and tetraploid clones

displayed considerable variation in gingerol concentrations. On average, the diploid

‘Queensland’ clones contained higher levels of all three gingerols but the differences were

not statistically significant.

Table 4-4. Mean±SE (range) concentrations of gingerols in two

commercial ‘Queensland’ clones and 12 tetraploid clones (µg/g).

[6]-gingerol [8]-gingerol [10]-gingerol

‘Queensland’ clones (n=2)

245.2±77.7 (139.3-339.4)

92.2±49.9 (50.9-176.6)

83.2±39.9 (45.2-147.9)

Tetraploid clones (n=12)

202.8±37.9 (131.6-252.5)

68.1±18.2 (40.2-92.0)

67.7±18.4 (36.6-98.5)

p-value (t-test) 0.058 0.279 0.471

Page 128: Phytochemistry and pharmacology of plants from the ginger family

104

4.4 Discussion

An extensive survey of fresh rhizomes of five diploid ginger cultivars, eleven experimental

tetraploid clones, and one recently released tetraploid clone grown in Australia was

conducted with respect to their content of pungent gingerols and shogaols. Ethanol was

chosen as the extraction solvent, as it is the solvent of choice for herbal medicine

preparations.

4.4.1 Gingerols

The three pungent compounds, [6]-, [8]- and [10]-gingerol, were identified and quantified in

all samples. [6]-Gingerol was the most abundant gingerol in all clones, which is in

accordance with the literature (Bartley, 1995; Chen et al., 1986a; Govindarajan, 1982a). The

mean ratio of [6]-gingerol:[8]-gingerol:[10]-gingerol was 3:1:1 across all 17 clones. A

strong, positive, linear correlation between levels of the three gingerols was found in all

clones, reflecting the close biosynthetic relationship between these compounds.

The mean content of gingerols obtained in the present study are considerably higher than

those found by Bartley in a supercritical CO2 extract of Australian ginger (Bartley, 1995),

but lower than levels reported from other parts of world (Table 4-5). This variability may

reflect genetic differences between clones in different regions or physiological responses to

environmental factors such as climate, soil characteristics or predation, but they may also be

due to differences in extraction and analytical methodologies.

Page 129: Phytochemistry and pharmacology of plants from the ginger family

105

Table 4-5. Literature data on gingerol content of fresh ginger rhizomes.

Values are µg per gram fresh rhizome. n.d. = not determined.

Country (reference)

Solvent/ extraction

Analytical method

[6]-gingerol [8]-gingerol [10]-gingerol

Australia (present study)

Ethanol HPLC 215 75 72

Hawaii (Zhang et al., 1994)

Methanol HPLC 2100 288 533

United States (Hiserodt et al., 1998)

Methylene chloride

HPLC 880 93 120

Taiwan (Young et al., 2002)

Ground; acetate buffer solution (pH 4.0) added

HPLC 806 n.d. n.d.

Australia (Bartley, 1995)

Supercritical CO2

NIES-MS

120 19 24

The cultivar ‘Jamaican’ contained the highest concentration of all three gingerols on a fresh

weight basis and was therefore the most pungent of the clones assayed. It also contained

higher levels of [8]- and [10]-gingerol relative to [6]-gingerol than any other clone.

‘Jamaican’ may thus be suitable for commercial production of highly pungent ginger

rhizomes with potential application in both the pharmaceutical and flavour industries, even

though, eventually, the viability of commercial production of this clone will depend on

biomass yield.

Repeated analyses of ethanolic extracts five months apart showed that [6]- and [8]-gingerol

did not degrade during this period when stored at 4º C. Concentrations of [10]-gingerol

showed a small but statistically significant decrease (5%). It is not known whether [10]-

gingerol is in fact less stable than [6]- and [8]-gingerol under these conditions or this finding

represents a type 1 error resulting from the small sample size.

Page 130: Phytochemistry and pharmacology of plants from the ginger family

106

4.4.2 Shogaols

Neither [6]- nor [8]-shogaol were identified in these samples, which were prepared at

ambient temperature from fresh rhizomes. In contrast, [6]-shogaol was identified in fresh

rhizome extracts prepared by hot Soxhlet extraction (data not shown, refer to Chapter 3).

These findings support the hypothesis that shogaols are not native constituents of fresh

ginger rhizomes but form from gingerols by dehydration as a result of heat treatment or

alkaline or acidic conditions (Connell & Sutherland, 1969; Vesper et al., 1995; Zhang et al.,

1994). Earlier reports of shogaols in fresh ginger extracts analyzed by GC-MS (Bartley,

1995; Bartley & Jacobs, 2000) can probably be explained by the high temperatures samples

are exposed to during this form of analysis, resulting in the formation of shogaols as artifacts

of analysis.

4.4.3 Ploidy

The mean concentrations of all three gingerols were lower for the tetraploid clones than for

the parent diploid ‘Queensland’ cultivar, although the differences were not statistically

significant. This observation is in marked contrast to the findings of Nakasone and

colleagues who reported that tetraploid clones of three Japanese ginger cultivars contained

higher concentrations of total gingerols ([6]-, [8]- and [10]-gingerol) and in particular of

[10]-gingerol than did their parent diploid genotypes (Nakasone et al., 1999).

Comparative quantitative studies of secondary metabolites in diploid versus polyploid

genotypes have been conducted on numerous medicinal plants. Although polyploidy often

appears to result in increased expression of secondary metabolites, this is not always the

case, and the effects of polyploidy are not predictable (Evans, 2002).

The findings of this study do not therefore preclude the possibility of identifying a tetraploid

clone with elevated gingerol biosynthesis. In this context it would be of particular interest to

monitor experimental tetraploid clones derived from the cultivar ‘Jamaican’, which are

currently under development at the Maroochy Research Station.

This study has described the variability of gingerol compounds in Australian commercial and

experimental ginger clones. When combined with agronomic data, the present information

Page 131: Phytochemistry and pharmacology of plants from the ginger family

107

should allow for selection of clones with specific levels of pungency, a characteristic which,

along with the aroma produced by the essential oil, determines the flavour characteristic of

ginger.

Page 132: Phytochemistry and pharmacology of plants from the ginger family

108

5. ESSENTIAL OIL COMPOSITION OF SEVENTEEN GINGER (ZINGIBER OFFICINALE) CLONES

5.1 Introduction

Ginger owes its unique flavour properties to the combination of pungency and aroma. The

pungency is provided by non-volatile phenolic compounds, whereas the essential oil gives

ginger its characteristic aroma. Ginger rhizome yields two primary extracts: oleoresin and

essential (or volatile) oil. The oleoresin is a solvent extract (usually in acetone or ethanol)

containing both essential oil and the phenolic compounds responsible for the pungency of

ginger, chiefly [6]-gingerol and to a lesser extent [8]- and [10]-gingerol. The corresponding

shogaols, which are dehydration products of gingerols formed in heat-treated ginger, are also

found in the oleoresin. Ginger oleoresin is used extensively as a flavouring agent in the food

and beverage industries.

Commercial ginger oil is normally extracted by steam distillation from dried rhizomes.

Typical ginger oil is characterized by a high content of sesquiterpene hydrocarbons, in

particular zingiberene, ar-curcumene, β-bisabolene, and β-sesquiphellandrene, while

important monoterpenoids normally include geranial, neral, and camphene (Lawrence, 1997;

Lawrence, 2000; Martins et al., 2001; Vernin & Parkanyi, 1994). Although these

compounds are characteristic of ‘typical’ ginger oils, the literature clearly shows that ginger

oil composition is highly variable (Asfaw & Abegaz, 1995; Connell & Jordan, 1971;

Ekundayo et al., 1988; Gurib-Fakim et al., 2002; Lawrence, 1995a; Lawrence, 1997;

Lawrence, 2000; Macleod & Pieris, 1984; Martins et al., 2001; Menut et al., 1994; Vernin &

Parkanyi, 1994). Factors such as geographical origin, whether the oil is distilled from fresh

or dried rhizome material, drying process and temperature, and analytical methodology may

all contribute to the disparity of published ginger oil analyses.

Ginger ‘oil’ obtained by supercritical fluid extraction using carbon dioxide is also

commercially available, but this product differs radically from steam-distilled oil due to the

presence of the pungent gingerols and shogaols (Bartley & Jacobs, 2000).

Page 133: Phytochemistry and pharmacology of plants from the ginger family

109

Ginger oil is used in the beverage and fragrance industries, and the world production of

ginger oil was estimated at 100-200 t in 2000 (Weiss, 2002). In the classic work, Perfume

and Flavor Materials of Natural Origin, Arctander described the odour of ginger oil as

“warm, but fresh-woody, spicy and with a peculiar resemblance to orange, lemon, lemon-

grass, coriander weed oil, etc. in the initial, fresh topnotes […the] sweet and heavy undertone

is tenacious, sweet and rich, almost balsamic-floral” (Arctander, 1994).

This chapter reports the essential oil composition (analysed by GC-MS) of the clones of

ginger previously analysed for their gingerol content (refer to Chapter 4). This is the first

comprehensive survey of steam-distilled Australian ginger oils since the early work by

Connell and Jordan (Connell & Jordan, 1971).

5.2 Materials and Methods

5.2.1 Plant materials and distillation

Details of the seventeen commercial and experimental clones of ginger included in this study

were provided in Section 4.2.1 of the previous chapter. Rhizomes were harvested in late July

and stored at ambient temperature for approximately seven weeks before being distilled.

Prior to distillation, unpeeled rhizomes were washed and had any diseased tissue removed

before being chopped into pieces approximately 1 3 7 mm. Equal parts of rhizome

from three different plants were pooled, and approximately 200 g of this material was

hydrodistilled in a Clevenger distillation apparatus for three hours.

5.2.2 Chemical analysis

Oils were analyzed on an Agilent Technologies (Palo Alto, California) 6890/5973 GC-MSD

system using helium as the carrier gas at a constant linear flow velocity of 29 cm/s. Oil

samples (150 µL) were diluted with pentane (1500 µL) and 1 µL of this solution was

injected. The column was a 50 m 0.22 mm capillary column, 1.00 µm film thickness

(BPX-5, SGE Ltd., Melbourne). The split ratio was 25:1. The column oven was

programmed from 70°C to 280°C at a ramp rate of 4°C/min (final hold time 4 minutes), and

Page 134: Phytochemistry and pharmacology of plants from the ginger family

110

the injector temperature was 250°C. Composition values were recorded as percentage area

based on the total ion current chromatogram. Retention indices were determined using a C-8

to C-22 n-alkane mixture. Compound identification was based on comparisons with mass

spectra and retention indices of authentic reference compounds where possible and by

reference to WILEY275, NBS75K and Adams terpene library (Adams, 1995) and published

data. Myrcene, 1,8-cineole, linalool, α-terpineol, β-citronellol, citral, geraniol, (–)-bornyl

acetate, citronellyl acetate, geranyl acetate and (E)/(Z)-nerolidol were obtained from Aldrich

Chemical Co. Inc. (Milwaukee, WI); (–)-borneol and (Z)-nerolidol were obtained from Fluka

Chemie (Buchs, Switzerland); 6-methyl-5-hepten-2-one was obtained from ants

(Formicidae) (Tomalski et al., 1987) and (E,E)-α-farnesene from the peel of ‘Granny Smith’

apples (Pechous & Whitaker, 2004). Germacrene-D and elemol were identified by

comparison with these compounds in authentic clary sage (Salvia sclarea L.) oil

(Anonymous, 2004) and elemi (Canarium luzonicum (Miq.) Asa Gray; Berjé Inc.,

Bloomfield NJ) oil (Villanueva et al., 1993), respectively.

5.2.3 Statistical analysis

Statistical analyses were performed using SPSS (Chicago, Illinois) for Windows Release

11.5. Mean values, standard deviations and ranges were calculated for major oil

constituents. The oil composition of the 17 clones was the subject of principal components

analysis and cluster analysis based on the ten most abundant constituents. The relationship

between neral and geranial was examined by way of a scatter plot and Pearson’s correlation

coefficient.

5.3 Results

5.3.1 Oil composition

The essential oil composition for the 17 clones is shown in Table 5-1. The mean, standard

deviation and range for the 14 most abundant constituents in the 16 homogenous or ‘typical’

clones are shown in Table 5-2, which also gives the percentage content for the atypical oil

from the cultivar ‘Jamaican’ (Z46). The ‘typical’ oils had a mean citral (neral + geranial)

Page 135: Phytochemistry and pharmacology of plants from the ginger family

111

content of 58%, whilst the five major sesquiterpene hydrocarbons typically found in ginger

oil (ar-curcumene, (E,E)-α-farnesene, zingiberene, β-bisabolene and β-sesquiphellandrene

(Lawrence, 1995b)) made up only 17%. In contrast, the oil from ‘Jamaican’ had a

comparatively low citral content (28%) and contained 35% of the main sesquiterpene

hydrocarbons. Some of the major oil constituents are shown in Fig. 5-1.

H3C

curcumene

zingiberene

H

β-sesquiphellandrene

β-bisabolene

(E,E)-α-farnesene

OH

borneol

CH2OH

geraniol geranial neral

CHO

OH

α-terpineol

OH

citronellol

CHO

Fig. 5-1. Volatile constituents from Zingiber officinale essential oil.

Page 136: Phytochemistry and pharmacology of plants from the ginger family
Page 137: Phytochemistry and pharmacology of plants from the ginger family

113

Table 5-2. Content of 14 constituents in essential oils of 16 ‘typical’ clones of ginger and one ‘atypical’ clone, ‘Jamaican’ (Z46).

Mean, standard deviation and range are shown for the 16 clones. All values are percentage content.

16 ‘typical’ clones mean±SD (range)

‘Jamaican’ (Z46)

1,8-cineole 1.52±0.52 (0.39-2.63) 0.79

linalool 1.21±0.20 (0.97-1.55) 1.02

borneol 2.41±0.39 (1.84-2.94) 3.91

α-terpineol 1.81±0.21 (1.49-2.18) 1.13

citronellol 1.93±0.28 (1.48-2.49) 1.09

neral 21.44±1.63 (19.39-26.49) 10.60

geraniol 4.91±1.28 (2.73-7.30) 1.54

geranial 36.50±3.26 (31.29-44.31) 17.51

geranyl acetate 2.02±0.92 (0.52-3.45) 0.26

ar-curcumene 3.24±0.73 (2.43-5.31) 5.72

(E,E)-α-farnesene 3.02±0.68 (2.10-4.30) 4.35

zingiberene 4.82±2.03 (1.86-9.00) 11.24

β-bisabolene 1.51±0.31 (0.97-2.16) 4.05

β-sesquiphellandrene 4.36±0.85 (2.93-5.62) 9.40

5.3.2 Principal components analysis

The percentage composition of the 17 oil samples was subjected to principal components

analysis based on the ten most abundant oil constituents (borneol, neral, geraniol, geranial,

geranyl acetate, ar-curcumene, (E,E)-α-farnesene, zingiberene, β-bisabolene and β-

sesquiphellandrene). This analysis revealed the presence of three components with

eigenvalues exceeding one. These components explained 61.6%, 18.8% and 10.7% of the

variability, respectively. Two components (together explaining 80% of the variability) were

retained for further investigation, and Varimax rotation was performed to assist in the

interpretation (Table 5-3).

Page 138: Phytochemistry and pharmacology of plants from the ginger family

114

Table 5-3. Varimax rotated component matrix for two component solution for the ten most abundant ginger essential oil constituents.

Component

1 2

neral -.96 -.25

geranial -.96 -.15

β-bisabolene .89 .39

borneol .87 -.21

β-sesquiphellandrene .78 .61

ar-curcumene .61 .30

zingiberene .55 .54

geraniol -.23 -.89

geranyl acetate .12 -.87

(E,E)-α-farnesene .45 .74

Percentage of variance explained 49.3% 31.2%

The rotated solution revealed a complex structure in which both components had several

strong loadings (coefficients relating the variables to the components), but some compounds

loaded substantially on both components. Neral, geranial, β-bisabolene, borneol and β-

sesquiphellandrene were strongly associated with component 1, indicating a high degree of

interrelationship (positive or negative) between the concentrations of these compounds.

Similarly, geraniol, geranyl acetate and (E,E)-α-farnesene were strongly associated with

component 2. More broadly, an inverse relationship between levels of citral (geranial +

neral) and the sesquiterpene hydrocarbons (zingiberene, ar-curcumene, β-

sesquiphellandrene, β-bisabolene and (E,E)-α-farnesene) was evident by inspection of the

component plot in Fig. 5-2.

Page 139: Phytochemistry and pharmacology of plants from the ginger family

115

Component 1

1.0.50.0-.5-1.0

Com

pone

nt 2

1.0

.5

0.0

-.5

-1.0

geranialneral

geraniol geranyl acetate

borneol

b-sesquiphellandrenezingiberene

b-bisabolenear-curcumene

(E,E)-a-farnesene

Fig. 5-2. Component plot in rotated space showing Varimax rotated data on two components based on the ten most abundant constituents of ginger essential oil.

5.3.3 Cluster analysis

In order to examine the degree of similarity displayed by the 17 clones in terms of oil

composition, a hierarchical, between-groups linkage, cluster analysis based on the ten most

abundant constituents was performed. This is a multivariate procedure that allows for the

classification of cases (or variables) into groups based on Euclidian distances between cases.

For each constituent, percentage values were rescaled to have a mean value of one, so that all

constituents were equally weighted for the purpose of the analysis, and Euclidian distances

were calculated between pairs of clones.

Fig. 5-3 shows a dendrogram of the 17 essential oils. This diagram confirms the unique

nature of the essential oil from the cultivar ‘Jamaican’ (Z46) when compared with oils from

the other 16 clones. The dendrogram also shows that the oil from the clone Z22 stands out

from the others. This oil had an extremely high citral content (71%). The remaining 15

Page 140: Phytochemistry and pharmacology of plants from the ginger family

116

clones fall into two clusters. The similarity between these two clusters was examined using a

multivariate general linear model, comparing the set of clones in Cluster 1 (Z25, Z32, Z33,

Z27, Z26, Z31, Z30, Z58, Z28, Z29, Z45) and Cluster 2 (Z24, Z44, Z47, Z23). Cluster 1

and Cluster 2 showed a near-significant difference (Wilks’ Lambda = 0.064; F = 5.838; df =

10 and 4; p = 0.052). In terms of individual constituents, six (borneol, geraniol, geranyl

acetate, (E,E)-α-farnesene, β-bisabolene, β-sesquiphellandrene) were significantly different

between clusters at p ≤ 0.05 and four (neral, geranial, ar-curcumene, zingiberene) were not.

Fig. 5-3. Dendrogram of hierarchical cluster analysis of 17 essential oils of ginger.

The diagram shows average linkage (between groups), and values shown along the horizontal axis are Euclidian distances rescaled to an arbitrary scale showing the levels of relative similarity where clusters join.

5.3.4 Citral content

The ginger oils analyzed in the present study had citral contents ranging from 28% in the

‘Jamaican’ cultivar (Z46) to 71% in the tetraploid clone Z22. The mean citral content of the

16 ‘typical’ ginger oils (‘Jamaican’ excluded) was 57.9±4.9% (range: 50.7-70.8%), which

was more than double the corresponding value for ‘Jamaican’ (28%). A similar trend was

Page 141: Phytochemistry and pharmacology of plants from the ginger family

117

evident for geraniol, which is a precursor to citral, with a mean concentration of 4.9%±1.3%

in the ‘typical’ oils compared with 1.5% in ‘Jamaican’.

5.3.5 Neral to geranial ratio

The 17 ginger oils contained neral and geranial in a ratio that was remarkably constant. The

neral to geranial ratio ranged from 0.55 to 0.64, with a mean value of 0.61±0.01. This fixed,

linear relationship between the two isomers, which persisted regardless of the percentage

content of citral, is illustrated in Fig. 5-4. The very strong correlation between the two

compounds is quantified by the Pearson’s correlation coefficient of 0.987 (p<0.001).

It is particularly interesting to note that the neral to geranial ratio also was 0.6 in the cultivar

‘Jamaican’ (Z46), which otherwise yielded an oil that was distinctly different from those of

the other 16 clones.

Neral (%)

302010

Ger

ania

l (%

)

50

40

30

20

10

'Jamaican'

Z22

Fig. 5-4. Scatter plot showing the relationship between the percentage content of the stereoisomers neral and geranial in essential oils from 17 clones of ginger.

Page 142: Phytochemistry and pharmacology of plants from the ginger family

118

5.4 Discussion

Essential oils of 17 clones of Australian ginger were prepared by hydrodistillation of

rhizomes and analyzed by GC-MS. Compared with values in the literature (Asfaw &

Abegaz, 1995; Connell & Jordan, 1971; Ekundayo et al., 1988; Gurib-Fakim et al., 2002;

Lawrence, 1995b; Lawrence, 1997; Lawrence, 2000; Macleod & Pieris, 1984; Martins et al.,

2001; Menut et al., 1994; Vernin & Parkanyi, 1994), all the samples were characterized by a

very high citral content (28% or greater) and a relatively low content of sesquiterpene

hydrocarbons.

5.4.1 Oil composition

There was no distinct difference in the composition of the oils of the tetraploid clones

compared with the diploid parent cultivar ‘Queensland’ or the cultivars ‘Canton’ and

‘Brazilian’. These oils were characterized by very high levels of citral (geranial + neral) (51-

71%) and relatively low levels of the sesquiterpene hydrocarbons characteristic of ginger oil

(Lawrence, 1995b), namely zingiberene (4.8±2.0%), ar-curcumene (3.2±0.7%), β-

sesquiphellandrene (4.4±0.9%), β-bisabolene (1.5±0.3%) and (E,E)-α-farnesene (3.0±0.7%).

These values contrast with those reported in 1971 by Connell and Jordan, whose analysis of

35 Australian ginger oils found citral levels ranging from 4 to 30% and much higher

concentrations of sesquiterpene hydrocarbons (zingiberene 20-28%, ar-curcumene 6-10%, β-

sesquiphellandrene 7-11% and β-bisabolene 5-9% (but no (E,E)-α-farnesene) (Connell &

Jordan, 1971). The results obtained in the present study are better aligned with more recent

analyses of ginger oils from Mauritius (Gurib-Fakim et al., 2002), Sao Tomé e Príncipe

(Martins et al., 2001) and Nigeria (Ekundayo et al., 1988) by being high in citral, relatively

low in the major sesquiterpene hydrocarbons and by containing (E,E)-α-farnesene. In

general, the published values for these compounds vary greatly; whether this is due to true

natural variability or differences in raw material (fresh or dried, time of harvest), distillation

conditions or analytical methodologies is uncertain.

An inverse relationship between levels of citral on the one hand and β-bisabolene, β-

sesquiphellandrene and borneol on the other was demonstrated by principal components

analysis. Likewise, an inverse relationship was shown to exist between geraniol/geranyl

Page 143: Phytochemistry and pharmacology of plants from the ginger family

119

acetate and (E,E)-α-farnesene. Terpenoids are derived from C5 isoprene units in the form of

the diphosphate (pyrophosphate) esters dimethylallyl diphosphate and isopentenyl

diphosphate and are products of the mevalonate pathway (Dewick, 1997). The common

biosynthetic origin of monoterpenoids (such as geranial, neral, geraniol and geranyl acetate)

and sesquiterpenoids (including β-bisabolene, β-sesquiphellandrene and (E,E)-α-farnesene)

provides a possible explanation for the inverse relationship seen between mono- and

sesquiterpenoids, as the pathways to mono- and sesquiterpenes would compete for common

precursors. Similarly, the fact that citral and borneol share the monoterpene precursor

geranyl pyrophosphate could explain the inverse relationship observed between these

compounds.

The oils analysed in this study were prepared from fresh rhizomes seven weeks post-harvest

by a short distillation process and analysed within a short period of time. They may differ

significantly from oils prepared from dried rhizome material by longer distillation processes

followed by prolonged storage under conditions that favour oxidation. It has been reported

that both zingiberene and β-sesquiphellandrene can undergo oxidation to ar-curcumene

(Connell & Jordan, 1971; Govindarajan, 1982a), and high levels of this compound may

therefore be indicative of a degraded oil.

The essential oil from the cultivar ‘Jamaican’ (Z46) differed distinctly in composition from

all the others, primarily by having lower citral content and higher sesquiterpene hydrocarbon

content. This oil contained 28% citral while the total amount of the five main sesquiterpene

hydrocarbons was 35%, which is more than double the mean concentration found in the 16

other clones. The distinctiveness of the oil of ‘Jamaican’ was confirmed by hierarchical

cluster analysis. As reported in Chapter 4, this cultivar was also the most pungent of the 17

clones tested due to its high content of gingerols. The dissimilar essential oil composition

coupled with its high pungency gives ‘Jamaican’ flavour and aroma qualities that are distinct

from the other clones examined.

Cluster analysis identified one other distinct clone (Z22), characterized by extremely high

(71%) citral content, while the remaining clones fell into two clusters. These two clusters

did not appear to reflect breeding history, as the different tetraploid clones produced from the

same parent cultivar (‘Queensland’) were split between the two clusters, as were the other

two cultivars, ‘Brazilian’ and ‘Canton’.

Page 144: Phytochemistry and pharmacology of plants from the ginger family

120

5.4.2 Citral content

The ginger oils analysed in this study contained remarkably high levels of citral (above 50%

in all samples except ‘Jamaican’ and up to 71% in one sample). The highest citral content

described previously in ginger oil was 56% in a sample from Fiji (Smith & Robinson, 1981).

Although Australian ginger has a reputation for its ‘lemony’ aroma, a 1971 study of 35

Australian ginger oils by Connell and Jordan found considerably lower citral levels, ranging

from 4% to 30% (Connell & Jordan, 1971). The same study found only very low citral

levels in oils from other parts of the world, and many other studies have reported very low

citral levels in non-Australian ginger oils (Asfaw & Abegaz, 1995; Humphrey et al., 1993;

Lawrence, 1995b; Lawrence, 2000; Macleod & Pieris, 1984). Conversely, several more

recent investigations have found substantial citral levels in ginger oils from other parts of the

world, for example from the Central African Republic (35%) (Menut et al., 1994), Mauritius

(27%) (Gurib-Fakim et al., 2002), Sao Tomé e Príncipe (22-25%) (Martins et al., 2001),

Nigeria (24%) (Ekundayo et al., 1988) and Tahiti (14-25%) (Lawrence, 1997; Lechat-

Vahirua et al., 1996).

There are several possible, not mutually exclusive, explanations for these somewhat

incongruent findings regarding citral content. One relates to the changes in gas

chromatography technology (especially detection systems) since the early 1970s, which

might explain some of the differences between the results of the current study and those of

Connell and Jordan (Connell & Jordan, 1971). It is also likely that ginger oils distilled from

fresh rhizomes generally have a higher citral content than oils obtained from dried rhizomes.

The notion that sun drying and storage of ginger could cause a loss of citral was put forward

by Govindarajan more than two decades ago (Govindarajan, 1982a), and studies from Sri

Lanka (Macleod & Pieris, 1984) and Nigeria (Ekundayo et al., 1988) have documented citral

losses in ginger oils ranging from 40% to 74%, when rhizomes were dried before distillation.

A significant difference in citral content has also been observed in supercritical fluid extracts

of fresh and dried Australian ginger, which contained 19.9% and 6.2% citral, respectively

(Bartley & Jacobs, 2000). These data would seem to suggest that many ginger oils contain

considerable amounts of citral, provided they are distilled from fresh rhizomes or rhizomes

dried at low temperature. On the other hand, the existence of genuine low-citral genotypes

of ginger is supported by a recent study from India, where the essential oil distilled from

fresh rhizomes contained less than 4% citral (Singh et al., 2005).

Page 145: Phytochemistry and pharmacology of plants from the ginger family

121

A number of other factors could also have contributed to the unexpectedly high citral values

obtained in this study. They include the particular growth period (8 months), the harvest time

(mid-winter), the 7-week delay between harvest and distillation, and the conditions of

cultivation (including climate) at Southern Cross University, which is located some 250 km

south of the commercial ginger-growing area in Australia.

5.4.3 Neral to geranial ratio

Neral (Z-citral) and geranial (E-citral) are stereoisomers occurring in many plants in mixtures

often referred to simply as citral. The neral to geranial ratios found in ginger oils in this

study were remarkably constant and close to 2:3 (mean 0.61±0.01, range 0.55-0.64). The

ratios reported in the literature for ginger oil vary greatly, from 0.3 in a Taiwanese oil

(Lawrence, 1995b) to 4.7 in a Chinese oil (Lawrence, 2000). However, a ratio of 0.6 was

also reported for a Mauritian oil made from fresh rhizomes (Gurib-Fakim et al., 2002), a

Nigerian oil also prepared from fresh rhizomes (Ekundayo et al., 1988) and an oil made from

dried rhizomes from Sao Tomé e Príncipe (Martins et al., 2001), while oils from Madagascar

(Möllenbeck et al., 1997) and the Central African Republic (Menut et al., 1994) were

reported to contain neral to geranial ratios of 0.4 and 0.9, respectively. The data obtained in

the current study support the hypothesis that neral and geranial occur in ginger in a relatively

fixed ratio of approximately 2:3 and calls into question the validity of analyses suggestive of

very different ratios.

Neral and geranial also occur in other plants in a similar ratio. Species of lemongrass

produce essential oil with a citral content typically exceeding 70%. Nine samples of West

Indian lemongrass (Cymbopogon citratus, family Poaceae) oil from Africa, China, New

Zealand and Cuba had neral to geranial ratios ranging from 0.61 (New Zealand) to 0.96

(China) with a mean of 0.76, while four Indian cultivars of East Indian lemongrass (C.

flexuosus) produced oils with a neral to geranial ratio ranging from 0.61 to 0.67 (Lawrence,

2002).

This study has shown that essential oil from the three ginger cultivars ‘Queensland’,

‘Canton’ and ‘Brazilian’ and twelve tetraploid clones derived from ‘Queensland’ had a high

degree of compositional similarity and were characterized by a very high citral content.

Tetraploid clones were shown to have retained the characteristics of the parent cultivar in

Page 146: Phytochemistry and pharmacology of plants from the ginger family

122

terms of essential oil composition. In contrast, the cultivar ‘Jamaican’ produced an oil that

was distinctly different, characterized by a lower citral content and higher levels of

sesquiterpene hydrocarbons. The distinctive aroma profile of this cultivar, which also

contains significantly higher levels of pungent gingerols, should make it of commercial

interest to the flavour and fragrance industries.

Page 147: Phytochemistry and pharmacology of plants from the ginger family

123

6. PHARMACOLOGICAL ACTIVITY OF ZINGIBERACEAE SPECIES

6.1 Introduction

Forty-one taxa were screened for inhibition of prostaglandin E2 (PGE2) production. Some of

the samples were also screened for antioxidant activity in the oxygen radical absorbance

(ORAC) assay, for inhibition of nitric oxide production, and for modulation of natural killer

cell activity. This work should be regarded as a preliminary survey, and it was constrained

by the cost of some of the assays.

Prostaglandin E2 is a largely pro-inflammatory eicosanoid produced from arachidonic acid

by the action of cyclooxygenase (COX). COX, in particular the isoform COX-2, is an

important pharmacological target for the mitigation of inflammation, which is an underlying

pathological mechanism in many chronic diseases including arthritic disorders,

cardiovascular disease, diabetes, some types of cancer, and osteoporosis (Libby, 2007).

Salicylates and other non-steroidal anti-inflammatory drugs target COX, and many plant

extracts and compounds have been shown to inhibit COX and the generation of PGE2

(Calixto et al., 2003; Calixto et al., 2004). Prostaglandins and COX were described in more

detail in Chapter 1.

Reactive oxygen species and oxidative stress have been linked to a large number of

pathologies including inflammation, cancer, cardiovascular disease, Parkinson’s disease and

Alzheimer’s disease, and the importance of antioxidant defenses (both endogenous and

exogenous) is well recognised (Gilbert, 1999). Higher plants are rich sources of antioxidant

compounds (well known examples are carotenoids and polyphenols), and plant extracts with

potent antioxidant activity are of interest as sources of potential therapeutic substances.

Nitric oxide (NO) is a free radical with numerous functions in both normal and pathological

processes. Nitric oxide is produced by the oxidation of L-arginine by inducible nitric oxide

synthase (iNOS) and is involved in pathological processes associated with an overproduction

of nitric oxide including inflammatory disorders of the joint, gut and lungs. Nitric oxide

functions as a messenger molecule in the immune system, and many immune cells produce

Page 148: Phytochemistry and pharmacology of plants from the ginger family

124

and respond to nitric oxide (Morikawa et al., 2005; Sharma et al., 2007; Tripathi et al.,

2007). Nitric oxide and iNOS play key roles in the up-regulation of the inflammatory

response, and nitric oxide mediates many of the destructive effects of interleukin-1 (IL-1)

and tumour necrosis factor-alpha (TNF-α) on cartilage in osteoarthritis. The inhibition of

nitric oxide and iNOS therefore represent an attractive therapeutic target in inflammatory

conditions (Cuzzocrea, 2006; Sharma et al., 2007; Vuolteenaho et al., 2007).

Inhibition of nitric oxide production and/or iNOS expression has been demonstrated for a

number of medicinal plants and plant compounds in recent years, for example Andrographis

paniculata (Batkhuu et al., 2002), Salvia miltiorrhiza (Jang et al., 2003), Hypericum

perforatum (Di Paola et al., 2007), Citrus reticulata (Jung et al., 2007) as well as the reishi

mushroom (Ganoderma lucidum) (Woo et al., 2005). It is widely considered that this

property likely contributes to the anti-inflammatory activity of these plants.

Natural killer (NK) cells are specialised bone marrow-derived lymphocytes that form part of

the innate immune system. NK cells have the ability to lyse virally infected and tumour

cells. They comprise approximately 5-20% of the peripheral lymphocyte population in the

peripheral blood, liver and spleen and are present at lower frequencies in the lymph nodes,

bone marrow, and thymus. The anticancer activity of NK cells in the form of tumour

rejection has been well established in a variety of animal models of cancer. In addition to

their anti-tumour and anti-viral activity, NK cells are also involved in other

pathophysiological processes such as inflammation and autoimmunity, and stimulated NK

cells can produce cytokines including interferon-γ (IFN-γ), tumour necrosis factor-α (TNF-

α), and granulocyte-macrophage colony stimulating factor (GM-CSF) (Wallace & Smyth,

2005; Yokoyama et al., 2004).

Little has been published on the effects of medicinal plant extracts on NK cell activity.

Medicinal plants for which an ability to increase NK cell activity has been demonstrated

include garlic (Allium sativum) (Abuharfeil et al., 2001), Viscum album (Schink, 1997),

Echinacea spp. (Currier & Miller, 2000; Zhai et al., 2007) and Nigella sativa (Abuharfeil et

al., 2001).

Page 149: Phytochemistry and pharmacology of plants from the ginger family

125

6.2 Materials and Methods

6.2.1 Plant materials

A total of 53 samples representing rhizome extracts of 43 Zingiberaceae taxa were included

in this study, although not all taxa were screened in all assays. Plant material was obtained

from a variety of sources. Nineteen samples were originally obtained from the live

collection in the George Brown Darwin Botanic Garden, Darwin, Northern Territory

(latitude S 12° 26'; longitude E 130° 50'), and a number of these were cultivated in pots at

Southern Cross University, Lismore, New South Wales (latitude S 28° 49'; longitude E 153°

18') before harvest of the rhizome material. Twelve samples were obtained from the living

collection in the Flecker Botanic Gardens, Cairns, Queensland (latitude S 16° 54'; longitude

E 145° 45'). Z129 Alpinia arctiflora, Z127 Alpinia modesta, Z126 Hornstedtia scottiana,

Z128 Pleuranthodium racemigerum, and Z125 Tapeinochilos ananassae were collected in

the wild in the Wet Tropics region of North Queensland under a permit issued by the

Queensland Environmental Protection Agency (voucher specimens were deposited in the

Queensland Herbarium; voucher number AQ736264, AQ736265, AQ736255, AQ736263,

AQ736256, respectively). Z107 Etlingera australasica was obtained from a tub specimen

grown in Brisbane, Queensland from wild stock collected at Bog Hole Creek, Silver Plains,

Queensland (latitude S 13° 42'; longitude E 143° 29') by Paul Forster of the Queensland

Herbarium (PIF30425 (BRI)). Z45 Zingiber officinale ‘Queensland’ rhizome stock was

obtained from the Queensland Department of Primary Industries & Fisheries Maroochy

Research Station at Nambour, Queensland (latitude S 26° 37’; longitude E 152° 57’) and

grown in raised beds at Southern Cross University. J-2 and J-4-4 Zingiber officinale

‘Jamaican’ were grown at the Maroochy Research Station. Z108/9 Zingiber officinale was

obtained from a produce store in Lismore, NSW. Twelve samples were obtained from the

living collections of the Medicinal Plant Garden and the Lismore campus grounds of

Southern Cross University.

Table 6-1 provides a list of the samples, their origin and methods of extraction.

Page 150: Phytochemistry and pharmacology of plants from the ginger family

126

Table 6-1. Zingiberaceae samples screened for biological activity.

ID Taxon Source Extraction method±

Z129$ Alpinia arctiflora (F. Muell.) Benth. Danbulla State Forest, c. 18 km WSW of Gordonvale, North Qld. (Qld Herbarium AQ736264)

3

Z13^ Alpinia caerulea (R. Br.) Benth. Darwin Botanic Garden (D94418) 1 Z102# Alpinia caerulea (R. Br.) Benth. Southern Cross University (NCM03-029) 2 Z49^ Alpinia calcarata Roscoe Southern Cross University 1 Z07^ Alpinia galanga (L.) Swartz Darwin Botanic Garden (D94418) 1 Z103# Alpinia galanga (L.) Swartz Southern Cross University (NCM04-002) 2, 5 Z111 Alpinia luteocarpa Elmer Flecker Botanic Gardens, Cairns 3 Z48^ Alpinia malaccensis Roscoe Southern Cross University 1 Z127$ Alpinia modesta Schumann Nyleta Creek, 16 km NNE of Tully, North

Qld. (Qld Herbarium AQ736265) 3

Z08* Alpinia mutica Roxb. Darwin Botanic Garden (D950346) 1 Z11* Alpinia purpurea ‘Eileen McDonald’ Darwin Botanic Garden (D93105) 1 Z53^ Alpinia spectabile Griff. ‘Giant Orange’ Southern Cross University 1 Z52^ Alpinia zerumbet (Pers.) B.L. Burtt & R.

M. Sm. Southern Cross University 1

Z104#+ Boesenbergia rotunda (L.) Mansf. Southern Cross University (NCM04-121) 2, 6 Z03^ Costus barbatus Suess. Darwin Botanic Garden (D971488) 1 Z112 Costus barbatus Suess. Flecker Botanic Gardens, Cairns 3 Z113 Costus leucanthus Maas Flecker Botanic Gardens, Cairns 3 Z114 Costus malortieanus H. Wendl. Flecker Botanic Gardens, Cairns 3 Z14^ Costus productus Gleason ex Maas Darwin Botanic Garden (D971499) 1 Z116 Costus productus Gleason ex Maas Flecker Botanic Gardens, Cairns 3 Z115 Costus pulverulentus C. Presl Flecker Botanic Gardens, Cairns 3 Z15* Costus tappenbeckianus J. Braun et K.

Schum. Darwin Botanical Garden (D971491) 1

Z59# Curcuma australasica J.D. Hook Southern Cross University (NCM03-036) 1 Z101# Curcuma australasica J.D. Hook Southern Cross University (NCM03-036) 2, 5 Z117 Curcuma cordata Wallich Flecker Botanic Gardens, Cairns 3 Z18^ Curcuma longa L. Darwin Botanic Garden (D970132) 1 Z106# Curcuma longa L. Southern Cross University (NCM04-013) 2, 5 Z118 Curcuma parviflora Wallich Flecker Botanic Gardens, Cairns

(NCM05-040) 3

Z12^ Elettaria cardamomum (L.) Maton Darwin Botanic Garden (D95162) 1 Z107 Etlingera australasica (R.M. Smith) R.M.

Smith Brisbane (PIF30425 (BRI)) 2, 5

Page 151: Phytochemistry and pharmacology of plants from the ginger family

127

Z04* Etlingera elatior (Jack) R.M. Smith ‘Burma Torch’

Darwin Botanic Garden (D95335) 1

Z10^ Hedychium coronarium Koenig Darwin Botanic Garden (D960188) 1 Z126$ Hornstedtia scottiana (F.Muell.)

Schumann Nyleta Creek, 16 km NNE of Tully, North Qld. (Qld Herbarium AQ736255)

3

Z05* Kaempferia galanga L. Darwin Botanic Garden (D970234) 1 Z120 Kaempferia rotunda L. Flecker Botanic Gardens, Cairns 3 Z128$ Pleuranthodium racemigerum (F. Muell.)

R.M. Smith Gillies Range, c. 13 km NE of Yungaburra, North Qld. (Qld Herbarium AQ736263)

3

Z20* Pleuranthodium racemigerum (F. Muell.) R.M. Smith

Darwin Botanic Garden (D971598) 1

Z121 Renealmia cernua (Sw. ex Roem. & Schult.) J.F. Macbr.

Flecker Botanic Gardens, Cairns 3

Z01* Scaphochlamys biloba (Ridley) Holttum Darwin Botanic Garden (D950355) 1 Z16* Scaphochlamys kunstleri (Baker)

Holttum Darwin Botanic Garden (D950403) 1

Z122 Scaphochlamys kunstleri (Baker) Holttum

Flecker Botanic Gardens, Cairns 3

Z09* Tapeinochilos ananassae (Hassk.) K. Schum.

Darwin Botanic Garden (D971993) 1

Z125$ Tapeinochilos ananassae (Hassk.) K. Schum.

Mt. Bellenden Ker, North Qld. (Qld Herbarium AQ736256)

3

Z06* Zingiber longipedunculatum Ridley Darwin Botanic Garden (D970235) 1 Z45^ Zingiber officinale Roscoe ‘Queensland’ Maroochy Research Station, Nambour, Qld. 1 Z108/9 Zingiber officinale Roscoe Purchased fresh in produce store, Lismore

(NCM04-207) 2, 5

J-2 Zingiber officinale Roscoe ‘Jamaican’, diploid

Maroochy Research Station, Nambour, Qld. 4

J-4-4 Zingiber officinale Roscoe ‘Jamaican’, tetraploid

Maroochy Research Station, Nambour, Qld. 4

Z124 Zingiber ottensii Valeton Flecker Botanic Gardens, Cairns 3 Z02* Zingiber montanum (J. König) Link ex.

A. Dietr. Darwin Botanic Garden (D970237) 1

Z105 Zingiber montanum (J. König) Link ex. A. Dietr.

Southern Cross University (NCM04-122) 2, 5

Z17 Zingiber spectabile Darwin Botanic Garden (D95357) 1 Z50 Zingiber spectabile Southern Cross University 1 Z55^ Zingiber ‘Dwarf Apricot’ Southern Cross University 1 Z19* Zingiber/Etlingera ‘Aniseed Ginger’ Darwin Botanic Garden (D95331) 1

± refer to Section 6.2.2 for details of extraction methods ; ^: cultivated in tubs at Southern Cross University; #: cultivated in the grounds of Southern Cross University; *: extract prepared from frozen rhizome material; +: ex George Brown Botanical Garden, Darwin; $: collected in the wild under permit.

Page 152: Phytochemistry and pharmacology of plants from the ginger family

128

6.2.2 Extraction methods

Although it would have been ideal for all extracts to have been prepared in the exact same

manner, this was not possible for practical reasons. Plant material was obtained over a

period of several years from various sources, in fresh, frozen or dried form. In some

situations it was also decided to compare extracts made with different solvents. This section

details the extraction methods employed.

6.2.2.1 Extraction method 1 (ethanolic)

All extracts were prepared from frozen rhizome material, except that of Curcuma

australasica, which was prepared from freshly harvested material.

Where possible, material that had been grown at Southern Cross University under uniform

conditions was used, but in cases where plants did not grow successfully, a frozen sample of

the material grown in Darwin was used (Table 1).

Rhizomes were cleaned of dirt and old leaf bases, where present, were removed. A healthy

tissue sample comprising a transection of the rhizome was chopped finely into pieces

approximately 1 mm3 in size. To chopped samples weighing between 0.9 g and 7.0 g was

added double the mass of 99% ethanol. Samples were then sonicated for 20 minutes

(Ultrasonic Cleaner 50 Hz, Unisonics Pty Ltd, Manly Vale, Australia) and subsequently

centrifuged for 10 minutes at 4000 rpm (Hettich Universal 16A Centrifuge, Tuttlingen,

Germany). The supernatant was retained and filtered through Zymark/Millipore Automation

Certified Filters (glassfiber APFB 1.0 µm prefilter, hydrophilic PTFE membrane 0.45 µm).

One mL aliquots of the extracts were taken to dryness on a rotational vacuum concentrator

(Christ Alpha 2-4, Osterode, Germany) 1 h at 35°C, 240 mbar; 1 h at 35°C, 150 mbar; 1 h at

35°C, 90 mbar; 4.5h at 45°C, 15 mbar). The resulting dry extracts were dissolved in DMSO

to a final concentration of 50 mg/mL.

Page 153: Phytochemistry and pharmacology of plants from the ginger family

129

6.2.2.2 Extraction method 2 (ethanolic)

Fresh rhizomes were sliced and dried at 40°C (except sample Z109 which was dried at 90°C

in order to examine the effect of drying temperature). The dried rhizomes were ground to a

coarse powder in a Retsch SM 2000 hammermill (Haan, Germany) fitted with a 4 mm mesh

and extracted with 10 parts (by mass) ethanol. The plant material was macerated for 24

hours in a stoppered conical flask while being agitated on a Bioline Orbital Shaker BL 4236.

The resulting extract was filtered through filter paper (Whatman No. 1) in a Buchner

apparatus and the solvent removed under vacuum on a Büchi rotary evaporator (Rotavapor

R114, Büchi Labortechnik, Flawil, Switzerland). The dry extract was weighed, redissolved

in ethanol to a known concentration and stored at –20°C until use. The dry extract yield as

percentage of dry herb mass was calculated.

6.2.2.3 Extraction method 3 (ethanolic)

Fresh rhizomes were sliced and dried at 40°C. The dried rhizomes were ground to a coarse

powder in a Waring blender and extracted with 4 parts (by mass) ethanol. The plant material

was macerated for 24 hours in a stoppered conical flask while being agitated on a Bioline

Orbital Shaker BL 4236. The resulting extract was filtered through filter paper (Whatman

No. 1) in Buchner apparatus and stored –20°C. Prior to testing in bioassays a small quantity

of extract was dried under a nitrogen stream and resolved to a known concentration.

6.2.2.4 Extraction method 4 (ethanolic)

Approximately 100 g fresh rhizome material (from 3-4 different rhizomes and incorporating

different rhizome parts) was chopped finely (approximately 1 mm3 cubes) and macerated

with 2 parts (by mass) ethanol in an agitated conical flask for 24 hours as described above.

The resulting extract was filtered through filter paper (Whatman No. 1) in Buchner apparatus

and stored –20°C. Prior to testing in bioassays a small quantity of extract was dried under a

nitrogen stream and resolved to a known concentration.

Page 154: Phytochemistry and pharmacology of plants from the ginger family

130

6.2.2.5 Extraction method 5 (aqueous)

Dried milled rhizome material (5.0 g) was macerated with 25 mL Milli-Q water in a

stoppered conical flask placed on an orbital shaker for 6 hours. The resulting extracts were

stored at –20°C. Prior to use, the extracts were spun down in a centrifuge, and filtered

through a double-layered glass filter-paper and a Millipore Millex-HN Nylon 0.45 µm filter.

6.2.2.6 Extraction method 6 (aqueous)

Freshly harvested rhizome and root tuber (5:1) were crushed in a mortar with a little water

and left to macerate for 15 min. The resulting aqueous juice was stored and filtered as

described for Extraction method 5.

6.2.3 Bioassay methods

6.2.3.1 Inhibition of PGE2 production

The bioassay was carried out in 96-well plates using Swiss albino mouse embryo fibroblast

cells 3T3 (American Type Culture Collection (ATTC), Manassas, VA) and a PGE2 enzyme

immuno-assay kit (Prostaglandin E2 EIA Kit – Monoclonal, Cayman Chemical Company,

Ann Arbor, MI, Catalog No. 514010).

Cell culture. Murine 3T3 fibroblasts were grown in 96-well plates (Nunclon, Thermo

Fischer Scientific, Rochester, NY). The growth medium consisted of Dulbecco’s Modified

Eagle Medium (colourless) containing 2 mM L-glutamine, 100 mM pyruvate, 4.5 g/L

glucose, 10% (v/v) foetal bovine serum and 100 U/mL each of penicillin G and

streptomycin. The concentration of the cell suspension was determined using a cell counter

(Beckman Coulter ActDiff Haematology Analyser, Fullerton, CA) and the suspension

diluted with growth medium to give a cell concentration of 2 105 cells/mL. 90 µL of this

suspension was pipetted into each well and the plate incubated at 37°C and 5% CO2 (Sanyo

CO2 MCO-17 AIC Incubator). After incubation for approximately 20 hours, plant extracts

and controls were added to the plate. In experiment series A plant extracts were tested a two

concentrations (50 mg/mL and 0.5 mg/mL, giving final well concentrations of 5000 µg/mL

and 500 µg/mL. In experiment series B plant extract were tested at three concentrations (1

Page 155: Phytochemistry and pharmacology of plants from the ginger family

131

mg/mL, 0.5 mg/mL and 0.1 mg/mL, giving final well concentrations of 100 µg/mL, 50

µg/mL and 10 µg/mL). Ethanolic plant extracts were diluted in the growth medium to an

ethanol concentration of 1% before being added to the wells.

Acetylsalicylic acid (Sigma A5376) was used as positive control. After addition of samples

and controls the plate was incubated for a further 3 hours. After 3 hours incubation, 1 µL

calcium ionophore A21387 (calcimycin; Sigma C7522) solution (5 mM in DMSO) was

added to each well to stimulate PGE2 synthesis, the plate shaken gently on an orbital shaker

for 20 seconds and incubated for a final 20 minutes. Immediately following incubation, the

plate was centrifuged (Hettich Universal 16A Centrifuge, Tuttlingen, Germany) for 3

minutes at 1000 RCF and the supernatant transferred to a clean 96-well plate, sealed and

frozen at –80°C until the enzyme immunoassay was performed.

Enzyme immunoassay. The PGE2 enzyme immunoassay was carried out using the thawed

cell culture supernatant, which was diluted 500-fold with the buffer provided with the kit.

50µL of diluted sample was added to each well. In experiment series A samples and controls

were run in duplicate, in experiment series B in triplicate. A PGE2 standard was run at 8

concentrations in duplicate on each plate in all experiments. PGE2 acetylcholine esterase

tracer and PGE2 monoclonal antibody were added (50 µL of each to each well) and the

sealed plate incubated at 4ºC for 18-20 hours. Following incubation the plate was washed 5

times with the wash buffer provided in the kit, dried and Ellman’s Reagent was added (200

µL per well). The plate was sealed, covered in alfoil and placed on an orbital shaker

(moderate speed) for 75-90 minutes, after which the absorbance at 405 nm of the developed

plate was read in a Wallac Victor 2 Plate Reader (Perkin Elmer, Waltham, MA). All

samples, standards and controls were assayed in triplicate.

Data analysis. Raw data were obtained as concentrations of PGE2 (pg/well). Standard

curves for PGE2 were plotted and R2 values calculated to confirm linearity. Mean, standard

deviation and coefficient of variance were calculated for replicates of samples, standards and

controls. Inhibition of PGE2 release by stimulated 3T3 cells was reported as percentage

inhibition compared with stimulated cells treated with solvent controls.

Page 156: Phytochemistry and pharmacology of plants from the ginger family

132

6.2.3.2 Oxygen Radical Absorbance Capacity (ORAC)

This assay is based on the work of Cao and Prior (Cao & Prior, 1999).

Ethanolic plant extracts were taken to dryness under a nitrogen stream and resolved and

diluted in AWA solvent (acetone 70.0 : Milli-Q water 29.5 : acetic acid 0.5) to give

concentrations of 100 µg/mL, 50 µg/mL, 25 µg/mL and 12.5 µg/mL.

Trolox was used as a standard and BHT as a positive control; stock solutions of both in 75

mM phosphate buffer pH 7.4 were diluted in AWA solvent to give Trolox concentrations of

100 µM, 50 µM, 25 µM, and 12.5 µM and butylated hydroxytoluene (BHT) concentrations

of 500 µM, 250 µM, 125 µM, and 62.5 µM.

The assay was carried out in black 96-well fluorescence assay plates (Perkin Elmer

Optiplate). Into each well was added 10 µL fluorescein solution (6.0 x 10-7 M in 75 mM

phosphate buffer [NaH2PO4 17.25 g : Na2HPO4 86.25 g, adjusted to pH 7.4]); 20 µL sample,

Trolox, BHT or solvent control solution; and 170 µL AAPH (2,2’-azobis-2-methyl-

propanimidamide, dihydrochloride; Cayman) solution (20 mM in 75 mM phosphate buffer,

pH 7.4). Fluorescein control wells received 10 µL fluorescein solution and 190 µL

phosphate buffer. Immediately after the addition of the AAPH solution, the assay plates

were placed in Wallac Victor 2 Plate Reader (Perkin Elmer, Waltham, MA) and the

fluorescence recorded at 37°C every minute for 35 minutes with the integral of these

readings representing the total antioxidant capacity. All samples, standards and controls

were assayed in quadruplicate.

Data analysis. Mean and standard deviation were calculated for replicates. Standard curves

for Trolox standards were plotted and R2 values calculated to confirm linearity.

6.2.3.3 Inhibition of nitric oxide production

This assay measures the release of nitric oxide by murine monocytes stimulated by

lipopolysaccharide (LPS).

Cell culture. RAW264 murine leukemic monocyte-macrophages (ATCC, Manassas, VA)

were grown in clear 96-well plates (Nunclon). The growth medium was the same as that

Page 157: Phytochemistry and pharmacology of plants from the ginger family

133

used in the PGE2 assay (see above). 100 µL of cell suspension (106 cells/mL) was added to

each well and the plate incubated at 37°C and 5% CO2 (Sanyo CO2 MCO-17 AIC Incubator).

After incubation (approx. 20 hours), 15 µL of plant extracts and controls were added to the

wells and the plate incubated for a further 1 hour. Ethanolic plant extracts were diluted in

the growth medium to an ethanol concentration of 1% before being added to the wells.

Extracts were screened at 10% and 20% of stock extract concentrations, which ranged from

35 – 50 mg/mL. One active extract was also assayed at 2% and 1% of stock extract

concentration. After incubation (1 hour), 5 µL of LPS solution (1 µg/mL in growth medium)

was added to each well and the plate returned to the incubator for a further 18 – 20 hours. At

the completion of the incubation period, the plate was centrifuged (3 min. at 1500 RCF) and

50 µL of the supernatants transferred to a clear assay plate (Spectra Plate, PerkinElmer).

Nitrite assay. Nitrite standards (7) were prepared by serial dilution of sodium nitrite in Milli-

Q water; standards were further diluted 10-fold in growth medium (final nitrite well

concentrations ranging from 1 to 500 µM). 50 µL of each standard concentration was

transferred to the designated wells. 50 µL of Griess Reagent (0.1% N-1-

napthylethylenediamine dihydrochloride, 1% sulphanilic acid in 5% phosphoric acid) was

added to all wells on the plate, the content mixed gently by placing the plate on an orbital

shaker (ensuring no bubbles were present), and the plate incubated at room temperature for

15 – 20 min protected from light. Following incubation the absorbance at 550 nm was read

in a Wallac Victor 2 Plate Reader. All samples, standards and controls were assayed in

triplicate.

Data analysis. Mean and standard deviation were calculated for replicates. Standard curves

for nitrite were plotted and R2 values calculated to confirm linearity. The nitric oxide

(nitrite) production in sample wells was calculated as a percentage of the production in

solvent control wells.

6.2.3.4 Modulation of natural killer cell activity

Target cells

Target cells were K562 cells (ATCC, Monassas, VA, USA), a leukemic cell line derived

from an individual with chronic myeloid leukaemia in terminal blast crisis. K562 cells lack

Page 158: Phytochemistry and pharmacology of plants from the ginger family

134

major-histocompatibility-complex class I and II antigens, which makes them a sensitive

target for human NK cells. K562 cells were stored under liquid nitrogen and prior to use

rapidly thawed in 37°C water bath, then incubated in medium (RPMI-1640 (Invitrogen,

Auckland, New Zealand) 87%, FBS 10% (Invitrogen), L-glutamine (Invitrogen) 1%,

Penicillin/Streptomycin (Invitrogen) 2%) for 1 hour at 37°C and 5% CO2 (Sanyo CO2 MCO-

17 AIC Incubator). Following incubation, the cell suspension was centrifuged for 5 minutes

at 120 RCF at room temperature, the supernatant discarded, and the cell pellet resuspended

in 1 mL RPMI-1640. Cells were then stained with 5 µL of a fluorescent membrane dye

(DiO Vybrant Cell Labelling Solution, Molecular Probes, Eugene, OR, USA, cat. no. V-

22886), incubated for a further 5 minutes at 37°C and 5% CO2, then centrifuged for 5

minutes at 120 RCT and washed twice with medium to remove excess dye before being

resuspended in 1 mL medium. The cell concentration of the suspension was determined

using an electronic cell counter (Beckman Coulter ActDiff Haematology Analyser, Fullerton,

CA) and adjusted with medium to 0.1 109 cells/L.

Effector cells

Human peripheral blood mononuclear cells (PBMCs) were isolated from fresh blood

collected by venesection into 4-mL Lithium-Heparin collection tubes. The whole blood was

diluted 1:1 with sterile phosphate-buffered saline (PBS), layered in 15-mL Falcon tubes on a

cushion of Isopaque-Ficoll (Amersham Biosciences, Uppsala, Sweden), and centrifuged for

20 minutes at 700 RCT (no brake, 0 deceleration). Following centrifugation, the upper

serum layer was removed, the PBMC layer collected and transferred to a clean 15-mL

Falcon tube, to which 12 mL of PBS was added. This cell suspension was centrifuged for 10

minutes at 250 RCF, the supernatant removed, and the PBMC pellet resuspended in 1 mL

medium. The cell concentration of this suspension was determined and adjusted with

medium to between 1 109 and 5 109 cells/L, dependent on the NK cell activity of the

donor blood used.

Page 159: Phytochemistry and pharmacology of plants from the ginger family

135

Pre-incubation of effector cells with extracts

In the case of aqueous extracts, 196 µL of cell suspension was incubated with 4 µL extract.

In order to minimise the solvent effects of the ethanolic extracts, 398 µL of PBMC

suspension was incubated with 2 µL extract for 2 hours at 37°C and 5% CO2.

Incubation of effector cells with target cells

Following the pre-incubation of effector cells with test extracts, effector and target cell

suspensions were combined in 5-mL Falcon tubes and incubated for 2 hours at 37°C/5%

CO2. For the ethanolic extracts, 200 µL of the pre-incubated effector cell suspension was

combined with 200 µL target cell suspension, resulting in a final effector to target cell ratio

of approximately 25:1 (subsequently reduced to 12:1 due to high activity of NK cells) and a

solvent concentration of 0.25%. For the aqueous extracts, 100 µL effector cell suspension

was combined with 200 µL target cell suspension and 100 µL medium, resulting in an

effector : target cell ratio of approximately 12:1 and a solvent concentration of 0.5%. No-

treatment and solvent controls were included in each experiment. After incubation for 2

hours the tubes were placed on ice (protected from light) followed by the addition of 100 µL

DNA stain (propidium iodide, Molecular Probes, Eugene, OR, cat. no. P-3566; 20 µL

stock/mL in PBS, final propidium iodide concentration 4 µg/mL) to each tube, after which

the tubes were again placed on ice for 5 minutes, then assayed within 30 minutes.

Control tubes containing target cell suspension and test extract but no effector cells were run

for all samples and the no-treatment and solvent controls in order to monitor any mortality of

the target cells caused by the extracts and/or solvent, independent of effector cell activity.

The kill rate of the control was subtracted from the kill rate of each test sample (see below).

Flow cytometric assay

The assay was carried out on a FACSCalibur (Becton Dickinson, Franklin Lakes, NJ) flow

cytometer linked to a MacIntosh computer OS 9.1 running CellQuest Pro Software (Becton

Page 160: Phytochemistry and pharmacology of plants from the ginger family

136

Dickinson). The percentage kill of target cells was based on a total count of 2500 target

cells. The specific cytotoxicity was calculated as follows:

Specific cytotoxicity (%) = Dead target cells(Sample) (%) − Dead target cells(Control)

6.3 Results

6.3.1 Inhibition of PGE2 production

A total of 42 samples representing 41 taxa and 39 species from a total of 14 genera were

screened for inhibition of PGE2 production. Samples were tested at two or three

concentrations. The data thus generated did not allow for accurate determination of IC50

values, but estimates were made where possible based on the available data (Table 6-2). For

the full data set, refer to Appendix B.

Only 9 of the 42 samples demonstrated 50% inhibition of PGE2 in the concentration range

tested. The species showing the most potent and dose-dependent inhibition were Alpinia

galanga, Boesenbergia rotunda, Curcuma australasica, C. longa, C. parviflora, Kaempferia

galanga, Pleuranthodium racemigerum, Zingiber officinale and Z. montanum. Most of these

are known medicinal plants, but nothing has been reported previously about the biological

activity of the two native Australian species, C. australasica and P. racemigerum.

Page 161: Phytochemistry and pharmacology of plants from the ginger family

137

Table 6-2. Inhibition of PGE2 production in 3T3 murine fibroblasts.

Estimated IC50 values are shown. NA: the estimated maximum inhibition was less than 50%. Concentrations refer to final concentrations in assay well. n=3 expect: * (n=2), + (n=4) and ^ (n=6).

Taxon ID Estimated IC50 (µg/mL) Curcuma longa^ Z106 c. 10 Curcuma australasica^ Z101 c. 50 Curcuma parviflora Z118 c. 50 Alpinia galanga Z103 >92 Boesenbergia rotunda Z104 c. 94 Pleuranthodium racemigerum^ Z128 50-100 Zingiber officinale Z108 50-100 Zingiber montanum Z105 50-100 Zingiber montanum* Z02 c. 500 Kaempferia galanga* Z05 c. 500 Alpinia calcarata* Z49 >100 Alpinia mutica* Z08 >500 Alpinia purpurea 'Eileen McDonald'* Z11 >500 Alpinia zerumbet* Z52 >500 Elettaria cardamomum* Z12 >500 Hedychium coronarium* Z10 >500 Scaphochlamys biloba* Z01 >500 Zingiber longipedunculatum* Z06 >500 Zingiber spectabile Z17 >500 Zingiber/Etlingera (Aniseed ginger)* Z19 >500 Alpinia arctiflora Z129 NA Alpinia caerulea Z102 NA Alpinia luteocarpa Z111 NA Alpinia malaccensis* Z48 NA Alpinia modesta Z127 NA Alpinia spectabile 'Giant Orange'* Z53 NA Costus barbatus Z112 NA Costus leucanthus Z113 NA Costus malortieanus Z114 NA Costus productus Z116 NA Costus pulverulentus Z115 NA Costus tappenbeckianus* Z15 NA Curcuma cordata Z117 NA Etlingera australasica Z107 NA Etlingera elatior 'Burma torch'* Z04 NA Hornstedtia scottiana Z126 NA Kaempferia rotunda Z120 NA Renealmia cernua Z121 NA Scaphochlamys kunstleri* Z122 NA Tapeinochilos ananassae Z125 NA Zingiber ottensii Z124 NA Zingiber sp. 'Dwarf Apricot'+ Z55 NA

Page 162: Phytochemistry and pharmacology of plants from the ginger family

138

6.3.2 Oxygen Radical Absorbance Capacity (ORAC)

Fifteen species were tested for antioxidant activity in the ORAC assay (Table 6-3).

Table 6-3. Oxygen radical absorbance capacity (ORAC).

Antioxidant activity is expressed in micromol Trolox equivalents (TE). Values shown are means±SEM for between 2 and 4 concentrations; each concentration was done in 4 replicates. ^: Drying temperature for plant material.

Taxon ID TE (µmol) per g extract

TE (µmol) per g dry herb equivalent

Curcuma longa Z106 3504±1141 528±172

Zingiber officinale (40°C)^ Z108 4166±63 262±4

Zingiber montanum Z105 2707±509 156±29

Boesenbergia rotunda Z104 2889±338 155±18

Zingiber officinale (90°C)^ Z109 3323±65 131±3

Alpinia galanga Z103 933±36 61±2

Curcuma australasica Z101 2357±479 54±11

Hornstedtia scottiana Z126 2284±275 47±6

Zingiber ottensii Z124 757±37 40±2

Tapeinochilos ananassae Z125 1257±125 37±4

Curcuma parviflora Z118 1452±191 37±5

Etlingera australasica Z107 778±75 36±4

Pleuranthodium racemigerum Z128 1418±221 34±5

Alpinia caerulea Z102 390±36 23±2

Alpinia luteocarpa Z111 1540±253 22±4

Scaphochlamys kunstleri Z122 944±60 18±1

The samples most active in this assay were Curcuma longa, Zingiber officinale, Z.

montanum and Boesenbergia rotunda. This was the case whether the samples were ranked

by Trolox Equivalence (TE, µmol) per gram extract or TE per gram dried herb equivalent.

Page 163: Phytochemistry and pharmacology of plants from the ginger family

139

6.3.3 Inhibition of nitric oxide production

Eight species were tested for their ability to inhibit nitric oxide production in RAW264 cells

(Table 6-4).

Table 6-4. Inhibition of nitric oxide production in LPS-stimulated RAW264 macrophages.

Values are percent inhibition compared with solvent control ± SEM. Concentrations shown are final concentrations in assay well.

Taxon ID Concentration (µg/mL)

Percent inhibition of nitric oxide

Alpinia galanga Z103 4.6 57±4 9.2 78±2 23.0 100±2 46.0 92±3 92.1 97±2 Curcuma australasica Z101 34.7 86±7 69.4 95±1 Zingiber officinale Z109 50.2 85±5 100.3 95±1 Boesenbergia rotunda Z104 47.3 67±1 94.52 100±3 Alpinia caerulea Z102 47.6 57±9 95.1 63±1 Zingiber montanum Z105 50.2 35±2 100.3 72±1 Curcuma longa Z106 49.7 35±1 99.3 51±2 Etlingera australasica Z107 46.4 0±0 92.7 16±1

Several of the tested extracts showed potent inhibition on nitric oxide at the concentrations

tested. At a concentration of 95-100 µg mL-1 Boesenbergia rotunda, Curcuma australasica

and Zingiber officinale caused almost complete inhibition of nitric oxide production. The

most potent of the extracts was Alpinia galanga, which demonstrated dose-dependent

inhibition in the range 5-92 µg mL-1 with complete inhibition achieved at 23 µg mL-1.

Page 164: Phytochemistry and pharmacology of plants from the ginger family

140

6.3.4 Modulation of natural killer cell activity

The results of the screening of 9 ethanolic extracts for modulation of NK cell activity are

shown in Table 6-5.

Table 6-5. Effect of ethanolic extracts on natural killer cell activity against K562 leukaemia cells.

^: Drying temperature for plant material.

Taxon ID Sample conc.

(µg/mL)

Conc. in pre-incubation tube with

effector cells (µg/mL)

Increase in specific cytotoxicity

compared with solvent control (%)

Untreated control 88.6 Solvent control (ethanol)

0.5% 0.0

Boesenbergia rotunda Z104 188.4 0.94 18.5 Curcuma longa Z106 198.0 0.99 10.8 Zingiber officinale (90°C)^

Z109 200.0 1.00 8.6

Etlingera australasica Z107 184.8 0.92 6.8 Zingiber officinale (40°C)^

Z108 173.6 0.87 0.0

Zingiber montanum Z105 200.0 1.00 -4.6 Curcuma australasica Z101 138.4 0.69 -4.9 Alpinia caerulea Z102 189.6 0.95 -4.9 Alpinia galanga Z103 91.8 0.46 -28.4

None of the ethanolic extracts tested produced a strong increase in NK cell activity. The

most effective extracts were Boesenbergia rotunda and Curcuma longa, which increased NK

cell activity by 18.5% and 10.8%, respectively. Alpinia galanga (tested at half the

concentration of the other extracts) caused an almost 30% decrease in activity.

Eight aqueous extract were also assayed for modulation of NK cell activity (Table 6-6).

Page 165: Phytochemistry and pharmacology of plants from the ginger family

141

Table 6-6. Effect of aqueous extracts on natural killer cell activity against K562 leukaemia cells.

*: percentage of stock extract. ^: Drying temperature for plant material.

Taxon ID Sample conc.*

Increase in specific cytotoxicity compared

with solvent control (%)

Untreated control -2.0 Solvent control (ethanol) 0.0 Curcuma longa Z106 0.1% 11.3 0.2% 10.7 0.4% 12.8 1.0% 6.5 10.0% 8.0 Curcuma australasica Z101 0.1% 2.0 1.0% 0.3 10.0% -0.5 Etlingera australasica Z107 0.1% -0.9 1.0% -9.9 10.0% 1.2 Zingiber officinale (40°C)^ Z108 0.1% 1.0 1.0% -11.9 10.0% 7.0 Zingiber officinale (90°C)^ Z109 0.1% -1.9 1.0% 9.9 10.0% -0.7 Zingiber montanum Z105 0.1% -0.3 1.0% 2.9 10.0% -11.9 Boesenbergia rotunda Z104 0.1% -6.0 1.0% -3.9 10.0% -18.8 Alpinia galanga Z103 0.1% 1.5 0.2% 9.2 0.4% 4.4 1.0% 0.2 10.0% -77.0

Page 166: Phytochemistry and pharmacology of plants from the ginger family

142

None of the aqueous extracts stimulated NK cell activity against K562 cells in a substantial

or dose-dependent manner. Alpinia galanga and to a lesser extent Boesenbergia rotunda

dramatically decreased the activity of natural killer cells at the highest concentration tested

(10% of stock extract), while these concentrations did not kill the target cells.

6.4 Discussion

6.4.1 Inhibition of PGE2 production

This study has for the first time demonstrated in vitro PGE2 inhibition by extracts of the two

native Australian Zingiberaceae, Curcuma australasica and Pleuranthodium racemigerum.

The estimated IC50 for the two extracts were c. 50 µg mL-1 and c. 75 µg mL-1, respectively,

which was similar to that of Zingiber officinale. These findings suggested the presence of

compounds with potent biological activity, and the extracts were therefore subsequently the

subject of bioactivity-guided fractionation (see Chapter 7). This is also the first report of

PGE2 inhibition by the South-east Asian species Curcuma parviflora.

Panduratin A isolated from Boesenbergia rotunda (syn. Kaempferia pandurata) has

previously been show to inhibit PGE2 and nitric oxide production as well as iNOS and COX-

2 expression in RAW264.7 cells (Yun et al., 2003), but the present study is the first to

demonstrate PGE2 inhibition by a crude extract of the plant.

PGE2 inhibition has also previously been shown for an extract of Zingiber montanum (syn. Z.

cassumunar) and several phenylbutanoid compounds isolated from this species (Jeenapongsa

et al., 2003; Jiang et al., 2006b; Panthong et al., 1997).

Extracts of Zingiber officinale, Curcuma longa and to a lesser degree Alpinia galanga were

included in this assay primarily as positive controls, as they are well established inhibitors of

PGE2 production in vitro and hence confirmed the validity of the assay.

The inhibitory effects of ginger (Z. officinale) and several ginger constituents on COX

activity and subsequent PGE2 production are well known and have been demonstrated in

isolated enzyme assays, cell-based models and animals. The early work by Kiuchi and

colleagues found that [6]-gingerol, [6]-gingerdione, [10]-gingerdione, [6]-

dehydrogingerdione and [10]-dehydrogingerdione isolated from fresh ginger were potent

Page 167: Phytochemistry and pharmacology of plants from the ginger family

143

inhibitors of COX in vitro with IC50 values comparable to that of indomethacin (Kiuchi et

al., 1982). Subsequent work by the same group reported potent inhibition by [10]-gingerol,

[6]-shogaol, [10]-gingerol, [6]-acetylgingerol, [6]-shogaol, [6]-dehydrogingerdione, [10]-

dehydrogingerdione, [6]-gingerdione and [10]-gingerdione, many of which had IC50 values

<2.5 µM (Kiuchi et al., 1992), and [8]-paradol was shown to be a potent inhibitor of COX in

vitro with an IC50 of 4 µM (Nurtjahja-Tjendraputra et al., 2003).

Although gingerols (in particular [6]-gingerol) are commonly referred to as the main active

component in ginger extracts, there are clearly numerous other compounds with anti-

inflammatory activity present. This was demonstrated by a recent study from the University

of Arizona showing that no correlation existed between PGE2 inhibitory activity of ginger

extracts and their gingerol content (Jiang et al., 2006b). This study found IC50 values for

methanolic ginger extracts ranging from 0.058 to 0.629 µg mL-1, which is much lower than

the estimated IC50 obtained in the present study (50-100 µg mL-1). This difference may be

due to the different cell lines and assay methods used in the two studies (Jiang and

colleagues used the human promonocytic cell line U937 and an immunoassay kit from R&D

System, while the present study used murine 3T3 fibroblasts and the PGE2 immunoassay kit

from Cayman Chemical Co.). The magnitude of difference seems unlikely to be due to

phytochemical differences between the plant materials used in the two studies.

The same University of Arizona study also found methanolic extracts of Alpinia galanga,

Zingiber montanum and Z. spectabile to be potent inhibitors of PGE2 production (IC50 values

of 0.06, 7.68 and 1.17 µg mL-1, respectively (Jiang et al., 2006b). In the present study A.

galanga and Z. montanum were both found to have some inhibitory activity (albeit again at

much higher concentrations), while Z. spectabile was not active.

Curcuma longa was the most potent inhibitor of PGE2 activity in the present study (IC50 = c.

10 µg mL-1). Rhizome extracts have been shown to inhibit PGE2 production in HL-60 cells

(IC50 = 0.92 µg mL-1) (Lantz et al., 2005), and the principal constituent, curcumin, is a well

documented inhibitor of COX expression and PGE2 production (Funk et al., 2006; Hong et

al., 2004; Ireson et al., 2001).

Page 168: Phytochemistry and pharmacology of plants from the ginger family

144

6.4.2 Oxygen Radical Absorbance Capacity (ORAC)

The most potent plant materials in terms of oxygen radical absorbance capacity (ORAC)

were turmeric (Curcuma longa) and ginger (Zingiber officinale) with values of 528 µmol TE

and 262 µmol TE per gram dry herb equivalent, respectively. The antioxidant properties of

both of these plants are well known (refer to Chapter 2, Sections 2.2.1.3.1 and 2.3.1.2.2.1).

Another study from the same laboratory reported a very similar ORAC value (557 µmol TE

per gram dry herb equivalent) for a turmeric extract prepared using a sequential three-solvent

extraction process (Wojcikowski et al., 2007). Another study reported an ethanolic extract

of fresh turmeric as having an ORAC value of 19.5 µmol TE per gram fresh weight (Tilak et

al., 2004). With fresh turmeric rhizome typically having a water content of about 75% (data

not shown), this is equivalent to approximately 80 µmol TE per gram dry herb. This value is

considerably lower than the one obtained in the present study (528 µmol TE), but differences

in extraction method and efficiency (stirring versus sonication) may at least partly explain

the discrepancy. Quantitative phytochemical differences between the two plant samples may

also have contributed.

One study has reported an ORAC value of 148 µmol TE per gram for fresh ginger (Zingiber

officinale) (Ninfali et al., 2005). Given that ginger rhizome typically has a water content

close to 90% (data not shown), this is equivalent to an ORAC value of approximately 1500

µmol TE per gram dry herb equivalent, which is an order of magnitude greater than the value

obtained in the present study (262 µmol TE per gram for ginger dried at 40° C). This

difference may reflect that the fresh rhizome has significantly greater antioxidant capacity

than the dried, but phytochemical differences in the plant material assayed and differences in

experimental methodology could also be factors.

It is interesting to note the difference in ORAC value for the two ginger samples, which were

prepared from the same raw material. Z108 was dried at 40° C, while Z109 was dried at 90°

C. Z108 had twice the ORAC value of Z109 (262 versus 131 µmol TE per gram dry herb

equivalent), clearly demonstrating that antioxidant activity is lost when plant material is

dried at high temperature. This is not unexpected, as oxidative processes catalysed by high

temperature are likely to deplete antioxidant capacity.

Page 169: Phytochemistry and pharmacology of plants from the ginger family

145

Curcuminoids isolated from Zingiber montanum (syn. Z. cassumunar) have been reported to

possess potent antioxidant activity (Masuda & Jitoe, 1994; Nagano et al., 1997). In the

present study the crude rhizome extract was found to be quite potent in the ORAC assay

(ORAC value of 156 µmol TE per gram dry herb equivalent).

Boesenbergia rotunda had an ORAC value similar to that of Z. montanum. B. rotunda (syn.

B. pandurata) has previously demonstrated potent antioxidant activity in a rat brain

homogenate assay, and several antioxidant compounds including panduratin A have been

isolated (Shindo et al., 2006).

Galangal (Alpinia galanga) has been shown to exert antioxidant activity in raw minced beef

(Cheah & Abu, 2000) but has not previously been tested in the ORAC assay. The observed

activity (61 µmol TE per gram dry herb equivalent) was quite modest.

None of the other species tested have been reported as having antioxidant activity. The two

native Australian species that displayed good activity in the PGE2 assay, Curcuma

australasica and Pleuranthodium racemigerum, had only modest antioxidant activity in the

ORAC assay (54 and 34 µmol TE per gram dry weight equivalent, respectively).

6.4.3 Inhibition of nitric oxide production

The ethanolic extract of Alpinia galanga was found to be a potent inhibitor of nitric oxide

production. At a concentration of 4.6 µg mL-1 it caused 57% inhibition, which corresponds

well with a reported IC50 = 7.3 µg mL-1 for an 80% aqueous acetone extract in LPS-activated

mouse peritoneal macrophages (Morikawa et al., 2005). 1’S-1’-acetoxychavicol acetate

isolated from A. galanga was found to be active in a peritoneal macrophage assay (Matsuda

et al., 2005).

Inhibition of nitric oxide production in vitro has been demonstrated for several other Alpinia

species and a considerable number of compounds isolated from them. These include

diarylheptanoids, a curcuminoid and flavonoids from A. blepharocalyx (Kadota et al., 1996;

Prasain et al., 1998), diarylheptanoids from A. officinarum (Lee et al., 2006; Matsuda et al.,

2006) and sesquiterpenes from A. oxyphylla (Muraoka et al., 2001). Inhibition of inducible

Page 170: Phytochemistry and pharmacology of plants from the ginger family

146

nitric oxide synthase (iNOS) has been reported for isolates from A. katsumadai (Hong et al.,

2002), A. officinarum (Yadav et al., 2003) and A. oxyphylla (Chun et al., 2002).

A methanolic extract of Curcuma longa (10 µg mL-1) was reported to cause 88% inhibition

of nitric oxide production in LPS-stimulated RAW264.7 cells (Hong et al., 2002). In

comparison, the present study found only 35% inhibition by C. longa at a concentration of

50 µg mL-1 (and 51% inhibition at 99 µg mL-1).

Inhibition of nitric oxide production and suppression of iNOS expression has previously

been demonstrated for methanolic extracts of C. longa, C. comosa and C. zedoaria

(Camacho-Barquero et al., 2007; Hong et al., 2002; Jang et al., 2004; Jantaratnotai et al.,

2006). Active compounds identified from these species include curcumin from C. longa and

two sesquiterpenoids from C. zedoaria. The sesquiterpenoid xanthorrhizol isolated from C.

xanthorrhiza has also been shown to inhibit iNOS activity and expression (Chung et al.,

2007; Lee et al., 2002).

In the present study ginger (Zingiber officinale) extract was a potent inhibitor of nitric oxide

at the concentrations tested (50 and 100 µg mL-1), while Z. montanum was less potent.

While inhibitory effects on nitric oxide production have not previously been reported for the

latter, this activity has been described for Z. officinale in osteoarthrotic sow chondrocytes

(Shen et al., 2005), and [6]-gingerol (the major pungent compound in ginger) has shown

dose-dependent inhibition of nitric oxide production and reduction in iNOS in LPS-

stimulated J774.1 mouse macrophages (Ippoushi et al., 2003). From other Zingiber species,

sesquiterpenoids from Z. zerumbet and the labdane diterpene aframodial from Z. mioga have

been shown to inhibit the expression of iNOS (Kim et al., 2005a; Murakami et al., 2002).

6.4.4 Modulation of natural killer cell activity

Only modest increases in NK cell activity were observed in this study for certain extracts,

while others were inactive. Best results were seen for the ethanolic extracts of Boesenbergia

rotunda (18.5% increase compared with solvent control) and Curcuma longa (10.8%

increase). The aqueous extract of C. longa also produced a modest increase (about 10%), but

this was not dose dependent in the concentration range tested. In contrast, both the ethanolic

extract of Alpinia galanga and the aqueous extract of fresh Boesenbergia rotunda caused

Page 171: Phytochemistry and pharmacology of plants from the ginger family

147

marked inhibition of NK cell activity (by 28.4% and 18.8%, respectively). There are no

previous reports concerning the effects of A. galanga or B. rotunda on NK cell activity.

Curcumin (0.01 µg mL-1) has previously been shown to significantly increase the

cytotoxicity of human NK cells ex vivo, using K562 cells as a target (Yadav et al., 2005).

However, in that study effector cells were incubated with curcumin (solvent unknown) for 18

hours as opposed to 2 hours in the present study, and the effector cells were subsequently

incubated with the target cells for 4 hours compared with 2 hours in the present study.

Curcumin was also found to partially reverse breast tumour exosome-mediated inhibition of

NK cell tumour cytotoxicity in vitro (Zhang et al., 2007). However, NK cell activity did not

change significantly compared with controls in rats fed dietary curcumin (1, 20 or 40 mg per

kg bodyweight) for 5 weeks (South et al., 1997).

In the present study the presence of ethanol even at a very low concentration (0.25%) in the

target cell suspension caused a significant (47%) decrease in specific NK cell cytotoxicity

compared with the untreated control. The sizeable effect of the solvent may have affected

the results of this assay for the ethanolic extracts. Further work should explore the use of

alternative extraction solvents (less polar than water) that may interfere less with the assay

than does ethanol.

6.4.5 Conclusion

This study has confirmed the Zingiberaceae as a rich source of compounds with interesting

biological and pharmacological actions. Of some 40 species tested in at least one assay, the

established medicinal species turmeric (Curcuma longa) and ginger (Zingiber officinale)

emerged as the most active. Two other plants with traditional medicinal uses, galangal

(Alpinia galanga) and Boesenbergia rotunda also showed good activity in several assays.

Of the 41 taxa tested turmeric was the most potent inhibitor of PGE2, and it was also the best

antioxidant in the ORAC assay. It caused only moderate inhibition of nitric oxide

production, but other studies suggest that this may be an aberrant result. The well known

ability of ginger to inhibit PGE2 production in vitro was confirmed, and ginger also showed

good antioxidant capacity, although rhizome material dried at 90° C had only half the

activity of material dried at 40° C. Ginger was also a potent inhibitor of nitric oxide

Page 172: Phytochemistry and pharmacology of plants from the ginger family

148

production, but like turmeric did not affect NK cell activity significantly. Boesenbergia

rotunda inhibited PGE2 production and showed antioxidant activity. The ethanolic extract of

the dried material caused an increase in NK cell activity, while an aqueous extract of the

fresh material caused a marked decrease. The explanation for this is unknown and warrants

further investigation.

Of particular interest was the screening of seven native Australian species in the PGE2 assay,

as none of these had previously been investigated for pharmacological activity. Two of these

species, Curcuma australasica and Pleuranthodium racemigerum, showed good activity in

this assay and were selected for further investigations (see Chapter 7).

Page 173: Phytochemistry and pharmacology of plants from the ginger family

149

7. BIOACTIVITY-GUIDED FRACTIONATION OF TWO NATIVE AUSTRALIAN ZINGIBERACEAE

7.1 Introduction

In the previous chapter, two native Australian species, Curcuma australasica and

Pleuranthodium racemigerum, were identified as having significant PGE2 inhibitory activity

in vitro. In this respect they demonstrated potency similar to well known medicinal species

such as Curcuma longa and Alpinia galanga.

Due to the paucity of information about the chemistry and pharmacological activity of both

Curcuma australasica and Pleuranthodium racemigerum, these species were selected for

further work with a view to extending the knowledge of the Zingiberaceae and about native

Australian plants.

The process of bioactivity-guided fractionation was applied and led to the isolation of

pharmacologically active compounds from both species. Inhibition of PGE2 was used as the

primary bioassay in this process, but fractions with high activity in this assay were also

tested for cytotoxic properties, primarily to ensure that the activity observed in the cell-based

PGE2 assay was not simply due to cytotoxic effects. Testing for cytotoxicity included the

murine cell line P388D1, which in the laboratory’s experience is highly sensitive to cytotoxic

agents. It should be noted that significant cytotoxic activity may potentially be of interest in

itself.

From C. australasica two known bioactive sesquiterpene compounds, zederone and

furanodien-6-one were isolated. In the case of P. racemigerum a novel curcuminoid

compound with significant pharmacological activity was isolated and structurally

characterised. The isolated compounds could potentially become candidates for further work

and preclinical testing.

Page 174: Phytochemistry and pharmacology of plants from the ginger family

150

7.2 Materials and Methods

7.2.1 Plant material

Fresh rhizome of Curcuma australasica was obtained from the George Brown Botanic

Gardens in Darwin (S 12° 27’; E 130° 50’) as a gift from Dr Greg Leach who also identified

the material.

Pleuranthodium racemigerum was collected in rainforest (828 m above sea level) in the Wet

Tropics Region of North Queensland in Gillies Range (S 17° 13'; E 145° 40') under a permit

issued by the Queensland Environmental Protection Agency. Voucher specimens have been

deposited in the Medicinal Plant Herbarium at Southern Cross University (NCM05-047) and

the Queensland Herbarium (AQ736263).

7.2.2 Extraction methods

7.2.2.1 Curcuma australasica

Fresh rhizomes were cleaned, sliced and extracted in double the mass of ethanol (99.7%) in a

sonicating bath for 30 min. The extract was filtered through filter paper (Whatman No. 3) in

a Buchner apparatus. The biomass was again covered with ethanol, left to steep for 22 hours

at 5° C, sonicated for 10 min, then filtered. The two filtrates were combined, and the extract

taken to dryness under vacuum on a Büchi Rotavapor R-114 (Switzerland) with the water

bath temperature at 40° C. The resulting semi-dry extract was dried on a rotational vacuum

concentrator for 48 hours at –88° C and 0.140 mbar (Christ Alpha 2-4, Osterode, Germany).

The dry extract was stored at –20° C until use.

7.2.2.2 Pleuranthodium racemigerum

Fresh rhizomes were cleaned, sliced and dried at 40° C. The dried rhizomes were ground to

a coarse powder in a Waring blender and extracted with 4 parts (by mass) ethanol. The plant

material was steeped for 24 hours in a stoppered conical flask while being agitated on a

Bioline Orbital Shaker BL 4236 (Edwards Instrument Company, Australia). The resulting

Page 175: Phytochemistry and pharmacology of plants from the ginger family

151

extract was filtered through a Quickfit glass filter and stored at –20° C. Prior to

fractionation, the extract was dried under vacuum on a Büchi Rotavapor R-114 (Switzerland)

with the water bath temperature at 45° C. The resulting extract, which was oily, was

redissolved in 91% aqueous methanol and partitioned with hexane in a separating funnel.

The methanolic phase was dried under vacuum. Prior to testing in bioassays a small quantity

of extract was dried under a nitrogen stream and redissolved to a known concentration.

7.2.3 Fractionation by preparative HPLC

7.2.3.1 Curcuma australasica

The dry extract was redissolved in 75% aqueous methanol (1 part extract to 10 parts solvent

by mass) and filtered through an Acrodisc CR PTFE 0.45 µm microfilter (Pall Gelman Lab.,

Ann Arbor, MI). Fractionation was carried out on a Gilson Preparative HPLC system fitted

with a Gilson FC204 Fraction Collector and an Altima C-18 5 µm column (150 mm long;

internal diameter 22 mm) (Alltech, Kentucky). Mobile phase A consisted of HPLC-grade

water obtained from an in-house Milli-Q system (Waters, Milford, MA), mobile phase B

consisted of HPLC-grade methanol (EM Science, Gibbstown, NJ); both contained 0.05%

trifluoroacetic acid (TFA). The mobile phase gradient was A:B (70:30, v/v) to A:B (5:95,

v/v) over 20 min followed by A:B (5:95, v/v) for 10 min, with a flow rate of 20 mL/min.

UV-VIS detection was at 210 nm and 360 nm. Twenty-nine 0.8-minute fractions were

collected, commencing at 2 min. Another fractionation of the same extract was carried out,

this time collecting 18 0.5-min fractions, starting at 12 min (3 1-mL injections of sample

were fractionated). Fractions were dried on a rotational vacuum concentrator for 48 hours at

–88° C and 0.140 mbar (Christ Alpha 2-4, Osterode, Germany).

7.2.3.2 Pleuranthodium racemigerum

The methanolic extract residue (700 mg) was redissolved in 3.5 mL acetonitrile with a few

drops of water and fractionated on a Gilson Preparative HPLC system as described in the

previous section. Mobile phase A consisted of HPLC-grade water, mobile phase B consisted

of HPLC-grade acetonitrile; both contained 0.05% TFA. The gradient eluting mobile phase

Page 176: Phytochemistry and pharmacology of plants from the ginger family

152

was A:B (70:30, v/v) to A:B (5:95, v/v) over 20 min followed by A:B (5:95, v/v) for 5 min,

with a flow rate of 20 mL/min. UV-VIS detection was at 210 nm and 360 nm. Twenty-nine

1-minute fractions were collected, commencing at 1 min. Two 1-mL injections of the

sample were fractionated. Another fractionation of the same extract was carried out, again

collecting 29 1-min fractions. Fractions were dried on a rotational vacuum concentrator as

described in the previous section.

7.2.4 Nuclear magnetic resonance spectroscopy

Nuclear magnetic resonance (NMR) spectra were obtained using a Bruker AVANCE

DRX500 (1H at 500.13 MHz; 13C at 125.77 MHz; 5mm QNI probe) spectrometer with

XWin-NMR software. The 1H and 13C NMR spectra were recorded using deuterated

chloroform (CDCl3) with the residual solvent peaks as reference (7.27 ppm for 1H and 77.2

ppm for 13C). In one case deuterated pyridine (pyridine-d5) was used instead (residual

solvent peaks 8.74 ppm for 1H and 150.35 ppm for 13C). The chemical shifts were expressed

in parts per million (ppm) as δ values and the coupling constants (J) in Hertz (Hz). All

experiments were carried out using the Bruker pulse programs.

7.2.4.1 One-dimensional NMR spectroscopy

7.2.4.1.1 1H NMR

The 1H NMR spectra were obtained for the isolated compounds. The chemical shifts,

coupling constants, peak intensities and splitting patters of each of the proton signals

provided information about the type of protons present in the molecule and their chemical

environments (Williams & Fleming, 1989).

7.2.4.1.2 J-modulated 13C NMR

The carbon resonances were distinguished according to their proton attachments (C, CH,

CH2, CH3) using the APT pulse sequence. The J-modulated 13C NMR spectra displayed the

Page 177: Phytochemistry and pharmacology of plants from the ginger family

153

CH3 and CH resonances arbitrarily pointing down and inverted with respect to the CH2 and

the quaternary carbons, which were arbitrarily pointing up (Sanders & Hunter, 1987).

7.2.4.2 Two-dimensional homonuclear correlation NMR spectroscopy

In the two-dimensional experiments, the 1H NMR spectra were displayed along both axes

and the proton correlations were indicated in the contour plot by cross-peaks along the

diagonal.

7.2.4.2.1 1H-1H Correlation spectroscopy (COSY)

This two-dimensional technique was used to correlate the chemical shifts of 1H nuclei that

were coupled to one another. A magnitude sequence with a final pulse angle of 45°

(COSY45) was used to obtain a COSY spectrum. The cross-peaks produced indicated which 1H nuclei were J-coupled.

7.2.4.3 Two-dimensional heteronuclear correlation spectroscopy

7.2.4.3.1 Heteronuclear single quantum correlation (HSQC) spectroscopy

This two-dimensional method was used to determine which 1H nuclei were bonded to which 13C nuclei (1JCH) in a molecule using an INEPT pulse sequence.

7.2.4.3.2 Heteronuclear multiple-bond correlation (HMBC) spectroscopy

This two-dimensional method was used for determining long-range 1H-13C connectivities

and was useful in determining the C assignments in a molecule through its correlation with a

proton.

Page 178: Phytochemistry and pharmacology of plants from the ginger family

154

7.2.5 Determination of accurate mass

A sample of the isolated Compound 3 from Pleuranthodium racemigerum was sent to the

School of Molecular and Microbial Sciences, University of Queensland, where determination

of accurate mass by high-resolution mass spectroscopy was carried out by Mr Graham

Macfarlane.

Electron impact ionization (EI) experiments were conducted on a Kratos MS25 RFA

instrument via a direct insertion probe at 70 eV and source temperature of 200° C.

Perfluorokerosene (PFK) was used as reference for magnet scan accurate mass in EI.

Electrospray ionization (ESI) experiments were conducted on a Finnigan MAT 900 XL-Trap

instrument with a Finnigan API III electrospray source, using MeOH as the solvent and

polypropylene glycol and polyethylene glycol as references for accurate mass data, acquired

by electric sector scan.

7.2.6 UV spectroscopy

UV absorbance data were obtained on a Hewlett-Packard Spectrophotometer 8453 (Palo

Alto, CA) controlled by UV-Visible ChemStation software (Rev. A.0803; Dayton, Ohio).

Absorbance data were acquired in the range 190-1100 nm at 1 nm intervals. The novel

Compound 3 from Pleuranthodium racemigerum was dissolved in methanol at a

concentration of 0.014 mg/mL (4.723 10–5 M). The determination of peak absorption

(λmax) was based on the mean value of five measurements. The extinction coefficient

(wavelength-dependent molar absorptivity coefficient), ε, was calculated from the Beer-

Lambert Law, A = ε c d, where A is absorbance, c the molar concentration and l the

distance (cuvette path) in cm.

7.2.7 Cytotoxicity assays

Cytotoxicity in 3T3 murine fibroblasts, P388D1 murine lymphoblasts, Caco-2 human

colonic adenocarcinoma, PC3 human prostate adenocarcinoma, HepG2 human hepatocyte

carcinoma, and MCF7 human mammary adenocarcinoma (all cell lines originating from the

Page 179: Phytochemistry and pharmacology of plants from the ginger family

155

American Type Culture Collection, ATCC) was assayed in 96-well plates using the ATPLite

kit (PerkinElmer, Waltham, MA) with chlorambucil and curcumin as reference compounds.

Cells were cultivated in the appropriate medium and test and reference compounds added at

five different concentrations and incubated for 24 hours (5% CO2). All samples were

assayed in triplicate.

The ATPLite assay was carried out according to the manufacturer’s instructions. Briefly, the

kit components were equilibrated to room temperature. The lyophilised substrate solution

was reconstituted with buffer and this reagent added to the plate containing the cells

incubated with test substances and standards. The plate was then shaken on an orbital

microplate shaker (700 rpm) for 2 minutes before luminescence was measured on a Wallac

Microbeta Scintillation and Luminescence Counter (PerkinElmer, Waltham, MA). ATP was

quantified using an ATP standard curve based on luminescence measurements of ATP

standards. Median lethal dose (LD50) values were calculated using Excel 2003 (Microsoft

Corporation, Redmond, WA) and GraphPad Prism 4 (San Diego, CA).

7.2.8 PGE2 assay

Inhibition of PGE2 production by 3T3 murine fibroblast cells (ATCC) was assayed using the

Prostaglandin E2 EIA Kit (Cayman Chemical, Ann Arbor, MI) as described in the previous

chapter.

7.3 Results

7.3.1 Curcuma australasica

7.3.1.1 Activity-guided fractionation

The extract was profiled by LC-MS prior to fractionation (Fig. 7-1).

Page 180: Phytochemistry and pharmacology of plants from the ginger family

156

min2 4 6 8 10 12 14 16 18 20

mAU

-100

102030405060

DAD1 A, Sig=210,8 Ref=off (D:\020821A\011-0101.D)

min2 4 6 8 10 12 14 16 18 20

mAU

-5

0

5

10

15

DAD1 C, Sig=280,8 Ref=off (D:\020821A\011-0101.D)

min2 4 6 8 10 12 14 16 18 20

0

20000

40000

60000

80000

MSD1 TIC, MS File (D:\020821A\011-0101.D) APCI, Pos, Scan, Frag: 150, "pos 150"

Fig. 7-1. LC-MS chromatogram of Curcuma australasica extract redissolved in methanol. Top: 210 nm; centre: 280 nm; bottom: total ion chromatogram.

Based on the chromatograms of the whole extract and the individual fractions, some

fractions were combined before being tested for their ability to inhibit PGE2 production in

murine fibroblasts. The results of this assay are shown in Table 7-1.

Page 181: Phytochemistry and pharmacology of plants from the ginger family

157

Table 7-1. Inhibition of PGE2 production in 3T3 murine fibroblast cells by fractions of Curcuma australasica extract. The final concentration of test substances in the assay well was 1mg/mL.

Fraction no. Percent inhibition

(mean±SD)

1 16±8

2 12±9

3-9 18±11

10 19±3

11-13 13±11

14 26±5

15 47±6

16 80±10

17 82±6

18 64±4

19 57±8

20 61±9

21-28 17±15

6-gingerol 94±4

8-gingerol 96±3

10-gingerol 92±2

8-shogaol 82±5

Two fractions, F16 and F17, showed strong inhibition of PGE2 production (80% and 82%

inhibition, respectively). F17 proved to contain a single major compound, Compound 1

(Figs. 7-2, 7-3), while F16 contained multiple compounds and included overlap with the

compound in F17. On this basis it was decided to proceed with structural elucidation of the

major compound in F17.

Page 182: Phytochemistry and pharmacology of plants from the ginger family

158

min16 16.5 17 17.5 18 18.5 19 19.5 20 20.5

mAU

0250500750

100012501500

DAD1 A, Sig=210,8 Ref=off (D:\020830\001-0101.D)

min16 16.5 17 17.5 18 18.5 19 19.5 20 20.5

mAU

050

100150200250300

DAD1 C, Sig=280,8 Ref=off (D:\020830\001-0101.D)

min16 16.5 17 17.5 18 18.5 19 19.5 20 20.50

2000000

4000000

6000000

8000000

10000000

MSD1 TIC, MS File (D:\020830\001-0101.D) APCI, Pos, Scan, Frag: 150, "pos 150"

Fig. 7-2. LC-MS chromatogram for Compound 1 (Fraction 17) from Curcuma australasica.

m/z100 200 300

0

20

40

60

80

100

*MSD1 SPC, time=19.660:19.785 of D:\020830\001-0101.D APCI, Pos, Scan, F

Max: 3.24134e+006

139

.2

201

.3

229

.2 2

30.1

245

.1 2

47.2

248

.1

nm200 225 250 275 300 325 350 375

mAU

0

200

400

600

800

1000

DAD1, 19.742 (1158 mAU,Dn1) of 001-0101.D

Fig. 7-3. Mass spectrum (M+1; left) and UV spectrum (right) for Compound 1 (Fraction 17) from Curcuma australasica.

Another active fraction, F20, was found to consist of another almost pure compound

(Compound 2), which was also the subject of structural elucidation, although it displayed

less potent activity in the PGE2 assay (61% inhibition). The LC-MS chromatogram for F20

is shown in Fig. 7-4 and the mass and UV spectra in Fig. 7-5.

Page 183: Phytochemistry and pharmacology of plants from the ginger family

159

min18 19 20 21 22 23

mAU

0

500

1000

1500

2000

2500

DAD1 A, Sig=210,8 Ref=off (D:\020829\008-0801.D)

min18 19 20 21 22 23

mAU

0

500

1000

1500

2000

DAD1 C, Sig=280,8 Ref=off (D:\020829\008-0801.D)

min18 19 20 21 22 230

10000002000000300000040000005000000600000070000008000000

MSD1 TIC, MS File (D:\020829\008-0801.D) APCI, Pos, Scan, Frag: 150, "pos 150"

Fig. 7-4. LC-MS chromatogram for Compound 2 (Fraction 20) from Curcuma australasica.

m/z100 200 300

0

20

40

60

80

100

*MSD1 SPC, time=22.680:22.804 of D:\020829\008-0801.D APCI, Pos, Scan, F

Max: 4.14733e+006

149

.1

213

.3

231

.3 2

32.2

nm200 225 250 275 300 325 350 375

mAU

0

200

400

600

800

1000

1200

1400

DAD1, 22.818 (1588 mAU,Dn1) of 008-0801.D

Fig. 7-5. Mass spectrum (M+1; left) and UV spectrum (right) for Compound 2 (Fraction 20) from Curcuma australasica.

Page 184: Phytochemistry and pharmacology of plants from the ginger family

160

7.3.1.2 Cytotoxic activity of Compound 1

In order to ensure that the observed activity in the PGE2 was not simply due to cytotoxic

effects, Compound 1 was also tested for cytotoxicity in the sensitive P388D1 murine

lymphoma cell line and compared to a number of other compounds and extracts (Table 7-2).

Table 7-2. Cytotoxicity of Compound 1 from Curcuma australasica in P388D1 murine lymphoma cells.

Values are mean±SD percentage survival compared to DMSO-treated (n=3).

Test substance Percent survival (mean±SD) 500 µg/mL 150 µg/mL 100 µg/mL 50 µg/mL 10 µg/mL

DMSO (control) 100±8 100±8 100±8 100±8 100±8 Alpinia galanga 1±0 - - 48±7 - Alpinia caerulea 68±2 - - 78±6 - Zingiber officinale 48±6 - - 68±5 - Curcuma longa 28±1 - - 62±1 - Curcuma australasica 34±3 - - 72±3 - Compound 1 - - 25±1 - 77±6 Curcumin - - 1±0 - 61±4 6-shogaol - - 1±0 - 50±5 Chlorambucil - 40±2 - - -

Compound 1 was found to be considerably less cytotoxic than curcumin and [6]-shogaol.

The cytotoxicity of the crude extract of Curcuma australasica was similar to that of Zingiber

officinale and less than that of C. longa.

7.3.1.3 Structural elucidation of Compound 1 and 2

Compounds 1 and 2 were the subjects of structural elucidation experiments by NMR

spectroscopy.

Page 185: Phytochemistry and pharmacology of plants from the ginger family

161

Compound 1 was identified as the sesquiterpene ketone zederone (Fig. 7-6) by comparison

of its spectral data with those reported in the literature (Hikino et al., 1966; Shibuya et al.,

1987). These data are presented in Table 7-3.

O

OO

43

21

10

15

98

7

6

12

13

115

14

Fig 7-6. Structure of zederone (Compound 1).

Page 186: Phytochemistry and pharmacology of plants from the ginger family

162

Table 7-3. 1H and 13C NMR spectral data of Compound 1 (zederone) from Curcuma australasica.

δ = chemical shift (ppm), int = integration, mult = multiplicity, J = coupling constant, s = singlet, d = doublet, dd = doublet of a doublet, ddd = doublet of a doublet of a doublet, dddd = doublet of a doublet of a doublet of a doublet, dt = doublet of a triplet, dq = doublet of a quartet, m = multiplet, br = broad

* Assignment could have been interchanged.

Compound 2 was identified as another sesquiterpene ketone, 1(10)E,4E-furanodien-6-one,

also by comparison with spectral data from the literature (Dekebo et al., 2000; Hikino et al.,

1975; Makabe et al., 2006). These data are presented in Table 7-4.

Differences greater than 1 ppm exist between the observed δC values and those reported by

Dekebo et al. (2000) for C-6 (1.2 ppm), C-7 (1.8 ppm), C-9 (1.3 ppm) and C-11 (1.2 ppm),

but these can be reasonably explained by the differences in instrumentation. It should be

Compound 1 CDCl3

Zederone (Shibuya et al. 1987) CDCl3

Position δC 126 MHz δH 500 MHz (int, mult, J in Hz)

δH, 500 MHz

1 131.5 5.49, 1H, dd (11.9, 4.0) 5.48, dd (11.5, 4.0) 2 24.9 2.25, 1H, m

2.53, 1H, m 2.23, br d 2.52, dddd (11.5, 12.0, 3.5, 13.5)

3 38.2 1.29, 1H, m 2.31, 1H, dt (13.0, 3.5)

1.29, ddd (13.5, 4.0, 13.0) 2.29, ddd (13.0, 3.5, 3.5)

4 64.2 5 66.8 3.82, 1H, s 3.81, s 6 192.5 7 123.5* 8 157.4 9 42.1 3.73, 2H, dd (16.5) 3.69, 3.75 (16.5)

10 131.3 11 122.5* 12 138.3 7.10, 1H, s 7.08, q (1.2) 13 10.5 2.13, 3H, d (1.2) 2.11, d (1.2) 14 15.4 1.35, 3H, s 1.34, br s 15 15.9 1.61, 3H, s 1.60, br s

Page 187: Phytochemistry and pharmacology of plants from the ginger family

163

noted that one of the coupling constants of the olefinic proton H-1 was found to be 11.5 Hz.

Dekebo et al. (2000) and Hikino et al. (1975) reported coupling constants of 6.6 and 4.5 Hz,

and 7.5 and 7.5 Hz, respectively. The large J value observed with this compound (2) has to

be a coupling of H-1 to one of the protons at the 2-position (H-2), probably the proton

resonating at 2.32 ppm (CDCl3) and for the J value to be that high, both H-1 and one of the

methylene protons must have assumed an axial-axial orientation. With the methyl group at

the 10-position trans to H-1, and the methyl group at the 4-position trans to H-5, the

structure could have formed a rigid configuration making H-1 almost axial to one of the

methylene protons of H-2.

Page 188: Phytochemistry and pharmacology of plants from the ginger family
Page 189: Phytochemistry and pharmacology of plants from the ginger family

165

The observed 1H and 13C chemical shift assignments for this compound (2) agree with the

literature values for 1(10)E,4E-furanodien-6-one as shown in Table 7-4. The structure of

1(10)E,4E-furanodien-6-one is shown in Fig. 7-7.

O

O

1

2

34 5 6

78

9

10

11

12

1314

15

Fig. 7-7. Structure of 1(10)E,4E-furanodien-6-one (Compound 2) from Curcuma australasica.

There are other reported isomers of this compound such as the 1(10)E,4Z isomer, which is

isofuranodienone (Hikino et al., 1975) and the 1(10)Z,4Z isomer (Brieskorn & Noble, 1983).

The NMR spectral data for these two isomers are summarised in Table 7-5. It is evident

from inspection of Table 7-4 and Table 7-5 that the NMR data of the isolated compound (2)

are consistent with those of the E,E isomer.

Page 190: Phytochemistry and pharmacology of plants from the ginger family

166

Table 7-5. 1H NMR spectral data from the literature for the isomers isofuranodienone and 1(10)Z,4Z-furanodien-6-one.

δ = chemical shift (ppm)

Isofuranodiene

Hikino et al. (1975) CDCl3

1(10)Z,4Z-furanodien-6one

Brieskorn & Noble (1983) CDCl3

Position δH 100 MHz δH, 90 MHz

1 5.20 5.29 5 6.10 5.88 9 3.13, 3.49 3.34

12 7.00 7.08 13 1.91 2.20 14 1.89 1.76 15 1.55 1.65

Page 191: Phytochemistry and pharmacology of plants from the ginger family

167

7.3.2 Pleuranthodium racemigerum

7.3.2.1 Activity-guided fractionation

An HPLC chromatogram of the extract redissolved in 90% aqueous methanol is shown in

Fig. 7-8.

min2 4 6 8 10 12 14 16 18 20

mAU

0200400600800

100012001400

DAD1 A, Sig=210,8 Ref=off (Z:\CHROMO~1\001-0101.D)

min2 4 6 8 10 12 14 16 18 20

mAU

0

200

400

600

800

1000

DAD1 C, Sig=280,8 Ref=off (Z:\CHROMO~1\001-0101.D)

min2 4 6 8 10 12 14 16 18 20

mAU

-15

-10

-5

0

5

DAD1 E, Sig=360,8 Ref=off (Z:\CHROMO~1\001-0101.D)

Fig. 7-8. LC-MS chromatogram of Pleuranthodium racemigerum extract redissolved in 90% aqueous methanol. Top: 210 nm; centre: 280 nm; bottom: 360 nm.

Based on the chromatograms of the whole extract and the individual fractions, fractions were

combined prior to testing for their ability to inhibit PGE2 production in murine fibroblasts.

The results of this assay are shown in Table 7-6.

Page 192: Phytochemistry and pharmacology of plants from the ginger family

168

Table 7-6. Inhibition of PGE2 production in 3T3 murine fibroblast cells by combined fractions of Pleuranthodium racemigera extract. Fractions and reference compounds were dissolved in DMSO at a concentration of 10mg/mL prior to being assayed. * Concentrations are final concentrations of test substance in assay well.

Combined fraction no. (fractions) Percent inhibition (mean±SD)

1µg/mL* 5µg/mL* 10µg/mL*

C1 (1-8) -35±3 -20±3 -33±3

C2 (9-13) 4±0 17±2 27±1

C3 (14-16) 10±1 24±6 28±3

C4 (17-18) 15±1 41±9 47±4

C5 (19-22) -1±0 19±2 -4±0

C6 (23-29) -5±0 13±1 -11±2

Aspirin (0.05mM*) 30±8

The most potent combined fraction, C4, was found to consist of a pure compound

(Compound 3). At a concentration of 10 µg/mL (subsequently calculated to be equivalent to

33.7 µM), this compound inhibited PGE2 production in 3T3 cells by close to 50%.

The LC-MS chromatogram for Compound 3 is shown in Fig. 7-9 and the mass and UV

spectra in Fig. 7-10.

Page 193: Phytochemistry and pharmacology of plants from the ginger family

169

min10 11 12 13 14 15 16 17

mAU

0

500

1000

1500

2000

DAD1 A, Sig=210,8 Ref=off (C:\DOCUME~1\ALLUSE~1\DOCUME~1\NEWCPP~1\DATA20~3\0509DATA\050929\021-0301.D)

min10 11 12 13 14 15 16 17

mAU

0

500

1000

1500

2000

DAD1 C, Sig=280,8 Ref=off (C:\DOCUME~1\ALLUSE~1\DOCUME~1\NEWCPP~1\DATA20~3\0509DATA\050929\021-0301.D)

min10 11 12 13 14 15 16 170

5000

10000

15000

20000

25000

30000

MSD1 TIC, MS File (C:\DOCUME~1\ALLUSE~1\DOCUME~1\NEWCPP~1\DATA20~3\0509DATA\050929\021-0301.D) APCI, Pos, Scan, Fra

Fig. 7-9. LC-MS chromatogram for Compound 3 from Pleuranthodium racemigerum.

m/z100 200 300 400

0

20

40

60

80

100

*MSD1 SPC, time=15.620:15.637 of C:\DOCUME~1\ALLUSE~1\DOCUME~1\NEWCPP~1\DA

Max: 6092 297

.2

107

.2 1

21.2

298

.4 2

87.2

147

.2

122

.4

189

.2

105

.2

213

.2

288

.4

351

.2

nm200 225 250 275 300 325 350 375

mAU

0

250

500

750

1000

1250

1500

1750

DAD1, 15.722 (1948 mAU,Dn1) of 021-0301.D

Fig. 7-10. Mass spectrum (M+1; top) and UV spectrum (bottom) for Compound 3 from Pleuranthodium racemigerum.

Compound 3 was subjected to further testing for cytotoxic activity and structural elucidation.

Page 194: Phytochemistry and pharmacology of plants from the ginger family

170

7.3.2.2 Cytotoxic activity of Compound 3

Compound 3 was screened for cytotoxic effect in 3T3 murine fibroblast cells. As

considerable cytotoxicity was detected (LD50 = 52.8 µM) after incubation for 3 hours (Fig 7-

11), further testing was undertaken.

Cytotoxicity of Compound 3 against 3T3 cells after 3 h

-20%

0%

20%

40%

60%

80%

100%

120%

0 20 40 60 80 100

Conc (ug/mL)

% In

hibi

tion

Fig. 7-11. Cytotoxic effect of fraction Compound 3 from Pleuranthodium racemigerum on 3T3 murine fibroblasts.

Compound 3 was tested for cytotoxic activity against one other murine cells line (P388D1

lymphoblast) and four human cell lines (Caco-2 colonic adenocarcinoma, PC3 prostate

adenocarcinoma, HepG2 hepatocyte carcinoma, and MCF7 mammary adenocarcinoma).

The cytotoxicity of Compound 3 was compared with that of the related compound curcumin

and the chemotherapeutic drug chlorambucil. The results are shown in Fig. 7-12.

Page 195: Phytochemistry and pharmacology of plants from the ginger family

171

Caco-2 (Human colonic adenocarcinoma)

-20%

0%

20%

40%

60%

80%

100%

0 1 10 100 1000

ug/mL

% in

hibi

tion

Compound 3CurcuminChlorambucil

PC3 (Human prostate adenocarcinoma)

-20%

0%

20%

40%

60%

80%

100%

0 1 10 100 1000

ug/mL

% in

hibi

tion

Compound 3CurcuminChlorambucil

HepG2 (Human hepatocyte carcinoma)

-20%

0%

20%

40%

60%

80%

100%

0 1 10 100 1000

ug/mL

% in

hibi

tion

Compound 3CurcuminChlorambucil

MCF7 (Human mammary adenocarcinoma)

-20%

0%

20%

40%

60%

80%

100%

1 10 100 1000

ug/mL

% in

hibi

tion

Compound 3CurcuminChlorambucil

P388 (Murine lymphoblast)

-20%

0%

20%

40%

60%

80%

100%

1 10 100 1000

ug/mL

% in

hibi

tion

Compound 3CurcuminChlorambucil

Fig. 7-12. Dose-response curves for cytotoxic activity of Compound 3 from Pleuranthodium racemigerum against five cell lines.

Curcumin was included for comparison and chlorambucil as a positive control.

Page 196: Phytochemistry and pharmacology of plants from the ginger family

172

LD50 values based on the dose-response curves were calculated for Compound 3 and curcumin (Table 7-7).

Table 7-7. LD50 (µM) values for cytotoxic activity of Compound 3 and curcumin against five cancer cell lines. 95% confidence intervals are shown in brackets

Cell line Compound 3 Curcumin

Caco-2 44.8 (32.6-61.8) 48.6 (38.9-60.4)

PC3 23.6 (20.6-26.9) 22.7 (17.7-28.9)

HepG2 40.6 (28.0-58.9) 43.8 (25.6-75.2)

MCF7 56.9 (45.3-71.1) 51.5 (34.8-76.0)

P388D1 117.0 (53.4-255.6) 75.7 (65.9-86.8)

The cytotoxic effects of Compound 3 closely reflected those of curcumin in the four human

cancer cell lines. In the case of the murine P388D1 cell line the LD50 for Compound 3 was

55% higher than for curcumin.

7.3.2.3 Structural elucidation of Compound 3

The NMR data obtained for Compound 3 from Z128 Pleuranthodium racemigerum are

summarised in Table 7-8.

Page 197: Phytochemistry and pharmacology of plants from the ginger family

173

Table 7-8. 1H and 13C NMR data from one-dimensional (1H and J-modulated 13C NMR) and two-dimensional correlation (COSY, HSQC and HMBC) NMR spectroscopy experiments on Compound 3 from Pleuranthodium racemigerum. δ = chemical shift (ppm), int = integration, mult = multiplicity, J = coupling constant, s = singlet, d = doublet, t = triplet, dt = doublet of a triplet, tt = triplet of a triplet

Position Chemical shift JHH (COSY) HMBC (ppm)

δH, 500 MHz (int, mult, J in Hz)

δC 126 MHz

2JCH 3JCH 4JCH

1 3.29 d (6.4) 38.3 H-2 133.4 129.6

129.6 131.6

32.5

2 5.55 dt (15.2, 6.4) 129.6 H-1, H-3 38.3 131.6

32.5 133.4

3 5.51 dt (15.2, 6.4) 131.6 H-2, H-4 32.5 129.6

29.2 38.3

4 2.07 dt (6.4, 7.3) 32.5 H-3, H-5 29.2 131.6

31.4 129.6

38.3

5 1.43 tt (7.3, 7.7) 29.2 H-4, H-6 32.5 35.0 131.6

6 1.61 dt (7.7, 7.8) 31.4 H-5, H-7 29.2 35.0

32.5 135.1

7 2.55 t (7.8) 35.0 H-6 31.4 135.1

29.2 129.6

1’ - 135.1 - 2’ 7.04 d (8.5) 129.6 H-3’ 115.3 35.0

129.6 153.8

3’ 6.76 d (8.5) 115.3 H-2’ 153.8 115.3 135.1

4’ - 153.8 - 5’ 6.76 d (8.5) 115.3 H-6’ 153.8 115.3

135.1

6’ 7.04 d (8.5) 129.6 H-5’ 115.3 35.0 129.6 153.8

1” - 133.4 - 2” 7.11 d (8.7) 129.6 H-3” 114.0 38.3

129.6 158.0

3” 6.86 d (8.7) 114.0 H-2” 158.0 114.0 133.4

4” 158.0 - 5” 6.86 d (8.7) 114.0 H-6” 158.0 114.0

133.4

6” 7.11 d (8.7) 129.6 H-5” 114.0 38.3 129.6 158.0

–OCH3 3.82 s 55.5 - 158.0

Page 198: Phytochemistry and pharmacology of plants from the ginger family

174

The 1H NMR spectrum showed signals in the region of 6.7-7.2 ppm, indicative of aromatic

protons. Signals were consistent with two aromatic rings.

Ring A. It was observed that the proton at 6.76 ppm (H-3’/H-5’) showed coupling to the

proton at 7.04 ppm (H-2’/H-6’) with a J value (coupling constant) of 8.5 Hz, suggesting

ortho-coupling. These signals each integrated to two protons, suggesting two sets of

identical protons in an aromatic ring.

The HSQC showed that the protons H-3’/H-5’ and H-2’/H-6’ were directly correlated to the

carbon signals at 115.3 and 129.6 ppm, respectively. HMBC showed long-range correlations

of the protons to the quaternary carbon signals at 153.8 ppm (C-4’) and 135.2 ppm (C-1’),

which made up the aromatic ring system with the carbon at position 4’ having a hydroxyl

substituent, as signified by its resonance in the downfield region (Fig. 7-13).

Fig. 7-13. Heteronuclear correlations in Ring A.

Ring B. Similar to what was observed for Ring A, the proton at 6.86 ppm (H-3”/H-5”)

showed coupling to the proton at 7.11 ppm (H-2”/H-6”) with a J value of 8.7 Hz, indicative

of ortho-coupling. Again, these signals integrated to two protons each, suggesting the

presence of a second aromatic ring.

Analogous to the findings for Ring A, the HSQC showed protons H-3”/H-5” and H-2”/H-6”

to be directly correlated to the carbon signals at 114.0 ppm and 129.6 ppm, respectively, and

HMBC showed long-range correlations to quaternary carbon signals at 158.0 ppm (C-4”)

and 133.4 ppm. In contrast to Ring A, a singlet that integrated to 3H was observed in the 1H

spectrum resonating at 3.82 ppm, consistent with the presence of a methoxy substituent.

HMBC showed correlation of the methoxy protons to the quaternary carbon at 158.0 ppm

(C-4”), placing the methoxy group at the C-4” position (Fig. 7-14).

1' 2' 3'4'

5'6'OH

HH

HH

Page 199: Phytochemistry and pharmacology of plants from the ginger family

175

Fig. 7-14. Heteronuclear correlations in Ring B.

Carbon bridge. The proton signals at 5.55 ppm (H-2) and 5.51 ppm (H-3) indicated olefinic

protons coupling to each other in an E position (J = 15.2 Hz). Five methylene proton signals

(H-1, H-4, H-5, H-6 and H-7) were also present in the region from 1.4 ppm to 3.3 ppm, and

the COSY experiment showed that H-1 and H-4 were vicinal to H-2 and H-3, respectively.

H-5 also showed JHH correlation to H-4 and H-6, and H-6 to H-7, as shown in Fig. 7-15.

HMBC showed correlation of H-1 to the quaternary carbon at 133.4 ppm (C-1”), which

connected C-1 to C-1”. Likewise, H-7 showed correlation to the quaternary carbon at 135.1

ppm (C-1’), linking C-7 to C-1’.

The structure of Compound 3 as shown in Fig. 7-15 was given the systematic name 1-(4”-

methoxyphenyl)-7-(4’-hydroxyphenyl)-2E-heptene. The APCI-MS data showed a (M+1)+

value of 297.2, which is consistent with a molecular formula of C20H24O2.

H3CO OH

1

2

3

4

5

6

71'

2'3'

4'

5'6'

1"

2"3"

4"

5"6"

Fig. 7-15. Chemical structure of Compound 3 from Pleuranthodium racemigerum, 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-2E-heptene.

OH3C

1"2"

3"4"

5" 6"

H

HH

H

Page 200: Phytochemistry and pharmacology of plants from the ginger family

176

7.3.2.4 Accurate Mass

The accurate mass for the Compound 3 was experimentally determined to be 319.1683,

which is consistent with the formula C20H24O2Na and the theoretical mass of 319.1674 as

shown in Table 7-9. This confirmed the molecular formula of the novel Compound 3 as

C20H24O2.

Table 7-9. Accurate mass determination for Compound 3.

Mass determined by MS C20H24O2Na

319.1683

Theoretical mass* C20H24O2Na

319.1674

Difference (δ) 0.0009 Exact mass of C20H24O2* 296.18 Molecular weight of C20H24O2* 296.40

* Data from ChemDraw Pro v. 4.01 (CambridgeSoft Corporation, Cambridge, Massachusetts, 1997)

7.3.2.5 UV spectrophotometric data and extinction coefficient

The UV spectrum of the novel Compound 3, 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-

2E-heptene, is shown in Fig. 7-16. The peak absorption (λmax) of the compound was at 203

nm, with other absorption maxima at 225 and 278 nm.

Page 201: Phytochemistry and pharmacology of plants from the ginger family

177

Wavelength (nm)200 225 250 275 300 325 350

Abs

orba

nce

(AU

)

0

0.2

0.4

0.6

0.8

1

1.2 203

225

278

334

245

214

Fig. 7-16. UV spectrum of 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-2E-heptene (Compound 3).

The extinction coefficient (wavelength-dependent molar absorptivity coefficient), ε, was

calculated for the novel Compound 3 from Pleuranthodium racemigerum using ε = A/(c

l), where A is the absorbance, c the molar concentration (4.7233468 10–5 M), and l the

distance (cuvette path; 1 cm). The peak absorbance readings and extinction coefficients ε for

Compound 3, 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-2E-heptene at 203 nm, 225 nm

and 278 nm are shown in Table 7-10.

Table 7-10. Peak absorption (λmax) and extinction coefficients (ε) of Compound 3 at 203 nm, 225 nm and 278 nm.

Abs203nm Abs225nm Abs278nm

1 1.19440 0.82368 0.16267 2 1.19610 0.82355 0.16243 3 1.18990 0.82464 0.16273 4 1.19790 0.82434 0.16282 5 1.19570 0.82616 0.16307 Mean 1.19480 0.82447 0.16274 SD 0.00301 0.00105 0.00023 ε 2.53 x 104 1.75 x 104 3.45 x 103

log(ε) 4.40 4.24 3.54

Page 202: Phytochemistry and pharmacology of plants from the ginger family

178

7.4 Discussion

Biologically active compounds were successfully isolated from two native Australian

Zingiberaceae species using bioactivity-guided fractionation. Bioactivity-guided

fractionation is a widely used method in natural products research. It is an effective means

of isolating the most active major compounds in a complex mixture, while it might be less

suitable for the identification of highly active compounds that occur in very low

concentrations. As the present work was of a preliminary nature, carried out on species that

had never been studied previously, bioactivity-guided fractionation was deemed an

appropriate methodology.

7.4.1 Curcuma australasica

From Curcuma australasica the sesquiterpene ketones zederone (Compound 1, Fig. 7-6) and

furanodien-6-one (Compound 2, Fig. 7-7) were isolated for the first time. Zederone has

previously been reported from the rhizomes of C. zedoaria (Christm.) Roscoe (Hikino et al.,

1966; Makabe et al., 2006) and C. phaeocaulis Valeton (Hou et al., 1997), and from

Chloranthus serratus (Thunb.) Roem. et Schult. (Chloranthaceae) (Kawabata et al., 1985),

while furanodien-6-one has been reported from C. zedoaria (Hikino et al., 1975; Makabe et

al., 2006), Commiphora molmol Engler (myrrh) (Brieskorn & Noble, 1983) and other

Commiphora species (Dekebo et al., 2000).

Zederone was isolated from the almost pure fraction of Curcuma australasica that most

potently inhibited the production of PGE2, suggesting this compound confers anti-

inflammatory activity to C. australasica. Zederone was found to be moderately cytotoxic

against P388D1 cells, but its toxicity was less than that of both curcumin and [6]-shogaol.

Compared with zederone, furanodien-6-one was a less potent inhibitor of PGE2. These

findings are in contrast to those of Makabe et al. (2006), who tested 11 sesquiterpenoids

isolated from C. zedoaria for in vivo anti-inflammatory activity in the 12-O-

tetradecanoylphorbol-13-acetate (TPA)-induced mouse ear oedema model and found

furanodien-6-one but not zederone to be a potent topical anti-inflammatory agent. As

oedema induced by TPA is reduced or delayed by COX inhibitors (Young et al., 1983), an

anti-inflammatory effect of zederone in this model would have been expected, given the

Page 203: Phytochemistry and pharmacology of plants from the ginger family

179

good in vitro activity in the PGE2 seen in the present study. Given the purity of Fraction 17

from which zederone was isolated, it seems unlikely that a compound other than zederone

contributed significantly to the inhibition of PGE2. It is possible that the seemingly

incongruent results obtained in the two studies reflect differences between an in vitro model

and an in vivo situation, or that stability issues played a role.

7.4.2 Pleuranthodium racemigerum

From Pleuranthodium racemigerum, the extract of which was a potent inhibitor of PGE2

production (see Chapter 6), a novel bioactive curcuminoid compound was isolated

(Compound 3, Fig. 7-15). This compound, 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-2E-

heptene, also strongly inhibited PGE2 production (IC50 ≅ 34 µM). In addition, it displayed

considerable dose-dependent cytoxicity against four human and two murine cells lines. Its

potency as a cytotoxic agent was similar to that of the related curcuminoid compound,

curcumin. These findings suggest this novel compound could warrant further investigations,

both as a potential anti-inflammatory agent and as a cytotoxic agent. Whether its

cytotoxicity precludes it from being used as an anti-inflammatory agent would need to be

determined in the first instance by animal studies, but the fact that the cytotoxic potency

resembled that of curcumin suggests that it may exert anti-inflammatory activity in vivo at

sub-toxic concentrations.

This is the first report concerning the chemistry and pharmacology of Pleuranthodium

racemigerum. As this species grows in rainforest and has a limited distribution, it is doubtful

that wild stock would be able to sustain a potential development of this plant as a source of a

new medicinal substance, but the plant could potentially be cultivated in tropical areas,

should further work indicate that an efficacious and safe medicine can be derived from it.

Page 204: Phytochemistry and pharmacology of plants from the ginger family

180

8. CONCLUDING REMARKS This chapter discusses the findings in terms of the research questions and aims of the project

(refer to Chapter 2, Section 2.16). It also identifies areas of future research arising from the

present work.

8.1 Phytochemical investigations of ginger (Zingiber officinale)

This part of the work made a significant and original contribution to the knowledge of the

phytochemistry of ginger. It included the first extensive survey of Australian commercial

and experimental ginger clones in terms of the main pungent compounds, and also provided

the first survey of steam-distilled Australian ginger oils for more than three decades.

Seventeen clones of ginger were examined for gingerol and shogaol content and in terms of

essential oil composition (Chapters 4 and 5). The aim of identifying a clone yielding

particularly high levels of pharmacologically active gingerols was achieved. The cultivar

known as ‘Jamaican’ produced significantly higher concentrations of [6]-, [8]- and [10]-

gingerol than did any other clone, and it contained these compounds in a somewhat different

ratio. A linear relationship existed between the content of [6]-, [8]- and [10]-gingerol, which

occurred in the approximate ratio of 3:1:1 in all clones except ‘Jamaican’, which contained

relatively higher levels of [8]- and [10]-gingerol.

Given that gingerols appear to confer many of the pharmacological effects of ginger (refer to

Chapter 2, Section 2.2.1.3), ‘Jamaican’ may well prove to be of particular pharmaceutical

interest, although agronomic studies would be required to confirm its suitability as a viable

source of raw material for pharmaceutical use.

‘Jamaican’ also yielded an essential oil that was distinct from that of other clones. It was

characterised by a relatively high content of sesquiterpene hydrocarbons (incl. zingiberene,

β-sesquiphellandrene and ar-curcumene) and a relatively low content of monoterpenoids

such as geranial, neral and geranyl acetate, compared with the other clones. The

combination of high gingerol levels (pungency) with distinct essential oil composition

(aroma) should make ‘Jamaican’ a cultivar of interest also to the flavour and fragrance

industry.

Page 205: Phytochemistry and pharmacology of plants from the ginger family

181

The question of whether significant phytochemical differences that are genetically

determined exist between genotypes of ginger was answered in the affirmative with the

identification of the cultivar ‘Jamaican’ and its distinct phytochemical profile, which set it

apart from 16 other clones that were grown, harvested and extracted under identical

conditions.

Apart from the ‘Jamaican’, none of the clones stood out phytochemically. This included the

experimental tetraploid clones, which did not differ significantly from their parent cultivar in

terms of gingerol content or essential oil composition. This is noteworthy, because a

previous study in Japan found tetraploid gingers to contain increased levels of gingerols

(Nakasone et al., 1999), and it demonstrates that the effects of polyploidy on secondary

metabolites are unpredictable.

With the exception of ‘Jamaican’, the steam-distilled essential oils were characterised by

very high levels (>50%) of citral (neral + geranial) and correspondingly low levels of

sesquiterpenes, when compared with published ginger oil analyses generally and the

previous survey of Australian ginger essential oil (Connell & Jordan, 1971) in particular.

These findings confirm the reputation of Australian ginger as having a ‘lemony’ aroma due

to its high citral content. Neral and geranial occurred in a fixed ratio of 2:3 in all clones.

Both aqueous and ethanolic extracts of ginger were shown to inhibit COX-1 in vitro (Chapter

3), which is in accordance with the literature (refer to Chapter 2, Section 2.2.1.3.2). A strong

positive correlation existed between inhibitory activity and the content of both [6]- and [8]-

gingerol, confirming that these compounds inhibit COX-1. Ethanol extracted gingerols more

effectively than did water. Hot water was more effective than cold water, but when ethanol

was used as a solvent, temperature did not affect the extraction efficiency, suggesting that

ethanol at ambient temperature is a highly efficient solvent of gingerols.

Most of the phytochemical investigations of ginger in the present study were carried out on

extracts prepared from fresh rhizomes at ambient temperature. Shogaols were not detected

in these extracts; [6]-shogaol was identified only in hot Soxhlet extracts. This would appear

to confirm that shogaols are not native constituents of ginger rhizome, but form as

degradation products of gingerols.

Page 206: Phytochemistry and pharmacology of plants from the ginger family

182

8.2 Screening Zingiberaceae for pharmacological activity

In this part of the work 41 taxa were screened for in vitro inhibition of PGE2 production, and

a number of the samples were also tested for antioxidant activity, inhibition of nitric oxide

production, and for modulation of natural killer cell activity. The work succeeded in making

a substantial and original contribution to the knowledge of the Zingiberaceae as medicinal

plants and a source of pharmacologically active compounds.

The hypothesis that the combination of ethnobotanical and taxonomic information is a

productive strategy to identify previously unrecognised plants with therapeutic potential was

supported by the work, which identified two native Australian species that had not

previously been known to possess pharmacological activity.

The work also provided new insights into the pharmacological activity of several known

medicinal plants. Inhibition of PGE2 was demonstrated for the first time for extracts of the

South-east Asian medicinal plants Boesenbergia rotunda and Curcuma parviflora. This

supports the traditional use of B. rotunda in the treatment of rheumatism and muscular pains

(bin Jantan et al., 2001). The inhibition of nitric oxide by Zingiber montanum was shown for

the first time. This property supports the use of the species in traditional Thai medicine for

the treatment of asthma (Piromrat et al., 1986), given that nitric oxide acts as an

inflammatory mediator in the lung, and concentrations of nitric oxide correlate with airway

inflammation in this condition (Stewart & Katial, 2007). Few Zingiberaceae species have

previously been tested for modulation of natural killer cell activity, and the finding that both

Alpinia galanga and the aqueous extract of fresh B. rotunda caused inhibition of natural

killer cell activity is novel and warrants further investigation. It would also be desirable to

conduct further work to confirm whether ethanolic extracts of B. rotunda and C. longa do in

fact significantly increase the activity of NK cells, as suggested by the present work.

Two native Australian species, Curcuma australasica and Pleuranthodium racemigerum,

were identified in this work as being potent inhibitors of PGE2 (IC50 values being similar to

that of Zingiber officinale). Neither species has a recorded history of use as a medicinal

plant, nor has either of them been the subject of previous pharmacological or phytochemical

investigations. From C. australasica two known sesquiterpene ketones, zederone and

furanodienone, were isolated. Both inhibited PGE2 production, zedorone more potently so.

These compounds have previously been reported from other Curcuma species. From P.

Page 207: Phytochemistry and pharmacology of plants from the ginger family

183

racemigerum a novel curcuminoid compound, 1-(4”-methoxyphenyl)-7-(4’-hydroxyphenyl)-

2E-heptene, was isolated. This compound was a potent inhibitor of PGE2 production (IC50 ≅

34 µM), and its structure was successfully elucidated using NMR techniques. It also

displayed dose-dependent cytotoxicity against six different cell lines, its potency being

similar to that of the related compound, curcumin. Curcuminoids and other diarylheptanoids

are widespread in some Zingiberaceae genera, such as Curcuma, Zingiber and Alpinia.

Thus this work has contributed to the understanding of the pharmacology of the

Zingiberaceae. It has also successfully identified species with previously unrecognised

pharmacological activity from genera with recognised medicinal species (Pleuranthodium

racemigerum was previously placed in the genus Alpinia) and isolated active compounds that

were chemically related to other active compounds in related species.

8.3 Direction of future research

The Zingiberaceae, with about 1100 species in more than 50 genera, clearly represents a rich

source of secondary metabolites. For most of these species little or no information about

their phytochemistry or pharmacological properties exist, so there is vast scope for further

investigations of the family as a whole. In terms of future research arising more directly

from the present work, there is also significant scope.

The ginger cultivar ‘Jamaican’ was identified as being phytochemically unique among the

clones investigated, but whether it would prove a viable source of raw material for the

pharmaceutical and/or flavour and fragrance industries would depend on its agronomic

attributes. Agronomic trials with ‘Jamaican’ would be able to determine its suitability as a

commercial crop in terms of yield, resistance to disease, etc.

Although none of the tetraploid clones included in this work proved to have altered

phytochemical characteristics, it would be worthwhile to monitor new experimental

polyploid clones, in particular with respect to increased production of gingerols.

It would also be of interest to investigate the environmental effects, in particular climate and

latitude, as well as the effect of growth period and harvest time, on the phytochemistry of

Page 208: Phytochemistry and pharmacology of plants from the ginger family

184

ginger. Such work could clarify, for example, if any of these parameters impact on the citral

content of the essential oil.

With respect to the pharmacological activity of the species included in this work, there is

clearly much more to be done. The PGE2 assay employed was expensive and time

consuming to the point of being limiting. With the initial screening done, the species that

showed inhibition greater than 50% should be subjected to further work in order to establish

accurate IC50 values. Several other bioassays relevant to potential anti-inflammatory activity

could be applied. These include inhibition of phospholipase A2 (PLA2), 5-lipoxygenase (5-

LOX) and TNF-α, and effects on COX-1 and COX-2 expression and on the key transcription

factor nuclear factor kappa B (NFκB).

The most promising samples in these assays could be targeted for bioassay-guided

fractionation provided the active compounds were not already identified. As only a subset of

species were included in the assays for antioxidant activity, inhibition of nitric oxide and

modulation of natural killer cell activity, the remainder could be screened in these assays. It

would also be desirable to screen them for cytotoxic activity.

In terms of the two native Australian species that were found to possess good PGE2

inhibitory activity, further pharmacological investigations as outlined above would be highly

desirable. Both species should also undergo further phytochemical studies in an attempt to

isolate and characterise more novel bioactive compounds. Given the paucity of information

on any Australian Zingiberaceae, it would be highly desirable to carry out a comprehensive

survey using bioassay-guided fractionation of the remainder 12 native species.

oooOooo

Page 209: Phytochemistry and pharmacology of plants from the ginger family
Page 210: Phytochemistry and pharmacology of plants from the ginger family
Page 211: Phytochemistry and pharmacology of plants from the ginger family
Page 212: Phytochemistry and pharmacology of plants from the ginger family
Page 213: Phytochemistry and pharmacology of plants from the ginger family
Page 214: Phytochemistry and pharmacology of plants from the ginger family

190

APPENDIX B: QUALITY ASSESSMENT OF CLINICAL TRIALS OF GINGER IN OSTEOATHRITIS

Assessment of the quality of the reporting of two randomised trials of ginger in osteoarthritis (with ginger extract as the sole active intervention). Checklist based on the revised CONSORT Statement (Moher et al., 2001) except items marked with an asterisk (*), which were adapted from the CONSORT Statement elaborated for the reporting of herbal interventions (Gagnier et al., 2006). Each item was given a score of 0, 0.5 or 1, depending on whether the information was provided not at all or to an unsatisfactory extent (0), to some extent (0.5), or to a satisfactory or mostly satisfactory extent (1). PAPER SECTION And topic

Item Description

Bliddal et al. 2000

Wigler et al. 2003

TITLE & ABSTRACT

1 How participants were allocated to interventions (e.g., "random allocation", "randomized", or "randomly assigned").

1

1

INTRODUCTION Background

2 Scientific background and explanation of rationale.

0 0

METHODS Participants

3 Eligibility criteria for participants and the settings and locations where the data were collected.

0.5

0.5

Interventions 4 Precise details of the interventions intended for each group and how and when they were actually administered.

Elaborated in 4A-4F

Herbal medicinal product name

4A* E.g. Latin binomial, extract name, brand name, name of manufacturer

0.5 1

Characteristics of the herbal product

4B* Incl. plant parts, fresh or dried or extract, solvent details, method of authentication, lot number of raw material, details of voucher specimen or retention sample

0

0.5

Dosage regimen and qualitative description

4C* Dosage, duration of administration, how these were determined; content of quantified herbal constituents per dosage unit; details of additives; for standardised products, the quantity of each active/marker per dosage unit

0

1

Qualitative testing 4D* Chemical fingerprint and methods for obtaining this; description of any special/purity testing; standardisation: what to standardise and how

0

0

Placebo/control group

4E* The rationale for the type of control/placebo used

1 1

Practitioner 4F* A description of the practitioners (e.g., training and practice experience) that are a part of the intervention

0.5

0

Objectives 5 Specific objectives and hypotheses. 0.5 0.5 Outcomes 6 Clearly defined primary and secondary

outcome measures and, when applicable, any methods used to enhance the quality of measurements (e.g., multiple observations, training of assessors).

1

1

Sample size 7 How sample size was determined and, when applicable, explanation of any interim analyses and stopping rules.

1

0

Page 215: Phytochemistry and pharmacology of plants from the ginger family

191

Randomization -Sequence generation

8 Method used to generate the random allocation sequence, including details of any restrictions (e.g., blocking, stratification)

1

1

Randomization -Allocation concealment

9 Method used to implement the random allocation sequence (e.g., numbered containers or central telephone), clarifying whether the sequence was concealed until interventions were assigned.

0

0

Randomization - Implementation

10 Who generated the allocation sequence, who enrolled participants, and who assigned participants to their groups.

0

0

Blinding (masking) 11 Whether or not participants, those administering the interventions, and those assessing the outcomes were blinded to group assignment. When relevant, how the success of blinding was evaluated.

0.5

0.5

Statistical methods 12 Statistical methods used to compare groups for primary outcome(s); Methods for additional analyses, such as subgroup analyses and adjusted analyses.

1

1

RESULTS Participant flow

13 Flow of participants through each stage (a diagram is strongly recommended). Specifically, for each group report the numbers of participants randomly assigned, receiving intended treatment, completing the study protocol, and analyzed for the primary outcome. Describe protocol deviations from study as planned, together with reasons.

1

1

Recruitment 14 Dates defining the periods of recruitment and follow-up.

0 0

Baseline data 15 Baseline demographic and clinical characteristics of each group.

1 1

Numbers analyzed 16 Number of participants (denominator) in each group included in each analysis and whether the analysis was by "intention-to-treat". State the results in absolute numbers when feasible (e.g., 10/20, not 50%).

1

1

Outcomes and estimation

17 For each primary and secondary outcome, a summary of results for each group, and the estimated effect size and its precision (e.g., 95% confidence interval).

1

1

Ancillary analyses 18 Address multiplicity by reporting any other analyses performed, including subgroup analyses and adjusted analyses, indicating those pre-specified and those exploratory.

1

-

Adverse events 19 All important adverse events or side effects in each intervention group.

1 1

DISCUSSION Interpretation

20 Interpretation of the results, taking into account study hypotheses, sources of potential bias or imprecision and the dangers associated with multiplicity of analyses and outcomes.

1

1

Generalizability 21 Generalizability (external validity) of the trial findings.

1 1

Overall evidence 22 General interpretation of the results in the context of current evidence.

1 1

Total score 17.5 17.0

Page 216: Phytochemistry and pharmacology of plants from the ginger family

192

APPENDIX C: INHIBITION OF PGE2 IN 3T3 CELLS Concentrations shown are final concentrations in assay well. Mean value±SEM; n=3 expect * (n=2), + (n=4) and ^ (n=6).

Taxon ID Concentration (µg/mL)

Percent inhibition of PGE2

Alpinia arctiflora Z129 100.0 -18.51±1.81 50.0 12.08±1.39 10.0 1.41±0.26 Alpinia caerulea Z102 94.8 33.81±8.99 47.4 35.37±5.61 9.5 -15.16±2.85 Alpinia calcarata* Z49 500.0 54.96±4.66 5.0 8.52±0.25 Alpinia galanga Z103 91.8 47.02±0.53 45.9 42.88±4.34 9.2 22.80±4.23 Alpinia luteocarpa Z111 100.0 0.77±0.08 50.0 20.31±0.35 10.0 1.03±0.07 Alpinia malaccensis* Z48 500.0 28.98±1.62 5.0 -22.37±0.29 Alpinia modesta Z127 100.0 -3.60±0.73 50.0 23.65±1.49 10.0 -10.67±0.80 Alpinia mutica* Z08 500.0 48.04±0.17 5.0 -29.26±2.99 Alpinia purpurea 'Eileen McDonald'* Z11 500.0 47.36±0.99 5.0 15.09±0.13 Alpinia spectabile 'Giant Orange'* Z53 500.0 18.23±1.49 5.0 -18.08±0.76 Alpinia zerumbet* Z52 500.0 42.98±2.80 5.0 18.70±2.47 Boesenbergia rotunda Z104 94.2 50.52±1.69 47.1 45.60±2.53 9.4 21.77±0.84

Page 217: Phytochemistry and pharmacology of plants from the ginger family

193

Costus barbatus Z112 100.0 -45.67±2.76 50.0 -30.74±1.78 10.0 -59.31±2.29 Costus leucanthus Z113 100.0 -28.14±1.20 50.0 -7.79±0.29 10.0 -25.76±1.79 Costus malortieanus Z114 100.0 -33.55±2.37 50.0 -16.45±2.00 10.0 -17.10±1.25 Costus productus Z116 100.0 -0.22±0.05 50.0 6.28±1.00 10.0 -27.06±1.34 Costus pulverulentus Z115 100.0 -40.48±2.31 50.0 7.14±0.16 10.0 -20.56±0.76 Costus tappenbeckianus* Z15 500.0 -4.66±0.52 5.0 -24.54±1.38 Curcuma australasica^ Z101 100.0 56.21±7.41 50.0 51.11±11.81 10.0 19.90±3.53 Curcuma cordata Z117 100.0 7.63±0.51 50.0 14.61±0.10 10.0 9.33±0.51 Curcuma longa^ Z106 100.0 60.15±4.64 50.0 57.69±6.45 10.0 49.83±14.91 Curcuma parviflora Z118 100.0 66.72±4.75 50.0 51.38±1.59 10.0 30.11±1.79 Elettaria cardamomum* Z12 500.0 44.29±0.00 5.0 14.91±0.01 Etlingera australasica Z107 92.4 28.50±6.56 46.2 22.41±9.84 9.2 -14.12±3.28 Etlingera elatior 'Burma torch'* Z04 500.0 -0.37±0.01 5.0 -14.36±0.80 Hedychium coronarium* Z10 500.0 41.60±0.96 5.0 -9.02±0.15

Page 218: Phytochemistry and pharmacology of plants from the ginger family

194

Hornstedtia scottiana Z126 100.0 0.97±0.09 50.0 10.15±0.52 10.0 3.33±0.22 Kaempferia galanga* Z05 500.0 61.31±5.48 5.0 7.39±0.35 Kaempferia rotunda Z120 100.0 -39.61±2.58 50.0 -25.32±1.49 10.0 -52.60±3.67 Pleuranthodium racemigerum^ Z128 100.0 59.40±3.76 50.0 46.85±6.85 10.0 39.00±5.69 Renealmia cernua Z121 100.0 17.45±1.63 50.0 18.26±0.30 10.0 11.36±0.53 Scaphochlamys biloba* Z01 500.0 35.21±2.07 5.0 -33.13±4.92 Scaphochlamys kunstleri* Z122 100.0 26.13±3.43 50.0 31.09±0.22 10.0 24.67±1.17 Tapeinochilos ananassae Z125 100.0 4.30±0.31 50.0 16.31±0.68 10.0 14.28±0.65 Zingiber longipedunculatum* Z06 500.0 26.99±0.50 5.0 -4.23±0.11 Zingiber officinale Z108 100.0 50.44±1.69 50.0 47.62±5.69 10.0 11.02±2.53 Zingiber ottensii Z124 100.0 -28.17±2.77 50.0 -19.11±4.05 10.0 -3.62±0.18 Zingiber montanum* Z02 500.0 64.72±5.97 5.0 9.40±0.33 Zingiber montanum Z105 100.0 49.35±4.12 50.0 47.54±1.80 10.0 22.93±1.80 Zingiber spectabile Z17 500.0 33.99±1.91 5.0 -17.48±0.62

Page 219: Phytochemistry and pharmacology of plants from the ginger family

195

Zingiber sp. 'Dwarf Apricot'+ Z55 500.0 11.65±1.82 5.0 -7.49±0.48 Zingiber/Etlingera (Aniseed ginger)* Z19 500.0 38.83±0.90 5.0 10.92±0.50

Page 220: Phytochemistry and pharmacology of plants from the ginger family

196

REFERENCES Abe, Y, Hashimoto, S, Horie, T (1999). Curcumin inhibition of inflammatory cytokine

production by human peripheral blood monocytes and alveolar macrophages. Pharmacological Research 39: 41-47.

Abourashed, EA, Muhammad, I, Resch, M, Bauer, R, el-Feraly, FS, Hufford, CD (2001).

Inhibitory effects of maesanin and analogs on arachidonic acid metabolizing enzymes. Planta Medica 67: 360-361.

Abuharfeil, NM, Salim, M, Von Kleist, S (2001). Augmentation of natural killer cell activity

in vivo against tumour cells by some wild plants from Jordan. Phytotherapy Research 15: 109-113.

Adams, PA (1995). Identification of essential oil components by gas chromatography/mass

spectroscopy. Allured: Carol Stream, Illinois. Afzal, M, Al-Hadidi, D, Menon, M, Pesek, J, Dhami, MS (2001). Ginger: an ethnomedical,

chemical and pharmacological review. Drug Metabolism & Drug Interactions. 18: 159-190.

Agrawal, AK, Rao, CV, Sairam, K, Joshi, VK, Goel, RK (2000). Effect of Piper longum

Linn, Zingiber officinalis Linn and Ferula species on gastric ulceration and secretion in rats. Indian Journal of Experimental Biology 38: 994-998.

Agrawal, PK, Singh, SB, Thakur, RS (1984). 13C NMR spectral analysis of spirostanol and

furostanol glycosides of Paris polyphylla and Costus speciosus. Indian Journal of Pharmaceutical Sciences 46: 158-160.

Ahmed, RS, Sharma, SB (1997). Biochemical studies on combined effects of garlic (Allium

sativum Linn) and ginger (Zingiber officinale Rosc) in albino rats. Indian Journal of Experimental Biology 35: 841-843.

Ahmed, RS, Seth, V, Banerjee, BD (2000a). Influence of dietary ginger (Zingiber officinale

Rosc) on antioxidant defense system in rat: comparison with ascorbic acid. Indian Journal of Experimental Biology 38: 604-606.

Ahmed, RS, Seth, V, Pasha, ST, Banerjee, BD (2000b). Influence of dietary ginger (Zingiber

officinale Rosc) on oxidative stress induced by malathion in rats. Food & Chemical Toxicology 38: 443-450.

Ahsan, H, Parveen, N, Khan, NU, Hadi, SM (1999). Pro-oxidant, anti-oxidant and cleavage

activities on DNA of curcumin and its derivatives demethoxycurcumin and bisdemethoxycurcumin. Chemico-Biological Interactions 121: 161-175.

Akiyama, K, Kikuzaki, H, Aoki, T, Okuda, A, Lajis, NH, Nakatani, N (2006). Terpenoids

and a diarylheptanoid from Zingiber ottensii. Journal of Natural Products 69: 1637-1640.

Page 221: Phytochemistry and pharmacology of plants from the ginger family

197

Al-Zuhair, H, El-Sayeh, B, Ameen, HA, Al-Shoora, H (1996). Pharmacological studies of

cardamom oil in animals. Pharmacological Research 34: 79-82. Ali, A, Kaur, G, Hamid, H, Abdullah, T, Ali, M, Niwa, M, Alam, MS (2003). Terminoside

A, a new triterpene glycoside from the bark of Terminalia arjuna inhibits nitric oxide production in murine macrophages. Journal of Asian Natural Products Research 5: 137-142.

Altman, RD, Marcussen, KC (2001). Effects of a ginger extract on knee pain in patients with

osteoarthritis. Arthritis & Rheumatism. 44: 2531-2538. Anonymous (2001). Japanese Pharmacopoeia (English version). 14th edn.

<http://jpdb.nihs.go.jp/jp14e/>. Anonymous (2002). WHO Traditional Medicine Strategy 2002-2005. World Health

Organization: Geneva. Anonymous (2004). British Pharmacopoeia 2004. The Stationary Office: London. Anonymous (2005). Pharmacopoeia of the People's Republic of China. English edn. People's

Medical Publishing House: Beijing. Anonymous (2007). British Pharmacopoeia. The Stationary Office: London. Anonymous (2008). Global Biodiversity Information Facility.

<http://data.gbif.org/welcome.htm>. Anto, RJ, Mukhopadhyay, A, Denning, K, Aggarwal, BB (2002). Curcumin

(diferuloylmethane) induces apoptosis through activation of caspase-8, BID cleavage and cytochrome c release: its suppression by ectopic expression of Bcl-2 and Bcl-xl. Carcinogenesis 23: 143-150.

Antunes, LM, Araujo, MC, Darin, JD, Bianchi, ML (2000). Effects of the antioxidants

curcumin and vitamin C on cisplatin-induced clastogenesis in Wistar rat bone marrow cells. Mutation Research 465: 131-137.

Antunes, LM, Darin, JD, Bianchi, NdL (2001). Effects of the antioxidants curcumin or

selenium on cisplatin-induced nephrotoxicity and lipid peroxidation in rats. Pharmacological Research 43: 145-150.

Apariman, S, Ratchanon, S, Wiriyasirivej, B (2006). Effectiveness of ginger for prevention

of nausea and vomiting after gynecological laparoscopy. Journal of the Medical Association of Thailand 89: 2003-2009.

Arctander, S (1994). Perfume and Flavor Materials of Natural Origin. Allured Publishing:

Carol Stream, IL.

Page 222: Phytochemistry and pharmacology of plants from the ginger family

198

Arfeen, Z, Owen, H, Plummer, JL, Ilsley, AH, Sorby-Adams, RA, Doecke, CJ (1995). A double-blind randomized controlled trial of ginger for the prevention of postoperative nausea and vomiting. Anaesthesia & Intensive Care 23: 449-452.

Asai, A, Nakagawa, K, Miyazawa, T (1999). Antioxidative effects of turmeric, rosemary and

capsicum extracts on membrane phospholipid peroxidation and liver lipid metabolism in mice. Bioscience, Biotechnology & Biochemistry 63: 2118-2122.

Asfaw, N, Abegaz, B (1995). Chemical constituents of the essential oils of Zingiber

officinale Roscoe cultivated in Ethiopia. Sinet, an Ethiopian Journal of Science 18: 133-137.

Atsumi, T, Fujisawa, S, Tonosaki, K (2005a). Relationship between intracellular ROS

production and membrane mobility in curcumin- and tetrahydrocurcumin-treated human gingival fibroblasts and human submandibular gland carcinoma cells. Oral Diseases 11: 236-242.

Atsumi, T, Murakami, Y, Shibuya, K, Tonosaki, K, Fujisawa, S (2005b). Induction of

cytotoxicity and apoptosis and inhibition of cyclooxygenase-2 gene expression, by curcumin and its analog, alpha-diisoeugenol. Anticancer Research 25: 4029-4036.

Azmi Muda, M, Nor Azma, Y, Lindayani, Halijah, I, Norzulaani, K, Norsaadah Abdul, R

(2002). Establishment of tissue culture and antibacterial properties of selected species in the family Zingiberaceae. Towards modernisation of research and technology in herbal industries. Proceedings of the Seminar on Medicinal and Aromatic Plants, 24-25 July 2001. 2002. pp 273-276.

Balick, M (1990). Ethnobotany and the identification of therapeutic agents from the

rainforest. In: Chadwick, DJ, Marsh, J (eds). Bioactive Compounds from Plants. Wiley: Chichester, U. K.

Banerjee, M, Tripathi, LM, Srivastava, VML, Puri, A, Shukla, R (2003). Modulation of

inflammatory mediators by ibuprofen and curcumin treatment during chronic inflammation in rat. Immunopharmacology & Immunotoxicology 25: 213-224.

Banjerdpongchai, R, Wilairat, P (2005). Effects of water-soluble antioxidants and

MAPKK/MEK inhibitor on curcumin-induced apoptosis in HL-60 human leukemic cells. Asian Pacific Journal of Cancer Prevention: Apjcp 6: 282-285.

Baolin, L, Inami, Y, Tanaka, H, Inagaki, N, Iinuma, M, Nagai, H (2004). Resveratrol inhibits

the release of mediators from bone marrow-derived mouse mast cells in vitro. Planta Medica 70: 305-309.

Barclay, LR, Vinqvist, MR, Mukai, K, Goto, H, Hashimoto, Y, Tokunaga, A, Uno, H (2000).

On the antioxidant mechanism of curcumin: classical methods are needed to determine antioxidant mechanism and activity. Organic Letters 2: 2841-2843.

Bartley, JP (1995). A new method for the determination of pungent compounds in ginger

(Zingiber officinale). Journal of the Science of Food & Agriculture 68: 215-222.

Page 223: Phytochemistry and pharmacology of plants from the ginger family

199

Bartley, JP, Jacobs, AL (2000). Effects of drying on flavour compounds in Australian-grown ginger (Zingiber officinale). Journal of the Science of Food & Agriculture 80: 209-215.

Batkhuu, J, Hattori, K, Takano, F, Fushiya, S, Oshiman, K-i, Fujimiya, Y (2002).

Suppression of NO production in activated macrophages in vitro and ex vivo by neoandrographolide isolated from Andrographis paniculata. Biological & Pharmaceutical Bulletin 25: 1169-1174.

Baum, L, Ng, A (2004). Curcumin interaction with copper and iron suggests one possible

mechanism of action in Alzheimer's disease animal models. Journal of Alzheimer's Disease 6: 367-377; discussion 443-369.

Bengmark, S (2006). Curcumin, an atoxic antioxidant and natural NFkappaB,

cyclooxygenase-2, lipooxygenase, and inducible nitric oxide synthase inhibitor: a shield against acute and chronic diseases. Jpen: Journal of Parenteral & Enteral Nutrition 30: 45-51.

Betancor-Fernandez, A, Perez-Galvez, A, Sies, H, Stahl, W (2003). Screening

pharmaceutical preparations containing extracts of turmeric rhizome, artichoke leaf, devil's claw root and garlic or salmon oil for antioxidant capacity. Journal of Pharmacy & Pharmacology 55: 981-986.

Bhandari, U, Sharma, JN, Zafar, R (1998). The protective action of ethanolic ginger

(Zingiber officinale) extract in cholesterol fed rabbits. Journal of Ethnopharmacology 61: 167-171.

Bhattarai, S, Tran, VH, Duke, CC (2001). The stability of gingerol and shogaol in aqueous

solutions. Journal of Pharmaceutical Sciences. 90: 1658-1664. Bhattarai, S, Tran, VH, Duke, CC (2007). Stability of [6]-gingerol and [6]-shogaol in

simulated gastric and intestinal fluids. Journal of Pharmaceutical and Biomedical Analysis 45: 648-653.

Bierhaus, A, Zhang, Y, Quehenberger, P, Luther, T, Haase, M, Muller, M, Mackman, N,

Ziegler, R, Nawroth, PP (1997). The dietary pigment curcumin reduces endothelial tissue factor gene expression by inhibiting binding of AP-1 to the DNA and activation of NF-κB. Thrombosis and Haemostasis 77: 772-782.

bin Jantan, I, Basni, I, Ahmad, AS, Ali, NAM, Ahmad, AR, Ibrahim, H (2001). Constituents

of the rhizome oils of Boesenbergia pandurata (Roxb.) Schlecht from Malaysia, Indonesia and Thailand. Flavour & Fragrance Journal 16: 110-112.

Biswas, SK, McClure, D, Jimenez, LA, Megson, IL, Rahman, I (2005). Curcumin induces

glutathione biosynthesis and inhibits NF-kappaB activation and interleukin-8 release in alveolar epithelial cells: mechanism of free radical scavenging activity. Antioxidants & Redox Signaling 7: 32-41.

Bliddal, H, Rosetzsky, A, Schlichting, P, Weidner, MS, Andersen, LA, Ibfelt, HH,

Christensen, K, Jensen, ON, Barslev, J (2000a). A randomized, placebo-controlled,

Page 224: Phytochemistry and pharmacology of plants from the ginger family

200

cross-over study of ginger extracts and ibuprofen in osteoarthritis. Osteoarthritis & Cartilage 8: 9-12.

Bliddal, H, Rosetzsky, A, Schlichting, P, Weidner, MS, Andersen, LA, Ibfelt, HH,

Christensen, K, Jensen, ON, Barslev, J (2000b). A randomized, placebo-controlled, cross-over study of ginger extracts and ibuprofen in osteoarthritis. Osteoarthritis & Cartilage. 8: 9-12.

Blumenthal, ME (1998). The Complete German Commission E Monographs. American

Botanical Council: Austin, Texas. Bone, ME, Wilkinson, DJ, Young, JR, McNeil, J, Charlton, S (1990). Ginger root - a new

antiemetic. The effect of ginger root on postoperative nausea and vomiting after major gynaecological surgery. Anaesthesia 45: 669-671.

Bordia, A, Verma, SK, Srivastava, KC (1997). Effect of ginger (Zingiber officinale Rosc.)

and fenugreek (Trigonella foenum-graecum L.) on blood lipids, blood sugar and platelet aggregation in patients with coronary artery disease. Prostaglandins Leukotrienes & Essential Fatty Acids 56: 379-384.

Boukouvalas, J, Wang, J-X, Marion, O, Ndzi, B (2006). Synthesis and stereochemistry of the

antitumor diterpenoid (+)-zerumin B. The Journal of Organic Chemistry 71: 6670-6673. Brater, DC, Harris, C, Redfern, JS, Gertz, BJ (2001). Renal effects of COX-2-selective

inhibitors. American Journal of Nephrology 21: 1-15. Bremner, P, Heinrich, M (2002). Natural products as targeted modulators of the nuclear

factor-kappa B pathway. Journal of Pharmacy and Pharmacology 54: 453-472. Brennan, P, O'Neill, LA (1998). Inhibition of nuclear factor kappaB by direct modification in

whole cells--mechanism of action of nordihydroguaiaritic acid, curcumin and thiol modifiers. Biochemical Pharmacology 55: 965-973.

Brieskorn, CH, Noble, P (1983). Furanosesquiterpenes from the essential oil of myrrh.

Phytochemistry 22: 1207-1211. Brouet, I, Ohshima, H (1995). Curcumin, an anti-tumour promoter and anti-inflammatory

agent, inhibits induction of nitric oxide synthase in activated macrophages. Biochemical & Biophysical Research Communications 206: 533-540.

Bruhn, JG, Holmstedt, B (1981). Ethnopharmacology: objectives, principles and

perspectives. In: Beal, JL, Reinhard, E (eds). Natural Products as Medicinal Agents. Hippokrates Verl.: Stuttgart. pp 405-430.

Bucar, F, Resch, M, Bauer, R, Burits, M, Knauder, E, Schubert-Zsilavecz, M (1998). 5-

Methylflavasperone and rhamnetin from Guiera senegalensis and their antioxidative and 5-lipoxygenase inhibitory activity. Pharmazie 53: 875-878.

Cady, RK, Schreiber, CP, Beach, ME, Hart, CC (2005). Gelstat Migraine (sublingually

administered feverfew and ginger compound) for acute treatment of migraine when

Page 225: Phytochemistry and pharmacology of plants from the ginger family

201

administered during the mild pain phase. Medical Science Monitor: International Medical Journal of Experimental and Clinical Research 11: PI65-69.

Calder, PC (2005). Polyunsaturated fatty acids and inflammation. Biochemical Society

Transactions 33: 423-427. Calixto, JB, Otuki, MF, Santos, AR (2003). Anti-inflammatory compounds of plant origin.

Part I. Action on arachidonic acid pathway, nitric oxide and nuclear factor kappa B (NF-kappaB). Planta Medica 69: 973-983.

Calixto, JB, Campos, MM, Otuki, MF, Santos, AR (2004). Anti-inflammatory compounds of

plant origin. Part II. modulation of pro-inflammatory cytokines, chemokines and adhesion molecules. Planta Medica 70: 93-103.

Camacho-Barquero, L, Villegas, I, Sanchez-Calvo, JM, Talero, E, Sanchez-Fidalgo, S,

Motilva, V, Alarcon de la Lastra, C (2007). Curcumin, a Curcuma longa constituent, acts on MAPK p38 pathway modulating COX-2 and iNOS expression in chronic experimental colitis. International Immunopharmacology 7: 333-342.

Cao, G, Prior, RL (1999). Measurement of oxygen radical absorbance capacity in biological

samples. Methods in Enzymology, Vol. Volume 299. Academic Press. pp 50-62. Careddu, P (1999). Motion sickness in children: results of a double-blind study with ginger

(Zintona®) and dimenhydrinate. Healthnotes Review of Complementary and Integrative Medicine 6: 102-107.

Carlson, TJS (2002). Medical ethnobotanical research as a method to identify bioactive

plants to treat infectious diseases. In: Iwu, MM, Wootton, JC (eds). Ethnomedicine and Drug Discovery. Elsevier: Amsterdam. pp 45-53.

Chan, MM-Y (1995). Inhibition of tumor necrosis factor by curcumin, a phytochemical.

Biochemical Pharmacology 49: 1551-1556. Chan, MM-Y, Ho, C-T, Huang, H-I (1995). Effects of three dietary phytochemicals from tea,

rosemary and turmeric on inflammation-induced nitrite production. Cancer Letters 96: 23-29.

Chan, MM, Huang, HI, Fenton, MR, Fong, D (1998). In vivo inhibition of nitric oxide

synthase gene expression by curcumin, a cancer preventive natural product with anti-inflammatory properties. Biochemical Pharmacology 55: 1955-1962.

Chang, CP, Chang, JY, Wang, FY, Chang, JG (1995). The effect of Chinese medicinal herb

Zingiberis rhizoma extract on cytokine secretion by human peripheral blood mononuclear cells. Journal of Ethnopharmacology 48: 13-19.

Chauhan, DP (2002). Chemotherapeutic potential of curcumin for colorectal cancer. Current

Pharmaceutical Design 8: 1695-1706. Cheah, PB, Abu, HN (2000). Natural antioxidant extract from galangal (Alpinia galanga) for

minced beef. Journal of the Science of Food & Agriculture 80: 1565-1571.

Page 226: Phytochemistry and pharmacology of plants from the ginger family

202

Chen, CC, Kuo, MC, Ho, CT (1986a). High performance liquid chromatographic

determination of pungent gingerol compounds of ginger (Zingiber officinale Roscoe). Journal of Food Science 51: 1364-1365.

Chen, CC, Kuo, MC, Wu, CM, Ho, CT (1986b). Pungent compounds of ginger (Zingiber

officinale Roscoe) extracted by liquid carbon dioxide. Journal of Agricultural & Food Chemistry 34: 477-480.

Chen, H, Zhang, ZS, Zhang, YL, Zhou, DY (1999). Curcumin inhibits cell proliferation by

interfering with the cell cycle and inducing apoptosis in colon carcinoma cells. Anticancer Research 19: 3675-3680.

Chen, J, Wanming, D, Zhang, D, Liu, Q, Kang, J (2005). Water-soluble antioxidants

improve the antioxidant and anticancer activity of low concentrations of curcumin in human leukemia cells. Pharmazie 60: 57-61.

Chen, YC, Shen, SC, Chen, LG, Lee, TJ, Yang, LL (2001). Wogonin, baicalin, and baicalein

inhibition of inducible nitric oxide synthase and cyclooxygenase-2 gene expressions induced by nitric oxide synthase inhibitors and lipopolysaccharide. Biochemical Pharmacology 61: 1417-1427.

Chevallier, A (2001). Encyclopedia of Medicinal Plants. Revised edn. Dorling Kindersley:

St. Leonards, NSW. Chirangini, P, Sharma, GJ, Sinba, SK (2004). Sulfur free radical reactivity with curcumin as

reference for evaluating antioxidant properties of medicinal Zingiberales. Journal of Environmental Pathology Toxicology and Oncology 23: 227.

Chittumma, P, Kaewkiattikun, K, Wiriyasiriwach, B (2007). Comparison of the effectiveness

of ginger and vitamin B6 for treatment of nausea and vomiting in early pregnancy: a randomized double-blind controlled trial. Journal of the Medical Association of Thailand 90: 15-20.

Cho, J-W, Park, K, Kweon, GR, Jang, B-C, Baek, W-K, Suh, M-H, Kim, C-W, Lee, K-S,

Suh, S-I (2005). Curcumin inhibits the expression of COX-2 in UVB-irradiated human keratinocytes (HaCaT) by inhibiting activation of AP-1: p38 MAP kinase and JNK as potential upstream targets. Experimental & Molecular Medicine 37: 186-192.

Chrubasik, S, Pittler, MH, Roufogalis, BD (2005). Zingiberis rhizoma: a comprehensive

review on the ginger effect and efficacy profiles. Phytomedicine 12: 684-701. Chun, KS, Kang, JY, Kim, OH, Kang, H, Surh, YJ (2002). Effects of yakuchinone A and

yakuchinone B on the phorbol ester-induced expression of COX-2 and iNOS and activation of NF-kappaB in mouse skin. Journal of Environmental Pathology, Toxicology & Oncology. 21: 131-139.

Chung, WY, Park, JH, Kim, MJ, Kim, HO, Hwang, JK, Lee, SK, Park, KK (2007).

Xanthorrhizol inhibits 12-O-tetradecanoylphorbol-13-acetate-induced acute inflammation and two-stage mouse skin carcinogenesis by blocking the expression of

Page 227: Phytochemistry and pharmacology of plants from the ginger family

203

ornithine decarboxylase, cyclooxygenase-2 and inducible nitric oxide synthase through mitogen-activated protein kinases and/or the nuclear factor-kappa B. Carcinogenesis 28: 1224-1231.

Connell, DW (1969). The pungent principles of ginger and their importance in certain ginger

products. Food Technology in Australia 21: 570-575. Connell, DW, Sutherland, MD (1969). A re-examination of gingerol, shogaol and zingerone,

the pungent principles of ginger (Zingiber officinale Roscoe). Australian Journal of Chemistry 22: 1033-1043.

Connell, DW, Jordan, RA (1971). Composition and distinctive volatile flavour

characteristics of the essential oil from Australian-grown ginger (Zingiber officinale). Journal of the Science of Food & Agriculture 22: 93-95.

Connell, DW (1986). Chemical studies of some constituents of Zingiber officinale,

University of Queensland, Brisbane. (PhD thesis). Cotton, CM (1996). Ethnobotany: Principles and Applications. John Wiley & Sons:

Chichester, U.K. Cox, P (1994). The ethnobotanical approach to drug discovery: strengths and limitations. In:

Chadwick, DJ, Marsh, J (eds). Ethnobotany and the Search for New Drugs. Wiley: Chichester, U. K.

Cragg, GM, Newman, DJ (2002). Drugs from nature: past achievements, future prospects.

In: Iwu, MM, Wootton, JC (eds). Ethnomedicine and Drug Discovery. Elsevier: Amsterdam. pp 23-37.

Currier, NL, Miller, SC (2000). Natural killer cells from aging mice treated with extracts

from Echinacea purpurea are quantitatively and functionally rejuvenated. Experimental Gerontology 35: 627-639.

Cuzzocrea, S (2006). Role of nitric oxide and reactive oxygen species in arthritis. Current

Pharmaceutical Design 12: 3551-3570. Daniel, S, Limson, JL, Dairam, A, Watkins, GM, Daya, S (2004). Through metal binding,

curcumin protects against lead- and cadmium-induced lipid peroxidation in rat brain homogenates and against lead-induced tissue damage in rat brain. Journal of Inorganic Biochemistry 98: 266-275.

Das, KC, Das, CK (2002). Curcumin (diferuloylmethane), a singlet oxygen ((1)O(2))

quencher. Biochemical & Biophysical Research Communications 295: 62-66. Daswani, L, Bohra, A (2000). Antibacterial potential of Elettaria cardamomum (Chotti

elaichi) against human pathogenic bacteria, Staphylococcus aureus. Journal of Eco-Physiology 3: 135-137.

Davis, PB, Drumm, ML (2004). Some like it hot: curcumin and CFTR. Trends in Molecular

Medicine 10: 473-475.

Page 228: Phytochemistry and pharmacology of plants from the ginger family

204

de Kraker, JW, Franssen, MCR, de Groot, A, Shibata, T, Bouwmeester, HJ (2001).

Germacrenes from fresh costus roots. Phytochemistry 58: 481-487. Deans, KA, Sattar, N (2006). "Anti-inflammatory" drugs and their effects on type 2 diabetes.

Diabetes Technology & Therapeutics 8: 18-27. Dedov, VN, Tran, VH, Duke, CC, Connor, M, Christie, MJ, Mandadi, S, Roufogalis, BD

(2002). Gingerols: a novel class of vanilloid receptor (VR1) agonists. British Journal of Pharmacology 137: 793- 798.

Dekebo, A, Dagne, E, Hansen, LK, Gautun, OR, Aasen, AJ (2000). Crystal structures of two

furanosesquiterpenes from Commiphora sphaerocarpa. Tetrahedron Letters 41: 9875-9878.

Denniff, P, Macleod, I, Whiting, DA (1981). Synthesis of the (±)-[n]-gingerols (pungent

principles of ginger) and related compounds through regioselective aldol condensations: relative pungency assays. Journal of the Chemical Society, Perkin Transactions I: 82-87.

Dewick, PM (1997). Medicinal Natural Products - a biosynthetic approach. John Wiley &

Sons: Chichester. Di Paola, R, Mazzon, E, Muià , C, Crisafulli, C, Genovese, T, Di Bella, P, Esposito, E,

Menegazzi, M, Meli, R, Suzuki, H, Cuzzocrea, S (2007). Protective effect of Hypericum perforatum in zymosan-induced multiple organ dysfunction syndrome: relationship to its inhibitory effect on nitric oxide production and its peroxynitrite scavenging activity. Nitric Oxide: Biology And Chemistry 16: 118-130.

Dick, H (1992). Edible plants. Red ginger, Hornstedtia scottiana; green ginger, Amomum

dallachyi; blue ginger, Alpinia caerulea; the ginger family. Australian Plants 16: 239-243, 249, 253.

Divya, CS, Pillai, MR (2006). Antitumor action of curcumin in human papillomavirus

associated cells involves downregulation of viral oncogenes, prevention of NFkB and AP-1 translocation, and modulation of apoptosis. Molecular Carcinogenesis 45: 320-332.

Dixit, VK, Varma, KC (1975). Anthelmintic properties of essential oils from rhizomes of

Hedychium coronarium Koeing and Hedychium spicatum Koeing. Indian Journal of Pharmacy 37: 143-144.

Dixit, VK, Varma, KC, Vashisht, VN (1990). Studies on essential oils of Hedychium

spicatum Koenig and Hedychium coronarium Koenig. Indian Journal of Pharmacy 39: 58-60.

Dos Santos, AF, Sant'Ana, AE (2000). The molluscicidal activity of plants used in Brazilian

folk medicine. Phytomedicine 6: 431-438. Dugenci, SK, Arda, N, Candan, A (2003). Some medicinal plants as immunostimulant for

fish. Journal of Ethnopharmacology 88: 99-106.

Page 229: Phytochemistry and pharmacology of plants from the ginger family

205

Duncan, BB, Schmidt, MIs (2006). The epidemiology of low-grade chronic systemic

inflammation and type 2 diabetes. Diabetes Technology & Therapeutics 8: 7-17. Durgaprasad, S, Pai, CG, Vasanthkumar, Alvres, JF, Namitha, S (2005). A pilot study of the

antioxidant effect of curcumin in tropical pancreatitis. Indian Journal of Medical Research 122: 315-318.

Duvoix, A, Blasius, R, Delhalle, S, Schnekenburger, M, Morceau, F, Henry, E, Dicato, M,

Diederich, M (2005). Chemopreventive and therapeutic effects of curcumin. Cancer Letters 223: 181-190.

Eberhart, CE, Dubois, RN (1995). Eicosanoids and the gastrointestinal tract.

Gastroenterology 109: 285-301. Eberhart, LH, Mayer, R, Betz, O, Tsolakidis, S, Hilpert, W, Morin, AM, Geldner, G, Wulf,

H, Seeling, W (2003). Ginger does not prevent postoperative nausea and vomiting after laparoscopic surgery. Anesthesia & Analgesia. 96: 995-998, table of contents.

Eikelenboom, P, Veerhuis, R, Scheper, W, Rozemuller, AJM, van Gool, WA, Hoozemans,

JJM (2006). The significance of neuroinflammation in understanding Alzheimer's disease. Journal of Neural Transmission 113: 1685-1695.

Ekundayo, O, Laakso, I, Hiltunen, R (1988). Composition of ginger (Zingiber officinale

Roscoe) volatile oils from Nigeria. Flavour and Fragrance Journal 3: 85-90. El Tahir, KEH, Shoeb, H, Al-Shora, H (1997). Exploration of some pharmacological

activities of cardamom seed (Elettaria cardamomum) volatile oil. Saudi Pharmaceutical Journal 5: 96-102.

Evans, WC (2002). Trease and Evans Pharmacognosy. 15th edn. WB Saunders: Edinburgh. Eybl, V, Kotyzova, D, Bludovska, M (2004). The effect of curcumin on cadmium-induced

oxidative damage and trace elements level in the liver of rats and mice. Toxicology Letters 151: 79-85.

Eybl, V, Kotyzova, D, Koutensky, J (2006). Comparative study of natural antioxidants -

curcumin, resveratrol and melatonin - in cadmium-induced oxidative damage in mice. Toxicology 225: 150-156.

Farombi, EO, Ekor, M (2006). Curcumin attenuates gentamicin-induced renal oxidative

damage in rats. Food & Chemical Toxicology 44: 1443-1448. Fischer-Rasmussen, W, Kjaer, SK, Dahl, C, Asping, U (1991). Ginger treatment of

hyperemesis gravidarum. European Journal of Obstetrics, Gynecology, & Reproductive Biology 38: 19-24.

Fitzgerald, GA (2004). Coxibs and cardiovascular disease. N Engl J Med 351: 1709-1711.

Page 230: Phytochemistry and pharmacology of plants from the ginger family

206

Flynn, DL, Rafferty, MF, Boctor, AM (1986). Inhibition of human neutrophil 5-lipoxygenase activity by gingerdione shogaol capsaicin and related pungent compounds. Prostaglandins Leukotrienes & Medicine 24: 195-198.

Foster, RW (1996). Basic Pharmacology. 4th edn. Butterworth-Heinemann: Oxford. Foungbe, S, Sawadogo, D, Declume, C (1987). Experimental study of uterine-relaxant

activity of Alstonia boonei (Apocynaceae) and Costus lucanusianus (Zingiberaceae) traditionally used as anti-abortion drug in Ivory Coast. Annales Pharmaceutiques Francaises 45: 373-377.

Fuhrman, B, Rosenblat, M, Hayek, T, Coleman, R, Aviram, M (2000). Ginger extract

consumption reduces plasma cholesterol, inhibits LDL oxidation and attenuates development of atherosclerosis in atherosclerotic, apolipoprotein E-deficient mice. Journal of Nutrition. 130: 1124-1131.

Funk, JL, Frye, JB, Oyarzo, JN, Kuscuoglu, N, Wilson, J, McCaffrey, G, Stafford, G, Chen,

G, Lantz, RC, Jolad, SD, Solyom, AM, Kiela, PR, Timmermann, BN (2006). Efficacy and mechanism of action of turmeric supplements in the treatment of experimental arthritis. Arthritis & Rheumatism 54: 3452-3464.

Gaedeke, J, Noble, NA, Border, WA (2005). Curcumin blocks fibrosis in anti-Thy 1

glomerulonephritis through up-regulation of heme oxygenase 1. Kidney International 68: 2042-2049.

Gagnier, JJ, Boon, H, Rochon, P, Moher, D, Barnes, J, Bombardier, C, Group, C (2006).

Reporting randomized, controlled trials of herbal interventions: an elaborated CONSORT statement. Annals of Internal Medicine 144: 364-367.

Garcea, G, Berry, DP, Jones, DJL, Singh, R, Dennison, AR, Farmer, PB, Sharma, RA,

Steward, WP, Gescher, AJ (2005). Consumption of the putative chemopreventive agent curcumin by cancer patients: assessment of curcumin levels in the colorectum and their pharmacodynamic consequences. Cancer Epidemiology, Biomarkers & Prevention 14: 120-125.

Gautschi, M, Yang, X, Eilerman, RG, Frater, G (1999). Flavor chemicals with pungent

properties. In: Teranishi (ed). Flavor Chemistry: 30 Years of Progress. Kluwer Academic/Plenum Publishers: New York.

Ghayur, MN, Gilani, AH (2005). Ginger lowers blood pressure through blockade of voltage-

dependent calcium channels. Journal of Cardiovascular Pharmacology 45: 74-80. Ghoneim, AI, Abdel-Naim, AB, Khalifa, AE, El-Denshary, ES (2002). Protective effects of

curcumin against ischaemia/reperfusion insult in rat forebrain. Pharmacological Research 46: 273-279.

Ghosh, SK, Bhaskar, S, Satyabrata, G, Subhendu, G (2000). Screening of some angiospermic

plants for antimicrobial activity. Journal of Mycopathological Research 38: 19-22.

Page 231: Phytochemistry and pharmacology of plants from the ginger family

207

Gilbert, DL (1999). Reactive oxygen species in biological systems: an interdisciplinary approach. Kluwer Academic Publishers: Dordrecht.

Giri, J, Devi, TKS, Meerarani, S (1984). Effect of ginger on serum cholesterol levels. Indian

Journal of Nutrition & Dietetics 21: 433-436. Go, EJ, Kim, HR, Chung, SY, Jeong, YH, Kim, NH, Han, AR, Seo, EK, Lee, HJ (2004).

Evaluation on the P-glycoprotein inhibitory activity of Indonesian medicinal plants. Natural Product Sciences 10: 85-88.

Goel, A, Boland, CR, Chauhan, DP (2001). Specific inhibition of cyclooxygenase-2 (COX-

2) expression by dietary curcumin in HT-29 human colon cancer cells. Cancer Letters 172: 111-118.

Govindarajan, VS (1982a). Ginger - chemistry, technology, and quality evaluation. Part 1.

Critical Reviews in Food Science & Nutrition 17: 1-96. Govindarajan, VS (1982b). Ginger-chemistry, technology, and quality evaluation: Part 2.

Critical Reviews in Food Science & Nutrition 17: 189-258. Goyal, RK, Kadnur, SV (2006). Beneficial effects of Zingiber officinale on goldthioglucose

induced obesity. Fitoterapia 77: 160-163. Griggs, B (1997). New green pharmacy - the story of western herbal medicine. Vermillion:

London. Grontved, A, Hentzer, E (1986). Vertigo-reducing effect of ginger root. A controlled clinical

study. Orl; Journal of Oto-Rhino-Laryngology & its Related Specialties 48: 282-286. Grontved, A, Brask, T, Kambskard, J, Hentzer, E (1988). Ginger root against seasickness. A

controlled trial on the open sea. Acta Oto-Laryngologica 105: 45-49. Grzanna, R, Lindmark, L, Frondoza, CG (2005). Ginger - an herbal medicinal product with

broad anti-inflammatory actions. Journal of Medicinal Food 8: 125-132. Gu, JQ (2002). Activity-guided isolation of constituents of Renealmia nicolaioides with the

potential to induce the phase II enzyme quinone reductase. Journal of Natural Products 65.

Guh, JH, Ko, FN, Jong, TT, Teng, CM (1995). Antiplatelet effect of gingerol isolated from

Zingiber officinale. Journal of Pharmacy & Pharmacology 47: 329-332. Gukovsky, I, Reyes, CN, Vaquero, EC, Gukovskaya, AS, Pandol, SJ (2003). Curcumin

ameliorates ethanol and nonethanol experimental pancreatitis. American Journal of Physiology - Gastrointestinal & Liver Physiology 284: G85-95.

Gupta, MM, Farooqui, SU, Lal, RN (1981). Distribution and variation of diosgenin in

different parts of Costus speciosus. Journal of Natural Products 44: 486-489.

Page 232: Phytochemistry and pharmacology of plants from the ginger family

208

Gurib-Fakim, A, Gueho, J, Bissoondoyal, MD (1997). Plantes Medicinales de Maurice. Vol. 3. Vol. 3. Editions de l'Ocean Indien: Rose Hill, Mauritius.

Gurib-Fakim, A, Maudarbaccus, N, Leach, D, Doimo, L, Wohlmuth, H (2002). Essential oil

composition of Zingiberaceae species from Mauritius. Journal of Essential Oil Research 14: 271-273.

Habsah, M, Amran, M, Mackeen, MM, Lajis, NH, Kikuzaki, H, Nakatani, N, Rahman, AA,

Ghafar, Ali, AM (2000a). Screening of Zingiberaceae extracts for antimicrobial and antioxidant activities. Journal of Ethnopharmacology 72: 403-410.

Habsah, M, Amran, M, Mackeen, MM, Lajis, NH, Kikuzaki, H, Nakatani, N, Rahman, AA,

Ghafar, Ali, AM (2000b). Screening of Zingiberaceae extracts for antimicrobial and antioxidant activities. Journal of Ethnopharmacology 72: 403-410.

Hald, A, Van Beek, J, Lotharius, J (2007). Inflammation in Parkinson's disease: causative or

epiphenomenal? Sub-Cellular Biochemistry 42: 249-279. Halvorsen, BL, Holte, K, Myhrstad, MCW, Barikmo, I, Hvattum, E, Remberg, SF, Wold, A-

B, Haffner, K, Baugerod, H, Andersen, LF, Moskaug, O, Jacobs, DR, Jr., Blomhoff, R (2002). A systematic screening of total antioxidants in dietary plants. Journal of Nutrition 132: 461-471.

Han, A-R, Kim, M-S, Jeong, YH, Lee, SK, Seo, E-K (2005). Cyclooxygenase-2 inhibitory

phenylbutenoids from the rhizomes of Zingiber cassumunar. Chemical & Pharmaceutical Bulletin 53: 1466-1468.

Han, AR, Min, HY, Windono, T, Jeohn, GH, Jang, DS, Lee, SK, Seo, EK (2004). A new

cytotoxic phenylbutenoid dimer from the rhizomes of Zingiber cassumunar. Planta Medica 70: 1095-1097.

Hansson, GK (2005). Inflammation, atherosclerosis, and coronary artery disease. New

England Journal of Medicine 352: 1685-1695. Haraguchi, H, Kuwata, Y, Inada, K, Shingu, K, Miyahara, K, Nagao, M, Yagi, A (1996).

Antifungal activity from Alpinia galanga and the competition for incorporation of unsaturated fatty acids in cell growth. Planta Medica 62: 308-313.

Harshberger, JW (1896). The purposes of ethnobotany. Botanical Gazette 21: 146-154. Hata, Y, Pancho, LR, Nojima, H, Kimura, I (1998). Endothelium-dependent potentiation of

prostaglandin F2α-induced contractions by (±)-[6]-gingerol is inhibited by cyclooxygenase- but not lipoxygenase-inhibitors in mouse mesenteric veins. Biological & Pharmaceutical Bulletin 21: 792-794.

He, X-G, Bernart, MW, Lian, L-S, Lin, L-Z (1998). High-performance liquid

chromatography-electrospray mass spectrometric analysis of pungent constituents of ginger. Journal of Chromatography 796: 327-334.

Page 233: Phytochemistry and pharmacology of plants from the ginger family

209

Heinrich, M, Gibbons, S (2001). Ethnopharmacology in drug discovery: an analysis of its role and potential contribution. Journal of Pharmacy and Pharmacology 53: 425-432.

Heinrich, M, Barnes, J, Gibbons, S, Williamson, EM (2004). Fundamental of

Pharmacognosy and Phytotherapy. Churchill Livingstone: Edinburgh. Hersh, EV, Lally, ET, Moore, PA (2005). Update on cyclooxygenase inhibitors: has a third

COX isoform entered the fray? Current Medical Research And Opinion 21: 1217-1226. Higdon, J (2007). An evidence-based approach to dietary phytochemicals. Thieme: New

York. Hikino, H, Takahashi, H, Sakurai, Y, Takemoto, T, Bhacca, NS (1966). Structure of

zederone. Chemical and Pharmaceutical Bulletin 14: 550-551. Hikino, H, Konno, C, Agatsuma, K, Takemoto, T, Horibe, I, Tori, K, Ueyama, M, Takeda, K

(1975). Sesquiterpenoids. Part XLVII. Structure, configuration, conformation, and thermal rearrangement of furanodienone, isofuranodienone, curzerenone, epicurzerenone, and pyrocurzerenone, sesquiterpenoids of Curcuma zedoaria. Journal of the Chemical Society, Perkins Transactions 1: 478-484.

Hikino, H (1985). Recent research on Oriental medicinal plants. In: Wagner, H, Hikino, H,

Farnsworth, NR (eds). Economic and medicinal plant research Vol. 1. Academic Press: London.

Hiserodt, RD, Franzblau, SG, Rosen, RT (1998). Isolation of 6-, 8-, and 10-gingerol from

ginger rhizome by HPLC and preliminary evaluation of inhibition of Mycobacterium avium and Mycobacterium tuberculosis. Journal of Agricultural & Food Chemistry 46: 2504-2508.

Hnatiuk, RJ (1990). Census of Australian Vascular Plants. Australian Government

Publishing Service: Canberra. Holt, PR, Katz, S, Kirshoff, R (2005). Curcumin therapy in inflammatory bowel disease: a

pilot study. Digestive Diseases & Sciences 50: 2191-2193. Holtmann, S, Clarke, AH, Scherer, H, Hohn, M (1989). The anti-motion sickness mechanism

of ginger. A comparative study with placebo and dimenhydrinate. Acta Oto-Laryngologica 108: 168-174.

Hong, CH, Hur, SK, Oh, OJ, Kim, SS, Nam, KA, Lee, SK (2002). Evaluation of natural

products on inhibition of inducible cyclooxygenase (COX-2) and nitric oxide synthase (iNOS) in cultured mouse macrophage cells. Journal of Ethnopharmacology 83: 153-159.

Hong, J, Bose, M, Ju, J, Ryu, J-H, Chen, X, Sang, S, Lee, M-J, Yang, CS (2004). Modulation

of arachidonic acid metabolism by curcumin and related beta-diketone derivatives: effects on cytosolic phospholipase A(2), cyclooxygenases and 5-lipoxygenase. Carcinogenesis 25: 1671-1679.

Page 234: Phytochemistry and pharmacology of plants from the ginger family

210

Hou, Y-C, Hsieh, Y-S, Chen, C-C, Chao, P-D (1997). Reverse transcriptase inhibitors from Curcuma phaeocaulis. Chinese Pharmaceutical Journal 49: 119-125.

Houghton, PJ (2002). Traditional plant medicines as a source of new drugs. In: Evans, WC

(ed). Trease and Evans Pharmacognosy. 15th edn. W. B. Saunders: Edinburgh. Huang, MT, Lysz, T, Ferraro, T, Abidi, TF, Laskin, JD, Conney, AH (1991). Inhibitory

effects of curcumin on in vitro lipoxygenase and cyclooxygenase activities in mouse epidermis. Cancer Research 51: 813-819.

Huang, Y-B, Hsu, L-R, Wu, P-C, Ko, H-M, Tsai, Y-H (1993). Crude drug (Zingiberaceae)

enhancement of percutaneous absorption of indomethacin: In vitro and in vivo permeation. Kao-Hsiung i Hsueh Ko Hsueh Tsa Chih [Kaohsiung Journal of Medical Sciences] 9: 392-400.

Humphrey, AM, Harris, JR, Mansfield, CA, Michalkiewicz, DM, Milchard, M, Moyler, DA,

Osbiston, A, Smith, S, Starr, B, Stevens, TM, Wilson, JJ (1993). Application of gas liquid chromatography to the analysis of essential oils. 16. Monographs for 5 essential oils. Analyst 118: 1089-1098.

Inoue, K, Kobayashi, S, Noguchi, H, Sankawa, U, Ebizuka, Y (1995). Spirostanol and

furostanol glycosides of Costus speciosus (Koenig.) SM. Natural Medicines 49: 336-339. Ippoushi, K, Azuma, K, Ito, H, Horie, H, Higashio, H (2003). [6]-Gingerol inhibits nitric

oxide synthesis in activated J774.1 mouse macrophages and prevents peroxynitrite-induced oxidation and nitration reactions. Life Sciences 73: 3427-3437.

Iqbal, M, Sharma, SD, Okazaki, Y, Fujisawa, M, Okada, S (2003). Dietary supplementation

of curcumin enhances antioxidant and phase II metabolizing enzymes in ddY male mice: possible role in protection against chemical carcinogenesis and toxicity. Pharmacology & Toxicology 92: 33-38.

Ireson, C, Orr, S, Jones, DJ, Verschoyle, R, Lim, CK, Luo, JL, Howells, L, Plummer, S,

Jukes, R, Williams, M, Steward, WP, Gescher, A (2001). Characterization of metabolites of the chemopreventive agent curcumin in human and rat hepatocytes and in the rat in vivo, and evaluation of their ability to inhibit phorbol ester-induced prostaglandin E2 production. Cancer Research 61: 1058-1064.

Ito, C, Itoigawa, M, Furukawa, A, Hirano, T, Murata, T, Kaneda, N, Hisada, Y, Okuda, K,

Furukawa, H (2004). Quinolone alkaloids with nitric oxide production inhibitory activity from Orixa japonica. Journal of Natural Products 67: 1800-1803.

Itokawa, H, Morita, H, Sumitomo, T (1987). Antitumour principles from Alpinia galanga.

Planta Medica 53: 32-33. Iwu, MM (2002). Ethnobotanical approach to pharmaceutical drug discovery: strengths and

limitations. In: Iwu, MM, Wootton, JC (eds). Ethnomedicine and Drug Discovery. Elsevier: Amsterdam. pp 309-320.

Page 235: Phytochemistry and pharmacology of plants from the ginger family

211

Jagetia, G, Baliga, M, Venkatesh, P (2004). Ginger (Zingiber officinale Rosc.), a dietary supplement, protects mice against radiation-induced lethality: Mechanism of action. Cancer Biotherapy and Radiopharmaceuticals 19: 422-435.

Jagetia, GC, Baliga, MS, Venkatesh, P, Ulloor, JN (2003). Influence of ginger rhizome

(Zingiber officinale Rosc) on survival, glutathione and lipid peroxidation in mice after whole-body exposure to gamma radiation. Radiation Research 160: 584.

Jagetia, GC, Rajanikant, GK (2004). Role of curcumin, a naturally occurring phenolic

compound of turmeric in accelerating the repair of excision wound, in mice whole-body exposed to various doses of gamma-radiation. Journal of Surgical Research 120: 127-138.

Jala, VR, Haribabu, B (2004). Leukotrienes and atherosclerosis: new roles for old mediators.

Trends in Immunology 25: 315-322. Jang, MK, Lee, HJ, Kim, JS, Ryu, J-H (2004). A curcuminoid and two sesquiterpenoids from

Curcuma zedoaria as inhibitors of nitric oxide synthesis in activated macrophages. Archives of Pharmacal Research 27: 1220-1225.

Jang, S-I, Jeong, S-I, Kim, K-J, Kim, H-J, Yu, H-H, Park, R, Kim, H-M, You, Y-O (2003).

Tanshinone IIA from Salvia miltiorrhiza inhibits inducible nitric oxide synthase expression and production of TNF-alpha, IL-1beta and IL-6 in activated RAW 264.7 cells. Planta Medica 69: 1057-1059.

Jantaratnotai, N, Utaisincharoen, P, Piyachaturawat, P, Chongthammakun, S, Sanvarinda, Y

(2006). Inhibitory effect of Curcuma comosa on NO production and cytokine expression in LPS-activated microglia. Life Sciences 78: 571-577.

Jeenapongsa, R, Yoovathaworn, K, Sriwatanakul, KM, Pongprayoon, U, Sriwatanakul, K

(2003). Anti-inflammatory activity of (E)-1-(3,4-dimethoxyphenyl) butadiene from Zingiber cassumunar Roxb. Journal of Ethnopharmacology 87: 143-148.

Jeyakumar, SM, Nalini, N, Menon, VP (1999). Antioxidant activity of ginger (Zingiber

officinale Rosc) in rats fed a high fat diet. Medical Science Research 27: 341-344. Jian, Y-T, Mai, G-F, Wang, J-D, Zhang, Y-L, Luo, R-C, Fang, Y-X (2005). Preventive and

therapeutic effects of NF-kappaB inhibitor curcumin in rats colitis induced by trinitrobenzene sulfonic acid. World Journal of Gastroenterology 11: 1747-1752.

Jiang, H, Deng, C-S, Zhang, M, Xia, J (2006a). Curcumin-attenuated trinitrobenzene

sulphonic acid induces chronic colitis by inhibiting expression of cyclooxygenase-2. World Journal of Gastroenterology 12: 3848-3853.

Jiang, H, Xie, Z, Koo, HJ, McLaughlin, SP, Timmermann, BN, Gang, DR (2006b).

Metabolic profiling and phylogenetic analysis of medicinal Zingiber species: Tools for authentication of ginger (Zingiber officinale Rosc.). Phytochemistry 67: 1673-1685.

Page 236: Phytochemistry and pharmacology of plants from the ginger family

212

Jitoe, A, Masuda, T, Tengah, IGP, Suprapta, DN, Gara, IW, Nakatani, N (1992). Antioxidant activity of tropical ginger extracts and analysis of the contained curcuminoids. Journal of Agricultural & Food Chemistry 40: 1337-1340.

Jitoe, A, Masuda, T, Nakatani, N (1993). Phenylbutenoid dimers from the rhizomes of

Zingiber cassumunar. Phytochemistry 32: 357-363. Jobin, C, Bradham, CA, Russo, MP, Juma, B, Narula, AS, Brenner, DA, Sartor, RB (1999).

Curcumin blocks cytokine-mediated NF-kappa B activation and proinflammatory gene expression by inhibiting inhibitory factor I-kappa B kinase activity. Journal of Immunology 163: 3474-3483.

Joe, B, Vijaykumar, M, Lokesh, BR (2004). Biological properties of curcumin-cellular and

molecular mechanisms of action. Critical Reviews in Food Science & Nutrition 44: 97-111.

Johnson, T (1999). CRC Ethnobotany Desk Reference. CRC Press: Boca Raton. Jolad, SD, Lantz, RC, Solyom, AM, Chen, GJ, Bates, RB, Timmermann, BN (2004). Fresh

organically grown ginger (Zingiber officinale): composition and effects on LPS-induced PGE2 production. Phytochemistry 65: 1937.

Jolad, SD, Lantz, RC, Chen, GJ, Bates, RB, Timmermann, BN (2005). Commercially

processed dry ginger (Zingiber officinale): composition and effects on LPS-stimulated PGE2 production. Phytochemistry 66: 1614-1635.

Jovanovic, SV, Boone, CW, Steenken, S, Trinoga, M, Kaskey, RB (2001). How curcumin

works preferentially with water soluble antioxidants. Journal of the American Chemical Society 123: 3064-3068.

Jung, EM, Lim, JH, Lee, TJ, Park, J-W, Choi, KS, Kwon, TK (2005). Curcumin sensitizes

tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis through reactive oxygen species-mediated upregulation of death receptor 5 (DR5). Carcinogenesis 26: 1905-1913.

Jung, KH, Ha, E, Kim, MJ, Won, H-J, Zheng, LT, Kim, HK, Hong, SJ, Chung, JH, Yim, S-V

(2007). Suppressive effects of nitric oxide (NO) production and inducible nitric oxide synthase (iNOS) expression by Citrus reticulata extract in RAW 264.7 macrophage cells. Food And Chemical Toxicology: An International Journal Published For The British Industrial Biological Research Association 45: 1545-1550.

Jung, KK, Lee, HS, Cho, JY, Shin, WC, Rhee, MH, Kim, TG, Kang, JH, Kim, SH, Hong, S,

Kang, SY (2006). Inhibitory effect of curcumin on nitric oxide production from lipopolysaccharide-activated primary microglia. Life Sciences 79: 2022-2031.

Kadnur, SV, Goyal, RK (2005). Beneficial effects of Zingiber officinale Roscoe on fructose

induced hyperlipidemia and hyperinsulinemia in rats. Indian Journal of Experimental Biology 43: 1161-1164.

Page 237: Phytochemistry and pharmacology of plants from the ginger family

213

Kadota, S, Prasain, JK, Li, JX, Basnet, P, Dong, H, Tani, T, Namba, T (1996). Blepharocalyxins A and B, novel diarylheptanoids from Alpinia blepharocalyx, and their inhibitory effect on NO formation in murine macrophages. Tetrahedron Letters 37: 7283-7286.

Kalpana, C, Menon, VP (2004). Inhibition of nicotine-induced toxicity by curcumin and

curcumin analog: a comparative study. Journal of Medicinal Food 7: 467-471. Kang, BY, Song, YJ, Kim, KM, Choe, YK, Hwang, SY, Kim, TS (1999). Curcumin inhibits

Th1 cytokine profile in CD4+ T cells by suppressing interleukin-12 production in macrophages. British Journal of Pharmacology 128: 380-384.

Kang, G, Kong, P-J, Yuh, Y-J, Lim, S-Y, Yim, S-V, Chun, W, Kim, S-S (2004). Curcumin

suppresses lipopolysaccharide-induced cyclooxygenase-2 expression by inhibiting activator protein 1 and nuclear factor kappab bindings in BV2 microglial cells. Journal of Pharmacological Sciences 94: 325-328.

Kang, H-C, Nan, J-X, Park, P-H, Kim, J-Y, Lee, SH, Woo, SW, Zhao, Y-Z, Park, E-J, Sohn,

DH (2002). Curcumin inhibits collagen synthesis and hepatic stellate cell activation in-vivo and in-vitro. Journal of Pharmacy & Pharmacology 54: 119-126.

Kaperonis, EA, Liapis, CD, Kakisis, JD, Dimitroulis, D, Papavassiliou, VG (2006).

Inflammation and atherosclerosis. European Journal Of Vascular And Endovascular Surgery 31: 386-393.

Kaphai, BK, Kapur, SK, Sarin, YK (1977). Studies on Costus speciosus Sims. Part II:

Pharmacognostic characters of rhizome drug. Indian Journal of Pharmacy 39: 74-76. Kaplan, MA, Pugialli, HR, Lopes, D, Gottlieb, HE (2000). The stereochemistry of ledol from

Renealmia chrysotrycha: an NMR study. Phytochemistry 55: 749-753. Karunagaran, D, Rashmi, R, Kumar, TR (2005). Induction of apoptosis by curcumin and its

implications for cancer therapy. Current Cancer Drug Targets 5: 117-129. Katiyar, SK, Agarwal, R, Mukhtar, H (1996). Inhibition of tumor promotion in SENCAR

mouse skin by ethanol extract of Zingiber officinale rhizome. Cancer Research 56: 1023-1030.

Kaur, C, Kapoor, HC (2002). Anti-oxidant activity and total phenolic content of some Asian

vegetables. International Journal of Food Science and Technology 37: 153-161. Kawabata, J, Y., F, Tahara, S, Mizutani, J (1985). Structures of novel sesquiterpene ketones

from Chloranthus serratus (Chloranthaceae). Agricultural and Biological Chemistry 49: 1479-1485.

Kawamori, T, Lubet, R, Steele, VE, Kelloff, GJ, Kaskey, RB, Rao, CV, Reddy, BS (1999).

Chemopreventive effect of curcumin, a naturally occurring anti- inflammatory agent, during the promotion/progression stages of colon cancer. Cancer Research 59: 597-601.

Page 238: Phytochemistry and pharmacology of plants from the ginger family

214

Keating, A, Chez, RA (2002). Ginger syrup as an antiemetic in early pregnancy. Alternative Therapies in Health & Medicine. 8: 89-91.

Kelm, MA, Nair, MG, Strasburg, GM, DeWitt, DL (2000). Antioxidant and cyclooxygenase

inhibitory phenolic compounds from Ocimum sanctum Linn. Phytomedicine 7: 7-13. Kempaiah, RK, Srinivasan, K (2004). Influence of dietary curcumin, capsaicin and garlic on

the antioxidant status of red blood cells and the liver in high-fat-fed rats. Annals of Nutrition & Metabolism 48: 314-320.

Kikuzaki, H, Nakatani, N (1993). Antioxidant effects of some ginger constituents. Journal of

Food Science 58: 1407-1410. Kikuzaki, H (2000). Ginger for drug and spice purposes. In: Mazza, G, Oomah, BD (eds).

Herbs, Botanicals and Teas. Technomic Publishing Co.: Lancaster. pp 75-105. Kim, HW, Murakami, A, Abe, M, Ozawa, Y, Morimitsu, Y, Williams, MV, Ohigashi, H

(2005a). Suppressive effects of mioga ginger and ginger constituents on reactive oxygen and nitrogen species generation, and the expression of inducible pro-inflammatory genes in macrophages. Antioxidants & Redox Signaling 7: 1621-1629.

Kim, HY, Park, EJ, Joe, E-H, Jou, I (2003). Curcumin suppresses Janus kinase-STAT

inflammatory signaling through activation of Src homology 2 domain-containing tyrosine phosphatase 2 in brain microglia. Journal of Immunology 171: 6072-6079.

Kim, SO, Kundu, JK, Shin, YK, Park, JH, Cho, MH, Kim, TY, Surh, YJ (2005b). [6]-

Gingerol inhibits COX-2 expression by blocking the activation of p38 MAP kinase and NF-ΚB in phorbol ester-stimulated mouse skin. Oncogene 24: 2558-2567.

Kimura, I, Kimura, M, Pancho, LR (1989). Modulation of eicosanoid-induced contraction of

mouse and rat blood vessels by gingerols. Japanese Journal of Pharmacology 50: 253-261.

Kirana, C, Record, IR, McIntosh, GH, Jones, GP (2003). Screening for antitumor activity of

11 species of Indonesian Zingiberaceae using human MCF-7 and HT-29 cancer cells. Pharmaceutical Biology 41: 271-276.

Kishore, N, Dwivedi, RS (1992). Zerumbone: a potential fungitoxic agent isolated from

Zingiber cassumunar Roxb. Mycopathologia 120: 155-159. Kiuchi, F, Shibuya, M, Sankawa, U (1982). Inhibitors of prostaglandin biosynthesis from

ginger. Chemical & Pharmaceutical Bulletin 30: 754-757. Kiuchi, F, Iwakami, S, Shibuya, M, Hanaoka, F, Sankawa, U (1992). Inhibition of

prostaglandin and leukotriene biosynthesis by gingerols and diarylheptanoids. Chemical & Pharmaceutical Bulletin 40: 387-391.

Kobayashi, M, Shoji, N, Ohizumi, Y (1987). Gingerol, a novel cardiotonic agent, activates

the Ca2+-pumping ATPase in skeletal and cardiac sarcoplasmic reticulum. Biochimica et Biophysica Acta 903: 96-102.

Page 239: Phytochemistry and pharmacology of plants from the ginger family

215

Kobayashi, T, Hashimoto, S, Horie, T (1997). Curcumin inhibition of Dermatophagoides

farinea-induced interleukin-5 (IL-5) and granulocyte macrophage-colony stimulating factor (GM-CSF) production by lymphocytes from bronchial asthmatics. Biochemical Pharmacology 54: 819-824.

Koo, KL, Ammit, AJ, Tran, VH, Duke, CC, Roufogalis, BD (2001). Gingerols and related

analogues inhibit arachidonic acid-induced human platelet serotonin release and aggregation. Thrombosis Research. 103: 387-397.

Kosuge, T, Yokota, M, Sugiyama, K (1985). Studies on anticancer principles in Chinese

medicines. II. Cytotoxic principles in Biota orientalis (L.) Endl. and Kaempferia galanga L. Chemical & Pharmaceutical Bulletin 33: 5565-5567.

Kubota, K, Murayama, KN, Kobayashi, A, Amaike, M (1998). Acetoxy-1,8-cineoles as

aroma constituents of Alpinia galanga Willd. Journal of Agricultural & Food Chemistry 46: 5244-5247.

Kubota, K, Someya, Y, Yoshida, R, Kobayashi, A, Morita, T, Koshino, H (1999).

Enantiomeric purity and odor characteristics of 2- and 3-acetoxy-1,8-cineoles in the rhizomes of Alpinia galanga Willd. Journal of Agricultural & Food Chemistry 47: 685-689.

Kulmatycki, KM, Jamali, F (2006). Drug disease interactions: role of inflammatory

mediators in depression and variability in antidepressant drug response. Journal of Pharmacy & Pharmaceutical Sciences 9: 292-306.

Kumar, A, Dhawan, S, Hardegen, NJ, Aggarwal, BB (1998). Curcumin (diferuloylmethane)

inhibition of tumor necrosis factor (TNF)-mediated adhesion of monocytes to endothelial cells by suppression of cell surface expression of adhesion molecules and of nuclear factor-kappaB activation. Biochemical Pharmacology 55: 775-783.

Kumar, PA, Suryanarayana, P, Reddy, PY, Reddy, GB (2005). Modulation of alpha-

crystallin chaperone activity in diabetic rat lens by curcumin. Molecular Vision 11: 561-568.

Kuo, P-C, Damu, AG, Cherng, C-Y, Jeng, J-F, Teng, C-M, Lee, EJ, Wu, T-S (2005).

Isolation of a natural antioxidant, dehydrozingerone from Zingiber officinale and synthesis of its analogues for recognition of effective antioxidant and antityrosinase agents. Archives of Pharmacal Research 28: 518-528.

Kuster, RM, Mpalantinos, MA, De Holanda, MC, Lima, P, Brand, ET, Parente, JP (1999).

GC-MS determination of kava-pyrones in Alpinia zerumbet leaves. HRC-Journal of High Resolution Chromatography 22: 129-130.

Kuttan, R, Sudheeran, PC, Josph, CD (1987). Turmeric and curcumin as topical agents in

cancer therapy. Tumori 73: 29-31. Lans, C, Harper, T, Georges, K, Bridgewater, E (2001). Medicinal and ethnoveterinary

remedies of hunters in Trinidad. BMC Complementary & Alternative Medicine 1: 10.

Page 240: Phytochemistry and pharmacology of plants from the ginger family

216

Lantz, RC, Chen, GJ, Solyom, AM, Jolad, SD, Timmermann, BN (2005). The effect of

turmeric extracts on inflammatory mediator production. Phytomedicine 12: 445-452. Lantz, RC, Chen, GJ, Sarihan, M, Solyom, AM, Jolad, SD, Timmermann, BN (2007). The

effect of extracts from ginger rhizome on inflammatory mediator production. Phytomedicine 14: 128-128.

Lapworth, A, Pearson, LK, Royle, RA (1917). The pungent principles of ginger. Part 1. The

chemical character and decomposition products of Tresh's "gingerol". Journal of the Chemical Society: 777.

Lawrence, BM (1995a). Progress in Essential oils: Ginger oil. Perfumer and Flavorist 20:

54-56. Lawrence, BM (1995b). Progress in essential oils. Perfumer & Flavorist 20: 49-59. Lawrence, BM (1997). Progress in essential oils. Perfumer & Flavorist 22: 71-83. Lawrence, BM (1998). Progress in essential oils. Perfumer & Flavorist 23: 47-57. Lawrence, BM (2000). Progress in essential oils. Perfumer & Flavorist 25: 46-57. Lawrence, BM (2002). Progress in essential oils. Perfumer & Flavorist 27: 46-64. Lechat-Vahirua, I, François, P, Menut, C, Lamaty, G, Bessiere, JM (1993). Aromatic plants

of French Polynesia. I. Constituents of the essential oils of rhizomes of three Zingiberaceae: Zingiber zerumbet Smith, Hedychium coronarium Koenig and Etlingera cevuga Smith. Journal of Essential Oil Research 5: 55-59.

Lechat-Vahirua, I, Menut, C, Lamaty, G, Bessiere, JM (1996). Huiles essentielles de

Polynesie Francais. Rivista Italiana EPPOS: 627-638. Leclercq, IA, Farrell, GC, Sempoux, C, dela Pena, A, Horsmans, Y (2004). Curcumin

inhibits NF-kappaB activation and reduces the severity of experimental steatohepatitis in mice. Journal of Hepatology 41: 926-934.

Lee, E, Surh, YJ (1998). Induction of apoptosis in HL-60 cells by pungent vanilloids, [6]-

gingerol and [6]-paradol. Cancer Letters 134: 163-168. Lee, HJ, Kim, JS, Ryu, J-H (2006). Suppression of inducible nitric oxide synthase expression

by diarylheptanoids from Alpinia officinarum. Planta Medica 72: 68-71. Lee, J-W, Min, H-Y, Han, A-R, Chung, H-J, Park, E-J, Park, HJ, Hong, J-Y, Seo, E-K, Lee,

SK (2007). Growth inhibition and induction of G1 phase cell cycle arrest in human lung cancer cells by a phenylbutenoid dimer isolated from Zingiber cassumunar. Biological & Pharmaceutical Bulletin 30: 1561-1564.

Lee, J, Im, Y-H, Jung, HH, Kim, JH, Park, JO, Kim, K, Kim, WS, Ahn, JS, Jung, CW, Park,

YS, Kang, WK, Park, K (2005a). Curcumin inhibits interferon-alpha induced NF-kappaB

Page 241: Phytochemistry and pharmacology of plants from the ginger family

217

and COX-2 in human A549 non-small cell lung cancer cells. Biochemical & Biophysical Research Communications 334: 313-318.

Lee, KW, Kim, J-H, Lee, HJ, Surh, Y-J (2005b). Curcumin inhibits phorbol ester-induced

up-regulation of cyclooxygenase-2 and matrix metalloproteinase-9 by blocking ERK1/2 phosphorylation and NF-kappaB transcriptional activity in MCF10A human breast epithelial cells. Antioxidants & Redox Signaling 7: 1612-1620.

Lee, SK, Hong, CH, Huh, SK, Kim, SS, Oh, OJ, Min, HY, Park, KK, Chung, WY, Hwang,

JK (2002). Suppressive effect of natural sesquiterpenoids on inducible cyclooxygenase (COX-2) and nitric oxide synthase (iNOS) activity in mouse macrophage cells. Journal of Environmental Pathology, Toxicology & Oncology 21: 141-148.

Leu, T-H, Maa, M-C (2002). The molecular mechanisms for the antitumorigenic effect of

curcumin. Current Medicinal Chemistry - Anti-Cancer Agents 2: 357-370. Lev-Ari, S, Maimon, Y, Strier, L, Kazanov, D, Arber, N (2006). Down-regulation of

prostaglandin E2 by curcumin is correlated with inhibition of cell growth and induction of apoptosis in human colon carcinoma cell lines. Journal of the Society for Integrative Oncology 4: 21-26.

Leverington, RE (1975). Ginger technology. Food Technology in Australia 27: 309-313. Li, L, Braiteh, FS, Kurzrock, R (2005). Liposome-encapsulated curcumin: in vitro and in

vivo effects on proliferation, apoptosis, signaling, and angiogenesis. Cancer 104: 1322-1331.

Libby, P (2007). Inflammatory mechanisms: the molecular basis of inflammation and

disease. Nutrition Reviews 65: S140-146. Lien, HC, Sun, WM, Chen, YH, Kim, H, Hasler, W, Owyang, C (2003). Effects of ginger on

motion sickness and gastric slow-wave dysrhythmias induced by circular vection. American Journal of Physiology - Gastrointestinal & Liver Physiology. 284: G481-489.

Lim, GP, Chu, T, Yang, F, Beech, W, Frautschy, SA, Cole, GM (2001). The curry spice

curcumin reduces oxidative damage and amyloid pathology in an Alzheimer transgenic mouse. Journal of Neuroscience 21: 8370-8377.

Lin, J-K (2004). Suppression of protein kinase C and nuclear oncogene expression as

possible action mechanisms of cancer chemoprevention by curcumin. Archives of Pharmacal Research 27: 683-692.

Lipsky, PE (1999). Specific COX-2 inhibitors in arthritis, oncology, and beyond: where is

the science headed? Journal of Rheumatology 26 Suppl 56: 25-30. Liu, H, Danthi, SJ, Enyeart, JJ (2006). Curcumin potently blocks Kv1.4 potassium channels.

Biochemical & Biophysical Research Communications 344: 1161-1165.

Page 242: Phytochemistry and pharmacology of plants from the ginger family

218

Liu, JH, Zschocke, S, Reininger, E, Bauer, R (1998). Inhibitory effects of Angelica pubescens f. biserrata on 5-lipoxygenase and cyclooxygenase. Planta Medica 64: 525-529.

Lu, Y, Sun, C, Wang, Y, Pan, Y (2005). Preparative isolation and purification of two

phenylbutenoids from the rhizomes of Zingiber cassumunar by upright counter-current chromatography. Journal Of Chromatography. A 1089: 258-262.

Luger, P, Weber, M, Dung, NX, Tuyet, NTB (1996). Ethyl p-methoxycinnamate from

Kaempferia galanga L in Vietnam. Acta Crystallographica - Section C - Crystal Structure Communications 52: 1255-1257.

Mabberley, DJ (1997). The plant-book: a portable dictionary of the vascular plants. 2nd edn.

Cambridge University Press: Cambridge. Macleod, AJ, Pieris, NM (1984). Volatile aroma constituents of Sri Lankan ginger.

Phytochemistry 23: 353-360. Mahady, GB, Pendland, SL, Yun, GS, Lu, Z-Z, Stoia, A (2003). Ginger (Zingiber officinale

Roscoe) and the gingerols inhibit the growth of CagA+ strains of Helicobacter pylori. Anticancer Research 23: 3699.

Mahesh, T, Balasubashini, MS, Menon, VP (2005). Effect of photo-irradiated curcumin

treatment against oxidative stress in streptozotocin-induced diabetic rats. Journal of Medicinal Food 8: 251-255.

Maheshwari, RK, Singh, AK, Gaddipati, J, Srimal, RC (2006). Multiple biological activities

of curcumin: a short review. Life Sciences 78: 2081-2087. Majithiya, JB, Balaraman, R (2005). Time-dependent changes in antioxidant enzymes and

vascular reactivity of aorta in streptozotocin-induced diabetic rats treated with curcumin. Journal of Cardiovascular Pharmacology 46: 697-705.

Makabe, H, Maru, N, Kuwabara, A, Kamo, T, Hirota, M (2006). Anti-inflammatory

sesquiterpenes from Curcuma zedoaria. Natural Product Research 20: 680-685. Mall, M, Kunzelmann, K (2005). Correction of the CF defect by curcumin: hypes and

disappointments. Bioessays 27: 9-13. Mann, J (1992). Murder, magic and medicine. Oxford University Press: Oxford. Manusirivithaya, S, Sripramote, M, Tangjitgamol, S, Sheanakul, C, Leelahakorn, S,

Thavaramara, T, Tangcharoenpanich, K (2004). Antiemetic effect of ginger in gynecologic oncology patients receiving cisplatin. International Journal of Gynecological Cancer 14: 1063-1069.

Marnett, LJ, Rowlinson, SW, Goodwin, DC, Kalgutkar, AS, Lanzo, CA (1999). Arachidonic

acid oxygenation by COX-1 and COX-2. Mechanisms of catalysis and inhibition. J Biol Chem 274: 22903-22906.

Page 243: Phytochemistry and pharmacology of plants from the ginger family

219

Martin, GJ (1995). Ethnobotany: a methods manual. Chapman & Hall: London. Martins, AP, Salgueiro, L, Goncalves, MJ, da Cunha, AP, Vila, R, Canigueral, S, Mazzoni,

V, Tomi, F, Casanova, J (2001). Essential oil composition and antimicrobial activity of three Zingiberaceae from S.Tome e Principe. Planta Medica 67: 580.

Mascolo, N, Jain, R, Jain, SC, Capasso, F (1989). Ethnopharmacologic investigation of

ginger (Zingiber officinale). Journal of Ethnopharmacology 27: 129-140. Masuda, T, Jitoe, A (1994). Antioxidative and antiinflammatory compounds from tropical

gingers - isolation, structure determination, and activities of cassumunins A, B and C, new complex curcuminoids from Zingiber cassumunar. Journal of Agricultural & Food Chemistry 42: 1850-1856.

Masuda, T, Jitoe, A (1995). Phenylbutenoid monomers from the rhizomes of Zingiber

cassumunar. Phytochemistry 39: 459-461. Masuda, Y, Kikuzaki, H, Hisamoto, M, Nakatani, N (2004). Antioxidant properties of

gingerol related compounds from ginger. Biofactors 21: 293-296. Matsuda, H, Shimoda, H, Yoshikawa, M (1999). Structure-requirements of isocoumarins,

phthalides, and stilbenes from Hydrangeae Dulcis Folium for inhibitory activity on histamine release from rat peritoneal mast cells. Bioorganic & Medicinal Chemistry 7: 1445-1450.

Matsuda, H, Ando, S, Morikawa, T, Kataoka, S, Yoshikawa, M (2005). Structure-activity

relationships of 1'S-1'-acetoxychavicol acetate for inhibitory effect on NO production in lipopolysaccharide-activated mouse peritoneal macrophages. Bioorganic & Medicinal Chemistry Letters 15: 1949-1953.

Matsuda, H, Ando, S, Kato, T, Morikawa, T, Yoshikawa, M (2006). Inhibitors from the

rhizomes of Alpinia officinarum on production of nitric oxide in lipopolysaccharide-activated macrophages and the structural requirements of diarylheptanoids for the activity. Bioorganic & Medicinal Chemistry 14: 138-142.

McCance, K, Huether, S (2002). Pathophysiology. 4th edn. Mosby: St. Louis. Menut, C, Lamaty, G, Bessiere, JM, Koudou, J (1994). Aromatic plants of tropical central

Africa. XIII. Rhizomes volatile components of two Zingiberales from the Central African Republic. Journal of Essential Oil Research 6: 161-164.

Mesa, MD, Aguilera, CM, Ramirez-Tortosa, CL, Ramirez-Tortosa, MC, Quiles, JL, Baro, L,

Martinez de Victoria, E, Gil, A (2003). Oral administration of a turmeric extract inhibits erythrocyte and liver microsome membrane oxidation in rabbits fed with an atherogenic diet. Nutrition 19: 800-804.

Meyer, K, Schwartz, J, Crater, D, Keyes, B (1995). Zingiber officinale (ginger) used to

prevent 8-MOP associated nausea. Dermatology Nursing 7: 242-244.

Page 244: Phytochemistry and pharmacology of plants from the ginger family

220

Micklefield, GH, Redeker, Y, Meister, V, Jung, O, Greving, I, May, B (1999). Effects of ginger on gastroduodenal motility. International Journal of Clinical Pharmacology & Therapeutics 37: 341-346.

Miles, S (1980). The ginger connection. Australian Geographic 2: 69-71. Mills, S, Bone, K (2000). Principles and Practice of Phytotherapy. Churchill Livingstone:

Edinburgh. Mishra, B, Priyadarsini, KI, Bhide, MK, Kadam, RM, Mohan, H (2004). Reactions of

superoxide radicals with curcumin: probable mechanisms by optical spectroscopy and EPR. Free Radical Research 38: 355-362.

Moeslinger, T, Friedl, R, Volf, I, Brunner, M, Koller, E, Spieckermann, PG (2000).

Inhibition of inducible nitric oxide synthesis by the herbal preparation Padma 28 in macrophage cell line. Canadian Journal of Physiology & Pharmacology 78: 861-866.

Moher, D, Schulz, KF, Altman, DG (2001). The CONSORT statement: revised

recommendations for improving the quality of reports of parallel-group randomised trials. Lancet 357: 1191-1194.

Mohtar, M, Shaari, K, Ali, NAM, Ali, AM (1998). Antimicrobial activity of selected

Malaysian plants against micro-organisms related to skin infections. Journal of Tropical Forest Products 4: 199-206.

Möllenbeck, S, Konig, T, Schreier, P, Schwab, W, Rajaonarivony, J, Ranarivelo (1997).

Chemical composition and analyses of enantiomers of essential oils from Madagascar. Flavour & Fragrance Journal 12: 63-69.

Moon, Y, Glasgow, WC, Eling, TE (2005). Curcumin suppresses interleukin 1beta-mediated

microsomal prostaglandin E synthase 1 by altering early growth response gene 1 and other signaling pathways. Journal of Pharmacology & Experimental Therapeutics 315: 788-795.

Morikawa, T, Ando, S, Matsuda, H, Kataoka, S, Muraoka, O, Yoshikawa, M (2005).

Inhibitors of nitric oxide production from the rhizomes of Alpinia galanga: structures of new 8-9' linked neolignans and sesquineolignan. Chemical & Pharmaceutical Bulletin 53: 625-630.

Mowrey, DB, Clayson, DE (1982). Motion sickness, ginger, and psychophysics. Lancet 1:

655-657. Murakami, A, Kondo, A, Nakamura, Y, Ohigashi, H, Koshimizu, K (1993). Possible anti-

tumour promoting properties of edible plants from Thailand, and identification of an active constituent, cardamonin, of Boesenbergia pandurata. Bioscience Biotechnology & Biochemistry 57: 1971-1973.

Murakami, A, Ohigashi, H, Koshimizu, K (1994). Possible anti-tumour promoting properties

of traditional Thai food items and some of their active constituents. Asia Pacific Journal of Clinical Nutrition 3: 185-191.

Page 245: Phytochemistry and pharmacology of plants from the ginger family

221

Murakami, A, Takahashi, D, Kinoshita, T, Koshimizu, K, Kim, HW, Yoshihiro, A,

Nakamura, Y, Jiwajinda, S, Terao, J, Ohigashi, H (2002). Zerumbone, a Southeast Asian ginger sesquiterpene, markedly suppresses free radical generation, proinflammatory protein production, and cancer cell proliferation accompanied by apoptosis: the alpha,beta-unsaturated carbonyl group is a prerequisite. Carcinogenesis 23: 795-802.

Muraoka, O, Fujimoto, M, Tanabe, G, Kubo, M, Minematsu, T, Matsuda, H, Morikawa, T,

Toguchida, I, Yoshikawa, M (2001). Absolute stereostructures of novel norcadinane- and trinoreudesmane-type sesquiterpenes with nitric oxide production inhibitory activity from Alpinia oxyphylla. Bioorganic and Medicinal Chemistry Letters 11: 2217-2220.

Murugaiah, JS, Namasivayam, N, Menon, VP (1999). Effect of ginger (Zingiber officinale

R.) on lipids in rats fed atherogenic diet. Journal of Clinical Biochemistry and Nutrition 27: 79.

Nagano, T, Oyama, Y, Kajita, N, Chikahisa, L, Nakata, M, Okazaki, E, Masuda (1997). New

curcuminoids isolated from Zingiber cassumunar protect cells suffering from oxidative stress: A flow-cytometric study using rat thymocytes and H2O2. Japanese Journal of Pharmacology 75: 363-370.

Nakamura, H, Yamamoto, T (1983). The active part of the [6]-gingerol molecule in

mutagenesis. Mutation Research 122: 87-94. Nakasone, Y, Yonaha, M, Wada, K, Adaniya, S, Kikuzaki, H, Nakatani, N (1999).

[Evaluation of pungency in the diploid and tetraploid types of ginger (Zingiber officinale Roscoe) and antioxidative activity of their methanol extracts]. Japanese Journal of Tropical Agriculture 43: 71-75.

Nakatani, N, Kikuzaki, H, Yamaji, H, Yoshio, K, Kitora, C, Okada, K, Padolina, WG (1994).

Labdane diterpenes from rhizomes of Hedychium coronarium. Phytochemistry 37: 1383-1388.

Nakazawa, T, Ohsawa, K (2002). Metabolism of [6]-gingerol in rats. Life Sciences 70: 2165-

2175. Nanji, AA, Jokelainen, K, Tipoe, GL, Rahemtulla, A, Thomas, P, Dannenberg, AJ (2003).

Curcumin prevents alcohol-induced liver disease in rats by inhibiting the expression of NF-kappa B-dependent genes. American Journal of Physiology - Gastrointestinal & Liver Physiology 284: G321-327.

Nanthakomon, T, Pongrojpaw, D (2006). The efficacy of ginger in prevention of

postoperative nausea and vomiting after major gynecologic surgery. Journal of the Medical Association of Thailand 89 Suppl 4: S130-136.

Naora, K, Ding, G, Hayashibara, M, Katagiri, Y, Kano, Y, Iwamoto, K (1992).

Pharmacokinetics of [6]-gingerol after intravenous administration in rats with acute renal or hepatic failure. Chemical & Pharmaceutical Bulletin 40: 1295-1298.

Page 246: Phytochemistry and pharmacology of plants from the ginger family

222

Narayan, S (2004). Curcumin, a multi-functional chemopreventive agent, blocks growth of colon cancer cells by targeting beta-catenin-mediated transactivation and cell-cell adhesion pathways. Journal of Molecular Histology 35: 301-307.

Needleman, P, Turk, J, Jakschik, BA, Morrison, AR, Lefkowith, JB (1986). Arachidonic acid

metabolism. Annual Review of Biochemistry 55: 69-102. Ninfali, P, Mea, G, Giorgini, S, Rocchi, M, Bacchiocca, M (2005). Antioxidant capacity of

vegetables, spices and dressings relevant to nutrition. British Journal of Nutrition 93: 257-266.

Noro, T, Miyase, T, Kuroyanagi, M (1983). Monoamine oxidase inhibitor from the rhizomes

of Kaempferia galanga L. Chemical & Pharmaceutical Bulletin 31: 2708-2711. Nurtjahja-Tjendraputra, E, Ammit, AJ, Roufogalis, BD, Tran, VH, Duke, CC (2003).

Effective anti-platelet and COX-1 enzyme inhibitors from pungent constituents of ginger. Thrombosis Research 111: 259.

Ohtsuki, T, Tamaki, M, Toume, K, Ishibashi, M (2008). A novel sesquiterpenoid dimer

parviflorene F induces apoptosis by up-regulating the expression of TRAIL-R2 and a caspase-dependent mechanism. Bioorganic & Medicinal Chemistry 16: 1756-1763.

Olszanecki, R, Jawie, J, Gajda, M, Mateuszuk, L, Gebska, A, Korabiowska, M, Chlopicki, S,

Korbut, R (2005). Effect of curcumin on atherosclerosis in apoE/LDLR-double knockout mice. Journal of Physiology & Pharmacology 56: 627-635.

Omata, A, Yomogida, K, Teshima, Y, Nakamura, S, Hashimoto, S, Arai, T, Furukawa, K

(1991). Volatile components of ginger flowers (Hedychium coronarium Koenig). Flavour & Fragrance Journal 6: 217-220.

Ono, K, Hasegawa, K, Naiki, H, Yamada, M (2004). Curcumin has potent anti-

amyloidogenic effects for Alzheimer's beta-amyloid fibrils in vitro. Journal of Neuroscience Research 75: 742-750.

Onogi, T, Minami, M, Kuraishi, Y, Satoh, M (1992). Capsaicin-like effect of (6)-shogaol on

substance P-containing primary afferents of rats: a possible mechanism of its analgesic action. Neuropharmacology 31: 1165-1169.

Otero, R, Nunez, V, Barona, J, Fonnegra, R, Jimenez, SL, Osorio, RG, Saldarriaga, M, Diaz,

A (2000). Snakebites and ethnobotany in the northwest region of colombia. Part III: neutralization of the haemorrhagic effect of Bothrops atrox venom. Journal of Ethnopharmacology 73: 233-241.

Ozaki, Y, Kawahara, N, Harada, M (1991). Anti-inflammatory effect of Zingiber

cassumunar Roxb. and its active principles. Chemical & Pharmaceutical Bulletin 39: 2353-2356.

Pal, S, Bhattacharyya, S, Choudhuri, T, Datta, GK, Das, T, Sa, G (2005). Amelioration of

immune cell number depletion and potentiation of depressed detoxification system of tumor-bearing mice by curcumin. Cancer Detection & Prevention 29: 470-478.

Page 247: Phytochemistry and pharmacology of plants from the ginger family

223

Pancho, LR, Kimura, I, Unno, R, Kurono, M, Kimura, M (1989). Reversed effects between

crude and processed ginger extracts on PGF2 alpha-induced contraction in mouse mesenteric veins. Japanese Journal of Pharmacology 50: 243-246.

Panthong, A, Kanjanapothi, D, Niwatananant, W, Tuntiwachwuttikul, P, Reutrakul, V

(1997). Anti-inflammatory activity of Compound D ((E)-4-(3',4'-dimethoxyphenyl)but-3-en-2-ol) isolated from Zingiber cassumunar Roxb. Phytomedicine 4: 207-212.

Park, C, Kim, GY, Kim, GD, Choi, BT, Park, Y-M, Choi, YH (2006). Induction of G2/M

arrest and inhibition of cyclooxygenase-2 activity by curcumin in human bladder cancer T24 cells. Oncology Reports 15: 1225-1231.

Park, EJ, Jeon, CH, Ko, G, Kim, J, Sohn, DH (2000). Protective effect of curcumin in rat

liver injury induced by carbon tetrachloride. Journal of Pharmacy & Pharmacology 52: 437-440.

Park, KK, Chun, KS, Lee, JM, Lee, SS, Surh, YJ (1998). Inhibitory effects of [6]-gingerol, a

major pungent principle of ginger, on phorbol ester-induced inflammation, epidermal ornithine decarboxylase activity and skin tumor promotion in ICR mice. Cancer Letters 129: 139-144.

Pechous, SW, Whitaker, BD (2004). Cloning and functional expression of an (E,E)-α-

farnesene synthase cDNA from peel tissue of apple fruit. Planta 219: 84-94. Penna, SC, Medeiros, MV, Aimbire, FS, Faria-Neto, HC, Sertie, JA, Lopes-Martins, RA

(2003). Anti-inflammatory effect of the hydralcoholic extract of Zingiber officinale rhizomes on rat paw and skin edema. Phytomedicine. 10: 381-385.

Pereira, dS, Bernardo, RR, Paz, PJ (1999). A new steroidal saponin from the rhizomes of

Costus spicatus. Planta Medica 65: 285-287. Pfeiffer, E, Heuschmid, FF, Kranz, S, Metzler, M (2006). Microsomal hydroxylation and

glucuronidation of [6]-gingerol. Journal of Agricultural & Food Chemistry 54: 8769-8774.

Phan, PV, Sohrabi, A, Polotsky, A, Hungerford, DS, Lindmark, L, Frondoza, CG (2005).

Ginger extract components suppress induction of chemokine expression in human synoviocytes. Journal of Alternative & Complementary Medicine 11: 149-154.

Phillips, S, Hutchinson, S, Ruggier, R (1993a). Zingiber officinale does not affect gastric

emptying rate. A randomised, placebo-controlled, crossover trial. Anaesthesia 48: 393-395.

Phillips, S, Ruggier, R, Hutchinson, SE (1993b). Zingiber officinale (ginger) - an antiemetic

for day case surgery. Anaesthesia 48: 715-717. Piper, JT, Singhal, SS, Salameh, MS, Torman, RT, Awasthi, YC, Awasthi, S (1998).

Mechanisms of anticarcinogenic properties of curcumin: The effect of curcumin on

Page 248: Phytochemistry and pharmacology of plants from the ginger family

224

glutathione linked detoxification enzymes in rat liver. International Journal of Biochemistry & Cell Biology 30: 445-456.

Piromrat, K, Tuchinda, M, Geadsomnuig, S, Koysooko, R, Bunjob, M (1986).

[Antihistaminic effect of Plai Zingiber cassumunar Roxb. on histamine skin test in asthmatic children]. Siriraj Hospital Gazette 38: 251-256.

Pithayanukul, P, Tubprasert, J, Wuthi-Udomlert, M (2007). In vitro antimicrobial activity of

Zingiber cassumunar (Plai) oil and a 5% Plai oil gel. Phytotherapy Research 21: 164-169.

Platel, K, Srinivasan, K (1996). Influence of dietary spices or their active principles on

digestive enzymes of small intestinal mucosa in rats. International Journal of Food Sciences & Nutrition 47: 55-59.

Platel, K, Srinivasan, K (2000). Influence of dietary spices and their active principles on

pancreatic digestive enzymes in albino rats. Nahrung. 44: 42-46. Plummer, SM, Holloway, KA, Manson, MM, Munks, RJ, Kaptein, A, Farrow, S, Howells, L

(1999). Inhibition of cyclo-oxygenase 2 expression in colon cells by the chemopreventive agent curcumin involves inhibition of NF-kappaB activation via the NIK/IKK signalling complex. Oncogene 18: 6013-6020.

Polasa, K, Naidu, AN, Ravindranath, I, Krishnaswamy, K (2004). Inhibition of B(a)P

induced strand breaks in presence of curcumin. Mutation Research 557: 203-213. Pongprayoon, U, Tuchinda, P, Claeson, P, Sematong, T, Reutrakul, V, Soontornsaratune, P

(1997). Topical antiinflammatory activity of the major lipophilic constituents of the rhizome of Zingiber cassumunar. Part II: Hexane extractives. Phytomedicine 3: 323-326.

Pongrojpaw, D, Chiamchanya, C (2003). The efficacy of ginger in prevention of post-

operative nausea and vomiting after outpatient gynecological laparoscopy. Journal of the Medical Association of Thailand. 86: 244-250.

Pongrojpaw, D, Somprasit, C, Chanthasenanont, A (2007). A randomized comparison of

ginger and dimenhydrinate in the treatment of nausea and vomiting in pregnancy. Journal of the Medical Association of Thailand 90: 1703-1709.

Portnoi, G, Chng, LA, Karimi-Tabesh, L, Koren, G, Tan, MP, Einarson, A (2003).

Prospective comparative study of the safety and effectiveness of ginger for the treatment of nausea and vomiting in pregnancy. American Journal of Obstetrics & Gynecology 189: 1374-1377.

Prasad, PN, Ammal, EKJ (1983). Costus malortieanus H. Wendl a new source for diosgenin.

Current Science 52: 825-826. Prasain, JK, Li, JX, Tezuka, Y, Tanaka, K, Basnet, P, Dong, H, Namba, T, Kadota, S (1998).

Calyxin H, epicalyxin H, and blepharocalyxins A and B, novel diarylheptanoids from the seeds of Alpinia blepharocalyx. Journal of Natural Products 61: 212-216.

Page 249: Phytochemistry and pharmacology of plants from the ginger family

225

Priyadarsini, KI, Maity, DK, Naik, GH, Kumar, MS, Unnikrishnan, MK, Satav, JG, Mohan, H (2003). Role of phenolic O-H and methylene hydrogen on the free radical reactions and antioxidant activity of curcumin. Free Radical Biology & Medicine 35: 475-484.

Puri, A, Sahai, R, Singh, KL, Saxena, RP, Tandon, JS, Saxena, KC (2000). Immunostimulant

activity of dry fruits and plant materials used in indian traditional medical system for mothers after child birth and invalids. Journal of Ethnopharmacology 71: 89-92.

Qureshi, S, Shah, AH, Ageel, AM (1992). Toxicity studies on Alpinia galanga and Curcuma

longa. Planta Medica 58: 124-127. Ramaswami, G, Chai, H, Yao, Q, Lin, PH, Lumsden, AB, Chen, C (2004). Curcumin blocks

homocysteine-induced endothelial dysfunction in porcine coronary arteries. Journal of Vascular Surgery 40: 1216-1222.

Ramirez-Tortosa, MC, Mesa, MD, Aguilera, MC, Quiles, JL, Baro, L, Ramirez-Tortosa, CL,

Martinez-Victoria, E, Gil, A (1999). Oral administration of a turmeric extract inhibits LDL oxidation and has hypocholesterolemic effects in rabbits with experimental atherosclerosis. Atherosclerosis 147: 371-378.

Ranjan, D, Johnston, TD, Wu, G, Elliott, L, Bondada, S, Nagabhushan, M (1998). Curcumin

blocks cyclosporine A-resistant CD28 costimulatory pathway of human T-cell proliferation. Journal of Surgical Research 77: 174-178.

Rao, CV, Rivenson, A, Simi, B, Reddy, BS (1995). Chemoprevention of colon

carcinogenesis by dietary curcumin, a naturally occurring plant phenolic compound. Cancer Research 55: 259-266.

Resch, M, Steigel, A, Chen, ZL, Bauer, R (1998). 5-Lipoxygenase and cyclooxygenase-1

inhibitory active compounds from Atractylodes lancea. Journal of Natural Products 61: 347-350.

Ribenfeld, D, Borzone, L (1999). Ranodomized double-blind study comparing ginger

(Zintona®) and dimenhydrinate in motion sickness. Healthnotes Review of Complementary and Integrative Medicine 6: 98-101.

Ringbom, T, Segura, L, Noreen, Y, Perera, P, Bohlin, L (1998). Ursolic acid from Plantago

major, a selective inhibitor of cyclooxygenase-2 catalyzed prostaglandin biosynthesis. Journal of Natural Products 61: 1212-1215.

Ringman, JM, Frautschy, SA, Cole, GM, Masterman, DL, Cummings, JL (2005). A potential

role of the curry spice curcumin in Alzheimer's disease. Current Alzheimer Research 2: 131-136.

Rodrigues, EB (2007). Inflammation in dry age-related macular degeneration.

Ophthalmologica 221: 143-152. Ruffin, MT, Normolle, DP, Heath, DD (2003). Dose escalation of curcumin in healthy

adults. Cancer Epidemiology Biomarkers & Prevention 12: 1324S.

Page 250: Phytochemistry and pharmacology of plants from the ginger family

226

Rui, CL, Lacaille-Dubois, M-A, Hanquet, B, Correia, M, Chauffert, B (1997). New diosgenin glycosides from Costus afer. Journal of Natural Products 60: 1165-1169.

Rukkumani, R, Aruna, K, Varma, PS, Rajasekaran, KN, Menon, VP (2004). Comparative

effects of curcumin and an analog of curcumin on alcohol and PUFA induced oxidative stress. Journal of Pharmacy & Pharmaceutical Sciences 7: 274-283.

Ryu, HS, Kim, HS (2004). [Effect of Zingiber officinale Roscoe extracts on mice immune

cell activation]. Korean Journal of Nutrition 37: 23-30. Sajithlal, GB, Chithra, P, Chandrakasan, G (1998). Effect of curcumin on the advanced

glycation and cross-linking of collagen in diabetic rats. Biochemical Pharmacology 56: 1607-1614.

Salh, B, Assi, K, Templeman, V, Parhar, K, Owen, D, Gomez-Munoz, A, Jacobson, K

(2003). Curcumin attenuates DNB-induced murine colitis. American Journal of Physiology - Gastrointestinal & Liver Physiology 285: G235-243.

Sanders, JKM, Hunter, BK (1987). Modern NMR spectroscopy. Oxford University Press:

Oxford. Sautebin, L (2000). Prostaglandins and nitric oxide as molecular targets for anti-

inflammatory therapy. Fitoterapia 71: S48-S57. Scheffer, JJC, Baerheim, SA, Gani, A (1981). Antifungal activity of Alpinia galanga. Planta

Medica 42: 140-141. Schink, M (1997). Mistletoe therapy for human cancer: the role of the natural killer cells.

Anti-Cancer Drugs 8 Suppl 1: S47-51. Schneider, I, Bucar, F (2005). Lipoxygenase inhibitors from natural plant sources. Part 2:

medicinal plants with inhibitory activity on arachidonate 12-lipoxygenase, 15-lipoxygenase and leukotriene receptor antagonists. Phytotherapy Research 19: 263-272.

Sekiguchi, M (2001). Pacovatinins A-C, new labdane diterpenoids from the seeds of

Renealmia exaltata. Journal of Natural Products 64: 1102-1106. Sekiguchi, M (2002). Renealtins A and B, new diarylheptanoids with a tetrahydrofuran ring

from the seeds of Renealmia exaltata. Journal of Natural Products 65: 375-376. Sekiwa, Y, Kubota, K, Kobayashi, A (2000). Isolation of novel glucosides related to

gingerdiol from ginger and their antioxidative activities. Journal of Agricultural & Food Chemistry 48: 373-377.

Sertie, JAA, Basile, AC, Oshiro, TT, Silva, FD, Mazella, AAG (1992). Preventive anti-ulcer

activity of the rhizome extract of Zingiber officinale. Fitoterapia 63: 55-59. Shah, BH, Nawaz, Z, Pertani, SA, Roomi, A, Mahmood, H, Saeed, SA, Gilani, AH (1999).

Inhibitory effect of curcumin, a food spice from turmeric, on platelet-activating factor-

Page 251: Phytochemistry and pharmacology of plants from the ginger family

227

and arachidonic acid-mediated platelet aggregation through inhibition of thromboxane formation and Ca2+ signaling. Biochemical Pharmacology 58: 1167-1172.

Sharma, C, Kaur, J, Shishodia, S, Aggarwal, BB, Ralhan, R (2006). Curcumin down

regulates smokeless tobacco-induced NF-kappaB activation and COX-2 expression in human oral premalignant and cancer cells. Toxicology 228: 1-15.

Sharma, JN, Srivastava, KC, Gan, EK (1994). Suppressive effects of eugenol and ginger oil

on arthritic rats. Pharmacology 49: 314-318. Sharma, JN, Ishak, FI, Yusof, APM, Srivastava, KC (1997). Effects of eugenol and ginger

oil on adjuvant arthritis and the kallikreins in rats. Asia Pacific Journal of Pharmacology 12: 9-14.

Sharma, JN, Al-Omran, A, Parvathy, SS (2007). Role of nitric oxide in inflammatory

diseases. Inflammopharmacology 15: 252-259. Sharma, RA, Euden, SA, Platton, SL, Cooke, DN, Shafayat, A, Hewitt, HR, Marczylo, TH,

Morgan, B, Hemingway, D, Plummer, SM, Pirmohamed, M, Gescher, AJ, Steward, WP (2004). Phase I clinical trial of oral curcumin: biomarkers of systemic activity and compliance. Clinical Cancer Research 10: 6847-6854.

Sharma, RA, Gescher, AJ, Steward, WP (2005). Curcumin: the story so far. European

Journal of Cancer 41: 1955-1968. Sharma, RK, Sarma, TC, Leclercq, PA (2002). Essential oils of Zingiber officinale Roscoe

from North East India. Journal of Essential Oil-Bearing Plants 5: 71-76. Sharma, SS, Gupta, YK (1998). Reversal of cisplatin-induced delay in gastric emptying in

rats by ginger (Zingiber officinale). Journal of Ethnopharmacology 62: 49-55. Shen, C-L, Hong, K-J, Kim, SW (2005). Comparative effects of ginger root (Zingiber

officinale Rosc.) on the production of inflammatory mediators in normal and osteoarthrotic sow chondrocytes. Journal of Medicinal Food 8: 149-153.

Shibuya, H, Hamamoto, Y, Cai, Y, Kitagawa, I (1987). A reinvestigation of the structure of

zederone, a furanogermacrane-type sesquiterpene from zedoary. Chemical and Pharmaceutical Bulletin 35: 924-927.

Shindo, K, Kato, M, Kinoshita, A, Kobayashi, A, Koike, Y (2006). Analysis of antioxidant

activities contained in the Boesenbergia pandurata Schult. rhizome. Bioscience, Biotechnology, and Biochemistry 70: 2281-2284.

Shishodia, S, Potdar, P, Gairola, CG, Aggarwal, BB (2003). Curcumin (diferuloylmethane)

down-regulates cigarette smoke-induced NF-kappaB activation through inhibition of IkappaBalpha kinase in human lung epithelial cells: correlation with suppression of COX-2, MMP-9 and cyclin D1. Carcinogenesis 24: 1269-1279.

Page 252: Phytochemistry and pharmacology of plants from the ginger family

228

Shoba, G, Joy, D, Joseph, T, Majeed, M, Rajendran, M, Srinivas, PS (1998). Influence of piperine on the pharmocokinetics of curcumin in animals and human volunteers. Planta Medica 64: 353-356.

Shoji, N, Iwasa, A, Takemoto, T, Ishida, Y, Ohizumi, Y (1982). Cardiotonic principles of

ginger (Zingiber officinale Roscoe). Journal of Pharmaceutical Sciences 71: 1174-1175. Sidhu, GS, Mani, H, Gaddipati, JP, Singh, AK, Seth, P, Banaudha, KK, Patnaik, GK,

Maheshwari, RK (1999). Curcumin enhances wound healing in streptozotocin induced diabetic rats and genetically diabetic mice. Wound Repair & Regeneration 7: 362-374.

Silva, BP, Bernardo, RR, Parente, JP (1998). New furostanol glycosides from Costus

spicatus. Fitoterapia 69: 528-532. Simon, LS, Weaver, AL, Graham, DY, Kivitz, AJ, Lipsky, PE, Hubbard, RC, Isakson, PC,

Verburg, KM, Yu, SS, Zhao, WW, Geis, GS (1999). Anti-inflammatory and upper gastrointestinal effects of celecoxib in rheumatoid arthritis. JAMA 282: 1921-1928.

Singh, G, Maurya, S, Catalan, C, de Lampasona, MP (2005). Studies on essential oil, Part

42: chemical, antifungal, antioxidant and sprout suppressant studies on ginger essential oil and its oleoresin. Flavour and Fragrance Journal 20: 1-6.

Singh, S, Dube, NK, Tripathi, SC, Singh, SK (1984). Fungitoxicity of some essential oils

against Aspergillus flavus. Indian Perfumer 28: 164-166. Singh, S, Aggarwal, BB (1995). Activation of transcription factor NF-kappa B is suppressed

by curcumin (diferuloylmethane). Journal of Biological Chemistry 270: 24995-25000. Singhal, SS, Awasthi, S, Pandya, U, Piper, JT, Saini, MK, Cheng, JZ, Awasthi, YC (1999).

The effect of curcumin on glutathione-linked enzymes in K562 human leukemia cells. Toxicology Letters 109: 87-95.

Sirat, HM (1994). Study on the terpenoids of Zingiber ottensii. Planta Medica 60: 497-497. Sirat, HM, Masri, D, Rahman, AA (1994). The distribution of labdane diterpenes in the

Zingiberaceae of Malaysia. Phytochemistry 36: 699-701. Sirat, HM, Nordin, AB (1994). Essential oil of Zingiber ottensii Valeton. Journal of

Essential Oil Research 6: 635-636. Sirat, HM, Leh, NHN (2001). The rhizome oil of Zingiber spectabile Valet. Journal of

Essential Oil Research 13: 256-257. Sirirugsa, P (1999). Thai Zingiberaceae: species diversity and their uses. 1997 IUPAC

Symposium; Pukhet. IUPAC. 1999. Smith, C, Crowther, C, Willson, K, Hotham, N, McMillian, V (2004a). A randomized

controlled trial of ginger to treat nausea and vomiting in pregnancy. Obstetrics & Gynecology 103: 639-645.

Page 253: Phytochemistry and pharmacology of plants from the ginger family

229

Smith, MK, Hamill, SD (1996). Field evaluation of micropropagated and conventionally propagated ginger in subtropical Queensland. Australian Journal of Experimental Agriculture 36: 347-354.

Smith, MK, Hamill, SD (2002). PBR 2000/161 Zingiber officinale Ginger 'Buderim Gold'.

Plant Varieties Journal 15: 85. Smith, MK, Hamill, SD, Gogel, BJ, Severn-Ellis, AA (2004b). Ginger (Zingiber officinale)

autotetraploids with improved processing quality produced by an in vitro colchicine treatment. Australian Journal of Experimental Agriculture 44: 1065-1072.

Smith, RM, Robinson, AM (1981). The essential oil of ginger from Fiji. Phytochemistry 20:

203-206. Smith, RM (1987). Zingiberaceae-Costaceae. Flora of Australia Volume 45. Australian

Government Publishing Service: Canberra. pp 19-37. Solecki, RS (1975). Shanidar IV, a neanderthal flower burial in northern Iraq. Science 190:

880-881. Someya, Y, Kobayashi, A, Kubota, K (2001). Isolation and identification of trans-2-and

trans-3-hydroxy-1,8-cineole glucosides from Alpinia galanga. Bioscience Biotechnology & Biochemistry 65: 950-953.

Song, Y, Sonawane, ND, Salinas, D, Qian, L, Pedemonte, N, Galietta, LJV, Verkman, AS

(2004). Evidence against the rescue of defective DeltaF508-CFTR cellular processing by curcumin in cell culture and mouse models. Journal of Biological Chemistry 279: 40629-40633.

South, EH, Exon, JH, Hendrix, K (1997). Dietary curcumin enhances antibody response in

rats. Immunopharmacology and Immunotoxicology 19: 105-119. Sreepriya, M, Bali, G (2006). Effects of administration of embelin and curcumin on lipid

peroxidation, hepatic glutathione antioxidant defense and hematopoietic system during N-nitrosodiethylamine/phenobarbital-induced hepatocarcinogenesis in Wistar rats. Molecular & Cellular Biochemistry 284: 49-55.

Srinivasan, K, Sambaiah, K (1991). The effect of spices on cholesterol 7 alpha-hydroxylase

activity and on serum and hepatic cholesterol levels in the rat. International Journal for Vitamin & Nutrition Research 61: 364-369.

Sripramote, M, Lekhyananda, N (2003). A randomized comparison of ginger and vitamin B6

in the treatment of nausea and vomiting of pregnancy. Journal of the Medical Association of Thailand 86: 846-853.

Srivastava, KC (1984a). Aqueous extracts of onion, garlic and ginger inhibit platelet

aggregation and alter arachidonic acid metabolism. Biomedica Biochimica Acta 43: S335-S346.

Page 254: Phytochemistry and pharmacology of plants from the ginger family

230

Srivastava, KC (1984b). Effects of aqueous extracts of onion, garlic and ginger on platelet aggregation and metabolism of arachidonic acid in the blood vascular system: in vitro study. Prostaglandins Leukotrienes & Medicine 13: 227-235.

Srivastava, KC (1986). Isolation and effects of some ginger components of platelet

aggregation and eicosanoid biosynthesis. Prostaglandins Leukotrienes & Medicine 25: 187-198.

Srivastava, KC, Mustafa, T (1989). Ginger (Zingiber officinale) and rheumatic disorders.

Medical Hypotheses 29: 25-28. Srivastava, KC, Mustafa, T (1992). Ginger (Zingiber officinale) in rheumatism and

musculoskeletal disorders. Medical Hypotheses 39: 342-348. Stewart, JJ, Wood, MJ, Wood, CD, Mims, ME (1991). Effects of ginger on motion sickness

susceptibility and gastric function. Pharmacology 42: 111-120. Stewart, L, Katial, R (2007). Exhaled nitric oxide. Immunology and Allergy Clinics of North

America 27: 571-586. Strasser, E-M, Wessner, B, Manhart, N, Roth, E (2005). The relationship between the anti-

inflammatory effects of curcumin and cellular glutathione content in myelomonocytic cells. Biochemical Pharmacology 70: 552-559.

Su, C-C, Chen, G-W, Lin, J-G, Wu, L-T, Chung, J-G (2006). Curcumin inhibits cell

migration of human colon cancer colo 205 cells through the inhibition of nuclear factor kappa B /p65 and down-regulates cyclooxygenase-2 and matrix metalloproteinase-2 expressions. Anticancer Research 26: 1281-1288.

Subehan, Usia, T, Iwata, H, Kadota, S, Tezuka, Y (2006). Mechanism-based inhibition of

CYP3A4 and CYP2D6 by Indonesian medicinal plants. Journal of Ethnopharmacology 105: 449-455.

Suekawa, M, Ishige, A, Yuasa, K, Sudo, K, Aburada, M, Hosoya, E (1984). Pharmacological

studies on ginger. I. Pharmacological actions of pungent constitutents, (6)-gingerol and (6)-shogaol. Journal of Pharmacobio-Dynamics 7: 836-848.

Suekawa, M, Aburada, M, Hosoya, E (1986). Pharmacological studies on ginger. II. Pressor

action of (6)-shogaol in anesthetized rats, or hindquarters, tail and mesenteric vascular beds of rats. Journal of Pharmacobio-Dynamics 9: 842-852.

Sugiyama, T, Nagata, J-i, Yamagishi, A, Endoh, K, Saito, M, Yamada, K, Yamada, S,

Umegaki, K (2006). Selective protection of curcumin against carbon tetrachloride-induced inactivation of hepatic cytochrome P450 isozymes in rats. Life Sciences 78: 2188-2193.

Sun, Y-M, Zhang, H-Y, Chen, D-Z, Liu, C-B (2002). Theoretical elucidation on the

antioxidant mechanism of curcumin: a DFT study. Organic Letters 4: 2909-2911.

Page 255: Phytochemistry and pharmacology of plants from the ginger family

231

Surh, YJ, Lee, SS (1994a). Enzymic reduction of [6]-gingerol, a major pungent principle of ginger, in the cell-free preparation of rat liver. Life Sciences 54: L321-L326.

Surh, YJ, Lee, SS (1994b). Enzymatic reduction of xenobiotic alpha,beta-unsaturated

ketones: formation of allyl alcohol metabolites from shogaol and dehydroparadol. Research Communications in Chemical Pathology & Pharmacology 84: 53-61.

Suryanarayana, P, Krishnaswamy, K, Reddy, GB (2003). Effect of curcumin on galactose-

induced cataractogenesis in rats. Molecular Vision 9: 223-230. Suzuki, M, Nakamura, T, Iyoki, S, Fujiwara, A, Watanabe, Y, Mohri, K, Isobe, K, Ono, K,

Yano, S (2005). Elucidation of anti-allergic activities of curcumin-related compounds with a special reference to their anti-oxidative activities. Biological & Pharmaceutical Bulletin 28: 1438-1443.

Swarnakar, S, Ganguly, K, Kundu, P, Banerjee, A, Maity, P, Sharma, AV (2005). Curcumin

regulates expression and activity of matrix metalloproteinases 9 and 2 during prevention and healing of indomethacin-induced gastric ulcer. Journal of Biological Chemistry 280: 9409-9415.

Takahashi, M, Koyano, T, Kowithayakorn, T, Hayashi, M, Komiyama, K, Ishibashi, M

(2003). Parviflorene A, a novel cytotoxic unsymmetrical sesquiterpene-dimer constituent from Curcuma parviflora. Tetrahedron Letters 44: 2327-2329.

Tandan, SK, Chandra, S, Gupta, S, Lal, J (1997). Analgesic and anti-inflammatory effects of

Hedychium spicatum. Indian Journal of Pharmaceutical Sciences 59: 148-150. Taroeno, Brophy, JJ, Zwaving, JH (1991). Analysis of the essential oil of Zingiber

cassumunar Roxb. from Indonesia. Flavour & Fragrance Journal 6: 161-164. Tewari, A, Pant, AK, Mengi, N, Patra, NK (1999). A review on Alpinia species: chemical,

biocidal and pharmacological aspects. Journal of Medicinal and Aromatic Plant Sciences 21: 1155-1168.

Tewtrakul, S, Subhadhirasakul, S (2007). Anti-allergic activity of some selected plants in the

Zingiberaceae family. Journal of Ethnopharmacology 109: 535-538. Thangapazham, RL, Sharma, A, Maheshwari, RK (2006). Multiple molecular targets in

cancer chemoprevention by curcumin. AAPS Journal 8: E443-449. Thomson, M, Al-Qattan, KK, Al-Sawan, SM, Alnaqeeb, MA, Khan, I, Ali, M (2002). The

use of ginger (Zingiber officinale Rosc.) as a potential anti-inflammatory and antithrombotic agent. Prostaglandins Leukotrienes & Essential Fatty Acids. 67: 475-478.

Thubthimthed, S, Limsiriwong, P, Rerk-am, U, Suntorntanasat, T (2005). Chemical

composition and cytotoxic activity of the essential oil of Zingiber ottensii. III WOCMAP Congress on Medicinal and Aromatic Plants - Volume 1: Bioprospecting and Ethnopharmacology. pp 107-109.

Page 256: Phytochemistry and pharmacology of plants from the ginger family

232

Tilak, JC, Banerjee, M, Mohan, H, Devasagayam, TP (2004). Antioxidant availability of turmeric in relation to its medicinal and culinary uses. Phytotherapy Research 18: 798-804.

Tiwawech, D, Hirose, M, Futakuchi, M, Lin, C, Thamavit, W, Ito, N, Shirai, T (2000).

Enhancing effects of Thai edible plants on 2-amino-3,8-dimethylimidazo(4,5-f)quinoxaline-hepatocarcinogenesis in a rat medium-term bioassay. Cancer Letters 158: 195-201.

Tjendraputra, E, Tran, VH, Liu-Brennan, D, Roufogalis, BD, Duke, CC (2001). Effect of

ginger constituents and synthetic analogues on cyclooxygenase-2 enzyme in intact cells. Bioorganic Chemistry 29: 156-163.

Tomalski, MD, Blum, MS, Jones, TH, Fales, HM, Howard, DF, Passera, L (1987).

Chemistry and functions of exocrine secretions of the ants Tapinoma melanocephalum and T. erraticum. Journal of Chemical Ecology 13: 253-263.

Toume, K, Takahashi, M, Yamaguchi, K, Koyano, T, Kowithayakorn, T, Hayashi, M,

Komiyama, K, Ishibashi, M (2004). Parviflorenes B-F, novel cytotoxic unsymmetrical sesquiterpene-dimers with three backbone skeletons from Curcuma parviflora. Tetrahedron 60: 10817-10824.

Toume, K, Sato, M, Koyano, T, Kowithayakorn, T, Yamori, T, Ishibashi, M (2005).

Cytotoxic dimeric sesquiterpenoids from Curcuma parviflora: isolation of three new parviflorenes and absolute stereochemistry of parviflorenes A, B, D, F, and G. Tetrahedron 61: 6700-6706.

Trakoontivakorn, G, Nakahara, K, Shinmoto, H, Takenaka, M, Onishi-Kameyama, M, Ono,

H, Yoshida, M, Nagata, T, Tsushida, T (2001). Structural analysis of a novel antimutagenic compound, 4-hydroxypanduratin A, and the antimutagenic activity of flavonoids in a Thai spice, fingerroot (Boesenbergia pandurata Schult) against mutagenic heterocyclic amines. Journal of Agricultural & Food Chemistry 49: 3046-3050.

Tresh, JC (1879). Proximate analysis of the rhizome of Zingiber officinalis, and comparative

examination of typical specimens of commercial gingers. Pharmaceutical Journal 15: 171.

Tripathi, P, Tripathi, P, Kashyap, L, Singh, V (2007). The role of nitric oxide in

inflammatory reactions. FEMS Immunology And Medical Microbiology 51: 443-452. Tunstall, RG, Sharma, RA, Perkins, S, Sale, S, Singh, R, Farmer, PB, Steward, WP, Gescher,

AJ (2006). Cyclooxygenase-2 expression and oxidative DNA adducts in murine intestinal adenomas: modification by dietary curcumin and implications for clinical trials. European Journal of Cancer 42: 415-421.

Ukil, A, Maity, S, Karmakar, S, Datta, N, Vedasiromoni, JR, Das, PK (2003). Curcumin, the

major component of food flavour turmeric, reduces mucosal injury in trinitrobenzene sulphonic acid-induced colitis. British Journal of Pharmacology 139: 209-218.

Page 257: Phytochemistry and pharmacology of plants from the ginger family

233

Valli, J (2003). The use of leukotriene receptor antagonists to treat allergic rhinitis. PharmaNote 18: 9.

Vasudevan, K, Vembar, S, Veeraraghavan, K, Haranath, PS (2000). Influence of intragastric

perfusion of aqueous spice extracts on acid secretion in anesthetized albino rats. Indian Journal of Gastroenterology. 19: 53-56.

Vaughan, JG, Geissler, A (1997). The New Oxford Book of Food Plants. Oxford University

Press: Oxford. Venkatesan, N, Punithavathi, V, Chandrakasan, G (1997). Curcumin protects bleomycin-

induced lung injury in rats. Life Sciences 61: PL51-58. Venkatesan, N (1998). Curcumin attenuation of acute adriamycin myocardial toxicity in rats.

British Journal of Pharmacology 124: 425-427. Venkatesan, N, Punithavathi, D, Arumugam, V (2000). Curcumin prevents adriamycin

nephrotoxicity in rats. British Journal of Pharmacology 129: 231-234. Venkateshwarlu, V (1997). Cyclo-oxygenase inhibitors from spices. Indian Drugs 34: 427-

432. Verma, SK, Singh, M, Jain, P, Bordia, A (2004). Protective effect of ginger, Zingiber

officinale Rosc on experimental atherosclerosis in rabbits. Indian Journal of Experimental Biology 42: 736-738.

Vernin, G, Parkanyi, C (1994). Ginger oil (Zingiber officinale Roscoe). In: Charalambous, G

(ed). Spices, Herbs and Edible Fungi. Elsevier Science: Amsterdam. pp 579-594. Vesper, H, Kindsmueller, C, Meurens, M, Kollmannsberger, H, Nitz, S (1995). Isolation and

quantification of the pungent principles of ginger. Chemie Mikrobiologie Technologie der Lebensmittel 17 (3-4) 17: 114-117.

Viel, TA, Domingos, CD, Monteiro, APD, Lima-Landman, MTR, Lapa, AJ, Souccar, C

(1999). Evaluation of the antiurolithiatic activity of the extract of Costus spiralis Roscoe in rats. Journal of Ethnopharmacology 66: 193-198.

Villanueva, MA, Torres, RC, Baser, KHC, Özek, T, Kürkçüoglu, M (1993). The composition

of Manila elemi oil. Flavour and Fragrance Journal 8: 35-37. Vimala, S, Norhanom, AW, Yadav, M (1999). Anti-tumour promoter activity in Malaysian

ginger rhizobia used in traditional medicine. British Journal of Cancer 80: 110-116. Visalyaputra, S, Petchpaisit, N, Somcharoen, K, Choavaratana, R (1998). The efficacy of

ginger root in the prevention of postoperative nausea and vomiting after outpatient gynaecological laparoscopy. Anaesthesia 53: 506-510.

Volate, SR, Davenport, DM, Muga, SJ, Wargovich, MJ (2005). Modulation of aberrant crypt

foci and apoptosis by dietary herbal supplements (quercetin, curcumin, silymarin, ginseng and rutin). Carcinogenesis 26: 1450-1456.

Page 258: Phytochemistry and pharmacology of plants from the ginger family

234

Vuolteenaho, K, Moilanen, T, Knowles, RG, Moilanen, E (2007). The role of nitric oxide in

osteoarthritis. Scandinavian Journal of Rheumatology 36: 247-258. Vutyavanich, T, Kraisarin, T, Ruangsri, R (2001). Ginger for nausea and vomiting in

pregnancy: randomized, double-masked, placebo-controlled trial. Obstetrics & Gynecology. 97: 577-582.

Wallace, ME, Smyth, MJ (2005). The role of natural killer cells in tumor control - effectors

and regulators of adaptive immunity. Springer Seminars in Immunopathology: Published online. 10.1007/s00281-00004-00195-x.

Wei, Q-Y, Ma, J-P, Cai, Y-J, Yang, L, Liu, Z-L (2005). Cytotoxic and apoptotic activities of

diarylheptanoids and gingerol-related compounds from the rhizome of Chinese ginger. Journal of Ethnopharmacology 102: 177-184.

Weidner, MS, Sigwart, K (2001). Investigation of the teratogenic potential of a Zingiber

officinale extract in the rat. Reproductive Toxicology 15: 75-80. Weiss, EA (2002). Spice Crops. CAB International: Oxon, UK. Weyerstahl, P, Marschall, H, Thefeld, K, Subba, GC (1998). Constituents of the essential oil

from the rhizomes of Hedychium gardnerianum Roscoe. Flavour & Fragrance Journal 13: 377-388.

WHO (1989). Medicinal Plants in China. World Health Organization: Manila. WHO (1999). WHO Monographs on Selected Medicinal Plants Volume 1. World Health

Organisation: Geneva. Wigler, I, Grotto, I, Caspi, D, Yaron, M (2003). The effects of Zintona EC (a ginger extract)

on symptomatic gonarthritis. Osteoarthritis and Cartilage 11: 783-789. Wijesekera, ROB (1991). The Medicinal Plant Industry. CRC Press: Boca Raton, Florida. Wilasrusmee, C, Siddiqui, J, Bruch, D, Wilasrusmee, S, Kittur, S, Kittur, DS (2002). In vitro

immunomodulatory effects of herbal products. American Surgeon. 68: 860-864. Wilkinson, JM (2000). Effect of ginger tea on the fetal development of Sprague-Dawley rats.

Reproductive Toxicology. 14: 507-512. Willetts, KE, Ekangaki, A, Eden, JA (2003). Effect of a ginger extract on pregnancy-induced

nausea: a randomised controlled trial. Australian & New Zealand Journal of Obstetrics & Gynaecology 43: 139-144.

Williams, DH, Fleming, I (1989). Spectroscopic methods in organic chemistry. 4th edn.

McGraw-Hill: London. Williamson, EM (2002). Major Herbs of Auyrveda. Churchill Livingstone.

Page 259: Phytochemistry and pharmacology of plants from the ginger family

235

Willuhn, G, Pretzsch, G (1985). Diosgenin and sterols from Costus spiralis. Planta Medica 51: 185-187.

Wilson, PG, Hastings, SM (1993). Zingiberaceae. In: Harden, GJ (ed). Flora of New South

Wales Volume 4. New South Wales University Press: Kensington, NSW. pp 254-256. Wojcikowski, K, Stevenson, L, Leach, D, Wohlmuth, H, Gobe, G (2007). Antioxidant

capacity of 55 medicinal herbs traditionally used to treat the urinary system: a comparison using a sequential three-solvent extraction process. The Journal of Alternative and Complementary Medicine 13: 103-110.

Woo, CWH, Man, RYK, Siow, YL, Choy, PC, Wan, EWY, Lau, CS, O, K (2005).

Ganoderma lucidum inhibits inducible nitric oxide synthase expression in macrophages. Molecular and Cellular Biochemistry 275: 165-171.

Wright, CW (2002). Antiprotozoal natural products. In: Evans, WC (ed). Trease and Evans

Pharmacognosy. 15th edn. W. B. Saunders: Edinburgh. Wube, AA, Gibbons, S, Asres, K, Streit, B, Adams, M, Bauer, R, Bucar, F (2006). In vitro

12(S)-HETE and leukotriene metabolism inhibitory activity of sesquiterpenes of Warburgia ugandensis. Planta Medica: 754-756.

Xu, J, Fu, Y, Chen, A (2003). Activation of peroxisome proliferator-activated receptor-

gamma contributes to the inhibitory effects of curcumin on rat hepatic stellate cell growth. American Journal of Physiology - Gastrointestinal & Liver Physiology 285: G20-30.

Yadav, PN, Liu, Z, Rafi, MM (2003). A diarylheptanoid from lesser galangal (Alpinia

officinarum) inhibits proinflammatory mediators via inhibition of mitogen-activated protein kinase, p44/42, and transcription factor nuclear factor-κB. Journal of Pharmacology and Experimental Therapeutics 305: 925-931.

Yadav, VS, Mishra, KP, Singh, DP, Mehrotra, S, Singh, VK (2005). Immunomodulatory

effects of curcumin. Immunopharmacology and Immunotoxicology 27: 485 - 497. Yamahara, J, Miki, K, Chisaka, T, Sawada, T, Fujimura, H, Tomimatsu, T, Nakano, K,

Nohara, T (1985). Cholagogic effect of ginger and its active constituents. Journal of Ethnopharmacology 13: 217-225.

Yamahara, J, Mochizuki, M, Rong, HQ, Matsuda, H, Fujimura, H (1988). The anti-ulcer

effect in rats of ginger constituents. Journal of Ethnopharmacology 23: 299-304. Yamahara, J, Huang, QR, Li, YH, Xu, L, Fujimura, H (1990). Gastrointestinal motility

enhancing effect of ginger and its active constituents. Chemical & Pharmaceutical Bulletin 38: 430-431.

Yamazaki, R, Hatano, H, Aiyama, R, Matsuzaki, T, Hashimoto, S, Yokokura, T (2000).

Diarylheptanoids suppress expression of leukocyte adhesion molecules on human vascular endothelial cells. European Journal of Pharmacology. 404: 375-385.

Page 260: Phytochemistry and pharmacology of plants from the ginger family

236

Yang, F, Lim, GP, Begum, AN, Ubeda, OJ, Simmons, MR, Ambegaokar, SS, Chen, PP, Kayed, R, Glabe, CG, Frautschy, SA, Cole, GM (2005). Curcumin inhibits formation of amyloid beta oligomers and fibrils, binds plaques, and reduces amyloid in vivo. Journal of Biological Chemistry 280: 5892-5901.

Yang, SW (1999). A new labdane diterpenoid from Renealmia alpinia collected in the

Suriname rainforest. Journal of Natural Products 62: 1173-1174. Yang, X, Eilerman, RG (1999). Pungent principal of Alpinia galangal (L.) Swartz and its

applications. Journal of Agricultural & Food Chemistry 47: 1657-1662. YawBin, H, JiaYou, F, ChenHsun, H, PaoChu, W, YiHung, T (1999). Cyclic monoterpene

extract from cardamom oil as a skin permeation enhancer for indomethacin: in vitro and in vivo studies. Biological & Pharmaceutical Bulletin 22: 642-646.

Ye, Y, Li, B (2006). 1'S-1'-Acetoxychavicol acetate isolated from Alpinia galanga inhibits

human immunodeficiency virus type 1 replication by blocking Rev transport. Journal of General Virology 87: 2047-2053.

Yokoyama, WM, Kim, S, French, AR (2004). The dynamic life of natural killer cells.

Annual Review of Immunology 22: 405-429. Yoon, JH, Baek, SJ (2005). Molecular targets of dietary polyphenols with anti-inflammatory

properties. Yonsei Medical Journal 46: 585-596. Yoshikawa, M, Hatakeyama, S, Taniguchi, K, Matuda, H, Yamahara, J (1992). 6-

Gingesulfonic acid, a new anti-ulcer principle, and gingerglycolipids A, B, and C, three new monoacyldigalactosylglycerols, from zingiberis rhizoma originating in Taiwan. Chemical & Pharmaceutical Bulletin 40: 2239-2241.

Youn, HS, Saitoh, SI, Miyake, K, Hwang, DH (2006). Inhibition of homodimerization of

Toll-like receptor 4 by curcumin. Biochemical Pharmacology 72: 62-69. Young, H-Y, Chiang, C-T, Huang, Y-L, Pan, FP, Chen, G-L (2002). Analytical and stability

studies of ginger preparations. Journal of Food and Drug Analysis 10: 149-153. Young, HY, Luo, YL, Cheng, HY, Hsieh, WC, Liao, JC, Peng, WH (2005). Analgesic and

anti-inflammatory activities of [6]-gingerol. Journal of Ethnopharmacology 96: 207-210. Young, JM, Wagner, BM, Spires, DA (1983). Tachyphylaxis in 12-0-tetradecanoylphorbol

acetate- and arachidonic acid-induced ear edema. Journal of Investigative Dermatology 80: 48-52.

Yoysungnoen, P, Wirachwong, P, Bhattarakosol, P, Niimi, H, Patumraj, S (2006). Effects of

curcumin on tumor angiogenesis and biomarkers, COX-2 and VEGF, in hepatocellular carcinoma cell-implanted nude mice. Clinical Hemorheology & Microcirculation 34: 109-115.

Yun, J-M, Kwon, H, Hwang, J-K (2003). In vitro anti-inflammatory activity of panduratin A

isolated from Kaempferia pandurata in RAW264.7 cells. Planta Medica: 1102-1108.

Page 261: Phytochemistry and pharmacology of plants from the ginger family

237

Yun, J-M, Kwon, H, Mukhtar, H, Hwang, J-K (2005). Induction of apoptosis by panduratin

A isolated from Kaempferia pandurata in human colon cancer HT-29 cells. Planta Medica: 501-507.

Zhai, Z, Liu, Y, Wu, L, Senchina, DS, Wurtele, ES, Murphy, PA, Kohut, ML, Cunnick, JE

(2007). Enhancement of innate and adaptive immune functions by multiple Echinacea species. Journal of Medicinal Food 10: 423-434.

Zhang, F, Altorki, NK, Mestre, JR, Subbaramaiah, K, Dannenberg, AJ (1999). Curcumin

inhibits cyclooxygenase-2 transcription in bile acid- and phorbol ester-treated human gastrointestinal epithelial cells. Carcinogenesis 20: 445-451.

Zhang, H-G, Kim, H, Liu, C, Yu, S, Wang, J, Grizzle, WE, Kimberly, RP, Barnes, S (2007).

Curcumin reverses breast tumor exosomes mediated immune suppression of NK cell tumor cytotoxicity. Biochimica et Biophysica Acta 1773: 1116-1123.

Zhang, M, Deng, C, Zheng, J, Xia, J, Sheng, D (2006). Curcumin inhibits trinitrobenzene

sulphonic acid-induced colitis in rats by activation of peroxisome proliferator-activated receptor gamma. International Immunopharmacology 6: 1233-1242.

Zhang, X, Iwaoka, WT, Huang, AS, Nakamoto, ST, Wong, R (1994). Gingerol decreases

after processing and storage of ginger. Journal of Food Science 59: 1338. Zhang, Z, Rigas, B (2006). NF-kappaB, inflammation and pancreatic carcinogenesis: NF-

kappaB as a chemoprevention target (review). International Journal Of Oncology 29: 185-192.

Zhou, BN (1997). Bioactive labdane diterpenoids from Renealmia alpinia collected in the

Suriname rainforest. Journal of Natural Products 60: 1287-1293.

oooOooo