23.12.2012 Views

Essential Oil- Bearing Grasses

Essential Oil- Bearing Grasses

Essential Oil- Bearing Grasses

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Essential</strong> <strong>Oil</strong>-<br />

<strong>Bearing</strong> <strong>Grasses</strong><br />

The genus Cymbopogon


Medicinal and Aromatic Plants — Industrial Profiles<br />

Individual volumes in this series provide both industry and academia with in-depth coverage of<br />

one major genus of industrial importance.<br />

Series Edited by Dr. Roland Hardman<br />

Volume 1<br />

Valerian, edited by Peter J. Houghton<br />

Volume 2<br />

Perilla, edited by He-ci Yu,<br />

Kenichi Kosuna, and Megumi Haga<br />

Volume 3<br />

Poppy, edited by Jenö Bernáth<br />

Volume 4<br />

Cannabis, edited by David T. Brown<br />

Volume 5<br />

Neem, edited by H.S. Puri<br />

Volume 6<br />

Ergot, edited by Vladimír Kˇren and<br />

Ladislav Cvak<br />

Volume 7<br />

Caraway, edited by Éva Németh<br />

Volume 8<br />

Saffron, edited by Moshe Negbi<br />

Volume 9<br />

Tea Tree, edited by Ian Southwell and<br />

Robert Lowe<br />

Volume 10<br />

Basil, edited by Raimo Hiltunen and<br />

Yvonne Holm<br />

Volume 11<br />

Fenugreek, edited by<br />

Georgios Petropoulos<br />

Volume 12<br />

Ginkgo biloba, edited by<br />

Teris A. Van Beek<br />

Volume 13<br />

Black Pepper, edited by P.N. Ravindran<br />

Volume 14<br />

Sage, edited by Spiridon E. Kintzios<br />

Volume 15<br />

Ginseng, edited by W.E. Court<br />

Volume 16<br />

Mistletoe, edited by Arndt Büssing<br />

Volume 17<br />

Tea, edited by Yong-su Zhen<br />

Volume 18<br />

Artemisia, edited by Colin W. Wright<br />

Volume 19<br />

Stevia, edited by A. Douglas Kinghorn<br />

Volume 20<br />

Vetiveria, edited by Massimo Maffei<br />

Volume 21<br />

Narcissus and Daffodil, edited by<br />

Gordon R. Hanks<br />

Volume 22<br />

Eucalyptus, edited by<br />

John J.W. Coppen<br />

Volume 23<br />

Pueraria, edited by Wing Ming Keung<br />

Volume 24<br />

Thyme, edited by E. Stahl-Biskup and<br />

F. Sáez<br />

Volume 25<br />

Oregano, edited by Spiridon E. Kintzios<br />

Volume 26<br />

Citrus, edited by Giovanni Dugo and<br />

Angelo Di Giacomo<br />

Volume 27<br />

Geranium and Pelargonium, edited by<br />

Maria Lis-Balchin<br />

Volume 28<br />

Magnolia, edited by Satyajit D. Sarker<br />

and Yuji Maruyama<br />

Volume 29<br />

Lavender, edited by Maria Lis-Balchin<br />

Volume 30<br />

Cardamom, edited by P.N. Ravindran


Volume 31<br />

Hypericum, edited by Edzard Ernst<br />

Volume 32<br />

Taxus, edited by H. Itokawa and<br />

K.H. Lee<br />

Volume 33<br />

Capsicum, edited by Amit Krish De<br />

Volume 34<br />

Flax, edited by Alister Muir and<br />

Niel Westcott<br />

Volume 35<br />

Urtica, edited by Gulsel Kavalali<br />

Volume 36<br />

Cinnamon and Cassia, edited by<br />

P.N. Ravindran, K. Nirmal Babu, and<br />

M. Shylaja<br />

Volume 37<br />

Kava, edited by Yadhu N. Singh<br />

Volume 38<br />

Aloes, edited by Tom Reynolds<br />

Volume 39<br />

Echinacea, edited by<br />

Sandra Carol Miller<br />

Assistant Editor: He-ci Yu<br />

Volume 40<br />

Illicium, Pimpinella and Foeniculum,<br />

edited by Manuel Miró Jodral<br />

Volume 41<br />

Ginger, edited by P.N. Ravindran<br />

and K. Nirmal Babu<br />

Volume 42<br />

Chamomile: Industrial Profiles,<br />

edited by Rolf Franke and<br />

Heinz Schilcher<br />

Volume 43<br />

Pomegranates: Ancient Roots to Modern<br />

Medicine, edited by Navindra P.<br />

Seeram, Risa N. Schulman, and<br />

David Heber<br />

Volume 44<br />

Mint, edited by Brian M. Lawrence<br />

Volume 45<br />

Turmeric, edited by P. N. Ravindran,<br />

K. Nirmal Babu, and K. Sivaraman


<strong>Essential</strong> <strong>Oil</strong>-<br />

<strong>Bearing</strong> <strong>Grasses</strong><br />

The genus Cymbopogon<br />

Edited by Anand Akhila<br />

Medicinal and Aromatic Plants — Industrial Profiles


CRC Press<br />

Taylor & Francis Group<br />

6000 Broken Sound Parkway NW, Suite 300<br />

Boca Raton, FL 33487-2742<br />

© 2010 by Taylor and Francis Group, LLC<br />

CRC Press is an imprint of Taylor & Francis Group, an Informa business<br />

No claim to original U.S. Government works<br />

Printed in the United States of America on acid-free paper<br />

10 9 8 7 6 5 4 3 2 1<br />

International Standard Book Number: 978-0-8493-7857-7 (Hardback)<br />

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been<br />

made to publish reliable data and information, but the author and publisher cannot assume responsibility for the validity<br />

of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright<br />

holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this<br />

form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may<br />

rectify in any future reprint.<br />

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized<br />

in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying,<br />

microfilming, and recording, or in any information storage or retrieval system, without written permission from the<br />

publishers.<br />

For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://<br />

www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923,<br />

978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For<br />

organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged.<br />

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for<br />

identification and explanation without intent to infringe.<br />

Library of Congress Cataloging-in-Publication Data<br />

<strong>Essential</strong> oil-bearing grasses : the genus Cymbopogon / editor: Anand Akhila.<br />

p. cm. -- (Medicinal and aromatic plants--industrial profiles ; v. 46)<br />

Includes bibliographical references and index.<br />

ISBN 978-0-8493-7857-7 (hardcover : alk. paper)<br />

1. Cymbopogon. 2. Cymbopogon--Industrial applications. 3. Cymbopogon--Therapeutic use. I.<br />

Akhila, Anand. II. Title. III. Series: Medicinal and aromatic plants--industrial profiles ; v. 46.<br />

QK495.G74E745 2010<br />

661’.806--dc22 2009024407<br />

Visit the Taylor & Francis Web site at<br />

http://www.taylorandfrancis.com<br />

and the CRC Press Web site at<br />

http://www.crcpress.com


Contents<br />

Preface to the Series ..........................................................................................................................ix<br />

Preface...............................................................................................................................................xi<br />

Author/Editor Page ........................................................................................................................ xiii<br />

Contributors .....................................................................................................................................xv<br />

Chapter 1 The Genus Cymbopogon: Botany, Including Anatomy, Physiology,<br />

Biochemistry, and Molecular Biology ..........................................................................1<br />

Cinzia M. Bertea and Massimo E. Maffei<br />

Chapter 2 Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon ...............25<br />

Anand Akhila<br />

Chapter 3 The Cymbopogons: Harvest and Postharvest Management ..................................... 107<br />

A. K. Pandey<br />

Chapter 4 Biotechnological Studies in Cymbopogons: Current Status and Future Options .... 135<br />

Ajay K. Mathur<br />

Chapter 5 The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s ...................................... 151<br />

Rakesh Tiwari<br />

Chapter 6 In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon<br />

Species ...................................................................................................................... 167<br />

Watcharee Khunkitti<br />

Chapter 7 Thrombolysis-Accelerating Activity of <strong>Essential</strong> <strong>Oil</strong>s ............................................. 185<br />

Hiroyuki Sumi and Chieko Yatagai<br />

Chapter 8 Analytical Methods for Cymbopogon <strong>Oil</strong>s .............................................................. 195<br />

Ange Bighelli and Joseph Casanova<br />

Chapter 9 Citral from Lemongrass and Other Natural Sources: Its Toxicology<br />

and Legislation .........................................................................................................223<br />

David A. Moyler<br />

Index .............................................................................................................................................. 239<br />

vii


Preface to the Series<br />

There is increasing interest in industry, academia, and the health sciences in medicinal and aromatic<br />

plants. In passing from plant production to the eventual product used by the public, many sciences<br />

are involved. This series brings together information that is currently scattered through an everincreasing<br />

number of journals. Each volume gives an in-depth look at one plant genus about which<br />

an area specialist has assembled information ranging from the production of the plant to market<br />

trends and quality control.<br />

Many industries are involved, such as forestry, agriculture, chemical, food, flavor, beverage,<br />

pharmaceutical, cosmetic, and fragrance. The plant raw materials are roots, rhizomes, bulbs,<br />

leaves, stems, barks, wood, flowers, fruits, and seeds. These yield gums, resins, essential (volatile)<br />

oils, fixed oils, waxes, juices, extracts, and spices for medicinal and aromatic purposes. All these<br />

commodities are traded worldwide. A dealer’s market report for an item may say “drought in the<br />

country of origin has forced up prices.”<br />

Natural products do not mean safe products, and account of this has to be taken by the above<br />

industries, which are subject to regulation. For example, a number of plants that are approved for<br />

use in medicine must not be used in cosmetic products.<br />

The assessment of “safe to use” starts with the harvested plant material, which has to comply<br />

with an official monograph. This may require absence of, or prescribed limits of, radioactive material,<br />

heavy metals, aflatoxin, pesticide residue, as well as the required level of active principle. This<br />

analytical control is costly and tends to exclude small batches of plant material. Large-scale, contracted,<br />

mechanized cultivation with designated seed or plantlets is now preferable.<br />

Today, plant selection is not only for the yield of active principle, but for the plant’s ability to<br />

overcome disease, climatic stress, and the hazards caused by mankind. Such methods as in vitro<br />

fertilization, meristem cultures, and somatic embryogenesis are used. The transfer of sections of<br />

DNA is giving rise to controversy in the case of some end uses of the plant material.<br />

Some suppliers of plant raw material are now able to certify that they are supplying organically<br />

farmed medicinal plants, herbs, and spices. The Economic Union directive CVO/EU No. 2092/91<br />

details the specifications for the obligatory quality controls to be carried out at all stages of production<br />

and processing of organic products.<br />

Fascinating plant folklore and ethnopharmacology lead to medicinal potential. Examples are<br />

the muscle relaxants based on the arrow poison curare from species of Chondrodendron, and the<br />

antimalarials derived from species of Cinchona and Artemisia. The methods of detection of pharmacological<br />

activity have become increasingly reliable and specific, frequently involving enzymes<br />

in bioassays and avoiding the use of laboratory animals. By using bioassay-linked fractionation of<br />

crude plant juices or extracts, compounds can be specifically targeted which, for example, inhibit<br />

blood platelet aggregation, or have antitumor, antiviral, or any other required activity. With the assistance<br />

of robotic devices, all the members of a genus may be readily screened. However, the plant<br />

material must be fully authenticated by a specialist.<br />

The medicinal traditions of ancient civilizations such as those of China and India have a large<br />

armamentarium of plants in their pharmacopoeias that are used throughout Southeast Asia. A similar<br />

situation exists in Africa and South America. Thus, a very high percentage of the world’s population<br />

relies on medicinal and aromatic plants for their medicine. Western medicine is also responding.<br />

Already in Germany all medical practitioners have to pass an examination in phytotherapy before<br />

being allowed to practice. It is noticeable that medical, pharmacy, and health-related schools<br />

throughout Europe and the United States are increasingly offering training in phytotherapy.<br />

ix


x Preface to the Series<br />

Multinational pharmaceutical companies have become less enamored of the single compound,<br />

magic-bullet cure. The high costs of such ventures and the endless competition from “me-too” compounds<br />

from rival companies often discourage the attempt. Independent phytomedicine companies<br />

have been very strong in Germany. However, by the end of 1995, 11 (almost all) had been acquired<br />

by the multinational pharmaceutical firms, acknowledging the lay public’s growing demand for<br />

phytomedicines in the Western world.<br />

The business of dietary supplements in the Western world has expanded from the health store to<br />

the pharmacy. Alternative medicine includes plant-based products. Appropriate measures to ensure<br />

their quality, safety, and efficacy either already exist or are being answered by greater legislative<br />

control by such bodies as the U.S. Food and Drug Administration and the recently created European<br />

Agency for the Evaluation of Medicinal Products based in London.<br />

In the United States, the Dietary Supplement and Health Education Act of 1994 recognized the<br />

class of phytotherapeutic agents derived from medicinal and aromatic plants. Furthermore, under<br />

public pressure, the U.S. Congress set up an Office of Alternative Medicine, which in 1994 assisted<br />

the filing of several Investigational New Drug (IND) applications required for clinical trials of some<br />

Chinese herbal preparations. The significance of these applications was that each Chinese preparation<br />

involved several plants and yet was handled as a single IND. A demonstration of the contribution<br />

to efficacy of each ingredient of each plant was not required. This was a major step toward more<br />

sensible regulations in regard to phytomedicines.<br />

I always look forward to the journal HerbalGram (HG), of the American Botanical Council,<br />

which was founded by Mark Blumenthal in 1988. He continues as the Executive Director and<br />

as the Editor of HG. In it he regularly justifies the status of a medicinal plant and challenges<br />

official bodies when necessary. In HG Number 80 (2008), he tells the U.S. Food and Drug<br />

Administration to rescind the 1991IMPORT ALERT on the herb stevia (Vol. 19 in this series)<br />

because the United Nations and the World Health Organization have concluded that stevia<br />

extract, containing 95% stevia glycosides, is safe for human use as a sweetening agent, in the<br />

range 4 mg/kg body weight per day. This has paved the way for regulatory approval around the<br />

world for the use of this low-cost noncaloric material and notably for those obese persons facing<br />

cardiovascular disease and diabetes.<br />

For this volume, I thank its editor, Dr Anand Akhila, for his dedicated work and the chapter<br />

contributors for their authoritative information. My thanks are also due to Barbara Norwitz of CRC<br />

Press and her staff for their unfailing help.<br />

Roland Hardman, BPharm, BSc (Chem), PhD (London), FR Pharm S.


Preface<br />

The essential oils of the grasses of species of Cymbopogon have an industrial profile; they are used<br />

in beverages, foodstuffs, fragrances, household products, personal care products, pharmaceuticals,<br />

and in tobacco.<br />

These oils are sourced from around the world by, for example, Fuerest Day Lawson (FDL) Ltd. It<br />

is fascinating that this company is on the same site where plant raw material has arrived for the past<br />

400 years—close to the Tower of London and the River Thames. No longer in a warehouse but in a<br />

modern tower block, FDL has all the equipment to test the efficacy of an essential oil for a particular<br />

purpose. For the producer of the oil, FDL will advise on the development of a modern production<br />

process (in the country of origin if required), even to the stage of the final saleable product.<br />

FDL’s technical director, David A. Moyler, regularly attends meetings of the European Union<br />

in Brussels concerned with the regulations covering the commercial production and sale of such<br />

products, and he has contributed a relevant chapter to this book.<br />

The genus Cymbopogon (family Gramineae) has many species of grasses that grow in tropical<br />

and subtropical regions around the world from mountains to grasslands to arid zones. These plants<br />

produce essential oils with pleasant aromas in their leaves.<br />

Five species yield the three oils of main commercial importance: lemongrass from C. citratus of<br />

Malaysian origin (West Indian lemongrass) and C. flexuosus (East Indian lemongrass) from India,<br />

Sri Lanka, Burma, and Thailand; palmarosa oil from C. martinii; citronella oil from C. nardus (Sri<br />

Lanka), and C. winterianus (Java).<br />

This book describes the considerable ethno-botanical, phytochemical, and pharmacological<br />

knowledge that is associated with the multidimensional uses of the oils of the cymbopogons.<br />

Lemongrass originated in Asia and is an ingredient of its herbal teas, soups, and innumerable<br />

other food recipes found in South East Asia, Cambodia, and Vietnam. In Europe the oil is used in<br />

spiced wines and herbal beers. Citronella oil gives a pleasant, refreshing aroma to personal care<br />

products including mosquito repelling lotions, etc. Palmarosa oil supplies a rose note to fine perfumes<br />

and to, for example, perfumed candles and herbal pillows. These last two oils may result in a<br />

higher income for a farmer than from the traditional food crops.<br />

Extraction of the oils is by steam distillation and the aqueous distillates (hydrosols), after separation<br />

of the oils, are said to have antiviral, antibacterial, and antifungal properties. The spent grass,<br />

after the extraction of the oil, is used as a cattle fodder, or for paper making, or as a fuel in the next<br />

round of distillation.<br />

These oils are a source of precursors for the production of vitamin A and potentially for other<br />

compounds; as “green” factories these plants provide alternative synthetic routes to the petrochemical<br />

ones.<br />

For those in academia, both teachers and research students, those in agriculture and other industries,<br />

and those in business, this book provides an account of the botany, taxonomy, chemistry, and<br />

biogenesis of the oils, and their extraction and analytical methods, biotechnology, storage, legislation,<br />

and trade with all the accompanying references.<br />

Professor Massimo Maffei is already a notable contributor to this series having edited a<br />

related Graminea volume (20), Genus Vetiveria, and contributed chapters to several other volumes.<br />

I thank him and all the other contributors for their kind cooperation with me and particularly for<br />

their expert data. My special thanks are due to Dr. Roland Hardman for his continuous encouragement<br />

from the very beginning of writing this book, despite his busy schedule. His help and persistent<br />

enthusiasm have been a great inspiration to me.<br />

Anand Akhila<br />

M.Sc., Ph.D. (University of London)<br />

xi


Author/Editor Page<br />

Dr. Anand Akhila was born in 1955. He has an outstanding<br />

academic career, obtaining his Ph.D. from<br />

University College London in 1980 in the area of Natu<br />

ral Product Chemistry, particularly the biosynthetic<br />

pathways occurring in nature. For the last 27 years he<br />

has been working as a senior scientist at the Central<br />

Institute of Medicinal and Aromatic Plants, a reputed<br />

national laboratory of the government of India. During<br />

the course of his long career, he has published over<br />

60 research papers, book chapters, and review articles,<br />

besides giving presentations at symposia and conferences.<br />

Anand Akhila was presented with the CSIR<br />

Young Scientist Award in 1988, a national honor in recognition<br />

of outstanding work, awarded by the Council<br />

of Scientific and Industrial Research, the premier scientific<br />

organization of the government of India. He is<br />

a Member of the Royal Society of Chemistry, United<br />

Kingdom, and many other Indian scientific societies.<br />

His current area of research has been the study of<br />

metabolic pathways of compounds of medicinal and<br />

aromatic values in plants such as Azadirachta indica, Artemisia annua, Cymbopogon species,<br />

Mentha species, and others.<br />

xiii


Contributors<br />

Anand Akhila<br />

Central Institute of Medicinal and Aromatic<br />

Plants<br />

Lucknow, India<br />

Cinzia M. Bertea<br />

Department of Plant Biology and Centre of<br />

Excellence CEBIOVEM<br />

Turin, Italy<br />

Ange Bighelli<br />

Equipe Chimie-Biomasse<br />

Université de Corse<br />

Ajaccio, France<br />

Joseph Casanova<br />

Equipe Chimie-Biomasse<br />

Université de Corse<br />

Ajaccio, France<br />

Watcharee Khunkitti<br />

Faculty of Pharmaceutical Sciences<br />

Khon Kaen University<br />

Massimo E. Maffei<br />

Department of Plant Biology and Centre of<br />

Excellence CEBIOVEM<br />

Turin, Italy<br />

Ajay K. Mathur<br />

Division of Plant Tissue Culture<br />

Central Institute of Medicinal and Aromatic<br />

Plants (CIMAP)<br />

Lucknow, India<br />

David A. Moyler<br />

Fuerst Day Laws on Ltd<br />

London, U.K.<br />

A. K. Pandey<br />

Tropical Forest Research Institute<br />

(Indian Council of Forestry Research and<br />

Education)<br />

NWFP Division, Tropical Forest Research<br />

Institute<br />

Jabalpur, India<br />

Chieko Yatagai<br />

Kurashiki University of Science and the Arts<br />

Department of Physiological Chemistry<br />

Kurashiki, Japan<br />

Hiroyuki Sumi<br />

Kurashiki University of Science and the Arts<br />

Department of Physiological Chemistry<br />

Kurashiki, Japan<br />

Rakesh Tiwari<br />

Central Institute of Medicinal and Aromatic<br />

Plants<br />

Lucknow, India<br />

xv


1<br />

contents<br />

1.1 IntroductIon<br />

The Genus Cymbopogon<br />

Botany, Including Anatomy,<br />

Physiology, Biochemistry,<br />

and Molecular Biology<br />

Cinzia M. Bertea and Massimo E. Maffei<br />

1.1 Introduction ..............................................................................................................................1<br />

1.2 Anatomy ...................................................................................................................................4<br />

1.2.1 General Considerations .................................................................................................4<br />

1.2.1.1 Leaf Anatomy ................................................................................................5<br />

1.3 Biochemistry .............................................................................................................................7<br />

1.3.1 Characterization of the Photosynthetic Variant ............................................................8<br />

1.3.2 NADP + Inhibition of NADP-MDH ...............................................................................9<br />

1.3.3 pH and Temperature Dependence of NADPH-MDH and NADP-ME .........................9<br />

1.3.4 CO 2 Assimilation and Stomatal Conductance ............................................................ 10<br />

1.4 Molecular Biology .................................................................................................................. 11<br />

1.4.1 Randomly Amplified Polymorphic DNA (RAPD) Markers ...................................... 11<br />

1.4.2 Simple Sequence Repeat Markers (SSRs) .................................................................. 13<br />

1.5 Physiology and Ecophysiology ............................................................................................... 14<br />

1.5.1 Cymbopogon martinii ................................................................................................. 14<br />

1.5.2 Cymbopogon flexuosus ............................................................................................... 16<br />

1.5.3 Cymbopogon winterianus ........................................................................................... 17<br />

1.5.4 Other Cymbopogon Species ....................................................................................... 18<br />

References ........................................................................................................................................20<br />

Among monocots forming the family of Gramineae some grasses produce essential oils that are<br />

a valuable source for the flavor industry. Two grasses are known for their industrial potential for<br />

essential oil production: Vetiveria zizanioides Stapf, which has been the subject of a monograph in<br />

the series Medicinal and Aromatic Plants—Industrial Profiles (Maffei 2002) and Cymbopogon,<br />

which is the subject of this book and which updates the monograph edited by Kumar et al. (2000).<br />

Cymbopogon is a genus comprising about 180 species, subspecies, varieties, and subvarieties. It<br />

is native to warm temperate and tropical regions of the Old World and Oceania. Table 1.1 lists the<br />

several species, subspecies, varieties, and subvarieties as reported by the International Plant Names<br />

Index (2004), published on the Internet http://www.ipni.org (accessed September 10, 2007).<br />

The name Cymbopogon was introduced by Sprengel in 1815 (Sprengel 1815) and at that time<br />

the genus consisted of a few species, which were then moved to the genus Andropogon. In fact,<br />

both Cymbopogon and Andropogon belong to the tribe Andropogoneae, a monophyletic tribe that<br />

1


2 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 1.1<br />

list of Cymbopogon species, subspecies, Varieties, and subvarieties as reported<br />

in the International Plant names Index (2004)<br />

Cymbopogon acutispathaceus De Wild C. exaltatus A. Camus.<br />

C. afronardus Stapf C. exaltatus var. ambiguus Domin.<br />

C. ambiguus A. Camus. C. exaltatus var. exaltatus (RBr) Domin.<br />

C. andongensis Rendle C. exaltatus var. genuinus Domin.<br />

C. angustispica Nakai C. exaltatus var. gracilior Domin.<br />

C. annamensis A. Camus. C. exaltatus var. lanatus (RBr) Domin.<br />

C. arabicus Nees ex Steud. C. exarmatus Stapf<br />

C. arriani Aitch. C. excavatus Stapf<br />

C. arundinaceus Schult. C. familiaris De Wild<br />

C. bagirmicus Stapf C. figarianus Chiov.<br />

C. bassacensis A. Camus. C. filipendulus Rendle<br />

C. bequaertii De Wild C. finitimus Rendle<br />

C. bhutanicus Noltie C. flexuosus Stapf<br />

C. bombycinus A. Camus. C. flexuosus var. assamensis S.C. Nath and K.K. Sarma<br />

C. bombycinus var. bombycinus (RBr) Domin. C. floccosus Stapf<br />

C. bombycinus var. townsvillensis Domin. C. foliosus Roem and Schult.<br />

C. bombycinus var. typicus Domin. C. gazensis Rendle<br />

C. bracteatus Hitchcock C. gidarba [Buch–Ham. ex Steud.] Haines<br />

C. caesius (Hook and Arn.) Stapf C. giganteus Chiov.<br />

C. caesius subsp. giganteus (Chiov) Sales C. glandulosus Spreng.<br />

C. calcicola C.E. Hubb. C. glaucus Schult.<br />

C. calciphilus Bor C. globosus Henrard.<br />

C. cambodgiensis E.G. Camus and A. Camus C. goeringii A. Camus<br />

C. chevalieri A. Camus C. goeringii var. hongkongensis S. Soenarko<br />

C. chrysargyreus Stapf C. gratus Domin.<br />

C. circinnatus Hochst. ex Hookf. C. hamatulus A Camus<br />

C. citratus Stapf C. hirtus Stapf ex Burtt Davy<br />

C. citriodorus Link C. hirtus subsp. villosum (Pignatti) Pignatti<br />

C. claessensii Robyns C. hispidus Griff.<br />

C. clandestinus Stapf C. hookeri (Munro ex Hackel) Stapf ex Bor<br />

C. coloratus Stapf C. humboldtii Spreng.<br />

C. commutatus Stapf C. iwarancusa Schult.<br />

C. commutatus var. jammuensis (Gupta) H.B. Naithani C. jinshaensis R. Zhang and C.H. Li<br />

C. condensatus Spreng. C. jwarancusa subsp. olivieri (Boiss.) S. Soenarko<br />

C. confertiflorus Stapf C. kapandensis De Wild<br />

C. connatus Chiov. C. khasianus (Hackel) Stapf ex Bor<br />

C. cyanescens Stapf C. ladakhensis B.K. Gupta<br />

C. cymbarius Rendle. C. lanatus Roberty<br />

C. densiflorus Stapf C. laniger Duthie<br />

C. dependens B.K. Simon C. lecomtei Rendle<br />

C. dieterlenii Stapf ex Phillips C. lepidus (Nees) Chiov.<br />

C. diplandrus De Wild. C. liangshanensis S.M. Phillips and S.L. Chen<br />

C. distans (Nees ex Steud.) Will Watson. C. lividus (Thwaites) Willis<br />

C. divaricatus Stapf C. luembensis De Wild<br />

C. eberhardtii A. Camus. C. mandalaiaensis Soenarko<br />

C. effusus A. Camus. C. marginatus Stapf ex Burtt Davy<br />

C. elegans Spreng. C. martinii Stapf


The Genus Cymbopogon 3<br />

table 1.1 (continued)<br />

list of Cymbopogon species, subspecies, Varieties, and subvarieties as reported<br />

in the International Plant names Index (2004)<br />

C. martinianus Schult. C. pruinosus Chiov.<br />

C. mekongensis A. Camus C. pubescens (vis) Fritsch.<br />

C. melanocarpus Spreng. C. queenslandicus S.T. Blake<br />

C. micratherus Pilg. C. quinhonensis (A. Camus) S.M. Phillips and S.L. Chen<br />

C. microstachys (Hookf.) S. Soenarko<br />

[transferred to Andropogon (Phillips and Hua 2005)]<br />

C. microthecus A. Camus C. ramnagarensis B.K. Gupta<br />

C. minor B.S. Sun and R. Zhang ex S.M. Phillips and S.L. Chen C. rectus A. Camus<br />

C. minutiflorus S. Dransf. C. reflexus Roem and Schult.<br />

C. modicus De Wild C. refractus A. Camus<br />

C. motia B.K. Gupta C. rufus Rendle<br />

C. munroi (C.B. Clarke) Noltie C. ruprechtii Rendle<br />

C. nardus (L.) Rendle C. scabrimarginatus De Wild<br />

C. nardus subvar. bombycinus (R.Br.) Roberty C. schimperi Rendle<br />

C. nardus var. confertiflorus (Steud.) Stapf ex Bor C. schoenanthus Spreng.<br />

C. nardus subvar. exaltatus (R.Br.) Roberty C. schoenanthus subsp. velutinus Cope<br />

C. nardus subvar. grandis Roberty. C. schultzii Roberty<br />

C. nardus subvar. lanatus (R.Br.) Roberty C. sennaarensis Chiov.<br />

C. nardus var. luridus (Hookf.) Gavade and M.R. Almeida C. setifer Pilg.<br />

C. nardus subvar. procerus (R.Br.) Roberty C. siamensis Bor<br />

C. nardus subvar. refractus (R.Br.) Roberty C. solutus Stapf<br />

C. nardus subvar. schultzii Roberty C. stipulatus Chiov.<br />

C. nervatus A. Camus C. stolzii Pilg.<br />

C. nyassae Pilg. C. stracheyi (Hookf.) Raizada and Jain<br />

C. obtectus S.T. Blake C. strictus Bojer.<br />

C. olivieri (Boiss.) Bor C. stypticus Fritsch.<br />

C. osmastonii R. Parker C. suaveolens Pilger<br />

C. pachnodes (Trin.) Will Watson C. subcordatifolius De Wild<br />

C. papillipes (Hochst. ex A. Rich) Chiov. C. tamba Rendle<br />

C. parkeri Stapf C. tenuis Gilli<br />

C. pendulus Stapf C. thwaitesii (Hookf.) Willis<br />

C. phoenix Rendle C. tibeticus Bor<br />

C. pilosovaginatus De Wild C. tortilis (Presl.) A. Camus<br />

C. pleiarthron Stapf C. tortilis subsp. goeringii (Steud.) TKoyama<br />

C. plicatus Stapf C. traninhensis (A. Camus) SSoenarko<br />

C. plurinodis Stapf ex Burtt Davy C. transvaalensis Stapf ex Burtt Davy<br />

C. polyneuros Stapf C. travancorensis Bor<br />

C. pospischilii (K. Schum) C.E. Hubb C. tungmaiensis L. Liu<br />

C. princeps Stapf C. umbrosus Pilg.<br />

C. procerus A. Camus C. validus Stapf ex Burtt Davy<br />

C. procerus var. genuinus Domin. C. vanderystii De Wild<br />

C. procerus var. procerus (R.Br.) Domin. C. versicolor (Nees ex Steud.) Will Watson<br />

C. procerus var. schultzii Domin. C. virgatus Stapf ex Rhind.<br />

C. prolixus (Stapf) Phillips C. virgatus Stapf ex Bor<br />

C. prostratus Sweet C. welwitschii Rendle<br />

C. proximus Stapf C. winterianus Jowitt<br />

C. proximus var. sennarensis (Hochst.) Tackholm C. xichangensis R. Zhang and B.S. Sun<br />

Source: Published on the Internet http://www.ipni.org [accessed September 10, 2007].


4 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

includes 85 genera. Mathews and coworkers (2002) found strong support for a core Andropogoneae<br />

that includes, among others, Andropogon and Cymbopogon, and support for its relationship with an<br />

expanded Saccharinae that includes Microstegium. The limited difference in the plant traits between<br />

Andropogon and Cymbopogon, has argued the possibility that species belonging to Cymbopogon<br />

might be a subgenus of Andropogon. Most of Andropogoneae have pairs of spikelets in the inflorescence,<br />

one sessile and one on a pedicel, although in some species one or the other of these spikelets<br />

appear to be suppressed. The inflorescences form is also highly variable (Mathews et al. 2002).<br />

Morphologically, the main difference in the genus Cymbopogon is the presence of some pair of<br />

spikelets, for each spike, with unisexual male flowers, whereas in the Andropogon spikelets are<br />

usually sessile and often sterile. Cymbopogon plants are tall (up to and above 1 m) perennial plants,<br />

with narrow and long leaves that are mostly characterized by the presence of silica thorns aligned<br />

on the leaf edges. Leaves bear glandular hairs, usually each with a basal cell that is wider than the<br />

distal cell (see Section 1.2). Representative of the Andropogoneae exhibit C 4 photosynthesis, with<br />

NADP-ME as the primary decarboxylating enzyme (Mathews et al. 2002), they usually have a chromosome<br />

number of five, with ploidy levels ranging from tetraploids to 24-ploid. Polyploidy, either<br />

as alloploidy or segmental alloploidy, is frequent. Representative specimens of various species of<br />

the genus Cymbopogon have been cytogenetically studied by Spies and coworkers (Spies et al.<br />

1994). The monophyly of Cymbopogon has also been clearly demonstrated, and the genus is sister<br />

to Heteropogon (Mathews et al. 2002).<br />

Among the several aromatic species belonging to the genus Cymbopogon the most important<br />

in terms of essential oil production are C. martinii, also known as palmarosa; C. citratus, better<br />

known as lemongrass; and the so-called East Indian lemongrass, Cymbopogon flexuosus, native to<br />

India, Sri Lanka, Burma, and Thailand; whereas, for the related West Indian C. citratus, a Malesian<br />

origin is generally assumed. C. nardus and C. winterianus produce the famous citronella from Sri<br />

Lanka and Java, respectively. Also known to produce essential oils are C. schoenanthus, or camel<br />

grass; C. caesius, or inchi/kachi grass; C. afronardus, C. clandestinus; C. coloratus; C. exaltatus;<br />

C. goeringii; C. giganteus; C. jwarancusa; C. polyneuros; C. procerus; C. proximus; C. rectus;<br />

C. sennaarensis; C. stipulatus; and C. virgatus (Guenther 1950b). The main constituents of<br />

Cymbopogon essential oils will be described in other chapters of the book.<br />

Many laboratories in several countries are deeply involved in studying various aspects of cymbopogons,<br />

using variously derived genetic resources. The work already done covers a wide array of<br />

topics, including botanical identification, plant description, cytogenetics, and cell, tissue, and organ<br />

in vitro cultures. Physiology and biochemistry of stress tolerance and essential oil biosynthesis,<br />

genetics and biotechnology, and agrotechnology involved in crop production and disease and pest<br />

control chemistry of terpenes, biological activities of essential oil terpenoids and trade and marketing<br />

aspects (reviewed by Kumar et al. 2000).<br />

The major objective of this monograph is to update the literature and give references on the aforementioned<br />

topics of cymbopogons, with particular attention to industrial aspects. In the following<br />

sections of this introductory chapter, we will explore the anatomy and biochemistry of the photosynthetic<br />

apparatus, also considering the most recent advances in molecular biology of Cymbopogon.<br />

We will conclude the chapter with some physiological and ecophysiological considerations.<br />

1.2 anatomy<br />

1.2.1 Ge n e r a l Co n s i d e r at i o n s<br />

Higher plants can be divided into two groups, C 3 and C 4, based on the mechanism utilized for photosynthetic<br />

carbon assimilation related to anatomical and ultrastructural features. A cross section<br />

of a typical C 3 leaf reveals essentially one type of photosynthetic, chloroplast-containing cell, the<br />

mesophyll, and in these plants atmospheric CO 2 is fixed directly by the primary carbon-fixation<br />

enzyme ribulose 1,5-bisphosphate carboxylase/oxygenase (Rubisco). In contrast, a typical C 4 leaf


The Genus Cymbopogon 5<br />

has two distinct chloroplast-containing cell types, the mesophyll and the bundle sheath (or Kranz)<br />

cells, and they differ in photosynthetic activities (Hatch 1992; Maurino et al. 1997). The operation<br />

of the C 4 photosynthetic mechanism requires the cooperative effort of both cell types, connected by<br />

an extensive network of plasmodesmata that provides a pathway for the flow of metabolites between<br />

the cells.<br />

The C 4 pathway is a complex adaptation of the C 3 pathway that overcomes the limitation of photorespiration<br />

and is found in a diverse collection of species, many of which grow in hot climates. It<br />

was first discovered in tropical grasses (e.g., sugarcane and maize) and is now known to occur in<br />

16 plant families. It occurs in both monocotyledonous and dicotyledonous plants, and is particularly<br />

prominent in species of the Gramineae, Chenopodiaceae, and Cyperaceae (Edwards and Walker<br />

1983). About half of the species of the Poaceae are included among the C 4 plants (Smith and Brown<br />

1973). The key feature of C 4 photosynthesis is the compartmentalization of activities into two specialized<br />

cell and chloroplast types. Rubisco and C 3 photosynthetic carbon reduction (PCR) cycle<br />

are found in the inner ring of bundle sheath cells. These cells are separated from the mesophyll<br />

and from the air in the intercellular spaces by a lamella that is highly resistant to the diffusion of<br />

CO 2 (Hatch 1988). Thus, by virtue of this two-stage CO 2 fixation pathway, the mesophyll-located<br />

C 4 cycle acts as a biochemical pump to increase the concentration of CO 2 in the bundle sheath an<br />

estimated 10-fold over atmospheric concentrations. The net result is that the oxygenase activity of<br />

Rubisco is effectively suppressed, and the PCR cycle operates more efficiently.<br />

C 4 plants have two chloroplast types, each found in a specialized cell type. Leaves of C 4 plants<br />

show extensive vascularization, with a ring of bundle sheath (BS) cells surrounding each vein and<br />

an outer ring of mesophyll (M) cells surrounding the bundle sheath. CO 2 fixation in these plants is<br />

a two-step process.<br />

There are three variants on the basic C 4 pathway, and the biochemical distinctions are correlated<br />

with the ultrastructural differences of Kranz cells (Gutierrez et al. 1974; Hatch et al. 2007; Hatch<br />

et al. 1975). The three C 4 variants can be distinguished ultrastructurally by using combinations<br />

of two characters of bundle sheath cell chloroplasts, by the degree of granal stacking, and by the<br />

chloro plasts position (Gutierrez et al. 1974).<br />

For this reason, comparative grass leaf anatomy has become the object of intensive investigation<br />

in relation to photosynthesis along with biochemical studies.<br />

1.2.1.1 leaf anatomy<br />

First descriptions of the genus Cymbopogon were given by Breakwell (1914) on C. bombycinus<br />

and C. refractus under the name of Andropogon bombycinus R. Br. and A. refractus, respectively,<br />

followed by a leaf structure description done by Vickery (1935) and Prat (1937). Further studies were<br />

con ducted by Metcalfe (1960). The descriptions given by these authors are very similar and still valid.<br />

1.2.1.1.1 Generic Characters<br />

Both adaxial and abaxial epidermises of Cymbopogon species contain short-cells, over the veins,<br />

solitary, paired or in short or long rows, the proportion of each type varying with the species. Silica<br />

bodies are located over the veins, mostly crossed to dumbell shaped. Microhairs are present usually<br />

each with the basal cell wider than the distal cell, the latter frequently tapering to a pointed apex,<br />

or hemispherical. Stomata with subsidiary cells range from low or tall dome-shaped to triangular,<br />

the proportions of each type varying in different species and sometimes in separate preparations<br />

from a single species. The vascular bundles are small, mostly angular, but less conspicuous in some<br />

species than in others. The mesophyll presents a distinctly radial chlorenchyma. Bundle sheaths are<br />

single (Metcalfe 1960).<br />

Figure 1.1A shows an electron scanning micrograph of a C. citratus leaf blade. It is possible to<br />

observe developed prickle hairs with rather elongated bases over the veins. The abaxial epidermis<br />

also reveals several stomata, with narrow guard cells associated with subsidiary cells, typical of<br />

grasses (Figure 1.1B).


6 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

A<br />

C<br />

E<br />

G<br />

FIgure 1.1 (A) Electron scanning micrograph of C. citratus leaf blade. Prickle hairs are evident on the<br />

veins (white arrows) (80×). (B) Electron scanning micrograph of a stoma seen from the surface. Guard cells<br />

are narrow in the middle and enlarged at the end, while subsidiary cells are triangular (2000×). (C) Semi-thin<br />

cross section of C. citratus leaf stained with toluidine blue. The vascular bundle in a minor vein is surrounded<br />

by a layer of sheath cells, with chloroplasts arranged in a centrifugal position. (D) Bundle sheath chloroplast<br />

without grana and with few starch grains. (E) Mesophyll chloroplast with most of the thylakoid stacked in<br />

grana and without starch grains. (F) Bundle sheath chloroplast showing immunolabeling against Rubisco.<br />

Colloidal gold particles (white arrows in the high magnification) strongly label the stroma. (G) Enlargement<br />

of plasmodesmata connecting mesophyll and bundle sheath cells, providing metabolite flow between the two<br />

photosynthetic tissues.<br />

B<br />

D<br />

F


The Genus Cymbopogon 7<br />

1.2.1.1.1.1 Leaf Ultrastructure of Cymbopogon citratus The structure of parenchymatic bundle<br />

sheath (BS) cells is particularly important in distinguishing C 3 and C 4 species. A commonly mentioned<br />

anatomical feature of C 4 plants is the orderly arrangement of mesophyll cells with reference<br />

to the BS, the two together forming concentric layers around the vascular bundle. The anatomical<br />

differences between plants exhibiting a C 4 photosynthetic carbon assimilation pathway can be<br />

disclosed by electron microscopical observations of the BS (Chapman and Hatch 1983; Edwards<br />

and Walker 1983; Gutierrez et al. 1974; Hatch 1988; Hatch et al. 1975; Jenkins et al. 1989). In the<br />

NADP-ME type, chloroplasts are peripherally arranged and grana are deficient or absent in bundle<br />

sheath cells. Two other distinctive features are the presence or absence of a mestome sheath, a layer of<br />

cells intervening between metaxylem vessel elements and laterally adjacent BS Kranz cells, and the<br />

presence or absence of a cell wall suberized lamella (SL). The mestome sheath occurs in NAD-ME<br />

and PCK species, while the SL is present in bundle sheath cell walls of several NADP-ME and PCK<br />

species (Eastman et al. 1988; Hattersley and Watson 1976). The number of mitochondria in the<br />

NADP-ME subtype is lower than in the NAD-ME one because, in the latter, the enzymes involved<br />

in the transformation of aspartate to CO 2 and pyruvate are present in these organelles (Hatch et al.<br />

1975). Anatomical studies conducted on C. citratus leaves indicated the presence of the C 4 Kranz<br />

anatomy in this plant along with several ultrastructural features typical of NADP-ME species.<br />

In C. citratus cross sections of minor veins, the vascular bundle appears surrounded by one layer<br />

of sheath cells, in which the chloroplasts are located in a centrifugal position (Figure 1.1C). No<br />

mestome sheath between metaxylem vessel elements and laterally adjacent Kranz cells is observed.<br />

In the bundle sheath, ultrastructural analyses show a suberized cell wall lamella, and the presence<br />

of agranal chloroplasts, containing numerous starch grains (Figure 1.1D). Figure 1.1E shows mesophyll<br />

chloroplasts with most of the thylakoids stacked in grana. The two photosynthetic tissues are<br />

connected by a certain number of plasmodesmata that provide a pathway for the flow of metabolites<br />

between the mesophyll and bundle sheath cells (Figure 1.1G).<br />

These observations are in accordance with previous anatomical descriptions related to the genus<br />

Cymbopogon reported by Rajendrudu and Das (1981). These authors reported on the leaf anatomy<br />

and photosynthetic carbon assimilation in five species of Cymbopogon (C. flexuosus, C. martinii var.<br />

motia, C. nardus, C. pendulus, and C. winterianus) a Kranz-type leaf anatomy with a centrifugal<br />

position of starch-containing chloroplasts in the bundle sheath cells. Starch was exclusively localized<br />

in the bundle sheath cells that were typically elongated parallel to the veins and nearly twice<br />

as long as wide in the species of Cymbopogon. A narrow leaf interveinal distance was a common<br />

feature among the five Cymbopogon species. A xylem-mestome sheath of cells between metaxylem<br />

vessels and laterally adjacent bundle sheath cells of primary vascular bundles was totally absent in<br />

the five Cymbopogon species (Rajendrudu and Das 1981).<br />

1.2.1.1.2 Rubisco Immunolocalization<br />

High-resolution immunolocalization of Rubisco by electron microscopy showed that labeling<br />

occurred only in the bundle sheath chloroplasts of C. citratus. For these experiments, purified rabbit<br />

polyclonal antibodies raised against Rubisco were employed. Bound antibodies were then visualized<br />

by linking conjugated gold-labeled goat antirabbit polyclonal antibodies. Gold particles<br />

appeared to be uniformly distributed throughout the stroma (Figure 1.1F). These studies conducted<br />

on C. citratus leaves provide evidence for the localization of Rubisco in the stroma of bundle sheath<br />

chloroplasts, as expected for a C 4 plant (Bertea et al. 2003).<br />

1.3 bIochemIstry<br />

In a preliminary physiological study conducted by (Maffei et al. 1988) on C. citratus grown in<br />

humid temperate climates, some of the PEP-carboxylase kinetic characteristics, and Rubisco and<br />

glycolate oxidase activities were found to be comparable to those of C 4 plants.


8 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

The C 4 mechanism was also confirmed by the 13 C/ 12 C stable isotope ratio analyses (δ 13 C =<br />

−13.0). These results are in accordance with δ 13 C values measured on other species of Cymbopogon<br />

(Rajendrudu and Das 1981) in which δ 13 C value of −11.0 for C. flexuosus, −9.7 for C. martinii, −11.6<br />

for C. nardus, −10.3 for C. pendulus, and −11.3 for C. winterianus, respectively, were recorded.<br />

From a biochemical point of view, the three types of the basic C 4 pathway differ mainly in the<br />

C 4 acid transported into the bundle sheath cells (malate and aspartate) and in the way in which<br />

it is decarboxylated; they are named (based on the enzymes that catalyse their decarboxylation)<br />

NADP-dependent malic enzyme (NADP-ME) found in the chloroplasts, NAD-dependent malic<br />

enzyme (NAD-ME) found in mitochondria, and phosphoenolpyruvate (PEP) carboxykinase (PCK),<br />

found the cytosol of the bundle sheath cells (Edwards and Walker 1983; Ghannoum et al. 2001;<br />

Hatch et al. 1975; Jenkins et al. 1989; Huang et al. 2001). Furthermore, a characteristic leaf anatomy,<br />

biochemistry, and physiology are associated with each of the C 4 types (Dengler and Nelson 1999;<br />

Hattersley and Watson 1976). A clear indication of the C 4 photosynthetic pathway of C. citratus and<br />

the variant to which it belongs was obtained by estimating the activities of NADP-ME (EC 1.1.1.40),<br />

NADP-MDH (EC 1.1.1.82), PPDK (EC 2.7.9.1), NAD-ME (EC 1.1.1.39), and PCK (EC 4.1.1.49) as<br />

well as some kinetic characteristics of NADP-ME and NADP-MDH. Adaptation to a particular<br />

environment is a complex process involving a number of physiological, morphological, and ecological<br />

factors (Ghannoum et al. 2001; Huang et al. 2001).<br />

Therefore, enzyme activities were recorded at the low and high temperatures typical of humidtemperate<br />

climates, in order to evaluate the adaptability of C. citratus. In order to estimate increases<br />

or decreases in the reaction rate due to changes in the protonation state, groups involved in the catalysis<br />

and/or binding of substrates as a consequence of pH fluctuations, activities were also recorded<br />

at different pH values.<br />

1.3.1 Ch a r a C t e r i z at i o n o f t h e Ph o t o s y n t h e t i C Va r i a n t<br />

Further studies dealing with the characterization of the C 4 variant indicated an NADP-dependent<br />

malic enzyme photosynthetic pathway in C. citratus.<br />

The biochemical subtype was established through the estimation of the highest activities of<br />

NADP-dependent malic enzyme (NADP-ME, EC 1.1.1.40), NADP-dependent malate dehydrogenase<br />

(NADP-MDH, E.C. 1.1.1.82), pyruvate, orthophosphate dikinase (PPDK, E.C. 2.7.9.1), NADdependent<br />

malic enzyme (NAD-ME, E.C. 1.1.1.39), and phosphoenolpyruvate carboxykinase (PCK,<br />

E.C. 4.1.1.49) and some kinetic, along with some chemical-physical parameters of NADP-ME and<br />

NADP-MDH.<br />

Extraction and partial purification sequentially involved precipitation with crystalline ammonium<br />

sulfate, dialysis, and anion exchange (DEAE-Sephacell). Both, extraction and assays were<br />

conducted according to Ashton (1990). The low activity values of PPDK (90.28 nKat mg −1 prot),<br />

PCK (


The Genus Cymbopogon 9<br />

Km values obtained for OAA and NADPH (NADP-MDH) were 29.0 (±0.014) mM and 31,67<br />

(±0.09) mM, respectively. With regard to NADP-ME, the apparent Km values for NADP + and malate<br />

were 19.40 (±0.08) and 242.0 (±0.008) mM, respectively. In the case of NADP-MDH, Vmax values<br />

for OAA and NADPH were 12.52 (±0.021) and 14.97 (±0.012) mKat mg −1 prot, respectively.<br />

NADP-ME Vmax values for malate and NADP + were 8.63 (±0.507) and 18.60 (±0.007) mKat<br />

mg −1 prot, respectively. In general, relatively high activities of NADP-MDH and NADP-ME allowed<br />

a partial characterization of these enzymes and provided evidence for an NADP-ME subtype for<br />

C. citratus. The apparent kinetic properties of both enzymes were comparable to those of plants<br />

belonging to this subtype (Ashton 1990; Hatch et al. 2007; Hatch et al. 1975), and were consistent<br />

with a high photosynthetic activity, even when the plant was cultivated in a temperate climate.<br />

1.3.2 nadP + in h i b i t i o n o f nadP-Mdh<br />

Inhibition studies were carried out by measuring the NADP-MDH-catalyzed reaction at varying<br />

concentrations of NADP + and constant concentrations of OAA (1.0 mM) and NADPH (0.2 mM).<br />

NADP-MDH activity was increasingly inhibited by increasing NADP + concentrations. The activity<br />

value recorded in the presence of 0.25 mM NADP + was only 38% of the activity measured in<br />

absence of the oxidized coenzyme.<br />

In Zea mays, activation of NADP-MDH is regulated by oxidation and reduction of cysteine<br />

residues (thioredoxin-mediated system) (Lunn et al. 1995), and interconversion of the reduced and<br />

oxidized forms is influenced by the NADPH/NADP + ratio (Trevanion et al. 1997). A high NADPH/<br />

NADP + ratio leads to a more active enzyme; thus, high rates of OAA reduction only occur in reduced<br />

conditions. The percentage of inhibition caused increasing NADP + concentration in our DEAE<br />

preparations was in accordance with the observations reported earlier. The relatively high activities<br />

of NADP-ME detected enable us to characterize the enzyme in C. citratus. A low Km value was<br />

calculated for free NADP + . These results confirm the high affinity of NADP + for its binding site in<br />

all isoforms of this enzyme (Rothermel and Nelson 1989). A higher Km value was calculated for<br />

malate in accordance with literature data.<br />

1.3.3 Ph a n d te M P e r at u r e dePendenCe o f nadPh-Mdh a n d nadP-Me<br />

pH studies were carried out by using a buffer system that contained an equimolar mixture of buffers<br />

adjusted to different pH values with KOH (Bertea et al. 2001).<br />

Assays were performed at different pH values, 6.0 to 10.5 for NADP-MDH, and 6.0 to 10 for<br />

NADP-ME, using the DEAE-preparation. Maximal activity of NADP-MDH enzyme was observed<br />

at pH 8.3, in agreement with the published data (Ashton 1990). An increase in activity was<br />

observed starting from the lowest pH value (6.0) up to pH 8.3. At pH 7.0−7.5, the activity was comparable<br />

to that at pH values ranging between 9.5 and 10.0. At pH 10.5 the activity was comparable<br />

to that recorded at pH 6.0. Temperature changes also affected the reaction rate. The influence of<br />

temperature on enzyme activities was determined for both NADP-MDH and NADP-ME by adding<br />

the substrates to standard assay mixtures equilibrated at the appropriate temperatures. The assays<br />

were performed by using the DEAE-preparation. Apparent activation energy was calculated from<br />

Arrhenius plots.<br />

When enzyme activity was measured using the standard assay system at temperatures ranging<br />

from 20°C to 49°C, maximal activity was detected at 35°C, while at 49°C activity was lower, but<br />

still much higher than that at 20°C, in accordance with the typical behavior of C 4 photosynthetic<br />

enzymes. From a linear Arrhenius plot of the data, in a temperature range from 20°C to 38°C, the<br />

activation energy of the reaction was calculated to be 6970.8 cal mol –1 . The highest NADP-ME<br />

activity was recorded at pH 8.3, in accordance with the enzyme characteristics (Edwards and


10 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Andreo 1992), whereas at pH 10.0 the activity was higher than the activity recorded at pH 6.0 and<br />

6.5. With regard to temperature, maximal activity was measured at 45°C, the lowest at 20°C. Also<br />

in this case, activity response to temperature changes was typical of C 4 plants (Edwards and Andreo<br />

1992). The activation energy of the reaction, calculated in a temperature range from 20°C to 45°C,<br />

was 7605.2 cal mol –1 .<br />

Climatic conditions exert an evident effect on the physiological status of photosynthetic enzymes,<br />

and variations in light, temperature, moisture, etc., may influence the cytosolic and stromal pH<br />

(Ashton 1990). In C 4 plants, NADP-MDH has subunits of 42 kDa, and the native enzyme apparently<br />

occurs as either a tetramer or a dimer. The tetramer is the more active form; it is stable at alkaline<br />

pH values and at high temperatures (Ashton 1990). NADP-ME is a tetramer with a molecular weight<br />

of about 280 kDa, and it is more stable at pH values above 8.0 (Edwards and Andreo 1992).<br />

This enzyme exists as a dimer and a monomer, both of which are active. Differences in pH can<br />

dramatically alter the activities of these photosynthetic enzymes. Temperature is another critical<br />

parameter. When it is low, photosynthetic rates of C 4 plants may fall below those of C 3 ones. The<br />

response of enzyme activities to such changes depends on the photosynthetic pathway adopted,<br />

resulting in a different optimum range of temperatures over which the highest growth rate can be<br />

maintained (Fitter and Hay 1987).<br />

Because they originated in tropical and subtropical areas, the optimum temperature for photosynthesis<br />

in C 4 plants is 30°C–40°C, which is approximately 10°C higher than in C 3 plants (Leegood<br />

1993). However, C 4 photosynthesis is usually sensitive to low temperature; the minimum temperature<br />

for photosynthesis in several C 4 tropical grasses is 5°C–10°C (Casati et al. 1997). Activities at<br />

different temperatures and pH values of C. citratus NADP-MDH and NADP-ME indicated that this<br />

species is a C 4 NADP-ME plant, which is able to retain its photosynthetic mechanism even when<br />

cultivated in temperate climates.<br />

1.3.4 Co 2 as s iM i l a t i o n a n d st o M ata l Co n d u C ta n C e<br />

A very low compensation point (between 8 and 15 ppm CO 2) was calculated for C. citratus. This<br />

result is typical for a C 4 plant. Stomatal opening increased in response to CO 2 concentration up to<br />

157 ppm. However, at higher CO 2 values a decrease was recorded, thus indicating a clear effect<br />

of the CO 2-concentrating mechanism present in C 4 plants (data not shown). In order to evaluate<br />

changes in photosynthesis as a function of leaf age, CO 2 assimilation and stomatal conductance<br />

were also measured at different developmental stages of C. citratus leaves. A general increase for<br />

both parameters was observed, starting from primordial up to mature leaves, while a decrease in<br />

CO 2 assimilation and stomatal conductance was recorded in old leaves. Thus, primordial leaves<br />

presented the lowest CO 2 assimilation value (9.01 mmol CO 2 dm −2 s −1 ), while the highest values<br />

were recorded in young and mature leaves, without appreciable differences (22.71 and 23.94 mmol<br />

CO 2 dm −2 s −1 , respectively). With regard to stomatal conductance, the lowest value was measured in<br />

old leaves (55.00 mM H 2O dm −2 s −1 ), while the highest occurred in mature ones (165.27 mM H 2O<br />

dm −2 s −1 ).<br />

From a physiological point of view, the remarkable differences between the photosynthetic<br />

responses of C 3 and C 4 plants to CO 2 concentration become apparent when calculating the CO 2<br />

compensation point. In plants with CO 2-concentrating mechanisms, including C 4 plants, CO 2<br />

concentrations at the carboxylation sites are often saturating. Plants with C 4 metabolism have<br />

a CO 2 compensation point of or close to zero, reflecting their very low levels of photorespiration.<br />

The results obtained in C. citratus are in accordance with the values previously recorded on other<br />

species of Cymbopogon (Rajendrudu and Das 1981). In addition, the C 4 mechanism allows the plant<br />

to maintain high photosynthetic rates at lower partial CO 2 pressures in the intercellular spaces of<br />

the leaf, which require lower rates of stomatal conductance for a given rate of photosynthesis. For<br />

these reasons, measuring the CO 2 compensation point and stomatal conductance can be useful to<br />

distinguish between C 3 and C 4 pathways.


The Genus Cymbopogon 11<br />

1.4 molecular bIology<br />

The morphological variation and oil characteristics of various species and varieties of Cymbopogon<br />

have been reported, but such information is not sufficient to precisely define the relatedness among<br />

the morphotypes and chemotypes. For instance, C. martinii var. sofia and C. martinii var. motia are<br />

morphologically almost indistinguishable, but show distinct chemotypic characteristics in terms of<br />

oil constituents (Guenther 1950a). Conversely, phenotypically and taxonomically well distinguishable<br />

species produce oils of almost identical chemical compositions, such as lemongrass oils from<br />

C. citratus and C. flexuosus (Khanuja et al. 2005). Such phenotypic traits, whether morphological<br />

or chemotypic, are basically the phenotypic expression of the genotype, while DNA markers are<br />

independent of environment, age, and tissue, and expected to reveal the genetic variation more<br />

conclusively in assessing such variations. Introgression of various traits, intermittent mutations,<br />

and selection through human intervention may lead to variation in chemotypic characters across<br />

geographical distributions (Kuriakose 1995). While natural hybridization may lead to the formation<br />

of morphological or chemotypic intermediates, defining taxa purely on this basis may not be appropriate.<br />

Molecular markers provide extensive polymorphism at DNA level used for differentiating<br />

closely related genotypes (Pecchioni et al. 1996) and also to find out the extent of genetic diversity<br />

(Jain et al. 2003).<br />

Different types of molecular markers have been developed and used in various plant species<br />

including grasses in the recent past years.<br />

1.4.1 ra n d o M ly aMPlified Po ly M o r P h iC dna (raPd) Ma r k e r s<br />

DNA-based markers such as randomly amplified polymorphic DNA (RAPD) (Welsh and<br />

McClelland 1990) have been employed not only for cultivar identification but also for phylogenetic<br />

and pedigree studies in a number of food, forage, and fiber crops (Chalmers et al. 1992; Kresovich<br />

et al. 1994). RAPDs have provided rare biotype specific markers in medicinal and aromatic plants<br />

such as vetiver (Adams and Dafforn 1997) and Artemisia annua (Sangwan et al. 1999).<br />

Randomly primed polymerase chain reaction provides a simple and fast approach to detecting<br />

DNA polymorphism, with allelic RAPD marker variations being detected as a plus or minus allele<br />

(Welsh and McClelland 1990). In particular, the approach provides multilocus profiling of DNA<br />

sequence differences of genotypes when genetic knowledge is lacking. Several studies have been<br />

carried out on Cymbopogon species by employing the RAPD approach.<br />

A study using RAPD markers was carried out by Shasany and coworkers (2000) to trace the<br />

ancestors of cultivar Java II within C. winterianus. The species C. winterianus Jowitt is believed to<br />

have originated from the well-known species C. nardus, type Maha Pengiri, referred to as Ceylonese<br />

(Sri Lankan) commercial citronella. It was introduced into Indonesia and became commercially<br />

known as the Javanese citronella. The Javanese type C. winterianus material was introduced into<br />

India for the commercial cultivation of this crop during 1959. Varieties of this species have been<br />

developed later by the use of breeding procedures from the same introduced material. The authors<br />

carried out a comparative analysis of the morphological characters, chemical traits (oil percentage<br />

and constituents), and RAPD profiles to assess the diversity and relationships among the Java<br />

citronella cultivated forms, which were systematically developed for their suitability to different<br />

climatic regions, and also their differences and similarities to the believed parent species C. nardus<br />

(Purseglove 1975). All these accessions were analyzed at the molecular level for the similarity and<br />

genetic distances through RAPD profiling, using 20 random primers. More than 50% divergence<br />

was observed for all the C. winterianus accessions in relation to C. nardus accession CN2. The clustering<br />

based on the similarity matrices showed a major cluster of six accessions, consisting of two<br />

subclusters. The accession C. nardus CN2 got carved out along with two C. winterianus accessions,<br />

CW2 and CW6. On the other hand, the accessions CW2 and CW6 demonstrated distinct identities<br />

compared to CN2 at the DNA level (Shasany et al. 2000).


12 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

The same approach was used by Sangwan et al. (2001) on eleven elite and popular Indian cultivars<br />

of Cymbopogon aromatic grasses of essential oil trade types—citronella, palmarosa, and<br />

lemongrass. They were characterized by means of RAPDs to discern the extent of diversity at the<br />

DNA level between and within the oil biotypes. Primary allelic variability and the genetic bases of<br />

the cultivated germplasm were computed through parameters of gene diversity, expected heterozygosity,<br />

allele number per locus, SENA, and Shannon’s information indices. The allelic diversity<br />

was found to be in this order: lemongrass > palmarosa > citronella. Lemongrasses displayed higher<br />

(1.89) allelic variability per locus than palmarosa (1.63) and citronella (1.40). Also, RAPDs of diagnostic<br />

and curatorial importance were discerned as “stand-along” molecular descriptors. Principal<br />

component analysis (PCA) resolved the cultivars into four clusters: one each of citronella and palmarosa,<br />

and two of lemongrasses (one of C. flexuosus and another of C. pendulus and its hybrid<br />

with C. khasianus). Proximity of the two species-groups of lemongrasses was also revealed as they<br />

shared the same dimension in the three-dimensional PCA (Sangwan et al. 2001).<br />

The same authors analyzed the elite and popular cultivars of C. martinii for genomic and<br />

expressed molecular diversity using RAPD, enzyme, and SDS-PAGE protein polymorphisms. The<br />

allelic score at each locus of the enzymes, as well as presence and absence profiling in RAPDs,<br />

and overall occurrence of band types were subjected to computation of gene diversity, expected<br />

heterozygosity, allele number per locus, and similarity matrix. These, in turn, provide inputs to<br />

derive primary account of allelic variability, genetic bases of the cultivated germplasm, putative<br />

need for gene/trait introgression from the wild or geographically diverse habitat in elite selections.<br />

‘PRC1’ possessed the highest number of unique bands based on RAPD polymorphism. In variety<br />

‘IW31245E,’ diaphorase and glutamate oxaloacetate transaminase isozymes generated two unique<br />

bands as dia-III2 and got-II4. ‘RRL(B)77’ exhibited three unique bands; one produced by esterase<br />

as allele est-II1 and two by malic enzyme (me-III1,3). Only one unique band was generated by<br />

malic enzyme in variety ‘Trishna.’ But sofia had three unique bands, two contributed by diaphorase<br />

(dia-II3 and dia-II4) and one by glutamate oxaloacetate transaminase (got-II2). SDS-PAGE analysis<br />

revealed the presence of unique polypeptide fragments (97.7 to 31.6 kDa) in varieties ‘IW31245E,’<br />

‘RRL(B)77,’ ‘Tripta,’ ‘Trishna,’ ‘PRC1,’ and sofia, generated as a diagnostic marker. In general,<br />

molecular distinctions associated with varieties. motia and sofia were clearly noticed in C. martinii<br />

(Sangwan et al. 2003).<br />

Khanuja et al. (2005) analyzed 19 Cymbopogon taxa belonging to 11 species, 2 varieties, 1<br />

hybrid taxon, and 4 unidentified species for their essential oil constituents and RAPD profiles to<br />

determine the extent of genetic similarity and thereby the phylogenetic relationships among them.<br />

Remarkable variation was observed in the essential oil yield ranging from 0.3% in Cymbopogon<br />

travancorensis Bor to 1.2% in Cymbopogon martinii (Roxb.) Wats. var. motia. Citral, a major essential<br />

oil constituent, was employed as the base marker for chemotypic clustering. Based on genetic<br />

analysis, elevation of Cymbopogon flexuosus var. microstachys (Hook. F.) Soenarko to species status<br />

and separate species status for C. travancorensis Bor, which has been merged under C. flexuosus<br />

(Steud.) Wats., were suggested toward resolving some of the taxonomic complexes in Cymbopogon.<br />

The separate species status for the earlier proposed varieties of C. martinii (motia and sofia) is further<br />

substantiated by these analyses. The unidentified species of Cymbopogon have been observed<br />

as inter mediate forms in the development of new taxa (Khanuja et al. 2005).<br />

Somaclonal variants that arise through the tissue culture have been reported in a large number of<br />

species. The significance of somaclonal variation in crop improvement depends upon establishing a<br />

genetic basis for variation (Larkin and Scowcroft 1981). The use of the molecular marker is becoming<br />

widespread for the identification of somaclonal variant. In particular, RAPD markers have<br />

proved useful for this purpose owing to its ability to analyze DNA variation at many loci using small<br />

amounts of tissue (Munthali et al. 1996; Wallner et al. 1996). Screening of somaclonal variants with<br />

improved oil yield and quality have been reported in two species of Cymbopogon, C. winterianus<br />

and C. martinii (Mathur et al. 1988; Patnaik et al. 1999), but RAPDs were not used to establish


The Genus Cymbopogon 13<br />

the genetic basis of this somaclonal variation. The paper of Nayak and coworkers (2003) reports the<br />

quantitative and qualitative analysis of selected somaclones of jamrosa (a hybrid Cymbopogon) in<br />

the field, screening and selection of agronomically useful somaclonal variants with high oil yield<br />

and desirable quality, and detection of gross genetic changes through RAPD analysis.<br />

In this study, the high oil yield somaclonal variants SC1 and SC2 were subjected to RAPD analysis<br />

and the result was compared with RAPD profile of the control. A total of 22 arbitrary primers<br />

were utilized for initial screening for their amplifying ability. Of these, 12 primers successfully<br />

amplified jamrosa DNA with reproducible banding pattern. In general, 2-11 amplified fragments<br />

were scored, depending upon primers, ranging in molecular sizes from 266 bp to 1.9 Kb. The test<br />

samples SC1, SC2, and the control could be suitably distinguished by the presence of specific markers<br />

or by their absence. Out of the two somaclones analyzed, relatively less distinctness in the amplified<br />

DNA of SC2 was detected using the primers tested. Banding pattern of this somaclone SC2<br />

and the control plant was similar whereas in other variants, somaclone (SC1) DNA polymorphism<br />

was observed by having distinct banding pattern. As indicated by RAPDs gross genetic changes<br />

have occurred in somaclone (SC1). The results obtained by Nayak and coworkers are in agreement<br />

with detection of somaclonal variants by RAPD analysis in Populus deltoides (Rani et al. 1995),<br />

garlic (Al-Zahim et al. 1999) and in rice (Yang et al. 1999). Taylor et al. (1995) also reported that<br />

RAPD analysis proved suitable for detecting gross genetic changes occurring in sugarcane tissues<br />

subjected to prolonged in vitro culture. This work has demonstrated the scope of selecting improved<br />

clones of jamrosa with high oil yield and quality through somaclonal variation and suitability of<br />

RAPDs for detecting gross genetic changes in somaclonal variants at DNA level.<br />

1.4.2 siM P l e se q u e n C e re P e at Ma r k e r s (ssrs)<br />

Kumar and coworkers (2007) developed a set of simple sequence repeat markers from a genomic<br />

library of Cymbopogon jwarancusa to help in the precise identification of the species (including<br />

accessions) of Cymbopogon. For this purpose, they isolated 16 simple sequence repeats<br />

containing genomic deoxyribonucleic acid clones of C. jwarancusa, which contained a total of<br />

32 simple sequence repeats with a range of 1 to 3 simple sequence repeats per clone. The majority<br />

(68.8%) of the 32 simple sequence repeats comprised dinucleotide repeat motifs followed<br />

by simple sequence repeats with trinucleotide (21.8%) and other higher-order repeat motifs.<br />

Eighteen (81.8%) of the 22 designed primers for the above simple sequence repeats amplified<br />

products of expected sizes, when tried with genomic DNA of C. jwarancusa. Thirteen (72.2%)<br />

of the 18 functional primers detected polymorphism among the three species of Cymbopogon<br />

(C. flexuosus, C. pendulus, and C. jwarancusa) and amplified a total of 95 alleles (range 1–18<br />

alleles) with a PIC value of 0.44 to 0.96 per simple sequence repeat. Thus, the higher allelic<br />

range and high level of polymorphism demonstrated by the developed simple sequence repeat<br />

markers are likely to have many applications such as in improvement of essential oil quality by<br />

authentication of Cymbopogon species and varieties, and mapping or tagging the genes controlling<br />

agronomically important traits of essential oils, which can further be utilized in marker<br />

assisted breeding (Kumar et al. 2007). Considering the high reproducibility and polymorphic<br />

nature of the SSRs, the SSRs developed by Kumar and coworkers (2007) may be utilized for<br />

identification/authentication of superior accessions/species of the genus Cymbopogon with correctness<br />

and certainty to ensure production of high-quality oil. The SSR markers developed<br />

during this kind of study might also be used to resolve the taxonomic disputes, study the genetic<br />

diversity, and for genetic mapping and QTL (quantitative trait loci) analysis. The SSRs due to<br />

their codominant nature may be specifically useful for the identification of interspecific hybrids,<br />

which have been shown to be superior in terms of both their yielding high quantity and better<br />

quality of essential oils.


14 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

1.5 PhysIology and ecoPhysIology<br />

As discussed, Cymbopogon has a photosynthetic machinery that allows the plant to perform high<br />

rates of carbon assimilation and, at the same time, save water. In species that produce essential oil,<br />

the biogenesis of terpenoids relies on photosynthetic carbon dioxide reduction on the one hand and<br />

availability of water and nutrients, on the other. For this reason several studies have been conducted<br />

in order to assess which nutrients and at what conditions were required for an optimal production<br />

of both biomass and essential oils. In this section, we will discuss the most important Cymbopogon<br />

species in terms of yield of biomass and essential oil production as related to nutrition. Furthermore,<br />

when available, references to biotechnological applications will be also reported.<br />

1.5.1 Cy m b o p o g o n martinii<br />

Water requirement, productivity, and water use efficiency of palmarosa (C. martinii) were studied<br />

under different levels of irrigation (0.1, 0.3, 0.5, 0.7, 0.9, 1.1, 1.3, and 1.5 IW:CPE ratio). Growth,<br />

herb, and essential oil yield increased significantly up to 0.5 IW:CPE ratio. At 0.5 IW:CPE ratio<br />

palmarosa produced 47.3 t ha −1 yr −1 of fresh herb and 227.3 kg ha −1 yr −1 of essential oil. Further<br />

increase in irrigation levels caused an adverse effect on growth and yield of palmarosa. Irrigation<br />

levels did not affect the quality of oil in terms of its geraniol and geranyl acetate contents. Water<br />

requirement of palmarosa was worked out to be 89.1 cm. The highest water use efficiency of 2.97 kg<br />

ha −1 cm −1 oil was recorded at 0.1 IW:CPE ratio, at 0.5 IW:CPE ratio (optimum) it was 2.55 kg ha −1<br />

cm −1 oil. Irrigation scheduled at 0.5 IW:CPE ratio gave the highest net return of Rs 51 963 ha −1<br />

yr −1 (Singh et al. 1997). In C. martinii the application of 160 kg N/ha per year produced the highest<br />

amount of biomass and essential oil, and increased the net profit and NPK uptake by the crop<br />

(Rao et al. 1988); furthermore, dressing of 40 kg K/ha enhanced the yield of biomass by 13.6% and<br />

6.5% and that of oil by 12.9% and 6.1%, compared with 20 and 80 kg K/ha, respectively (Singh<br />

et al. 1992). In the same species, harvesting the crop at early seeding (112−115 days after planting)<br />

gave 25% more herbage and 51% more oil yield over harvesting vegetative stage, while the oil so<br />

produced had higher content (90.1%) geraniol (Maheshwari et al. 1992). Highest dry-matter yield,<br />

essential oil yield, and maximum net return of palmarosa were recorded by applying Azotobacter<br />

at 2 kg/ha together with 20 kg N + 20 kg P/ha under rainfed condition in a shallow black soil<br />

(Maheshwari et al. 1998). Intercropping of blackgram−blackgram or sorghum fodder−ratoon with<br />

palmarosa gave additional yields of 660 kg/ha seed and 16.6 t/ha fodder, respectively, compared<br />

with the sole crop of palmarosa (Rao et al. 1994). Moreover, sowing of pigeon pea in alternate rows<br />

parallel to palmarosa proved most efficient and economic, as it provided higher economic returns,<br />

bonus income, and monetary advantage, and the oil content and quality in terms of total geraniols<br />

of palmarosa were not adversely affected by adoption of intercropping (Maheshwari et al. 1995).<br />

However, in palmarosa–pigeon pea intercropping systems, competition exists mainly for light rather<br />

than for nutrients and moisture, possibly because the two crop components acquire their nutrients<br />

and moisture from different soil layers (Singh et al. 1998). Concerning essential oil production of<br />

palmarosa, changes in fresh weight, dry weight, chlorophyll, and essential oil content and its major<br />

constituents, such as geraniol and geranyl acetate, were examined for both racemes and spathe at<br />

various stages of spikelet development (Dubey et al. 2000). The essential oil content was maximal<br />

at the unopened spikelets stage and decreased significantly thereafter. At unopened spikelets stage,<br />

the proportion of geranyl acetate (58.6%) in the raceme oil was relatively greater compared with<br />

geraniol (37.2%), whereas the spathe oil contained more geraniol (61.9%) compared with geranyl<br />

acetate (33.4%). The relative percentage of geranyl acetate in both the oils, however, decreased significantly<br />

with development, and this is accompanied by a corresponding increase in the percentage<br />

of geraniol. Analysis of the volatile constituents from racemes and spathes (from mature spikelets)<br />

and seeds by capillary GC indicated 28 minor constituents besides the major constituent geraniol.<br />

(E)-Nerolidol was detected for the first time in an essential oil from this species. The geraniol


The Genus Cymbopogon 15<br />

content predominated in the seed oil, whereas the geranyl acetate content was higher in the raceme<br />

oil (Dubey et al. 2000).<br />

Biotechnology is a powerful and consolidated technique for understanding plant growth and<br />

development as well as for improving biomass and yield of crops. Callus could be induced from<br />

nodal explant of mature tillering plant of C. martinii in different basal media supplemented with<br />

2,4-dichlorophenoxy acetic acid (2,4-D) and kinetin (Kin). Shoot bud was regenerated from such<br />

calli in MS and B5 basal media modified with various combinations of phytohormones, vitamins,<br />

and amino acids. Root formation was induced either in white basal medium or half-strength MS or<br />

B5 media containing naphthalene acetic acid (NAA) or indole-3-butyric acid (IBA). High survival<br />

percentage of regenerated plants in soil was obtained after acclimatization in normal environment<br />

(Baruah and Bordoloi 1991). A detailed characterization of chromosomal status was carried out in<br />

callus, somatic embryos, and regenerants derived from in vitro cultured nodal and inflorescence<br />

explants of C. martinii (2n = 20). Both the callus lines revealed considerable ploidy variations (tetraploids<br />

to octoploids and hyperoctoploids), and the degree of polyploidization increased with the<br />

culture age. Frequencies of various polyploid cells were significantly higher in nodal callus lines<br />

(3.6% to 46.3%) than the inflorescence callus lines (1.9% to 23.6%) when analyzed over 520 days<br />

of culture. Somatic embryos derived from both the callus lines retained a predominantly diploid<br />

chromosome status throughout (99.0% to 93.1%). Root tip analysis of about 70 regenerants randomly<br />

taken from cultures of various ages (days 20 to 520) revealed only diploid chromosome<br />

numbers (2n = 20) implying a strong relative stability of diploidy among the regenerants (Patnaik<br />

et al. 1996). Chromosome counts of cells in suspensions, calli, and somatic embryos derived from<br />

cultures of different ages revealed the presence of diploids, tetraploids, and octaploids (Patnaik<br />

et al. 1997). Sodium chloride tolerant callus lines of C. martinii were obtained by exposing the callus<br />

to increasing concentrations of NaCl (0–350 mM) in the MS medium. The tolerant lines grew<br />

better than the sensitive wild-type lines in all concentrations of NaCl tested up to 300 mM. Callus<br />

survival and growth were completely inhibited, resulting in tissue browning and subsequent death<br />

at 350 mM NaCl. The selected lines retained their salt tolerance after 3–4 subcultures on salt-free<br />

medium, indicating the stability of the induced salt tolerance. The growth behavior, the Na + , K + ,<br />

and proline contents of the selected callus lines were characterized and compared with those of<br />

the NaCl-sensitive lines. The Na + levels increased sharply, while the K + level declined continuously<br />

with the corresponding increase in external NaCl concentrations in both lines, but the NaCltolerant<br />

callus lines always maintained higher Na + and K + levels than that of the sensitive lines.<br />

The NaCl-selected callus line accumulated high levels of proline under salt stress. The degree of<br />

NaCl tolerance of the selected lines was in negative correlation with the K + /Na + ratio and in positive<br />

correlation with proline accumulation (Patnaik and Debata 1997a). The embryogenic potential of<br />

NaCl-tolerant callus selected even at 300 mM could be improved significantly by the incorporation<br />

of gibberellic acid (GA (3)) and abscisic acid (ABA), in the medium where, with 2 mg/L of GA (3)<br />

and 1 mg/L of ABA, the highest rates of embryogenesis (44.5%, 28.8%, and 18.6%) were achieved<br />

against 17.5%, 8.2%, and 1.8% on medium devoid of GA (3), and ABA at 50%, 150%, and 250 mM<br />

of NaCl, respectively (Patnaik and Debata 1997b). Finally, plants regenerated from cell suspension<br />

cultures of palmarosa were analyzed for somaclonal variation in five clonal generations. A wide<br />

range of variation in important quantitative traits, for example, plant yield, height, tiller number,<br />

oil content and qualitative changes in essential oil constituents geraniol, geranyl acetate, geranyl<br />

formate, and linalool, were observed among the 120 somaclones screened. Eight somaclones were<br />

selected on the basis of high herb and oil yield over the donor line and high geraniol content in the<br />

oil. Based on performance in the field trials, three superior lines were selected, and maintained for<br />

five clonal generations. The superior lines exhibited a reasonable degree of stability in the traits<br />

selected (Patnaik et al. 1999).<br />

Palmarosa was also found to be associated with a vesicular-arbuscular mycorrhizal (VAM)<br />

fungus, Glomus aggregatum. Glasshouse experiments showed that inoculation of palmarosa with<br />

G. aggregatum caused a twofold and threefold biomass production as compared to nonmycorrhizal


16 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

plants. These findings indicate the potential use of VAM-fungi for improving the production of this<br />

essential oil-bearing plant (Gupta and Janardhanan 1991). Furthermore, when the interactive effects<br />

of phosphate solubilizing bacteria, N-2 or fixing bacteria, and arbuscular mycorrhizal fungi (AMF)<br />

were studied in a low phosphate alkaline soil amended with a tricalcium insoluble source of inorganic<br />

phosphate on the growth of C. martinii. The rhizobacteria behaved as a “mycorrhiza helper”<br />

and enhanced root colonization by G. aggregatum in presence of tricalcium phosphate at the rate of<br />

200 mg kg −1 soil (P1 level) (Ratti et al. 2001).<br />

A dramatic increase in PEP carboxylase activity and oil biosynthesis was observed under<br />

drought conditions in C. martinii (Sangwan et al. 1993). The physiological and biochemical<br />

basis of drought tolerance in C. martinii has been elucidated on the basis of growth and metabolic<br />

responses (Fatima et al. 2002).<br />

1.5.2 Cy m b o p o g o n f l e x u o s u s<br />

Cymbopogon flexuosus (also known as lemongrass) is a perennial, multicut aromatic grass that<br />

yields an essential oil used in perfumery and pharmaceutical industries and vitamin A. It has a<br />

long initial lag phase. The growth and herbage and oil production of C. flexuosus in response to<br />

different levels of irrigation water (IW) 0.1, 0.3, 0.5, 0.7, 0.9, 1.1, 1.3, and 1.5 times cumulative pan<br />

evaporation CPE evaluated on deep sandy soils showed that an increment in the level of irrigation<br />

increased the plant height up to 0.7 IW:CPE ratio. However, the response of irrigation levels<br />

on tiller pro duction of lemongrass differed with the season of harvest. <strong>Oil</strong> content had an inverse<br />

relationship with the levels of irrigation, whereas significantly higher herb and essential oil yields<br />

were recorded at 0.7 IW:CPE ratio, irrespective of season of harvest (Singh et al. 2000). Application<br />

of nitrogen (0, 50, 100, and 150 kg N ha −1 yr −1 ) and phosphorus to C. flexuosus crops maintained<br />

the fertility of the soil, while potassium depletion was noticed (Singh 2001). When the effects of<br />

phosphorus (at 0, 17.75, and 35.50 kg ha −1 yr −1 ), potassium (at 0, 33.2, 66.4, and 99.6 kg ha −1 yr −1 ) and<br />

nitrogen (at 100 and 200 kg ha −1 yr −1 ) and potassium (at 0, 33.2 and 66.4 kg ha −1 yr −1 ) were studied<br />

on herbage and oil yield of C. flexuosus, it was found that plants produced significantly higher herbage<br />

and oil yields compared with controls (Singh et al. 2005; Singh and Shivaraj 1999). Spraying of<br />

iron-complexed additives on C. flexuosus increased iron translocation and the dry-matter production.<br />

Application of iron chelates and salts increased the vegetative herb yield, and oil and citral<br />

content. While maximum geraniol and less citral were obtained in the chlorotic plants, Fe recovered<br />

plants possessed more citral and less geraniol. The maximum recovery of total chlorophyll<br />

and nitrate reductase activity were recorded in the crop when Fe-EDTA chelates were sprayed at<br />

22.4 ppm (Misra and Khan 1992). In C. flexuosus, a closer plant spacing of 45 × 45 cm resulted in<br />

higher herb and oil yields compared to wider spacing of 60 × 60 cm. Application of 150 kg N ha −1<br />

yr −1 resulted in higher herb and oil yields. Higher nitrogen applications also increased the plant<br />

height and number of tillers per clump. The oil content and quality were not influenced by spacing<br />

and nitrogen levels (Singh et al. 1996b).<br />

As for C. martinii, intercropping of C. flexuosus with the food legumes such as blackgram<br />

(Vigna mungo (L) Hepper), cowpea (Vigna unguiculata (L) Walp), or soybean (Glycine max (L)<br />

Merr.) prompted extra yields over and above that of pure cultures, without affecting the oil yield<br />

(Singh and Shivaraj 1998).<br />

The influence of different foliar applications of the triacontanol (Tria.)-based plant growth regulator<br />

Miraculan on growth, CO 2 exchange, and essential oil accumulation in C. flexuosus showed<br />

increased rates in plant height, tillers per plant, biomass yield, accumulation of essential oil, net<br />

CO 2, and exchange and transpiration compared to the untreated control, but the number of leaves per<br />

tiller remained unaffected. Application of Miraculan also increased micronutrient uptake and total<br />

chlorophyll and citral content but decreased chlorophyll a/b ratio and stomatal resistance. Increase<br />

in shoot biomass, photosynthesis, and chlorophyll were significantly correlated with essential oil<br />

content (Misra and Srivastava 1991). Only young and rapidly expanding C. flexuosus leaves were


The Genus Cymbopogon 17<br />

found to have the capacity to synthesize and accumulate essential oil and citral. The pattern of the<br />

ratio of the label incorporated in citral to that in geraniol, during leaf ontogeny, evinced parallelism<br />

with the geraniol dehydrogenase activity. The elevated levels of glucose-6-phosphate dehydrogenase,<br />

6-phosphogluconate dehydrogenase, NADP + -malic enzyme, and NADP + -isocitrate dehydrogenase<br />

coincided with the period of active essential oil biogenesis accompanying early leaf growth (Singh<br />

et al. 1990). Thus, there is an active involvement of oxidative pathways in essential oil biosynthesis.<br />

The time-course (12 h light followed by 12 h dark) monitoring of the C-14 radioactivity in starch and<br />

essential oil, after exposure of the immature (15 days after emergence) leaf to (CO)-C-14, revealed<br />

a progressive loss of label from starch and a parallel increase in radioactivity in essential oil. Thus,<br />

there was indication of a possible degradation of transitory starch serving as the source of carbon<br />

precursor for essential oil (monoterpene) biogenesis in the tissue (Singh et al. 1991).<br />

Biotechnological applications revealed that C. flexuosus plants derived from somatic embryoids<br />

were more uniform in all the characteristics examined when compared with the field performance<br />

of plants raised through slips by standard propagation procedures (Nayak et al. 1996).<br />

A fungal endophyte, Balansia sclerotica (Pat.) Hohn., has been found to establish a perennial<br />

association with the commercially grown East Indian C. flexuosus cv. Kerala local (syn. = OD-19).<br />

Endophyte-infected plants produced 195% more shoot biomass and 185% more essential oil than<br />

the endophyte-free control plants when grown experimentally under glasshouse conditions. The<br />

essential oil extracted from the endophyte-infected plants is qualitatively identical with that of<br />

endophyte-free plants and is free of toxic ergot alkaloids. Thus, B. sclerotica-infected East Indian<br />

C. flexuosus has potential for agricultural exploitation (Ahmad et al. 2001).<br />

1.5.3 Cy m b o p o g o n w i n t e r i a n u s<br />

Java citronella (C. winterianus) is a perennial, multiharvest aromatic grass, the shoot biomass of<br />

which, on steam distillation, yields an essential oil extensively used in fragrance and flavor industries.<br />

Fresh C. winterianus (Java citronella) herbage and essential oil yields were significantly influenced<br />

by application of N up to 200 kg ha −1 yr −1 , while tissue N concentration and N uptake increased only<br />

to 150 kg N ha −1 . The oil yields with neem cake-coated urea (urea granules coated with neem cake)<br />

and urea super granules were 22 and 9% higher over that with prilled urea, and urea supergranules<br />

were significantly increased up to 200 kg N ha −1 while with neem cake-coated urea, response was<br />

observed only to 150 kg N ha −1 ! Estimated recovery of N during two years from neem cake-coated<br />

urea, urea supergranules, and prilled urea were 38%, 31%, and 21%, respectively (Singh and Singh<br />

1992). The interaction between N doses and nitrification inhibitors was also significant. Nitrification<br />

inhibitors performed better at the highest N dose (450 kg N ha −1 yr −1 ), and the increase in the essential<br />

oil yields was to an extent of 27.3% to 34.6% when compared with “N alone” treatment. The<br />

nitrification inhibitors also increased the apparent N recoveries by citronella considerably. The oil<br />

content in the herb and its quality were not affected by the treatments. The nitrification inhibitors<br />

increased citronella yields and improved N economy (Puttanna et al. 2001). In Java citronella significant<br />

positive correlations were observed between fresh matter, citronellol content, dry and fresh<br />

matter yields, and total essential oil content (Omisra and Srivastava 1994). When the effect of depth<br />

(25, 37.5, and 50 mm) and methods (ridge and furrow, and broad bed and furrow method) of irrigation<br />

acid nitrogen levels (0, 200, and 400 kg N ha −1 yr −1 ) were studied on herb and oil yields of Java<br />

citronella, highest herb and oil yields were achieved with the application of 400 kg N, maintaining<br />

25 mm depth of irrigation, while the content and quality of oil were not affected either by irrigation<br />

or nitrogen (Singh et al. 1996a).<br />

Among food legumes, greengram (Vigna radiata (L.) Wilez.), and among vegetables, clusterbean<br />

(Cyamopsis psoraloides D. C., syn. Cyamopsis tetragonoloba (L.) Taub.), tomato (Lycopersicon<br />

esculentum Mill.) and lady’s finger (Abelmoschus esculentus Moench.) as intercrops of C. winterianus<br />

did not decrease its biomass and essential oil yield and produced bonus yields of these<br />

crops over and above that of Java citronella. Maximum monetary returns were recorded by Java


18 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

citronella intercropped with tomato or greengram. However, Java citronella intercropped with redgram<br />

(Cajanus cajan (L.) Millsp.), horsegram (Macrotyloma uniflorum (Lam.) Verd, syn. Dolichos<br />

biflorus Roxb.), and brinjal (Solanum melongena L.) suffered significant biomass and essential oil<br />

yield reductions. Horsegram proved to be the most competitive intercrop, producing least yields and<br />

minimum monetary returns (Rao 2000).<br />

Changes in the utilization pattern of primary substrate, viz. [U-C-14] acetate, (CO 2)-C-14<br />

and [U-C-14] saccharose, and the contents of C-14 fixation products in photosynthetic metabolites<br />

(sugars, amino acids, and organic acids) were determined in Fe-deficient Java citronella in<br />

relation to the essential oil accumulation. An overall decrease in photosynthetic efficiency of the<br />

Fe-deficient plants as evidenced by lower levels of incorporation into the sugar fraction and essential<br />

oil after (CO 2)-C-14 had been supplied was observed. When acetate and saccharose were fed<br />

to the Fe-deficient plants, despite a higher incorporation of label into sugars, amino acids, and<br />

organic acids, there was a lower incorporation of these metabolites into essential oils than in control<br />

plants. Thus, the availability of precursors and the translocation to a site of synthesis/accumulation,<br />

severely affected by Fe deficiency, is equally important for the essential oil biosynthesis in citronella<br />

(Srivastava et al. 1998). Lal and coworkers (2001) observed that improvement of oil quality<br />

with high citronellal content and low elemol content in Java citronella is believed to be achievable,<br />

although some compromise will have to be made in oil yield.<br />

Nutrient acquisition and growth of Java citronella was also studied in a P-deficient sandy soil<br />

to determine the effects of mycorrhizal symbiosis and soil compaction. When a pasteurized sandy<br />

loam soil was inoculated either with rhizosphere microorganisms excluding VAM fungi (nonmycorrhizal)<br />

or with the VAM fungus, Glomus intraradices Schenck and Smith (mycorrhizal) and supplied<br />

with 0, 50, or 100 mg P kg −1 soil, G. intraradices was found to substantially increase root and<br />

shoot biomass, root length, nutrient (P, Zn, and Cu) uptake per unit root length, and nutrient concentrations<br />

in the plant, compared to inoculation with rhizosphere microorganisms when the soil was at<br />

the low bulk density and not amended with P. Little or no plant response to the VAM fungus was<br />

observed when the soil was supplied with 50 or 100 mg P kg −1 soil and/or compacted to the highest<br />

bulk density. At higher soil compaction and P supply, the VAM fungus significantly reduced root<br />

length. Nonmycorrhizal plants at higher soil compaction produced relatively thinner roots and had<br />

higher concentrations and uptake of P, Zn, and Cu than at lower soil compaction, particularly under<br />

conditions of P deficiency (Kothari and Singh 1996).<br />

Pythium aphanidermatum was the predominant fungus recovered from the roots of Java citronella<br />

showing lethal yellowing in the northern part of India. Roots of infected plants showed marked<br />

discoloration, and the cortical region was completely disintegrated and sloughed from the vascular<br />

tissue. Diseased plants were chlorotic and stunted. Rotting was often found to spread from roots to<br />

stem, leading to severe chlorosis and death of the infected plants. The pathogenicity of the fungus<br />

was established. The disease is a potential constraint to citronella cultivation in nonarid climates<br />

where the crop is irrigated extensively (Alam et al. 1992). Another disease affecting commercial<br />

plantations of Java citronella is a collar rot and wilt disease. The causal organism was identified as<br />

Fusarium moniliforme, anamorph of Gibberella fujikuroi. Isolates of the pathogen differed in their<br />

pathogenicity on the host plant under glasshouse conditions. Differences were also observed in<br />

growth rates, pigment production, and sporulation between isolates (Alam et al. 1994).<br />

1.5.4 ot h e r Cy m b o p o g o n sP e C i e s<br />

Application of graded levels of lime up to 10 t/ha on acid soil (pH 4.2) raised the pH up to 6.7. It<br />

increased the dry herbage of C. khasianus linearly. Increase of soil pH decreased N, P, K, Fe, and<br />

Zn contents in dry herbage significantly but increased the Ca and Mg contents. Liming showed a<br />

positive effect on the uptake of N, P, K, and Ca. However, Fe and Mg declined beyond lime levels<br />

of 23 and 5.0 t, respectively. Uptake of Zn was found fluctuating. <strong>Oil</strong> content (2.00%–2.07%; DWB)<br />

and geraniol (80.2%–81.0%) in the oil were unaffected by the lime treatments (Choudhury and


The Genus Cymbopogon 19<br />

Bordoloi 1992). Experiments were also conducted to measure the rate of C. caesius litter decomposition<br />

and to identify fungal flora associated with the litter during different stages of decomposition<br />

in a tropical grassland. Rate of litter decomposition was several times higher than in temperate<br />

grasslands. Buried litter decayed more rapidly, and this rate was not influenced by climatic conditions.<br />

In contrast, surface litter recorded a lower decomposition rate, which was dependent on temporal<br />

(seasonal) fluctuations. Total nitrogen, available phosphorus, and potassium contents of the<br />

stem litter decreased during the initial stages of incubation.<br />

Thirty-five species of fungi were isolated from the litter during the different stages of litter<br />

degradation. Most belonged to Hyphomycetes, which are active decomposers (Senthilkumar et al.<br />

1992). C. nardus var. confertiflorus and C. pendulus were grown under mild and moderate water<br />

stress for 45 and 90 d to investigate the impact of in situ drought stress on plants in terms of relative<br />

water content, psi, concentration of proline, activities of PEP carboxylase and geraniol dehydrogenase,<br />

and geraniol and citral biogenesis. The results revealed that the species exhibited differential<br />

responses under mild and moderate stress treatments. In general, plant growth was reduced considerably,<br />

while the level of essential oils was maintained or enhanced.<br />

Significant induction in catalytic activity of PEP carboxylase under water stress was one of the<br />

consistent metabolic responses of the aromatic grasses. The major oil constituents, geraniol and<br />

citral, increased substantially in both the species. Activity of geraniol dehydrogenase was also<br />

modulated under moisture stress. The responses varied depending upon the level and duration of<br />

moisture stress. The observations have been analyzed in terms of possible relevance of some of<br />

these responses to their drought stress adaptability/tolerance (Singhsangwan et al. 1994). In vitro<br />

plants of C. citratus were established, starting from shoot apices derived from plants cultivated<br />

under field conditions. The effect of the immersion frequency (two, four, and six immersions per<br />

day) on the production of biomass in temporary immersion systems (TIS) of 1 L capacity was studied.<br />

The highest multiplication coefficient (12.3) was obtained when six immersions per day were<br />

used. The maximum values of fresh weight (FW; 62.2 and 66.2 g) were obtained with a frequency<br />

of four and six immersions per day, respectively. However, the values for dry weight (DW; 6.4 g)<br />

and height (8.97 cm) were greater in the treatment with four immersions per day. The TIS used in<br />

this work for the production of lemongrass biomass may offer the possibility of manipulating the<br />

culture parameters, which can influence the production of biomass and the accumulation of secondary<br />

metabolites. We describe for the first time the in vitro production of Cymbopogon citratus<br />

biomass in TIS. In vitro regeneration of C. polyneuros was obtained through callus culture using<br />

leaf base, node, and root as explants. Callus was induced from different explants with 2–5 mg/L<br />

alpha-naphthalene acetic acid (NAA) and 1–2 mg/L kinetin in Murashige and Skoog’s (MS) basal<br />

medium. High frequency shoots were noticed from leaf-base callus supplemented with 3.5 mg/L<br />

6-benzylaminopurine (BA), l-arginine, adenine, and a low level of NAA (0.2 mg/L). About 80–85<br />

shoot buds were obtained from ca. 200 mg of callus per culture. The individual shoots produced<br />

root in the presence of 0.5–3 mg/L indole 3-butyric acid or its potassium salt.<br />

Regenerated plants were cytologically and phenotypically stable. Regenerants were transplanted<br />

into soil and subsequently transferred to the field (Das 1999). C. nardus could be propagated via<br />

tissue culture using axillary buds as explants. The aseptic bud explants obtained using double sterilization<br />

methods produced stunted abnormal multiple shoots when they were cultured on Murashige<br />

and Skoog (MS) medium supplemented with 1.0 mg L −1 or 2.0 mg L −1 benzyladenine (BA). Stunted<br />

shoots that cultured on MS + 1.0 mg L −1 RA + 1.0 mg L −1 N-6-isopentenyl-adenine (2iP) could<br />

induce elongation of shoots from about 60% of the stunted shoots. Normal multiple shoots could<br />

be induced at the highest (19.7 shoots per bud) from the bud explants within 6 weeks when cultured<br />

on proliferation medium consisted of MS supplemented with 0.3 mg L −1 BA and 0.1 mg L −1<br />

indole-3-butyric acid (IBA). The separated individual shoot produced roots when transferred to<br />

basic MS solid medium. The essential oils that were contained in the mature plants namely citronellal,<br />

geraniol, and citronellol, were also found in the in vitro C. nardus plantlets. Citronellal was<br />

the main essential oil component in the matured plants, while geraniol was the main component in


20 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

the in vitro plantlets (Chan et al. 2005). The occurrence, mode of infection, and the extent of damage<br />

caused by Psilocybe kashmeriensis sp. nov. Abraham on oil grass C. jawarancusa in Kashmir<br />

valley is discussed herein. A brief description of the new agaric species is also offered (Abraham<br />

1995). Dormant vegetative slips of jamrosa (C. nardus var. confertiflorus × C. jwarancusa) were<br />

subjected to various doses of gamma rays. Plants raised from them were screened with a view to isolate<br />

improved clones of the crop. Five mutant clones isolated exhibited variation in quality/ quantity<br />

of essential oil. These changes in oil characters were attributed to microlevel mutations induced by<br />

gamma rays (Kak et al. 2000).<br />

reFerences<br />

Abraham SP. 1995. Notes on the occurrence of an unusual agaric on Cymbopogon in Kashmir Valley. Nova<br />

Hedwigia 60: 227–232.<br />

Adams RP, Dafforn MR. 1997. DNA sampling of the pan-tropical vetiver grass uncovers genetic uniformity in<br />

erosion control germplasm. Diversity 13: 27, 28.<br />

Ahmad A, Alam M, Janardhanan KK. 2001. Fungal endophyte enhances biomass production and essential oil<br />

yield of East Indian lemongrass. Symbiosis 30: 275–285.<br />

Al-Zahim MA, Ford-Lloyd BV, Newbury HJ. 1999. Detection of somaclonal variation in garlic (Allium sativum<br />

L.) using RAPD and cytological analysis. Plant Cell Reports 18: 473–477.<br />

Alam M, Chourasia HK, Sattar A, Janardhanan KK. 1994. Collar rot and wilt—a new disease of Java-citronella<br />

(Cymbopogon winterianus) caused by Fusarium-moniliforme sheldon. Plant Pathology 43: 1057–1061.<br />

Alam M, Sattar A, Janardhanan KK, Husain A. 1992. Lethal yellowing of Java citronella (Cymbopogon winterianus)<br />

caused by Pythium aphanidermatum. Plant Disease 76: 1074–1076.<br />

Ashton AR. 1990. Enzymes of C 4 photosynthesis. In Dey PMHJB, Ed. Methods in Plant Biochemistry. London:<br />

Academic Press, pp. 39–72.<br />

Baruah A, Bordoloi DN. 1991. Growth and regeneration of palmarosa (Cymbopogon martinii [Roxb.] Wats.)<br />

callus tissues under varied nutritional status. Indian Journal of Experimental Biology 29: 582, 583.<br />

Bertea CM, Scannerini S, D’Agostino G, Mucciarelli M, Camusso W, Bossi S, Buffa G, Maffei M. 2001. Evidence<br />

for a C4 NADP-ME photosynthetic pathway in Vetiveria zizanioides Stapf. Plant Biosystems 135: 249–262.<br />

Bertea CM, Tesio M, D’Agostino G, Buffa G, Camusso W, Bossi S, Mucciarelli M, Scannerini S, Maffei M.<br />

2003. The C–4 biochemical pathway, and the anatomy of lemongrass (Cymbopogon citratus (DC) Stapf.)<br />

cultivated in temperate climates. Plant Biosystems 137: 175–184.<br />

Breakwell E. 1914. A study of the leaf anatomy of some native species of the genus Andropogon, N.O.<br />

Gramineae. Proceedings of the Linnean Society, N.S.W. 39: 385–394.<br />

Casati P, Spampinato CP, Andreo CS. 1997. Characteristics and physiological function of NADP-malic enzyme<br />

from wheat. Plant Cell Physiology 38: 928–934.<br />

Chalmers KJ, Waugh R, Simons AJ, Powell W. 1992. Detection of genetic variations between and within populations<br />

of Gliricidia sepium and G. maculata using RAPD markers. Heredity 69: 465–472.<br />

Chan LK, Dewi PR, Boey PL. 2005. Effect of plant growth regulators on regeneration of plantlets from bud<br />

cultures of Cymbopogon nardus L. and the detection of essential oils from the in vitro plantlets. Journal<br />

of Plant Biology 48: 142–146.<br />

Chapman KSR, Hatch MD. 1983. Intracellular location of phosphoenolpyruvate carboxykinase and other<br />

C 4 photosynthetic enzymes in mesophyll and bundle sheath protoplasts of Panicum maxymum. Plant<br />

Science Letters 29: 145–154.<br />

Choudhury SN, Bordoloi DN. 1992. Effect of liming on the uptake of nutrients and yield performance of<br />

Cymbopogon khasianus in acid soils of Northeast India. Indian Journal of Agronomy 37: 518–522.<br />

Das AB. 1999. Effects of different physiochemical factors on regeneration of Cymbopogon polyneuros Stapf.<br />

via callus culture and subsequent chromosomal stability. Israel Journal of Plant Sciences 47: 195–198.<br />

Dengler NG, Nelson T. 1999. Leaf structure and development in C4 plants. In Sage RF, Monson RK, Eds. The<br />

Biology of C4 Plants. San Diego: Academic Press, pp. 133–172.<br />

Dubey VS, Mallavarapu GR, Luthra R. 2000. Changes in the essential oil content and its composition during<br />

palmarosa (Cymbopogon martinii (Roxb.) Wats. var. motia) inflorescence development. Flavour and<br />

Fragrance Journal 15: 309–314.<br />

Eastman PAK, Dengler NG, Peterson CA. 1988. Suberized bundle sheaths in grasses (Poaceae) of different<br />

photosynthetic types I. Anatomy, ultrastructure, and histochemistry. Protoplasma 142: 92–111.<br />

Edwards GE, Andreo CS. 1992. NADP-Malic enzyme from plants. Phytochemistry 31: 1845–1857.


The Genus Cymbopogon 21<br />

Edwards GE, Walker OA. 1983. C3, C4: Mechanism, and Cellular and Environmental Regulation of<br />

Photosynthesis. London: Blackwell Scientific Publications.<br />

Fatima S, Farooqi AHA, Sharma S. 2002. Physiological and metabolic responses of different genotypes of<br />

Cymbopogon martinii and C. winterianus to water stress. Plant Growth Regulation 37: 143–149.<br />

Fitter AH, Hay RKM. 1987. Environmental Physiology of Plants. London: Academic Press.<br />

Ghannoum O, van Caemmerer S, Conroy JP. 2001. Carbon and water economy of Australian NAD-ME and<br />

NADP-ME C4 grasses. Australian Journal of Plant Physiology 28: 213–223.<br />

Guenther E. 1950a. The <strong>Essential</strong> <strong>Oil</strong>s (IV). New York, Van Nostrand. Ref Type: Serial (Book, Monograph).<br />

Guenther E. 1950b. The <strong>Essential</strong> <strong>Oil</strong>s. Princeton: Van Nostrand.<br />

Gupta ML, Janardhanan KK. 1991. Mycorrhizal association of Glomus aggregatum with palmarosa enhances<br />

growth and biomass. Plant and Soil 131: 261–263.<br />

Gutierrez M, Gracen VE, Edwards GE. 1974. Biochemical and cytological relationships in C 4 plants. Planta<br />

119: 279–300.<br />

Hatch MD. 1988. C 4 photosynthesis: A unique blend of modified biochemistry, anatomy and ultrastructure.<br />

Biochimica et Biophysica Acta 895: 81–106.<br />

Hatch MD. 1992. C 4 photosynthesis: An unlikely process full of surprises. Plant and Cell Physiology 33:<br />

333–342.<br />

Hatch MD, Kagawa T, Craig S. 1975. Subdivision of C 4-pathway species based on differing C 4 acid decarboxylating<br />

systems and ultrastructural features. Australian Journal of Plant Physiology 2: 111–128.<br />

Hatch MD, Kagawa T, Craig S. 2007. Subdivision of C 4-pathway species based on differing C 4 acid decarboxylating<br />

systems and ultrastructural features. Australian Journal of Plant Physiology 2: 111–128.<br />

Hattersley PW, Watson L. 1976. C 4 grasses: An anatomical criterion for distinguishing between NADP-malic<br />

enzyme species and PCK or NAD-malic enzyme species. Australian Journal of Botany 24: 297–308.<br />

Huang Y, Street-Perrot FA, Metcalfe SE, Brenner M, Moreland M, Freeman KH. 2001. Climate change as the dominant<br />

control on glacial–interglacial variations in C3 and C4 plant abundance. Science 293: 1647–1651.<br />

Jain N, Shasany AK, Sundaresan V, Rajkumar S, Darokar MP, Bagchi GD, Gupta AK, Kumar S, Khanuja SPS.<br />

2003. Molecular diversity in Phyllanthus amarus assessed through RAPD analysis. Current Science 85:<br />

1454–1458.<br />

Jenkins CLD, Furbank RT, Hatch MD. 1989. Mechanism of C 4 photosynthesis. Plant Physiology 91:<br />

1372–1381.<br />

Kak SN, Bhan MK, Rekha K. 2000. Development of improved clones of jamrosa (Cymbopogon nardus [L.]<br />

Rendle var. confertiflorus [Steud.] Bor. x C. jwarancusa Jones Schult.) through induced mutations.<br />

Journal of <strong>Essential</strong> <strong>Oil</strong> Research 12: 108–110.<br />

Khanuja SPS, Shasany AK, Pawar A, Lal RK, Darokar MP, Naqvi AA, Rajkumar S, Sundaresan V, Lal N, Kumar<br />

S. 2005. <strong>Essential</strong> oil constituents and RAPD markers to establish species relationship in Cymbopogon<br />

Spreng. (Poaceae). Biochemical Systematics and Ecology 33: 171–186.<br />

Kothari SK, Singh UB. 1996. Response of Java citronella (Cymbopogon winterianus Jowitt) to VA mycorrhizal<br />

fungi and soil compaction in relation to P supply. Plant and Soil 178: 231–237.<br />

Kresovich S, Lamboy RL, Li R, Szewc-McFadden AK, Blick SM. 1994. Application of molecular methods<br />

and statistical analysis for discrimination of accessions and clones of vetiver grass. Crop Science 34:<br />

805–809.<br />

Kumar J, Verma V, Shahi AK, Qazi GN, Balyan HS. 2007. Development of simple sequence repeat markers in<br />

Cymbopogon species. Planta Medica 73: 262–266.<br />

Kumar S, Dwivedi S, Kukreja AK, Sharma JR, Bagchi GD. 2000. Cymbopogon: The Aromatic Grass. Lucknow,<br />

India: Central Institute of Medicinal and Aromatic Plants.<br />

Kuriakose KP. 1995. Genetic variability in East Indian lemongrass (Cymbopogon flexuosus Stapf). Indian<br />

Perfumer 39: 76–83.<br />

Lal RK, Sharma JR, Misra HO, Sharma S, Naqvi AA. 2001. Genetic variability and relationship in quantitative<br />

and qualitative traits of Java citronella (Cymbopogon winterianus Jowitt). Journal of <strong>Essential</strong> <strong>Oil</strong><br />

Research 13: 158–162.<br />

Larkin PJ, Scowcroft WR. 1981. Somaclonal variation a novel source of variability from cell cultures for plant<br />

improvement. Theoretical and Applied Genetics 60: 197–214.<br />

Leegood RC. 1993. Carbon dioxide-concentrating mechanisms. In Lea PJ, Leegood RC, Eds. Plant Biochemistry<br />

and Molecular Biology. Chichester, U.K: Wiley & Sons, pp. 47–72.<br />

Lunn JE, Agostino A, Hatch MD. 1995. Regulation of NADP-malate dehydrogenase in C 4 plants—activity and<br />

properties of maize thioredoxin M and the significance of non-active site thiol groups. Australian Journal<br />

of Plant Physiology 55: 577–584.<br />

Maffei M. 2002. Vetiveria—The Genus Vetiveria. London: Taylor & Francis.


22 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Maffei M, Codignola A, Fieschi M. 1988. Photosynthetic enzyme activities in lemongrass cultivated in temperate<br />

climates. Biochemical Systematics and Ecology 16: 263, 264.<br />

Maheshwari SK, Chouhan GS, Trivedi KC, Gangrade SK. 1992. Effect of irrigation and stage of crop harvest<br />

on oil yield and quality of palmarosa (Cymbopogon martinii) oil grass. Indian Journal of Agronomy 37:<br />

514–517.<br />

Maheshwari SK, Sharma RK, Gangrade SK. 1995. Effect of spatial arrangement on performance of palmarosa<br />

(Cymbopogon martinii var. motia)—pigeonpea (Cajanus cajan) intercropping on a black cotton soil<br />

(vertisol). Indian Journal of Agronomy 40: 181–185.<br />

Maheshwari SK, Sharma RK, Gangrade SK. 1998. Response of palmarosa (Cymbopogon martinii var. motia)<br />

to biofertilizers, nitrogen, and phosphorus in a shallow black soil under rainfed condition. Indian Journal<br />

of Agronomy 43: 175–178.<br />

Mathews S, Spangler RE, Mason-Gamer RJ, Kellogg EA. 2002. Phylogeny of Andropogoneae inferred from<br />

phytochrome B, GBSSI, and NDHF. International Journal of Plant Sciences 163: 441–450.<br />

Mathur AK, Ahuja PS, Pandey B, Kukreja AK, Mandal S. 1988. Screening and evaluation of somaclonal variations<br />

for quantitative and qualitative traits in aromatic grass, Cymbopogon winterianus Jowitt. Plant<br />

Breeding 101: 321–334.<br />

Maurino VG, Drincovich MF, Casati P, Andreo CS, Edwards GE, Ku MSB, Gupta SK, Franceschi VR. 1997.<br />

NADP-malic enzyme: Immunolocalization in different tissues of the C 4 plant maize and the C 3 plant<br />

wheat. Journal of Experimental Botany 48: 799–811.<br />

Metcalfe CR. 1960. Anatomy of the Monocotyledons. I. Gramineae. London: Oxford University Press.<br />

Misra A, Khan A. 1992. Correction of iron-deficiency chlorosis in lemongrass (Cymbopogon flexuosus Steud.)<br />

Watts. Agrochimica 36: 349–360.<br />

Misra A, Srivastava NK. 1991. Effect of the triacontanol formulation miraculan on photosynthesis, growth,<br />

nutrient uptake, and essential oil yield of lemongrass (Cymbopogon flexuosus) Steud. Watts. Plant<br />

Growth Regulation 10: 57–63.<br />

Munthali MT, Newbury HJ, Ford-Loyd BV. 1996. The detection of somaclonal variants of beet using RAPD.<br />

Plant Cell Reports 15: 474–478.<br />

Nayak S, Debata BK, Sahoo S. 1996. Rapid propagation of lemongrass (Cymbopogon flexuosus [Nees] Wats.)<br />

through somatic embryogenesis in vitro. Plant Cell Reports 15: 367–370.<br />

Nayak S, Debata BK, Srivastava VK, Sangwan NS. 2003. Evaluation of agronomically useful somaclonal variants<br />

in jamrosa (a hybrid Cymbopogon) and detection of genetic changes through RAPD. Plant Science<br />

164: 1029–1035.<br />

Omisra A, Srivastava NK. 1994. Influence of iron nutrition on chlorophyll contents, photosynthesis, and<br />

essential monoterpene oil(s) in Java citronella (Cymbopogon winterianus Jowitt). Photosynthetica 30:<br />

425–434.<br />

Patnaik J, Debata BK. 1997a. In vitro selection of NaCl tolerant callus lines of Cymbopogon martinii (Roxb.)<br />

Wats. Plant Science 124: 203–210.<br />

Patnaik J, Debata BK. 1997b. Regeneration of plantlets from NaCl tolerant callus lines of Cymbopogon martinii<br />

(Roxb.) Wats. Plant Science 128: 67–74.<br />

Patnaik J, Sahoo S, Debata BK. 1996. Cytology of callus, somatic embryos and regenerated plants of palmarosa<br />

grass, Cymbopogon martinii (Roxb.) Wats. Cytobios 87: 79–88.<br />

Patnaik J, Sahoo S, Debata BK. 1997. Somatic embryogenesis and plantlet regeneration from cell suspension<br />

cultures of palmarosa grass (Cymbopogon martinii). Plant Cell Reports 16: 430–434.<br />

Patnaik J, Sahoo S, Debata BK. 1999. Somaclonal variation in cell suspension culture-derived regenerants of<br />

Cymbopogon martinii (Roxb.) Wats. var. motia. Plant Breeding 118: 351–354.<br />

Pecchioni N, Faccioli P, Monetti A, Stanca AM, Terzi V. 1996. Molecular markers for genotype identification<br />

in small grain cereals. Journal of Genetic Breeding 50: 203–219.<br />

Phillips SM, Hua P. 2005. Notes on grasses (Poaceae) for the flora of China, V. New species in Cymbopogon.<br />

Novon 15: 471–473.<br />

Prat H. 1937. Caractères anatomique et histologiques de quelques Andropogonées de l’Afrique occidentale.<br />

Ann. Mus. Colon. Marseille 5: 25–28.<br />

Purseglove JW. 1975. Tropical Crops: Monocotyledons. London: Longman Group Ltd.<br />

Puttanna K, Gowda NMN, Rao EVSP. 2001. Effects of applications of N fertilizers and nitrification inhibitors<br />

on dry matter and essential oil yields of Java citronella (Cymbopogon winterianus Jowitt.). Journal of<br />

Agricultural Science 136: 427–431.<br />

Quiala E, Barbon R, Jimenez E, De Feria M, Chavez M, Capote A, Perez N. 2006. Biomass production of<br />

Cymbopogon citratus (DC) Stapf., a medicinal plant, in temporary immersion systems. In Vitro Cellular<br />

and Developmental Biology–Plant 42: 298–300.


The Genus Cymbopogon 23<br />

Rajendrudu G, Das VSR. 1981. C4 photosynthetic carbon metabolism in the leaves of aromatic tropical<br />

grasses—I. Leaf anatomy, CO 2 compensation point and CO 2 assimilation. Photosynthesis Research 2:<br />

225–233.<br />

Rani V, Parida A, Raina SN. 1995. Random amplified polymorphic DNA (RAPD) micropropagated plants of<br />

Populus deltoides Marsh. Plant Cell Reports 14: 459–462.<br />

Rao BRR. 2000. Biomass yield and essential oil yield variations in Java citronella (Cymbopogon winterianus<br />

Jowitt.), intercropped with food legumes and vegetables. Journal of Agronomy and Crop Science—<br />

Zeitschrift fur Acker und Pflanzenbau 185: 99–103.<br />

Rao EVSP, Singh M, Chandrasekhara G. 1988. Effect of nitrogen application on herb yield, nitrogen uptake<br />

and nitrogen recovery in Java citronella (Cymbopogon Winterianus Jowitt). Indian Journal of Agronomy<br />

33: 412–415.<br />

Rao EVSP, Singh M, Rao RSG. 1994. Performance of intercropping systems based on palmarosa (Cymbopogon<br />

martinii var. motia). Indian Journal of Agricultural Sciences 64: 442–445.<br />

Ratti N, Kumar S, Verma HN, Gautam SP. 2001. Improvement in bioavailability of tricalcium phosphate to<br />

Cymbopogon martinii var. motia by rhizobacteria, AMF and Azospirillum inoculation. Microbiological<br />

Research 156: 145–149.<br />

Rothermel BA, Nelson T. 1989. Primary structure of the maize NADP-dependent malic enzyme. Journal of<br />

Biological Chemistry 264: 19587–19592.<br />

Sangwan NS, Yadav U, Sangwan RS. 2001. Molecular analysis of genetic diversity in elite Indian cultivars of<br />

essential oil trade types of aromatic grasses (Cymbopogon species). Plant Cell Reports 20: 437–444.<br />

Sangwan NS, Yadav U, Sangwan RS. 2003. Genetic diversity among elite varieties of the aromatic grasses,<br />

Cymbopogon martinii. Euphytica 130: 117–130.<br />

Sangwan RS, Farooqi AHA, Bansal RP, Singhsangwan N. 1993. Interspecific variation in physiological and<br />

metabolic responses of 5 species of Cymbopogon to water stress. Journal of Plant Physiology 142:<br />

618–622.<br />

Sangwan RS, Sangwan NS, Jain DC, Kumar S, Ranade SA. 1999. RAPD profile-based genetic characterization<br />

of chemotypic variants of Artemisia annua L. Biochemistry and Molecular Biology International<br />

47: 935–944.<br />

Senthilkumar K, Udaiyan K, Manian S. 1992. Rate of litter decomposition in a tropical grassland dominated by<br />

Cymbopogon caesius in Southern India. Tropical Grasslands 26: 235–242.<br />

Shasany AK, Lal RK, Patra NK, Darokar MP, Garg A, Kumar S, Khanuja SPS. 2000. Phenotypic and RAPD<br />

diversity among Cymbopogon winterianus Jowitt accessions in relation to Cymbopogon nardus Rendle.<br />

Genetic Resources and Crop Evolution 47: 553–559.<br />

Singh A, Singh M, Singh K. 1998. Productivity and economic viability of a palmarosa-pigeonpea intercropping<br />

system in the subtropical climate of North India. Journal of Agricultural Science 130: 149–154.<br />

Singh K, Singh DV. 1992. Effect of rates and sources of nitrogen application on yield and nutrient-uptake of<br />

Java citronella (Cymbopogon winterianus Jowitt). Fertilizer Research 33: 187–191.<br />

Singh M. 2001. Long-term studies on yield, quality, and soil fertility of lemongrass (Cymbopogon flexuosus) in<br />

relation to nitrogen application. Journal of Horticultural Science and Biotechnology 76: 180–182.<br />

Singh M, Rao RSG, Ramesh S. 2005. Effects of nitrogen, phosphorus, and potassium on herbage, oil yield,<br />

oil quality, and soil fertility status of lemongrass in a semi-arid tropical region of India. Journal of<br />

Horticultural Science and Biotechnology 80: 493–497.<br />

Singh M, Rao RSG, Rao EVSP. 1996a. Effect of depth and method of irrigation and nitrogen application on<br />

herb and oil yields of Java citronella (Cymbopogon winterianus Jowitt) under semi-arid tropical conditions.<br />

Journal of Agronomy and Crop Science—Zeitschrift fur Acker und Pflanzenbau 177: 61–64.<br />

Singh M, Shivaraj B. 1998. Intercropping studies in lemongrass (Cymbopogon flexuosus) (Steud, Wats.).<br />

Journal of Agronomy and Crop Science—Zeitschrift fur Acker und Pflanzenbau 180: 23–26.<br />

Singh M, Shivaraj B. 1999. Effect of irrigation regimes on growth, herbage and oil yields of lemongrass<br />

(Cymbopogon flexuosus) under semi-arid tropical conditions. Indian Journal of Agricultural Sciences<br />

69: 700–702.<br />

Singh M, Shivaraj B, Sridhara S. 1996b. Effect of plant spacing and nitrogen levels on growth, herb and oil<br />

yields of lemongrass (Cymbopogon flexuosus (Steud) Wats. var. cauvery). Journal of Agronomy and Crop<br />

Science—Zeitschrift fur Acker und Pflanzenbau 177: 101–105.<br />

Singh N, Luthra R, Sangwan RS. 1990. Oxidative pathways and essential oil biosynthesis in the developing<br />

Cymbopogon flexuosus leaf. Plant Physiology and Biochemistry 28: 703–710.<br />

Singh N, Luthra R, Sangwan RS. 1991. Mobilization of starch and essential oil biogenesis during leaf ontogeny<br />

of lemongrass (Cymbopogon flexuosus Stapf). Plant and Cell Physiology 32: 803–811.


24 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Singh RS, Bhattacharyya TK, Kakti MC, Bordoloi DN. 1992. Effect of nitrogen, phosphorus and potash on<br />

essential oil production of palmarosa (Cymbopogon martinii var. motia) under rainfed condition. Indian<br />

Journal of Agronomy 37: 305–308.<br />

Singh S, Ram M, Ram D, Sharma S, Singh DV. 1997. Water requirement and productivity of palmarosa on<br />

sandy loam soil under a sub-tropical climate. Agricultural Water Management 35: 1–10.<br />

Singh S, Ram M, Ram D, Singh VP, Sharma S, Tajuddin. 2000. Response of lemongrass (Cymbopogon flexuosus)<br />

under different levels of irrigation on deep sandy soils. Irrigation Science 20: 15–21.<br />

Singhsangwan N, Farooqi AHA, Sangwan RS. 1994. Effect of drought stress on growth and essential oil metabolism<br />

in lemongrasses. New Phytologist 128: 173–179.<br />

Smith BN, Brown WV. 1973. The Kranz syndrome in the Gramineae as indicated by carbon isotopic ratios.<br />

American Journal of Botany 60: 505–513.<br />

Spies JJ, Troskie TH, Vandervyver E, Vanwyk SMC. 1994. Chromosome studies on African plants—11. The<br />

tribe Andropogoneae (Poaceae, Panicoideae). Bothalia 24: 241–246.<br />

Sprengel CPJ. 1815. Cymbopogon. Plantarum Minus Cognitarum Pugillus 2: 14.<br />

Srivastava NK, Misra A, Sharma S. 1998. The substrate utilization and concentration of C-14 photosynthates<br />

in citronella under Fe deficiency. Photosynthetica 35: 391–398.<br />

Taylor PWJ, Geijskes JR, Ko HL, Fraser TA, Henry RJ, Birch RG. 1995. Sensitivity of random amplified<br />

polymorphic DNA analysis to detect genetic change in sugarcane during tissue culture. Theoretical and<br />

Applied Genetics 90: 1165–1173.<br />

Trevanion SJ, Furbank RT, Ashton AR. 1997. NADP-Malate dehydrogenase in the C 4 plant Flaveria bidentis.<br />

Plant Physiology 113: 1153–1165.<br />

Vickery JW. 1935. The leaf anatomy and vegetative characters of the indigenous grasses of N.S. Wales.<br />

Proceedings of the Linnean Society, N.S.W. 60: 340–373.<br />

Wallner E, Weising R, Rompf G, Kahl G, Kopp B. 1996. Oligonucleotide finger printing and RAPD analysis<br />

of Achillea species: characterisation and long term monitoring of micropropagated clones. Plant Cell<br />

Reports 15: 647–652.<br />

Welsh JM, McClelland M. 1990. Fingerprinting genomes using PCR with arbitrary primers. Nucleic Acids<br />

Research 18: 7213–7218.<br />

Yang H, Tabei Y, Kamad H, Kayano T, Takaiwa F. 1999. Detection of somaclonal variation in cultured rice cell<br />

using digoxigenin-based random amplified polymorphic DNA. Plant Cell Reports 18: 520–526.


2<br />

contents<br />

Chemistry and Biogenesis<br />

of <strong>Essential</strong> <strong>Oil</strong> from<br />

the Genus Cymbopogon<br />

Anand Akhila<br />

2.1 Introduction ............................................................................................................................26<br />

2.2 Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from Cymbopogons ...........................................26<br />

2.3 Physicochemical Characteristics of the <strong>Essential</strong> <strong>Oil</strong>s from Cymbopogons ..........................27<br />

2.4 Chemistry and Uses of Cymbopogon <strong>Essential</strong> <strong>Oil</strong>s ..............................................................68<br />

2.4.1 Lemongrass <strong>Oil</strong>s .........................................................................................................68<br />

2.4.1.1 Cymbopogon flexuosus ................................................................................68<br />

2.4.1.2 Cymbopogon citratus ...................................................................................69<br />

2.4.1.3 Cymbopogon pendulus ................................................................................ 71<br />

2.4.2 Citronella <strong>Oil</strong>s ............................................................................................................. 71<br />

2.4.2.1 Cymbopogon winterianus ............................................................................ 71<br />

2.4.2.2 Cymbopogon nardus ....................................................................................72<br />

2.4.3 Palmarosa and Gingergrass <strong>Oil</strong>s ................................................................................73<br />

2.4.3.1 Cymbopogon martinii ..................................................................................73<br />

2.4.4 Cymbopogon jwarancusa ........................................................................................... 74<br />

2.4.5 Cymbopogon schoenanthus ........................................................................................ 75<br />

2.4.6 Other Cymbopogon Species ....................................................................................... 75<br />

2.4.6.1 Cymbopogon caesius ................................................................................... 75<br />

2.4.6.2 Cymbopogon coloratus ................................................................................ 75<br />

2.4.6.3 Cymbopogon confertiflorus ......................................................................... 75<br />

2.4.6.4 Cymbopogon densiflorus ............................................................................. 76<br />

2.4.6.5 Cymbopogon distans ................................................................................... 76<br />

2.4.6.6 Cymbopogon khasianus ...............................................................................77<br />

2.4.6.7 Cymbopogon ladakhensis ............................................................................77<br />

2.4.6.8 Cymbopogon microstachys ..........................................................................77<br />

2.4.6.9 Cymbopogon nervatus .................................................................................77<br />

2.4.6.10 Cymbopogon olivieri ...................................................................................77<br />

2.4.6.11 Cymbopogon parkeri ................................................................................... 78<br />

2.4.7 Some Lesser-Known Species ...................................................................................... 78<br />

2.4.7.1 Cymbopogon polyneuros ............................................................................. 78<br />

2.4.7.2 Cymbopogon procerus ................................................................................. 78<br />

2.4.7.3 Cymbopogon rectus ..................................................................................... 78<br />

2.4.7.4 Cymbopogon sennarensis ............................................................................ 78<br />

2.4.7.5 Cymbopogon stracheyi ................................................................................ 78<br />

2.4.7.6 Cymbopogon tortilis .................................................................................... 78<br />

2.4.7.7 Cymbopogon travancorensis ....................................................................... 78<br />

25


26 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

2.4.7.8 Cymbopogon goeringii ................................................................................79<br />

2.4.7.9 Cymbopogon asmastonii .............................................................................79<br />

2.4.7.10 Cymbopogon giganteus ...............................................................................79<br />

2.4.8 Biosynthesis of Terpenes in Cymbopogon Species ....................................................79<br />

2.4.9 Biological Activities....................................................................................................84<br />

2.4.9.1 Pain Reliever ................................................................................................89<br />

2.4.9.2 Activity against Leukemia and Malignancy ................................................ 91<br />

2.4.9.3 Activation of Male Hormones ...................................................................... 91<br />

2.4.9.4 Activity against Worms ............................................................................... 91<br />

2.4.9.5 Lowering Blood Sugar ................................................................................. 91<br />

2.4.9.6 Potential to Repel Mosquitoes and Kill the Larvae ..................................... 91<br />

2.4.9.7 Activity to Reduce Edema ...........................................................................92<br />

2.4.9.8 Potential to Control Aging Process ..............................................................92<br />

2.4.9.9 Activity against Pests ...................................................................................93<br />

2.4.9.10 Activity against Microbes ............................................................................93<br />

References ........................................................................................................................................95<br />

2.1 IntroductIon<br />

Aromatic grasses are one of the chief sources of essential oils. The genus Cymbopogon comprises<br />

a large number of species, out of which lemongrass, citronella, palmarosa, and few others produce<br />

oil of commercial importance (Gupta 1969; Gupta and Deniel 1982; Gupta and Jain 1978; Gupta<br />

et al. 1975). The chemical compounds present in the essential oils of Cymbopogon do not reflect<br />

the actual olfactory or other properties of the species (Gildemeister and Hoffmann 1956). There are<br />

instances when distinct species such as lemongrass oils from C. pendulus, C. citratus, and C. flexuosus<br />

(Anonymous 1958) produce oil of almost identical chemical composition. “Chemical characters<br />

are like other characters; they work when they work and they don’t work when they don’t work. Like<br />

all taxonomic characters they attain their value through correlation with other characters,” has been<br />

rightly quoted by Cronquits (1980). However, most of the Cymbopogon species found all around the<br />

world produce essential oils that differ widely in their physical properties and chemical constituents.<br />

The varieties motia and sofia of C. martinii are good examples of such variable characters.<br />

The Cymbopogon species has great prospects for producing quality essential oils (Arctander<br />

1960; Han et al. 1971), and it has direct relevance to the perfumery industry with economic benefit<br />

to humankind. However, the actual potential of cymbopogons has not been exploited to the fullest.<br />

Though tremendous work has been done regarding Cymbopogon chemistry, a lot more needs to be<br />

done to make use of the major and minor constituents present in its essential oils, particularly the<br />

mono- and sesquiterpenes (Chopra et al. 1956). An effort has been made in this chapter to present a<br />

comprehensive overview of most of the cultivated and wild species of cymbopogons.<br />

2.2 chemIstry and bIogenesIs oF essentIal oIl From cymboPogons<br />

About 25 to 30 species are reported in genus Cymbopogon, and many of them are very good<br />

sources of essential oils of commercial importance. The compounds present in these oils are characteristic,<br />

but cannot necessarily be used for identification, of the species. Several botanical races<br />

of these species produce essential oils that are entirely different in their constituents. The essential<br />

oils of the Cymbopogon species mainly comprises of mono- and sesquiterpenoids and, despite<br />

their importance, very few high-tech identification techniques (such as GC-MS high resolution)<br />

have been utilized to identify the minor and trace constituents present in them; further, only a few<br />

reports are available in the literature. An attempt has been made in this chapter to present a complete<br />

analysis of most of the essential oils obtained from these Cymbopogon species. Biosynthetic<br />

pathways to most of the mono- and sesquiterpenes have been discussed. Efforts have also been


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 27<br />

made to provide complete data on the physicochemical properties of the oil and spectroscopic data<br />

( 1 13 H2 CNMR) of its individual constituents (Table 2.2 and Table 2.3). During the years 1950 to<br />

1980, when GC and GC-MS techniques were not frequently available for the analysis of essential<br />

oils, the most significant data that could have been used was the density, specific gravity, refractive<br />

index, specific rotation, and solubility of the oil in aqueous alcohol.<br />

2.3 PhysIcochemIcal characterIstIcs oF the essentIal oIls<br />

From cymboPogons<br />

Physical and chemical properties of any essential oil are of prime importance, and chemists are<br />

now working in an era when highly sophisticated instruments are available for quality and quantity<br />

analysis. Still, the specific gravity, optical rotation, solubility in dilute alcohol, and the refractive<br />

index must be determined for all oils and liquid isolates. Before the availability of modern analytical<br />

techniques, the essential oil chemists were working using their ingenuity, a highly developed<br />

sense of smell and taste, and analytical ability. Besides the determination of physical and chemical<br />

properties, other tests have also been carried out, such as ester content, total alcohol determination,<br />

congealing point, and melting points in the case of solids, which is of great importance. The reported<br />

values of these constants for the essential oils of Cymbopogon species are shown in Table 2.1, which<br />

is self-explanatory. This chapter will cater to the needs of students besides researchers and, therefore,<br />

brief definitions of the physicochemical characteristics, which are highly relevant in testing the<br />

quality of the essential oils, have been provided.<br />

Specific gravity—Specific gravity is defined as the ratio of the density of a given solid or liquid<br />

substance to the density of water at a specific temperature and pressure, typically at 4°C (39°F)<br />

and 1 atm (14.7 psia). Substances with a specific gravity greater than 1 are denser than water, and<br />

so (ignoring surface tension effects) will sink in it, and those with a specific gravity less than 1 are<br />

less dense than water, and hence will float in it. Specific gravity is a special case of relative density,<br />

with the latter term often preferred in modern scientific writing. Specific gravity (SG) is expressed<br />

mathematically as<br />

density of thesubstance ρ<br />

SG =<br />

density of water<br />

ρl<br />

where ρ is the density of the substance, and ρ1 is the density of water. (By convention ρ, the Greek<br />

letter rho, denotes density.)<br />

Refractive index—The refractive index (or index of refraction) of a medium is a measure of<br />

how much the speed of light (or other waves such as sound waves) is reduced inside the medium.<br />

For example, typical glass has a refractive index of 1.5, which means that, in glass, light travels<br />

at 1/1.5 = 0.67 times the speed of light in a vacuum. Two common properties of glass and other<br />

transparent materials are directly related to their refractive index. First, light rays change direction<br />

when they cross the interface from air to the material, an effect that is used in lenses and glasses.<br />

Second, light reflects partially from surfaces that have a refractive index different from that of their<br />

surroundings.<br />

Definition: The refractive index n of a medium is defined as the ratio of the phase velocity c of<br />

a wave phenomenon such as light or sound in a reference medium to the phase velocity vp in the<br />

medium itself: n = c/νp.<br />

Specific rotation—The specific rotation of a chemical compound [α] is defined as the observed<br />

angle of optical rotation α when plane-polarized light is passed through a sample with a path length<br />

of 1 dm and a sample concentration of 1 g/dL. The specific rotation of a pure material is an intrinsic<br />

property of that material at a given wavelength and temperature. Values should always be accompanied<br />

by the temperature at which the measurement was performed and the solvent in which the


28 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.1 major and minor constituents and Physiochemical Properties of essential oils<br />

obtained from cymbopogon species<br />

Cymbopogon flexuosus (steud.) Wats.<br />

Major terpenes—citral-a (geranial; 40%–50%), citral-b (neral; 30%–35%), Atal and Bradu 1976a; De Martinez 1977<br />

borneol (0%–2%), citronellal (0.37%–8.04%), citronellol (0.44%–4.58%),<br />

citronellyl acetate (1.2%–3.6%), geraniol (1.73%–40.0%), geranyl acetate<br />

(1.95%–5.1%), limonene (2.4%–3.7%), and methyl eugenol (20%)<br />

Myrcene (0.1%–14.2%) Formacek and Kubeczka 1982; Borovik<br />

and Kuravskaya 1977; Bhattacharya et al.<br />

1997; Gonzalo and Villarrubia 1973;<br />

Guenther 1950; Gonzalo 1973; Taskinen<br />

et al. 1983; Thapa and Agarwal 1989;<br />

Thapa et al. 1976; Thapa et al. 1981<br />

Traces—α-bergamotene, β-bisabolene, τ-cadinene, α-cadinol, camphene, Atal and Bradu 1976a; Chiang et al. 1981;<br />

δ-3-carene, β-caryophyllene, β-caryophyllene oxide, 1,8-cineole,<br />

Jyrkit 1983; Le and Chu 1976; Foda et al.<br />

α-curcumene, p-cymene, n-decyldehyde, dipentene, β-elemene,<br />

1975; Mohammad et al. 1981a, 1981b;<br />

τ-elemene, elemicin, elemol, farnesol, geranyl formate, α- humulene, Nair et al. 1980a, 1980b; Sobti et al.<br />

isopulegol, linalool, linalyl acetate, p-menthane, methyl heptenol, methyl 1978c, 1982; Srikulvandhana et al. 1976;<br />

heptenone, τ-murolene, nerol, nerolidol, neryl acetate, 2-nonanone,<br />

Zaki et al. 1975; Jyrkit 1983<br />

cis-β-ocimene, τ-β-ocimene, perillene, phellandrene, α-pinene, β-pinene,<br />

piperitone, terpinen-4-ol, α-terpineol, and terpinolene<br />

origin specific gravity ηd [α] d solubility reference<br />

Cochin 0.899–0.905 at 15° 1.4883–1.488 at 20° +1°25′ to −5°0′ 1.5–3 vol. of Guenther 1950;<br />

Aldehyde content (a) Bisulfite method 70%–80%; (b) Neutral 70% alcohol Gildemeister and<br />

sulfite method 65%–80%<br />

Hoffmann 1956<br />

Ceylon 0.895–0.908 at 15° 1.483–1.489 at 20° 1° to 5° NA De Sylva 1959<br />

Aldehyde content 65% to 85%<br />

Jammu Regional 0.9212 at 18° 1.4895 at 20° −15.2° NA Thapa et al. 1976<br />

Research<br />

Laboratory<br />

(RRL)-57<br />

Acid value 2.69; ester value 73.01<br />

Jammu Regional 0.9137 at 18° 1.4887 at 20° −15.5° NA Thapa et al. 1976<br />

Research<br />

Laboratory<br />

(RRL)-59<br />

Acid value 2.52; ester value 62.79<br />

Kerala 0.899–0.905 1.480–1.486 at 35° +1°25′ to −5°0′ 75% alcohol Chakrabarti and<br />

at 35° Ghosh 1974<br />

Kumaon 0.9651 1.489 NA NA Baslas and Baslas<br />

Acid value 22.28; ester value 87.22<br />

1968<br />

Lucknow 0.8911 at 25° 1.4816 at 30° −1° 1.2 vol. of Virmani and Datta<br />

Aldehyde content 89%<br />

70% alcohol 1973<br />

Lucknow 0.892 at 30° 1.4825 at 30° −1.5° NA Sharma et al. 1972<br />

Aldehyde content 71.3%<br />

Odakali 0.8886 at 20° 1.4862 at 20° −0.40° NA Thapa et al. 1981<br />

Acid value 3.65; ester value 46.77<br />

Pantnagar 0.8975–0.899 1.483–1.488 1°25′ to 5° 1.2 vol. of Gulati et al. 1976<br />

Aldehyde content 81.39%<br />

70% alcohol<br />

West Bengal 0.90 1.4803 at 35° +1°30′ 75% alcohol Chakrabarti and<br />

at 35° Ghosh 1974


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 29<br />

table 2.1 (continued) major and minor constituents and Physiochemical Properties<br />

of essential oils obtained from cymbopogon species<br />

origin specific gravity η d [α] d solubility reference<br />

Travancore 0.895–0.908 1.483–1.489 +1°30′ to −5° 3 vol. of 70%<br />

alcohol<br />

Indian Standards<br />

Institution (ISI)<br />

0.892–0.902 at<br />

25°<br />

Anonymous 1950<br />

1.4802 at 20° −3° to +1° Anonymous 1952<br />

Cymbopogon jwarancusa (Jones) schult.<br />

Major terpenes—piperitone (20%–70%) and ∆ 4 -carene (20%–24%),<br />

citronellal (30%–40%), p- cymene (0.6%–3.5%), geraniol (0.04%–22.5%),<br />

β-pinene (3.5%), and γ-terpinene (7.5%)<br />

Traces—alloaromadendrene, cis- and γ-allo-ocimene, α- bisabolene,<br />

β-bisabolene, borneol, d-cadinene, calamene, camphene, camphor,<br />

β-caryophyllene, β-caryophyllene oxide, α -chamigrene, 1,8-cineole,<br />

citronellol, α-cubebene, cuprene, o-cymene, 5,6-dimethyl-5-norbornen-<br />

2-ol, dipentene, β-elemene, d -elemene, elemol, eucarvone, eudesmol,<br />

α -farnesene, β-farnesene, fenchone, geranyl acetate, geranyl formate,<br />

geranyl propionate, germacrene, α-humulene, iso-borneol, kasuralcohol,<br />

lavendulol, linalool, longifolene, p-mentha-2,8-dien-1-ol, cis- and<br />

γ-p-mentha -2-en-1-ol methyl heptenone, methyl thymyl ether,<br />

α-muurolene, myrcene, myrtenal, phellandrene, α-pinene, γ- and<br />

cis-peperitol, terpinen-4-ol, α-terpineol, terpinolene, γ-thuj -2-en-4-ol,<br />

verbenone, and β-ylangene<br />

Ansari and Quadry 1987; Balyan et al.<br />

1979; Dev et al. 1988; Dhar and Lattoo<br />

1985; Dhar et al. 1981; Dhar and Dhar<br />

1997; Guenther 1950; Liu et al. 1981;<br />

Maheshwari and Mohan 1985; Mathela<br />

and Pant 1988; Mathela et al. 1986; Nair<br />

et al. 1982; Saeed et al.1978; Shahi 1992;<br />

Shahi and Sen 1989; Sobti et al. 1982;<br />

Thapa et al. 1971; Shahi and Tava 1993<br />

origin specific gravity η d [α] d solubility reference<br />

Not<br />

specified<br />

0.9203–0.9228 at 30° 1.481–1.4858 at 30° +51°41′ to 42°48′ at 30° NA Guenther 1948,<br />

1950<br />

Acid value 0.7;<br />

ester value 12.0<br />

Hazara 0.9203 at 30° 1.481 at 30° +51.65° at 30° NA Anonymous 1950<br />

Sind 0.923 at 30° 1.4858 at 30° +42.8° at 30° NA Anonymous 1950<br />

UP 0.909 at 30° 1.4856 at 30° +25.7° at 30° NA Anonymous 1950<br />

Cymbopogon martinii (roxb.) Wats.<br />

Major terpenes—geraniol (65%–85%), citral (4%–12%), citronellol<br />

(6.4%), linalool (2.4%), and geranyl acetate (6%–12%)<br />

Traces— α-amorphine, β-betulenol, α-betulenol, bicyclogermacrene,<br />

β-bisabolene, γ-bisabolene, α-cadinene, γ-cadinene, cis- calamene,<br />

γ-calamene, calacorene, β-curcumene, o-cymene, p-cymene, m-cymene,<br />

dipentene, β-elemene, γ-elemene, β-farnesene, farnesol, farnesyl acetate,<br />

formaldehyde, geranyl-n-butyrate, germacrene-B, germacrene-D,<br />

β-helmiscapene, α-humulene, β-humulene, isovaleraldehyde, limonene,<br />

methyl heptenone, myrcene, γ-muurolene, nerolidol, 2-nonanol,<br />

α-phellandrene, α-pinene, β-pinene, selina-4,7-diene, α-selinene,<br />

β-selinene, d selinene, α-terpinene, γ-terpinene, α-terpineol, β-terpineol,<br />

and terpinolene<br />

Gaydou and Raudriamiharisoa 1987;<br />

Guenther 1950; Mallavarapu et al. 1998;<br />

Peyron 1972, 1973; Anon. 1973<br />

Anonymous 1980; Chiang et al. 1981; De<br />

Martinez 1977; Maheshwari and Mohan<br />

1985; Mohammad et al. 1981b; Nair et al.<br />

1980a, 1980b; Nigam et al. 1987;<br />

Oliveros-Belardo 1989; Sobti et al. 1981,<br />

1982; Bottani et al. 1987; Naves 1970,<br />

1971; Opdyke 1974<br />

(continued on next page)


30 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.1 (continued) major and minor constituents and Physiochemical Properties<br />

of essential oils obtained from cymbopogon species<br />

origin specific gravity η d [α] d solubility reference<br />

India 0.8903–0.8911 at 15° 1.4718–1.4738 at<br />

20°<br />

−0°5′ to<br />

1°20′<br />

Java 0.891–0.892 — +0°30′ to<br />

0°42′<br />

West<br />

Bengal<br />

1.5 vol. of<br />

70% alcohol<br />

3.5 vol. of<br />

60% alcohol<br />

Anonymous<br />

1950<br />

0.89 1.4702 at 35° −4° 75% alcohol Chakrabarti and<br />

Ghosh 1974<br />

ISI 0.8778–0.8898 at 25°C 1.4710–1.4755 at 25° −2° to +3° NA Anonymous<br />

Acid value 3.0; alcohol content 88%–94%<br />

1952<br />

Hyderabad 0.889–0.900 1.468–1.4729 −3° to +6° 80% alcohol Chakrabarti and<br />

Ghosh 1974<br />

Jammu 0.889 at 15° 1.475 at 20° NA NA Sobti et al.<br />

Acid value 0.5–3.0; ester value 12.48<br />

1981<br />

Lucknow 0.8810 at 25° 1.4722–1.4702 at 25° +1° 2 vol. of 70% Virmani and<br />

(Pre-winter<br />

harvest)<br />

Acid value 0.5; ester value 6.8; geraniol content 93.7%<br />

alcohol Dutta 1973<br />

Cymbopogon martinii stapf.<br />

Major terpenes—geraniol (36%–65%), perillyl alcohol (15%–27%),<br />

p-menthenols (40%–62%), γ- and cis -carveol (3%–25%), 1,8-cineole<br />

(9.8%), and iso piperitenol (13.4%)<br />

Traces—∆ 3 -carene, carveyl acetate, d,l-carvone, caryophyllene oxide,<br />

p-cymene, dihydrocarveol, γ- and cis-dihydrocarvone, dipentene,<br />

d-limonene oxide, d-α-phellandrene, α-pinene, piperitone oxide, and<br />

tricylene<br />

Boelens 1994; Guenther 1950; Kalia et al.<br />

1980; Nigam et al. 1965; Sobti et al. 1978;<br />

Thapa et al. 1971, 1981<br />

origin specific gravity ηd [α] d solubility reference<br />

ISI 0.8997–0.9287 at 1.4760–1.4910 −14° to +54° NA Anonymous 1952<br />

25°C<br />

at 25°<br />

Acid value 6.0; alcohol content 36%–60%<br />

Jammu (wild 0.9646 at 20° 1.4974 at 20° +22.56° NA Kalia et al. 1980<br />

collection) Acid value 22.4; ester value 92<br />

Madras 0.900–0.953 at 15° 1.4780–1.4930 −30° to +54° Soluble in 2.3 vol. Gildemeister and<br />

at 20°<br />

of 70% alcohol Hoffmann 1956;<br />

Acid value 6.2; ester value 8.0<br />

Guenther 1950<br />

Cymbopogon citratus (d.c.) stapf.<br />

Major terpenes—citral-a or geranial (10%–48%) and citral-b or<br />

neral (3%–43%), borneol (5%), geraniol (2.6%–40%), geranyl<br />

acetate (0.1%–3.0%), linalool (1.2%–3.4%), and nerol<br />

(0.8%–4.5%)<br />

Traces—camphene, camphor, α-camphorene, ∆-3-carene,<br />

caryophyllene, caryophyllene oxide, 1,8-cineole, citronellal,<br />

citronellol, n-decyldehyde, α,β-dihydropseudoionone, dipentene,<br />

β-elemene, elemol, farnesal, farnesol, fenchone, furfural,<br />

iso-pulegol, iso-valeraldehyde, limonene, linalyl acetate, menthol,<br />

menthone, methyl heptenol, ocimene, α-oxobisabolene,<br />

β-phellandrene, α-pinene, β-pinene, terpineol, terpinolene,<br />

2-undecanone, neral, nerolic acid, and geranic acid<br />

Abdullah et al. 1975; Abegaz et al. 1983; Brazil<br />

et al. 1971; Baruah et al. 1995; Guenther 1950;<br />

Idrissi et al. 1993; Thapa et al. 1981; Torres 1993;<br />

Zheng et al. 1993<br />

Beech 1977; Crawford et al. 1975; El Tawil and El<br />

Beih 1982; Hanson et al. 1976; Kusumov and<br />

Babaev 1983; Liu et al. 1981; Manjoor-i-Khuda<br />

et al. 1984; Mathela 1991; Neyberg 1953; Nigam<br />

et al. 1987; Olaniyi et al. 1975; Opdyke 1973;<br />

Oliveros-Belardo and Aureus 1978, 1979; Rabha<br />

et al. 1979; Rouesti and Voriate 1960; Sarer et al.<br />

1983; Sargenti and Lancas 1997; Zamureenka<br />

et al. 1981


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 31<br />

table 2.1 (continued) major and minor constituents and Physiochemical Properties<br />

of essential oils obtained from cymbopogon species<br />

origin specific gravity η d [α] d solubility reference<br />

India 0.865–0.914 at 15° −0°10′ to 2°40′ Anonymous 1950<br />

Belgium 0.8847 at 20° 1.4849 at 20° −0°18′ at 25° Neyberg 1953<br />

Citral content 71.3%<br />

Odakali 0.8986 at 20° 1.4910 at 20° −0.62° at 20° 75% alcohol Thapa et al. 1981<br />

Acid value 5.34; ester value 44.2<br />

Cymbopogon pendulus (nees ex steud.) Wats.<br />

Major terpenes—citral-a or geranial (30%–50%), citral-b or neral<br />

(20%–35%), geranyl acetate (3%–5%), β-caryophyllene (2.1%),<br />

elemol (2.2%), geraniol (2%–6%), and linalool (3.0%)<br />

Traces—camphene, ∆ 3 -carene, caryophyllene oxide, citronellal,<br />

citronellyl acetate, p-cymene, dipentene, β-elemene, methyl<br />

heptenone, myrcene, β-phellandrene, α-pinene, and β-pinene<br />

Atal and Bradu 1976a; Balyan et al. 1979; Gulati<br />

and Garg 1976; Manjoor-i-Khuda et al. 1984,<br />

1986; Nigam et al. 1975; Pino et al. 1996;<br />

Rajendrudu and Rama Das 1983; Sobti et al.<br />

1982; Thapa et al. 1981; Thapa and Agarwal 1989<br />

origin specific gravity ηd [α] d solubility reference<br />

Haldwani India 0.9002–0.9152 1.4905–1.4890 −0.36° Soluble in 1.9 Gulati et al. 1976<br />

Aldehyde content 88.73%<br />

vol. of alcohol<br />

Cymbopogon winterianus Jowitt<br />

Major terpenes—geraniol (20%–25%), citronellol (4%–10%),<br />

citronellal (30%–45%), caryophyllene (2.1%), citronellyl acetate<br />

(3.0%), elemol (6.0%), geranyl acetate (4.2%), linalyl acetate (2.0%),<br />

methyl-iso-eugenol (2.3%), and nerol (7.7%)<br />

Traces—borneol, cadinene, l-cadinol, l-camphene, 1-carvone, citral,<br />

citronellyl butyrate, cymbopol, dipentene, eugenol, farnesol, geranyl<br />

formate, l-limonene, linalool, methyl heptenone, methyl eugenol,<br />

α-pinene, sesquicitronellene, terpinene, terpinen-4-1, and thujyl alcohol<br />

Baslas 1970; Iruthayathas et al. 1977; Pino et al.<br />

1996; Razdan and Koul 1973; Siddiqui et al.<br />

1975; Wijesekera et al. 1973a, 1973b<br />

Anonymous 1973; Chiang et al. 1981; Ganguly<br />

et al. 1979; Kaul et al. 1977; Liu et al. 1981;<br />

Singh et al. 1970; Sobti et al. 1982.<br />

origin specific gravity ηd [α] d solubility reference<br />

ISI 0.8710–0.8870 at 30° 1.4610–1.4700<br />

at 30°<br />

−0°30′–6° Anonymous 1952<br />

Nilgiri Hills<br />

Total geraniol 85%–97%<br />

0.900–0.929 +2°11′ to<br />

+12°12′<br />

Anonymous 1950<br />

Wild 0.885–0.901 at 15° 1.463–1.475 at 20° −4° to +1°47′ Soluble in 1–2 Gildemeister and<br />

Total geraniol content 85%–96%; citronellol 25%–54% vol. of alcohol Hoffmann 1956;<br />

Guenther 1950<br />

Java 0.897 at 27° 1.4654 at 27.5° −2°34′ at 28° Soluble in Guenther 1948,<br />

Total geraniol 88.8%; citronellol content 42.7%<br />

2.5–7.5 vol. of<br />

70% alcohol<br />

1950<br />

Java 0.887–0.895 at15° 1.4685–1.4728 at −0°35′ to 5°6′ Soluble in 1–2 Guenther 1950<br />

20°<br />

vol. of 80%<br />

Total geraniol 82.3%–89.4%; citronellol content 28.8%–43.9% alcohol<br />

Java 0.900–0.920 1.479–1.494 at 20° −7° to +22° Soluble in 80% Chakrabarti and<br />

alcohol<br />

Ghosh 1974<br />

(continued on next page)


32 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.1 (continued) major and minor constituents and Physiochemical Properties<br />

of essential oils obtained from cymbopogon species<br />

origin specific gravity η d [α] d solubility reference<br />

Pantnagar 0.9009 1.471 Baslas 1970<br />

Total geraniol content 85.4%; citronellol 32.47%<br />

Kumaon 0.9095 1.471 Baslas 1968<br />

Total geraniol content 83.2%; citronellol 32%<br />

West Bengal 0.912 1.470 at 35° −14°5′ Soluble in 80%<br />

alcohol<br />

Bangalore 0.8870 at 24° 1.4660 at 24° −0°30′ Soluble in 1 vol.<br />

Total geraniol content 65%<br />

Jorhat 0.890–0.892 at 25° 1.4650–1.4714 at<br />

25°<br />

of 80% alcohol<br />

−3°3′ Soluble in 1.5<br />

vol. of 80%<br />

alcohol<br />

Total geraniol content 85 to 90%<br />

Lucknow 0.875 at 25° 1.462 at 20° −1°2′ Soluble in 1−2<br />

Total geraniol content 96%<br />

Cymbopogon nardus (l.) rendle<br />

Major terpenes—limonene (9%–28%), geraniol (20%–30%), citronellol<br />

(8%–20%), citronellal (5%–16%), caryophyllene (1.4%), camphene<br />

(5%–6%), citral (18%), citronellol (20%), geranyl acetate (7%–8%),<br />

linalool (8.0%), methyl eugenol (4.1%), α-phellandrene (16.2%), β-pinene<br />

(15%), sesquicitronellene (5%–6%), and thujyl alcohol (8%–30%)<br />

Traces—α -bergamotene, l-borneol, citronellyl acetate, citronellyl<br />

butyrate, dipentene, elemol, ethyl iso-eugenol, eugenol, farnesol,<br />

n-heptyl alcohol, iso-pulegol, iso-valeraldehyde, methyl heptenone,<br />

nerol, cis-ocimene, pelargonaldehyde, and tricyclene<br />

vol. of 80%<br />

alcohol<br />

Chakrabarti and<br />

Ghosh 1974<br />

Virmani and Dutta<br />

1971<br />

Virmani and Dutta<br />

1971<br />

Virmani and Dutta<br />

1971<br />

Bruns et al. 1981; Guenther 1950; Gupta and<br />

Chauhan 1970; Krishnarajah et al. 1985;<br />

Manjoor-i-Khuda et al. 1984; Opdyke 1976;<br />

Razdan 1984<br />

Bruns et al. 1981; Gulati and Sadgopal 1972;<br />

Herath et al. 1979; Lucius and Adler 1971;<br />

Thieme et al. 1980; Wijesekera et al. 1973a,<br />

1973b<br />

origin specific gravity ηd [α] d solubility reference<br />

ISI 0.8870–0.9080 at 30°<br />

Total alcohol 55%–65%<br />

1.4745–1.4805<br />

at 30°<br />

−9° to −18° Anonymous 1952<br />

Haldwani 0.9233 1.4820 at 25° +4.06° Gulati and<br />

Acid value 6.5; ester value 28.3<br />

Sadgopal 1972<br />

Nainital 0.8632 at 20° 1.479 at 20° Gupta and<br />

Acid value 2.83; ester value 26.38<br />

Chauhan 1970<br />

Lucknow 0.895 at 30° 1.478 at 30° −12° Sharma et al. 1972<br />

0.900–0.920 1.479–1.494<br />

at 20°<br />

Ceylon 0.899–0.908 at 15° 1.4792–1.4842<br />

at 20°<br />

Ceylon 0.898–0.908 at 15° 1.4785–1.4900<br />

at 20°<br />

Java 0.885–0.900 at 15.5° 1.465–1.473<br />

at 20°<br />

−7° to −22° Soluble in 1–2<br />

vol. of 80%<br />

alcohol<br />

−9°40′ to −14°40′<br />

at 20°<br />

Soluble in 1 vol.<br />

of 80% alcohol<br />

Gildemeister and<br />

Hoffmann 1956;<br />

Guenther 1950<br />

Guenther 1950<br />

−7° to −14° Anonymous 1950<br />

−5° to +1° at 20° Soluble in 3 vol.<br />

of 80% alcohol<br />

Anonymous 1950


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 33<br />

table 2.1 (continued) major and minor constituents and Physiochemical Properties<br />

of essential oils obtained from cymbopogon species<br />

Cymbopogon schoenanthus (l.) spreng subsp. proximus hochst.<br />

Major terpenes—piperitone (80% amongst monoterpenes), elemol (39%),<br />

eudesmol (20%) amongst sesquiterpene alcohol, cis-carveol (4.8%), citral-a<br />

(2.4%), citral- b (3.3%), dihydrocarveol (35%), limonene (3.12%), linalool (21.6%)<br />

Traces—α-pinene, β-elemene, β-selinene, calamenene, cadalene p hydroxycinnamic<br />

acid, and several sesquiterpene alcohols<br />

Dawidar et al. 1990; Elagamal and Wolff<br />

1987; El Tawil and El Beih 1982;<br />

Modawi et al. 1984<br />

Ahmed et al. 1970; Evans et al. 1982;<br />

Shahi et al. 1990; Siddiqui et al. 1980<br />

origin specific gravity ηd [α] d solubility reference<br />

0.9169 at 15°C 1.4831 at 20° −59°22′ Soluble in 0.8 vol. Gildemeister and<br />

Acid value 6.0; alcohol content 36%−60%<br />

of 70% alcohol Hoffmann 1956<br />

Cymbopogon caesius (nees) stapf<br />

Major terpenes—Perillyl alcohol (25.6%), geraniol (19.8%), limonene (7.2%),<br />

citronellol (6.8%), and citronellal (6.7%)<br />

Trace—Carvone (30%) Liu et al. 1981<br />

α-Thujene, α-pinene, terpinolene, linalool, isopulegol, borneol, terpineol,<br />

Kanjilal et al. 1995<br />

geraniol, bornyl acetate, eugenol, citronellyl acetate, geranyl acetate,<br />

β-caryophyllene, perillaldelyde, caryophyllene oxide, elemol, and guaiol<br />

origin specific gravity η d [α] d solubility reference<br />

Bangalore 0.9267–0.9339 at<br />

15°C<br />

1.484 to 1.4856 at<br />

25°<br />

−18.3° to −5.6° at<br />

25°<br />

Acid value 0.9–2.5; saponification value 13.2–24.0; sap. value after<br />

acetylation 15.0–164.0<br />

Cymbopogon coloratus (nees) stapf.<br />

Monoterpenes—myrcene, limonene, trans-β-ocimene, linalool, neral,<br />

geranial, geraniol (69.11%), geranyl acetate, and elemol<br />

Soluble in<br />

70% alcohol<br />

Gildemeister<br />

and Hoffmann<br />

1956<br />

Mallavarapu et al.<br />

1992<br />

origin specific gravity η d [α] d solubility reference<br />

Malabar<br />

District<br />

0.911–0.920 at<br />

15°C<br />

NA −7°43′ to −10°<br />

20′ at 25°<br />

Cymbopogon densiflorus (steud.) stapf.<br />

Monoterpenes—Flower-limonene (52.1%), trans-p-menth-2,8-dien-1-ol<br />

(10%), verbenol (9.7%), and perillyl alcohol (7.2%)<br />

Monoterpenes—Leaf—trans-p-mentha-2,8-dien-1-ol (22.4%), verbenol<br />

(18%), perillyl alcohol (17.2%), and cis-p-mentha-1-(7)-dien-2-ol (11.1%)<br />

Soluble in 1 vol.<br />

of 80% alcohol<br />

Gildemeister and<br />

Hoffmann 1956<br />

Boelens 1994;<br />

Chisowa 1997<br />

origin specific gravity ηd [α] d solubility reference<br />

Congo 0.9304 at 15°C 1.4683 at 20° +59°30′ Soluble in 0.5 vol. of Gildemeister and<br />

Sap. value 2.1; ester value 19.6; ester value 8.42 after acetylation 80% alcohol Hoffmann 1956<br />

Cymbopogon distans (nees) Wats.<br />

Major terpenes—terpineol (20%) Sobti et al. 1978<br />

Piperitone (30%–40%) and geraniol (10%) Liu et al. 1981;<br />

Thapa et al. 1971<br />

Limonene (29%) and methyl eugenol (13%), β-bisabolene (5.4%), α-bisabolol<br />

(3.0%), bornyl acetate (4.8%), γ-cadinene (3.6%), caryophyllene (4.7%),<br />

d-citronellal (4.0%), p-cymene (5.1%), farnesol (5.1%), d-menthone (10.4%),<br />

geranyl acetate (10%–12%), α-humulene (3.5%), limonene (5.8%–29.0%),<br />

methyl eugenol (13.4%), α-phellandrene (2.3%–6.0%), and α-pinene (3.5%)<br />

Singh and Sinha<br />

1976<br />

(continued on next page)


34 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.1 (continued) major and minor constituents and Physiochemical Properties<br />

of essential oils obtained from cymbopogon species<br />

Traces—amorphine, γ-α-bergamotene, borneol, d-cadinene, camphene,<br />

1-carbomenthone, citronellol, β-farnesene (a keto compound), β-muurolene,<br />

myrcene, neryl propionate, octanol β-phellandrene, β-pinene, sabinene,<br />

β-selinene, α-terpinene, γ-terpinene, and terpinolene<br />

material was dissolved. Often the temperature is not specified; in these cases it is assumed to be<br />

room temperature. The formal unit for specific rotation values is deg cm 2 g −1 , but scientific literature<br />

uses just degrees. A negative value means levorotatory rotation, and a positive value means dextrorotatory<br />

rotation.<br />

Optical rotation is measured with an instrument called a polarimeter. There is a linear relationship<br />

between the observed rotation and the concentration of optically active compound in the<br />

sample. There is a nonlinear relationship between the observed rotation and the wavelength of light<br />

used. Specific rotation is calculated using either of two equations, depending on the sample you are<br />

measuring. For pure liquids,<br />

[α] λ T = α/l × d<br />

In this equation, l is the path length in decimeters, and d is the density of the liquid in g/mL, for a<br />

sample at a temperature T (given in degrees Celsius) and wavelength λ (in nanometers). If the wavelength<br />

of the light used is 589 nanometer (the sodium D line), the symbol “D” is used. The sign of<br />

the rotation (+ or −) is always given: [α] D 20 = +6.2°. For solutions, a different equation is used:<br />

[α] λ T = 100α/l × d<br />

Balyan et al. 1979; Gupta and Daniel<br />

1982; Liu et al. 1981; Mathela et al.<br />

1988; Mathela and Joshi 1981; Mathela<br />

et al. 1990a; Melkani et al. 1985; Singh<br />

and Sinha 1976; Sobti et al. 1978c<br />

origin specific gravity ηd [α] d solubility reference<br />

Nainital 0.801 — — NA Mathela and Joshi 1981<br />

Acid value 1.15; ester value after acetylation 80.95<br />

Cymbopogon nervatus (hohst.) chiov.<br />

Terpenes—β-selinene, β-elemene, β-bergamotene, and germacrene-D Modawi et al.<br />

1984<br />

origin specific gravity η d [α] d solubility reference<br />

Kordofan Sudan 0.9405 at 15° 1.4946 at 20° +26°22′ Soluble in 0.5 vol.<br />

of 80% alcohol<br />

Ester value 9.3, ester value after acetylation 99.1<br />

Gildemeister and<br />

Hoffmann 1956<br />

When using this equation, the concentration and the solvent are always provided in parentheses<br />

after the rotation. The rotation is reported using degrees, and no units of concentration are given (it<br />

is assumed to be g/100 mL).<br />

Solubility—Solubility is a characteristic physical property referring to the ability of a given substance,<br />

the solute, to dissolve in a solvent. It is measured in terms of the maximum amount of solute<br />

dissolved in a solvent at equilibrium. The resulting solution is called a saturated solution. Certain<br />

liquids are soluble in all proportions with a given solvent, such as ethanol in water. This property is<br />

known as miscibility. Under certain conditions the equilibrium solubility can be exceeded to give a<br />

so-called supersaturated solution, which is metastable.


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 35<br />

table 2.2 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

acyclic monoterpene hydrocarbons<br />

1 h-nmr 13 c-nmr 1 h-nmr 13 c-nmr<br />

trans-Ocimene—(4Z,6E)-2,6-dimethylocta-2,4,6-triene; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

cis-Ocimene—(4Z,6E)-2,6-dimethylocta-2,4,6-triene; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

H 5.44<br />

1.71<br />

22.4<br />

1.71<br />

5.43 H<br />

16.4<br />

135.2<br />

135.1<br />

12.8<br />

135.1<br />

H 6.51<br />

6.44 H<br />

124.6<br />

135.2<br />

124.6<br />

1.71<br />

1.71<br />

12.8<br />

130.4<br />

125.4<br />

130.4<br />

5.99 H<br />

H 5.99<br />

125.4<br />

H 6.51<br />

6.44 H<br />

25.7<br />

19.7 143.3<br />

19.7<br />

143.3<br />

25.7<br />

1.71<br />

1.71<br />

1.71<br />

1.71<br />

β-Phellandrene—3-isopropyl-6-methylenecyclohex-1-ene; chemical<br />

chemical formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z:<br />

136.13 (100.0%), 137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

α-Phellandrene—5-isopropyl-2-methylcyclohexa-1,3-diene; chemical chemical<br />

formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13<br />

(100.0%), 137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

1.71<br />

109.7<br />

H 4.80<br />

4.88 H<br />

22.3<br />

131.0<br />

H 5.58<br />

5.97 H<br />

34.4<br />

143.9<br />

124.2<br />

125.6<br />

5.90<br />

2.01;1.91<br />

124.2<br />

28.1<br />

41.9<br />

134.1<br />

1.46;1.21<br />

31.1<br />

2.15<br />

134.1 40.3<br />

2.23;1.98<br />

5.80 2.13<br />

5.47 H<br />

32.0<br />

1.86<br />

32.0<br />

1.86<br />

21.1 21.1<br />

1.01 1.01<br />

21.2 21.2<br />

1.01 1.01<br />

(continued on next page)


36 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

acyclic oxygenated monoterpenes<br />

Citral-a-(Z)-3,7-dimethylocta-2,6-dienal; chemical formula: C10H16O; exact Citral-b-(E)-3,7-dimethylocta-2,6-dienal; chemical formula: C10H16O; exact<br />

mass: 152.12; molecular weight: 152.23, m/z: 152.12 (100.0%), 153.12<br />

mass: 152.12; molecular weight: 152.23, m/z: 152.12 (100.0%), 153.12<br />

(10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

(10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

9.68 H<br />

O<br />

(Z)<br />

2.00<br />

1.71<br />

2.00<br />

5.77<br />

H 5.20<br />

O<br />

23.0<br />

5.77 H 9.68<br />

164.5<br />

(Z)<br />

(E)<br />

33.2<br />

127.7<br />

2.00<br />

1.71<br />

26.1 191.1<br />

123.5 O<br />

2.00 H 5.20<br />

17.0<br />

164.5 191.1<br />

39.2 (E)<br />

127.7<br />

O<br />

26.1<br />

123.5<br />

132.0<br />

19.6 25.6<br />

132.0<br />

19.6 25.6<br />

1.71 1.71<br />

1.71 1.71


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 37<br />

Citronellal—3,7-dimethyloct-6-enal; chemical formula: C10H18O; exact mass:<br />

154.14; molecular weight: 154.25, m/z: 154.14 (100.0%), 155.14 (11.1%);<br />

elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

Citronellol—3,7-dimethyloct-6-en-1-ol; chemical formula: C10H20O; exact<br />

mass: 156.15; molecular weight: 156.27, m/z: 156.15 (100.0%), 157.15<br />

(10.8%); elemental analysis: C, 76.86; H, 12.90; O, 10.24<br />

1.06<br />

20.6<br />

O<br />

4.78<br />

OH<br />

1.06<br />

21.1<br />

202.1<br />

27.6<br />

9.72<br />

1.88<br />

60.3<br />

29.3<br />

3.53<br />

1.65<br />

1.29<br />

O<br />

37.5<br />

51.0<br />

2.48;2.23<br />

OH<br />

38.0<br />

1.44<br />

1.29<br />

39.9<br />

24.1<br />

H 5.20<br />

1.96<br />

24.4<br />

126.8<br />

H 5.20<br />

1.96<br />

126.8<br />

131.3<br />

19.6 25.6<br />

131.3<br />

19.6 25.6<br />

1.71 1.71<br />

1.71 1.71<br />

Citronellyl butyrate—8-butoxy-2,6-dimethyloct-2-ene; chemical formula:<br />

C14H28O; exact mass: 212.21; molecular weight: 212.37, m/z: 212.21<br />

(100.0%), 213.22 (15.5%), 214.22 (1.3%); elemental analysis: C, 79.18; H,<br />

13.29; O, 7.53<br />

Citronellyl acetate—3,7-dimethyloct-6-enyl acetate; chemical formula:<br />

C12H22O2; exact mass: 198.16; molecular weight: 198.3, m/z: 198.16 (100.0%),<br />

199.17 (13.3%), 200.17 (1.2%); elemental analysis: C, 72.68; H, 11.18; O,<br />

16.14<br />

14.1<br />

0.96<br />

2.01<br />

19.0<br />

32.2<br />

1.33<br />

O<br />

1.46<br />

20.8<br />

29.5<br />

O O<br />

1.06<br />

62.4<br />

72.1<br />

O<br />

69.9<br />

21.1<br />

1.06 3.37<br />

O<br />

1.65<br />

3.37<br />

1.42<br />

H 5.20<br />

170.2 20.7<br />

O<br />

36.6<br />

37.7<br />

4.08<br />

1.65<br />

29.6<br />

24.4<br />

1.53<br />

H 5.20<br />

1.29<br />

38.0<br />

37.7<br />

1.29<br />

126.8<br />

24.4<br />

1.96<br />

126.8<br />

1.96<br />

19.6 131.3 25.6<br />

19.6 131.3 25.6<br />

1.71 1.71<br />

1.71 1.71<br />

(continued on next page)


38 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

Lavandulol—5-methyl-2-(prop-1-en-2-yl)hex-4-en-1-ol; chemical formula:<br />

C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

Geraniol—(E)-3,7-dimethylocta-2,6-dien-1-ol; chemical formula: C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14 (100.0%), 155.14<br />

(11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

107.5<br />

5.56<br />

OH<br />

H 4.88<br />

4.63 H<br />

5.39 H<br />

65.2<br />

50.0<br />

3.65;3.40<br />

2.29<br />

4.78<br />

2.09;1.84<br />

HO<br />

4.20<br />

21.5<br />

149.1<br />

HO<br />

1.71<br />

25.7<br />

1.71<br />

2.00<br />

124.2<br />

H 5.20<br />

5.20 H<br />

2.00<br />

131.4<br />

19.6 25.6<br />

1.71 1.71<br />

1.71<br />

1.71<br />

Geranyl formate—(Z)-3,7-dimethylocta-2,6-dienyl formate; chemical<br />

formula: C11H18O2; exact mass: 182.13; molecular weight: 182.26, m/z:<br />

182.13 (100.0%), 183.13 (12.0%); elemental analysis: C, 72.49; H, 9.95; O,<br />

17.56<br />

Geranyl acetate—(Z)-3,7-dimethylocta-2,6-dienyl acetate; chemical formula:<br />

C12H20O2; exact mass: 196.15; molecular weight: 196.29, m/z: 196.15 (100.0%),<br />

197.15 (13.3%), 198.15 (1.2%); elemental analysis: C, 73.43; H, 10.27; O,<br />

16.30<br />

23.2<br />

O<br />

23.2<br />

1.71<br />

4.75<br />

5.39 H<br />

135.7<br />

135.7<br />

2.01<br />

O<br />

123.0<br />

33.4<br />

H<br />

1.71<br />

5.20<br />

1.71<br />

O<br />

123.0<br />

33.4<br />

160.7<br />

4.83<br />

2.00<br />

60.4<br />

123.5 O 170.2 20.8<br />

H 5.20<br />

1.71<br />

2.00<br />

26.4<br />

O<br />

8.04<br />

O<br />

26.4<br />

O O<br />

62.9<br />

123.5<br />

2.00<br />

2.00<br />

H<br />

5.39<br />

19.6 132.0 25.6<br />

19.6 132.0 25.6<br />

1.71<br />

1.71


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 39<br />

Linalyl acetate—3,7-dimethylocta-1,6-dien-3-yl acetate, chemical formula:<br />

C12H20O2; exact mass: 196.15; molecular weight: 196.29, m/z: 196.15<br />

(100.0%), 197.15 (13.3%), 198.15 (1.2%); elemental analysis: C, 73.43; H,<br />

10.27; O, 16.30<br />

Linalool—3,7-dimethylocta-1,6-dien-3-ol, chemical formula: C10H18O; exact<br />

mass: 154.14; molecular weight: 154.25, m/z: 154.14 (100.0%), 155.14<br />

(11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

28.7<br />

5.23<br />

H<br />

25.4<br />

OH<br />

170.2 21.4<br />

O<br />

1.71<br />

73.2<br />

H 5.89<br />

82.9<br />

2.01<br />

145.7<br />

O<br />

112.6<br />

O<br />

H<br />

1.60<br />

5.20<br />

145.7<br />

42.9<br />

OH 2.0<br />

5.24 H<br />

39.6<br />

O<br />

1.71<br />

18.5<br />

H 5.89<br />

1.96<br />

1.41<br />

1.48<br />

112.6<br />

18.3<br />

126.8<br />

1.57<br />

126.8<br />

H 5.20<br />

1.96<br />

H 5.24<br />

5.23 H<br />

19.6<br />

131.3<br />

25.6<br />

131.3<br />

19.6 25.6<br />

1.71 1.71<br />

Neryl acetate—(E)-3,7-dimethylocta-2,6-dien-1-ol; chemical formula:<br />

C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

Nerol—(Z)-3,7-dimethylocta-2,6-dien-1-ol; chemical formula: C10H18O; exact<br />

mass: 154.14; molecular weight: 154.25, m/z: 154.14 (100.0%), 155.14<br />

(11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.3<br />

19.7<br />

17.5<br />

1.71<br />

120.0<br />

2.01<br />

1.71<br />

58.9<br />

135.7<br />

H 5.20<br />

1.71<br />

129.7<br />

29.9<br />

OH<br />

123.0<br />

39.7<br />

O O<br />

H<br />

1.71<br />

5.20<br />

1.71<br />

2.00<br />

20.8<br />

2.00<br />

168.0<br />

O<br />

123.5<br />

26.1<br />

4.75<br />

2.00<br />

26.4<br />

H 5.39<br />

2.00<br />

123.5<br />

H 5.39<br />

1.71<br />

O<br />

19.6<br />

132.0<br />

25.6<br />

4.20<br />

19.6 132.0 25.6<br />

5.56 HO<br />

(continued on next page)


40 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

6-methylhept-5-en-2-one—chemical formula: C8H14O; exact mass: 126.1;<br />

molecular weight: 126.2, m/z: 126.10 (100.0%), 127.11 (8.9%); elemental<br />

analysis: C, 76.14; H, 11.18; O, 12.6<br />

2.09<br />

31.0<br />

207.7<br />

O<br />

44.8<br />

O<br />

2.49<br />

22.2<br />

H 5.20<br />

2.24<br />

123.5<br />

132.0<br />

19.6 25.6<br />

1.71 1.71<br />

65<br />

cyclic monoterpene hydrocarbons<br />

p-Cymene—chemical formula: C10H14; exact mass: 134.11; molecular weight: o-Cymene—chemical formula: C10H14; exact mass: 134.11; molecular weight:<br />

134.22, m/z: 134.11 (100.0%), 135.11 (10.8%); elemental analysis: C, 89.49; H, 134.22, m/z: 134.11 (100.0%), 135.11 (10.8%); elemental analysis: C, 89.49;<br />

10.51<br />

H, 10.51<br />

2.19<br />

24.3<br />

125.9<br />

7.06<br />

135.6<br />

125.5<br />

128.7<br />

6.99<br />

6.98<br />

6.98<br />

6.98<br />

128.7<br />

128.7<br />

7.01<br />

126.0<br />

7.19 7.19<br />

126.0<br />

126.0<br />

133.9<br />

2.35<br />

146.2 22.4<br />

145.5<br />

3.12<br />

33.8<br />

2.86<br />

36.3<br />

1.29 1.29<br />

1.19 1.19<br />

23.3 23.3<br />

23.6 23.6


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 41<br />

(−)-Limonene—(S)-1-methyl-4-(prop-1-en-2 yl)cyclohex-1-ene; chemical<br />

formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13<br />

(100.0%), 137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

(+)-Limonene—(R)-1-methyl-4-(prop-1-en-2 yl)cyclohex-1-ene; chemical<br />

formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13<br />

(100.0%), 137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

1.71<br />

23.1<br />

1.71<br />

23.1<br />

H 5.37<br />

H 5.37<br />

133.9<br />

2.01;1.91<br />

133.9<br />

2.01;1.91<br />

123.3<br />

31.0<br />

123.3<br />

31.0<br />

1.77;1.52 2.33 2.09;1.84<br />

2.09;1.84<br />

1.77;1.52 2.33<br />

31.1<br />

31.1<br />

42.3<br />

28.1<br />

42.3<br />

28.1<br />

H 4.63<br />

H 4.63<br />

149.1<br />

1.71<br />

149.1<br />

1.71<br />

21.5 107.5<br />

4.88<br />

H<br />

21.5 107.5<br />

H 4.88<br />

α-Terpinene—1-isopropyl-4-methylcyclohexa-1,3-diene; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

Myrcene—7-methyl-3-methyleneocta-1,6-diene; chemical formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%), 137.13<br />

(11.0%); elemental analysis: C, 88.16; H, 11.84<br />

23.2<br />

1.71<br />

116.1<br />

H 5.02<br />

138.3<br />

145.3<br />

H 5.68<br />

H 6.25<br />

143.9<br />

5.16 H<br />

120.1<br />

31.7<br />

2.15<br />

110.4<br />

38.2<br />

H 4.88<br />

115.5<br />

26.1<br />

2.15<br />

2.00<br />

26.5<br />

142.7<br />

H 5.68<br />

123.5<br />

H<br />

4.80<br />

2.00<br />

5.20 H<br />

34.5<br />

2.52<br />

21.9 21.9<br />

19.6 132.0 25.6<br />

1.11 1.11<br />

1.71<br />

1.71<br />

(continued on next page)


42 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

Terpinolene—1-methyl-4-(propan-2-ylidene) cyclohex-1-ene; chemical<br />

formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13<br />

(100.0%), 137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

γ-Terpinene—1-isopropyl-4-methylcyclohexa-1,4-diene; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

23.2<br />

1.71<br />

23.2<br />

1.71<br />

132.1<br />

H 5.21<br />

131.5<br />

H 5.16<br />

120.5<br />

31.7<br />

2.00<br />

121.3<br />

37.8<br />

2.63<br />

30.2<br />

125.8<br />

28.1<br />

2.63<br />

2.00<br />

32.2<br />

116.7<br />

2.63<br />

140.7<br />

5.16 H<br />

34.5<br />

2.52<br />

19.5 125.3 19.5<br />

21.9 21.9<br />

1.11 1.11<br />

1.71 1.71<br />

bicyclic monoterpene hydrocarbons<br />

∆4-Carene —3,7,7-trimethylbicyclo[4.1.0]hept-2-ene; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

∆3-Carene—3,7,7-trimethylbicyclo[4.1.0]hept-3-ene; chemical formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%), 137.13<br />

(11.0%); elemental analysis: C, 88.16; H, 11.84<br />

23.4<br />

1.71<br />

23.4<br />

1.71<br />

133.9<br />

133.9<br />

H 5.37<br />

5.37 H<br />

124.2<br />

30.9<br />

36.4<br />

2.01;1.91<br />

123.3<br />

2.04;1.79<br />

30.3<br />

25.3<br />

0.87<br />

1.74;1.49<br />

25.5<br />

30.3<br />

0.22<br />

2.04;1.79<br />

36.4<br />

0.22<br />

28.3<br />

0.22<br />

27.7<br />

24.9<br />

1.11<br />

27.4<br />

23.6<br />

1.11<br />

27.7<br />

1.11<br />

1.11<br />

27.4


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 43<br />

α-Pinene—2,6,6-trimethylbicyclo[3.1.1]hept-2-ene; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

21.2<br />

1.71<br />

Camphene—2,2-dimethyl-3 methylenebicyclo[2.2.1]heptane; chemical<br />

formula: C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13<br />

(100.0%), 137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

24.8<br />

1.21<br />

48.4<br />

1.52<br />

135.7<br />

42.3<br />

H 5.37<br />

1.60;1.35<br />

31.5<br />

1.21<br />

H 4.88<br />

1.60;1.35<br />

123.3<br />

48.9<br />

2.62<br />

24.8<br />

39.6<br />

1.63;1.38<br />

25.2<br />

41.5<br />

1.11<br />

2.04;1.79<br />

1.11<br />

25.2<br />

2.04;1.79<br />

34.5<br />

2.19<br />

30.3<br />

31.3<br />

165.0<br />

1.97<br />

109.1<br />

46.6<br />

H 4.63<br />

41.2<br />

Sabienene—1-isopropyl-4 methylenebicyclo[3.1.0]hexane; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

β-Pinene—6,6-dimethyl-2 methylenebicyclo[3.1.1]heptane; chemical formula:<br />

C10H16; exact mass: 136.13; molecular weight: 136.23, m/z: 136.13 (100.0%),<br />

137.13 (11.0%); elemental analysis: C, 88.16; H, 11.84<br />

109.1<br />

4.76 H H 4.62<br />

109.1<br />

4.88 H H 4.63<br />

35.4<br />

153.5<br />

37.4<br />

0.88<br />

2.01;1.91<br />

37.9<br />

148.0<br />

52.9<br />

20.6<br />

43.9<br />

2.01;1.91<br />

41.5<br />

2.62<br />

1.38;1.13 0.27;0.02<br />

25.2<br />

25.2<br />

1.11<br />

43.1<br />

1.11<br />

32.3<br />

31.3<br />

1.41;1.16<br />

2.04;1.79<br />

33.3<br />

1.81<br />

43.6<br />

1.93<br />

18.6 18.6<br />

1.01 1.01<br />

(continued on next page)


44 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

oxygenated monoterpenes<br />

Carvacrol—(R)-2-methyl-5-(prop-1-en-2-yl)cyclohex-1-enol; chemical cis-Carveol—2-methyl-5-(prop-1-en-2-yl)cyclohex-2-enol; chemical formula:<br />

formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12 C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12,<br />

(100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51 (100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

16.9<br />

1.71<br />

1.71<br />

(Z)<br />

136.3<br />

OH<br />

71.9<br />

5.37 3.69<br />

H (Z) OH 4.14<br />

OH 16.77<br />

2.0;1.9<br />

123.3<br />

2.09;1.84 2.33 1.92;1.67<br />

2.1;1.8<br />

1.8;1.5 2.3<br />

35.2<br />

31.4<br />

36.1<br />

H 4.63<br />

H 4.88<br />

149.1<br />

1.71<br />

1.71<br />

21.5 107.5<br />

H 4.88<br />

H 4.63


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 45<br />

Carvotanacetone—5-isopropyl-2-methylcyclohex-2-enone; chemical<br />

formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12<br />

(100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

Carveyl acetate—2-methyl-5-(prop-1-en-2-yl) cyclohex-2-enyl acetate;<br />

chemical formula: C12H18O2; exact mass: 194.13; molecular weight: 194.27,<br />

m/z: 194.13 (100.0%), 195.13 (13.1%); elemental analysis: C, 74.19; H, 9.34;<br />

O, 16.47<br />

15.9<br />

1.93<br />

1.71<br />

17.1<br />

135.4<br />

O<br />

2.01<br />

(Z) 4.43 O<br />

5.37 H<br />

137.3<br />

O H 6.35<br />

170.2 21.1<br />

O<br />

198.9<br />

75.7<br />

136.3<br />

(Z)<br />

123.3<br />

O<br />

2.09;1.84 2.33<br />

2.01;1.76<br />

30.1<br />

45.9<br />

40.5<br />

O 3.03;2.78 1.73 2.04;1.79<br />

31.1 36.3 31.9<br />

H 4.63<br />

149.1<br />

21.5 107.5<br />

1.71<br />

1.82<br />

31.6<br />

1.01 1.01<br />

H<br />

4.88<br />

20.8 20.8<br />

Eucarvone—(2Z,4Z)-2,6,6-trimethylcyclohepta-2,4-dienone; chemical<br />

formula: C10H14O; exact mass: 150.1; molecular weight: 150.22, m/z: 150.10<br />

(100.0%), 151.11 (11.0%); elemental analysis: C, 79.96; H, 9.39; O, 10.65<br />

Carvone—2-methyl-5-(prop-1-en-2-yl)cyclohex-2-enone; chemical formula:<br />

C10H14O; exact mass: 150.1; molecular weight: 150.22, m/z: 150.10 (100.0%),<br />

151.11 (11.0%); elemental analysis: C, 79.96; H, 9.39; O, 10.65<br />

1.93<br />

29.6<br />

15.9<br />

1.21<br />

1.21<br />

O<br />

135.5<br />

5.65 2.90<br />

198.9<br />

134.7<br />

H<br />

H 6.35<br />

O<br />

29.6<br />

151.0<br />

O<br />

2.09;1.84<br />

3.07;2.82 2.43<br />

O<br />

36.6<br />

59.8<br />

202.6<br />

137.5<br />

30.9<br />

43.2<br />

41.6<br />

119.8<br />

6.21 H<br />

1.93<br />

H 4.63<br />

16.3<br />

135.6<br />

H<br />

7.01<br />

1.71<br />

149.1<br />

21.2 107.5<br />

H<br />

4.88<br />

(continued on next page)


46 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

(−)-Dihdrocarveol—(1R,5R)-2-methyl-5-(prop-1-en-2-yl)cyclohexanol;<br />

chemical formula: C10H18O; exact mass: 154.14; molecular weight: 154.25,<br />

m/z: 154.14 (100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H,<br />

11.76; O, 10.37<br />

1,8-Cineole—1,3,3-trimethyl-2 oxabicyclo [2.2.2]octane; chemical formula:<br />

C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14 (100.0%),<br />

155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

14.5<br />

1.06<br />

1.31<br />

OH 4.81<br />

1.69<br />

OH<br />

40.1<br />

3.16<br />

1.52;1.27<br />

77.9<br />

27.8<br />

1.53<br />

1.65;1.40<br />

1.57;1.32 2.12 1.72;1.47<br />

O<br />

38.0<br />

40.1<br />

28.6<br />

1.52;1.27 1.74 1.40<br />

H 4.67<br />

1.71<br />

147.7<br />

H<br />

4.66<br />

1.26 1.26<br />

21.4 110.6<br />

6,7-Epoxy-3,7-dimethyl-1,3-octadiene—(Z)-2,2-dimethyl-3-(3-methylpenta-<br />

2,4-dienyl)oxirane; chemical formula: C10H16O; exact mass: 152.12; molecular<br />

weight: 152.23, m/z: 152.12 (100.0%), 153.12 (10.9%); elemental analysis: C,<br />

78.90; H, 10.59; O, 10.51<br />

22.3<br />

3,4-Epoxy-3,7-dimethyl-1,6-octadiene—2-methyl-3-(3-methylbut-2-enyl)-2vinyloxirane;<br />

chemical formula: C10H16O; exact mass: 152.12; molecular<br />

weight: 152.23, m/z: 152.12 (100.0%), 153.12 (10.9%); elemental analysis: C,<br />

78.90; H, 10.59; O, 10.5<br />

22.1<br />

1.71<br />

O<br />

1.41<br />

133.6<br />

60.2<br />

O<br />

H 6.25<br />

5.25 H<br />

138.3<br />

132.8<br />

145.7<br />

64.8<br />

H 5.89<br />

2.59<br />

116.1<br />

67.1<br />

26.3<br />

2.21;1.96<br />

H<br />

O 5.16<br />

112.6<br />

126.8<br />

26.3<br />

2.21;1.96<br />

5.20 H<br />

H 5.02<br />

2.55<br />

H 5.23<br />

5.24 H<br />

O<br />

58.8<br />

1.26<br />

1.26<br />

1.71<br />

1.71<br />

23.3 23.3<br />

131.3<br />

19.6 25.6


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 47<br />

cis-Isopiperitenol—(1S,6S)-3-methyl-6-(prop-1-en-2-yl)cyclohex-2-enol,<br />

chemical formula: C10H16O; exact mass: 152.12; molecular weight: 152.23,<br />

m/z: 152.12 (100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H,<br />

10.59; O, 10.51<br />

1.71<br />

23.4<br />

Isopiperitenone—(R)-3-methyl-6-(prop-1-en-2-yl)cyclohex-2-enone; chemical<br />

formula: C10H14O; exact mass: 150.1; molecular weight: 150.22, m/z: 150.10<br />

(100.0%), 151.11 (11.0%), elemental analysis: C, 79.96; H, 9.39; O, 10.65<br />

1.71<br />

133.9<br />

H 5.37<br />

H 5.85<br />

2.01;1.91<br />

2.01;1.91<br />

122.9<br />

31.3<br />

3.72<br />

1.77;1.52 2.41<br />

1.57;1.32 3.03<br />

71.3<br />

21.9<br />

OH 4.14<br />

O<br />

51.6<br />

OH<br />

H 4.63<br />

H 4.97<br />

1.71<br />

1.71<br />

149.1<br />

H<br />

4.88<br />

H 4.89<br />

21.8 107.5<br />

Isopulegol—(1S,2R,5S)-5-methyl-2-(prop-1-en-2-yl)cyclohexanol; chemical<br />

formula: C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

trans-Isopiperitenol—(1S,6R)-3-methyl-6-(prop-1-en-2-yl)cyclohex-2-enol;<br />

chemical formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z:<br />

152.12 (100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O,<br />

10.51<br />

21.0<br />

1.71<br />

1.06<br />

1.06<br />

H 5.37<br />

43.0<br />

28.4<br />

1.67;1.42<br />

1.61<br />

2.01;1.91<br />

34.3<br />

1.52;1.27<br />

1.67;1.42<br />

1.61<br />

1.52;1.27<br />

1.57;1.32 2.20 3.20<br />

3.72<br />

70.4<br />

1.77;1.52 2.41<br />

52.7<br />

22.1<br />

OH 4.81<br />

3.16<br />

1.52;1.27<br />

1.50<br />

OH 4.14<br />

OH<br />

H 4.67<br />

147.7<br />

1.71<br />

OH<br />

4.81<br />

H 4.63<br />

1.71<br />

21.7 110.6<br />

H<br />

4.66<br />

1.82<br />

1.01 1.01<br />

(continued on next page)<br />

H 4.88


48 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

∆3-Menthene-3-one—5-methylene-2-(prop-1-en-2-yl)cyclohexanone; chemical formula: C10H14O; exact mass: 150.1; molecular weight: 150.22,<br />

m/z: 150.10 (100.0%), 151.11 (11.0%); elemental analysis: C, 79.96; H, 9.39;<br />

O, 10.65<br />

Limonene oxide—(1R,4R,6S)-1-methyl-4-(prop-1-en-2-yl)-7-oxabicyclo[4.1.0]<br />

heptane; chemical formula: C10H16O; exact mass: 152.12; molecular weight:<br />

152.23, m/z: 152.12 (100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H,<br />

10.59; O, 10.51<br />

110.7<br />

4.89 H H 4.97<br />

20.7<br />

1.31<br />

O<br />

O<br />

2.01;1.91<br />

2.86<br />

1.65;1.40<br />

51.5<br />

147.4<br />

3.17;3.07<br />

64.6<br />

62.6<br />

37.8<br />

35.5<br />

1.74;1.49 3.03<br />

1.57;1.32 2.12 1.70;1.45<br />

207.1<br />

58.0<br />

25.2<br />

O<br />

31.4<br />

42.2<br />

24.2<br />

O<br />

4.97 H<br />

H 4.67<br />

1.71<br />

1.71<br />

143.7<br />

147.7<br />

20.9<br />

110.7<br />

H<br />

4.89<br />

110.6<br />

21.4<br />

H<br />

4.66<br />

C<br />

Menthyl acetate—(1R,2S,5R)-2-isopropyl-5-methylcyclohexyl acetate;<br />

chemical formula: C12H22O2; exact mass: 198.16; molecular weight: 198.3,<br />

m/z: 198.16 (100.0%), 199.17 (13.3%), 200.17 (1.2%); elemental analysis: C,<br />

72.68; H, 11.18; O, 16.14<br />

(−)-Menthol—(1R,2S,5R)-2-isopropyl-5 methylcyclohexanol; chemical<br />

formula: C10H20O; exact mass: 156.15; molecular weight: 156.27, m/z: 156.15<br />

(100.0%), 157.15 (10.8%); elemental analysis: C, 76.86; H, 12.90; O, 10.24<br />

21.0<br />

20.7<br />

1.06<br />

1.06<br />

42.9<br />

28.3<br />

34.2<br />

O<br />

39.6<br />

28.5<br />

33.9<br />

1.52;1.27 1.61 1.76;1.51<br />

O<br />

1.52;1.27 2.01 3.90<br />

O<br />

1.67;1.42<br />

1.61<br />

1.52;1.27<br />

75.3<br />

22.3 47.1<br />

72.2<br />

22.1 50.4<br />

1.52;1.27 1.50 3.16<br />

170.2<br />

O<br />

21.0<br />

OH<br />

OH 4.81<br />

2.01<br />

21.0 25.7 21.0<br />

1.01 1.82 1.01<br />

25.5<br />

1.01 1.82 1.01<br />

21.3<br />

21.3


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 49<br />

p-Menth-2-en-1-ol—(1R,4R)-4-isopropyl-1 methyl cyclohex-2-enol;<br />

chemical formula: C10H18O; exact mass: 154.14; molecular weight: 154.25,<br />

m/z: 154.14 (100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H,<br />

11.76; O, 10.37<br />

1.41<br />

OH 2.0<br />

28.6<br />

OH<br />

(−)-Menthone—(2S,5R)-2-isopropyl-5-methylcyclohexanone; chemical<br />

formula: C10H18O; exact mass: 154.14, Molecular Weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

20.2<br />

1.06<br />

132.4<br />

68.5<br />

37.4<br />

5.59<br />

49.6<br />

33.9<br />

1.88;1.63<br />

32.6<br />

2.31;2.06<br />

1.97<br />

1.88;1.63<br />

131.3<br />

42.1<br />

20.2<br />

1.74;1.49 1.94 5.59<br />

211.5<br />

59.3<br />

24.0<br />

1.88;1.63 2.20<br />

O<br />

O<br />

1.01<br />

1.86<br />

1.01<br />

25.5<br />

20.5 20.5<br />

1.01<br />

2.05<br />

1.01<br />

31.9<br />

21.1 21.1<br />

cis-p-Menth-1(7),8-dien-2-ol—(1S,5S)-2-methylene-5-(prop-1-en-2-yl)<br />

cyclohexanol; chemical formula: C10H16O; exact mass: 152.12; molecular<br />

weight: 152.23, m/z: 152.12 (100.0%), 153.12 (10.9%); elemental analysis: C,<br />

78.90; H, 10.59; O, 10.51<br />

p-Menth-8-en-1-ol—(1s,4s)-1-methyl-4-(prop-1-en-2-yl)cyclohexanol;<br />

chemical formula: C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z:<br />

154.14 (100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O,<br />

10.37<br />

4.96 H H 5.01<br />

1.31<br />

109.1<br />

27.3<br />

OH<br />

4.14<br />

OH 4.64<br />

OH<br />

OH<br />

1.67;1.42<br />

1.67;1.42<br />

3.90<br />

2.01;1.91<br />

42.1<br />

71.0<br />

75.8<br />

149.3<br />

31.8<br />

42.1<br />

1.61;1.36<br />

1.46;1.21 2.16<br />

1.57;1.32 2.12 1.57;1.32<br />

43.2<br />

32.5 38.9<br />

22.1<br />

46.6<br />

22.1<br />

H<br />

4.88<br />

H 4.67<br />

149.1<br />

21.4 107.5<br />

1.71<br />

1.71<br />

147.7<br />

4.88<br />

H<br />

4.63<br />

21.4 110.6<br />

H<br />

4.66<br />

(continued on next page)


50 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

Perillaldehyde—(R)-4-(prop-1-en-2-yl)cyclohex-1-enecarbaldehyde;<br />

chemical formula: C10H14O; exact mass: 150.1; molecular weight: 150.22,<br />

m/z: 150.10 (100.0%), 151.11 (11.0%); elemental analysis: C, 79.96; H, 9.39;<br />

O, 10.65<br />

p-Menth-2,8-dien-1-ol—(1R,4S)-1-methyl-4-(prop-1-en-2-yl)cyclohex-2-enol;<br />

chemical formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z:<br />

152.12 (100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O,<br />

10.51<br />

1.41<br />

O<br />

O<br />

9.68<br />

28.6<br />

OH 2.0<br />

192.9<br />

OH<br />

H 6.54<br />

141.3<br />

2.01;1.91<br />

5.59<br />

1.88;1.63<br />

132.4<br />

68.6<br />

145.6<br />

20.5<br />

37.5<br />

2.09;1.84<br />

1.77;1.52 2.33<br />

1.77;1.52 2.64 5.59<br />

30.9<br />

27.9 42.0<br />

131.7<br />

48.4<br />

21.4<br />

4.88 H<br />

H 4.67<br />

1.71<br />

146.3<br />

1.71<br />

107.5 149.1 21.5<br />

H<br />

4.63<br />

21.5 110.6<br />

H<br />

4.70<br />

Methyl thymyl ether—1-isopropyl-2-methoxy-4-methylbenzene; chemical<br />

formula: C11H16O; exact mass: 164.12; molecular weight: 164.24, m/z: 164.12<br />

(100.0%), 165.12 (11.9%); elemental analysis: C, 80.44; H, 9.82; O, 9.74<br />

24.6<br />

2.35<br />

136.6<br />

112.7<br />

121.0<br />

6.49<br />

6.84<br />

127.0<br />

7.18<br />

O<br />

155.6<br />

133.4<br />

O<br />

30.4<br />

56.1<br />

3.12<br />

3.73<br />

1.29 1.29<br />

23.6 23.6


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 51<br />

Perillyl alcohol—(4-(prop-1-en-2-yl)cyclohex-1-enyl)methanol; chemical<br />

formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12<br />

(100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

Perillene—3-(4-methylpent-3-enyl)furan; chemical formula: C10H14O; exact<br />

mass: 150.1; molecular weight: 150.22, m/z: 150.10 (100.0%), 151.11, (11.0%);<br />

elemental analysis: C, 79.96; H, 9.39; O, 10.65<br />

OH<br />

OH 5.56<br />

142.8<br />

110.7<br />

7.23<br />

6.13<br />

66.7<br />

4.20<br />

137.1<br />

H 5.58<br />

O<br />

125.0<br />

O<br />

2.01;1.91<br />

122.5<br />

24.8<br />

139.3<br />

28.7<br />

2.59<br />

2.09;1.84<br />

1.77;1.52 2.33<br />

7.11<br />

31.4<br />

28.4 42.3<br />

28.1<br />

H 5.20<br />

2.29<br />

4.88 H<br />

123.5<br />

1.71<br />

149.1<br />

21.5<br />

107.5<br />

H<br />

4.63<br />

19.6 132.0 25.6<br />

1.71 1.71<br />

Piperitenone—3-methyl-6-(propan-2-ylidene)cyclohex-2-enone; chemical<br />

formula: C10H14O; exact mass: 150.1; molecular weight: 150.22, m/z: 150.10<br />

(100.0%), 151.11 (11.0%); elemental analysis: C, 79.96; H, 9.39; O, 10.65<br />

cis-Piperitenol—(1S,6R)-3-methyl-6-(prop-1-en-2-yl)cyclohex-2-enol;<br />

chemical formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z:<br />

152.12 (100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O,<br />

10.51<br />

1.71<br />

22.6<br />

1.71<br />

23.4<br />

H 5.37<br />

166.8<br />

H 6.17<br />

133.9<br />

2.01;1.91<br />

126.4<br />

31.5<br />

2.00<br />

122.9<br />

31.3<br />

3.72<br />

1.77;1.52 2.41<br />

187.2<br />

15.5 123.9<br />

2.00<br />

71.3<br />

21.9 51.6<br />

OH 4.14<br />

O<br />

O<br />

OH<br />

H 4.63<br />

149.1<br />

1.71<br />

18.9<br />

150.2<br />

18.9<br />

1.71 1.71<br />

21.8 107.5<br />

H<br />

4.88<br />

(continued on next page)


52 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

trans-Piperitol—(1R,6R)-6-isopropyl-3-methylcyclohex-2-enol; chemical<br />

formula: C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

23.4<br />

1.71<br />

cis-Piperitol—(1S,6R)-6-isopropyl-3-methylcyclohex-2-enol; chemical<br />

formula: C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.3<br />

1.71<br />

23.4<br />

133.9<br />

H 5.37<br />

133.9<br />

H 5.37<br />

122.9<br />

31.2<br />

2.01;1.91<br />

122.9<br />

31.2<br />

2.01;1.91<br />

71.2<br />

19.1 47.4<br />

3.68<br />

1.71<br />

1.74;1.49<br />

19.1 47.4 71.2<br />

OH<br />

3.68<br />

1.74;1.49 1.71<br />

OH<br />

OH<br />

4.14<br />

OH<br />

4.14<br />

25.6<br />

21.4 21.4<br />

1.82<br />

1.01 1.01<br />

25.6<br />

21.4 21.4<br />

1.82<br />

1.01 1.01<br />

Piperitone oxide—3-isopropyl-6-methyl-7-oxabicyclo[4.1.0]heptan-2-one;<br />

chemical formula: C10H16O2; exact mass: 168.12; molecular weight: 168.23,<br />

m/z: 168.12 (100.0%), 169.12 (11.1%); elemental analysis: C, 71.39; H, 9.59;<br />

O, 19.02<br />

Piperitone—6-isopropyl-3-methylcyclohex-2-enone; chemical formula:<br />

C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12 (100.0%),<br />

153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

20.2<br />

1.31<br />

1.71<br />

O<br />

66.3<br />

64.4<br />

O<br />

H 5.85<br />

31.8<br />

3.38<br />

2.01;1.76<br />

2.01;1.91<br />

208.8<br />

16.3 53.1<br />

1.88;1.63 2.20<br />

1.52;1.27 2.33<br />

O<br />

O<br />

O<br />

25.8<br />

2.05<br />

2.05<br />

20.5 20.5<br />

1.01 1.01<br />

1.01 1.01


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 53<br />

Terpin-4-ol—1-isopropyl-4-methylcyclohex-3-enol; chemical formula:<br />

C10H18O; exact mass: 154.14, molecular weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

23.1<br />

Pulegone—5-methyl-2-(propan-2-ylidene)cyclohexanone; chemical formula:<br />

C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12 (100.0%),<br />

153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

1.06<br />

1.71<br />

133.9<br />

1.88<br />

H 5.37<br />

123.3<br />

24.7<br />

2.01;1.91<br />

3.03;2.78<br />

1.41;1.16<br />

30.3 71.7 40.8<br />

OH<br />

37.6<br />

1.88;1.63 2.19;1.94<br />

OH 4.64<br />

1.90<br />

2.01;1.91<br />

O<br />

14.9 14.9<br />

1.01 1.01<br />

1.71 1.71<br />

β-Terpineol—1-methyl-4-(prop-1-en-2-yl)cyclohexanol; chemical formula:<br />

C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14<br />

(100.0%), 155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

α-Terpineol—2-(4-methylcyclohex-3-enyl)propan-2-ol; chemical formula:<br />

C10H18O; exact mass: 154.14; molecular weight: 154.25, m/z: 154.14 (100.0%),<br />

155.14 (11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

1.31<br />

27.3<br />

OH 4.64<br />

23.1<br />

1.71<br />

OH<br />

71.0<br />

1.67;1.42<br />

1.67;1.42<br />

133.9<br />

H 5.37<br />

42.1<br />

42.1<br />

123.3<br />

31.2<br />

2.01;1.91<br />

1.57;1.32 2.12 1.57;1.32<br />

22.1<br />

46.6<br />

22.1<br />

19.1 46.5 24.1<br />

1.74;1.49 1.71 2.04;1.79<br />

H 4.67<br />

147.7<br />

1.71<br />

OH<br />

73.2<br />

OH 4.64<br />

21.4 110.6<br />

27.7 27.7<br />

1.26 1.26<br />

H<br />

4.66<br />

(continued on next page)


54 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

bicyclic oxygenated monoterpenes<br />

Borneol—1,7,7-trimethylbicyclo[2.2.1]heptan-2-ol; chemical formula: C10H18O; Bornyl acetate—1,7,7-trimethylbicyclo[2.2.1]heptan-2-yl-acetate; chemical<br />

exact mass: 154.14; molecular weight: 154.25, m/z: 154.14 (100.0%), 155.14 formula: C12H20O2; exact mass: 196.15; molecular weight: 196.29, m/z:<br />

(11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

196.15 (100.0%), 197.15 (13.3%), 198.15 (1.2%); elemental analysis: C,<br />

73.43; H, 10.27; O, 16.30<br />

1.16<br />

OH<br />

1.49;1.24<br />

3.15<br />

1.11 1.11<br />

1.52;1.27<br />

1.67;1.42<br />

1.42<br />

4.81<br />

21.0<br />

2.01<br />

13.3<br />

O<br />

O<br />

13.5<br />

1.16<br />

170.2<br />

52.7 OH<br />

O<br />

49.4 O<br />

30.0 50.4 76.1<br />

19.8 19.8<br />

1.49;1.24<br />

3.89<br />

30.2 50.6 81.8<br />

1.11 1.11<br />

23.6<br />

35.8<br />

19.5 19.5<br />

1.52;1.27<br />

1.76;1.51 23.3<br />

32.5<br />

45.2<br />

1.42<br />

45.4<br />

Fenchone—(1S,4R)-1,3,3-trimethylbicyclo[2.2.1]heptan-2-one; chemical<br />

formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12<br />

(100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

Camphor—(1S,4S)-1,7,7 trimethylbicyclo[2.2.1]heptan-2-one; chemical<br />

formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12<br />

(100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

21.2<br />

15.6<br />

1.26<br />

1.26<br />

O<br />

51.5<br />

O<br />

O<br />

61.6<br />

O<br />

224.4<br />

34.1<br />

37.1<br />

1.84;1.59<br />

2.02;1.77<br />

21.8<br />

220.0<br />

31.5<br />

1.11<br />

1.84;1.59<br />

21.8<br />

1.11<br />

21.2<br />

47.0<br />

28.4<br />

1.21<br />

1.60;1.35<br />

39.0<br />

2.14;1.89<br />

1.60;1.35<br />

1.99<br />

40.1<br />

28.4<br />

30.1<br />

1.21<br />

1.99<br />

21.2<br />

34.4


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 55<br />

Myrtenol—7,7-dimethylbicyclo[2.2.1]hept-2-en-2-yl)methanol; chemical<br />

formula: C10H16O; exact mass: 152.12; molecular weight: 152.23, m/z: 152.12<br />

(100.0%), 153.12 (10.9%); elemental analysis: C, 78.90; H, 10.59; O, 10.51<br />

56.0<br />

2.25<br />

66.0 127.0<br />

25.2 25.2<br />

H 5.59<br />

23.1<br />

1.11<br />

OH<br />

OH 5.56<br />

20.6<br />

149.4<br />

4.20<br />

1.63;1.38<br />

1.11<br />

1.63;1.38<br />

2.25<br />

Isoborneol—(1S,2S,4S)-1,7,7-trimethylbicyclo[2.2.1]heptan-2-ol; chemical<br />

formula: C10H18O; molecular weight: 154.25, m/z: 154.14 (100.0%), 155.14,<br />

(11.1%); elemental analysis: C, 77.87; H, 11.76; O, 10.37<br />

1.16<br />

OH<br />

3.15<br />

1.49;1.24<br />

1.11<br />

1.11<br />

1.52;1.27<br />

1.67;1.42<br />

4.81<br />

13.3<br />

52.7<br />

OH<br />

76.1<br />

30.0<br />

19.8<br />

19.8 50.4<br />

23.6 35.8<br />

64.8<br />

52.5<br />

45.2<br />

1.42<br />

β-Thujene alcohol—(1S,2S,5R)-5-isopropyl-2-methylbicyclo[3.1.0]<br />

hexan-2-ol; chemical formula: C10H18O; exact mass: 154.14; molecular<br />

weight: 154.25, m/z: 154.14 (100.0%), 155.14 (11.1%); elemental analysis:<br />

C, 77.87; H, 11.76; O, 10.37<br />

Thujyl alcohol—(1S,3S,4S,5R)-1-isopropyl-4-methylbicyclo[3.1.0]<br />

hexan-3-ol; chemical formula: C10H18O; exact mass: 154.14; molecular<br />

weight: 154.25 m/z: 154.14 (100.0%), 155.14 (11.1%); elemental analysis: C,<br />

77.87; H, 11.76; O, 10.37<br />

24.8<br />

HO<br />

1.31<br />

4.64 HO<br />

12.0<br />

1.06<br />

41.0<br />

84.5<br />

34.1<br />

OH<br />

81.6<br />

OH 4.81<br />

3.16<br />

1.67;1.42<br />

1.50<br />

51.7<br />

18.4<br />

1.69<br />

1.42<br />

32.1<br />

12.0<br />

1.52;1.27<br />

1.52;1.27<br />

49.7<br />

18.5<br />

1.67;1.42<br />

1.52;1.27<br />

31.2<br />

31.2<br />

33.5<br />

1.81<br />

33.5<br />

1.81<br />

18.5 18.5<br />

1.01 1.01<br />

18.5 18.5<br />

1.01 1.01<br />

(continued on next page)


56 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.2 (continued) 1 h and 13 c-nmr data of monoterpenes Found in cymbogon essential oils<br />

Verbenone—(1R,5R)-4,6,6-trimethylbicyclo[3.1.1]hept-3-en-2-one; chemical<br />

formula: C10H14O; exact mass: 150.1; molecular weight: 150.22, m/z: 150.10<br />

(100.0%), 151.11 (11.0%); elemental analysis: C, 79.96; H, 9.39; O, 10.65<br />

20.7<br />

1.71<br />

25.0<br />

165.0<br />

1.11<br />

H 5.85<br />

125.5<br />

52.6<br />

48.7<br />

2.62<br />

25.0<br />

200.8<br />

1.11<br />

25.2<br />

2.15;1.90<br />

O<br />

2.80<br />

O<br />

58.8


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 57<br />

table 2.3<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

Acoradiene—(4S)-1,8-dimethyl-4-(prop-1-en-2-yl)spiro[4.5]dec-7-ene; chemical formula: C 15H 24; exact mass: 204.19;<br />

molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

1H-NMR 1.06<br />

1.60;1.35<br />

1.66<br />

1.70;1.45<br />

1.63;1.38<br />

2.01;1.91<br />

2.17<br />

1.71<br />

2.00;1.75<br />

H<br />

H<br />

4.88<br />

H<br />

5.37<br />

4.63<br />

1.71<br />

13C-NMR 16.5<br />

32.5<br />

41.5<br />

28.6<br />

27.1<br />

67.4<br />

54.4<br />

36.4<br />

22.0<br />

147.7<br />

123.3<br />

110.6<br />

Alloaromadendrene—(1aR,4aS,7R,7bS)-1,1,7-trimethyl-4-methylenedecahydro-1H-cyclopropa[e]azulene; chemical<br />

formula: C 15H 24; exact mass: 204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%);<br />

elemental analysis: C, 88.16; H, 11.84<br />

1.60;1.35<br />

1.06<br />

4.88 H<br />

2.18<br />

H<br />

1.63;1.38<br />

1.51<br />

1.67 0.17<br />

H 4.63<br />

1.11<br />

2.01;1.91<br />

0.18<br />

1.41;1.16<br />

1.11<br />

34.9<br />

18.9<br />

29.8<br />

109.1<br />

29.1<br />

133.9<br />

H<br />

36.2<br />

148.7<br />

50.9<br />

30.4<br />

58.6<br />

39.4 24.1<br />

Aromadendrene—(1aR,4aR,7R,7bS)-1,1,7-trimethyl-4-methylenedecahydro-1H-cyclopropa[e]azulene; chemical<br />

formula: C 15H 24; exact mass: 204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%);<br />

elemental analysis: C, 88.16; H, 11.84<br />

4.88 H<br />

2.18<br />

H<br />

1.63;1.38<br />

1.60;1.35<br />

1.06<br />

1.51<br />

1.67 0.17<br />

H 4.63<br />

1.11<br />

2.01;1.91<br />

0.18<br />

1.41;1.16<br />

1.11<br />

34.9<br />

18.9<br />

29.8<br />

H<br />

109.1<br />

58.6<br />

39.4 24.1<br />

27.9<br />

148.7<br />

50.9<br />

27.9<br />

23.4<br />

19.2<br />

36.2<br />

23.4<br />

19.2<br />

30.4<br />

27.9<br />

27.9<br />

23.1<br />

(continued on next page)


58 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

α-Bergamotene—2,6-dimethyl-6-(4-methylpent-3-enyl)bicyclo[3.1.1]hept-2-ene; chemical formula: C 15H 24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

1.71<br />

1.71<br />

H<br />

5.20<br />

2.62<br />

2.04;1.79<br />

1.96<br />

1.25<br />

1.71<br />

1.97<br />

1.16<br />

H 5.37<br />

2.04;1.79<br />

25.6<br />

131.3<br />

19.6<br />

126.8<br />

46.8<br />

22.6<br />

31.6<br />

α-Bisabolene—(Z)-1-methyl-4-(6-methylhept-5-en-2-ylidene)cyclohex-1-ene; chemical formula: C15H24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

1.71<br />

17.4<br />

1.71<br />

2.00<br />

H<br />

1.71<br />

5.20<br />

2.00<br />

2.63<br />

2.00<br />

H<br />

5.21<br />

2.00<br />

1.71<br />

33.6<br />

26.7<br />

126.5<br />

30.5<br />

123.5<br />

40.3<br />

125.0<br />

132.0<br />

19.6 25.6<br />

β-Bisabolene—1-methyl-4-(6-methylhepta-1,5-dien-2-yl)cyclohex-1-ene; chemical formula: C15H24; exact mass: 204.19;<br />

molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

1.71<br />

H<br />

1.71<br />

2.00<br />

2.00<br />

2.33<br />

1.77;1.52<br />

2.01;1.91<br />

4.88<br />

4.63 H<br />

106.7<br />

35.6<br />

28.4<br />

40.2<br />

153.5<br />

31.0<br />

H<br />

5.20<br />

2.09;1.84<br />

H<br />

5.37<br />

1.71<br />

26.7<br />

135.7<br />

28.4<br />

120.5<br />

31.4<br />

123.5<br />

123.3<br />

132.0<br />

19.6 25.6<br />

21.2<br />

43.5<br />

39.1<br />

31.7<br />

23.1<br />

132.1<br />

133.9<br />

23.2<br />

23.1<br />

123.3<br />

30.6


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 59<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

Bisabolol—(2S)-6-methyl-2-(4-methylcyclohex-3-enyl)hept-5-en-2-ol; chemical formula: C 15H 26O; exact mass: 222.2;<br />

molecular weight: 222.37, m/z: 222.20 (100.0%), 223.20 (16.6%), 224.21 (1.3%); elemental analysis: C, 81.02; H, 11.79;<br />

O, 7.20<br />

1.71<br />

1.71<br />

1.96<br />

5.20<br />

H<br />

1.31 4.64<br />

OH<br />

1.44<br />

1.74;1.49<br />

1.71<br />

2.01;1.91<br />

2.04;1.79<br />

H<br />

5.37<br />

1.71<br />

40.3<br />

18.5<br />

25.6<br />

OH<br />

19.4<br />

44.4<br />

77.2<br />

24.4<br />

126.8<br />

131.3<br />

19.6 25.6<br />

α-Butenol—(5R,Z)-2,10,10-trimethyl-6-methylenebicyclo[7.2.0]undec-2-en-5-ol; chemical formula: C 15H 24O; exact<br />

mass: 220.18; molecular weight: 220.35, m/z: 220.18 (100.0%), 221.19 (16.5%), 222.19 (1.5%); elemental analysis: C,<br />

81.76; H, 10.98; O, 7.26<br />

2.28;2.03<br />

H<br />

5.20<br />

5.01 H<br />

4.14 OH H<br />

4.96<br />

3.94<br />

2.63<br />

2.00;1.75<br />

1.71<br />

1.11<br />

2.01;1.91<br />

1.96<br />

1.41;1.16<br />

1.11<br />

34.1<br />

124.6<br />

76.3<br />

144.4<br />

44.8<br />

141.8<br />

40.3<br />

21.2<br />

27.7<br />

β-Butenol—(Z)-6,10,10-trimethyl-2-methylenebicyclo[7.2.0]undec-5-en-3-ol; chemical formula: C 15H 24O; exact mass:<br />

220.18; molecular weight: 220.35, m/z: 220.18 (100.0%), 221.19 (16.5%), 222.19 (1.5%); elemental analysis: C, 81.76;<br />

H, 10.98; O, 7.26<br />

2.28;2.03<br />

HO<br />

4.14<br />

3.94<br />

5.20<br />

H<br />

2.63<br />

1.71<br />

2.00;1.75<br />

H 1.11<br />

4.96 H<br />

5.01<br />

2.01;1.91<br />

1.96<br />

1.41;1.16<br />

1.11<br />

HO<br />

40.1<br />

OH<br />

124.6<br />

123.3<br />

109.8<br />

135.1<br />

74.1 42.3<br />

154.8<br />

40.6<br />

109.8<br />

27.7<br />

23.4<br />

31.2<br />

31.2<br />

133.9<br />

53.7<br />

33.9<br />

39.7<br />

23.1<br />

27.2<br />

27.7<br />

54.0<br />

33.9<br />

26.9<br />

27.7<br />

(continued on next page)


60 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

Calamenene—8-isopropyl-2,5-dimethyl-1,2,3,4-tetrahydronaphthalene; chemical formula: C15H22; exact mass: 202.17;<br />

molecular weight: 202.34, m/z: 202.17 (100.0%), 203.18 (16.5%), 204.18 (1.3%); elemental analysis: C, 89.04; H, 10.96<br />

2.33<br />

22.4<br />

6.65<br />

2.90;2.80<br />

1.68;1.43<br />

6.68<br />

1.77<br />

2.93;2.68 1.06<br />

3.12<br />

1.29 1.29<br />

132.9<br />

125.7<br />

136.6<br />

30.8<br />

32.9<br />

123.0<br />

142.1 133.9<br />

41.2<br />

34.4<br />

23.6 23.6<br />

29.7<br />

20.3<br />

α-Cadinene—1-isopropyl-4,7-dimethyl-1,2,4a,5,6,8a-hexahydronaphthalene; chemical formula: C 15H 24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

5.37 H<br />

1.71<br />

2.04;1.79 1.66<br />

1.98<br />

1.77;1.52<br />

2.01;1.91<br />

2.31<br />

1.82<br />

1.01 1.01<br />

H<br />

5.37<br />

1.71<br />

123.3<br />

28.5<br />

135.7 26.2<br />

51.3<br />

30.0<br />

21.3<br />

41.4 36.2<br />

124.2<br />

21.5 21.5<br />

β-Cadinene—(1S)-1-isopropyl-7-methyl-4-methylene-1,2,3,4,4a,5,6,8a-octahydronaphthalene; chemical formula:<br />

C 15H 24; exact mass: 204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental<br />

analysis: C, 88.16; H, 11.84<br />

4.63 H<br />

2.01;1.91<br />

1.41;1.16<br />

1.01<br />

1.82<br />

2.32<br />

1.49<br />

1.97<br />

1.01<br />

H 4.88<br />

1.77;1.52<br />

H 5.37<br />

2.01;1.91<br />

1.71<br />

109.1<br />

48.5<br />

124.2<br />

31.3<br />

133.9<br />

23.4<br />

52.2<br />

26.2<br />

36.1 148.0 31.3<br />

30.5 51.2<br />

30.0<br />

21.4 21.4<br />

133.9<br />

23.4


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 61<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

γ-Cadinene—(1S)-1-isopropyl-4,7-dimethyl-1,2,3,5,6,8a-hexahydronaphthalene; chemical formula: C15H24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

1.71<br />

17.3<br />

2.01;1.91<br />

1.74;1.49 1.66<br />

2.05;1.95<br />

2.63<br />

H 5.21<br />

1.82<br />

1.01 1.01<br />

2.01;1.99<br />

1.71<br />

31.1<br />

126.7<br />

23.5 45.2<br />

26.2<br />

129.7<br />

44.8<br />

30.0<br />

21.5 21.5<br />

γ-Cadinol—(1S,4R)-4-isopropyl-1,6-dimethyl-1,2,3,4,4a,7,8,8a-octahydronaphthalen-1-ol; chemical formula: C15H26O; exact mass: 222.2; molecular weight: 222.37, m/z: 222.20 (100.0%), 223.20 (16.6%), 224.21 (1.3%); elemental analysis:<br />

C, 81.02; H, 11.79; O, 7.20<br />

1.31<br />

OH<br />

1.67;1.42<br />

1.74;1.49<br />

1.53<br />

2.01;1.91<br />

4.64<br />

25.5<br />

OH<br />

40.2<br />

76.2<br />

17.2<br />

56.1<br />

31.5<br />

1.52;1.27 1.45<br />

2.10<br />

H 5.37<br />

1.82<br />

1.01 1.01<br />

1.71<br />

20.3 51.9<br />

125.1<br />

35.4<br />

29.9<br />

21.4 21.4<br />

β-Caryophyllene—(Z)-4,11,11-trimethyl-8-methylenebicyclo[7.2.0]undec-4-ene; chemical formula: C 15H 24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

1.71<br />

2.01;1.91<br />

1.41;1.16<br />

1.11<br />

5.20<br />

H<br />

1.96 2.63<br />

1.11<br />

2.00;1.75<br />

2.01;1.99<br />

2.05;1.95<br />

4.63<br />

H<br />

H<br />

4.88<br />

23.4<br />

33.7<br />

26.9<br />

27.7<br />

135.1<br />

33.9<br />

124.6<br />

124.2<br />

53.7 48.5<br />

149.9<br />

27.7<br />

40.3<br />

27.8<br />

34.9<br />

32.0<br />

130.6<br />

133.9<br />

23.5<br />

23.4<br />

109.8<br />

(continued on next page)


62 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

β-Caryophyllene alcohol—(E)-4,11,11-trimethyl-8-methylenebicyclo[7.2.0]undec-4-en-5-ol; chemical formula:<br />

C 15H 24O; exact mass: 220.18; molecular weight: 220.35, m/z: 220.18 (100.0%), 221.19 (16.5%), 222.19 (1.5%);<br />

elemental analysis: C, 81.76; H, 10.98; O, 7.26<br />

1.71<br />

2.01;1.91<br />

1.41;1.16<br />

1.11<br />

16.77<br />

OH<br />

1.96 2.63<br />

1.11<br />

2.01;1.99<br />

2.05;1.95<br />

H<br />

2.00;1.75 4.63<br />

H<br />

4.88<br />

13.2<br />

29.5<br />

27.2<br />

27.7<br />

173.7<br />

106.5<br />

33.9<br />

149.9<br />

53.7 48.5<br />

β-Caryophyllene oxide—chemical formula: C 15H 24O; exact mass: 220.18; molecular weight: 220.35, m/z: 220.18<br />

(100.0%), 221.19 (16.5%), 222.19 (1.5%); elemental analysis: C, 81.76; H, 10.98; O, 7.26<br />

1.31<br />

1.50;1.25<br />

1.38;1.13<br />

1.11<br />

2.51<br />

O<br />

1.96 2.63<br />

1.11<br />

1.59;1.34<br />

2.01;1.91<br />

4.88<br />

H<br />

H<br />

2.00;1.75<br />

4.63<br />

39.0<br />

27.4<br />

27.7<br />

OH<br />

64.9<br />

60.7<br />

33.8<br />

40.3<br />

153.2<br />

53.3 48.4<br />

α-Chamigrene—1,5,5,9-tetramethylspiro[5.5]undeca-1,8-diene; chemical formula: C 15H 24; exact mass: 204.19;<br />

molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

2.01;1.91<br />

1.70;1.45<br />

1.11<br />

5.37<br />

H<br />

1.71<br />

1.74;1.49<br />

2.01;1.91<br />

2.04;1.79<br />

1.11<br />

1.71<br />

H<br />

5.37<br />

21.0<br />

27.7<br />

22.5<br />

36.4<br />

23.4<br />

123.3<br />

O<br />

27.7<br />

38.6 47.8<br />

30.3<br />

40.3<br />

19.4<br />

32.2<br />

29.5<br />

144.0<br />

37.1<br />

23.4<br />

123.3<br />

28.7<br />

33.2<br />

109.8<br />

109.1<br />

29.4<br />

133.9<br />

23.1


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 63<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

α-Cubebene—chemical formula: C 15H 24; exact mass: 204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19<br />

(16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

1.71<br />

5.38<br />

H<br />

2.33;2.08<br />

0.87<br />

1.06<br />

1.59<br />

1.44<br />

1.41<br />

1.82<br />

1.01 1.01<br />

1.52;1.27<br />

1.52;1.27<br />

20.9<br />

123.8<br />

44.4<br />

142.6 39.5<br />

22.6<br />

16.1<br />

39.9<br />

28.0<br />

45.7<br />

29.9<br />

29.3<br />

26.5<br />

21.3 21.3<br />

α-Cuparene—(R)-1-methyl-4-(1,2,2-trimethylcyclopentyl)cyclohexa-1,3-diene; chemical formula: C 15H 24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

1.56;1.46<br />

1.55;1.30<br />

1.26<br />

1.60;1.35 2.15<br />

1.11<br />

1.11<br />

H<br />

5.68<br />

2.15<br />

H<br />

5.68<br />

1.71<br />

21.1<br />

41.5<br />

36.1<br />

48.0<br />

22.7<br />

20.2 24.8<br />

155.2<br />

57.8<br />

32.0<br />

118.6 120.1<br />

22.7<br />

145.3<br />

23.2<br />

Dihydro-alpha-copaene-8-ol—chemical formula: C 15H 26O; exact mass: 222.2; molecular weight: 222.37, m/z: 222.20<br />

(100.0%), 223.20 (16.6%), 224.21 (1.3%); elemental analysis: C, 81.02; H, 11.79; O, 7.20<br />

H 5.30<br />

H 1.64<br />

HO<br />

4.81<br />

1.14<br />

H<br />

1.01<br />

3<br />

2<br />

1.16<br />

5 1.52;1.27<br />

4 1.40 10<br />

9 1.52;1.27<br />

H 1.46<br />

1.60<br />

1 6 8<br />

7 1.06<br />

H 1.59<br />

1.82<br />

H 1.48<br />

1.01<br />

48.5<br />

23.2<br />

37.5 5<br />

4 48.2<br />

70.3<br />

HO<br />

2 1<br />

51.6<br />

23.6<br />

31.9<br />

6<br />

56.7<br />

7<br />

21.6<br />

3<br />

21.6<br />

26.4<br />

10<br />

9<br />

34.2<br />

33.4<br />

8<br />

19.1<br />

(continued on next page)


64 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

β-Elemene—(1R,2S,4R)-1-methyl-2,4-di(prop-1-en-2-yl)-1-vinylcyclohexane; chemical formula: C 15H 24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

4.88<br />

H<br />

H<br />

5.79<br />

1.71<br />

H<br />

4.66<br />

4.95<br />

H<br />

1.26<br />

1.52;1.27<br />

2.14 2.12<br />

1.57;1.32<br />

H 4.67<br />

1.57;1.32 H 4.67<br />

1.71<br />

H 4.66<br />

115.7<br />

146.3<br />

22.1<br />

147.7<br />

110.6<br />

25.5<br />

40.3<br />

γ-Elemene—(S)-1-methyl-2,4-di(propan-2-ylidene)-1-vinylcyclohexane; chemical formula: C 15H 24; exact mass: 204.19;<br />

molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

4.92 H<br />

5.71 H<br />

1.71<br />

H<br />

1.36<br />

1.46;1.21<br />

4.95<br />

1.71<br />

2.68;2.58<br />

2.01;1.91<br />

1.71<br />

1.71<br />

20.2<br />

121.5<br />

40.0<br />

57.1<br />

112.6 28.0<br />

145.6<br />

20.2<br />

30.6<br />

45.7<br />

44.9<br />

33.4<br />

26.6<br />

26.6<br />

147.7<br />

132.5 129.1<br />

β-Eudesmol—2-((4aR)-4a-methyl-8-methylenedecahydronaphthalene-2-yl)propan-2-ol; chemical formula: C 15H 26O;<br />

exact mass: 222.2; molecular weight: 222.37, m/z: 222.20 (100.0%), 223.20 (16.6%), 224.21 (1.3%); elemental analysis:<br />

C, 81.02; H, 11.79; O, 7.20<br />

1.38;1.28<br />

2.01;1.91<br />

H<br />

4.66<br />

1.16<br />

1.34;1.09 1.49;1.24<br />

H<br />

4.67<br />

2.10 1.50<br />

1.52;1.27<br />

1.52;1.27<br />

1.26<br />

OH<br />

1.26 4.64<br />

23.9<br />

38.2 148.0<br />

23.0<br />

39.4 39.4<br />

109.1<br />

34.2<br />

57.0<br />

23.0<br />

21.4<br />

28.6<br />

20.3<br />

19.5<br />

110.6<br />

124.6<br />

19.5<br />

46.8 27.6<br />

73.5<br />

OH<br />

27.6


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 65<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

α-Farnesene—(3Z,6E)-3,7,11-trimethyldodeca-1,3,6,10-tetraene; chemical formula: C 15H 24; exact mass: 204.19;<br />

molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

5.26<br />

H<br />

2.63<br />

2.00<br />

2.00<br />

1.71<br />

H 5.21<br />

1.71<br />

H 5.20<br />

1.71 1.71<br />

H 5.16<br />

H 6.25<br />

H<br />

5.02<br />

39.8<br />

26.4<br />

136.5<br />

17.6<br />

123.5<br />

121.8<br />

24.7<br />

138.3<br />

116.1<br />

19.6<br />

132.0<br />

25.6<br />

β-Farnesene—(E)-7,11-dimethyl-3-methylenedodeca-1,6,10-triene; chemical formula: C 15H 24; exact mass: 204.19;<br />

molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

4.80<br />

H<br />

H<br />

4.88<br />

6.25<br />

H<br />

2.00<br />

2.00<br />

2.00<br />

5.02<br />

H<br />

2.00<br />

H<br />

5.16<br />

5.20<br />

H<br />

1.71<br />

H<br />

5.20<br />

1.71 1.71<br />

39.7<br />

26.4<br />

135.7<br />

17.5<br />

123.5<br />

122.7<br />

26.8<br />

138.3<br />

116.1<br />

132.0<br />

19.6 25.6<br />

131.0<br />

135.1<br />

38.2<br />

22.4<br />

143.9<br />

110.4<br />

(continued on next page)


66 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

Germacrene D—(1E,6E)-8-isopropyl-1-methyl-5-methylenecyclodeca-1,6-diene; chemical formula: C 15H; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

2.00<br />

2.00<br />

H<br />

4.88<br />

5.20<br />

H<br />

2.01;1.91<br />

H 4.80<br />

1.71<br />

H<br />

6.03 H 5.61<br />

1.46;1.21<br />

2.15<br />

1.86<br />

1.01<br />

1.01<br />

126.2<br />

39.4<br />

140.6<br />

30.5<br />

125.1 55.0<br />

32.4 21.1<br />

26.1<br />

37.7 17.4<br />

148.5<br />

136.4<br />

110.4 21.1<br />

α-Himachalene—(4aS,9aR)-3,5,5-trimethyl-9-methylene-2,4a,5,6,7,8,9,9a-octahydro-1H-benzo[7]annulene; chemical<br />

formula: C 15H 24; exact mass: 204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%);<br />

elemental analysis: C, 88.16; H, 11.84<br />

2.01;1.91<br />

1.71<br />

4.63 H<br />

2.32 H<br />

1.77;1.52<br />

H<br />

H 1.96<br />

1.11<br />

5.37<br />

H 4.88<br />

2.01;1.91<br />

1.38;1.28<br />

1.34;1.09<br />

1.11<br />

23.4<br />

31.3<br />

133.9<br />

109.1<br />

26.8<br />

H<br />

49.7<br />

148.7<br />

38.8<br />

46.9<br />

23.5<br />

124.2<br />

37.2<br />

45.4<br />

H<br />

25.9 25.9<br />

α-Humulene—(1Z,4Z,8Z)-2,6,6,9-tetramethylcycloundeca-1,4,8-triene; chemical formula: C 15H 24; exact mass: 204.19;<br />

molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

1.71<br />

2.00<br />

5.20 H<br />

2.00<br />

1.92<br />

1.21<br />

H 5.15<br />

1.71<br />

2.63<br />

H 5.37<br />

1.21<br />

H 5.43<br />

23.5<br />

36.6<br />

124.7<br />

24.5<br />

131.9<br />

124.8<br />

37.5<br />

42.0<br />

30.2<br />

133.1<br />

23.5<br />

39.0<br />

123.3<br />

144.7<br />

30.2


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 67<br />

table 2.3 (continued)<br />

sesquiterpenes Found in major, minor, and trace amounts in Cymbopogon oils (in alphabetical<br />

order)<br />

β-Humulene—(1Z,5Z)-1,4,4-trimethyl-8-methylenecycloundeca-1,5-diene; chemical formula: C 15H 24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

1.71<br />

5.20 H<br />

1.96<br />

1.37<br />

1.92<br />

1.21<br />

1.96<br />

H 4.70<br />

2.63<br />

H 4.67<br />

H 5.37<br />

1.21<br />

H 5.43<br />

23.4<br />

38.4<br />

124.1<br />

25.4<br />

132.6<br />

42.5<br />

139.5<br />

37.5<br />

42.0<br />

Longifolene—chemical formula: C 15H 24; exact mass: 204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19<br />

(16.5%), 206.19 (1.2%); elemental analysis: C, 88.16; H, 11.84<br />

1.34;1.24<br />

1.34;1.09<br />

1.38;1.13<br />

1.11<br />

1.26<br />

1.51<br />

1.50<br />

H 4.88<br />

1.11<br />

1.60;1.35<br />

2.18<br />

H 4.63<br />

1.63;1.38<br />

30.2<br />

165.0<br />

37.6<br />

33.0<br />

57.0<br />

20.0<br />

62.2<br />

43.4 32.5<br />

α-Murrolene—1-isopropyl-4,7-dimethyl-1,2,4a,5,6,8a-hexahydronaphthalene; chemical formula: C 15H 24; exact mass:<br />

204.19; molecular weight: 204.35, m/z: 204.19 (100.0%), 205.19 (16.5%), 206.19 (1.2%); elemental analysis: C, 88.16;<br />

H, 11.84<br />

5.37 H<br />

2.04;1.79<br />

1.71<br />

1.66<br />

1.77;1.52<br />

2.32<br />

2.01;1.91<br />

1.97<br />

1.82<br />

H<br />

1.01 1.01<br />

5.37<br />

1.71<br />

123.3<br />

26.1<br />

21.3<br />

28.5 41.4<br />

22.9<br />

135.7<br />

51.3<br />

30.0<br />

36.2<br />

21.5 21.5<br />

109.1<br />

26.1<br />

26.2<br />

124.2<br />

30.2<br />

48.1<br />

31.3<br />

110.7<br />

43.0<br />

123.3<br />

144.7<br />

30.2<br />

32.9<br />

133.9<br />

23.4


68 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

2.4 chemIstry and uses oF Cymbopogon essentIal oIls<br />

The trading of essential oils has always been dependent on the knowledge imparted regarding its<br />

quality. The components present in it and the odor value provided to it because of several physical<br />

and chemical parameters are of immense value. These have been important criteria since ancient<br />

and medieval times. Chemists had to develop methods of analysis of oils and determine their<br />

chemical composition because of possibilities of adulterations of cheaper essential oils with highly<br />

priced ones. Prior to the advent of gas liquid chromatography (GLC) (Ille 1986), chemists had to<br />

rely entirely on chemical analysis, but GLC analysis might be the most important factor in this<br />

regard, and of possible help. During the last two decades, the methodology has improved manyfold.<br />

Consequently, the data regarding major and minor constituents found in essential oils has also<br />

multiplied and helped in evaluating the quality of the oils besides providing information on the<br />

constituents that could be isolated and used in pure form.<br />

2.4.1 le M o n G r a s s oils<br />

Lemongrass oil is distilled from two morphologically different species of lemongrass, C. flexuosus<br />

(common name: East Indian lemongrass) and C. citratus (common name: West Indian lemongrass).<br />

The chemical composition of these oils is very similar, though the percentage of citral and other<br />

major monoterpenes vary to some extent. A high yield of citral has been reported in C. pendulus<br />

(common name: North Indian lemongrass), which is another wild species and under limited cultivation.<br />

It has also been a major commercial source of lemongrass oil.<br />

2.4.1.1 Cymbopogon flexuosus<br />

The East Indian lemongrass (C. flexuosus (Steud.) Wats.) oil is indigenous to South India, found<br />

in the Malabar and Cochin regions, and in the Malay Peninsula. A product of C. flexuosus comes<br />

from Ceylon, Myanmar, and adjacent countries, as well as from Mexico and the West Indies. It is<br />

commonly called Malabar lemongrass oil. The major and trace constituents reported by various<br />

workers have been tabulated in Table 2.1. In one of the reports, Nath et al. (1994) have identified<br />

25 components after examining the essential oil. Geraniol, citronellol, neral, and geranial have<br />

been reported as the major components. Neral is also referred in the books and literature as citralcis,<br />

citral-a, α-citral, (Z)-3,7-dimethylocta-2,6-dienal or citral-(Z). Similarly, geranial has also been<br />

termed as citral-trans, Citral-b, β-citral, (E)-3,7-dimethylocta-2,6-dienal or citral-(E). During several<br />

studies conducted in the last few years (Bhattacharya et al. 1997; Boelens 1994; Cherian et al.<br />

1993; Choudhary and Kaul 1979; Kulkarni et al. 1997; Kuriakose 1995; Mathela et al. 1996; Nair<br />

et al. 1984; Patra and Dutta 1986; Rao et al. 1995), many chemotypes/cultivars/variants have been<br />

reported. Rao et al. (1992) have identified α-bisabolol and methyl isoeugenol as major components<br />

in a chemotype. It has also been found that the quality of the herb deteriorates on storage of the herb<br />

and also affects oil quality (Singh et al. 1994). In a GC-MS analysis of the essential oil, 32 constituents<br />

were identified (Taskinen et al. 1983). Kulkarni et al. (1997) reported a variant resembling citronella<br />

that contained citronellol (9.5%), citronellal (6%), citronellyl acetate (11.2%), geraniol (11.1%),<br />

and geranyl acetate (25.9%), along with 20 other constituents. The profiles of essential oils from<br />

five C. flexuosus cultivars (OD-19, Pragati, Cauvery, SHK-7, and GRL-l); one C. pendulus cultivar<br />

(Praman); and one hybrid C. khasianus × C. pendulus cultivar (CKP-25) have been examined on<br />

the GLC capillary column. Besides cultivated species (Atal and Bradu 1976a), a few wild-growing<br />

strains of the species have also been investigated and designated as RRL-14 and RRL-59, which<br />

contained geranial (40%) and methyl isoeugenol (20%) as the major constituents of the essential oil<br />

pool. The results revealed that cultivar GRL-1 (Patra et al. 1990) is different from other cultivars<br />

because of the presence of a high amount of geraniol (80.2%) and relatively higher concentration<br />

of myrcene (3.97%) and geranyl acetate (4.6%) (Patra et al. 1997). C. sikkimensis, which was a new<br />

variety of C. flexuosus, revealed the presence of methyl isoeugenol (20.5%), methyl eugenol (23%),


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 69<br />

and d-limonene (16.5%) as major constituents. The hybrid lemongrass CKP-25 differed from other<br />

citral-rich varieties with respect to a number of minor compounds (Bhattacharya et al. 1997).<br />

The oil of lemongrass is widely used in soaps and detergents (Bhattacharya 1970; Guenther<br />

1950; Opdyke 1976). Citral (a mixture of geranial and neral) is the major component of the oil and<br />

is isolated in bulk to be used in flavors, cosmetics, and perfumes. Ionones is another group of very<br />

important synthetic aromatics that possesses a strong and lasting odor. They are synthesized from<br />

citral and are further used in the manufacture of synthetic vitamin A.<br />

The antifungal, antibacterial, and antioxidant properties of lemongrass oil has been widely utilized<br />

(Alam et al. 1994; Gyane 1976; Mehmood et al. 1997; Ramdan et al. 1972a, 1972b; Rao et al.<br />

1971; Shadab-Qamar et al. 1992; Singh et al. 1978; Wannissorn et al. 1996). Allergic contact dermatitis<br />

has been reported with the oil (Selvag et al. 1995). Some other uses have been reported as preservative<br />

(Arora and Pandey 1977) and in inhibition of sensitization reactions (Opdyke 1976). The<br />

leftover residue from lemongrass leaves has been successfully utilized as a source of raw material<br />

for cellulose pulp and paper production (Ciaramello et al. 1972; Ramirez et al. 1977; Siddique-Ullah<br />

et al. 1979). The other important use of the oil has been in the preparation of an insect-repellent<br />

complex (Anonymous 1973), which has been tested for insect repellent/attractant and nematicidal<br />

activities (Ansari and Razdan 1995).<br />

The cultivation, essential oil characteristics, chemical constituents, and uses (culinary, antibacterial,<br />

medicinal, aromatic, and others) of lemongrass, and the lemongrass industry in India have been<br />

discussed at length in an article by Gupta and Jain (1978) Ansari et al. (1996) in earlier publications.<br />

Data have been tabulated on the differences between C. flexuosus and C. citratus oils in terms<br />

of their physicochemical properties, the effects of drying the herbage in sunlight for up to 5 days<br />

before distillation on oil yields, and export of lemongrass oil from India during 2002–2007 have<br />

also been discussed in this book. Effect on quality of citral content on preservation of lemongrass<br />

oil has also been worked out (Kurian et al. 1984).<br />

2.4.1.2 Cymbopogon citratus<br />

The West Indian lemongrass oil (C. citratus (D.C.) Stapf) is rated inferior in quality to that of<br />

the East Indian type because of its low citral content. However, it attained importance after<br />

World War II when the latter was difficult to obtain. Its major constituents are listed in Table 2.1,<br />

but it differs from the East Indian type by the occurrence of substantial quantities of myrcene<br />

(12%–15%). The myrcene may undergo diene-condensation and polymerization on aging, and hence<br />

loses its solubility in 70% alcohol. The C. citratus is an important crop in Ethiopia, and its analysis<br />

has shown geraniol (40%), geranial and neral (13%–15%), and α-oxobisabolene (12%) as major<br />

constituents, which are different from usual West Indian lemongrass oil (Abegaz et al. 1983). The<br />

hydrodistilled essential oil from the leaves of C. citratus Stapf grown in Zambia was analyzed by<br />

GC and GC-MS. Sixteen compounds representing 93.4% of the oil were identified, of which citraltrans<br />

(39.0%), citral-cis (29.4%), and myrcene (18.0%) were the major components. Small amounts<br />

of geraniol (1.7%) and linalol (1.3%) were also detected (Esmort et al. 1998). The composition of<br />

the essential oil of the leaves growing on the campus of Lagos State University was determined<br />

by the use of GC and GC-MS. The oil gave 27 peaks, amounting to 98% of the total oil. Twentythree<br />

constituents amounting to 97.3% of the total oil were identified. The main constituents were<br />

geranial (33.7%), neral (26.5%), and myrcene (25.3%). Small amounts of neomenthol (3.3%), linalyl<br />

acetate (2.3%), Z-β-ocimene (1.0%), and E-β-ocimene were also detected (Adeleke et al. 2001). In<br />

one of the reports, citral (69.39%) has been cited as a major component along with other minor components,<br />

such as caryophyllene, citronellol, geraniol, α- and β-pinene, ethyl laurate, 1-8-cineole,<br />

limonene, phellandrene methyl heptenone, linalool, menthol, myrcene, terpineol, and citronellol<br />

(Torres and Ragadio 1992, 1996).<br />

In one of the studies, a method was developed to validate a HPLC procedure for the quantitative<br />

determination of citral in C. citratus volatile oil. The HPLC assay was performed using a<br />

Spherisorb ® CN column (250 mm × 4.6 mm, 5 μm), an n-hexane:ethanol (85:15) mobile phase,


70 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

and a UV detector (set at 233 nm). The following parameters were evaluated: linearity, precision,<br />

accuracy, specificity, quantification, and detection limits. The method showed linearity in the range<br />

of 10.0–30.0 μg mL −1 . Precision and accuracy were determined at the concentration of 20 μg mL −1 .<br />

The concentration of citral in C. citratus volatile oil obtained with this assay was 75%. The HPLC<br />

method developed in this study showed an excellent performance (linearity, precision, accuracy,<br />

and specificity), and can be applied to assay citral in volatile oil (Rauber et al. 2005). A total of<br />

34 compounds were identified in Moroccan C. citratus oil, constituting about 89% of the total oil,<br />

with geraniol and neral (39.8% and 32%, respectively) as major constituents. The oil yield was drastically<br />

reduced under rust disease indices reduction (Baruah et al. 1995). A reduction in geraniol<br />

content in contrast to an increase in neral and myrcene was observed. The plant extract yielded<br />

tanins in herbal tea (Blake et al. 1993), with cymbopogene, cymbopogonol, triacontanol, alkaloid,<br />

and saponin (Ansari et al. 1996) being identified. The presence of alkaloids was reported; however,<br />

it needs further confirmation. The crop growing in Nagcorlan and Laguna was reported to contain<br />

93.74% citral (citral-a, citral-b, and others in the ratio of 61:33:6) (Torres 1993).<br />

This oil has also been reported to be antimicrobial (Chalchat et al. 1997; Handique and Singh<br />

1990; Kokate and Verma 1971; Moris et al. 1979; Orafidiya 1993; Syeed et al. 1990; Yadav and<br />

Dubey 1994) and insecticidal (Sukari et al. 1992), insect repellant (Ansari and Razdan 1995), and<br />

found to have cytotoxic properties apart from its usual uses in perfumery, food flavor, and pharmaceutical<br />

industries (Guenther 1950; Opdyke 1976). C. citratus oil has also been tested for anticarcinogenic<br />

activities (Dubey et al. 1997; Zheng et al. 1993). Citral-a and citral-b have been shown to<br />

possess antibacterial activity in the oil (Onawunmi et al. 1984).<br />

In another study, the antibacterial properties of the essential oil have been recorded. These activities<br />

are shown in two of the three main components of the oil identified through CG and MS methods.<br />

Whereas the α-citral (geranial) and β-citral (neral) components individually elicit antibacterial<br />

action on Gram-negative and Gram-positive organisms, the third component, myrcene, did not show<br />

observable antibacterial activity on its own. However, myrcene provided enhanced activities when<br />

mixed with either of the other two main components identified (Grace et al. 1984). C. citratus is<br />

one of the most commonly used plants in Brazilian folk medicine for the treatment of nervous and<br />

gastrointestinal disturbances. It is also used in many other places to treat feverish conditions. The<br />

usual way to use it is by ingesting an infusion made by pouring boiling water on fresh or dried leaves<br />

(which is called “abafado” in Portuguese). Abafados obtained from lemongrass harvested in three<br />

different areas of Brazil (Ceará, Minas Gerais, and São Paulo states) were tested on rats and mice<br />

in an attempt to add experimental confirmation of its popular medicinal use. Citral, the main constituent<br />

of the essential oil in Brazilian lemongrass, was also studied for comparison. Oral doses of<br />

abafados up to 40 times (C 40) larger than the corresponding dosage taken by humans, or 200 mg/kg<br />

of citral, were unable to reduce the body temperature of normal rats and/or rats made hyperthermic<br />

by previous administration of pyrogen. However, both compounds acted when injected via the intraperitoneal<br />

route. Oral administration of doses C 20–C 100 of abafados and 200 mg/kg of citral did not<br />

change the intestinal transit of a charcoal meal in mice; neither did it decrease the defecation scores<br />

of rats in an open-field arena. Again, via the intraperitoneal route, both compounds were active. The<br />

possible central nervous system depressant effect of abafados was investigated by using batteries of<br />

12 tests designed to detect general depressant, hypnotic, neuroleptic, anticonvulsant, and anxiolytic<br />

effects. In all the tests employed, oral doses of abafados up to C 208 or citral up to 200 mg/kg were<br />

without effect. Only in a few instances did intraperitoneal doses demonstrate effects. These data<br />

do no lend support to the popular oral therapeutic use of lemongrass to treat nervous and intestinal<br />

ailments and feverish conditions (Carlini et al. 1986). Tea obtained from leaves of C. citratus (D.C.)<br />

Stapf is used for its anxiolytic, hypnotic, and anticonvulsant properties in Brazilian folk medicine.<br />

<strong>Essential</strong> oil (EO) from fresh leaves was obtained by hydrodistillation and orally administered to<br />

Swiss male mice 30 min before experimental procedures. EO at 0.5 or 1.0 g/kg was evaluated for<br />

sedative/hypnotic activity through pentobarbital sleeping time, for anxiolytic activity by elevated<br />

plus maze and light/dark box procedures, and for anticonvulsant activity through seizures induced


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 71<br />

by pentylenetetrazole and maximal electroshock. EO was effective in increasing the sleeping time,<br />

percentage of entries, and time spent in the open arms of the elevated plus maze as well as the<br />

time spent in the light compartment of the light/dark box. In addition, EO delayed clonic seizures<br />

induced by pentylenetetrazole and blocked tonic extensions induced by maximal electroshock, indicating<br />

the elevation of the seizure threshold and/or blockage of seizure spread. These effects were<br />

observed in the absence of motor impairment evaluated on the rota-rod and open-field tests. The<br />

results were in accord with the ethnopharmacological use of C. citratus, and, after complementary<br />

toxicological studies, it can support investigations assessing their use as an anxiolytic, sedative, or<br />

anticonvulsive agent (Blanco et al. 2007).<br />

Studies were conducted to investigate the hypoglycemic and hypolipidemic effects of the single,<br />

daily oral dosing of 125–500 mg/kg of fresh leaf aqueous extract of C. citratus Stapf in normal,<br />

male Wistar rats for 42 days. The average weights of rats per group were taken at 2-week intervals<br />

for 42 days. On day 43, blood samples from the rats were collected for fasting plasma glucose<br />

(FPG), total cholesterol, triglycerides, low-density lipoproteins (LDL-c), very low-density lipoprotein<br />

(VLDL-c), and high-density lipoprotein (HDL-c) assays through cardiac puncture under halothane<br />

anesthesia. Acute oral dose toxicity study of C. citratus was also conducted using the limit<br />

dose test of the Up and Down Procedure statistical program (AOT425StatPgm, Version 1.0) at a<br />

dose of 5000 mg/kg body weight/oral route. These results showed C. citrates to lower the FPG and<br />

lipid parameters dose dependently (p < 0.05) while raising the plasma HDL-c level (p < 0.05) in the<br />

same dose-related fashion but with no effect on the plasma triglycerides level (p > 0.05). The results<br />

of acute oral toxicity showed CCi to be of low toxicity and, as such, could be considered relatively<br />

safe on acute exposure, thus confirming its folkloric use and safety in suspected Type 2 diabetic<br />

patients (Adeneye and Agbaje 2007).<br />

2.4.1.3 Cymbopogon pendulus<br />

The North Indian lemongrass oil (C. pendulus (Nees ex Steud.) Wats.) occurs in wild areas of northern<br />

India such as Saharanpur (in the state of Uttar Pradesh) (Atal and Bradu 1976a; Muthuswami<br />

and Sayed 1980). The major and minor constituents are listed in Table 2.1. This is also a major<br />

source of lemongrass oil that is used in perfumery as much as the East Indian variety. In one of<br />

the reports, elemicin (53.7%) content was found to be very high in the essential oil obtained from<br />

this plant. It is noteworthy that it is the starting material for the synthesis of the antimalarial drug<br />

trimethoxyprim. (Z)-asarone (5.3%) is another valuable component of this oil, which is used as an<br />

antiallergic compound (Shahi et al. 1997).<br />

2.4.2 Ci t r o n e l l a oils<br />

C. winterianus and C. nardus are cultivated on a large scale, and these are closely related to each<br />

other in various aspects. Both species are distinguished morphologically by the shape and length of<br />

their leaves. The chemical composition of the essential oil obtained from them also differs considerably<br />

(Lawrence 1991; Wijesekara et al. 1973).<br />

2.4.2.1 Cymbopogon winterianus<br />

The Java citronella (C. winterianus Jowitt) is grown mainly in Java, Haiti, Honduras, Taiwan,<br />

Guatemala, and China, and is highly priced in comparison to the Ceylon type (see the following<br />

text) because its oil contains higher percentages of monoterpene alcohols and their esters. These<br />

are listed in Table 2.1. This oil is also known as Mahapengiri oil. The trace constituents identified<br />

from the essential oil include geranyl formate, borneol, farnesol, cadinene, l-cadinol, l-camphene,<br />

1-carvone, citral, citronellyl butyrate, cymbopol, dipentene, eugenol, l-limonene, linalool, methyl<br />

heptenone, methyl eugenol, α-pinene, sesquicitronellene, terpinene, terpinen-4-1, and thujyl alcohol.<br />

Conventional and nonconventional techniques have been applied while carrying research on<br />

these crops (Chauhan et al. 1976) because of their diversity and immense usefulness. The climatic


72 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

conditions and time of harvesting have been important factors in determining the chemical profile<br />

of the citronella oil (Singh et al. 1996), and this has been discussed in another chapter in this book.<br />

The pattern of accumulation of monoterpenes and effect of storage of herb prior to distillation<br />

have been reported (Singh et al. 1994, 1996). Soulari and Fanghaenel (1971) made a detailed study<br />

of essential oil of C. winterianus produced in Cuba. An improved variety of Java citronella was<br />

released by Ganguly et al. (1979).<br />

The hydrodistilled essential oil from the aerial parts of C. winterianus Jowitt, cultivated in<br />

Southern Brazil, was analyzed by GC-MS. Thirty-one components, representing 96% of the oil,<br />

were characterized. Enantiomeric ratios of limonene, linalool, citronellal, and β-citronellol were<br />

obtained by multidimensional gas chromatography, using a developmental model set up with two<br />

GC ovens. The enantiomeric distributions are discussed as indicators of the origin authenticity and<br />

quality of this oil (Lorenzo et al. 2000).<br />

Java citronella oil is one of the most important essential oils because of the high content of<br />

citronellal and is mainly used for the isolation of citronellal, which is converted into citronellol.<br />

Citronellol is further converted into citronellol esters, hydroxy citronellal, and synthetic menthol<br />

(Dev Kumar et al. 1977). Java citronella oil is usually employed in the scenting of soaps and all<br />

kinds of technical preparations as well as for the extraction of aromatic isolates.<br />

The essential oil steam-distilled from C. winterianus (Java citronella) of Cuban origin was analyzed<br />

by GC and GC-MS. Thirty-six compounds were identified, of which citronellal (25.04%),<br />

citronellol (15.69%), and geraniol (16.85%) were the major constituents (Pino et al. 1996). Using<br />

enantioselective multidimensional chromatography (enantio-MDGC) and the column combination<br />

polyethylene glycol/heptakis (2,3-di-O-acetyl-6-O-tert-butyldimethylsilyl)-beta-cyclodextrinin OV<br />

1701-vi, the chiral monoterpenoids cis/trans-rose oxides, linalol [linalool], citronellol, and terpinen-<br />

4-ol were stereoanalyzed simultaneously. The method was applied to chirality evaluation of these<br />

compounds in commercial and authentic Java (C. winterianus) and Ceylon (C. nardus) citronella<br />

oils. The enantiomeric distributions are discussed with reference to their uses as indicators of the<br />

authenticity of these essential oils (Kreis et al. 1994).<br />

2.4.2.2 Cymbopogon nardus<br />

Citronella oil is derived from C. nardus (L.) Rendle and is also called “Lanabatu oil.” The grass<br />

is mostly cultivated in Sri Lanka, and hence the oil obtained from it is known as Ceylon citronella<br />

oil. The main constituents of this oil have been presented in Table 2.1. The presence of phenolic<br />

derivatives (methyl eugenol and methyl isoeugenol) is the most significant difference between the<br />

Ceylon-type and Java-type oils. The wild varieties of citronella growing in Sri Lanka contain phenyl<br />

propanoids in abundance, whereas phenyl propanoids are present in traces in the Java-type<br />

oil, and this is the significant difference between them. The presence of elemol in the Ceylon-type<br />

has been suggested to be formed as an artifact. Wijesekara et al. (1973) isolated hedycaryol, a<br />

thermolabile precursor, by cold percolation of the macerated fresh grass. In one of the reports it<br />

was shown that the oil contained large amounts of monoterpene hydrocarbons, whereas the Java<br />

variety (Mahapengiri) contained only small amounts, mainly limonene. Both types contained comparable<br />

amounts of geraniol, and the Java type had more quantities of citronellol and citronellal.<br />

In addition, the Ceylon type contained tricyclene, methyl eugenol, methyl isoeugenol, eugenol, and<br />

l-borneol. The GLC profiles enable the identification of the type of oil and the detection of kerosene<br />

as a possible adulterant. The variety that grows wild in Sri Lanka (Mana) was quite different<br />

from both cultivated types (Wijesekera et al. 1973). The steam-distilled volatile oil obtained from<br />

partially dried grass (citronella grass) C. nardus (Linn.) Rendle (Syn. Andropogon nardus Linn.),<br />

cultivated in the Nilgiri Hills at Ooty, India, was analyzed by capillary GC and GC-MS. The partially<br />

dried grass contained 35 components, out of which 29 constituents were completely identified<br />

and comprised 92.7% of the oil. The oil contains 16 monoterpenes (79.8%), nine sesquiterpenes<br />

(11.5%), and four nonterpenic compounds (1.4%). The prominent monoterpenes were citronellal<br />

(29.7%), geraniol (24.2%), γ-terpineol (9.2%), and cis-sabinene hydrate (3.8%). The predominant


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 73<br />

sesquiterpenes were (E)-nerolidol (4.8%), β-caryophyllene (2.2%), and germacren-4-ol (1.5%)<br />

(Vijender and Mohammed 2002).<br />

A total of 36 compounds were identified in the steam-distilled essential oil of C. nardus collected<br />

from Zimbabwe in 1989. The major compound was trans-geraniol (29.47%), followed by its<br />

ester form geraniol formate (8.79%) (Moody et al. 1995).<br />

The Ceylon citronella oil was tested against Gram-positive bacteria and fungi, and it was<br />

found that, under in vitro conditions, the oil was as active as penicillin (Kokate and Verma 1971).<br />

C. winteri anus oil also finds use in providing scent and good smell to low-cost products such as<br />

soaps, sprays, disinfectants, polishes, and all kinds of technical preparations. Several products and<br />

formulations have been prepared using citronella oil (Chicopharma 1970; Kichiyoshi et al. 1981) for<br />

preventing thinner sniffing and slowing the release of the rapidly evaporating substances into the<br />

atmosphere. A number of patents have been filed on the uses of citronella oil. A patent describes the<br />

use of citronella grass after the distillation in papermaking and sulfate pulping (Bhaumik and Rao<br />

1978a, 1978b; Ciaramello et al. 1972). Another French patent by Tabakoff in 1969 and a Japanese<br />

publication by Fuji et al. in 1972 described the use of citronella oil in a composition that can be<br />

used with or without addition of water and does not corrode equipment or harm the hands. A joint<br />

Canadian and Indian Patent (Prasad and Jamwal 1971; Shaw 1971) described the use of citronella oil<br />

for the production of a formulation as mosquito repellent. Similarly, it was also found to be effective<br />

as housefly repellent (Osmani et al. 1972).<br />

2.4.3 Pa l M a r o s a a n d Gi n G e r G r a s s oils<br />

Palmarosa oil is obtained (Shylaraj and Thomas 1992) from the variety motia, which yields an oil<br />

of better quality having more geraniol and which is commercially much more important than the<br />

oil of gingergrass derived from the variety “sofia.” The oil obtained from variety sofia is sometimes<br />

used as an adulterant (Biaswara et al. 1976a, 1976b). The two important essential oils, palmarosa oil<br />

C. martinii (Roxb.) Wats. and gingergrass oil, are obtained from two of the most important grasses<br />

cultivated in India. The sofia variety has lower geraniol content.<br />

2.4.3.1 Cymbopogon martinii<br />

Palmarosa oil (C. martinii (Roxb.) Wats.), distilled from variety motia, has geraniol as the major<br />

component and is considered better in quality (Siddiqui et al. 1975, 1979). GC-MS analysis (Gaydou<br />

and Raudriamiharisoa 1987) of the hydrocarbons (4.75%) of the oil showed the presence of monoterpenes<br />

(45.9%), sesquiterpenes (52.2%), n-alkanes (1.6%), and unidentified compounds (0.4%). The<br />

main and trace constituents from the oil have been listed in Table 2.1. An interesting dihemiacetal<br />

bismonoterpenoid has been identified by Bottini et al. (1987). The x-ray diffraction method was used<br />

to establish its structure. Palmarosa oil is another that is extensively used more than any other oil in<br />

soaps, and it imparts a rose-like prominent and lasting odor (Guenther 1950; Opdyke 1974). This oil<br />

is also employed in the flavoring of tobacco and other mouth fresheners. There are several reports<br />

that the oil content increases by storing the herb prior to distillation (Singh et al. 1994). This has been<br />

discussed in another chapter by Aklandey in this book. Fungitoxic activity has also been reported<br />

from this oil (Singh et al. 1980) and, as in the case of the citronella crop, the leftover part after distillation<br />

of the plant is used in paper production (Ciaramello et al. 1972). The essential oil produced from<br />

the sofia variety of C. martinii Stapf is known as gingergrass oil. The major and minor constituents<br />

of the oil are listed in Table 2.1. The cis and trans forms of p-menth-2,8 diene-1-ol, p- menth1(7),8dien-2-ol,<br />

carveol, and piperitol, along with limonene (20%) and monoterpene alcohols, have been<br />

reported from the wild strain of C. martinii var. sofia growing in Kumaon hills (Mathela et al. 1988;<br />

Mathela and Pant 1988; Mathela et al. 1986). One new hemiacetal bismonoterpenoid compound<br />

cymbodi acetal was characterized in the oil of C. martinii (Bottini et al. 1987). The oil of gingergrass<br />

is used in low-cost perfume formulations and scenting of soaps and cosmetics.


74 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

The composition of the hydrocarbon fraction of the essential oil from C. martinii, which represents<br />

less than 5% of the oil, has been studied. Using well-established techniques, 11 monoterpenes<br />

(ca 46%), 28 sesquiterpenes (ca 52%), and 16 n-alkanes (ca 1.6%) have been identified. The<br />

major constituents are limonene, α-terpinene, myrcene, β-caryophyllene, α-humulene, and β- and<br />

δ-selinenes. The study of the n-alkanes of C. martinii revealed the presence of all members of the<br />

homologous series C 15–C 30 (Gaydou and Raudriamiharisoa 1986, 1987b).<br />

[2R-(2α,4αβ,5αβ,7α,9αβ,10αβ)]-Octahydro-2,7-bis(1-methylethenyl)5αH, 10αH-4α,<br />

9α-ethanodibenzo[b, e] [1,4]dioxin-5α,10α-diol (cymbodiacetal), isolated from the essential oil of<br />

C. martinii, was identified by means of x-ray diffraction of its 1:1 solvate with deuteriochloroform<br />

(Bottini et al. 1987). Only immature palmarosa C. martinii (Roxb.) Wats. var. motia inflorescence<br />

with unopened spikelets accumulated essential oil substantially. Geraniol and geranyl acetate together<br />

constituted about 90% of palmarosa oil. The proportion of geranyl acetate in the oil decreased significantly<br />

with a corresponding increase of geraniol during inflorescence development. An esterase<br />

enzyme activity, involved in the transformation of geranyl acetate to geraniol, was detected from<br />

the immature inflorescence using a GC procedure. The enzyme, termed as geranyl acetate-cleaving<br />

esterase (GAE), was found to be active in the alkaline pH range with the optimum at pH 8.5. The<br />

catalysis of geranyl acetate was linear up to 6 h, and after 24 h of incubation, 75% of the geranyl<br />

acetate incubated was hydrolyzed. The GAE enzymic preparation, when stored at 4°C for a week,<br />

was quite stable with only 40% loss of activity. The physiological role of GAE in the production of<br />

geraniol during palmarosa inflorescence development has been discussed (Dubey and Luthra 2001).<br />

2.4.4 Cy m b o p o g o n j w a r a nC u s a<br />

The roots of C. jwarancusa (Jones) Schult. also contain essential oil unlike other Cymbopogon species.<br />

The oil obtained from aerial parts of the plant is rich in monoterpene alcohols, imparting an excellent<br />

odor. Besides major and trace constituents reported in Table 2.1, several other reports are being mentioned<br />

here. Cis- and trans-p-menthenols (38%) have been reported as major components of the oil<br />

(Mathela and Pant 1988). Mathela et al. (1986) reported monoterpene alcohol (25%), piperitols (25%),<br />

and piperitinone (6.5%), along with α-thujene, camphene, p-cymene, piperitinone, umbellulone, and<br />

elemol in the essential oil of C. jwarancusa. Piperitone was reported to be the single major constituent<br />

(Saeed et al. 1978; Thapa et al. 1971), while a mixture of four isomers (3R,4S-(+)cis-p-menth-1-ene-<br />

3-ol-4 and 3S,4S-(−)-trans and two isomers (cis and trans) with same skeleton having p-menth-2-enl-ol<br />

constitute 56% of oil. The oil obtained from the roots was rich in monoterpene hydrocarbon and<br />

sesquiterpene alcohol, the main sesquiterpenoid component being agarospiral (Mathela et al. 1986).<br />

Two chemical races, C. jwarancusa subsp. jwarancusa and C. jwarancusa subsp. oliveri, have<br />

been identified from the Kumaon Himalayan region (Mathela et al. 1986); the first one being rich<br />

in piperitone and the latter having cyclic and monoterpene alcohols but low piperitone. The components<br />

of the oil of C. jwarancusa differed with growth conditions, particularly geographical locations.<br />

The composition of the essential oil of Khavi grass, C. jwarancusa, was investigated by glass<br />

capillary GC in combination with mass spectrometry. Sixty-four compounds were identified, 55 of<br />

which were reported for the first time. The oil contains a high percentage of piperitone (60%–70%),<br />

which is mainly responsible for the smell of Khavi grass (Talat et al. 1978).<br />

The grass found in the Indian Thar desert contained citral, geraniol, geranyl acetate, and piperitone<br />

as the major components in its essential oil (Shahi 1992; Shahi and Sen 1989, 1993). In an<br />

unusual finding, paramenthenol was reported to constitute 60% of the oil (Boelens 1994; Mathela<br />

1991). Piperitone is mainly employed as a starting material for the preparation of several valuable<br />

perfumery compounds, for example, menthol, thymol, etc. (Guenther 1950), and being the principal<br />

constituent of the essential oil and possessing a mint or camphor-like odor, this oil is used for scenting<br />

many technical preparations directly. In a study on growth performance of three cultivars of<br />

C. jwarancusa (Jorlab-C.j.5, Jorlab-C.j.3, and C.j.), it was found that cultivar Jorlab-C.j.5 proved to


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 75<br />

be the best among the three test cultivars in respect of herb yield (21.1 t/ha), oil content (1.6%), and<br />

major oil constituent, that is, piperitone (83%) (Singh and Pathak 1994).<br />

2.4.5 Cy m b o p o g o n s C h o e n a n t h u s<br />

C. schoenanthus (L.) Spreng subsp. proximus Hochst. is primarily native to East Africa and has<br />

been experimented under cultivation in the surrounding areas (Guenther 1950). The major and<br />

minor constituents of the oil have been discussed in Table 2.1. Cryptomeridiol, a component in<br />

the oil, has been found to be responsible for its antispasmodic activity. Diuretic and antihistaminic<br />

commercial preparations have been made for the oil. Apart from these uses, the oil finds its usual<br />

employment as a flavoring agent and in the perfumery industry.<br />

The insecticidal activity of the crude essential oil extracted from C. schoenanthus and its main<br />

constituent, piperitone, was assessed in different developmental stages of Callosobruchus maculatus<br />

(Ketoh et al. 2006). Piperitone was more toxic to adults with an LC 50 value of 1.6 μL/L versus<br />

2.7 μL/L obtained with the crude extract. Piperitone inhibited the development of newly laid eggs<br />

and neonate larvae, but was less toxic than the crude extract to individuals developing inside the<br />

seeds (Guillaume et al. 2005).<br />

2.4.6 ot h e r Cy m b o p o g o n sP e C i e s<br />

The essential oils of numerous wild-growing Cymbopogon species have been chemically examined,<br />

and the results reveal that some of them can be used as a source of valuable essential oils.<br />

2.4.6.1 Cymbopogon caesius<br />

C. caesius (Nees) Stapf is a loosely tufted perennial grass with erect culms, sometimes stilt-rooted,<br />

to 2½ m high; of deciduous savanna bushland and wooded grassland; abundant throughout the region<br />

and, in general, over all of tropical Africa. Rao and Sudborough (1925) investigated this oil for the<br />

first time and reported limonene, geraniol, perillyl alcohol, and dipentene as the chief constituents<br />

present in it. About 50 years later, Sinha and Mehra (1977) reported that carvacrol, d-perillaldehyde,<br />

and d-nerolidol were the main constituents of the oil that were effective against E. coli. However,<br />

carvone (30%) is the major constituent in the Chinese oil (Liu et al. 1981). Recently, GC-MS analysis<br />

of the oil from the plants growing in the Northeast region of India have shown 35 compounds,<br />

out of which 24 have been identified and listed in Table 2.1. The essential oil of C. caesius is well<br />

placed to be used in perfumery. Detailed analysis by GC-MS is being undertaken, and reports are<br />

likely to be available shortly.<br />

2.4.6.2 Cymbopogon coloratus<br />

A total of 33 compounds were reported by Mallavarapu et al. (1992a) in the chemical profile of the<br />

essential oil C. coloratus (Nees) Stapf. The main constituents of the oil are myrcene, limonene,<br />

trans-β-ocimene, linalool, neral, geranial, geraniol (69.11%), geranyl acetate, and elemol. Pilley<br />

et al. (1928) described the presence of borneol, limonene, and camphene. The oil is also used for<br />

perfuming soaps and cosmetics (Gupta and Daniel 1982). Bor (1954) was skeptical about the authenticity<br />

of the species.<br />

2.4.6.3 Cymbopogon confertiflorus<br />

The synonym of C. confertiflorus (Steud.) Stapf is reported as C. nardus, which is also known as<br />

Ceylon citronella as stated earlier. The oil of this particular species is similar to the Ceylonese<br />

variety but inferior in quality with less geraniol (35%–40%) as the principal constituent (Gupta and<br />

Daniel 1982).


76 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

2.4.6.4 Cymbopogon densiflorus<br />

C. densiflorus (Steud.) Stapf is a tufted perennial grass with culms 1.8 m high; of open spaces<br />

along roadsides and wooded grassland; native to central tropical Africa, Gabon to Zimbabwe, and<br />

introduced into the region (Nigeria and possibly other states) and into Brazil. It is grown in Gabon<br />

and Nigeria as an ornamental and for its aromatic oil. The leaves are avoided by browsing cattle in<br />

Zambia. The crushed leaves are used as treatment for rheumatism in Gabon. In Malawi, the flower<br />

head is smoked in a pipe as a cure for bronchial affections and, for the same complaints, the plant<br />

sap is used in the Congo (Brazzaville), where it is also given as treatment for asthma and to calm<br />

fits. It is macerated with Ocimum basilicum (Labiatae), and the compound is used for epilepsy in<br />

Zaïre. It is conjectured that any medicinal action is due to the camphoraceous volatile oil. The plant<br />

has also been recorded to be used as a tonic and styptic. It has fetish attributes as well. In Gabon,<br />

the inflorescence of C. densiflorus is burnt in fumigations required in certain rituals, for example,<br />

in incantations to chase away malignant spirits, to cleanse those who have lost their spouse, to rejuvenate<br />

and restore the efficiency of a fetish, as an amulet or talisman when the owner has violated<br />

a taboo. In Tanganyika, witch doctors smoke the flower panicle, either alone or with tobacco, to<br />

induce dreams to foretell the future. Huntsmen in Gabon use the plant as a fetish lure for game. The<br />

oil of the Brazilian plant was compared with the African oil by Koketsu et al. (1976), and olfactory<br />

analysis showed no noticeable difference. The monoterpenes found in the oil from the flowers and<br />

leaves of the Zambian plant are shown in Table 2.1. No remarkable difference in the quality of the<br />

oil from Brazilian and African plants could be recorded (Boelens 1994; Chisowa 1997).<br />

2.4.6.5 Cymbopogon distans<br />

The essential oil of C. distans (Nees) Wats. has been studied by GC-MS (Mathela and Joshi 1981;<br />

Mathela et al. 1989), which showed several monoterpenoids in addition to 19% sesquiterpene alcohols.<br />

Sobti et al. (1978) reported terpineol (20%) as the major constituent, whereas Thapa et al.<br />

(1971) and Liu et al. (1981) reported piperitone (30%–40%) and geraniol (10%) as the principal<br />

constituents of the oil. In another publication (Singh and Sinha 1976), limonene (29%) and methyl<br />

eugenol (13%) were reported to be the major constituents. Other major constituents have been<br />

reported in Table 2.1.<br />

C. distans has been reported to occur in nature in the form of several geographical races. The<br />

essential oils isolated from the leaves of C. distans chemotype loharkhet and the roots of C. jwarancusa<br />

(collected from India) were analyzed by GC, GC-MS, and liquid chromatography. Both oils<br />

were qualitatively very similar in sesquiterpenoid composition but contained different total concentrations<br />

of sesquiterpenoids (79.6% and 38.0% in the oils of C. distans and C. jwarancusa, respectively).<br />

The main sesquiterpenoids of the essential oil of C. distans were eudesmanediol (34.4%) and<br />

5-epi-7-epi-alpha-eudesmol (11.2%). The main sesquiterpenoid in the essential oil of C. jwarancusa<br />

was agarospirol (9.5%) (Beauchamp et al. 1996; Dunyan et al. 1992). Exhaustive studies have been<br />

done by Mathela et al. (1988a, 1989) and Mathela and Pant (1988) on the chemical investigations of<br />

the essential oil, and they characterized C. distans munsiyariensis, which contained eudesmanediol<br />

(34.4%) along with geraniol (22.89%), neryl acetate (18.34%), neral (14.74%), and limonene (12.08%)<br />

as the major constituents. Mathela and Pant (1988) reported four chemotypes from the Kumaon<br />

and Garhwal regions of Uttar Pradesh (India) having marker compounds α-oxobisabolene-1<br />

(chemotype I); citral, geraniol, and geranyl acetate (chemotype II); piperitone, limonene, and eudesmanediol<br />

(chemotype III); and sesquiterpene alcohol (chemotype IV) in their oils. One more chemotype<br />

with chemical marker p-menthol (66.5%) was reported later (Pande et al. 1997). A GC-MS<br />

study of the hydrocarbon fraction and the fraction containing oxygenated compounds showed the<br />

presence of 12 monoterpene hydrocarbons (28.4%), 13 sesquiterpene hydrocarbons (32.8%), 3 sesquiterpene<br />

alcohols (27.2%), 2 esters (7.2%), and 3 carbonyl compounds (4.4%) in the essential oil of<br />

C. distans. Of these, 27 compounds have been identified (Mathela and Joshi 1981).


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 77<br />

2.4.6.6 Cymbopogon khasianus<br />

The essential oil of C. khasianus (Hack.) Stapf ex Bor has been reported to contain citral (40%–60%)<br />

and geraniol (70%–80%) as the major constituents (Balyan et al. 1979; Choudhary and Leclercq 1995;<br />

Rabha et al. 1986; Rabha et al. 1989; Sobti et al. 1978a; Sobti et al. 1982; Thapa et al. 1971; Thapa<br />

et al. 1976; Verma et al. 1987). Methyl eugenol (75.82%) is the major constituent in C. khasianus.<br />

2.4.6.7 Cymbopogon ladakhensis<br />

Very little information is available on C. ladakhensis. However, Gupta and Daniel (1982) have<br />

reported that piperitone is the chief constituent of this oil.<br />

2.4.6.8 Cymbopogon microstachys<br />

The essential oil C. microstachys (Hook.s) Soenarke from the northeastern Indian state of Manipur<br />

was grown in Bhubaneswar in the state of Orissa in pots. The plants were found to grow well and,<br />

in 45 days, reached full growth. The essential oil thus obtained from this species was analyzed by<br />

GC and GC/MS and the results showed that the oil contained (E)-methyl isoeugenol (56.4%–60.7%)<br />

as the major constituent. This is the first time that an oil of C. microstachys has been found with<br />

(E)-methyl isoeugenol as the major constituent (Rout et al. 2005). The oil was found to contain about<br />

60% phenyl propenoids with methyl eugenol (19.5%), methyl isoeugenol (4.2%), elemicin (25.3%),<br />

and isoelemicin (11.0%). The oil now reported had quite a different composition with (E)-methyl<br />

isoeugenol as the main constituent (56.4%–60.7%). Other constituents in significant quantities were<br />

myrcene (7.8%–12.2%), (Z)- and (E)-β-ocimene (2.4%–2.9%), cis-α-bergamotene (0.8%–1.7%),<br />

trans-α-bergamotene (0.8%–3.4%), germacrene D (0.6%–2.7%), and (Z,E)-α-farnesene (0.4%–2.9%),<br />

besides an unidentified compound at RT 46.1 min (0.6%–6.0%). In all, 44 components were identified,<br />

constituting 91.3%–96.2% of the oil. The study shows that the aromatic grass collected in Manipur is<br />

a chemical variant of C. microstachys. The easy adaptability of the plant to Bhubaneswar conditions,<br />

high oil yield, and presence of methyl isoeugenol as the major phenyl propenoid may make this a<br />

commercially important essential oil. Earlier reports have indicated that it contains citral and geraniol<br />

(Gupta and Daniel 1982). Methyl eugenol, elemicin, and isoelemicin are the major constituents<br />

(Boelens 1994; Pant et al. 1990) found in a chemical investigation of the oil, along with more than two<br />

dozen minor constituents: aromadendrene; alloaromadendrene, caryophyllene oxide, 3,4-epoxy-3,7<br />

dimethyl-1,6-octadione, 6-7-epoxy-3,7-dimethyl-1,3-octadiene, cis-3-hexenol, humulene, limonene,<br />

cis-limonene oxide, linalool, 6-methyl-5-hepten-2-one methyl isoeugenol, myrcene, nerolidol,<br />

4-nonanone, trans-ocimene, β-phellendrene, α-pinene, sabinene, terpinen-4-ol, α-terpineol, tricyclene,<br />

trimethoxy benzaldelyde, and veratraldehyde (Mathela et al. 1990b).<br />

2.4.6.9 Cymbopogon nervatus<br />

The essential oil hydrocarbons from C. nervatus (Hochest.) Chiov. have been investigated (Modawi<br />

et al. 1984), and sesquiterpenes β-selinene, β-elemene, β-bergamotene, and germacrene-D have<br />

been reported. The antibacterial activity of essential oil of dried inflorescence of C. nervatus<br />

was investigated. The essential oil remarkably inhibited the growth of tested bacteria except for<br />

Salmonella typhi. The maximum activity was against Shigella dysenteriae and Klebsiella pneumoniae<br />

(Kamali et al. 2005).<br />

2.4.6.10 Cymbopogon olivieri<br />

The major components of the essential oil of C. olivieri (Boiss.) Bor are iso-pulegol (17.2%), β-pinene<br />

(24.4%), myrcene (17.9%), piperitone (6.6%), α-pinene (7.3%), and pulegone (10.9%) (Sharma et al.<br />

1980). The minor components include limonene, linalool, linalyl acetate, phellandrene, piperitenone,<br />

and terpinene (Bor 1954; Gupta and Daniel 1982). The oil has been tested to be fungitoxic<br />

(Singh et al. 1980).


78 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

2.4.6.11 Cymbopogon parkeri<br />

The major constituents of the oil of C. parkeri Stapf are nerol (32%), geraniol (33%), farnesol (3.75%),<br />

geranyl acetate (8.9%), and neryl acetate (3.7%) (Rizk et al. 1985, 1983). The trace constituents<br />

reported are decane, eudesmol, geraniol, geranyl heptanoate, geranyl hexanoate, geranyl octanoate,<br />

guaiol, β-gurjunene, limonene, linalool, 14-methyl-heptadecanone, 7-methyl-4-octanone, 12-methyltridecanone,<br />

neral, neryl butanoate, neryl hexanoate, neryl octanoate, piperitone, α-terpineol, and<br />

xylene (Gupta and Daniel 1982; Rizk et al. 1983, 1986). A study of the antifungal activity of C. parkeri<br />

essential oil was done on the growth of Rhizoctonia solani, Pyricularia orizea, and Fusarium<br />

oxysporum, three important phytopathogenic fungi.<br />

2.4.7 so M e lesser-kn o w n sP e C i e s<br />

2.4.7.1 Cymbopogon polyneuros<br />

The major constituents in this sweet-scented oil obtained from C. polyneuros (Steud.) Stapf. have<br />

been reported as limonene, perillaldehyde, and perillyl alcohol (Gupta and Daniel 1982; Thapa<br />

et al. 1971).<br />

2.4.7.2 Cymbopogon procerus<br />

Elemicin (34%) and pinene have been described by Gildemeister and Hoffmann (1956) as the major<br />

constituents of the oil from C. procerus (R. Br.) Domin.<br />

2.4.7.3 Cymbopogon rectus<br />

The major constituents of the oil of C. rectus A. Camus are geraniol (40%–60%), methylisoeugenol<br />

(30.5%), and α-pinene (Gildemeister and Hoffmann 1956).<br />

2.4.7.4 Cymbopogon sennarensis<br />

Gildemeister and Hoffmann (1956) have reported that the oil from C. sennarensis (Hochest.) Chiov.<br />

contains pinene and limonene (13%) and l-methenone-3 (45%).<br />

2.4.7.5 Cymbopogon stracheyi<br />

The essential oil of C. stracheyi (Hook. f.) Riaz and Jain bears a very strong aromatic note and contains<br />

geraniol, citral, geranyl acetate, citronellol, and piperitone (Gupta and Daniel 1982; Mathela<br />

1991; Thapa et al. 1971). Some analytical studies of plants growing in the Almora region of Uttar<br />

Pradesh exhibit the presence of piperitone and car-2-ene as major components, along with geraniol,<br />

α-copaene, β-elemene, caryophyllene, and calarene. Lohani et al. (1986) carried out detailed<br />

chemical investigation of its oil and revealed the presence of car-2-ene (29.4%) and piperitone<br />

(47.8%), along with several minor constituents such as acoradiene, β-bisabolene, α-cadinene, camphene,<br />

trans-caryophyllene, α-copaene, p- cymene, dihydro-α-capaene-8-ol, geraniol, β-elemene,<br />

α-himachalene, intermediol, juniper camphor, and nerolidol.<br />

2.4.7.6 Cymbopogon tortilis<br />

A group from China (Liu et al. 1981) reported that methyl eugenol (55%) is the principal component<br />

of the oil C. tortilis (Presl.) Hitche.<br />

2.4.7.7 Cymbopogon travancorensis<br />

Elimicin is the chief constituent of the oil from C. travancorensis Bor. Other constituents reported<br />

are borneol, elemol, camphene, limonene, and citral (Gupta and Daniel 1982; Mallavarapu et al.<br />

1992b; Menon 1956).


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 79<br />

2.4.7.8 Cymbopogon goeringii<br />

The essential oil obtained from C. goeringii (Steud.) A. Camus exhibits antiarrhythmic action on<br />

isolated guinea pig heart. The results from chemical analysis of the oil are not available (Liu and<br />

Feng 1989).<br />

2.4.7.9 Cymbopogon asmastonii<br />

The compounds reported from the essential oil of C. asmastonii are d-limonene (5.5%), carveol<br />

(69.5%), and a complex mixture of ketones (17.5%) (Manjoor-i-khuda et al. 1986).<br />

2.4.7.10 Cymbopogon giganteus<br />

The composition of the essential oil isolated by hydrodistillation from the leaves of C. giganteus<br />

Chiov. growing wild in Ivory Coast was determined by GC-RI, GC-MS, and 13 C-NMR after fractionation<br />

on silica gel. The oil was characterized by high contents of trans- and cis-p-mentha-<br />

2,8-dien-1-ols (18.4% and 8.7%, respectively), cis- and trans-p-mentha-1(7),8-dien-2-ols (16.0% and<br />

15.7%, respectively), and limonene (12.5%). A total of 46 components were identified, including<br />

25 compounds reported for the first time in the oils of this species (Boti et al. 2006). The essential<br />

oils of fresh flowers (2 samples), leaves, and stems of Cymbopogon giganteus (Hochest.) Chiov.<br />

from the Cameroon were investigated by GC and GC-MS. More than 55 components have been<br />

identified in the samples 1 (flowers sample 1), 2 (leaves), 3 (stems), and 4 (flowers sample 2) with<br />

main compounds possessing the p-menthane skeleton as follows: cis-p-mentha-1(7),8-then-2-ol (1:<br />

22.8%, 2: 27.7%, 3: 29.1%, 4: 20.5%), trans-p-mentha-1(7),8-dien-2-ol (1: 24.9%, 2: 21.6%, 3: 28.1%,<br />

4: 26.5%), trans-p-mentha-2,8-then-1-ol (1: 17.3%, 2: 22.1%, 3: 21.4%, 4: 16.3%), and cis-p-mentha-<br />

2,8-dien-1-ol (1: 8.3%, 2: 5.4%, 3: 4.6%, 4: 9.7%).<br />

The oils of several other species have also been distilled out, but their chemistry is still not well<br />

studied. These could well be potentially valuable aromatic crops. Further research is needed to<br />

reveal the chemistry of these uncommon grasses that can be brought under cultivation on a larger<br />

scale. The most interesting species are C. caesius, C. coloratus, C. distans, C. khasianus, C. microstachys,<br />

C. parkeri, and others. On the other hand, all species that have been chemically investigated<br />

or not need to be cultivated in a particular phytogeographical region and reinvestigated using<br />

modern spectroscopic methods such as GC-MS, IR, NMR, 13 C-NMR, etc. Popielas et al. (1991) and<br />

Vole et al. (1997) studied the oil from its inflorescence and identified 18 compounds, most of them<br />

belonging to oxygenated monoterpenes with p-menthadiene skeleton. It is expected for this species<br />

that chemotaxonomists will be enabled to classify all the species in a perfect and systematic manner<br />

so that the prevailing confusion regarding the correct chemotaxonomy of this large genus may be<br />

removed. Additional components in higher concentrations, responsible for the characteristic aroma<br />

impressions of the samples from C. giganteus, are especially limonene, trans-verbenol, and carvone<br />

as well as some other mono- and sesquiterpenes. Antimicrobial activities of the oils from the leaf,<br />

stem, and flowers were found against Gram-positive and Gram-negative bacteria as well as the yeast<br />

Candida albicans, and these results were discussed with the compositions of each sample (Leopold<br />

et al. 2007).<br />

2.4.8 bi o s y n t h e s i s o f te r P e n e s in Cy m b o p o g o n sP e C i e s<br />

Terpenoids are a class of compounds derived from the universal precursor isopentenyl diphosphate<br />

(IPP) and its allylic isomer dimethylallyldiphosphate (DMAPP), also called isoprene units<br />

(Scheme 2.1). Terpenoid building blocks are then formed through condensation of additional IPP<br />

moieties (C 5) via prenyltransferases. Monoterpenoids are derived from geranyl pyrophosphate (GPP,<br />

C 10), sesquiterpenoids are derived from farnesyl pyrophosphate (FPP, C 15), and diterpenoids are


80 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

H<br />

HO<br />

H<br />

H<br />

2<br />

*<br />

1<br />

* CHO<br />

6<br />

OH<br />

H<br />

OH<br />

OH<br />

Glucose-6-phosphate<br />

Glycolysis<br />

several steps<br />

HO<br />

C-2 in Acetyl-CoA and * in other structures are from C-1, C-6 of Glucose<br />

(a) Acetoacetyl-CoA thiolase (AACT)<br />

H<br />

O<br />

H<br />

Dihydroxyacetone<br />

phosphate<br />

H<br />

H<br />

1<br />

2<br />

O<br />

OH<br />

3<br />

* CH2OP Glyceraldehyde-3-phosphate<br />

NAD O<br />

+ NADH + H +<br />

O O<br />

(a)<br />

O<br />

C O<br />

C<br />

1<br />

2<br />

C<br />

1 SCoA<br />

C<br />

1 SCoA<br />

C<br />

Acetyl-CoA<br />

–<br />

1<br />

*<br />

*<br />

2<br />

CoA-SH CO2 2<br />

3<br />

CH3 O<br />

Acetoacetyl-CoA<br />

Pyruvate<br />

*6<br />

* CH2 OH<br />

(b)<br />

OH<br />

* 4<br />

3<br />

2 * (c)<br />

O<br />

5 1 (d)<br />

COOH<br />

SCoA<br />

3-hydroxy-3-methylglutaryl-CoA<br />

HO<br />

(b) HMG CoA Synthase<br />

(c) HMG-CoA reductase<br />

(d) NADPH<br />

6*<br />

* 4<br />

5<br />

3<br />

OH<br />

Mevalonate<br />

6 *<br />

3 5<br />

2<br />

*<br />

4<br />

*<br />

OPP<br />

Dimethylallyl diphosphate<br />

1<br />

* CH2OP 2 *<br />

(e) ATP<br />

1 (f) ATP<br />

COOH<br />

PPO<br />

scheme 2.1 MVA-pathway from glucose to IPP and DMAPP.<br />

(h)<br />

6*<br />

* 4<br />

5<br />

3<br />

OH<br />

2<br />

*<br />

1<br />

COOH<br />

MVA 5-diphosphate<br />

6 *<br />

(g)<br />

3 5<br />

2<br />

4<br />

*<br />

OPP<br />

*<br />

Isopentenyl diphosphate<br />

(e) Mevalonate kinase<br />

(f) Phosphomevalonate kinase<br />

(g) IPP synthase<br />

(h) IPP Isomerase


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 81<br />

derived from geranylgeranyl pyrophosphate (GGPP, C 20). Even higher-order terpenoids are possible<br />

through condensation of these intermediates to larger precursor moieties. For example, sterols are<br />

derived from the triterpenoid squalene (C 30), which contains six isoprene units through condensation<br />

of two molecules of FPP, and carotenoids (C 40) are largely formed through condensation of two<br />

molecules of GGPP to yield eight-isoprene-unit compounds. After the formation of the acyclic terpenoid<br />

structural building blocks (e.g., GPP, FPP, GGPP), terpene synthases act to generate the main<br />

terpene carbon skeleton. Additional transformations often involving oxidation, reduction, isomerization,<br />

and conjugation enzymes decorate or alter the main skeleton with varied functional groups<br />

to yield the tremendously diverse terpenoid family of compounds. The essential oils obtained from<br />

aroma-bearing plants mainly possess mono- and sesquiterpenoids in high percentage besides a few<br />

nonterpenoidal compounds. The essential oil obtained from various Cymbopogon species also contains<br />

a large number of mono- and sesquiterpenoids as discussed in this chapter.<br />

It has now been unequivocally proven that two distinct and independent biosynthetic routes exist<br />

to IPP and its allylic isomer DMAPP, the two building blocks for isoprenoids in plants. The cytosolic<br />

pathway is triggered by acetyl coenzyme A (Scheme 2.1) where the classical intermediate mevalonic<br />

acid is formed, which in turn converts into IPP and DMAPP. These further combine to elongate into<br />

sesquiterpenes (C 15) and triterpenes (C 30) (Newman and Chappell 1999). In contrast, the plastidial<br />

pathway (Eisenreich et al. 1998, 2001; Lichtenthaler 1999; Rohmer 1999) provides precursors for<br />

the biosynthesis of isoprene (C 5), mono- (C 10), di- (C 20), and tetraterpenes (C 40) (Lichtenthaler 1999;<br />

Eisenreich et al. 1997, Scheme 2.2). The pathway (Scheme 2.2) is initiated by the transketolase-type<br />

condensation of pyruvate (C-2 and C-3) and glyceraldehyde-3-phosphate to 1-deoxy-d-xylulose-5phosphate<br />

(DXP), followed by the isomerization and reduction of this intermediate to 2-C-methyl-derythritol-4-phosphate,<br />

formation of the cytidine 5′-diphosphate derivative, phosphorylation at C-2,<br />

and cyclization to 2-C-methyl-d-erythritol-2,4-cyclodiphosphate (MECDP) as the last defined step.<br />

This is further converted to 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate (HMBDP). HMBDP<br />

is converted to two C-5 units for further condensation (Scheme 2.3). Later, the genes encoding<br />

each enzyme of the plastid pathway up to formation of the cyclic diphosphate are isolated from<br />

plants and from eubacteria where the pathway exists (Takahashi et al. 1998; Bouvier et al. 1998;<br />

Rohdich et al. 1999, 2000; Sprenger et al. 1997; Lange and Croteau 1999; Little and Croteau 1999;<br />

Lange et al. 1998; Schwender et al. 1999; Kuzuyama et al. 2000a, 2000b; Lüttgen et al. 2000; Herz<br />

et al. 2000).<br />

The cymbopogons have been reported to possess about 100 monoterpenes and around 50 sesquiterpenes<br />

in all, and thus the MVA and DXP pathways are very likely to be present during the biosynthesis<br />

of these compounds in the cymbopogons. Not much biosynthetic studies have been carried<br />

out on these species. However, Akhila (1985) has conducted studies on the biosynthetic relationship<br />

of citral-trans and citral-cis using doubly labeled [ 14 C, 3 H] precursors. The results revealed that the<br />

leaf blades of C. flexuosus converted geraniol into citral-trans with the loss of pro-(1S)-hydrogen<br />

whereas nerol lost the pro-(1R)-hydrogen while being converted into citral-cis. Secondly, the citraltrans<br />

is converted into citral-cis and vice versa, and there is no separate route for the biosynthesis of<br />

either of the two aldehydes. The mechanism for interconversion has been shown in Scheme 2.4.<br />

The biosynthesis of three major components in C. winterianus has been studied by Akhila (1986)<br />

using 3 H- and 14 C-labeled precursors. Geraniol, citronellol, and citronellal formed in the blades of<br />

C. winterianus from doubly labeled mevalonic acid predominantly labeled only that C 5 moiety that<br />

was derived from IPP. This was believed to be due to the presence of a metabolic pool of DMAPP.<br />

These results later on support the newly discovered theory of a nonmevalonate pathway (DXP pathway)<br />

in which mevalonic acid is believed to be converted into IPP in cytosol. The very low incorporation<br />

of radioactivity into monoterpenes (C 10) may be the result of seepage of IPP through the<br />

plastidial membrane. Monoterpenes are believed to be biosynthesized in plastids. According to the<br />

reports, geraniol is converted to citronellol, which in turn is transformed into citronellal as per the<br />

mechanism shown in Scheme 2.5.


82 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

H<br />

H<br />

1<br />

2<br />

O<br />

OH<br />

O<br />

C O<br />

3 *<br />

CH2OP C<br />

–<br />

+<br />

1<br />

(a)<br />

1<br />

OH<br />

Glyceraldehyde<br />

moeity<br />

2<br />

O<br />

(b)<br />

OP<br />

Glyceraldehyde-3-phosphate<br />

3<br />

CH3 Pyruvate moeity O OH<br />

1-Deoxy-D-xylulose 5-phosphate<br />

Pyruvate<br />

1 *<br />

3<br />

2<br />

OH<br />

4<br />

OH OH<br />

(a) DXP Synthase<br />

(b) �iamine diposphate<br />

(c) DXP reductoisomerase<br />

(d) MEP-Cytidylyltransferase<br />

(e) CDP-Me kinase<br />

(f) MECDP Synthase<br />

(g) HMBDP Synthase<br />

5<br />

OP<br />

2-C-Methyl-D-Erythritol<br />

4-phosphate (MEP)<br />

1 *<br />

3<br />

2<br />

OH<br />

4<br />

OH OH<br />

5<br />

OPOPO<br />

1<br />

*<br />

HO<br />

O<br />

2<br />

HO<br />

4<br />

O<br />

3<br />

NH 2<br />

N<br />

Biosynthesis of several mono- and sesquiterpene skeletons and compounds has been worked out<br />

in various plant species (Akhila et al. 1980a, 1980b, 1980c, 1988a, 1988b, 1987a, 1987b, 1985, 1986,<br />

1990; Croteau et al. 1981; Crotaeau and Davis 2005). Though about 150 mono- and sesquiterpenes<br />

are present in all the Cymbopogon species collectively, biosynthetic experiments have not been carried<br />

out. Based on the literature, two comprehensive schemes (Schemes 2.6 and 2.7) have been made<br />

showing biosynthetic pathways to most of the mono- and sesquiterpene skeletons and compounds<br />

5<br />

N<br />

(c)<br />

OP<br />

H<br />

H<br />

OH OH<br />

4-(Cytidine 5'-diphospho)-2-C-methyl-D-Erythritol (CDP-Me)<br />

O<br />

1 *<br />

3<br />

2<br />

O<br />

P<br />

OH 4<br />

OH<br />

OH OH<br />

(d)<br />

O<br />

5<br />

P<br />

O<br />

H<br />

O<br />

-2-C-methyl-D-Erythritol<br />

2,4-cyclodiphosphate (MECDP)<br />

H<br />

1 *<br />

3<br />

2<br />

OP<br />

4<br />

O<br />

OH OH<br />

scheme 2.2 Nonmevalonate pathway (DXP) to monoterpenes.<br />

(f)<br />

(g)<br />

1<br />

*<br />

5<br />

2<br />

H<br />

2<br />

O<br />

(e)<br />

3<br />

O<br />

3<br />

OPOPO<br />

4<br />

H<br />

4<br />

5<br />

OH<br />

H<br />

OH<br />

O<br />

5<br />

H<br />

OP<br />

NH 2<br />

N<br />

H<br />

OH<br />

2-Phospho-4-(Cytidine 5'-diphospho)<br />

-2-C-methyl-D-Erythritol (CDP-Me2P)<br />

3<br />

1*<br />

* C is derived from<br />

Glyceraldehyde-3-phosphate<br />

is derived from Pyruvate<br />

2<br />

4<br />

5<br />

OPP<br />

OH<br />

1-hydroxy-2-methyl-2-(E)-butenyl<br />

-4-diphosphate (HMBDP)<br />

N<br />

O


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 83<br />

3<br />

3<br />

3<br />

OH<br />

H<br />

1<br />

*<br />

25<br />

1<br />

*<br />

2<br />

1<br />

*<br />

T<br />

OH<br />

T<br />

C<br />

*<br />

[1- 14 C, 3 H 1 ]-(1R)-Geraniol Citral-trans<br />

4<br />

4<br />

HMBDP<br />

DMAPP<br />

5<br />

*<br />

OPP<br />

OPP<br />

T<br />

IPP-isomerase<br />

scheme 2.3 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate (HMBDP) to IPP and DMAPP.<br />

OH<br />

H<br />

[1- 14 C, 3 H 1 ]-(1S)-Nerol Citral-cis<br />

derived therefrom. Possible biosynthetic pathways for cadinane, bisabolane, eudesmane, and gurjunane<br />

compounds have been shown in Schemes 2.8, 2.9, 2.10, and 2.11.<br />

cis-Farnesyl pyrophosphate (FPP), also known as 2,3-(Z)-farnesyl pyrophosphate (Scheme 2.12),<br />

has been considered as a universal intermediate starter for many sesquiterpenes. However, it isomerizes<br />

to its trans-isomer 2,3-(E)-FPP through enzymatic conversion, possibly via nerolidol. FPP also<br />

cyclizes to a major group of sesquiterpenes such as caryophyllenes and humulenes. For convenience<br />

and according to generally adopted nomenclature, the numbering of carbon atoms (1 to 15) has been<br />

O<br />

Unstable or enzyme<br />

bond intermediate<br />

T<br />

3<br />

*<br />

O<br />

T<br />

C<br />

*<br />

OH<br />

Free rotation at<br />

C-3<br />

3<br />

T<br />

Unstable or enzyme<br />

bond intermediate<br />

scheme 2.4 Biosynthesis of citral-trans and citral-cis from doubly labeled geraniol and nerol.<br />

3<br />

1<br />

*<br />

2<br />

4<br />

IPP<br />

5<br />

OPP<br />

*<br />

OH


84 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

PPO<br />

DMAPP<br />

IPP<br />

shown in FPP in Scheme 2.12. A nucleophilic attack by an enzyme at C-3 of trans-FPP triggers<br />

the cyclization process that forms the nine-member cyclic skeleton along with the cyclobutane ring<br />

with elimination of pyrophosphate from C-1. This is the penultimate precursor of β-caryophyllene<br />

and β-caryophyllene epoxide. Alternatively, an enzymatic attack at C-10 will facilitate the formation<br />

of the C11–C1 bond with the elimination of pyrophosphate from C-1 and formation of C 11-ring<br />

skeleton of α- and β-humulene.<br />

Alloaromadendrene and aromadendrene are a different group of sesquiterpenes present in<br />

the essential oils of cymbopogons. A biogenetic route has been proposed from the cis-FPP in<br />

Scheme 2.13. There are three double bonds in FPP inviting electrophiles (enzymes) to attack them<br />

at suitable positions depending on the accessibility of the enzyme to the site of the attack. This<br />

depends upon the stereochemistry and spatial arrangement of the hydrogens and other attached<br />

atoms such as oxygen and phosphorus. In this case, an enzyme is attracted at C-7 of ∆ 6,7 , which<br />

attaches to C-2 resulting in the formation of a cyclopentane ring, whereas the other enzyme attacks<br />

at C-10, enabling ∆ 10,11 to attack at C-1 and release FPP from there. This forms the aromadendrene<br />

skeleton. The re- and si-face attacks by ∆ 6,7 on C-2 generates the two isomers alloaromadendrene<br />

and aromadendrene. Another mechanism for the biosynthetic route to butenol and acoradienes has<br />

been illustrated in Scheme 2.14. These schemes are biogenetic speculations based on chemical considerations<br />

and mechanisms, and are very good experimental models to be verified by radiotracer<br />

techniques.<br />

2.4.9 bioloGiCal aC tiVitie s<br />

H<br />

OPP<br />

The essential oil obtained from various Cymbopogon species is widely used in the perfumery,<br />

cosmetic, food, and flavor industries. Besides, the oil possess several biological activities (Dikshit<br />

and Hussain 1984), which have been discussed in detail in Section (2.4.9), but some highlights are<br />

OPP<br />

Geranyl Pyrophosphate (GPP)<br />

Citronellal<br />

Geraniol<br />

Citronellol<br />

scheme 2.5 Biosynthesis of geraniol, citronellol, and citronellal from IPP + DMAPP in Cymbopogon<br />

winterianus.<br />

CHO<br />

OH<br />

OH


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 85<br />

(Z)<br />

(Z)<br />

(Z)<br />

(Z)<br />

O<br />

OH<br />

CHO<br />

(Z)<br />

(R)<br />

OH<br />

(E)<br />

(E)<br />

(R)<br />

Redoxinterconversion<br />

CHO<br />

(E)<br />

(Z)<br />

(a)<br />

(a)<br />

CH2OR (b)<br />

p-Cymene γ-Terpinene α-Terpinene Terpinolene o-Cymene<br />

Carvotanacetone<br />

(–)-Dihydrocarveol<br />

R=H, Geraniol<br />

Citral-trans<br />

Citral-cis<br />

R=H, Nerol; R=Ac, Neryl<br />

acetate;R=PP, Neryl<br />

pyrophosphate (b)<br />

7<br />

(Z)<br />

(Z)<br />

?<br />

(Z)<br />

2<br />

1<br />

(Z)<br />

O<br />

(Z)<br />

(Z)<br />

(S) O<br />

(a) 1,8-bond<br />

(S) OH<br />

(S)<br />

HO<br />

(Z)<br />

1<br />

(a)<br />

(S)<br />

6<br />

+<br />

3<br />

(S)<br />

4<br />

8<br />

(S)<br />

(R)<br />

4<br />

+<br />

(c)<br />

OR<br />

OH<br />

(a)<br />

+<br />

OH<br />

(S)<br />

Camphor<br />

(S)<br />

Isoborneol<br />

4S-(+)-Carvone<br />

(+)-cis-Carveol<br />

4R-(+)-Limonene<br />

Terpinyl carbonium ion/<br />

enzyme bond species<br />

(c)<br />

Terpinen-4-ol<br />

α-Terpineol<br />

(S) O<br />

8<br />

Terpinyl carbonium ion/<br />

enzyme bond species<br />

H2C H<br />

H3C OH<br />

1 (S)<br />

1 (S)<br />

+<br />

8<br />

8<br />

–<br />

R=PP Linalyl<br />

Pyropphosphate<br />

2<br />

1<br />

8<br />

1,2 Methyl<br />

Shift<br />

+<br />

2<br />

Skeleton<br />

rearrangement<br />

8<br />

1<br />

(Z)<br />

O<br />

(Z)<br />

HO<br />

(S)<br />

(Z)<br />

(Z)<br />

(Z)<br />

(a´)<br />

Camphene<br />

(R)<br />

Fenchone<br />

(S)<br />

(S)<br />

Enzyme bonded species<br />

(R)<br />

(R)<br />

(S)<br />

OR<br />

(E)<br />

(a´)<br />

O<br />

4R-(-)-Carvone<br />

(-)-trans-Carveol<br />

4S-(-)-Limonene<br />

Δ 4 −Carene<br />

Δ 3 −Carene<br />

OH<br />

7<br />

+ +<br />

2<br />

1<br />

Hydrogenase<br />

1,8-Cineole<br />

3<br />

Oxidase<br />

R=H, Geraniol; R=Ac,<br />

Geranyl acetate; R=PP,<br />

Geranyl pyrophosphate<br />

4<br />

HO 7<br />

1<br />

(S) (S) 2<br />

3<br />

4<br />

(R)<br />

OH<br />

7<br />

(S)<br />

1<br />

(S) 2<br />

3<br />

4<br />

(S)<br />

(R)<br />

CHO<br />

CH 2 OH<br />

8<br />

β-Terpineol<br />

8<br />

β-�ujene-4-ol<br />

8<br />

�ujyl alcohol<br />

Citronellal<br />

Citronellol<br />

scheme 2.6 Biosynthetic pathways to monoterpene hydrocarbons and oxygenated monoterpenes commonly found in various Cymbopogon species. The most widely<br />

occurring three possible precursors (C10) GPP, NPP, and LPP have been shown.


86 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

OH<br />

OPP<br />

Farnesyl Pyrophosphate<br />

α-Muurolene<br />

Cadalene<br />

β-Cadinene<br />

Cadinane Group<br />

Bisabolane Group<br />

α-bisabolene<br />

β-bisabolene<br />

Bisabolol<br />

O<br />

α-Cadinene γ-Cadinene<br />

Calamenene<br />

OH<br />

OH<br />

OH<br />

OH<br />

OH<br />

trans-trans-Farnesol<br />

Eudesmandiol Eudesmane Group<br />

β-Eudesmol<br />

α-oxibisabolene<br />

OH<br />

Cymbopol<br />

α-Cubebene<br />

γ-Cadinol<br />

CHO<br />

OH<br />

Farnesal<br />

Elemane Group<br />

Intermediol<br />

β-Elemene<br />

HO<br />

OH<br />

OH<br />

Dihydro−α-Copaene-8-ol<br />

α-Farenesene<br />

β-Farenesene<br />

trans-cis-Farnesol<br />

γ-elemene<br />

α-elemol<br />

scheme 2.7 Diversity in biosynthetic pathways to various sesquiterpene skeletons found in various Cymbopogon species from commonly known intermediate cisfarnesylpyrophosphate<br />

(FPP).


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 87<br />

(b)<br />

X (Enz:)<br />

(a) X (Enz:)<br />

H<br />

X<br />

Y(Enz:)<br />

OPP<br />

(a)<br />

(b)<br />

H<br />

X<br />

Y(Enz:)<br />

Y Cadinane skeleton<br />

Y<br />

Cadinane<br />

Compounds<br />

scheme 2.8 Cyclization of FPP to cadinane skeleton and compounds of cadinane group. X and Y denote<br />

attacking enzymes involved in biosynthesis.<br />

PPO<br />

PPO<br />

X –<br />

Enz<br />

cis-Farnesylpyrophosphate<br />

scheme 2.9 Cyclization of cis-FPP to monocyclic bisabolane compounds.<br />

cis-Farnesylpyrophosphate<br />

Y –<br />

H+<br />

H<br />

X<br />

H X<br />

Y a<br />

H2C H<br />

b<br />

a<br />

X<br />

b<br />

X<br />

Compounds of<br />

Eudesmane Group<br />

scheme 2.10 Possible biogenesis of compounds of eudesmane group in cymbopogons. X and Y denote<br />

enzymes or their biogenetic equivalents that make a nucleophilic attack on electron-deficient sites of the<br />

skeleton.<br />

Y<br />

H H<br />

Enz<br />

X<br />

H X H H H<br />

Compounds of Bisabolane<br />

Group are formed from<br />

elimination of Hydrogen,<br />

attack by OH<br />

Gurjunane<br />

compounds<br />

scheme 2.11 Biogenetic pathways to Gurjunane compounds after 1,2 methyl shifts and hydrogen loss<br />

while forming a cyclopropane ring.


88 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

13<br />

8<br />

9<br />

14<br />

7<br />

10<br />

11<br />

5<br />

5<br />

6<br />

1 (Z)<br />

2<br />

OPP<br />

4<br />

3<br />

15<br />

8<br />

9<br />

7<br />

10<br />

11<br />

6<br />

(E)<br />

2<br />

4<br />

Enz:<br />

3<br />

H<br />

15 C<br />

H2 12<br />

13<br />

12<br />

1<br />

OPP<br />

2,3-(Z)-Farnesyl<br />

pyrophosphate<br />

14<br />

2,3-(E)-Farnesyl<br />

pyrophosphate<br />

15 C<br />

H2 provided here to give the reader an overall view of the pharmacological properties of this oil. The<br />

important activities include antibacterial, antifungal, anticancer, pesticidal, anthelmintic, mosquito<br />

repellant, mosquito larvicidal, antiinflammatory, analgesic, and hypoglycemic. The main components<br />

such as monoterpenes are nonnutritive dietary components found in the essential oils of citrus<br />

fruits and other plants. A number of these dietary monoterpenes have antitumor activity. For example,<br />

d-limonene, which comprises >90% of orange peel oil, has chemopreventive activity against<br />

rodent mammary, skin, liver, lung mammary, lung and forestomach cancers when fed during the<br />

13<br />

8<br />

9<br />

14<br />

7<br />

10<br />

11<br />

6<br />

12<br />

Enzyme bonded species<br />

8<br />

14<br />

7<br />

6<br />

5<br />

4<br />

(S)<br />

7<br />

(R)<br />

O<br />

7<br />

9<br />

Enz:<br />

13<br />

10<br />

11<br />

12<br />

(E)<br />

2<br />

1<br />

3<br />

15C<br />

H2 OPP<br />

H<br />

13<br />

10<br />

11<br />

2<br />

1<br />

3<br />

13<br />

10<br />

11<br />

2<br />

1<br />

3<br />

2,3-(E)-Farnesyl<br />

pyrophosphate<br />

12<br />

β-Caryophyllene epoxide<br />

12<br />

β-Caryophyllene<br />

14<br />

14<br />

(Z)<br />

8<br />

H<br />

9<br />

H<br />

Enz<br />

7<br />

10<br />

11<br />

6<br />

5<br />

(E)<br />

2<br />

1<br />

4<br />

3<br />

15<br />

14<br />

(Z)<br />

13<br />

8<br />

7<br />

9<br />

10<br />

11<br />

12<br />

12<br />

6<br />

5<br />

4<br />

3<br />

(Z) 15<br />

7<br />

(Z)<br />

(R)<br />

HO<br />

10 2<br />

3<br />

13 12<br />

(Z)<br />

8<br />

9<br />

7<br />

6<br />

5<br />

α-Humulene<br />

(α-Caryophyllene) 13<br />

11 1<br />

10<br />

4<br />

12<br />

11<br />

3<br />

β-Caryophyllene alcohol<br />

13<br />

1 2<br />

(Z)<br />

15<br />

12<br />

β-Humulene<br />

scheme 2.12 Biosynthesis of caryophyllenes and humulenes.<br />

5<br />

2<br />

1<br />

4<br />

3<br />

Enz<br />

H


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 89<br />

Enz:<br />

8<br />

9<br />

14<br />

7<br />

10<br />

11<br />

6<br />

1<br />

5<br />

(Z)<br />

2<br />

OPP<br />

13 12<br />

2,3-(Z)-Farnesyl<br />

pyrophosphate<br />

3<br />

14<br />

H2C Enz<br />

8<br />

9<br />

7<br />

10<br />

11<br />

H<br />

H<br />

6<br />

1<br />

4<br />

3<br />

5<br />

(Z)<br />

2<br />

15<br />

13 12<br />

Enzyme bonded species<br />

1<br />

7<br />

11<br />

H<br />

12 13<br />

a-Himachalene<br />

4<br />

3<br />

15<br />

8<br />

9<br />

Enz:<br />

14<br />

Enz:<br />

7<br />

10<br />

11<br />

H<br />

1<br />

6<br />

2<br />

OPP<br />

3 H +<br />

initiation phase. In addition, perillyl alcohol has promotion phase chemopreventive activity against<br />

rat liver cancer, and germaniol has in vivo antitumor activity against murine leukemia cells. Perillyl<br />

alcohol and d-limonene also have chemotherapeutic activity against rodent mammary and pancreatic<br />

tumors. As a result, their cancer chemotherapeutic activities are under evaluation in Phase I<br />

clinical trials. Several mechanisms of action may account for the antitumor activities of monoterpenes<br />

(Crowell 1999).<br />

2.4.9.1 Pain reliever<br />

Cymbopogon winterianus (Poaceae) is used for its analgesic, anxiolytic, and anticonvulsant properties<br />

in Brazilian folk medicine. The cited report aimed to perform phytochemical screening and<br />

investigate the possible anticonvulsant effects of the essential oil from fresh leaves of C. winterianus<br />

in different models of epilepsy (Quintans-Júnior et al. 2008). Myrcene was identified as the active<br />

constituent responsible for the activity. The peripheral analgesic effect of myrcene was confirmed<br />

5<br />

(Z)<br />

13 12<br />

2,3-(Z)-Farnesyl<br />

pyrophosphate<br />

4<br />

15<br />

14<br />

H2C Enz<br />

H<br />

H<br />

5<br />

8<br />

7<br />

6 4<br />

9 H 1<br />

3<br />

10<br />

Enz<br />

2<br />

11<br />

13 12<br />

Enzyme bonded species<br />

Aromadendrene<br />

scheme 2.13 Biosynthesis of himachalene and aromadendrene.<br />

14<br />

H2C H<br />

Enz H<br />

8<br />

7<br />

6<br />

9 H 1<br />

10<br />

Enz<br />

11<br />

15<br />

13 12<br />

Enzyme bonded species<br />

H<br />

7<br />

5<br />

2<br />

Allo-aromadendrene<br />

4<br />

3<br />

15


90 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

8<br />

9<br />

13<br />

14<br />

7<br />

10<br />

11<br />

1<br />

6<br />

12<br />

Enz:<br />

5<br />

(Z)<br />

2<br />

OPP<br />

2,3-(Z)-Farnesyl<br />

pyrophosphate<br />

8<br />

9<br />

14<br />

7<br />

10<br />

11<br />

1<br />

6<br />

5<br />

2<br />

OPP<br />

4<br />

3<br />

H<br />

H<br />

15<br />

8<br />

9<br />

13<br />

13 12<br />

Enzyme bonded species or<br />

electronically motivated reaction<br />

14<br />

Enz:<br />

8<br />

9<br />

13<br />

7<br />

10<br />

11<br />

1<br />

6<br />

(Z)<br />

2<br />

OPP<br />

12<br />

5<br />

(Z)<br />

2,3-(Z)-Farnesyl<br />

pyrophosphate<br />

4<br />

3<br />

H<br />

4<br />

3<br />

H<br />

15<br />

15<br />

14<br />

7<br />

10<br />

11<br />

12<br />

Enz:<br />

12<br />

1<br />

6<br />

9<br />

8<br />

10<br />

11<br />

Acoradiene<br />

by testing a saturated compound preparation on the hyperalgesia induced by prostaglandin in the<br />

rat paw test, and upon the contortions induced by intraperitoneal (ip) infections of iloprost in mice<br />

(Lorenzetti et al. 1991). Oral administration of an infusion of C. citratus fresh leaves to rats produced<br />

a dose-dependent analgesia for the hyperalgesia induced by subplanter infections of either caragenin<br />

or prostaglandin E2, but did not affect that induced by dibutyryl cyclic AMP. These results indicated<br />

a peripheral site of action. In contrast to the central analgesic effect of morphine, myrcene<br />

did not cause tolerance on repeated infection in rats. This analgesic property supports the use of<br />

lemongrass tea as a sedative in folk medicine. It is suggested that terpenes such as myrcene may<br />

constitute a lead for the development of new peripheral analgesics with a profile of action different<br />

from that of aspirin-like drugs.<br />

5<br />

(Z)<br />

Enzyme bonded species<br />

Enz<br />

12<br />

C<br />

H2 scheme 2.14 Biosynthetic routes to butenol and acoradiene.<br />

13<br />

8<br />

9<br />

14<br />

7<br />

10<br />

11<br />

H<br />

2<br />

14<br />

7<br />

2<br />

1<br />

13<br />

4<br />

3<br />

H<br />

OH<br />

6<br />

3<br />

15<br />

15<br />

H<br />

5<br />

4<br />

10<br />

α-butenol β-butenol<br />

6<br />

1<br />

5<br />

(Z)<br />

2<br />

H<br />

4<br />

3<br />

15<br />

8<br />

9<br />

9<br />

13<br />

14<br />

8<br />

7<br />

10<br />

14<br />

2<br />

1<br />

11<br />

6<br />

6<br />

5<br />

15<br />

2<br />

12<br />

5<br />

4<br />

4<br />

3<br />

H<br />

15<br />

�is intermediate is not possible<br />

stereochemically hence another<br />

mechanism is proposed<br />

12<br />

10<br />

11<br />

7<br />

1<br />

13<br />

1<br />

6<br />

3<br />

OH<br />

15


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 91<br />

2.4.9.2 activity against leukemia and malignancy<br />

The essential oil from a lemongrass variety of Cymbopogon flexuosus (CFO) and its major chemical<br />

constituent sesquiterpene isointermedeol (ISO) were investigated for their ability to induce apoptosis<br />

in human leukemia HL-60 cells because dysregulation of apoptosis is the hallmark of cancer<br />

cells. CFO and ISO inhibited cell proliferation with 48 h IC 50 of ~30 and 20 μg/mL, respectively<br />

(Kumar et al. 2008). Two active compounds, d-limonene and geraniol, were isolated by glutathione-<br />

S-transferase (GST) assay and fractionation of lemongrass (C. citratus) oil. These were tested for<br />

their capacity to induce activity of the detoxifying enzyme GST in several tissues of the female A/J<br />

mice. d-Limonene increased GST activity two- to threefold than controls in the mouse liver and<br />

the mucosa of the small and large intestines. Geraniol showed high GST- inducing activity in the<br />

mucosa of the small and large intestines, which was about 2.5-fold greater than controls. Induction<br />

of increased GST activity, which is believed to be a major mechanism for chemical carcinogen<br />

detoxification, has been recognized as one of the characteristics of the action of anticarcinogens<br />

(Zheng et al. 1993).<br />

The essential oil from C. citratus and its isolated principal citral have been tested for cytotoxicity<br />

against P388 leukemia cells. The cytotoxicity of citral, IC 50 against P388 mouse leukemia cells was<br />

71 µg/mL (Dubey et al. 1997). In another experiment, IC 50 of C. citratus oil in P388 leukemia cells<br />

was found to be 5.7 µg/mL (Dubey et al. 1997).<br />

2.4.9.3 activation of male hormones<br />

The antimale sex hormone agent is a 5-reductase inhibitor that converts testosterone to active dihydrotestosterone.<br />

This agent is extracted from the leaves, stems, rhizomes, roots, or whole plant of<br />

C. flexuosus. The composition containing the antimale sex hormone agent is especially useful as<br />

hair growth stimulants (Kisaki et al. 1998).<br />

2.4.9.4 activity against Worms<br />

The essential oil from C. martinii var. motia, in varying concentrations (0–0.4%), have shown to<br />

have good to excellent anthelmintic activity against tapeworms, round worms, and earthworms in<br />

in vitro tests (Sangwan et al. 1985). The activity exceeded that of the drug piperazine phosphate<br />

(Siddiqui and Garg 1990). Helminthiasis is one of the most important groups of parasitic diseases<br />

in the Indo–Pakistan subcontinent, resulting in heavy production losses in livestock. A wide variety<br />

of anthelmintics is used for the treatment of helminths in animals. However, the development of<br />

resistance in hel minths against commonly used anthelmintics has always been a challenge faced by<br />

the animal health care professionals. Therefore, exploitation of the anthelmintic potential of plants<br />

indigenous to the Indo–Pakistan subcontinent is an area of research interest (Akhtar et al. 2000).<br />

2.4.9.5 lowering blood sugar<br />

A study was designed to investigate the hypoglycemic and hypolipidemic effects of the single daily<br />

oral dosing of 125–500 mg/kg of fresh leaf aqueous extract of Cymbopogon citratus Stapf (CCi) in<br />

normal male Wistar rats for 42 days (Adeneye and Agbaje 2007). In another report, Cymbopogon<br />

proximus herb was assessed by Eskandar and Won Jun (1995) for hypoglycemic and hyperinsulinemic<br />

action on alloxan diabetic rats. A dose of 1.5 mL of herb suspension/100 g B weight was orally<br />

administered to the rats for intervals of 4, 8, and 16 days. The results revealed that considerable<br />

hypoglycemic effect was exerted after 16 days. The level of serum insulin was also increased in<br />

diabetic rats.<br />

2.4.9.6 Potential to repel mosquitoes and Kill the larvae<br />

Ointment and cream formulations of lemongrass oil in different classes of base and the oil in liquid<br />

paraffin solution have been evaluated for mosquito repellency in a topical application. Mosquito


92 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

repellency was tested by determining the bite deterrence of product samples applied on an experimental<br />

bird’s skin against a 2-day-starved culture of Aedes aegypti L. mosquitoes. The 1% v/v<br />

solution and 15% v/w cream and ointment preparations of the oil exhibited ≥50% repellency lasting<br />

2–3 h, which may be attributed to citral, a major oil constituent. This activity was comparable to that<br />

of a commercial mosquito repellent. Base properties of the lemongrass oil formulations influenced<br />

their effectiveness. The oil demonstrated efficacy from the different bases in the order of hydrophilic<br />

base > emulsion base > oleaginous base (Oyedele et al. 2002). Ansari and Razdan (1995) studied<br />

various essential oils for their mosquito-repellent activity. <strong>Essential</strong> oils from C. martinii var. sofia,<br />

C. citratus, and C. nardus were found as effective as chemical base oil. The percent protection<br />

against Culex quinquefasciatus ranged third 95%–96%. Fractional distillation of Ceylon citronella<br />

(C. nardus) oil yielded 13 fractions. Monoterpene hydrocarbon fractions were highly lethal to late<br />

third instar Culex quinquefasciatus larvae. The results suggested that myrcene was responsible for<br />

this activity. Elemol and/or methyl iso-eugenol were identified as active larvicidal principles in the<br />

latter fractions. The residue after the fractional distillation also possessed considerable larvicidal<br />

activity (Ranaweera and Dayananda 1997).<br />

2.4.9.7 activity to reduce edema<br />

The species of Cymbopogon giganteus is widely used in traditional medicine against several diseases.<br />

This study reports the inhibitory effect produced by the chemical constituents of the essential<br />

oil from leaves of C. giganteus of Benin in vitro on 5-lipoxygenase, and has been found useful<br />

as an antiinflammatory agent. The scientists assayed the antiradical scavenging activity of the<br />

sample by the 1,1-diphenyl-2-picrylhydrazyl (DPPH) method (Alitonou et al. 2006). Earlier, the<br />

antiinflammatory activity of the oil was attributed to the inhibition of the prostaglandin pathway<br />

(Krishnamoorthy et al. 1998). Oral administration of essential oils extracted from C. martinii<br />

leaves produced dose-dependent inhibition of carrageenan-induced paw edema in experimental<br />

male albino rats.<br />

2.4.9.8 Potential to control aging Process<br />

Leaves from cultivated Cymbopogon citratus were extracted with methanol, 80% aqueous ethanol,<br />

and water (infusion and decoction), and the extracts were assessed for their antiradical capacity<br />

by 2,2-diphenyl-1-picrylhydrazyl (DPPHʹ) assay; the infusion extract exhibited the strongest<br />

activity. Tannins, phenolic acids (caffeic and p-coumaric acid derivatives), and flavone glycosides<br />

(apigenin and luteolin derivatives) were identified in three different fractions obtained from an<br />

essential-oil-free infusion, and a correlation with their scavenger capacity for reactive oxygen species<br />

was studied. The tannin and flavonoid fractions were the most active against species involved<br />

in oxidative damage processes. In the flavonoid fraction, representing 6.1% of the extract, 13 compounds<br />

(O- and C-glycosylflavones) were tentatively identified by high-performance liquid chromatography,<br />

coupled to photodiode-array and electrospray ionization mass spectrometry detectors<br />

(HPLC–PDA–ESI/MS), nine of which were identified for the first time in this plant, all of them<br />

being C-glycosylflavones (mono-C-, di-C- and O,C-diglycosylflavones). The potential beneficial<br />

and protective value of the identified polyphenols for human health is discussed (Figueirinha et al.<br />

2008). Hyalurodinase inhibitors are extracted from C. nardus and some other plants for the preparation<br />

of cosmetics. Hyalurodinase inhibitors prevented the degradation of aging-related hyaluronic<br />

acid (Namba et al. 1995). Antioxidant activity of C. schoenanthus was measured by DPPH assay.<br />

The results ranged from 36.0% to 73.8% (2 μL of essential oil per milliliter of test solution). The<br />

antioxidant activity was also assayed using the β-carotene–linoleic acid bleaching method. The best<br />

results (IC 50 = 0.47 ± 0.04 mg mL −1 ) were obtained with the fresh leaves of plants collected in the<br />

desert region. The greatest acetylcholinesterase inhibitory activity (IC 50 = 0.26 ± 0.03 mg mL −1 ) was<br />

exhibited by the essential oil of the fresh leaves from the mountain region (Khadri et al. 2008).


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 93<br />

2.4.9.9 activity against Pests<br />

The susceptibility of Spodoptera litura larvae to different concentrations (0.2%–0.8%) of the essential<br />

oil of C. citratus has been studied in relation to host plant resistance in peanut. Field trials indicated<br />

that larvae developing on the most susceptible variety had the lowest mortality due to biopesticide<br />

lemongrass oil. The larvae treated with the oil before feeding showed significant higher mortality on<br />

the diet containing resistant pods than on that containing susceptible pods (Rajapakse and Jayasena<br />

1991). The essential oils from C. martinii var. motia, C. flexuosus, and C. winterianus are reported<br />

to possess insect-repellant, nematicidal, and insect-attractant properties (Ahmad et al. 1993).<br />

A number of essential oils including citronella (C. winterianus) and palmarosa (C. martinii)<br />

showed pesticidal activity against the stored grain insect Tribolium castaneum (Naik et al. 1995).<br />

Within a storage period of 10 days, samples of maize and cowpea treated with lemongrass powder<br />

and essential oil showed no physical deterioration (Adegoke and Odesola 1996).<br />

2.4.9.10 activity against microbes<br />

The essential oil of C. martinii var. motia and its different dilutions have shown significant antibacterial<br />

activity against Staphylococcus aureus, S. pyagens, E. coli, and Corynebacterium ovis<br />

(Gangrade et al. 1990).<br />

The essential oils of lemongrass (C. citratus), palmarosa (C. martinii var. motia), and khavi<br />

grass (C. jwarancusa) were tested for antibacterial property against E. coli, S. aureus, Shigella flexneri,<br />

and Salmonella typhi. Lemongrass oil was the most active and caused complete inhibition of<br />

S. aureus at less than 400 ppm. Palmarosa was more active against S. flexneri and S. typhi, whereas<br />

khavi grass showed less activity than lemongrass and palmarosa. The activity of the oil might be<br />

attributed to the components citral, geraniol, and piperitone (Syeed et al. 1990).<br />

The essential oil from the leaves of C. martinii was tested for toxicity against Fusarium oxysporum.<br />

Toxicity was the strongest as the mycelial growth of the pathogen was inhibited. Fungitoxicity<br />

remained unchanged in temperature treatment after a long storage period. It had no effect on the<br />

Cajanus cajan plant (Shrivastava et al. 1990). The essential oil of C. nardus exhibited a very good<br />

order of antifungal activity (Lemos et al. 1994).<br />

The effect of auto-oxidation of lemongrass (C. citratus) oil on its antibacterial activity was studied.<br />

Using the active oxygen method, the oil was found to undergo rapid oxidation under accelerated<br />

test conditions. The oxidized oil samples were found to have reduced activity against bacteria. The<br />

activity was completely lost in extensively oxidized oil samples. Inclusion of antioxidants in the oil<br />

samples reduced the rate of oxidation and enhanced the antibacterial activity of the oil. The effects<br />

of the antioxidants were concentration-dependent, and at their effective concentration oxidation was<br />

completely prevented for the period of the test (Orafidiya 1993).<br />

The inhibitory effect of lemongrass (C. flexuosus) essential oil isolated from local and Thai<br />

cultivars against pathogenic fungi were reviewed. No significant difference was found. The oil completely<br />

inhibited the growth of Monilia sitophilia, Penicillium digilotum, Aspergillus parasiticus,<br />

A. niger, and A. fungis (Shadab-Qamar et al. 1992).<br />

The minimum inhibitory concentration (MIC) and minimum lethal concentration (MLC) of the<br />

oil obtained from lemongrass (C. citratus), and citral against 35 clinical isolates of four dermatophytes<br />

were determined by the agar dilution method. The MIC and MLC of lemongrass oil were<br />

found to be higher than those of citral. The mode of action of lemongrass oil and citral were proven<br />

to be fungicidal. A comparative study of efficacy of cream containing four different concentrations<br />

of lemongrass oil was performed in vitro by hole diffusion assay. The 2.5% lemongrass oil was<br />

demonstrated to be the minimum concentration for the preparation of an antifungal cream for subsequent<br />

clinical study (Wannissorn et al. 1996).<br />

The essential oil from several plants, including lemongrass (C. citratus), were tested for antimicrobial<br />

activities against Paenibacillus larvae, the causal agent of American Foul Broad (AFB)


94 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

disease of honeybees. Trials for determining the MIC of the oil revealed that lemongrass and<br />

thyme were most effective. The results indicated that lemongrass and thyme oils could be used as<br />

effective inhibitors of AFB in honeybee colonies (Allpi et al. 1996).<br />

Thirteen essential oils of African origin (mostly from the Cameroon) were correlated with the<br />

antimicrobial activities of the oil toward six microbial strains: S. aureus, B. coli, Proteus mirabilis,<br />

Klebsiella pneumoniae, Candida albicans, and Pseudomonas aeruginosa. The oil of C. citratus<br />

displayed noteworthy antifungal and antibacterial properties (Chalchat et al. 1997). Fresh oil as<br />

well as 2-, 7-, and 12-year-old oils of a local variety of lemongrass (C. citratus) were distilled and<br />

redistilled, and tested against B. coli, S. aureus, Shigella flexneri, Salmonella typhi, Para-A, and<br />

Klebsiella pneumonae. The oil that was kept for two years exhibited, after redistillation, maximum<br />

activity due to its high citral content. S. flexneri and S. typhi were inhibited effectively at low doses<br />

of the oil. The inhibition appeared to be mostly by the citral content of the oil (Syeed et al. 1990).<br />

During the screening of some aromatic plants for fungitoxicity of their volatile oils, C. pendulus<br />

var. Praman exhibited the strongest activity, completely inhibiting the mycelial growth of the test<br />

organisms Microsporuin gypseum and Trichophyton mentagrophytes. The volatile oil distilled from<br />

fresh leaves was found to be fungicidal at its MIC of 200 µg/mL, inhibiting heavy inocula of the<br />

test fungi. During the testing of its fungitoxic spectrum, it also inhibited mycelial growth of three<br />

other fungi and was found to be more active than some commercial drugs tested (Pandey et al.<br />

1996). <strong>Essential</strong> oils obtained from the leaves of 29 medicinal plants commonly used in Brazil were<br />

screened against 13 different E. coli serotypes. The oils were obtained by water distillation using a<br />

Clevenger-type system, and their MIC was determined by the microdilution method. <strong>Essential</strong> oil<br />

from C. martinii exhibited a broad inhibition spectrum, presenting strong activity (MIC between<br />

100 and 500 μg/mL) against 10 out of 13 E. coli serotypes: three enterotoxigenic, two enteropathogenic,<br />

three enteroinvasive, and two shiga-toxin producers. C. winterianus strongly inhibited two<br />

enterotoxigenic, one enteropathogenic, one enteroinvasive, and one shiga-toxin producer serotypes.<br />

Aloysia triphylla also shows good potential to kill E. coli with moderate-to-strong inhibition. Other<br />

essential oils showed antimicrobial properties, although with a more restricted action against the<br />

serotypes studied. Chemical analysis of C. martinii essential oil performed by GC and GC–MS<br />

showed the presence of compounds with known antimicrobial activity, including geraniol, geranyl<br />

acetate, and trans-cariophyllene, which, tested separately, indicated geraniol as antimicrobial active<br />

compound. The significant antibacterial activity of C. martinii oil suggests that it could serve as a<br />

source for compounds with therapeutic potential (Duarte et al. 2006).<br />

An essential oil from a lemongrass variety of C. flexuosus (CFO) and its major chemical constituent<br />

sesquiterpene isointermedeol (ISO) were investigated for their ability to induce apoptosis<br />

in human leukemia HL-60 cells because dysregulation of apoptosis is the hallmark of cancer cells.<br />

CFO and ISO inhibited cell proliferation with 48 h IC of ~30 and 20 μg/mL, respectively. Both<br />

induced concentration-dependent strong and early apoptosis as measured by various endpoints,<br />

for example, annexin V binding, DNA laddering, apoptotic bodies formation, and an increase in<br />

hypodiploid sub-G0 DNA content during the early 6 h period of study. This could be because<br />

of early surge in reactive oxygen species (ROS) formation with concurrent loss of mitochondrial<br />

membrane potential observed. Both CFO and ISO activated apical death receptors TNFR1, DR4,<br />

and caspase-8 activity. Simultaneously, both increased the expression of mitochondrial cytochrome<br />

c protein with its concomitant release to cytosol leading to caspase-9 activation, suggesting thereby<br />

the involvement of both the intrinsic and extrinsic pathways of apoptosis. Further, Bax translocation<br />

and decrease in nuclear NF-κB expression predict multi target effects of the essential oil and<br />

ISO while both appeared to follow similar signaling apoptosis pathways. The easy and abundant<br />

availability of the oil combined with its suggested mechanism of cytotoxicity makes CFO highly<br />

useful in the development of anticancer therapeutics (Kumar et al. 2008). The essential oil from a<br />

lemon grass variety of Cymbopogon flexuosus was studied for its in vitro cytotoxicity against twelve<br />

human cancer cell lines. The in vivo anticancer activity of the oil was also studied using both solid<br />

and ascitic Ehrlich and Sarcoma-180 tumor models in mice. In addition, the morphological changes


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 95<br />

in tumor cells were studied to ascertain the mechanism of cell death. The in vitro cytotoxicity studies<br />

showed dose-dependent effects against various human cancer cell lines (Sharma et al. 2009).<br />

C. winterianus (Poaceae) is used for its analgesic, anxiolytic, and anticonvulsant properties in<br />

Brazilian folk medicine and these reports are aimed to perform phytochemical screening and to<br />

investigate the possible anticonvulsant effects of the essential oil from fresh leaves of C. winterianus<br />

in different models of epilepsy. The phytochemical analysis of the oil showed the presence of<br />

geraniol (40.06%), citronellal (27.44%), and citronellol (10.45%) as the main compounds. A behavioral<br />

screening demonstrated that the essential oil (100, 200, and 400 mg/kg, ip) caused depressant<br />

activity on CNS. When administered concurrently (200 and 400 mg/kg, ip) it significantly reduced<br />

the number of animals that exhibited PTZ- and PIC-induced seizures in 50% of the experimental<br />

animals (p < 0.05). Additionally, EO (100, 200, and 400 mg/kg, ip) significantly increased (p <<br />

0.05) the latencies of clonic seizures induced by STR. Our results demonstrated a possible anticonvulsant<br />

activity of the essential oil (Quintans-Júnior et al. 2008).<br />

reFerences<br />

Abdullah MA, Foda YH, Saleh M, Zaki MSA, Mostafa MM. 1975. Identification of the volatile constituents<br />

of the Egyptian lemongrass oil I. Gas chromatographic analysis. Nahrung 19: 195–200 (C.A., 83,<br />

48089 u).<br />

Abegaz B, Yohannes PG, Deiter RK. 1983. Constituents of essential oil of Ethiopian Cymbopogon citratus<br />

Stapf. J Nat Prod 46: 424–426.<br />

Adegoke GO, Odesola BA. 1996. Storage of maize and cowpea and inhibition of microbial agents of biodeterioration<br />

using the powder and essential oil of lemongrass (C. citratus). Int Biodeterioration<br />

Biodegrad 37: 81–84.<br />

Adeleke AK, Adebola OO, Adeolu OA. 2001. Volatile leaf oil constituents of Cymbopogon citratus (DC) Stapf.<br />

Flav Frag J 16(5): 377, 378.<br />

Adeneye AA, Agbaje EO. 2007. Hypoglycemic and hypolipidemic effects of fresh leaf aqueous extract of<br />

Cymbopogon citratus Stapf. in rats. J Ethnopharmacol 112(3): 440–444.<br />

Ahmad R, Sheikh V, Ahmad A, Ahmad M. 1993. <strong>Essential</strong> oils as insect and animal attractant and repellants.<br />

Hamdard Med 36: 99–105.<br />

Ahmed ZF, Rizk AM, Hammouda FM. 1970. Postep Dziedzinie Leku Rosi. Proc Ref Dosw Wygloszone Symp<br />

20–23 (C.A.78: 94856-u).<br />

Akhila A, Francis MJO, Banthorpe DV. 1980a. Biosynthesis of carvone in Mentha spicata. Phytochemistry 19:<br />

1433–1437.<br />

Akhila A, Rani K, Thakur RS. 1990. Biosynthesis of artemisinic acid in Artemisia annua. Phytochemistry<br />

29(7): 2129–2132.<br />

Akhila A, Sharma PK, Thakur RS. 1988a. 1,2-Hydrogen shifts in the formation of Sesquithujane skeleton in<br />

ginger rhizomes. Phytochemistry 27(11): 3471–3473.<br />

Akhila A, Sharma PK, Thakur RS. 1987a. 1,2-Hydrogen shifts during the biosynthesis of Patchoulenes in<br />

Pogostemon cablin. Phytochemistry 26(10): 2705–2707.<br />

Akhila A, Sharma PK, Thakur RS. 1988b. Biosynthetic relationship in Patchouli alcohol, Seychellene and<br />

Cycloseychellene in Pogostemon cablin. Phytochemistry 27(7): 2105–2108.<br />

Akhila A. 1980b. Banthorpe DV: Biosynthetic origin of gem-methyls of geraniol. Phytochemistry 19:<br />

1429–1430.<br />

Akhila A. 1980c. Biosynthesis of the skeleton of Pulegone in Mentha pulegium. Pflanzenphysiol Zeitschrift fur<br />

99(3): 277–282.<br />

Akhila A. 1985. Biosynthetic relationship of citral-trans and citral-cis in Cymbopogon flexuosus (Lemongrass).<br />

Phytochemistry 24(11): 2585–2587.<br />

Akhila A. 1986. Biosynthesis of monoterpenes in Cymbopogon winterianus. Phytochemistry 25(2): 421–424.<br />

Akhila A. 1987b. Biosynthesis of menthol and related monoterpenes in Mentha arvensis. J Plant Physiol<br />

126(4/5): 379–386.<br />

Akhtar MS, Iqbal Z, Khan MN, Lateef M. 2000. Anthelmintic activity of medicinal plants with particular reference<br />

to their use in animals in the Indo–Pakistan subcontinent. Small Ruminant Res 38(2): 99–107.<br />

Alam K, Agua T, Manen H, Taie R, Rao KS, Burrows I, Huber ME, Rali T. 1994. Preliminary screening of sea<br />

weeds, sea grass for antimicrobial and antifungal activity. J Pharmacognosy 32: 169–399.


96 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Alitonou GA, Avlessi F, Sohounhloue DK, Agnaniet H, Bessiere JM, Menut C. 2006. Investigations on the<br />

essential oil of Cymbopogon giganteus from Benin for its potential use as an anti-inflammatory agent. Int<br />

J Aromatherapy 16(1): 37–41.<br />

Allpi AM, Ringuilet JA, Ceremele El, Re MS, Henning CP. 1996. Antimicrobial activity of some essential oil<br />

against Paenibacillus larvae, the causal agent of American foulbrood disease. J Herbs Spice Med Plants<br />

4: 9–16.<br />

Anonymous. 1950. The Wealth of India, Vol. II, pp. 411–419, CSIR, New Delhi.<br />

Anonymous. 1952. Indian Standard Specifications, Indian Standards Institutions, New Delhi.<br />

Anonymous. 1958. <strong>Essential</strong> <strong>Oil</strong>s and Aromatic Chemicals, Symposium at Dehradun, CSIR, New Delhi.<br />

Anonymous. 1973. Application of gas liquid chromatography to the analysis of essential oil III. Determination<br />

of geraniol in oils of citronella. Analyst 98: 823–829.<br />

Anonymous. 1978. Insect repellent complex. Pat Ger Offen 2, 752, 140, 24 May 1978 (C.A. 89: 175043 f).<br />

Anonymous. 1980. Terpenoids from palmarosa grass (Cymbopogon martinii var. motia) R.A K. Riechst. Aromen<br />

Kosmet 30: 20–22 (C.A. 92: 160599e).<br />

Ansari MA, Razdan RK. 1995. Relative efficacy of various oils in repelling mosquitoes. Indian J Mal 32:<br />

104–111.<br />

Ansari SH, Ali M, Siddiqui AA. 1996. Evaluation of chemical constituents and trade potential of Cymbopogon<br />

citratus (Lemongrass). Hamdard Med 39: 55–59.<br />

Ansari SH, Quadry JS. 1987. TLC and GLC studies on Khawi grass oil. Indian J Nat Prod 3: 10–12.<br />

Arctander S. 1960. Perfume and Flavour Materials of Natural Origin. Allured Publishing, Elizabeth, New Jersey.<br />

Arora R, Pandey GN. 1977. The application of essential oils and their isolates for blue mold decay control in<br />

Citrus reticulata Blanco. J Food Sci Technol 14: 14–16.<br />

Atal CK, Bradu BL. 1976a. Search for aroma chemicals of industrial value from genus Cymbopogon. Part II.<br />

C. pendulus (Nees ex Steud.) Wats. (Jammu lemongrass)—a new superior source of citral. Indian J<br />

Pharmacol 38: 61–63.<br />

Atal CK, Bradu BL. 1976b. Search for aroma chemicals of industrial value from genus Cymbopogon Pt. III:<br />

Cymbopogon pendulus (Nees ex Steud.) Wats. (Jammu lemongrass), a new superior source of citral.<br />

Indian Perfumer 20: 29–33.<br />

Atal CK, Bradu BL. 1976c. Search for aroma chemicals of industrial value from genus Cymbopogon Pt. IV.<br />

Chandi and Kolar grasses as source of methyl eugenol. Indian Perfumer 20: 35–38.<br />

Atal CK, Bradu BL. 1976d. Search for aroma chemicals of industrial value from Cymbopogon Part V. Chandi<br />

and Kolar grasses as source of methyl eugenol. Indian J Pharm 38: 63, 64.<br />

Baiswara RB, Nair KNG, Mathew TV. 1976a. Detection of adulteration of palmarosa oil in ginger grass oil by<br />

thin layer chromatography. Res Ind 21: 37–39.<br />

Baiswara RB, Nair KNG, Mathew TV. 1976b. Detection of ginger grass oils in lemongrass oil by thin layer<br />

chromatography. Res Ind 21: 39, 40.<br />

Balyan SS, Shahi AK, Choudhury SN, Singh A, Atal CK. 1979. Performance of five strains of genus<br />

Cymbopogon at Jammu. Indian Perfumer 23: 121–124.<br />

Baruah P, Mishra BP, Pathak MG, Ghosh AC. 1995. Dynamics of essential oil of Cymbopogon citrates (D.C.)<br />

Stapf. under rust disease indices. J Essen <strong>Oil</strong> Res 7: 337–338.<br />

Baslas RK, Baslas KK. 1968. <strong>Essential</strong> oil potentialities of Kumaon. Indian Perfumer 12: 3–6.<br />

Baslas RK. 1970. Chemistry of Indian essential oils. Part IX. Flav Industry 1: 475–478.<br />

Beauchamp PS, Dev V, Docter DR, Ehsani R, Vita G, Melkani AB, Mathela CS, Bottini AT. 1996. Comparative<br />

investigation of the sesquiterpenoids present in the leaf oil of Cymbopogon distans (Steud.)<br />

Wats. var. loharkhet and the root oil of Cymbopogon jwarancusa (Jones) Schult. J Essen <strong>Oil</strong> Res 8:<br />

117–121.<br />

Beech DF. 1977. Growth and oil production of lemongrass (C. citratus) in the Ord Irrigation Area, Western<br />

Australia. Aust J Exp Agric Anim Husb 17: 301–307.<br />

Bhattacharya AK, Kaul PN, Rao BRR, Mallavarapu GR, Ramesh S. 1997. Interspecific and inter cultivar variations<br />

in the essential oil profiles of lemongrass. J Essen <strong>Oil</strong> Res 9: 361–364.<br />

Bhattacharya SC. 1970. Perfumery chemicals from indigenous raw materials. J Indian Chem Soc 47:<br />

307–313.<br />

Bhaumik SK, Rao VS. 1978a. Studies on prehydrolysis sulfate pulp from Cymbopogon grass (oil extracted).<br />

Indian Pulp Pap 33: 3–5.<br />

Bhaumik SK, Rao VS. 1978b. Evaluation of citronella grass (waste) as a raw material for paper making. Indian<br />

Pulp Pap 33: 17–19.<br />

Blake O, Booth R, Carrigon D. 1993. The tanin content of herbal teas. Br J Phytotherapy 3: 124–127.


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 97<br />

Blanco MM, Costa C, Freire AO, Santos Jr. JG, Costa M. 2007. Neurobehavioral effect of essential oil of<br />

Cymbopogon citratus in mice. Phytomedicine 16: 265–270.<br />

Boelens MH. 1994. Sensory and chemical evaluation of tropical grass oils. Perfum Flavor 19: 29–45.<br />

Bor NL. 1954. The genus Cymbopogon Spreng. In India, Burma and Ceylon Part II. J Bombay Nat Hist Soc<br />

52: 149–183.<br />

Borovik VN, Kuravskaya IM. 1977. Flammability and explosiveness of some synthetic perfumes and intermediate<br />

products. Maslo-Zhir Prom-st 10: 34, 35 (C.A.89: 48794-b).<br />

Boti JB, Muselli A, Tomi F, Koukoua G, N’Guessan TY, Costa J, Casanova J. 2006. Combined analysis of<br />

Cymbopogon giganteus Chiov. leaf oil from Ivory Coast by GC/RI, GC/MS and 13 C-NMR C.R. Chimie<br />

9: 164–168.<br />

Bottini AT, Dev V, Garfagnoli DJ, David J, Hope H, Joshi P, Lohani H, Mathela CS, Nelson TB. 1987.<br />

Isolation and crystal structure of a novel dihemiacetal bis monoterpenoid from Cymbopogon martinii.<br />

Phytochemistry 26: 2301, 2302.<br />

Bouvier F, d’Harlingue A, Suire C, Backhaus RA, Camara B. 1998. Dedicated roles of plastid transketolases<br />

during the early onset of isoprenoid biogenesis in pepper fruits. Plant Physiol 117: 1423–1431.<br />

Brazil S, Gilberto A de A, Bauer L. 1971. <strong>Essential</strong> oil of Cymbopogon citratus. Rio Grande do Sul 52:<br />

193–196 (C.A., 76, l3l371-p).<br />

Bruns K, Heinrich B, Pagel I. 1981. Citronellol. Untersuchung von Handels und Hybridolen Verschiedener<br />

Provenienz in Vorkommen und Analytic Atherischer Ole, Band 2 (Kubeczkaeds KH, Ed.), G. Thieme<br />

Verlag, Stuttgart, Germany.<br />

Carlini EA, Contar JDP, Silva-Filho AR, Da Silveira-Filho NG, Frochtengarten ML, Bueno OFA. 1986.<br />

Pharmacology of lemongrass (Cymbopogon citratus Stapf.). I. Effects of teas prepared from the leaves<br />

on laboratory animals. J Ethnopharmacol 17(1): 37–64.<br />

Chakrabarti MM, Ghosh SK. 1974. <strong>Essential</strong> oils from plant grown in West Bengal. Indian Perfumer 18:<br />

35–39.<br />

Chalchat JC, Garry RP, Menut C, Lonaly G, Malhuret R, Chopineau J. 1997. Correlation between chemical<br />

composition and antimicrobial activity. VI. Activity of some African essential oils. J Essen <strong>Oil</strong> Res 9:<br />

67–75.<br />

Chauhan YS, Singh KK, Ganguly D. 1976. Improvement of Java citronella (C. winterianus Jowitt) by chemical<br />

mutagenesis. Indian Perfumer 20: 73–77.<br />

Cherian S, Chattattu GJ, Viswanathan TV. 1993. Chemical composition of lemongrass varieties. Indian<br />

Perfumer 37: 77–80.<br />

Chiang HC, Ma YC, Huang KF. 1981. Quantitative analysis of citronella oil and lemongrass oil by hydrogen<br />

NMR spectrometry. Taiwan Yao Huesch Ta Chin 33: 95–103 (C.A. 97, 11630-j).<br />

Chicopharma NV. 1970. Pat. Neth. No. 6909123 (C.A. 7479503-a).<br />

Chisowa BH. 1997. Chemical composition of flower and leaf oils of Cymbopogon densiflorus Stapf from<br />

Zambia. J Essen <strong>Oil</strong> Res 9: 469, 470.<br />

Chopra RN, Nayar SL, Chopra IC. 1956. Glossary of Indian Medicinal Plants. CSIR, New Delhi, p. 87.<br />

Choudhary DK, Kaul BL. 1979. Radiation-induced methyl eugenol deficient mutant of Cymbopogon flexuosus<br />

(Nees ex Steud.) Wats. Proc Indian Acad Sci Sect 88 B: 225–228.<br />

Choudhary SN, Leclercq PA. 1995. <strong>Essential</strong> oil of Cymbopogon khasianus (Munro ex Hack.) Bor from north<br />

eastern India. J Essen <strong>Oil</strong> Res 7: 555, 556.<br />

Ciaramello D, Azzini A, Pinto AJD, Guilherme M, Donalisio R. 1972. Use of citronella, lemongrass, palmarosa<br />

and vetiver leaves for the production of cellulose and paper. An Acad Bras Cienc 44(Suppl.): 430–441.<br />

(C.A. 83, 8 1640-x).<br />

Crawford M, Hanson AW, Koker MES. 1975. Structure of Cymbopogon, a novel triterpenoid from lemongrass.<br />

Tetrahedron Lett 35: 3099–3102.<br />

Cronquits A. 1980. Chemistry in plant taxonomy, In Chemosystematics Principles and Practice. (Bisby FA,<br />

Vaughan JC, and Wright CA, Eds), Academic Press, London. pp. 1–27.<br />

Croteau R, Felton M, Karp F, Kjonaas R. 1981. Relationship of camphor biosynthesis to leaf development in<br />

sage (Salvia officinalis). Plant Physiol 67(4): 820–824.<br />

Croteau R, Davis EM. 2005. (−)-Menthol biosynthesis and molecular genetics. Naturwissenschaften 92:<br />

562–567.<br />

Crowell PL. 1999. Prevention and therapy of cancer by dietary monoterpenes. J Nutr 129: 775–778.<br />

Dawidar AM, Esmiraly ST, Abdel MM. 1990. Sequiterpenes from Cymbopogon schoenanthus. Pharmazie 45:<br />

296, 297.


98 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

De Martinez MV. 1977. Adaptation of the determination of citral with hydroxyl aminehydrochloride to lemongrass<br />

oil. Arch Bioquim Quim Farm 20: 51–56 (C.A., 91,181225-j).<br />

De Sylva MG. 1959. Lemongrass oil from Ceylon. Mfg Chemist 30: 415, 416.<br />

Dev Kumar C, Narayana MR, Khan MNA. 1977. Synthetic products from oil of citronella. Indian Perfumer<br />

22: 139–145.<br />

Dev V, Melander D, Yee JT. 1988. Terpenoid composition of the essential oil from Cymbopogon jwarancusa.<br />

Dev Food Sci Flav Frag 18: 317–321.<br />

Dhar AK, Lattoo S. 1985. <strong>Essential</strong> oil content in Cymbopogon jwarancusa after cutting. Pafai J 7: 28–31.<br />

Dhar AK, Thapa RK, Atal CK. 1981. Variability in yield and composition of essential oil in Cymbopogon<br />

jwarancusa. Planta Med 41: 386–388.<br />

Dhar RS, Dhar AK, 1997. Ontogenetic variation in essential oil concentration and its constituents in five genotypes<br />

of Cymbopogon jwarancusa (Jones) Schult. J Essen <strong>Oil</strong> Res 9: 433–439.<br />

Dikshit A, Hussain A. 1984. Anti-fungal action of some essential oils against animal pathogens. Fitoterapia<br />

55: 171–176.<br />

Duarte MCT, Leme EE, Delarmelina C, Soares AA , Figueira GM, Sartoratto A. 2006. Activity of essential oils<br />

from Brazilian medicinal plants on Escherichia coli. J Ethnopharmacol 111(2): 197–201.<br />

Dubey NK, Kishore N, Varma J, Lee SY 1997. Cytotoxicity of the essential oil of Cymbopogon citratus and<br />

Ocimum gratissimum. Indian J Pharm Sci 59: 263–264.<br />

Dubey NK, Takeya K, Itakawa H. 1997. Citral: A cytotoxic principle isolated from the essential oil of<br />

Cymbopogon citratus against 1388 leukaemia cells. Curr Sci 73: 22–24.<br />

Dunyan Xue, Maosen Song, Ning Chen, Yaozu Chen. 1992. Chemical ingredients of the essential oil of<br />

Cymbopogon distans. Gaoden Xoxiao Huaxue X. Xuebao 13: 1551, 1552.<br />

Eisenreich W, Rohdich F, Bacher A. 2001. Deoxyxylulose phosphate pathway to terpenoids. Trends Plant Sci<br />

6: 78–84.<br />

Eisenreich W, Sagner S, Zenk MH, Bacher A. 1997. Monoterpenoid essential oils are not of mevalonoid origin.<br />

Tetrahedron Lett 38: 3889–3892.<br />

Eisenreich W, Schwarz M, Cartayrade A, Arigoni D, Zenk MH, Bacher A. 1998. The deoxyxylulose phosphate<br />

pathway of terpenoid biosynthesis in plants and microorganisms. Chem Biol 5: R221–R223.<br />

Elagamal MH, Wolff P. 1987. A further contribution to the sesquiterpenoid constituents of Cymbopogon proximus.<br />

Planta Med 53: 293, 294.<br />

EI Tawil BAH, EI Beih FK. 1982. Study of volatile oil of Saudi Cymbopogon citratus D.C. and Cymbopogon<br />

schoenanthus (L). J Chin Chem Soc (Taipei) 30: 281, 282 (C.A. 100,39433).<br />

Eskandar EF, Won Jun H. 1995. Hypoglycaemic and hyperinsulinenic effect of some Egyptian herbs used for<br />

the treatment of diabetes mellitus (type II) in rats. Egypt J Pharm Sci 36: 334–341.<br />

Esmort H, David RH, Dudley IF. 1998. Volatile constituents of the essential oil of Cymbopogon citratus Stapf.<br />

grown in Zambia. Flav Frag J 13(1): 29, 30.<br />

Evans FE, Miller DW, Cairus T, Baddeley GV, Wen KB. 1982. 13 C NMR of naturally occurring substances,<br />

Part 74. Structure analysis of proximadiol (Cryptomeridiol) by 13 C NMR spectroscopy. Phytochemistry<br />

21: 855–858.<br />

Figueirinha A, Paranhos A, Pérez-Alonso JJ, Santos-Buelga C, Batista MT. 2008. Cymbopogon citratus leaves:<br />

Characterization of flavonoids by HPLC-PDA-ESI/MS/MS and an approach to their potential as a source<br />

of bioactive polyphenols. Food Chem 110(3): 718–728.<br />

Foda YH, Abdullah MA, Zaki MS, Mostafa MM. 1975. Identification of the volatile constituents of the Egyptian<br />

lemongrass oil. II. IR Spectroscopy. Nahrung 19: 395–400 (C.A.83, 1683 17-w).<br />

Formacek K, Kubeczka KH. 1982. <strong>Essential</strong> <strong>Oil</strong>s Analysis by Capillary Chromatography and 13 C NMR<br />

Spectroscopy. John Wiley & Sons, New York.<br />

Fuji T, Furukawa S, Suzuki S. 1972. Compounded perfumes for toilet goods. Non-irritative compounded perfumes<br />

for soaps. Yukagaku 21: 904–908 (C.A. 78, 75786-e).<br />

Gagrade SK, Sri Vastava RD, Sharma OP, Maghe MN, Trivedi KC. 1990. Evaluation of some essential oil for<br />

antibacterial properties. Indian Perfumer 34: 204–208.<br />

Ganguly P, Singh KK, Bhagat SD, Upadhyay DN, Chauhan YS, Gupta NK, Singh HS. 1979. “RRLJOR-3-1970”<br />

an improved strain of Java citronella (Cymbopogon martinii). Phytochemistry 26: 183–185.<br />

Gaydou BM, Raudriamiharisoa RP. 1987. Hydrocarbon from the essential oil of Cymbopogon winterianus<br />

Jowitt. Indian Perfumer 23: 107–111.<br />

Gaydou BM, Raudriamiharisoa RP. 1987. Composition of palmarosa (cymbopogon martinii) essential from<br />

Madagascar. J Agric Food Chem 35(1): 62–66.


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 99<br />

Gaydou EM, Raudriamiharisoa RP. 1986. Hydrocarbons from the essential oil of Cymbopogon martinii.<br />

Phytochemistry 26(1): 183–185.<br />

Gildemeister B, Hoffmann F. 1956. Die Artherischen Ole, Vol. IV. Akademie Verlag, Berlin, pp. 307–419.<br />

Gonzalo VF, Villarrubia MM. 1974. <strong>Essential</strong> oil of lemongrass in Tucuman, Int Congr Essen <strong>Oil</strong>s, Allured<br />

Publishing, Oak Park, Ilinois. p. 110 (C.A.,84,49730-m).<br />

Gonzalo VF, Villarrubia MM. 1973. <strong>Essential</strong> oil of lemongrass in Tucuman. Arch Bioquim Quim Farm 18:<br />

51–57.<br />

Grace O, Yisak W, Ogunlana EO. 1984. Antibacterial constituents in the essential of cymbopogon citratus.<br />

(D.C) Sapf. Journal of Ethnopharmacology 12: 279–286.<br />

Guenther E. 1948. The <strong>Essential</strong> <strong>Oil</strong>s. Vol. 1, Van Nostrand, London. pp. 232–236.<br />

Guenther E. 1950. The <strong>Essential</strong> <strong>Oil</strong>s. Vol. IV, Van Nostrand, London. pp. 3–153.<br />

Guillaume KK, Honore KK, Isabella AG, Jacques H. 2006. Comparative effects of cymbopogon shoenanthus<br />

essential oil and piperitone on collasabuchus maculatur development. Fitoterapia 77(7–8): 506–510.<br />

Gulati BC, Sadgopal 1972. Chemical examination of oil of Cymbopogon nardus. Indian <strong>Oil</strong> Soap J 37:<br />

305–310.<br />

Gulati BC, Garg SN. 1976. <strong>Essential</strong> oil of Cymbopogon pendulus Wats. under the Tarai climate of Uttar<br />

Pradesh. Indian J Pharm 38: 78, 79.<br />

Gupta BK, Daniel P. 1982. Aromatic grasses of India and their utilization—a plea for further research. Pafai J<br />

Jan–March, 13–27.<br />

Gupta BK, Jam N. 1978. Cultivation and utilization of genus Cymbopogon in India. Indian Perfumer 21: 55–68.<br />

Gupta BK. 1969. Studies in the genus Cymbopogons. Perfum Essen <strong>Oil</strong> Res 70B: 241–247.<br />

Gupta PC, Singh R, Singh K, Pradhan K. 1975. Chemical composition and in vitro matter digestibility of some<br />

important grasses. Ann Arid Zone 14: 245–250 (C.A. 84, 178699-g).<br />

Gupta YN, Chauhan PNS. 1970. Effect of ultraviolet radiation on essential oils. Indian Perfumer 14: 63–66.<br />

Gyane DD. 1976. Preservation of shea butter. Drug Cosmet Ind 18: 36–38 and 138–140.<br />

Handique AK, Singh HB. 1990. Antifungal action of lemongrass oil on some soil-borne plant pathogens. Indian<br />

Perfumer 34: 232–234.<br />

Han IK, Kim KL, Park SH, Kim BH, Ahn BH. 1971. Nutritive values of the native grasses and legumes in Korea<br />

V. Chemical determination of the mineral content of native herbage plants. Han‘guk Chuksanak’hoe Chi<br />

13: 329–334 (C.A.81, 74938-f).<br />

Hanson SW, Crawford M, Koker MES, Menezes FA. 1976. Cymbopogonol: A new triterpenoid from C. citratus.<br />

Phytochemistry 15: 1074, 1075.<br />

Herath HMW, Iruthayathas EE, Ormrod DP. 1979. Temperature effects on essential oil composition of citronella<br />

selection. Econ Bot 33: 425–430.<br />

Herz S, Wungsintaweekul J, Schuhr CA, Hecht S, Lüttgen H, Sagner S, Fellermeier M, Eisenreich W, Zenk<br />

MH, Bacher A, Rohdich F. 2000. Biosynthesis of terpenoids: YgbB protein converts 4-diphosphocytidyl-<br />

2C-methyl-d-erythritol 2-phosphate to 2C-methyl-d-erythritol 2,4-cyclodiphosphate. Proc Natl Acad Sci<br />

USA 97: 2486–2490.<br />

Idrissi A, Bellakhdar J, Canigueral S, Iglesian J, Vila R. 1993 Composition of the essential oil of lemongrass<br />

(C. citratus D.C.) Stapf. cultivated in Morocco. Plantes Medicinales et Phytotherapie 24: 274–277.<br />

Ille K. 1986. Gas chromatographic analysis of some essential oils used in perfumery. Mezhdunar 4th Kongar<br />

Efirnym Maslam. Naumenko PV (Ed.), Moscow. pp. 112–123.<br />

Iruthayathas EE, Herath HMW, Wijesekara ROB, Jayawardene AL. 1997. Variations in the composition of oil<br />

in citronella. J Nat Sci Coun Sri Lanka 5: 133–146.<br />

Jirovetz L, Buchbauer G, Eller G, Ngassoum, MB, Maponmetsem PM. 2007. Composition and antimicrobial<br />

activity of Cymbopogon giganteus (Hochst.) Chiov. <strong>Essential</strong> flower, leaf and stem oils from Cameroon.<br />

J Essen <strong>Oil</strong> Res 19: 485–489.<br />

Jyrkit T. 1983. Composition of the essential oil of C. flexuosus. J Chrom 262: 364–366.<br />

Kalia NK, Sood RP, Padha CD, Jamwal RK, Chopra MM. 1980. Utilization of wild growing Cymbopogon<br />

martinii (var. sofia). Indian Perfumer 24: 182–184.<br />

Kamali HHEL, Hamza MA, Amir MYEL. 2005. Antibacterial activity of the essential oil from Cymbopogon<br />

nervatus inflorescence. Fitoterapia 76(5): 446–449.<br />

Kanjilal PB, Pathak MG, Singh RS, Ghosh AC. 1995. Volatile constituents of the essential oil of Cymbopogon<br />

caesius (Nees ex Hook. et Am) Stapf. J Essen <strong>Oil</strong> Res 7: 437–449.<br />

Kaul BL, Choudhary DK, Atal CK. 1977. Introduction and chemical evaluation of Burmese citronella in<br />

Jammu. Indian J Pharmacol 39: 42, 43.


100 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Ketoh GK, Koumaglo HK, Glitho IA, Huignard J. 2006. Comparative effects of Cymbopogon schoenanthus<br />

essential oil and piperitone on Callosobruchus maculatus development. Fitoterapia 77(7–8): 506–510.<br />

Khadri A, Serralheiro MLM, Nogueira JMF, Neffati M, Smiti S, Araújo MEM. 2008. Antioxidant and<br />

antiacetylcholinesterase activities of essential oils from Cymbopogon schoenanthus L. Spreng.<br />

Determination of chemical composition by GC-mass spectrometry and 13 C NMR. Food Chem<br />

109(3): 630–637.<br />

Kichiyoshi Koryo KK, Akira Sangy KK. 1981. An unpleasant odour composition for prevention of thinner<br />

sniffing. Jpn Tokyo Koho JP Pat 48538 (C.A., 96, 117345-q).<br />

Kisaki A, Matsuyama Y, Sakano T, Fujiwara N, Nakaguchi O. 1998. Antimale sex hormone agent and composition.<br />

Jpn Kokai Tokyo JP 17: 486.<br />

Kokate CK, Verma KC. 1971. Antimicrobial activity of volatile oils of Cymbopogon nardus and Cymbopogon<br />

citratus. Sci Cult 37: 196–198.<br />

Koketsu M, Moura LL, Magalhaes MT. 1976. <strong>Essential</strong> oil of Cymbopogon densiflorus Stapf. and Tagetes<br />

minuta L. grown in Brazil. An Acad Bras Cienc 48: 743–746 (C.A.88, 141480-k).<br />

Kreis P, Masandl A. 1994. Chiral compounds of essential oil. Part XVII. Simultaneous stereoanalysis of<br />

Cymbopogon oil constituents. Flav Frag J 9: 257–260.<br />

Krishnamoorthy G, Kavimani S, Loganthan C. 1998. Antiinflammatory activity of the essential oil of<br />

Cymbopogon martinii. Indian J Pharm Sci 60: 114–116.<br />

Krishnarajah SR, Ganesalingam VK, Senanayake UM. 1985. Repellancy and toxicity of some plant oils and<br />

their terpene components to Stotroga cereallelia (Olivier). Trop Sci 25: 249–252.<br />

Kulkarni RN, Mallavarapu GR, Bhaskaran K, Ramesh S, Kumar S. 1997. <strong>Essential</strong> oil composition of citronella-like<br />

variant of lemongrass. J Essen <strong>Oil</strong> Res 9: 393–395.<br />

Kumar A, Malik F, Bhushan S, Sethi VK, Shahi AK, Kaur J, Taneja SC, Ghulam N, Qazi GN, Singh J. 2007.<br />

An essential oil and its major constituent isointermedeol induce apoptosis by increased expression of<br />

mitochondrial cytochrome c and apical death receptors in human leukaemia HL-60 cells. Chem-Biol<br />

Interact 171(3): 332–347.<br />

Kuriakose KP. 1995. Genetic variability in East Indian lemongrass (C. flexuosus Stapf.). Indian Perfumer 39:<br />

76–83.<br />

Kurian A, Nair BVG, Rajan KC. 1984. Effect of antioxidants on the preservation of citral content of lemongrass<br />

oil. Indian Perfumer 28: 28–32.<br />

Kusumov FY, Babaev RI. 1983. Components of the essential oil of Cymbopogon citratus. Stapf Khim Prir<br />

Soedin 1: 108–109 (C.A., 98 204188-a).<br />

Kuzuyama T, Takagi M, Kaneda K, Dairi T, Seto H. 2000. Formation of 4-(cytidine-5′-diphospho)-2-C-methyld-erythritol<br />

from 2-C-methyl-d-erythritol 4-phosphate by 2-C-methyl-d-erythritol 4-phosphate cytidylyltransferase,<br />

a new enzyme in the non-mevalonate pathway. Tetrahedron Lett 41: 703–706.<br />

Kuzuyama T, Takagi M, Kaneda K, Watanabe H, Dairi T, Seto H. 2000 Studies on the nonmevalonate pathway:<br />

conversion of 4-(cytidine-5′-diphospho)-2-C-methyl-d-erythritol to its 2-phospho derivative by<br />

4-(cytidine-5′-diphospho)-2-C-methyl-d-erythritol kinase Tetrahedron Lett 41: 2925–2928.<br />

Lange BM, Croteau R. 1999. Isoprenoid biosynthesis via a mevalonate-independent pathway in plants: cloning<br />

and heterologous expression of 1-deoxy-d-xylulose-5-phosphate reductoisomerase from peppermint.<br />

Arch Biochem Biophys 365: 170–174.<br />

Lange BM, Wildung MR, McCaskill DG, Croteau R. 1998 A family of transketolases that directs isoprenoid<br />

biosynthesis via a mevalonate-independent pathway. Proc Natl Acad Sci USA 95: 2100–2104.<br />

Lawrence BM. 1991. Citronella oil: A monograph. In Lawrence Review of Natural Products, p. I.<br />

Le KT, Chu PNS. 1976. Chemical composition of lemongrass essence (Cymbopogon flexuosus Stapf.) from<br />

Vietnam. Tap Clii Hoc 14: 31–36.<br />

Lemos TLG, Monte EJO, Matos FJA, Alencer JW, Craveiro AA, Barobasa RCSB, Lima BO. 1994. Chemical<br />

composition and antimicrobial activity of essential oils from Brazilian plants. Fitoterapia 63: 226–228.<br />

Lichtenthaler HK. 1999. Annu Rev Plant Physiol Plant Mol Biol 50: 47–66.<br />

Little DB, Croteau R. 1999. Biochemistry of essential oil terpenes: A thirty year overview. In Flavour Chemistry:<br />

Thirty Years of Progress (Hornstein I et al., Eds). Kluwer Academic/Plenum Press, New York. pp.<br />

239–253.<br />

Liu C, Zhang J, Yiao R, Gan L. 1981. Chemical studies on ten essential oils of Cymbopogon genus. Huaxue<br />

Xuebao 241–247 (C.A.98, 104358-in).<br />

Liu M, Feng G. 1989. Effect of Cymbopogon goeringii (Steud.) A. Camus volatile oil on physiological properties<br />

of isolated guinea pig myocardium. Zlionguo Zliongyao Zazlii 14: 160–622.


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 101<br />

Lohani H, Bisht JC, Melkani AB, Mathela DK, Mathela CS, Dev V. 1986. Chemical composition of essential<br />

oil of Cymbopogon stracheyi (Hook f.) Riaz and Jani. Indian Perfumer 30: 447–452.<br />

Lorenzetti BB, Souza GEP, Sarti SJ, Santos ED, Frreira SH. 1991. Myrcene mimics the peripheral analgesic<br />

activity of lemongrass tea. J Ethnopharmacol 34: 43–48.<br />

Lorenzo D, Dellacassa E, Atti-Serafini L, Santos AC, Frizzo C, Paroul N, Moyna P, Mondello L, Dugo G. 2000.<br />

Composition and stereoanalysis of Cymbopogon winterianus Jowitt. oil from Southern Brazil. Flav<br />

Frag J 15(3): 177–181.<br />

Lucius G, Adler B. 1971. Mezhdunar Kangr. Effirnym Maslam (Mater) 4th, 1: 211–214. Naumenko PV, Ed.,<br />

“Pishcheuaya, Promyshlennot,” Moscow, USSR (C.A. 78,15 1557-y).<br />

Lüttgen H, Rohdich F, Herz S, Wungsintaweekul J, Hecht S, Schuhr CA, Fellermeier M, Sagner S, Zenk MH,<br />

Bacher A, Eisereich W. 2000. Biosynthesis of terpenoids: YchB protein of Escherichia coli phosphorylates<br />

the 2-hydroxy group of 4-diphosphocytidyl-2C-methyl-d-erythritol. Proc Natl Acad Sci USA 97:<br />

1062–1067.<br />

Maheshwari ML, Chandel KPS, Chien MJ. 1988. The composition of essential oil from Hyssopus officinalis<br />

and Cymbopogon jwarancusa (Jones) Schult. collected in the cold desert of Himalaya. Dev Food Sci Flav<br />

Frag 18: 171–176.<br />

Maheshwari ML, Mohan J. 1985. Geranyl formate and other esters in palmarosa oil. Pafai J 7: 21–26.<br />

Mallavarapu GR, Kulkami RN, Ramesh S. 1992a. Composition of the essential oil of Cymbopogon coloratus.<br />

Planta Med 58: 479, 480.<br />

Mallavarapu GR, Ramesh S, Kulkami RN, Shyamsunder KV. 1992b. Composition of essential oil of Cymbopogon<br />

travencorensis. Planta Med 58: 219–222.<br />

Mallavarapu GR, Rajeshwara Rao, Kaul PN, Ramesh S, Bhattacharya AK. 1998. Volatile constituents of the<br />

essential oil of the seeds and the herb of palmarosa (C.martinii (Roxb.) Wats. var. motia Burk. Flav<br />

Frag J 13: 161–169.<br />

Manjoor-i-Khuda M, Rahman M, Yusuf M, Choudhary J. 1984. <strong>Essential</strong> oils of Cymbopogon species of<br />

Bangladesh. J Bangladesh Acad Sci 8: 77–80.<br />

Manjoor-i-Khuda M, Rahman M, Yusuf M, Choudhary J. 1986. Studies on essential oil bearing plants of<br />

Bangladesh. Part II. Five species of Cymbopogon from Bangladesh and the chemical constituents of their<br />

essential oils. Bangladesh J Sci Ind Res 21: 70–82.<br />

Mathela CS, Joshi P. 1981. Terpenes from the essential oil of Cymbopogon distans. Phytochemistry 20: 2770,<br />

2771.<br />

Mathela CS, Melkani AB, Pant AK, Pande C. 1988a. Chemical variations in Cymbopogon distans and their<br />

chemosystematic implications. Biochem Syst Ecol 16:161–165.<br />

Mathela CS, Chittattu GI, Thomas J. 1996. OD-468. A Lemongrass chemotype rich in geranyl acetate. Indian<br />

Perfumer 40: 9–12.<br />

Mathela CS, Lohani H, Pande C, Mathela DK. 1988. Chemosystematics of terpenoids in Cymbopogon martini.<br />

Biochem Syst Ecol 16: 167–169.<br />

Mathela CS, Melkani AB, Pant AK, Vasu Dev, Nelson TE, Hope H, Bottini AT. 1989. Eudesmanediol from<br />

Cymbopogon distans. Phytochemistry 28: 1936–1938.<br />

Mathela CS, Pande C, Pant AK, Singh AK. 1990a. Terpenoid composition of Cymbopogon distans f. munsiyariensis.<br />

Herb Hung 27: 117–121.<br />

Mathela CS, Pande C, Pant AK, Singh AK. 1990b. Phenylpropanoid constituents of Cymbopogon micro stachys.<br />

J Indian Chem Soc 67: 526–528.<br />

Mathela CS, Pant AK, Melkani AB, Pant A. 1986. Aromatic grasses of U.P Himalaya: A new wild species as a<br />

source of aroma chemicals. Sci Cult 52: 342–344.<br />

Mathela CS, Pant AK. 1988. Production of essential oils from some new Himalayan Cymbopogon species.<br />

Indian Perfumer 32: 40–50.<br />

Mathela CS, Sinha GK. 1978. Study of the essential oil of C. nardus var. stracheyi. J Indian Chem Soc 55:<br />

621, 622.<br />

Mathela CS. 1991. Flavour constituents of Himalayan aromatic grasses. International Symposium on Newer<br />

Trends in <strong>Essential</strong> <strong>Oil</strong>s and Flavours, RRL Jammu, India, p. 27.<br />

Mehmood Z, Ahmad S, Mohammed F. 1997. Antifungal activity of some essential oils and their major constituents.<br />

Indian J Nat Products 13: 10–13.<br />

Melkani AB, Joshi P, Pant PK, Mathela CS, Dev V. 1985. Constituents of the essential oil from two varieties of<br />

Cymbopogon distans. J Nat Prod 46: 1995–1997.<br />

Menon TCK. 1956. The essential oil of Cymbopogon travancorensis. J Bombay Nat Hist Soc 53: 742, 743.


102 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Modawi BM, Magar HRY, Satti AM, Duprey RJH. 1984. Chemistry of Sudanese flora. Part III. Cymbopogon<br />

nervatus. J Nat Prod 47: 167–169.<br />

Mohammad F, Nigam MC, Rehman W. 1981a. Detection of new trace constituents in the essential oils of<br />

Cymbopogon flexuosus. Pafai 13: 22–24.<br />

Mohammad F, Nigam MC, Rehman W. 1981b. <strong>Essential</strong> oil of Cymbopogon martinii var. motia: detection of<br />

new trace constituents. Pafai J 3: 11–13.<br />

Moody JO, Adeleye SA, Gundidza MG, Wyllie G, Ajayi-Obe OO. 1995. Analysis of the essential oil of<br />

Cymbopogon nardus (L.) Rendle growing in Zimbabwe. Pharmazie 50(1): 74, 75.<br />

Moris JA, Khettry A, Seitz EW. 1979. Antimicrobial activity of aroma chemicals and essential oils. J Am <strong>Oil</strong><br />

Chem Soc 56: 595–603.<br />

Muthuswami S, Sayed S. 1980. Preliminary note on the extraction of lemongrass oil. Indian Perfumer 24: 226,<br />

227.<br />

Naik SN, Kumar A, Maheshwari RC, Kumar B, Chandra R, Guddewar MB. 1995. Evaluation of toxicity and<br />

growth-inhibiting activity of Tripolium castaneum (Herbst). Indian Perfumer 39: 1–8.<br />

Nair BVG, Chinnamma NP, Pushpakumari R. 1980a. Influence of varieties on the grass, oil yield and quality<br />

of oil of palmarosa (C. martinii Stapf. var. motia). Indian Perfumer 24: 22–24.<br />

Nair BVG, Chinnamma NP, Pushpakumari R. 1980b. Investigations on some types of lemongrass (C. flexuosus<br />

Stapf.) Indian Perfumer 24: 20, 21.<br />

Nair EVG, Mariam KA. 1978. The new promising aromatic plant of Kerala. Indian Perfumer 22: 300, 301.<br />

Nair RV, Siddiqui MS, Sen T, Nigam MC. 1982. Identification of three new species of Cymbopogon in perfumery<br />

value and their G.C. evaluation. Pafai J 4: 21–23.<br />

Namba T, Hatsutori Y, Shimomoura K, Nakamura M. 1995. Extraction of hyalurodinase inhibitors from<br />

Azadirachta indica or other plants for manufacturing cosmetics or for therapeutic use. Japan Kokai Tokai<br />

Tokyo Koho JP 07, 138,180 (95,138,180).<br />

Nath SC, Saha BN, Bordoloi DN, Mathur RK, Leclercq PA. 1994. The chemical composition of the essential<br />

oil of Cymbopogon flexuosus Steud. Wats. growing in northeast India. J Essen <strong>Oil</strong> Res 6: 85–87.<br />

Naves YR. 1970. Palmarosa oil compounds. Parfum Cosmet Savons 13: 354–357 (C.A.73, 59212–6).<br />

Naves YR. 1971. Components of palmarosa essential oil. In Meijhdunar Kongr. Efirnym Maslam (Mater),<br />

4th, 1: 239–244. Naumenko PV, Ed., “Pishchevaya Promyshlennost”, Moscow, USSR (C.A. 78, 16393<br />

1-d).<br />

Newman JD, Chappell J. 1999 Isoprenoid biosynthesis in plants: carbon partitioning within the cytoplasmic<br />

pathway. Crit Rev Biochem Mol Biol 34: 95–100.<br />

Neyberg AG. 1953. Some essential oil from the east of the colony. Bull Agr Conogo 44: 1–38 and 319–364<br />

(C.A. 48, 957-d).<br />

Ng TT. 1972. Growth performance and production potential of some aromatic grasses in Sarawak, Preliminary<br />

assessment. Trop Sci. 14: 47–58.<br />

Nigam MC, Datta SC, Bhattacharya AK, Duhan SPS. 1975. Utilization of citral rich oils for production of<br />

geraniol. Indian Perfumer 18: 55, 56.<br />

Nigam MC, Nigam IC, Levi L. 1965. <strong>Essential</strong> oil and their constituents XXVII: Composition of gingergrass.<br />

Can J Chem 43: 521–525.<br />

Nigam MC, Srivastava HK, Siddiqui MS. 1987. Chemistry of Cymbopogon and their essential oils. Pafai 19:<br />

13–20.<br />

Olaniyi AA, Sofowora EA, Oguntimehin BO. 1975. Phytochemical investigation of some Nigerian plants used<br />

against fevers II. Cymbopogon citratus. Planta Med 28: 186–189.<br />

Oliveros-Belardo L, Aureus B. 1978. <strong>Essential</strong> oil from C. citratus (D.C.) Stapf. growing in the Philippines.<br />

Asian J Pharm 3: 14–17 (C.A.90, 15698 1-q).<br />

Oliveros-Belardo L, Aureus E. 1979. <strong>Essential</strong> oil from Cymbopogon citrates (D.C.) Stapf. growing wild in the<br />

Phillippines. Int Congr <strong>Essential</strong> <strong>Oil</strong>s. 7: 166–168 (C.A. 92, 37768-g).<br />

Oliveros-Belardo L, Smith RM, Gupta ZB, Abadiano MJP. 1989. Leaf essential oil of wild Cymbopogon martinii<br />

(Roxb.) Wats. from Sorsogon, Philippines. Philippine J Sci 118: 31–58.<br />

Onawunmi GO, Yisak W, Ogunlana EO. 1984. Antibacterial constituents in the essential oil of Cymbopogon<br />

citratus (DC.) Stapf. J Ethnopharmacol 12(3): 279–286.<br />

Opdyke DLJ. 1973. Monographs on fragrance raw materials: Caryophyllene. Food Cosmet Toxicol 11:<br />

1059–1060.<br />

Opdyke DLJ. 1974. Monographs on fragrance raw materials: Palmarosa oil. Food Cosmet Toxicol 12 (Suppl.):<br />

947.<br />

Opdyke DLJ. 1976. Inhibition of sensitization reactions induced by certain aldehydes. Food Cosmet Toxicol<br />

14: 197, 198.


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 103<br />

Opdyke DLJ. 1976. Monographs on fragrance raw materials: Lemongrass oil: East Indian. Food Cosmet Toxicol<br />

14: 455.<br />

Opdyke DLJ. 1976. Monographs on fragrance raw materials: Lemongrass oil: West Indian. Food Cosmet<br />

Toxicol 14: 457.<br />

Orafidiya LO. 1993. The effect of autoxidation of lemongrass oil on its antibacterial activity. Phytotherapy Res<br />

7: 269–271.<br />

Osmani S, Anees I, Naidu MB. 1972. Insect repellent creams from essential oils. Pesticides 6: 19–21.<br />

Oyedele AO, Gbolade AA, Sosan MB, Adewoyin FB, Soyelu OL, Orafidiya OO. 2002. Formulation of an effective<br />

mosquito-repellent topical product from lemongrass oil. Phytomedicine 9(3): 259–262.<br />

Pande C, Melkani AB, Mathela CS. 1997. Cymbopogon distans: report on new chemotype. Indian Perfumer<br />

41: 35, 36.<br />

Pandey MC, Sharma lR, Dikshit A. 1996. Antifungal evaluation of the essential oil of Cymbogon pendulus<br />

(Nees ex Steud) Flav Frag J 11: 257–260..<br />

Pandey TC, Gupta YN. 1973. Effect of γ-radiations on citronella oils (Ceylon type). Indian Perfumer 17: 18, 19.<br />

Pant AK, Mathela CS, Pande N, Pant AK. 1990. Phenyl propenoid constituents of Cymbopogon microstachys<br />

J Indian Chem Soc 67: 526, 527.<br />

Patra NK, Singh HP, Kalra A, Singh HB, Mengi N, Singh VR, Naqvi AA, Kumar S. 1997. Isolation and development<br />

of geraniol-rich cultivar of citronella (C. winterianus). J Med Arom Plant Sci 19: 672–676.<br />

Patra NK, Srivastava RK, Chauhan SP, Ahmad A, Misra LN. 1990. Chemical features and productivity of a<br />

geraniol-rich variety (GRL-1) of Cymbopogon flexuosus. Planta Med 56: 239, 240.<br />

Patra P, Dutta PK. 1986. Evaluation of improved lemongrass strains for herb and oil yield and citral content<br />

under Bhubaneswar condition. Res Ind 31: 358–360.<br />

Peyron L. 1972. Essence produced in Mata Grasse. An Acad Bras Cienc 44 (Suppl.): 332–340 (C.A. 83, 136686-z).<br />

Peyron L. 1973. <strong>Essential</strong> oil from Mato Grasso. Perftimes Cosmet Savons 3: 371–378.<br />

Pilley PP, Rao P, Sanjiva B, Simonson JL. 1928. Constituents of some Indian essential oils, Part XXII. The<br />

essential oil from flower heads of C. coloratus Stapf. J lndian lnst Sci IIA: 181–186.<br />

Pino IA, Rosada A, Correa MT. 1996. Chemical composition of the essential oil of Cymbopogon winterianus<br />

Jowitt from Cuba. J Essen <strong>Oil</strong> Res 8: 693, 694.<br />

Popielas L, Moulis C, Keita A, Fouraste I, Bessiere IM. 1991. The essential oil of Cymbopogon giganteus.<br />

Planta Med 57: 586, 587.<br />

Prasad S, Jamwal R. 1971. Insecticidal composition. Indian Pat. 122,504 (C.A. 77, 30347-a).<br />

Quintans-Júnior LJ, Souza TT, Leite BS, Lessa NMN, Bonjardim LR, Santos MRV, Alves PB, Blank AF,<br />

Antoniolli AR. 2008. Phythochemical screening and anticonvulsant activity of Cymbopogon winterianus<br />

Jowitt. (Poaceae) leaf essential oil in rodents. Phytomedicine 15(8): 619–624.<br />

Rabha LC, Baruah AKS, Bordoloi DN. 1979. Search for aroma chemicals of commercial value from plant<br />

resources of North East India. Indian Perfumer 23: 178.<br />

Rabha LC, Hagarika AK, Bordoli DN. 1986. Cymbopogon khasianus, a new rich source of methyl eugenol.<br />

Indian Perfumer 30: 339–344.<br />

Rabha LC, Hazarika AK, Bordoloi. 1989. A chemotype of Cymbopogon khasianus (Hack.) Stapf. ex Bor: An<br />

additional information for a source of higher oil and geraniol from North India. Indian Perfumer 33:<br />

261–265.<br />

Rajapakse RHS, Jayasena W. 1991. Plant resistance and biopesticides from lemongrass oil for suppressing<br />

Spodoptera litura in peanut. Agri Int 43: 166, 167.<br />

Rajendrudu G, Rama Das VS. 1983. Interspecific differences in the constituents of essential oil of Cymbopogon.<br />

Proc Indian Acad Sci 92: 331–334.<br />

Ramdan FM, EI-Zanfaly HT, Allian AM, EI-Wakeil FA. 1972a. Antibacterial effects of some essential oils I.<br />

Chem Microbiol Technol Lebensm 1: 96–102 (C.A. 77, 122532-k).<br />

Ramdan FM, El-Zanfaly HT, Allian AM, El-Wakeil FA. 1972b. Antibacterial effects of some essential oils II.<br />

Chem. Microbiol Technol Lebensm 2: 51–55 (C.A. 77. 122533-m).<br />

Ramirez AR, Bressani R, Elias LG. 1977. Utilization of lemongrass Cymbopogon species begasse in ruminant<br />

nutrition. Int. Symp Feed Compas Anim Nutr Requir Comput Diets 198–203 (C.A. 91, 90167-r).<br />

Ranaweera SS, Dayananda KR. 1997. Mosquito-larvicidal activity of Ceylon citronella (C. nardus L.) Rendle<br />

oil fractions. J Nat Sci SriLanka 24: 247–252.<br />

Rao BGV, Narsimha JPL. 1971. Activity of some essential oils towards phytopathogenic fungi. Riechst. Aromen<br />

Koerperflegen 21: 405, 406 and 408–410 (C.A. 76, 81621-x).<br />

Rao BL, Kala S, Dhar KL, Kaul BS. 1992. New aromatic chemicals in Cymbopogon for future. Indian Perfumer<br />

36: 241–245.


104 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Rao BN, Pandita S, Kaul BL. 1995. Gamma radiation as a mean of inducing variation in Cymbopogon flexuosus<br />

a chemotype for α-bisabolol Indian Perfumer 39: 84–87.<br />

Rao BS, Sudborough JJ. 1925. Kachi grass oil. J Indian Inst Sci 8A: 9–27.<br />

Rao JJ, Grover GS, Saxena VD. 1989. Mutagenic property of Cymbopogon martinii essential oil. Parfum<br />

Kosmet 70: 488.<br />

Rao JS. 1957. Grass flora of Coimbatore district, South India, J Bombay Nat Hist Soc 54: 674–690.<br />

Rauber CS, Guterres SS, Schapoval EES. 2005. LC determination of citral in Cymbopogon citratus volatile oil.<br />

J Pharm Biomed Anal 37(3): 597–601.<br />

Razdan TK, Koul GL. 1973. Chromatographic studies of Indian essential oils: <strong>Oil</strong> of Citronella Java type III.<br />

Indian J Chem 8: 27–32.<br />

Razdan TK.1984. Calorimetric analysis of the constituents of essential oils. Parfum Kosmet 65: 190–195.<br />

Rizk AM, Heiba HI, Mashaly M, Sandra P. 1985. Constituents of plants growing in Qatar. X . Seasonal variations<br />

of the volatile oil of Cymbopogon parkeri. Qatar Univ Sci Bull 5: 71–76 (C.A. 106, 2908h).<br />

Rizk AM, Heiba HI, Sandra P, Mashaly M, Bicchi C. 1983. Constituents of the volatile oil of Cymbopogon<br />

parkeri. J Chrom 279: 145–150.<br />

Rizk AM, Rimpler H, Ghaleb H, Heiba HI. 1986. Constituents of plants growing in Qatar. IX. The antispasmodic<br />

components of Cymbopogon parkeri Stapf. Int J Crude Drug Res 24: 69–74.<br />

Rohdich F, Wungsintaweekul J, Fellermeier M, Sagner S, Herz S, Kis K, Eisenreich W, Bacher A, Zenk MH.<br />

1999. Cytidine 5′-triphosphate-dependent biosynthesis of isoprenoids: YgbP protein of Escherichia<br />

coli catalyzes the formation of 4-diphosphocytidyl-2-C-methylerythritol. Proc Natl Acad Sci USA 96:<br />

11758–11763.<br />

Rohdich F, Wungsintaweekul J, Lüttgen H, Fischer M, Eisenreich W, Schuhr CA, Fellermeier, M, Schramek<br />

N, Zenk MH, Bacher A. 2000. Biosynthesis of terpenoids: 4-Diphosphocytidyl-2-C-methyl-d-erythritol<br />

kinase from tomato. Proc Natl Acad Sci USA 97: 8251–8256 (First Published July 4; 10.1073/<br />

pnas.140209197).<br />

Rohmer M. 1999. The discovery of a mevalonate-independent pathway for isoprenoid biosynthesis in bacteria,<br />

algae and higher plants. Nat Prod Rep 16: 565–574.<br />

Rouesti P, Voriate G. 1960. Lemongrass extract in Italian Somali Land. Fr Parf 3: 39–44 (C.A. 55, 7765-b).<br />

Rout PK, Sahoo S, Rao YR. 2005. <strong>Essential</strong> oil composition of Cymbopogon microstachys (Hook.) Soenarke<br />

occurring in Manipur. J Essen <strong>Oil</strong> Res 17: 358–360.<br />

Saeed T, Sandra PJ, Verzele MJE. 1978. Constituents of the essential oil of Cymbopogon jwarancusa.<br />

Phytochemistry 17(8): 1433, 1434.<br />

Sangwan NK, Verma KK, Verma BS, Malik MS, Dhinsa KS. 1985. Nematicidal activity of essential oil of<br />

Cymbopogon grasses. Nematologica 31: 93–99.<br />

Sarer E, Scheffer JJC, Svendsen AB. 1983. Composition of the essential oil of Cymbopogon citratus Stapf.<br />

cultivated in Turkey. Sci Pharm 51: 58–63 (C.A. 99, 10685-k).<br />

Sargenti SR, Lancas FM. 1997. Supercritical fluid extraction of Cymbopogon citratus. Chromatographia 46:<br />

285–290.<br />

Schwander J, Müller C, Zeidler J, Lichtenthaler HK. 1999. Cloning and heterologous expression of a cDNA<br />

encoding 1-deoxy-D-xylulose-5-phosphate reductoisomerase of Arabidopsis thaliana. FEBS Lett 455:<br />

140–144.<br />

Selvag E, Holm IO, Thune P. 1995. Allergic contact dermatitis in an aroma therapist with multiple sensitization<br />

to essential oils. Contact dermatitis 33: 354, 355.<br />

Shadab-Qamar E, Hanif M, Choudhary FM. 1992. Antifungal activity of lemongrass essential oils. Pakistan J<br />

Sci Ind Res 35: 246–249.<br />

Shahi AK, Sen DN. 1993. <strong>Essential</strong> oil yielding grasses of the Indian Thar desert. Agric Equip Int 45: 62–65.<br />

Shahi AK, Sen DN, Mathela CS. 1990. Chemical studies on Cymbopogon schoenanthus (Linn). Spreng a grass<br />

of Indian Thar desert. Indian Drugs 28: 37, 38.<br />

Shahi AK, Sen DN. 1989. Note on Cymbopogon jwarancusa (Jones) Schult: Source of piperitone in Thar desert.<br />

Curr Agric 13: 99–101.<br />

Shahi AK, Sharma SN, Tava A. 1997. Composition of Cymbopogon pendulus (Nees ex Steud.) Wats., an elemicin<br />

rich oil grass grown in Jammu region of India. J Essen <strong>Oil</strong> Res 9: 561–563.<br />

Shahi AK, Tava A. 1993. <strong>Essential</strong> oil compositions of three Cymbopogon species of Indian Thar desert. J Essen<br />

<strong>Oil</strong> Res 5: 639–343.<br />

Shahi AK. 1992. A search of Cymbopogon jwarancusa (Jones) Schult. subsp. olivieri (Boiss.) Soenakro. A<br />

citral yielding grass from Indian Thar desert. Indian Perfumer 36: 182–184.<br />

Sharma ML, Pandey MB, Khanna KR, Kapoor LD. 1972. <strong>Essential</strong> oils from plants raised on alkaline soil.<br />

Indian Perfumer 16: 27–30.


Chemistry and Biogenesis of <strong>Essential</strong> <strong>Oil</strong> from the Genus Cymbopogon 105<br />

Sharma ML, Srivaastava GS, Singh A. 1980. <strong>Essential</strong> oil from C. olivieri (Boiss.) Bor.). Indian Perfumer 24:<br />

17–19.<br />

Sharma PR, Mondhe DM, Muthiah S, Pal HC, Shahi AK, Saxena AK, Qazi GN. 2009. Anticancer activity of an<br />

essential oil from cymbopogon flexosus Chemica Biological Interactions 179(2–3): 160–168.<br />

Shaw DW. 1971. Roll-on insect repellent. Can Pat. 865069 (C.A.74: 140098-d).<br />

Shrivastava S, Rothria A, Misra N. 1990. Evaluation of essential oil of Cymbopogon martinii (Roxb.) Wats. for<br />

the control of Fusarium oxysporurn f. sp. udum casual organism of wilt of Arhar (Cajnus cajon). J Nat<br />

Acad Sci India 60B: 291, 292.<br />

Shylaraj KS, Thomas J. 1992. Effect of irradiation on the oil content of palmarosa (C. martinii var. motia).<br />

Indian Perfumer 36: 168–170.<br />

Siddique-Ullah M, Haque MM, Kalimuddin M. 1979. Dissolving pulp from lemongrass stem (C. flexuosus<br />

Linn.) by prehydrolysis Kraft process. Bangladesh J Sci 14: 359–361.<br />

Siddiqui MS, Mishra LN, Nigam MC, Abu-AI-Futuh, IM. 1980. Chemotaxonomie von Cymbopogon: Gas<br />

chromatographische untersuchung des etherischem oil von Cymbopogon proximus. Parfum Kosmet 61:<br />

419, 420.<br />

Siddiqui MS, Mohammad F, Srivastava SK, Gupta RK. 1979. A thin layer chromatographic method for estimation<br />

of geraniol in oils of palmarosa. Perfum Flavorist 4: 19, 20.<br />

Siddiqui MS, Sen T, Nigam MC, Datta SC. 1975. Isolation of alcoholic constituents of the oil of citronella<br />

(Java) by sodium complex method. Parfum Kosmet 56: 193, 194.<br />

Siddiqui N, Garg SC. 1990. In vitro anthelmintic activity of some essential oils. Pak J Sci Ind Res 33: 536,<br />

537.<br />

Singh A, Balyan SS, Shahi AK. 1978. Harvest management studies and yield potentiality of Jammu lemongrass.<br />

Indian Perfumer 22: 189–191.<br />

Singh AK, Dikshit A, Sharma ML, Dixit SN. 1980. Fungitoxic activity of some essential oils. Econ Bot 34:<br />

186–190.<br />

Singh AK, Naqvi AA, Ram G, Singh K. 1994. Effect of bay storage on oil yield and quality of three Cymbopogon<br />

species (C. winterianus, C. martinii, and C. flexuosus) during different harvesting seasons. J Essen <strong>Oil</strong><br />

Res 6: 289–294.<br />

Singh AK, Ram G, Sharma S. 1996. Accumulation pattern of important monoterpenes in the essential oil of<br />

citronella Java (C. winterianus) during one year of crop growth. J Med Arom Plant Sci 18: 883–887.<br />

Singh AP, Sinha GK. 1976. Study of essential oil of C. distans Linn. Wats. Indian Perfumer 20: 67–70.<br />

Singh RS, Pathak MG. 1994. Herb yield and volatile constituents of cymbopogon jwaranusa (Jones) Schult<br />

cultivars. Industrial Crops and Products 2(3): 197–199.<br />

Singh RS, Pathak MG, Singh KK. 1970. Dynamics and diurnal changes in oil of Java citronella (C. winterianus<br />

Jowitt). Indian Perfumer 23: 116–120.<br />

Sinha AK, Mehra MS. 1977. Study of some aromatic medicinal plants. Indian Perfumer 22: 129–131.<br />

Sobti SN, Bradu BL, Rao BL, Verma V, Jamwal PS. 1978a. Search for aroma chemicals of industrial value from<br />

the genus Cymbopogon Pt. VI. C. khasianus from Arunanchal Pradesh—a new source of citral. Indian<br />

Perfumer 22: 222, 223.<br />

Sobti SN, Rao BL, Jani BB. 1978b. Genetic variability in oils of Cymbopogon: Scope for evolving new strains.<br />

Indian Perfumer 22: 281–283.<br />

Sobti SN, Rao BL, Pushpangadan P, Verma V. 1978c. Investigations in some newly introduced Cymbopogon<br />

species. Indian Perfumer 22: 278–280.<br />

Sobti SN, Rao BL, Sharma SN, Atal CK. 1981. Jamrosa, a new geraniol rich Cymbopogon. Indian Perfumer<br />

25: 116–118.<br />

Sobti SN, Verma V, Rao BL. 1982. Scope for development of new cultivars of cymbopogons as a source of terpene<br />

chemicals. In Cultivation and Utilization of Aromatic Plants (Atal CK, Kapur BM, Eds), Regional<br />

Reasearch Laboratory, Jammu, India.<br />

Soulari M, Fanghaenel E. 1971. <strong>Essential</strong> oil of C. winterianus (citronella oil) produced in Cuba. Rev Cenic<br />

Cienc Fis. 3: 79–91.<br />

Sprenger GA, Schörken U, Wiegert T, Grolle S, De Graaf AA, Taylor SV, Begley TP, Bringer-Meyer S, Sahm H.<br />

1997. Identification of a thiamin-dependent synthase in Escherichia coli required for the formation of the<br />

1-deoxy-d-xylulose 5-phosphate precursor to isoprenoids, thiamin, and pyridoxol. Proc Natl Acad Sci<br />

USA 94: 12847–12862.<br />

Srikulvandhana S, Jennings WG, Derafols W. 1976. Lemongrass oil from mutant clones of Cymbopogon flexuosus.<br />

Chem Mikrobiol Technol Lebensm 4: 129–131.<br />

Sukari MA, Rahmani M, Manas A, Takahashi S. 1992. Toxicity studies of plants extracts on insects and fish<br />

Partanika 15: 41–44.


106 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Syeed M, Khalid MR, Choudhary FM. 1990. <strong>Essential</strong> oil of Gramineae family having antibacterial activity.<br />

Part I. C. citratus, C. martinii, C. jwarancusa oils. Pak J Sci Ind Res 33: 529–531.<br />

Takahashi S, Kuzuyama T, Watanabe H, Seto H. 1998. A 1-deoxy-d-xylulose 5-phosphate reductoisomerase<br />

catalyzing the formation of 2-C-methyl-d-erythritol 4-phosphate in an alternative non-mevalonate pathway<br />

for terpenoid biosynthesis. Proc Natl Acad Sci USA 95: 9879–9884.<br />

Talat S, Patrick JS, Maurice JEV. 1978. Constituent of the essential oil of cymbopogon jwarancusa. Phytochemistry<br />

17(8): 1433–1434.<br />

Taskinen J, Mathela DK, Mathela CS. 1983. Composition of the essential oil of Cymbopogon flexuosus.<br />

J Chromatogr 262: 364–366.<br />

Thapa RK, Agarwal SG, Dhar KL, Atal CK. 1981. Citral containing Cymbopogon species. Indian Perfumer<br />

25: 15–18.<br />

Thapa RK, Agarwal SK. 1989. Cymbopogon flexuosus: A rich source for (±) α-bisabolol. J Essen <strong>Oil</strong> Res<br />

107–110.<br />

Thapa RK, Bradu BL, Vashist VN, Atal CK. 1971. Screening of Cymbopogon species for useful constituents.<br />

Flav Industry 2: 49–51.<br />

Thapa RK, Dhar KL, Atal CK. 1976. The essential oil of Cymbopogon flexuosus RRL-57 and RRL-59, Part II.<br />

Indian Perfumer 20: 39–44.<br />

Thieme H, Benecke R, Brotka J. 1980. Application of gas chromatography in determining the constituents<br />

of Oleam citronella. Zentralbl Pharm Pharmakother Laboratoriums Diagn 119: 953–957 (C.A. 94,<br />

36417-w).<br />

Torres RC, Ragadio AG. 1992. Chemical composition of the essential oil of Philippines Cymbopogon citratus<br />

D.C Stapf. The Asian symposium of Medicinal Plants, Spices and other Natural Products (ASOMAS<br />

VII), Manila, 2–7 February.<br />

Torres RC, Ragodio AG. 1996. Chemical composition of the essential oil of Philippine Cymbopogon citratus<br />

(D.C.) Stapf. Philippine J Sci 125: 147–156.<br />

Torres RC. 1993. Citral from Cymbopogon citratus (D.C.) Stapf. lemongrass oil. Philippine J Sci 122: 269–278.<br />

Verma V, Sobti SN, Atal CK.1987. Chemical composition and inheritance pattern of five Cymbopogon species.<br />

Indian Perfumer 31: 295–305.<br />

Vijender SM, Mohammed A. 2002. Volatile constituents of Cymbopogon nardus (Linn.) Rendle. Flav Frag J<br />

18(1): 73–76.<br />

Virmani OP, Datta SC. 1973. Scope of production of some essential oil in the Lucknow district. Indian Perfumer<br />

17: 35–41.<br />

Virmani OP, Datta SC. 1971. <strong>Essential</strong> oil of Cymbopogon winterianus (oil of citronella). Flav Ind 2: 595–602.<br />

Vole CD, Baeta J, Antonio P da Cunha. 1997. Comparative chemical study of inflorescence of Cymbopogon<br />

gigariteum. Bot Esc Farm Univ Coimbra Ed Cient 27: 17–52.<br />

Wannissom B, Jarikasen S, Soontorn TT. 1996. Antifungal activity of lemongrass oil and lemongrass oil cream.<br />

Phytother Res 10: 551–554.<br />

Wijesekera ROB, Jayewardene AL, Foneska BD. 1973a. Varietal difference in the constituents of citronella oil.<br />

Phytochemistry 12: 2697–2704.<br />

Wijesekera ROB, Jayewardene AL, Foneska BD. 1973b. The chemical composition and analysis of citronella<br />

oil. J Nat Sci Coun Sri Lanka 1: 67–81 (C.A. 81.111382-t).<br />

Yadav P, Dubey NK. 1994. Screening of some essential oils against ring worm fungi. Pharma Sci 56:<br />

227–230.<br />

Zaki MSA, Foda YH, Mustafa MM, Abd Allah MA. 1975. Identification of the volatile constituents of the<br />

Egyptian lemongrass oil II. Thin layer chromatography. Nahrung 19: 201–205.<br />

Zamureenka VA, Klyuev NA, Grandberg IH, Dmitriev LB, Esvandghya GA. 1981. Composition of essential oil<br />

of lemongrass (C. citratus D.C.). Izv Timiryazensk S-Kh Akad 2: 167–169 (C.A. 94, 145166-j).<br />

Zheng GQ, Kenney PM, Tam TDT. 1993. Potential anticarcinogenic natural products isolated from lemongrass<br />

oil and galanga root oil. J Agric Food Chem 41: 153–156.


3<br />

contents<br />

The Cymbopogons<br />

Harvest and Postharvest<br />

Management<br />

A. K. Pandey<br />

3.1 Introduction .......................................................................................................................... 108<br />

3.2 Lemongrass ........................................................................................................................... 109<br />

3.2.1 Harvesting ................................................................................................................. 109<br />

3.2.2 Yield .......................................................................................................................... 112<br />

3.2.3 Postharvest Management .......................................................................................... 112<br />

3.2.3.1 Predistillation Handling ............................................................................. 112<br />

3.2.3.2 Drying ........................................................................................................ 113<br />

3.2.3.3 Storage ....................................................................................................... 113<br />

3.2.3.4 Distillation ................................................................................................. 113<br />

3.2.3.5 Supercritical Fluid Extraction (SFE) ......................................................... 115<br />

3.2.3.6 Treatment of <strong>Oil</strong> Prior to Storage .............................................................. 116<br />

3.2.4 Quality Analysis ....................................................................................................... 116<br />

3.2.4.1 Standard Specifications .............................................................................. 117<br />

3.3 Palmarosa (Cymbopogon martinii var. motia (Roxb.) Wats.) ............................................... 117<br />

3.3.1 Uses ........................................................................................................................... 118<br />

3.3.2 Harvesting ................................................................................................................. 118<br />

3.3.3 Storage ...................................................................................................................... 119<br />

3.3.4 Yield .......................................................................................................................... 119<br />

3.3.5 Distillation ................................................................................................................120<br />

3.3.6 <strong>Oil</strong> Storage ................................................................................................................120<br />

3.3.7 <strong>Oil</strong> Content and Yield ...............................................................................................120<br />

3.3.8 Standard Specifications ............................................................................................ 121<br />

3.4 Citronella (Cymbopogon winterianus Jowitt) ....................................................................... 121<br />

3.4.1 Harvesting ................................................................................................................. 122<br />

3.4.2 Storage of C. winterianus Hay .................................................................................126<br />

3.4.3 Yield ..........................................................................................................................126<br />

3.4.4 Distillation Procedure ...............................................................................................126<br />

3.4.5 Standard Specifications ............................................................................................126<br />

3.5 Jamrosa (Cymbopogon nardus Rendle) ................................................................................ 127<br />

3.5.1 Uses ........................................................................................................................... 127<br />

3.5.2 Harvesting ................................................................................................................. 127<br />

3.5.3 Distillation ................................................................................................................128<br />

3.5.4 Yield ..........................................................................................................................128<br />

3.6 Conclusion ............................................................................................................................128<br />

References ...................................................................................................................................... 129<br />

107


108 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

3.1 IntroductIon<br />

Cymbopogon (Poaceae) represents an important genus of about 120 species that grow in tropical<br />

and subtropical regions around the world. On account of their diverse uses in pharmaceutical,<br />

cosmetics, food and flavor, and agriculture industries, Cymbopogon grasses are cultivated (medicultured)<br />

on a large scale, especially in the tropics and subtropics. There is a large worldwide demand<br />

for the essential oils of Cymbopogon species (Dutta 1982; Gunther 1956). They are well known as<br />

a source of commercially valuable compounds such as geraniol, geranyl acetate, citral (neral and<br />

geranial), citronellal, piperitone, eugenol, etc., which are either used as such in perfumery and allied<br />

industries or as starting materials for the synthesis of other products commonly used in perfumery<br />

(Shahi and Tava 1993). Distillation of the grass produces an essential oil and a hydrosol (distillate<br />

water) that have powerful antibiotic, antiviral, and antifungal properties which are used effectively<br />

against infectious and inflammatory symptoms. Several Cymbopogon species are being cultivated<br />

in different parts of world. Lemongrass, palmarosa, and citronella essential oils are the main raw<br />

material products of the cultivated cymbopogons. However, other Cymbopogon species are also<br />

grown in other parts of the world (Oyen and Dung 1999).<br />

Volatile constituents of the essential oil of C. caesius were studied by Kanjilal et al. (1995). The<br />

main constituents were perillyl alcohol (25.61%), geraniol (19.80%), and limonene (7.26%) along<br />

with 21 other compounds. Choudhury and Leclercq (1995) also studied the essential oil composition<br />

of C. khasianus (Munro ex Hack.) Bor from northeastern India. C. pendulus, an elemicin-rich<br />

aromatic grass of the Meghalaya region of India, grew well under the subtropical climatic conditions<br />

at Jammu, India. The essential oil obtained from this plant was rich in elemicin (53.7%),<br />

a starting material for developing the antibacterial drug trimethoxyprim (Shahi et al. 1997).<br />

Chowdhury et al. (1998) studied the essential oil of Cymbopogon species growing in Bangladesh.<br />

<strong>Essential</strong> oil of C. nardus (L.) Rendle growing in Zimbabwe was studied by Moody et al. (1995).<br />

Comparative investigation of the sesquiterpenoids present in the leaf oil of C. distans (Steud.) Wats.<br />

var. loharkhet and the root oil of C. jwarancusa (Jones) Schult. was performed by Beauchamp et al.<br />

(1996). Chisowa (1997) studied the chemical composition of flower and leaf oils of C. densiflorus<br />

Stapf from Zambia.<br />

A wide range of variation has been observed in the oil content of Cymbopogon species, and this<br />

is influenced by genetic, agronomic, and geoclimatic factors (Rao et al. 1980; Patra et al. 1990;<br />

Pandey and Chowdhury 2000). It is reported that oil content is lower during the month of heavy<br />

rainfall compared to the dry months (Guenther 1961). Similarly, monthly variation in the oil content<br />

in lemongrass over a year has been studied (Handique et al. 1984). It has also been reported<br />

that in some aromatic crops, the factors photoperiod, intensity of light, temperature (Voirin et al.<br />

1990; Lincoln and Langenhein 1978; Clark and Menary 1980), and seasons or months of harvesting<br />

(Rudloff and Underhill 1965; Adams 1970; Singh et al. 2000) exert a profound influence on the<br />

essential oil content and terpenoid composition of these crops. Diseases such as iron chlorosis significantly<br />

reduced biomass, essential oil yields, and total chlorophyll content of the leaves of Java<br />

citronella (C. winterianus), lemongrass (C. flexuosus var. flexuosus), and palmarosa (C. martinii var.<br />

motia) plants.<br />

The presence, yield, and composition of essential oils have been affected in a number of ways<br />

by various factors, from their formation in plants to their final isolation. Several of the factors<br />

of influence have been studied, particularly for commercially important crops, to optimize the<br />

cultivation conditions and time of harvest and to obtain higher yields of high-quality essential<br />

oils that fit market requirements. Knowledge of factors that determine the oil yield and chemical<br />

variability of aromatic plants species are thus very important. These include (1) physiological<br />

variations, (2) environmental conditions, (3) geographic variations, (4) genetic factors and<br />

evolution, (5) political/social conditions, and also (6) harvest time and technique (Figueiredo<br />

et al. 2008; Singh et al. 1996; Dhar et al. 1996b, 1996c; Costa et al. 2005). This chapter gives an


The Cymbopogons 109<br />

account of the harvesting and distillation practices that will optimize yield of quality essential<br />

oil from Cymbopogon species (Cassel and Vergas Rubem 2006).<br />

3.2 lemongrass<br />

Lemongrass, a perennial herb widely cultivated in the tropics, encompasses three different species:<br />

C. flexuosus (Steud.) Wats. (East Indian), C. citratus Stapf (West Indian), and C. pendulus (North<br />

Indian). The common name lemongrass has been given to these species because of the typical<br />

strong lemongrass-like odor of the essential oil present in the leaves. Two species, C. flexuosus and<br />

C. pendulus, are cultivated in India, whereas C. citratus is cultivated in the West Indies, Guatemala,<br />

Brazil, etc. C. citratus is a tufted perennial grass with numerous stiff leafy stems arising from<br />

short rhizomatous rootstocks. The aboveground parts, which contain the oil, grow to 2 m in height.<br />

A number of cultivars are acknowledged, which differ considerably in yield and citral content (Nair<br />

et al. 1979).<br />

Lemongrass has been used fresh, dried, or powdered. The fresh stalks are found in Asian markets<br />

and now in many health food markets. Lemongrass is widely used in Thai and Vietnamese cooking.<br />

This aromatic herb is used in Caribbean and many types of Asian cooking, and has become very<br />

popular in the United States. Lemongrass has been used for centuries in Indonesia and Malaysia by<br />

herbalists, and it is also used in Ayurvedic herbalism. It is used in teas to combat depression and bad<br />

moods, and to fight fever and combat nervous and digestive disorders. Studies show that lemongrass<br />

has antibacterial (Burt 2004; Hussain 1994) and antifungal properties (Dikshit and Hussain 1984;<br />

Wannissorn et al. 1996; Pandey et al. 1996; Paranagama et al. 2003; Mahanta et al. 2007). The oil<br />

is used to cleanse oily skin, and in aromatherapy, it is used as a relaxant. Valued for its exotic citrus<br />

fragrance, it is commercially used in soaps, perfumes, cosmetics, and disinfectants, and is a raw<br />

material for manufacturing ionones and vitamin A.<br />

The leaves yield aromatic oil, containing 70%–90% citral (the aldehyde responsible for the lemon<br />

odor). The quality of lemongrass is generally determined by its citral content (Chisowa et al. 1998).<br />

Citral consists of the cis-isomer geranial and the trans-isomer neral. These two are normally present<br />

in the ratio of about 2 to 1. C. flexuosus has higher citral content than C. citratus (Weiss 1997;<br />

Taskinen et al. 1983).<br />

Lemongrass grows wild across the tropics, and the content and quality of the oil varies among<br />

provenances. It prefers a warm climate with well-distributed rainfall and well-drained soil. Usually,<br />

it grows on poor, gravelly soils. Lemongrass is a perennial grass mainly cultivated on hill slopes as<br />

a rainfed crop (Pandey et al. 2001). The crop provides maximum yield from the second to fourth<br />

year of planting and economic yield up to the fifth year. Thereafter the yield declines considerably.<br />

Different cultivars of lemongrass, for example, CKP-25, OD 19, Cauvery, Pragati, and Praman, were<br />

evaluated for herbage and oil yield. Brief information about various cultivars is given in Table 3.1.<br />

Cauvery recorded the highest oil yield (Singh 1997; Rao and Lala, 1992).<br />

3.2.1 ha r V e s t i n G<br />

Harvesting time is one of the most important factors influencing optimum superior-quality oil yield.<br />

Depending on soil and climatic conditions, a lemongrass plantation lasts, on average, for 3–5 years<br />

only. The yield of biomass and oil is less during the first year, but it increases in the second year and<br />

reaches a maximum in the third year; after this, the yield declines. Correct harvesting procedure is<br />

very important. The essential oil content varies considerably during the development of the plant. If<br />

the plant is harvested at the wrong time, the oil yield can be severely reduced. The oil is usually contained<br />

in oil glands, veins, or hairs that are often very fragile. Handling will break these structures<br />

and release the oils. This is why a strong smell is given off when these plants are handled. Therefore,


110 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 3.1<br />

currently grown Varieties of lemongrass and their description<br />

Variety description<br />

Sugandhi (OD 19) It is adapted to a wide range of soil and climatic conditions.<br />

A red-stemmed variety with a plant height of 1 to 1.75 m and profuse tillering.<br />

The oil yield ranges from 80 to 100 kg per hectare with 85%–88% of total citral produced under<br />

rainfed conditions (with life-saving irrigation).<br />

Pragati It is a tall-growing variety with dark purple leaf sheath suitable for North Indian Plains and the Tarai<br />

belt of subtropical and tropical climate.<br />

Average oil content is 0.63% with 75%–82% citral.<br />

Praman It has evolved through clonal selection and belongs to the species C. pendulus.<br />

It is a medium-sized variety with erect leaves and profuse tillering.<br />

The oil yield is high with 82% citral.<br />

Jama rosa It is very hardy, with vigorous growth.<br />

The variety yields about 35 t of herbage per hectare, containing 0.4% oil (FWB).<br />

The variety yields up to 300 kg oil in 4–5 cuts in 16–18 months of growing period.<br />

RRL 16 Average herbage yield of this variety is 15–20 t/ha/annum, giving 100–110 kg oil.<br />

The oil content varies from 0.6% to 0.8% (fresh weight basis) with 80% citral.<br />

CKP 25 This is a hybrid between C. khasianum × C. pendulus.<br />

It gives 60 t/ha herbage in North Indian plains under irrigation.<br />

The oil contains 82.85% citral.<br />

Other varieties OD-408, Cauvery (OD-408 is a white-stemmed selection of OD-19 and is an improvement in yield in<br />

terms of oil and citral content. Cauvery needs high soil moisture for luxuriant growth and was<br />

evolved for river valley tracts.)<br />

these plants have to be handled very carefully to prevent valuable oils from being lost. Harvesting is<br />

done with the help of sickles; the plants are cut 10 cm above ground level and are allowed to wilt in<br />

the field before being transported to the distillation site. Seasonal influence on lemongrass has been<br />

reported (Handique et al. 1984; Thomas et al. 1980).<br />

During the first year of planting, three cuttings are obtained and, subsequently, five to six cuttings<br />

per year are taken, subject to weather conditions (Rana et al. 1996). The harvesting season<br />

begins in May and continues until the end of January. The first harvest is generally obtained after<br />

4–6 months of transplanting seedlings. Subsequent harvests are done at intervals of 60–70 days,<br />

depending on the fertility of the soil and other seasonal factors. Under normal conditions, three<br />

harvests are possible during the first year and three to four in subsequent years, depending on<br />

the management practices followed. The optimum interval between harvests to obtain maximum<br />

quantity of oil is 40–45 days for local types of lemongrass. For OD-19, the optimum interval was<br />

found to be 60–65 days when grown in hilltops and 45–55 days in valleys and lower areas (Jha et al.<br />

2004). Singh et al. (2000) have reported that the first harvest should be taken 90 days after planting<br />

to boost the development of tillerings in lemongrass. Under Jammu (India) conditions, more tillers<br />

with fewer harvests of lemongrass was also reported (Pal et al. 1990; Singh et al. 1978).<br />

Rao et al. (2005) conducted an experiment in Bengaluru (formerly Bangalore), Karnataka, India,<br />

during 2001–2003 to study the effect of harvest intervals on oil and citral accumulation in C. flexuosus<br />

cv. Krishna. The highest percentages of oil (4.8) and citral (87.1) were obtained in lemongrass<br />

harvested on February 17, 2002, whereas the highest percentages of geraniol (9.3) and geranyl<br />

acetate (2.5) were obtained with July 21, 2003, harvest. Variations in major chemical constituents<br />

in the oil of C. flexuosus were found in different seasons under Brahmaputra valley agroclimatic<br />

conditions (Sarma et al. 2003). Effect of leaf position and age on quality of oil has been discussed<br />

(Singh et al. 1989) whereas sucrose metabolism to components of essential oil have been studied by<br />

Singh and Luthra in 1988.


The Cymbopogons 111<br />

table 3.2<br />

effect of different harvesting Intervals on herb yield, oil<br />

content, and oil yield of lemongrass<br />

cutting<br />

Fresh herb yield (q/ha) oil content (%) oil yield (l/ha)<br />

75 a 100 125 75 a 100 125 75 a 100 125<br />

First 69.4 136.0 147.6 0.50 0.53 0.56 35.1 72.2 82.7<br />

Second 152.3 141.6 50.0 0.56 0.61 1.02 85.9 86.8 51.7<br />

Third 77.4 — — 0.77 — — 58.0 — —<br />

CD 5% 62.0 NS 35.4 0.16 NS 0.12 NS NS 19.2<br />

Note: NS = Not significant, quintal.<br />

a Harvesting intervals (in days).<br />

Source: Gill B S et al. 2007. India Perfumer 51: 23–27.<br />

table 3.3<br />

biomass yield and essential oil concentration in different Varieties<br />

of lemongrass at different harvests<br />

Varieties<br />

First<br />

harvest<br />

biomass yield (t/ha) essential oil concentration (%)<br />

second<br />

harvest<br />

third<br />

harvest<br />

total<br />

harvest<br />

First<br />

harvest<br />

second<br />

harvest<br />

third<br />

harvest<br />

OD-19 7.9 4.3 16.0 28.2 0.62 0.75 0.66<br />

Cauvery 6.3 3.5 25.1 34.9 0.88 1.05 1.02<br />

Pragati 7.7 3.4 22.5 33.6 0.71 0.80 0.70<br />

SHK-7 12.5 4.0 24.4 40.9 0.42 0.51 0.57<br />

Praman 2.2 3.8 26.6 32.6 0.63 0.71 0.61<br />

LSD (P = 0.05) 4.5 NS 7.2 11.9 0.18 0.04 0.14<br />

Note: NS = Not significant.<br />

Source: Rajeswara Rao B R et al. 1998. Journal of Medicinal and Aromatic Plant Sciences 20:<br />

407–412.<br />

Harvesting intervals are determined by infrastructure, management, climate, and cultivar. If<br />

harvested too often, the productivity of the plant will be reduced, and the plant may even die. If the<br />

plant is allowed to grow too large, the oil yield will be reduced. It should be 1.2 m high with four to<br />

five leaves. The grass should be harvested early in the morning if it is not raining. Manual harvesting<br />

is common, but harvesting methods adopted depend on infrastructure. The bush is cut from<br />

7 cm (manually) to 25 cm (mechanically) above ground. Weiss (1997) reports that each worker in<br />

Sri Lanka harvests 2 t of fresh grass daily.<br />

An experiment was conducted in Punjab, India, to determine harvest intervals in lemongrass<br />

(Gill et al. 2007). The results given in Table 3.2 reveal that maximum herb and oil yields were<br />

obtained when the cutting was taken 125 days after general harvest (in the first week of March)<br />

and for subsequent cuttings taken after 75-day harvesting intervals. The oil content was maximum<br />

75 days after general harvesting, and the oil content decreased with delay in harvesting, that is, from<br />

75 to 125 days in the first as well as in the second and third cuttings. The decrease in oil content<br />

with delay in harvesting could be attributed to the fact that as the age of the crop increases, the plant<br />

becomes woodier and lower leaves become dry.<br />

The effects of plant age (60, 67, 74, 81, 88, 95, 102, 109, and 116 days) on C. citratus biomass<br />

production and essential oil yield were studied in Campos dos Goytacazes, Rio de Janeiro, Brazil.


112 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

The essential oil yield decreased as plant age increased. Nevertheless, the increase in dry-matter<br />

production with plant age resulted in an increase in the total essential oil production (Leal et al.<br />

2003). <strong>Essential</strong> oil from different plant parts of lemongrass grown in Brazil was studied by Ming<br />

et al. (1995). The study revealed that leaf blade contains 0.42% essential oil, whereas leaf sheath has<br />

0.13% essential oil.<br />

A study was conducted in the Western Cape, which has a Mediterranean climate. The plot<br />

was first harvested manually when the plants were 6 months old and every month thereafter. Leaf<br />

growth was significantly slower in winter; therefore, a monthly cut during the colder periods was not<br />

possible. Studies suggest that oil yield and citral content increase in hot dry seasons (Weiss 1997).<br />

Frequent cutting of C. flexuosus can increase total oil yield (Weiss 1997); a similar scenario exists<br />

for C. citratus.<br />

Experiments with irrigated lemongrass (C. citratus) in Western Australia have shown that the<br />

highest oil yield of 419 L/ha over a 360-day period was obtained when the plants were cut at 60-day<br />

intervals and at a height of 20 cm. Longer intervals and higher cutting heights gave lower oil yields,<br />

although in some cases, fresh and dry-matter yields were increased. Studies on the effect of water<br />

stress showed that time between irrigations in the dry season should not be more than 10 days if oil<br />

yields are to be maintained. Wilting of cut lemongrass in the dry season was shown to result in a<br />

loss of oil, with losses increasing with the duration of wilting up to 11 h (Beech 1997). The effects of<br />

seasonal variation and harvest time on the foliar content of essential oil from lemongrass C. citratus<br />

were also studied by Beech (1990) and Leal et al. (2001).<br />

The effects of harvesting time (07.00, 09.00, 11.00, 13.00, 15.00, or 17.00 h) on the essential oil<br />

content of lemongrass (C. citratus) and on the citral and myrcene contents of the essential oil were<br />

studied. The highest essential oil content was obtained at 09.00 h (5.59 mL/kg dry matter) and 11.00 h<br />

(5.31 mL/kg dry matter). The average citral (61.23%) and myrcene (24.14%) contents suggested that<br />

the essential oil was probably of the West Indian type. The results indicated that lemon grass may be<br />

harvested between 09.00 and 11.00 h for optimum essential oil, citral, and myrcene yields.<br />

Outbreaks of Puccinia nakanishikii on commercial plantations of C. citratus caused reductions<br />

in the yield of essential oils. Healthy, uninfected leaves yielded 0.80% oil, while leaves with a<br />

disease index of 60%–75% yielded 0.50% oil. This reduction in essential oil yield was also accompanied<br />

by a decrease in the content of geraniol, and increases in the contents of neral and myrcene<br />

(Boruah et al. 1995).<br />

3.2.2 yield<br />

The grass yield during the first year was about 10 t/ha, which gives about 28 kg of oil. From the second<br />

year onward, the grass yield was about 25 t/ha, giving about 75 kg of oil. On an average, 25–30 t<br />

of fresh herbage are harvested per hectare per annum from four to six cuttings, which yields about<br />

80 kg of oil. Under irrigated conditions, from newly bred varieties, an oil yield of 100–150 kg/ha<br />

is obtained. The average recovery of oil is 0.30%–0.35% with 70% citral for local types of lemongrass,<br />

while OD-19 variety gives 0.40%–0.45% oil recovery and 85%–90% citral content. C. citratus<br />

yields 30–50 t/ha and continues around this level for its 4–6-year plantation life. <strong>Oil</strong> yield from<br />

fresh herbage is 0.25%–0.5% and even 0.4%–0.6% with good management. This data is from Weiss<br />

(1997), but a wide range of variation in herbage and oil yield has been reported with infrastructure<br />

and management playing a significant role.<br />

3.2.3 PostharVest Ma n a G e M e n t<br />

3.2.3.1 Predistillation handling<br />

The cut grass may be distilled fresh, but some natural reduction in the moisture content by withering<br />

in the sun allows greater still vessel packing and oil recoveries per batch distillation. Wilting is done


The Cymbopogons 113<br />

in the field where grass lies after cutting, rather than in heaps, but it should not exceed 24 h or oil<br />

losses may occur through brittleness and evaporation. Around 250 kg of partially wilted grass can be<br />

packed into 1000 L of a still vessel. The crop is chopped into small pieces before filling the stills.<br />

3.2.3.2 drying<br />

Drying techniques influence the essential oil yield and composition. The grass is allowed to wilt for<br />

24 h before distillation, as it reduces the moisture content by 30% and improves oil yield. In Brazil,<br />

a study was conducted to analyze the influence of drying on yield and composition of essential<br />

oil. The temperature varied from 40°C to 60°C, and the air velocities investigated were 0.2, 0.5,<br />

and 0.8 m/s. The highest yield was obtained at 60°C, the highest temperature investigated, and at<br />

0.8 m/s. At air velocities near 0.2 m/s, the lowest masses of essential oil were extracted. The differences<br />

between air velocities did not influence the composition of the essential oils, but did influence<br />

the quantity of the components in the fractions and the time of drying. The essential oil extracted<br />

from the wet plants present fewer components than others, and it can be explained by the fact that<br />

water molecules solvate the components (Peisíno et al. 2005).<br />

Experiments were also conducted to study the effects of drying temperature on the amount and<br />

quality of essential oils extracted from C. citratus (Buggle et al. 1997). Leaf blades were cut into<br />

small parts (about 1–1.5 cm in length) and dried for several days at 30°C, 50°C, 70°C, or 90°C, until<br />

a constant weight was achieved. A higher amount of oil was collected at lower drying temperatures;<br />

but at 30°C, leaves were affected by fungal (Aspergillus sp., Penicillium sp., Rhizopus sp.,<br />

Cladosporium sp., Trichoderma sp., and Alternaria sp.) growth. The analysis of the oils by gas chromatography-mass<br />

spectrometry (GC-MS) showed variations in citral concentration (86.1%–95.2%).<br />

The best results were obtained at a drying temperature of 50°C (1.43% oil content).<br />

The effects of drying method (oven drying at 40°C or drying at room temperature using a moisture<br />

drier) and fragment size (powder obtained from the mill, and 1.0 cm or 20.0 cm fragments) on<br />

the yield and composition of the essential oil of lemongrass (C. citratus) were studied. The essential<br />

oil was extracted using Clevenger’s modified apparatus for 2 h. Higher essential oil and citral<br />

contents were obtained when the leaves were dried at room temperature. The fragment size had no<br />

significant effect on the evaluated parameters.<br />

3.2.3.3 storage<br />

The safe limit of herb storage varied according to the species and storage conditions. Storage of<br />

C. flexuosus herbage always caused a reduction in oil content except during the summer, when it<br />

was not affected by 3 days of storage under shade. Little variation in the geranial and neral contents<br />

of the essential oils of C. flexuosus leaves was observed during storage for 15 days. Temperature<br />

and humidity were found to play a vital role in biosynthesis accumulation of essential oils in stored<br />

herbs (Singh et al. 1994).<br />

3.2.3.4 distillation<br />

Distillation represents a dynamic part of a whole process in which the ethereal oils contained within<br />

a plant’s aromatic sac or glands are liberated through heat and pressure and transformed into a liquid<br />

essence of sublime beauty. The recovery of essential oils (the value-added product) from the raw<br />

botanical starting material is very important since the quality of the oil is greatly influenced during<br />

this step. There are different methods for obtaining volatile oils from plants, such as steam distillation,<br />

aqueous infusion, solvent extraction, cold or hot expression, and supercritical fluid extraction<br />

(SFE) with carbon dioxide. The chemical composition of the oil, both quantitative and qualitative,<br />

differs according to the technique used to obtain it from the plant parts (Sandra and Bicchi 1987).<br />

A comprehensive review of various techniques employed to obtain essential oil from the material<br />

in which it occurs was prepared by Weurman (1969). Harvesting the plant at the appropriate<br />

time and endeavoring to distill its essence is the art and craft of the distiller. The grass is either<br />

distilled afresh or is allowed to wilt for 24 h. Wilting reduces the moisture content and allows a


114 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

larger quantity of grass to be packed into the still, thereby economizing the fuel use. The current<br />

method of distillation adopted in different parts of India is primitive and gives oil of poor quality,<br />

as it is based on hydrodistillation or direct-fired still. It is being distilled in the same distilleries<br />

used for Japanese mint in India. Variation in the essential oil composition of rose-scented geranium<br />

(Pelargonium sp.) distilled by different distillation techniques, for example, water distillation, water<br />

steam distillation, and steam distillation, was studied by Kiran et al. (2004). For good-quality oil, it<br />

is advisable to adopt steam distillation. The equipment for distillation consists of a boiler to produce<br />

steam, a distillation tub, a condenser, and one to three separators.<br />

3.2.3.4.1 Steam Boiler<br />

A steam boiler is made of steel or galvanized sheets for steam distillation of essential oil. In developed<br />

countries, an oil- or gas-fired package boiler is used. Biomass-fired boilers are used in India;<br />

they are also quite common in remote third world countries. However, there are some distilleries that<br />

purchase wood, as it is a cheap local fuel, or burn fuel oil, while the spent biomass of the process is<br />

discarded, primarily for being too wet.<br />

3.2.3.4.2 Stills<br />

Stills come in all sizes, shapes, and materials of construction. A still is basically a tank with some<br />

means of injecting steam at the bottom in a way that allows its uniform distribution, such as perforated<br />

crosses or plates, false bottoms, manifolds, etc. This method is known as hydrodiffusion, as<br />

opposed to hydrodistillation. In the latter, the still is filled with the material submersed in water, and<br />

the oil is “boiled” out of the aromatic raw material. The opening of the still can be a simple manhole<br />

cover or a full-sized lid with the same diameter as the tank, depending on the unloading method.<br />

The steam/oil vapors exit at the upper ends of the still or through an opening in the lid, which is<br />

sometimes fitted with a coarse filter.<br />

Most stills operate at atmospheric pressure, but some are designed to withstand higher pressures,<br />

usually in the 2 bar range. Stills that operate under pressure are sometimes unloaded under pressure<br />

through a large opening at the end of a cone-shaped bottom. Materials that require frequent loading<br />

and unloading are processed in stills mounted on pivots that allow the still to swivel to the upsidedown<br />

position and dump the entire contents. The distillation tub is made of mild steel or copper<br />

and has a perforated bottom on which the grass rests. The tub has a steam inlet pipe at the bottom.<br />

A removable lid is fitted to the top. Charging and discharging can be done in perforated cages with<br />

iron chains, which can be lowered in the tub with the help of a chain-pulley block.<br />

3.2.3.4.3 Condensers<br />

Different types of condensers are available, but tubular condensers are better than others. The condenser<br />

is provided with an inlet and an outlet through which cold water is made to flow through the<br />

chamber to cool the pipes when the distillate flows through them. Condensers are of various types;<br />

they range from truck radiators to copper coils, shell and tube heat exchangers, pipes submersed in<br />

river-fed canals, air-cooled condensers, tube condensers inside sprinkler towers, etc., depending on<br />

the location, climate, available space, and resources.<br />

3.2.3.4.4 <strong>Oil</strong> Separators<br />

The oil separator is the one component that is most critical to overall product recovery and profitability<br />

of the plant, whether conventional or continuous. Except in modern facilities, the separator<br />

often seems to get the least engineering attention from distillery operators in the field. The separator,<br />

too, comes in a wide range of homemade designs, although the main idea is that of a continuous<br />

decanter, sometimes referred to as a Florentine flask. Its efficiency is governed by a number of wellknown<br />

variables such as oil and water specific gravity differential at various temperatures; phase<br />

viscosities versus ascending and descending cross-sectional velocities at various distillate flow<br />

rates and tank diameters; coalescing effects of different packing materials; emulsification effects;


The Cymbopogons 115<br />

oil solubility at various temperatures; chemical composition and polarity of the oil and its effect on<br />

solubility; etc.<br />

The stills in a typical distillery, usually two in number, are 6–9 ft high and 3–6 ft in diameter. After<br />

the still is tightly filled with grass, the lid is fastened on and steam let in at the bottom. The steam passes<br />

up through the grass, carrying off the oil through a water-cooled coil. The distillate consisting of water<br />

and oil is collected in steel or copper tanks about 3 ft in diameter and 18 in. deep. When the tank is<br />

nearly full, a siphon attachment begins to discharge the water in the lower level of the tank. The oil,<br />

which is lighter, floats and, when a quantity has been collected, is drawn out into bottles or drums.<br />

To obtain the maximum yield of oil and to facilitate its release, the grass is chopped into shorter<br />

lengths. Chopping the grass has further advantages in that more grass can be charged into the still<br />

and even packing is facilitated. It can be stored for up to 3 days under shade without any adverse<br />

effects on yield or quality of oil. <strong>Oil</strong> is obtained through steam distillation. The grass should be<br />

packed firmly as this prevents the formation of steam channels. The steam is allowed to pass into<br />

the still with a steam pressure of 18–32 kg in the boiler. The mixture of vapors of water and lemongrass<br />

oil passes into the condenser. As the distillation proceeds, the distillate collects in the<br />

separator. The oil, being lighter than water and insoluble, floats on the top of the separator and is<br />

continuously drawn off. It is then decanted and filtered. Small cultivators can use direct-fire stills,<br />

but in such cases, properly designed stills should be used. These stills are provided with a boiler at<br />

the bottom of the tub. This is separated by a false bottom from the rest of the tub. Water is poured<br />

at the bottom of the tub, and grass is charged in the top portion. In the still, the water does not come<br />

in contact with the grass. Providing a perforated disk just above the water level in the copper still<br />

will be helpful in producing oil of better quality. This method is known as water and steam method.<br />

The quality of oil suffers because of the crude method of production. To get a maximum yield of<br />

good-quality oil, it is advisable to use steam distillation.<br />

Thick stems should be removed before distillation as these are devoid of oil. Time required for<br />

one distillation is about 4 h, including the time required for charging and discharging, provided the<br />

firewood is well dried and of good quality. For one distillation, about 40 kg of firewood is required.<br />

A light yellow, lemon-scented volatile oil is obtained. When crop area is large enough, the steam<br />

method is found to be more economical. Coal can also be used as fuel. The oil has a strong lemonlike<br />

odor. The oil is yellowish in color, having 75%–85% citral and small amounts of other minor<br />

aromatic compounds. The oil content recovered from the grass ranges from 0.5%–0.8%.<br />

3.2.3.5 supercritical Fluid extraction (sFe)<br />

SFE of C. citratus yielded oil at par with steam distillation but with more compounds (Sargenti<br />

and Lancas 1997). A comparative analysis of the oil and supercritical CO 2 extract of C. citratus<br />

was done by Marongiu et al. (2006). Dried and ground lemongrass leaves were used as a matrix<br />

for supercritical extraction of essential oil with CO 2. The objective of this study was to analyze<br />

the influence of pressure on the supercritical extraction. A series of experiments were carried out<br />

for 360 min, at 50°C, and at different pressures: 90, 100, 110, and 120 bar. Extraction conditions<br />

were chosen so as to maximize the citral content in the extracted oil. The collected extracts were<br />

analyzed by GC-MS, and their composition was compared with that of the essential oil isolated by<br />

hydrodistillation and steam distillation. At higher solvent densities, the extract’s aspect changes,<br />

passing from a characteristic yellow-colored essential oil to a yellowish semisolid mass, because<br />

of the extraction of high-molecular-mass compounds. The optimum conditions for citral extraction<br />

were found to be 90 bar and 50°C. At these conditions, citral represents more than 68% of the<br />

essential oil and the extraction yield was 0.65%, while the yield obtained from hydrodistillation<br />

was 0.43% with a citral content of 73%. Lemongrass oil was also extracted by the SFE method. The<br />

essential oil yield did not increase as expected, but decreased (Rozzi et al. 2002).<br />

Lemongrass essential oil was extracted with dense carbon dioxide at 23°C–50°C and 85–120 bar.<br />

Liquid carbon dioxide extracts had a larger quantity of coextracted waxes than the supercritical<br />

extracts. The process condition of 120 bar and 40°C was considered ideal for the extraction of


116 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

lemon grass essential oil, as a good-quality product was obtained together with good extraction rate<br />

and yield (Carlson et al. 2001).<br />

3.2.3.6 treatment of oil Prior to storage<br />

<strong>Oil</strong> recovered from each distillation run is added to a settling tank, where it should be left for<br />

1–2 days to allow occluded water to sink to the bottom and be run off. The oil is then filtered to<br />

remove any solid debris and is next transferred to a bulking or storage tank. Bulking is an important<br />

operation since this allows for variation in oil composition between individual distillations and the<br />

supply to a buyer of an oil of reasonably consistent quality. Clean, new containers (steel/aluminum<br />

drums; UN standard) of 200 L capacity are used for shipment. Since the specific gravity of the oil is<br />

around 0.9, each drum will contain approximately 180 kg. For international shipment by land, sea,<br />

or air, drums must be labeled with appropriate identification and hazard codes.<br />

3.2.3.6.1 Purification of <strong>Oil</strong><br />

The insoluble particles present in the oil are removed by the simple filtration method after mixing<br />

it with anhydrous sodium sulfate and keeping it overnight or for 4–5 h. In case the color of the oil<br />

changes because of rusting, then it should be cleaned by the steam rectification process.<br />

3.2.3.6.2 Storage and Packing of <strong>Oil</strong><br />

The oil should be stored in glass bottles or containers made of stainless steel or aluminum or galvanized<br />

iron, depending on the quantity of oil to be stored. The oil should be filled up to the brim,<br />

and the containers should be kept away from direct heat and sunlight in cool or shaded places. The<br />

oil should be stored in well-sealed glass bottles, at 5°C–25°C, and in a dry, well-ventilated area<br />

away from direct heat and sunlight. Lemongrass oil can be stored for up to 3 years without affecting<br />

the quality of oil, if kept in aluminum containers sealed airtight using wax. Containers should be<br />

completely filled to exclude any air and protect the oil from sunlight as air and sunlight affect the<br />

citral content.<br />

In a study conducted at Lucknow, India, the freshly distilled, light-yellow-colored oils of palmarosa<br />

(C. martinii), citronella (C. nardus), and lemongrass (C. flexuosus) were stored separately<br />

in 1 L containers made of amber glass, plain glass, aluminum, iron, and high-density polyethylene<br />

(HDPE). The containers were placed in a dark room at room temperature for up to 26 months. The<br />

oil samples were drawn from the storage containers periodically and analyzed for the color and percentage<br />

composition of the major constituents of the oil. The oil changed color only when stored in<br />

iron containers. There was little, if any, change in the percentage composition of major constituents<br />

in the three oils stored in other containers (Raina et al. 1998). The following points are taken into<br />

consideration while storing oils:<br />

• Care is taken to ensure that the essential oil does not contain any water before storage.<br />

Amber-colored bottles are convenient for small quantities of oil. For large quantities, steel<br />

or aluminum drums are widely used.<br />

• The oils are left to stand for some time so that water can settle down. If the oil is still turbid,<br />

a small amount of sodium sulfate is added and the oil is filtered. The containers are<br />

filled up to the brim, tightly capped, and stored in a cool, dry, and dark place.<br />

• Exposure to air, light, and water causes deterioration of the quality of essential oil.<br />

3.2.4 qu a l i t y an a ly s i s<br />

Identification and estimation of various constituents of essential oils are carried out by gas<br />

chromatography.


The Cymbopogons 117<br />

3.2.4.1 standard specifications<br />

<strong>Oil</strong>s Association (BEOA) in its Chemical Hazard Information and Packaging (CHIP) Regulations<br />

list of 1999 gives the following details:<br />

Hazard symbol l: Xn<br />

Risk phrase : R65<br />

H/C % : 15<br />

Safety phrase : S62<br />

In the EU, all member countries today follow the standards published by the International<br />

Organization for Standardization (ISO 3217-1974). The main physiochemical requirements of this<br />

standard for lemongrass oil are the following:<br />

Relative density at 20°C/20°C : 0.872–0.897<br />

Optical rotation at 20°C : −3º to +1º<br />

Refractive index at 20°C : 1.483–1.489<br />

Carbonyl compounds as citral mininum : 75%<br />

Solubility in ethanol (70% v/v) at 20°C : soluble<br />

In the United States, the Fragrance Manufacturers’ Association (FMA) has published a standard<br />

(CAS # 08007-02-1) with very similar requirements. Both the ISO and FMA standards include<br />

gas chromatography analysis fingerprints for West Indian type lemongrass oil, and this analytical<br />

technique is the first of its kind used on a sample received by a buyer. The older physicochemical<br />

analyzes are used when adulteration or other quality deficiencies are suspected. It is important to<br />

recognize that the published standard specifications are the minimum requirements of buyers and<br />

users. More demanding in-house quality criteria may be set by end users, and these will include<br />

subjectively assessed odor characteristics.<br />

3.3 Palmarosa (Cymbopogon martinii Var. motia (roxb.) Wats.)<br />

Palmarosa (Cymbopogon martinii) is a widely distributed plant in India that yields a sweet, fragrant,<br />

aromatic oil. In India, palmarosa oil is mostly obtained from wild-growing grass in the states<br />

of Madhya Pradesh, Maharashtra, Andhra Pradesh, and Karnataka. A plantation of motia variety<br />

was started in Punjab in 1924. The late Prof. Puran Singh, Chief Chemist, Forest Research Institute<br />

and Colleges, Dehra Dun, succeeded in cultivating the grass at Jaranwala (Lyallpur) over an area of<br />

93 ha in a short period of 4–5 years. He put up a steam distillation plant, and 1350–1600 kg of oil<br />

was produced annually. It was later cultivated near Dehra Dun by Purandad <strong>Essential</strong> <strong>Oil</strong> Plantation<br />

(Industry), and oil of good quality has been produced. Palmarosa is adapted to marginal areas and<br />

poor soils and can be grown under dense canopies of trees and used for soil conservation.<br />

Sahoo (1994) selected cultivars, namely, RRL(b)69, RRL(B)77, IW31245, IW3630, CI8041,<br />

and HR89 for commercial production of high-quality palmarosa oil. Recently, its cultivation has<br />

also been taken up in the states of Karnataka, Maharashtra, Madhya Pradesh, and Uttar Pradesh.<br />

Palmarosa is also flourishing on a red sandy loam soil in the semiarid tropical climate of South<br />

India under rainfed conditions (Rajeswara Rao 2001). The use of farmyard manure (FYM) and<br />

nitrogen fertilizer had a positive impact on biomass and essential oil yield. Palmarosa grass has<br />

also been cultivated as an intercrop with pigeon pea. The oil content and quality, in terms of total<br />

geraniol, of palmarosa were not adversely affected by intercropping (Maheshwari et al. 1995). The<br />

flowering tops and foliage contain a sweet-smelling oil that emits a rose-like odor and is widely used


118 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

in soaps, cosmetics, and perfumery industries. The oil is also used as a raw material for producing<br />

geraniol, which is extensively used in the perfumery industry.<br />

Mainly, oil of palmarosa is obtained from flowering shoots and aboveground parts of the motia<br />

variety of C. martinii. The variety is also referred to as rosha grass or russa grass and yields an<br />

oil of high geraniol content (75%–90%), which is also called East Indian geranium oil or rosha oil.<br />

Another variety, sofia, is also found growing wild in India, which yields an oil of lower geraniol<br />

content known as gingergrass oil. The oil is of inferior grade and fetches much less price than palmarosa<br />

oil. <strong>Oil</strong> of palmarosa is one of the most important essential oils of India that is exported,<br />

and once, India was the principal supplier of this oil to the world. The market for export has fallen<br />

because of the deterioration of the quality of oil, competition with other countries, and appearance<br />

of synthetic geraniol in the market.<br />

<strong>Essential</strong> oil composition of palmarosa grass from different places of India was studied by<br />

Raina et al. (2003). Geraniol (67.6%–83.6%) was the major constituent and, although the composition<br />

of the three oils was similar, quantitative differences in the concentrations of some constituents<br />

were observed.<br />

3.3.1 uses<br />

<strong>Oil</strong> of palmarosa is used in perfumery, particularly for flavoring tobacco and for the blending of<br />

soaps, owing to the lasting rose note it imparts to the blend. In soap perfumes, it has a special importance<br />

because geraniol remains stable in contact with alkali. It also serves as a source for very high<br />

grade geraniol. Geraniol is highly valued as a perfume and as a starting material for a large number<br />

of aromatic chemicals, such as geranyl esters, that have a permanent rose-like odor.<br />

The essential oil is distributed in all parts of the grass, such as flower heads, leaves, and stems,<br />

with the flower heads containing the major portion. Usually, the grass is cut at a height of 5–8 cm<br />

from the ground level, and the whole plant is used for distillation. The maximum yield of oil is<br />

obtained when the entire plant is at the full-flowering stage. The flowering tops of palmarosa consist<br />

of spikelets, and each spikelet is further composed of racemes and a leaf-like structure called<br />

a spathe. Both racemes and spathe contain essential oil. Changes in the essential oil content and<br />

composition during inflorescence development were studied by Dubey et al. (2000) and Dubey<br />

and Luthra (2001). Changes in fresh weight, dry weight, chlorophyll and essential oil content, and its<br />

major constituents, that is, geraniol and geranyl acetate, were examined for both racemes and spathe<br />

at various stages of spikelet development. The essential oil content was maximum at the unopened<br />

spikelets stage and decreased significantly thereafter.<br />

At the unopened spikelets stage, the proportion of geranyl acetate (58.6%) in the raceme oil was<br />

relatively greater compared to geraniol (37.2%), whereas the spathe oil contained more geraniol<br />

(61.9%) compared to geranyl acetate (33.4%). The relative percentage of geranyl acetate in both<br />

the oils, however, decreased significantly with development, and this was accompanied by a corresponding<br />

increase in the percentage of geraniol. Analysis of the volatile constituents from racemes,<br />

spathes (from mature spikelets), and seeds by capillary gas chromatography (GC) indicated<br />

28 minor constituents besides the major constituent geraniol. Harvesting time, stage, and duration<br />

of harvest play a major role in determining the herbage, quality (chemical contents), and productivity<br />

of the herb.<br />

3.3.2 ha r V e s t i n G<br />

The number of harvests depends on the climatic condition of the place of cultivation and the method<br />

of crop management. In the first year, usually one crop is obtained during October–November,<br />

whereas two to three crops are obtained in the subsequent years in subtropical areas of the North<br />

Indian plains. Four harvests are taken in the tropical areas of South and Northeast India. By about<br />

3½ to 4 months, the plants attain a height of 150–200 cm, and they start producing inflorescence.


The Cymbopogons 119<br />

The grass is cut 1 week after flowering. Generally, two cuttings are made during the first year of<br />

planting. From the second year onward, three to five cuttings are possible. It is recommended to harvest<br />

the crop 7–10 days after the opening of flowers. Usually, the grass is cut at a height of 5–8 cm<br />

from the ground level, and the whole plant is used for distillation. The maximum yield of oil is<br />

obtained when the entire plant is at the full-flowering stage. The harvested herbage is spread in the<br />

field for 4–6 h to reduce its moisture by 50%, and such semidry produce can be stacked in shady,<br />

cool spaces for a few days without much loss of its oil.<br />

Palmarosa crop should be harvested at the full-flowering stage to seeding stage to obtain a high<br />

oil yield of good quality. During this period, the aerial parts, that is, the stem, leaf, inflorescence,<br />

and the whole herb, yielded 0.05%, 0.6%, 1.0%, and 0.5% oil with 78.5% to 88.50% geraniol. It was<br />

observed that the essential oil, rich in geraniol, had the least geranyl acetate content (Akhila et al.<br />

1984). The delay or postponement of harvesting affected the yield and quality of oil. The delay led<br />

to an increase in leaf:stem ratio of the crop.<br />

An experiment conducted in Punjab, India, revealed that each delay in harvesting, from 75<br />

days harvesting interval to 125 days harvesting interval, increased the oil yield in the first cutting<br />

(Table 3.4). An oil yield of 100.6 L/ha was produced when the crop was harvested at 125 days’ harvesting<br />

interval as compared to 87.0 and 46.6 L/ha of 100 and 75 days’ harvesting intervals. The oil<br />

content in fresh herb decreased with delay in harvesting from 75 days to 125 days (Table 3.4). The<br />

oil content in the herb also decreased in the second cutting in all harvesting intervals as compared to<br />

the first cutting. Shahidullah et al. (1996) also reported that in palmarosa, essential oil content was<br />

highest when harvested in May compared to November. Kuriakose (1989) reported that maximum<br />

oil content was obtained when the crop was harvested at 90 days’ interval as compared to 50, 60,<br />

70, and 80 days’ interval in palmarosa.<br />

3.3.3 st o r a G e<br />

Hay storage of C. martinii (during summer) either in the shade or in the open increased the essential<br />

oil content. A slight difference in geraniol and geranyl acetate contents of the essential oils of<br />

C. martinii leaves was observed.<br />

3.3.4 yield<br />

table 3.4<br />

effects of different harvesting Intervals on the herbage, oil<br />

content, and oil yield of Palmarosa<br />

cutting<br />

Fresh herb yield (q/ha) oil content (%) oil yield (l/ha)<br />

75 a 100 125 75 a 100 125 75 a 100 125<br />

First 60.3 151.8 254.5 0.78 0.58 0.44 46.6 87.0 100.6<br />

Second 291.1 231.2 166.4 0.51 0.25 0.27 148.4 57.8 45.1<br />

Third 149.5 101.7 — 0.27 0.31 — 88.8 40.5 32.3<br />

CD 5% 67.8 36.1 NS NS 0.18 NS 32.1 20.8 18.9<br />

Note: NS = Not significant.<br />

a Harvesting intervals (in days).<br />

Source: Gill B S et al. 2007. India Perfumer 51: 23–27.<br />

Palmarosa plantation remains productive for about 8 years. However, the yield of grass and oil starts<br />

decreasing from the fourth year onward. It is, therefore, recommended that the plantation be kept<br />

only for 4 to 5 years. Normally, 200–250 q/ha of fresh herbage is obtained in the first cutting, and


120 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

between 250–320 q/ha in second and subsequent harvests up to 3 years under irrigated conditions.<br />

On average, 200 kg of oil are received during the growing period of 15–16 months. The yield of oil<br />

for the first 4 years is as follows:<br />

3.3.5 distill ation<br />

First year: 60 kg/ha<br />

Second year: 80 kg/ha<br />

Third year: 80 kg/ha<br />

Fourth year: 80 kg/ha<br />

<strong>Oil</strong> of palmarosa is generally obtained by steam or hydrodistillation, similar to the process for<br />

lemongrass mentioned earlier. It takes 2 h to complete one distillation. The average recovery of oil<br />

is 0.40%–0.45%. Allowing the cut grass to wilt in the shade for 24 h during premonsoon months<br />

and 48 h during the postmonsoon months increases oil recovery. Small cultivators use direct-fire<br />

stills, but properly designed stills should be used. From the quality point of view, the grass should<br />

be distilled as fresh as possible. <strong>Oil</strong> obtained from dry or fermented grass is of inferior quality. For<br />

economic production of the oil, it is advisable that the harvested material be allowed to dry for a<br />

short period. The distillation unit should be clean, rust free, and free of any other odor.<br />

During steam distillation, a part of the essential oil becomes dissolved in the condensate or<br />

distillation water and is lost when this water is discarded. A method was developed to recover<br />

the dissolved essential oil from condensate water. The distillation water of palmarosa mixed with<br />

hexane in 10:1 proportion was thoroughly shaken for 30 min to trap the dissolved essential oil.<br />

Hexane was then distilled to yield “secondary” or “recovered” oil. In palmarosa, the “primary” or<br />

decanted oil (obtained directly by distilling the crop biomass) accounted for 92%, and the recovered<br />

oil accounted for 8% of the total oil yield. The solvent loss in this process was 4%–7%. Experiments<br />

conducted in the laboratory with the essential oil showed that the water solubility of palmarosa oil<br />

ranged from 0.12% to 0.15% at 31°C and 0.15% to 0.20% at 80°C. Hexane recovered up to 97% of<br />

the dissolved essential oil in water. The recovered essential oil was richer in organoleptically important<br />

oxygenated compounds, such as linalool (2.6%–3.8%), geraniol (91.8%–92.8%), and geranial<br />

(1.8%–2.0%), compared to the primary oil (Rajeswara Rao et al. 2005).<br />

3.3.6 oil st o r a G e<br />

The oil should be free of sediments, suspended matter, and moisture before storage. The container<br />

should be clean and rust free.<br />

3.3.7 oil Co n t e n t a n d yield<br />

The content and yield of the oil depend on many factors, such as climatic conditions of the place of<br />

cultivation; time of harvesting; maturity of the grass; nature of material being distilled, that is, fresh<br />

material or wilted material; method of distillation; etc. All parts of the plant contain the essential<br />

oil, the maximum oil being present in flowers and the stalks containing a negligible quantity. On<br />

average, the oil content in the various parts of the plant is as follows:<br />

Plant Parts essential oil (%)<br />

Whole plant 0.10–0.40<br />

Stalks 0.01–0.03<br />

Flowering tops 0.45–0.52<br />

Leaves 0.16–0.25


The Cymbopogons 121<br />

table 3.5<br />

Indian standard specifications for Palmarosa oil<br />

sl. no. characteristics requirements no.<br />

1. Solubility Soluble in 2 volumes of ethyl alcohol (70% by volume)<br />

2. Color Light yellow to yellow<br />

3. Odor Rosaceous with a characteristic grassy background<br />

4. Specific gravity 0.8740 to 0.8860 at 30°C/30°C<br />

5. Optical rotation −2° to +3°<br />

6. Refractive index 1.4690 to 1.4735 at 30°C<br />

7. Acid value (maximum) 3<br />

8. Ester value 9 to 36<br />

9. Ester value after acetylation 266 to 280<br />

10. Total alcohols, calculated as geraniol<br />

percent (minimum)<br />

90.0<br />

3.3.8 st a n d a r d sPeCifiCations<br />

Indian Standard Specifications for the oil of palmarosa (IS: 526-1986) are given in Table 3.5.<br />

The characteristic features of oil of palmarosa are the following: first, its sweet odor; and second,<br />

its solubility test in 70% alcohol (solubility of oil in 2.2–4.2 volumes of alcohol indicates<br />

a higher percentage of free geraniol). <strong>Oil</strong> of palmarosa chiefly contains 75.0%–95.0% alcohols,<br />

calculated as geraniol, and a small but varying amount of esters of the same alcohol, principally<br />

acetic and caproic acids. Java oils also have almost the same geraniol content, but their ester<br />

content is higher.<br />

3.4 cItronella (Cymbopogon winterianus JoWItt)<br />

Citronella oil is one of the industrially important essential oils obtained from different species of<br />

Cymbopogon. The oil is widely used in perfumery, soaps, detergents, industrial polishes, cleaning<br />

compounds, and other industrial products. It is classified in the trade into two types: Ceylon<br />

citronella oil, which is extracted from C. nardus; and Java citronella oil, which is obtained from<br />

C. winterianus. The major difference between these oils is the proportion of geraniol and citronellal.<br />

The higher proportion of geraniol and citronellal in Java-type oil makes it an important source<br />

of various derivatives such as citronellol and hydroxycitronellal, which are extensively used in compounding<br />

high-grade perfumes. The Ceylon-type citronella oil, which contains a relatively low proportion<br />

of geraniol and citronellal, is mainly used in cheaper products rather than for the extraction<br />

of derivatives (Pino and Corrla 1996). The citronella oil has export potential, and its production can<br />

utilize rural-sector participation (Coronel et al. 1984).<br />

Java citronella (Bordoloi 1982, C. winterianus) is a stoloniferous perennial that may grow up to<br />

1 m high. Young shoots, growing from the axillary leaves of the mother plant, develop into large<br />

clumps with leaves bending outward. Cultivation is normally undertaken in tropical areas up to<br />

an altitude of 600 m, and a well-distributed rainfall of 1500 to 2500 mm is preferable. Citronella<br />

adapts to a wide range of soils, but will not withstand waterlogging. Sandy soils that are fairly<br />

fertile are ideal. Very fertile soils provide a high biomass yield, but a low oil yield. Java citronella<br />

generally flourishes in areas where lemongrass is cultivated and, also, it is less susceptible to<br />

humidity-induced rust. Citronella showed positive correlation with irrigation and nitrogen application<br />

(Singh et al. 1996a). The beneficial response is obtained with moderate nitrogen application,<br />

while a heavy dosage enhances biomass but not oil product yields. The essential oil from C. winterianus<br />

of Cuban origin contains citronellal (25.04%), citronellol (15.69%), and geraniol (16.85%)


122 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

as major constituents. Brazil is classified as one of the major world producers of essential oils and<br />

citronella essential oil.<br />

The oil is used mostly in perfumery, both directly and indirectly. Soaps, soap flakes, detergents,<br />

household cleansers, technical products, insecticides, etc., are often perfumed exclusively<br />

with this oil. It is also a valuable constituent in perfumery for soaps and detergents (Hussain et al.<br />

1988). Citronellal is occasionally used in traces in flower compositions of citrus, cherry, ginger,<br />

etc. However, the greatest importance of citronellal lies in its role as a starting material for further<br />

derivatives. Hydroxycitronellal can be prepared from citronellal, and it is a key ingredient in compounding.<br />

Hydroxycitronellal is one of the most frequently used floralizing perfume materials. It<br />

finds its way into almost every type of floral fragrance and a great many nonfloral ones. For soap<br />

perfumes, a slightly rougher grade is used. High grade is used in flavor compositions. Java citronella<br />

is rich in geraniol (36.0%) and citronellal (42.7%) and shows repellent, antimycotic, and acaricide<br />

activities. It is reported to be an air freshener (Guenther 1992; Lawrence 1996).<br />

Cultivation of Java citronella was first introduced in the northeastern region of India by the<br />

Regional Research Laboratory (RRL), Jorhat, in 1965. RRL, Jammu, introduced a C. winterianus<br />

clone that yielded 60–70 t herbage per hectare, 0.6%–0.8% oil, and 60%–70% citronellal, which is<br />

extensively cultivated in the northeastern region (Sobti et al. 1982; Lal et al. 1999).<br />

The biosynthesis of secondary metabolites, although controlled genetically, is strongly affected<br />

by environmental, harvest, and postharvest factors. Agricultural factors have a critical effect on the<br />

quantitative and qualitative characteristics of Java citronella, which affect plant growth and yield.<br />

Precipitation, temperature, light, and humidity influence volatile oil yield and the main constituents’<br />

content (citronellol, citronellal, and geraniol) (Sarma 2002). Sarma et al. (2001) cultivated Java<br />

citronella between 1998 and 2000 in Northeast India. They found a large variation in volatile oil<br />

yield and content, depending on seasonal changes and harvesting time. Highest yield and citronellal<br />

content were obtained during September and October. Under drought conditions, the geraniol<br />

and citronellal contents reduced, whereas citronellol content increased (Fatima et al. 2002). Saikia<br />

et al. (2006) studied the variation of essential oil content in leaf and inflorescence of Java citronella.<br />

Maximum essential oil accumulation was recorded at a crop age of 69 days during the summer<br />

(April–June), when the crop experienced the highest air temperature and the lowest relative humidity.<br />

However, the lowest air temperature and moderate relative humidity experienced by the crop<br />

during winter (October–January) provided good conditions for citronella oil quality (accumulation<br />

of citronellal, citronellol, and geraniol), which was best at a crop age of 63 days (Singh et al. 1996a).<br />

Java citronella cultivated at an altitude of 60 m showed elevated biomass production (24.5 t/ha)<br />

and volatile oil content (1.3%). An increase in citronellal content and a decrease in citronellol and<br />

geraniol content were observed at this altitude (Sarma et al. 2001). Misra and Srivastava (1994)<br />

studied the influence of iron nutrition on chlorophyll contents, photosynthesis, and essential monoterpene<br />

oils in Java citronella. Significant positive correlations were observed between herbage,<br />

total essential oil yield, and citronellol content. Yellowing and crinkling disease influenced the<br />

essential oil yield and composition of citronella oil from sea level. The disease decreased biomass<br />

yield in the first and second years of harvesting by 62.80 and 82.70, respectively. The corresponding<br />

decreases in essential oil yield per plant were 62.80 and 79.00 (Rajeswara et al. 2004). However,<br />

only few studies have been conducted on the effects of harvest time and drying throughout the different<br />

harvest seasons in the northeast region of Brazil (Arrigoni-Blank et al. 2005; Carvalho-Filho<br />

et al. 2006).<br />

3.4.1 ha r V e s t i n G<br />

The time of harvesting affects the yield and quality of the oil. The first harvest is generally obtained<br />

after 4–6 months of transplanting. Subsequent harvests take place at intervals of 50–60 days,<br />

depending on the fertility of the soil and seasonal factors. Under normal conditions, two to three<br />

harvests are possible during the first year and three to four in subsequent years, depending on the


The Cymbopogons 123<br />

table 3.6<br />

effects of different harvesting Intervals on Fresh herb<br />

yield, oil content, and oil yield of citronella<br />

cutting<br />

Fresh herb yield (q/ha) oil content (%) oil yield (1/ha)<br />

75 a 100 125 75 a 100 125 75 a 100 125<br />

First 26.0 29.7 33.0 1.84 1.54 0.84 46.0 44.6 28.5<br />

Second 63.7 89.0 90.1 1.45 1.12 1.05 92.3 99.3 94.8<br />

Third 50.5 10.8 — 1.19 1.18 — 61.4 12.7 —<br />

CD 5% 22.9 16.8 8.6 NS NS NS NS 15.5 NS<br />

Note: NS = Not significant.<br />

a Harvesting intervals (in days).<br />

Source: Gill B S et al. 2007. India Perfumer 51: 23–27.<br />

management practices followed. Harvesting is done with the help of sickles, the plants being cut<br />

close to their bases about 10 cm above ground level. At harvest time, the grass is usually around 1 m<br />

high. The cut should be made above the first node in order to avoid the risk of dieback. Harvesting<br />

should be undertaken before flowering of the crop as it reduces the oil yields. Some researchers<br />

advocate harvesting when the stem bears six adult leaves and the seventh is rolled up. Others recommend<br />

cutting when the leaf tips begin to dry. The maximum yield is obtained in the third year, and<br />

after the fifth year the yields diminish rapidly. Mechanical harvesting is possible, but more commonly<br />

the grass is cut by hand. A dry day is preferable for the operation. After cutting, the grass<br />

should be left to wilt for a day, but care must be exercised to prevent fermentation. The oil obtained<br />

from this partially dry grass is more fragrant and of better quality than that obtained from grass<br />

distilled immediately after cutting.<br />

Studies were conducted at Punjab, India, to standardize the harvest intervals for citronella grass<br />

(Gill et al. 2007). The first cutting yielded 46.0, 44.6, and 28.5 L/ha of citronella oil at 75, 100, and<br />

125 days’ harvesting intervals (Table 3.6). Citronella should be harvested at 75-day interval for<br />

optimum yield. Shahidullah et al. (1996) reported that in citronella, essential oil content was higher<br />

when harvested in May compared to November. Weiss (1997) also reported that oil content of leaves<br />

differs significantly, and young leaves synthesize and accumulate most of the essential oil in citronella.<br />

Virmani et al. (1979) reported that in Java citronella, delay in harvesting causes the leaves to<br />

dry up, decreasing the oil yield.<br />

The grass is cut at all times of the year, but the yield varies with season; it is highest during the<br />

hot period, and low during the wet season and flowering period. The highest fresh biomass yield<br />

was obtained during summer (9326 kg/ha), fall (8174 kg/ha), and spring (8352 kg/ha) (Table 3.7),<br />

and the lowest yield during winter (3788 kg/ha). Seasons had significant effects on dry herbage biomass.<br />

Higher dry biomass production was achieved during fall and summer (5363 and 4897 kg/ha,<br />

respectively). A lower dry biomass production was obtained during winter and spring (1625 and<br />

3189 kg/ha, respectively); however, proportionally higher moisture was observed (61.82% and<br />

57.11%, respectively; Blank et al. 2007). Higher concentrations of fresh and dry biomass also were<br />

observed during dry seasons with irrigation.<br />

Another important factor influencing Java citronella volatile oil production is harvesting time.<br />

The results obtained are presented in Table 3.8. Most volatile oils were produced at 0900h (2.71%,<br />

2.22%, 2.36%, and 4.24% for summer, fall, winter, and spring seasons, respectively), because volatile<br />

oil content was generally found to peak at that time, except during fall, when there were no<br />

significant differences between harvest times. The percentage content of volatile oil in dried herbage<br />

was not significantly influenced by harvesting time. However, higher contents of volatile oil<br />

were obtained during spring (4.24%, fresh biomass; and 3.40%, dry biomass) and fall (3.17%, dry


124 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 3.7<br />

effects of seasons on Java citronella (Cymbopogon winterianus Jowitt)<br />

Fresh and dry biomass and moisture a<br />

season Fresh biomass yield (kg/ha) dry biomass yield (kg/ha) moisture content (%)<br />

Summer 9326 a 4897 a 34.4<br />

Fall 8174 a 5363 a 47.5<br />

Winter 3788 b 1625 c 61.8<br />

Spring 8352 a 3189 b 57.1<br />

a Biomass obtained in different seasons is not significantly different (P < 0.05).<br />

b About the same biomass.<br />

c Lowest biomass.<br />

Source: Blank Arie F et al. 2007. Brazilian Journal of Pharmacognosy 17(4): 557–564.<br />

table 3.8<br />

Influence of seasons and harvest times on Volatile<br />

oil content (In Percentage) From Fresh and dry leaves<br />

of C. winterianus a<br />

season<br />

oil content (%) oil yield (l/ha)<br />

0900 h 1200 h 1500 h 0900 h 1200 h 1500 h<br />

Fresh leaves<br />

Summer 2.71 b A 2.35 b AB 1.98 a B 162.50 a 114.82 a 97.17 a<br />

Fall 2.22 b A 1.84 b A 1.71 a A 119.10 a 91.47 a 98.46 a<br />

Winter 2.36 b A 1.98 b AB 1.67 a B 38.39 b 32.13 b 27.13 b<br />

Spring 4.24 a A 3.36 a B 2.43 a C 135.29 a 107.25 a 77.42 a<br />

dry leaves<br />

Summer 2.60 b A 2.47 a A 3.10 a A 127.32 b 120.79 b 151.80 a<br />

Fall 3.17 ab A 3.00 a A 3.20 a A 169.82 a 160.88 a 171.61 a<br />

Winter 2.57 b A 2.67 a A 2.33 b A 41.70 c 43.32 c 37.91 c<br />

Spring 3.40 a A 3.20 a A 3.50 a A 108.43 b 102.05 b 111.62 b<br />

a Means of oil content in same season followed by the same lowercase letter in each<br />

column and by the same uppercase letter in each line are not significantly different.<br />

Source: Blank Arie F et al. 2007. Brazilian Journal of Pharmacognosy 17(4): 557–564.<br />

biomass). Harvest at 1200h resulted in quite similar volatile oil content throughout all seasons. In<br />

general, 1500h harvest significantly reduced volatile oil content, showing the lowest value during<br />

winter season (2.33%, dry biomass), which is characterized by higher precipitation volumes<br />

(Table 3.8). In accordance with studies on C. citratus conducted by Nascimento et al. (2003),<br />

harvests performed at 0900h and 1100h provided higher volatile oil contents (5.59% and 5.31%,<br />

respectively). The chemotype citral/limonene of Lippia alba (Mill) N.E.Br. produces lower volatile<br />

oil content during the rainy season, which accounts for a potential metabolic decrease caused by<br />

reduction of solar radiation (Nagao et al. 2004). In general, winter harvest from fresh or dry leaves<br />

significantly reduced yield independently of harvest time (Table 3.8).<br />

Drying significantly increased volatile oil content and yield; this is probably due to the drying<br />

process affecting cell membrane resistance, assisting volatile oil release during hydrodistillation.


The Cymbopogons 125<br />

However, other species, such as Ocimum basilicum L., did not show significant differences between<br />

fresh and dry biomass (Carvalho-Filho et al. 2006), and Lippia alba showed peak volatile oil content<br />

at 1500h in both rainy and dry seasons. Specifically during the rainy season, volatile oil content<br />

increased throughout midday, decreasing by 1700h. Daily variations in temperature and luminosity<br />

may account for these results. The major components of C. winterianus volatile oil were identified<br />

as limonene, citronellal, citronellol, neral, geraniol, geranial, and farnesol. The GC-MS results<br />

showed that geraniol is an important volatile constituent of the volatile oil of C. winterianus, as<br />

well as citronellal. The content of these two compounds in fresh biomass varied with the season.<br />

However, no significant variation was observed for dry biomass.<br />

In the rainy period, the citronellal in fresh biomass reached the lowest values at 0900h and<br />

1200h, while the geraniol content increased. However, in the dry period (summer), the content of<br />

citronellal increased to 30.54% at 1500h, while the content of geraniol decreased to its lowest value<br />

(21.78%). Different results were observed by Rocha et al. (2002): maximum production of citronellal<br />

was between 0900h and 1100h. The least limonene content (1.17%) was observed in fresh leaves,<br />

during fall at 1200h. However, the highest limonene yield and content was achieved in plants harvested<br />

during winter. The content of citronellol in fresh biomass significantly decreased at 1500h<br />

during winter. The content of citronellol in dry biomass exhibited no significant variation.<br />

Drying provided an increase in citronellol during summer at 1200h, while in the rainy period<br />

a reduction was observed at 0900h and 1200h. In general, no significant differences in the content<br />

of neral were found in fresh and dry biomass throughout the harvest time and seasons, except for<br />

winter at 0900h when neral was absent. Higher content of neral was obtained by drying biomass<br />

harvested at 0900h during winter. Similarly, the content of neral increased after drying Melissa<br />

officinalis L. for 5 days at 40°C (Blank et al. 2007). Amounts of geranial in fresh and dry biomass<br />

were relatively low and not significantly different, except for dry herbage harvested at 1200h during<br />

the fall. Peak farnesol content was achieved only on fresh herbage during winter at 0900h,<br />

which decreased after drying. After drying, herbage harvested at 0900h and 1200h during summer<br />

showed the highest farnesol content (5.69% and 6.13%, respectively).<br />

According to Sarma (2002), Java citronella showed better results under high precipitation<br />

(100–200 mm), temperatures between 20°C and 30°C, and high moisture. In India, citronellal is<br />

obtained in a higher concentration during September and October, periods with favorable conditions<br />

locally. Chemical composition is affected in different seasons and length of drying. Time of<br />

harvest had little influence on composition. Java citronella showed lower volatile oil yield during<br />

winter. Morning harvests provide higher yield.<br />

Variations in citronella oil and its major constituents due to seasonal changes and stages of the<br />

crop were studied under the climatic conditions of Jorhat in Assam, India. Rainfall, temperature,<br />

sunshine, and relative humidity have a cumulative effect on oil yield and major constituents of<br />

the oil, namely, citronellal, citronellol, and geraniol. Postmonsoon months seemed to be favorable,<br />

contributing higher oil yield. Citronellal content was higher during September (44.3%) and October<br />

(45.7%). It was observed that light rainfall (100–200 mm), moderate temperature (20°C–30°C),<br />

sunshine for 5 to 6 h, and high humidity (90%–95%) were the favorable meteorological parameters<br />

for higher oil yields and citronellal content in citronella oil. The growing period or crop growth<br />

stage also had a profound effect on oil yield and citronella content. Older crops with highly matured<br />

leaves were found to yield higher oil and less citronellal. Alcohol content in citronella oil was not<br />

affected by seasonal variation. However, the total alcohol percentage was found to decrease, while<br />

aldehyde percentage was found to increase (Sarma 2002).<br />

A study was conducted in Brazil to optimize essential oil extraction processes from twigs and<br />

leaves by steam distillation. The process variables evaluated in this study were extraction time<br />

and raw material state (dry or natural). The essential oil compositions under different conditions<br />

were obtained with the objective of analyzing the influence of the variable value on the composition<br />

of citronella oil. The maximum yield, 0.942%, was obtained under the following conditions: extraction<br />

time—0400h and state—natural plant, and the results obtained from the factorial experimental


126 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

planning indicate that the variable that influences the essential oil yield more is the state. The principal<br />

compounds identified in the citronella essential oil were citronellal, citronellol, geraniol, geranyl<br />

acetate, and α-cadinol.<br />

3.4.2 st o r a G e o f C. w i n t e r i a n u s hay<br />

Hay storage of C. winterianus (during rainy season), either in the shade or in the open, increased<br />

the essential oil content of the leaves. Slight differences in the citronellal and citronellol content<br />

of the essential oils of C. winterianus leaves were observed.<br />

3.4.3 yield<br />

Depending on soil and climatic conditions, a citronella plantation lasts, on average, for 6 years.<br />

The yield of oil is less during the first year. It increases in the second year and reaches a<br />

maximum in the third and fourth years, after which it declines. For economy, the plantation is<br />

maintained only for 6 years. Up to 50 t/ha of fresh material is possible annually, but on average,<br />

25–30 t of fresh herbage is harvested per hectare per annum from four to five cuttings,<br />

which yield about 80 kg of oil. The oil content of the freshly cut material ranges from about<br />

0.5%–1.0%, and it is enhanced as the grass dries out through natural wilting. The annual oil<br />

yield under very favorable conditions can exceed 200 kg/ha, but 150 kg/ha is regarded as good<br />

in most situations.<br />

3.4.4 distill ation Pr o C e d u r e<br />

The same distillation units may be used for extraction of citronella essential oil as well as lemongrass<br />

oil. The grass is steam-distilled for better recovery of oil and for economical purposes. The<br />

harvested grass sometimes contains dead leaves, which should be removed. The remaining leaves<br />

are cut into shorter lengths. This reduces the volume of the grass, and facilities firm and even packing<br />

within the still. Further, chopping the grass gives a higher yield of oil than with uncut grass.<br />

Generally, distillation is complete within 2½ to 3 h under normal pressure starting from the initial<br />

condensation of the oil. About 80% of the total oil yield is recovered in the first hour, 19% in the second<br />

hour, and about 1% in the third hour of distillation. Larger percentages of the major components<br />

of the total oil, such as citronellal, geraniol, citronellol, and geranyl acetate, are recovered in the first<br />

hour of distillation. Steam distillation of citronella gave a maximum yield of 0.85% essential oil at<br />

20 psig saturated steam pressure, within 30–120 min distilling time, and at 42% moisture content.<br />

The moisture content in citronella has some influence on the amount of essential oil and steam utilized,<br />

while the quality of the oil is determined by the distilling time.<br />

Growers cultivating smaller areas can make use of properly designed direct-fired stills, in case<br />

they are not able to purchase a boiler. In such cases, the lower portion of the distillation tub is filled<br />

with water, and this functions as a boiler. The water in the boiler is separated from the remaining<br />

part of the still by means of a false perforated bottom on which the grass rests. In the still, the water<br />

does not come in contact with the grass. The tub is heated from below either by wood or coal, and<br />

the steam thus produced passes through the grass placed above in the tub, carrying oil vapors with<br />

it. However, distillation in such a direct-fired still takes a little more time, and the quality of the oil<br />

is also inferior.<br />

3.4.5 st a n d a r d sPeCifiCations<br />

In the EU, buyers use the ISO standard for Java citronella oil (ISO 3848-1976), and its main physicochemical<br />

requirements are summarized as follows:


The Cymbopogons 127<br />

Relative density at 20°C/20°C 0.880–0.895<br />

Refractive index 1.466–1.473<br />

Optical rotation −5° to 0°<br />

Carbonyl value Minimum 127, equivalent to 35% as citronellal<br />

Solubility in 80% ethanol at 20°C 1 in 2<br />

Ester value after acetylation Minimum 250, equivalent to 85% expressed as geraniol<br />

Source: Bureau of Product Standards, Department of Trade and Industry.<br />

In the United States, the FMA has published a standard (FMA #2308) with requirements very<br />

similar to those of ISO. Both the ISO and FMA standards include gas chromatography analysis fingerprints<br />

for Java citronella oil, and this analytical technique is the first one performed on a sample<br />

received by a buyer. The older physicochemical analyzes are used when adulteration or other quality<br />

deficiencies are suspected. It is important to recognize that the published standard specifications are<br />

the minimum requirements of buyers and users. More demanding in-house quality criteria may be<br />

set by end users, and these will include subjectively assessed odor characteristics.<br />

3.5 Jamrosa (Cymbopogon nardus rendle)<br />

Jamrosa, a hybrid between palmarosa and citronella, provides an essential oil that is used to impart<br />

a rosy-grassy note to natural perfumes. It is a drought-resistant hardy grass that attains a height<br />

of 1.5–2.5 m and has a hairy and fibrous shallow root system with long linear lanceolate leaves. It<br />

produces large, fawn-colored inflorescence with white, hairy, star-like spiked flowers. It is cultivated<br />

in the Indian states of Uttar Pradesh, Chhattisgarh, Madhya Pradesh, Rajasthan, Karnataka,<br />

Maharashtra, and Tamil Nadu (Shahi and Tava 1993). A well-drained sandy loam soil, free from<br />

waterlogging, with a soil pH of 7.5–8.5 is ideal for cultivation. A warm tropical climate and up<br />

to 300 m elevations in the foothills are suitable for cultivation of jamrosa. Temperatures ranging<br />

from 10°C–36°C with annual rainfall around 1000–1500 mm and ample sunshine are congenial for<br />

its growth. A moist and warm climate throughout the year accelerates its growth. Areas that are<br />

affected by severe frost are not suitable, as the frost kills the grass and reduces the oil content.<br />

3.5.1 uses<br />

<strong>Oil</strong> of jamrosa is used in perfumery, particularly for flavoring tobacco and for blending soaps,<br />

because of the lasting rose note it imparts to the blend. It also serves as a source of very high-grade<br />

geraniol. Geraniol is highly valued as a perfume and as a starting material for large chemicals, such<br />

as geranyl esters, that have a permanent rose-like odor.<br />

3.5.2 ha r V e s t i n G<br />

The essential oil is distributed in all parts of the grass, that is, flower heads, leaves, and stems, with<br />

the flower heads having the major content. It is recommended to harvest the crop 7–10 days after<br />

the opening of flowers. The number of harvests depends on the climatic condition of the place of<br />

cultivation and method of crop management. During the first year, usually one crop is obtained<br />

during October–November, whereas two to three crops are obtained in the subsequent years in the<br />

subtropical areas of the North Indian Plains. Four harvests are taken in the tropical areas of South<br />

and Northeast India. Usually, the grass is cut at a height of 5–8 cm from the ground level, and the<br />

whole plant is used for distillation. The maximum yield of oil is obtained when the entire plant is at<br />

the full-flowering stage. The harvested herbage is spread in the field for 4–6 h to reduce its moisture


128 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

by 50%, and this semidry produce can be stacked in shady, cool spaces for a few days without much<br />

loss of its oil.<br />

The effect of seasonal changes on oil content of a new clone of jamrosa, RL-931 (C. nardus var.<br />

confertiflorus × C. jwarancusa), was investigated by Bhan et al. (2003). The time of harvesting was<br />

found to be very crucial for the productivity of oil per unit area. Biosynthesis of oil coupled with its<br />

main constituents is directly associated with the timing of harvest. RL-931 showed a good deal of<br />

variation in oil content from the first harvest to the fourth harvest of the season. The premonsoon<br />

and onset of the monsoon period (May–July) was characterized by higher oil content (1.0%), whereas<br />

the postmonsoon and winter period (August–December) showed comparatively lower oil content<br />

(0.7%). These findings were in accordance with the results obtained for lemongrass (Handique et al.<br />

1984b). The quality of the oil is best in May–July. High temperature and low humidity favor the<br />

accumulation of geraniol, geranyl acetate, and a low citral content in the oil. Maximum oil content<br />

was recorded when both maximum and minimum temperatures were high.<br />

Central India is more suitable for growing aromatic grasses because of the large area under<br />

degraded forest lands. Many commercial houses in Chhattisgarh have entered into agreements<br />

to buy back aromatic plant produce, and the state government has announced Chhattisgarh as an<br />

herbal state because it has the climatic conditions most suited to aromatic plant production. To promote<br />

aromatic plant production, the chief emphasis is to be given to the establishment of processing<br />

units for oil extraction. Leaf collection will generate about 1 million man-days, and this will provide<br />

enormous employment opportunities. Patchouli and jamrosa keep yielding for 2 years with minimal<br />

expenditure in the second year, which will further enhance the income of farmers.<br />

3.5.3 distill ation<br />

Jamrosa grass oil can be obtained by the method mentioned previously for lemongrass.<br />

3.5.4 yield<br />

Jamrosa plantations remain productive for about 8 years. However, the yield of grass and oil starts<br />

decreasing from the fourth year onward. It is, therefore, recommended that the plantation be kept<br />

only for 4 years. Normally, 200–250 q/ha of fresh herbage is obtained in the first cutting and between<br />

250 and 320 q/ha in the second and subsequent harvests for up to 3 years under irrigated conditions.<br />

On average, 200 kg of oil is received during the growing period of 15–16 months.<br />

The yield of oil (in kg/ha) for the first 4 years is as follows:<br />

3.6 conclusIon<br />

First year 60<br />

Second year 80<br />

Third year 80<br />

Fourth year 80<br />

India is considered to be one of the leading producers of essential oils. However, the oil distilled<br />

by traditional raw methods does not fetch a good value in the international market. Improvements<br />

in extraction technology and the use of oil as a raw material can change this scene. To increase<br />

India’s participation in the international market, it is necessary to improve extraction technology in<br />

order to obtain products having international quality standards. Research on improving distillation<br />

technology is required to increase oil quality and distillation efficiency. Intermediate solutions to<br />

save energy using existing equipment, such as methods to insulate the stainless steel drums, should<br />

be explored.


The Cymbopogons 129<br />

reFerences<br />

Adams R P. 1970. Seasonal variation of terpenoid constituents in natural population of Juniperus pinchotii<br />

Sudw. Phytochemistry 9: 397–402.<br />

Akhila A, Tyagi B R, Naqvi A. 1984. Variation of essential oil constituents in Cymbopogon martinii Wats. var.<br />

motia at different stages of plant growth. Indian Perfumer 28: 126–128.<br />

Anonymous. Handbook of Medicinal and Aromatic Plants, Citronella (Cymbopogon winterianus Jowitt).<br />

NEDFI, pp. 13–18.<br />

Arrigoni-Blank M F, Silva-Mann R, Campos D A, Silva P A, Antoniolli A R, Caetano L C, Sant’Ana A E G,<br />

Blank A F. 2005. Morphological, agronomical and pharmacological characterization of Hyptis pectinata<br />

(L.) Poit germplasm. Revista Brasileira Farmacognosia 15: 298–303.<br />

Beauchamp P S, Dev V, Docter D R, Ehsani R, Vita G, Melkani A B, Mathela C S, Bottini A T. 1996.<br />

Comparative investigation of the sesquiterpenoids present in the leaf oil of Cymbopogon distans (Steud.)<br />

Wats. var. loharkhet and the root oil of Cymbopogon jwarancusa (Jones) Schult. Journal of <strong>Essential</strong> <strong>Oil</strong><br />

Research 8(2): 117–121.<br />

Beech D F. 1997. Growth and oil production of lemongrass (Cymbopogon citratus) in the Ord Irrigation Area,<br />

Western Australia. Australian Journal of Experimental Agriculture and Animal Husbandry 17(85):<br />

301–307.<br />

Beech D F. 1990. The effect of carrier and rate of nitrogen application on the growth and oil production of<br />

lemongrass (Cymbopogon citratus) in the Ord Irrigation Area, Western Australia. Australian Journal of<br />

Experimental Agriculture 30(2): 243–250.<br />

Bhan M K, Rekha Kanti, Kak S N, Agarwal S G, Dhar P L, Thappa R K. 2003. Variation of oil content in a new<br />

clone of Jamrosa ‘RL-931’ (Cymbopogon nardus var. confertiflorus × C. jwarancusa) during one year of<br />

crop growth. New Zealand Journal of Crop and Horticultural Science 31: 187–191.<br />

Blank Arie F, Costa Andressa G, Arrigoni-Blank Maria de Fátima, Cavalcanti Sócrates C H, Alves Péricles B,<br />

Innecco Renato, Ehlert Polyana A D, de Sousa Inajá Francisco. 2007. Influence of season, harvest<br />

time and drying on Java citronella (Cymbopogon winterianus Jowitt) volatile oil. Brazilian Journal of<br />

Pharmacognosy 17(4): 557–564.<br />

Bordoloi D N. 1982. Citronella oil industry in North East India. In Cultivation and Utilization of Aromatic<br />

Plants, Regional Research Laboratory, Jammu, Tawi, India, pp. 318–324.<br />

Boruah P, Misra B P, Pathak M G, Ghosh A C. 1995. Dynamics of essential oil of Cymbopogon citratus (DC)<br />

Stapf. under rust disease indices. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 7(3): 337, 338.<br />

Buggle V, Ming L C, Furtado E L, Rocha S F R, Marques M O M. 1997. Influence of different drying<br />

temperatures on the amount of essential oils and citral content in Cymbopogon citratus (DC)<br />

Stapf., Proceedings of the Second World Congress on Medicinal and Aromatic Plants, WOCMAP-2.<br />

Biological resources, sustainable use, conservation and ethnobotany, Mendoza, Argentina, November<br />

10–15, 1997 [edited by Caffini N, Bernath J, Craker L, Jatisatienr A, Giberti G]. Acta Horticulturae,<br />

1999, No. 571–574.<br />

Carlson Luiz H C, Machado R A F, Spricigo C B, Pereira L K, Bolzan A. 2001. Extraction of lemongrass essential<br />

oil with dense carbon dioxide. Journal of Supercritical Fluids 21(1): 33–39.<br />

Carvalho-Filho J L S, Blank A F, Alves P B, Ehlert P A D, Melo A S, Cavalcanti S C H, Arrigoni-Blank M F,<br />

Silva-Mann R. 2006. Influence of the harvesting time, temperature and drying period on basil (Ocimum<br />

basilicum L.) essential oil. Revista Brasileira Farmacognosia 16: 24–30.<br />

Cassel E, Vergas Rubem M F. 2006. Experiments and modelling of the Cymbopogon winterianus essential oil<br />

extraction by steam distillation. Journal of the Mexican Chemical Society 50(3): 126–129.<br />

Chisowa, E H. 1997. Chemical composition of flower and leaf oils of Cymbopogon densiflorus Stapf. from<br />

Zambia. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 9(4): 469, 470.<br />

Chisowa E H, Hall D R, Farman D I. 1998. Flavour and Fragrance Journal 13: 29, 30.<br />

Chowdhury J U, Mohammed Yusuf, Jaripa Begum, Mondello L, Previti P, Dugo G. 1998. Studies on the essential<br />

oil bearing plants of Bangladesh. Part IV. Composition of the leaf oils of three Cymbopogon species:<br />

C. flexuosus (Nees ex Steud.) Wats., C. nardus (L.) Rendle var. confertiflorus (Steud.) N. L. Bor and<br />

C. martinii (Roxb.) Wats. var. martinii. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 10(3): 301–306.<br />

Choudhury S N, Leclercq P A. 1995. <strong>Essential</strong> oil of Cymbopogon khasianus (Munro ex Hack.) Bor from<br />

northeastern India. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 7(5): 555, 556.<br />

Clark R J, Menary R C. 1980. Environmental effects on peppermint (Mentha piperita L.). I. Effect of day<br />

length, photon flux density, night temperature and day temperature on the yield and composition of peppermint<br />

oil. Australian Journal of Plant Physiology 7: 685–692.


130 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Coronel V Q, Anzaldo F E, Recaña M P. 1984. Effect of moisture content on the essential oil yield of lemongrass<br />

and citronella. NSTA Technology Journal, Vol. IX No. 3.<br />

Costa L C do B, Correa R M, Cardoso J C W, Pinto J E B P, Bertolucci S K V, Ferri P H. 2005. Yield and composition<br />

of essential oil of lemongrass under different drying and fragmentation conditions. Horticultura<br />

Brasileira 23(4): 956–959.<br />

Dhar A K, Dhar R S, Rekha K, Koul S. 1996. Effect of spacings and nitrogen levels on herb and oil yield, oil<br />

concentration and composition in three selections of Cymbopogon jwarancusa (Jones) Schultz. Journal<br />

of Spices and Aromatic Crops 5(2): 120–126.<br />

Dikshit A, Husain A. 1984. Antifungal action of some essential oils against animal pathogens, Fitoterapia<br />

55(3): 171–176.<br />

Dubey V S, Mallavarapu G R, Luthra R. 2000. Changes in the essential oil content and its composition during<br />

palmarosa (Cymbopogon martinii (Roxb.) Wats. var. motia) inflorescence development. Flavour and<br />

Fragrance Journal 15(5): 309–314.<br />

Dubey V S, Luthra R. 2001. Biotransformation of geranyl acetate to geraniol during palmarosa (Cymbopogon<br />

martinii, Roxb. Wats. var. motia) inflorescence development, Phytochemistry 57(5): 675–680.<br />

Dutta S C. 1982. Cultivation of Cymbopogon winterianus Jowitt for production of citronella (Java) oil. In Cultivation<br />

and Utilization of Aromatic Plants, Regional Research Laboratory, Jammu, Tawi, India, pp. 325–329.<br />

Fatima S, Farooqi A H A, Sharma S. 2002. Physiological and metabolic responses of different genotypes of<br />

Cymbopogon martinii and C. winterianus to water stress. Plant Growth Regulation 37: 143–149.<br />

Figueiredo A C, Barroso J G, Pedro L G, Scheffer J J C. 2008. Factors affecting secondary metabolite production<br />

in plants: volatile components and essential oils. Flavour and Fragrance Journal 23(4): 213–226.<br />

Gill B S, Salaria A, Kaur Satvinder, Gill G S. 2007. Harvesting management studies in lemongrass, palmarosa<br />

and citronella, India Perfumer 51: 23–27.<br />

Guenther E. 1950. <strong>Oil</strong>s of citronella, essential of citronella, aceite essential citronella, citronellol, oleum citronellac.<br />

In The <strong>Essential</strong> <strong>Oil</strong>s, Vol. IV. pp. 65–82. Van Nostrand, London.<br />

Guenther E. 1961. The <strong>Essential</strong> <strong>Oil</strong>s, Vol. IV. D. Van Nostrand, New York.<br />

Guenther E. 1992. The <strong>Essential</strong> <strong>Oil</strong>s. D. Van Nostrand, Princeton, New Jersey.<br />

Handique A K, Gupta R K, Bordoloi D N. 1984a. Variation of oil content in lemongrass and its genetic aspects.<br />

Indian Perfumer 24: 13–16.<br />

Handique A K, Gupta R K, Bordoloi D N. 1984b. Variation of oil content in lemongrass as influenced by seasonal<br />

changes and its genetic aspects. Indian Perfumer 28: 54–63.<br />

Hussain A, Virmoni O P, Sharma A, Mishra L N. 1988. Citronella oil (Java) Cymbopogon winterianus Jowitt.<br />

In Major <strong>Essential</strong> <strong>Oil</strong> <strong>Bearing</strong> Plants of India, pp. 52–57.<br />

Hussain A. 1994. <strong>Essential</strong> <strong>Oil</strong> Plants and Their Cultivation. Central Institute of Medicinal and Aromatic Plants,<br />

Lucknow, India.<br />

Jha B K, Kumar D, Joshi P P. 2004. Harvesting frequency in lemongrass var. OD-19 under semi-arid ecosystem<br />

of Gujarat. Indian Perfumer 48(3): 317–321.<br />

Kanjilal P B, Pathak M G, Singh R S, Ghosh A C. 1995. Volatile constituents of the essential oil of Cymbopogon<br />

caesius (Nees ex Hook. et Arn.) Stapf. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 7(4): 437–439.<br />

Kiran G, Babu, D, Kaul V K. 2004. Variation in essential oil composition of rose-scented geranium (Pelargonium<br />

sp.) distilled by different distillation techniques. Flavour and Fragrance Journal 20(2): 222–231.<br />

Kuriakose K P. 1989. Stage of harvest trial in palmarosa. Indian Perfumer 33(1): 21, 22.<br />

Lal R K, Sharma J R, Misra H O. 1999. Development of new variety Manjari of citronella Java (Cymbopogon<br />

winterianus). Journal of Medicinal and Aromatic Plant Sciences 21(3): 727–729.<br />

Lawrence B M. 1996. Progress in essential oils. Perfumer and Flavorist 21(1): 37–45.<br />

Leal T C A B, Freitas S P, Silva J F, Carvalho A J C. 2001. Evaluation of the effect of season variation and<br />

harvest time on the foliar content of essential oil from lemongrass (Cymbopogon citratus (DC) Stapf.).<br />

Crop Husbandry 48(278): 445–453.<br />

Leal T C A B, Freitas S P, Silva J F, Carvalho A J C. 2003. Production of biomass and essential oil in plants of<br />

lemongrass (Cymbopogon citratus (DC) Stapf) of various ages. Revista Brasileira de Plantas Medicinais<br />

5(2): 61–64.<br />

Lincoln D E, Langenheim J H. 1978. Effect of light and temperature on monoterpenoid yield and composition<br />

in Satureja douglasii. Biochemical Systematics and Ecology 24: 627–635.<br />

Mahanta J J, Chutia M, Bordoloi M, Pathak M G, Adhikary R K, Sarma T C. 2007. Cymbopogon citratus L.<br />

essential oil as a potential antifungal agent against key weed moulds of Pleurotus spp. spawns. Flavour<br />

and Fragrance Journal 22(6): 525–530.


The Cymbopogons 131<br />

Maheshwari S K, Sharma R K, Gangrade S K. 1995. Effect of spatial arrangement on performance of palmarosa<br />

(Cymbopogon martini var. motia)-pigeonpea (Cajanus cajan) intercropping on a black cotton soil<br />

(vertisol). Indian Journal of Agronomy 40(2): 181–185.<br />

Marongiu B, Piras A, Porcedda S, Tuveri E. 2006. Comparative analysis of the oil and supercritical CO(2)<br />

extract of Cymbopogon citratus Stapf. Natural Product Research 20(5): 455–459.<br />

Ming L C, Figueiredo R O, Machado S R, Andrade R M C. 1995. Yield of essential oil of and citral content in<br />

different parts of lemongrass leaves (Cymbopogon citratus (D.C.) Stapf.) Poaceae. International symposium<br />

on medicinal and aromatic plants, Amherst, Massachusetts, August 27–30, 1995 [edited by Craker<br />

L E, Nolan L, Shetty K] Acta Horticulturae (1996) (No. 426): 555–559.<br />

Misra A, Srivastava N K. 1994. Influence of iron nutrition on chlorophyll contents, photosynthesis and essential<br />

monoterpene oil(s) in Java citronella (Cymbopogon winterianus Jowitt). Photosynthetica 30(3):<br />

425–434.<br />

Moody J O, Adeleye S A, Gundidza M G, Wyllie G, Ajayi-Obe O O. 1995. Analysis of the essential oil of<br />

Cymbopogon nardus (L.) Rendle growing in Zimbabwe. Pharmazie 50(1): 74, 75.<br />

Nagao E O, Innecco R, Mattos S H, Filho S M, Marco C A. 2004. Effect of the harvest time on content and<br />

major constituent of essential oil of Lippia alba (Mill) N.E.Br., chemotype citral-limonene. Cienc Agron<br />

35: 355–360.<br />

Nair E C G, Chinnamma N P, Puspha Kumari R. 1979. A quarter century of research on lemongrass at lemongrass<br />

research station. Indian Perfumer 23(3): 218, 219.<br />

Nascimento I B do, Innecco R, Marco C A, Mattos S H, Nagao E O. 2003. Effects of cutting time on the essential<br />

oil of lemon grass. Revista Ciencia Agronomica 34(2): 169–172.<br />

Oyen L P A, Dung N X. 1999. Plant Resources of South-East Asia No. 19: <strong>Essential</strong> <strong>Oil</strong> Plants. Backhuys<br />

Publishers, Leiden.<br />

Pal S, Chandra S, Singh S, Balyan S, Kaul B L. 1990. Harvest management studies on lemongrass (CKP–25)—a<br />

new hybrid strain. Indian Perfumer 34(3): 213–216.<br />

Panda H. 2005. Cultivation and Utilization of Aromatic Plants. Asia Pacific Business Press Delhi, 600 pp.<br />

Pandey A K, Chowdhury A R. 2000. Chemical composition of essential oil of Cymbopogon flexuosus from<br />

Central India. Proceedings of Centennial conference on Spices and Aromatic Plants, September 20–23,<br />

2000, Calicut, Kerala, India.<br />

Pandey A K, Pandey B K, Banerjee S K, Shukla P K. 2001. Oushdhiya Paudhon Ki Kheti-Lemongrass<br />

(Cymbopogon flexuosus), 73 ICFRE-BR-15,CFRHRD-BR-4, Chhindwara, Madhya Pradesh, India.<br />

Pandey M C, Sharma J R, Dikshit A. 1996. Antifungal evaluation of the essential oil of Cymbopogon pendulus<br />

(Nees ex Steud.) Wats. cv. Praman. Flavour and Fragrance Journal 11(4): 257–260.<br />

Paranagama P A, Abeysekera K H T, Abeywickrama K, Nugaliyadde L. 2003. Fungicidal and anti-aflatoxigenic<br />

effects of the essential oil of Cymbopogon citratus (DC.) Stapf. (lemongrass) against Aspergillus flavus<br />

Link. isolated from stored rice. Letters in Applied Microbiology 37(1): 86–90.<br />

Patra D D, Singh R, Singh D V. 1990. Agronomy of Cymbopogon species. Current Research on Medicinal and<br />

Aromatic Plants 12: 72–84.<br />

Peisíno Alan Luppi, Alberto Diogo Laranja, Mendes Marisa, Calçada Luís Américo. 2005. Study of the frying<br />

effects in the composition of the essential oil of Lemon grass (Cymbopogan citratus). 2nd Mercosur<br />

Congress on Chemical Engineering, ENPROMER 2005. Rio de Janeiro, Brazil.<br />

Pino J A, Rosado A, Correa M T. 1996. Chemical composition of the essential oil of Cymbopogon winterianus<br />

Jowitt from Cuba. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 8(6): 693, 694.<br />

Rana V, Prakash Om, Joshi C S. 1996. Effect of nitrogen application and cutting frequencies on productivity of<br />

lemongrass (Cymbopogon flexuosus) under Poplar (Populus deltoids) plantation. Indian Perfumer 40(3):<br />

67–72.<br />

Raina V K, Srivastava S K, Tandon S, Mandal S, Aggarwal K K, Kahol A P, Sushil Kumar. 1998. Effect of<br />

container materials and long storage periods on the quality of aromatic grass oils. Journal of Medicinal<br />

and Aromatic Plant Sciences 20(4): 1013–1017.<br />

Raina V K, Srivastava S K, Aggarwal K K, Syamasundar K V, Khanuja S P S. 2003. <strong>Essential</strong> oil composition of<br />

Cymbopogon martinii from different places in India. Flavour and Fragrance Journal 18(4): 312–315.<br />

Rajeswara Rao B R, Chand S, Bhattacharya A K, Kaul P N, Singh C P, Singh K. 1998. Response to lemongrass<br />

(Cymbopogon flexuosus) cultivars to spacing and NPK fertilizers under irrigated conditions in semi-arid<br />

tropics. Journal of Medicinal and Aromatic Plant Sciences 20: 407–412.<br />

Rajeswara Rao B. R. 2001. Biomass and essential oil yields of rainfed palmarosa (Cymbopogon martinii (Roxb.)<br />

Wats. var. motia Burk.) supplied with different levels of organic manure and fertilizer nitrogen in semiarid<br />

tropical climate. Industrial Crops and Products 14(3): 171–178.


132 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Rajeswara Rao B R, Bhattacharya A K, Mallavarapu G R, Ramesh S. 2004. Yellowing and crinkling disease<br />

and its impact on the yield and composition of the essential oil of citronella (Cymbopogon winterianus<br />

Jowitt). Flavour and Fragrance Journal, 19(4): 344–350.<br />

Rajeswara Rao B R, Kaul P N, Syamasundar K V, Ramesh S. 2005. Chemical profiles of primary and secondary<br />

essential oils of palmarosa (Cymbopogon martinii (Roxb.) Wats var. motia Burk.). Industrial Crops and<br />

Products 21(1): 121–127.<br />

Rao B L, Verma V, Sobti S N. 1980. Genetic variability in citral containing Cymbopogon khasianus. Indian<br />

Perfumer 24: 13–16.<br />

Rao B L, Lala S. 1992. Optimization of oil yield in CKP-25. Indian Perfumer 36(4): 246, 247.<br />

Rao E V S P, Rao R S G, Puttanna K, Ramesh S. 2005. Significance of harvest intervals on oil content and<br />

citral accumulation in variety Krishna of lemongrass (Cymbopogon flexuosus). Journal of Medicinal and<br />

Aromatic Plant Sciences 27(1): 1–3.<br />

Rao P G. 2006. Citronella. Regional Research Laboratory, Jorhat, India.<br />

Reverchon E. 1997. Supercritical fluids extraction and fractionation of essential oils and related products.<br />

Journal of Supercritical Fluids 10: 1–37.<br />

Rocha M F A, Borges N S S, Inneco R, Mattos S H, Nagao E O. 2002. Influencia do Larario de Corte Sobre o<br />

Citronellol do oleo essencial (Cymbopogon winterisnus). Hortio Bras 20(Suppl.): 1–4.<br />

Rozzi N L, Phippen W, Simon J E, Singh R K. 2002. Supercritical fluid extraction of essential oil components<br />

from lemon-scented botanicals. Lebensmittel-Wissenschaft und–Technologie, 35(4): 319–324.<br />

Rudloff V E, Underhill W E. 1965. Seasonal variation in the volatile oil from Tanacetum vulgare L.<br />

Phytochemistry 4: 11–17.<br />

Sahoo S. 1994. Variability, selection and mutation for palmarosa oil production in India. Journal of Herbs,<br />

Spices & Medicinal Plants 2(3): 49–57.<br />

Saikia R C, Sarma A, Sarma T C, Baruah P K. 2006. Comparative study of essential oils from leaf and inflorescence<br />

of Java citronella (Cymbopogon winterianus Jowitt). Journal of <strong>Essential</strong> <strong>Oil</strong> <strong>Bearing</strong> Plants 9(1): 85–87.<br />

Sandra P, Bicchi C. 1987. Capillary Gas Chromatography in <strong>Essential</strong> <strong>Oil</strong>, Heidelberg, Basel, New York.<br />

Sargenti S R, Lancas F M. 1997. Supercritical fluid extraction of Cymbopogon citratus (DC.) Stapf. Chromatographia<br />

465: 6285.<br />

Sarma A, Sarma T C, Handique A, Baruah A K S. 2003. Variation in major chemical constituents in the oil of<br />

lemon grass (Cymbopogon flexuous [Stud. Wats]), association in different seasons under Brahmaputra<br />

valley agro-climatic conditions, FAFAI Journal 5(2): 43–49.<br />

Sarma T C, Sharma R K, Adhikari R K, Saha B N. 2001. Effect of altitude and age on herb yield, oil and its<br />

major constituents of Java citronella (Cymbopogon winterianus Jowitt.). Journal of <strong>Essential</strong> <strong>Oil</strong> Bear ing<br />

Plants 4: 77–81.<br />

Sarma T C. 2002. Variation in oil and its major constituents due to season and stage of the crop in Java citronella<br />

(Cymbopogon winterianus Jowitt). Journal of Spices and Aromatic Crops 11(2): 97–100.<br />

Shahi A K, Tava A. 1993. <strong>Essential</strong> oil composition of three Cymbopogon species of Indian Thar Desert.<br />

Journal of <strong>Essential</strong> <strong>Oil</strong> Research 5: 639–643.<br />

Shahi A K, Sharma S N, Tava A. 1997. Composition of Cymbopogon pendulus (Nees ex Steud) Wats, an<br />

elemicin-rich oil grass grown in Jammu region of India. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 9(5):<br />

561–563.<br />

Shahidullah M, Huq M F, Islam U, Karim M A, Begum F. 1996. Effect of number of harvests on herbage yield<br />

and oil content of citronella and palmarosa grasses. Bangladesh Journal of Scientific and Industrial<br />

Research 31(1): 127–135.<br />

Singh A, Balyan S S, Shahi A K. 1978. Harvest management studies and yield potentiality of Jammu lemongrass.<br />

Indian Perfumer 22(3): 189–191.<br />

Singh A K, Naqvi A A, Ram G, Singh K. 1994. Effect of hay storage on oil yield and quality in three Cymbopogon<br />

species (C. winterianus, C. martinii and C. flexuosus) during different harvesting seasons. Journal of<br />

<strong>Essential</strong> <strong>Oil</strong> Research 6(3): 289–294.<br />

Singh A K, Ram G, Sharma S. 1996. Accumulation pattern of important monoterpenes in the essential oil of<br />

citronella Java (Cymbopogon winterianus) during one year of crop growth. Journal of Medicinal and<br />

Aromatic Plant Sciences 18(4): 803–807.<br />

Singh K, Kothari S K, Singh D V, Singh V P, Singh P P. 2000. Agronomic studies in cymbopogons—a review.<br />

Journal of Spices and Aromatic Crops 9(1): 13–22.<br />

Singh M, Chandrasekhara G D, Rao E V S P. 1996b. <strong>Oil</strong> and herb yields of Java citronella (Cymbopogon winterianus<br />

Jowitt) in relation to nitrogen and irrigation regimes. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 8(5):<br />

531–534.


The Cymbopogons 133<br />

Singh M. 1997. Growth, herbage, oil yield, nitrogen uptake and nitrogen utilization efficiency of different<br />

cultivars of lemongrass (Cymbopogon flexuosus) as affected by water regimes. Journal of Medicinal and<br />

Aromatic Plant Sciences 19(3): 695–699.<br />

Singh M, Shivaraj B, Sridhara S. 1996c. Effect of plant spacing and nitrogen levels on growth, herb and oil<br />

yields of lemongrass (Cymbopogon flexuosus (Steud.) Wats. var. Cauvery). Journal of Agronomy and<br />

Crop Science 177(2): 101–105.<br />

Singh M. 2001. Long-term studies on yield, quality and soil fertility of lemongrass (Cymbopogon flexuosus) in<br />

relation to nitrogen application. Journal of Horticultural Science and Biotechnology 76(2): 180–182.<br />

Singh N, Luthra R, Sangwan R S. 1989. Effect of leaf position and age on the essential oil quantity and quality<br />

in lemongrass (Cymbopogon flexuosus) Planta Medica 55(3): 254–256.<br />

Singh N, Luthra R. 1988. Sucrose metabolism and essential oil accumulation during lemongrass (Cymbopogon<br />

flexuosus Stapf.) leaf development. Plant Science 57: 127–133.<br />

Sobti S N, Verma V, Rao B L. 1982. Scope for development of new cultivars of Cymbopogon as a source of<br />

terpene chemicals. In Cultivation and Utilization of Aromatic Plants, Regional Research Laboratory,<br />

Jammu, Tawi, India, pp. 302–307.<br />

Sushil K, Samresh D, Kukreja A K, Sharma J R, Bagchi G D. 2000. Cymbopogon: The Aromatic Grass<br />

Monograph. Central Institute of Medicinal and Aromatic Plants, Lucknow, India, 380 pp.<br />

Taskinen J, Mathela D K, Mathela C S. 1983. Composition of the essential oil of Cymbopogon flexuosus.<br />

Journal of Chromatography A 262: 364–366.<br />

Thomas J, Geetha K, Joy P P. 1980. Comparative performance of lemon grass species. Indian Perfumer 34:<br />

171, 172.<br />

Virmani O P, Singh K K, Singh P. 1979. Java Citronella and its Cultivation in India. Central Institute of<br />

Medicinal and Aromatic Plants, Farm Bulletin No. 1, Lucknow, India.<br />

Voirin B, Brun N, Bayet C. 1990. Effect of day length on the monoterpene composition of leaves of Mentha<br />

piperita. Phytochemistry 29: 749–755.<br />

Wannissorn B, Jarikasem S, Soontorntanasart T. 1996. Antifungal activity of lemon grass oil and lemon grass<br />

oil cream. Phytotherapy Research 10(7): 551–554.<br />

Weurman C. 1969. Journal of Agricultural and Food Chemistry 17: 370.<br />

Weiss E A. 1997. Gramineae. In <strong>Essential</strong> <strong>Oil</strong> Crops. CAB International, Wallingford, U.K., pp. 59–137.


4<br />

contents<br />

Biotechnological Studies<br />

in Cymbopogons<br />

Current Status and Future Options<br />

Ajay K. Mathur<br />

4.1 Introduction .......................................................................................................................... 135<br />

4.2 Cymbopogon: The Aromatic Genus ..................................................................................... 136<br />

4.3 Biotechnological Studies in Cymbopogons .......................................................................... 137<br />

4.3.1 Tissue Culture Studies .............................................................................................. 137<br />

4.3.2 Phylogenetic Studies ................................................................................................. 139<br />

4.3.3 Metabolic Engineering Studies................................................................................. 141<br />

4.4 Future Options in Biotechnology of Cymbopogons ............................................................. 142<br />

4.5 Conclusion ............................................................................................................................ 144<br />

References ...................................................................................................................................... 145<br />

4.1 IntroductIon<br />

Human dependence on plants for food, shelter, flavors, fragrance, colors, and health is prehistoric.<br />

Recent years have witnessed an unprecedented preference for natural herbals in health, nutraceuticals,<br />

and cosmetic sectors, and the gap between demand and supply is fast widening. Since the<br />

majority of medicinal and aromatic plants are still collected from the wild, several natural populations<br />

are being threatened or extinction. This changing scenario demands two immediate measures:<br />

(1) bringing more and more medicinal and aromatic plants under organized cultivation/domestication<br />

and (2) upgrading conventional plant improvement techniques with modern tools of plant<br />

biotechnology (Galili 2002; Gomez-Galera et al. 2007; Kole 2007). The human quest to improve<br />

the yield and quality of bioresources dates back to more than 10,000 years, with simple selection<br />

of traits quantifiable at the naked-eye level (Jauhar 2006). Such empirical selective breeding<br />

approaches then moved toward channelization of useful genes through conscious hybridization to<br />

induction of novel characters through deliberate mutation and polyploid breeding. With the advent<br />

of cell and tissue culture approaches coupled with tools of genetic engineering, the state of the art<br />

has reached to a level where crops are being designed by borrowing genes of interest from across<br />

the taxonomic boundaries, spread over plant, animal and microbial kingdoms (Vasil 2005; Datta<br />

2007; Jullien 2007). Flava-savor tomato, golden rice and Bt-cotton are just a few examples of the<br />

tremendous advantages that these tools of plant biotechnology can offer if properly amalgamated<br />

with traditional breeding approaches.<br />

Cymbopogons, belonging to the aromatic genus Cymbopogon of the angiosperms, are an attractive<br />

plant resource for the aroma industry because of their ability to grow under diverse environments,<br />

possession of significant chemical polymorphism in their essential oils, and the ease of<br />

low-input cultivation in even the most neglected rural setting (Croteau and Gershenzon 1994; Khanuja<br />

et al. 2005). The cymbopogons, despite their perennial nature coupled with erratic flowering and<br />

135


136 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

seed-setting behavior, an extremely narrow genetic base, and high susceptibility to environmental<br />

stress, constitute a storehouse of useful terpenes of high premium that value make these aromatic<br />

grasses an ideal candidate for biotechnological prospecting and scrutiny. Biotechnological<br />

approaches that are being vigorously pursued in Cymbopogon species today include the following:<br />

(1) somaclonal breeding for varietal improvement, (2) in vitro cloning of elites, (3) molecular<br />

taxonomy, (4) enzymatic/genetic dissection of the biogenetic pathways associated with terpenoid<br />

synthesis, (5) isolation, cloning, and expression of pathway related genes, (6) production of designer<br />

genotypes through pathway engineering and transgenic route, and (7) high-throughput screening of<br />

various constituents of the essential oils for novel bioactivities and their diversified usage (Mathur<br />

et al. 1988a, 1988b; Sreenath and Jagadishchandra 1991; Patnaik et al. 1999; Fiehn 2002; Hall et al.<br />

2002; Khanuja et al. 2005; Oksman-Caldentey and Saito 2005; Shasany et al. 2007; Kumar et al.<br />

2007a, 2007b; Tyo et al. 2007; Edris 2007; Gomez-Galera et al. 2007).<br />

This chapter provides an overview of the developments in biotechnological research carried out<br />

in Cymbopogon species over the last two decades. An attempt has also been made here to highlight<br />

the unaddressed or unfinished challenges in the study of this important group of aromatic grasses<br />

that plant biotechnologists should take up first.<br />

4.2 Cymbopogon: the aromatIc genus<br />

Stapf (1906) gave Cymbopogon the status of genus under the tribe Andropogoneae of the family<br />

Poaceae. Taxonomically, 140 species of Cymbopogon are known today. They have been classified<br />

into three series: Schoenanthi, Rusae, and Citrati (Soenarko 1977). Members of these three series<br />

are distinct in possessing thin, subcordate, or lanceolate type of leaves, respectively. Nearly 60<br />

species of Cymbopogon are odoriferous grasses that are naturally adapted to hot-moist regions<br />

of the oriental tropics, subtropical Africa, and Southeast Asia (Jagadishchandra 1975; Soenarko<br />

1977; Robbins 1983; Sreenath and Jagadishchandra 1991; Kuriakose 1995; Shasany et al. 2000).<br />

Most of the Cymbopogon species are recognized by the characteristic odor note of their essential<br />

oils. Lemongrass oil, citronella oil, palmarosa oil, gingergrass oil, and karnkusa oil obtained from<br />

C. flexuosus and C. winterianus, C. martinii var. motia/C.martinii var. sofia, and C. jwarancusa,<br />

respectively, are the five most widely traded essential oils in the aroma sector.<br />

Endowed with a differential blend of more than 50 terpenoidal constituents, these oils are in<br />

high demand in the aroma industry as a perfumery agent per se or as a source of lead molecules to<br />

derive more useful value-added products required for high-grade cosmetics or drugs. For example,<br />

lemongrass oil, which is used in a range of industrial products where a lemon flavor is required, can<br />

also provide citral that can be modified into β-ionone and methyl-ionone and serves as a starting<br />

precursor for vitamins A and E synthesis (Robbins 1983). Similarly, citronella oil is a basic ingredient<br />

in soaps and mosquito-repellent preparations, and is also a good source of citronellal, which can<br />

impart a lily aroma upon hydroxylation or can be converted into l-menthol for the toothpaste industry.<br />

Likewise, palmarosa oil rich in geraniol often fetches a high price as a permissible adulterant of<br />

costly rose oil. Tremendous value addition can be made in geraniol by converting it into nerol and<br />

laevo-citronellol (Thapa et al. 1971; Maheshwari and Mohan 1985). The essential oil of gingergrass<br />

is a good source of perillyl alcohols, which can be chemically converted into perillyl acetate to get<br />

a spearmint aroma (Zutchi 1982). Karnkusa oil from C. jwarancusa is used for its antipyretic and<br />

asthmolytic activity. Piperitone obtained from this oil can now be synthetically converted into thymol<br />

to fetch very high returns (Shasany et al. 2007).<br />

Among the six commercially cultivated species of Cymbopogon, namely, C. flexuosus (East<br />

Indian lemongrass), C. winterianus (Java citronella), C. martinii (palmarosa), C. nardus (Ceylon citronella),<br />

C. citratus (West Indian lemongrass), and C. pendulus (North Indian lemongrass), the first<br />

three are most widely used as a primary source of citral, geraniol and citronellol, citronellal, linalool,<br />

1,8-cineole, limonene, beta-caryophyllene, geranyl acetate, and geranyl formate in the perfumery<br />

world (Sobti et al. 1982; Sharma and Ram 2000; McGarvey and Croteau 1995; Aharoni et al.


Biotechnological Studies in Cymbopogons 137<br />

2005). The majority of these cultivated species are predominantly propagated through vegetative<br />

slips, except C. martinii, which is multiplied by transplanted seedlings. However, seed progeny thus<br />

raised in case of C. martinii show a lot of heterogeneity because of its outcrossing nature (Soenarko<br />

1977) and, hence, vegetative propagation is also resorted to in this species when clonal multiplication<br />

is desired.<br />

<strong>Essential</strong> oil content in Cymbopogon species has been shown to be a quantitative trait with a high<br />

degree of narrow sense heritability and distinct coinheritance and correlation patterns with other<br />

plant traits such as height, tiller number, biomass yield, etc. (Kulkarni 1994; Kole 1985; Kulkarni<br />

and Rajgopal 1986; Sharma and Ram 2000; Tripathy et al. 2007). As a result of these tight linkages,<br />

coupled with problems associated with erratic flowering and seed setting, traditional breeding<br />

methods are often found inadequate to fix genes through selfing to develop a superior genotype<br />

in Cymbopogon species. Genetic improvement in these grasses is therefore confined to recurrent<br />

selection and introduction of superior clones out of the vegetative populations. Though a few workers<br />

have resorted to induced mutation breeding approach, success has been limited and the genetic<br />

advantage was meager (Sharma and Ram 2000).<br />

4.3 bIotechnologIcal studIes In cymboPogons<br />

4.3.1 ti s s u e Cu lt u r e st u d i e s<br />

On account of the problems associated with their difficult reproductive behavior in applying traditional<br />

breeding approaches, the genetic base of cymbopogons is shrinking at a fast pace. Comple<br />

mentation of these limited breeding options with modern tools of plant biology was, therefore,<br />

considered necessary in the early 1980s and the first biotechnological tool pressed into the service<br />

was plant tissue culture (Jagadishchandra 1982; Jagadishchandra and Sreenath 1986; Mathur et al.<br />

1988a, 1988b, 1989a; Yadav et al. 2000; Zheng et al. 2007). In the initial phase, tissue culture<br />

techniques were applied to standardized methods for in vitro cloning of elite selections via axillary<br />

shoot culturing in C. martinii, C. nardus, C. citratus, and C. jwarancusa (Jagadishchandra 1982).<br />

Both rhizome and nodal explants readily responded on a hormone-free Murashige and Skoog (1962)<br />

medium. Additional supplementation of 6-benzyladenine in the medium at the 0.1–0.5 mg/L level<br />

was found beneficial for the growth of 5–8 new leaves from the preformed meristems present in the<br />

explants in 3–4 weeks’ time.<br />

The focus of the plant tissue culturist was then shifted toward the callus cultures, with two primary<br />

goals: (1) to develop somatic embryogenesis-based rapid clonal propagation system and (2) to enhance<br />

the spectrum of variability through in vitro mutagenesis, followed by calli cloning to produce somaclonal<br />

variations for desired traits. A variety of explants such as zygotic embryos, mesocotyl pieces<br />

of seedling, stem, rhizome, and leaf sheath base were tested in different Cymbopogon species for<br />

the purpose of callus induction (see Sreenath and Jagadishchandra 1991). MS, B5 (Gamborg et al.<br />

1968), and LS (Linsmaier and Skoog 1965) basal media were found suitable for inducing a callus<br />

response from these explants. Young unemerged inflorescences with early stages of flower ontogeny<br />

were also found to be responsive for callus initiation in C. martinii (Baruah and Bordoloi 1989).<br />

Most of the workers faced initial problems in initiating an axenic culture in Cymbopogon species.<br />

Leaching of phenolics from the injured ends of the explants and high degree of contamination due<br />

to endophytic microbes present deep inside the explants were the two major hurdles encountered.<br />

Addition of antioxidants such as ascorbic acid (40–80 mg/L) in the callus initiation medium and<br />

frequent shifting of explants onto the fresh medium during the initial 2 weeks of culturing were<br />

found to be very effective in circumventing the problem of browning of explants (Mathur et al.<br />

1988a, 1989a). These workers also showed that soaking of explants in aqueous solution of 50 mg/L<br />

chloramphenicol for 8–12 h prior to their plating on the culture medium could also bring about a<br />

reduction in contamination frequency. Among the various growth regulators tested, 2,4-dichlorophenoxyacetic<br />

acid (2,4-D) at 1.0–5.0 mg/L was found to be most effective in eliciting a callus


138 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

response in cymbopogons. BAP and α-naphthaleneacetic acid (NAA) at low doses (0.1–1.0 mg/L)<br />

in 2, 4-D containing medium were often found helpful in making the induced callus organogenetic<br />

in subsequent passages. Once induced, the callus of all Cymbopogon species was found to be very<br />

fast proliferating with high regenerative potential (Mathur et al. 1988a, 1988b; Nayak et al. 1996;<br />

Patnaik et al. 1999). Considering the reports published on callus induction in Cymbopogon species<br />

so far, the leaf sheath base and seedling explants were found most totipotent for this response.<br />

Mature leaves, young rolled leaves, and internodal stem explants were found least responsive. Callus<br />

initiated from former set of explants normally sustained their regeneration potential for a longer<br />

time (2–4 years). Callus originating from rhizome pieces, on the other hand, depicted a very low<br />

organogenetic potential.<br />

Callus cultures of cymbopogons normally showed two types of morphology: (1) organogenic and<br />

(2) nonorganogenic. The organogenic callus is usually nodular, fragile, yellowish, and opaque in<br />

external appearance, whereas their nonorganogenic counterpart is nonnodular, pale white, slimy, and<br />

translucent (Mathur et al. 1989a; Jagadishchandra 1982; Sreenath and Jagadishchandra 1989, 1991).<br />

Nonorganogenic calli are further characterized by the presence of loosely arranged isodiametric<br />

parenchyma. In comparison, the organogenic callus has a central core of parenchyma surrounded by<br />

a peripheral nodular zone consisting of meristematic cells. Friable callus also goes fast into a suspension<br />

mode (Sreenath and Jagadishchandra 1980; Patnaik et al. 1995, 1996, 1997, 1999). A 1:5 dilution<br />

every 2–3 weeks was found suitable for maintaining these suspension cultures in vitro.<br />

High-frequency regeneration of somatic embryos or adventive shoot buds is a common occurrence<br />

in Cymbopogon calli. While C. martinii and C. citratus calli predominantly regenerate via<br />

the somatic embryogenic route, C. winterianus callus showed a regeneration mode involving de<br />

novo formation of shoot bud primordia (Nayak et al. 1996; Patnaik et al. 1999; Mathur et al. 1988a).<br />

Cymbopogon jwarncusa calli showed both routes of plant regeneration. Though regeneration can<br />

often occur on the callus multiplication medium itself, removal of 2,4-D from the media was found<br />

essential in some Cymbopogon species (Mathur et al. 1989a, 1989b). These workers observed that<br />

C. winterianus callus cultures did not enter the shoot regeneration stage unless they were shifted from<br />

2,4-D containing medium to a medium supplemented with 3-indole acetic acid (IAA) and BAP or<br />

kinetin (Kn). The regenerated shoots showed root initiation on the shoot multiplication medium, but<br />

a short exposure of individually excised shoots on medium with 0.1–0.5 mg/L NAA or IAA alone<br />

was found to enhance the root quality and subsequent establishment of plantlets in soil (Mathur<br />

et al. 1988a, 1988b). Following these protocols, efficient plant regeneration has been reported in palmarosa<br />

grass (Baruah and Bordoloi 1989; Patnaik et al. 1995, 1997, 1999). Plants derived through<br />

somatic embryos, in general, showed better genetic uniformity in comparison to those obtained<br />

through shoot bud organogenesis. Though a few workers have also reported a wide range of chromosomal<br />

variations in palmarosa callus (Patnaik et al. 1996; Sreenath and Jagadishchandra 1991)<br />

but embryogenesis and subsequent plantlet regeneration were found to occur from cells with normal<br />

diploid constitution only. These protocols have been employed for large-scale clonal propagation of<br />

this aromatic herb and have important implications for in vitro maintenance of inbreds and male<br />

sterile lines of palmarosa for harnessing the hybrid vigor inbreeding program involving sexual<br />

hybrids, synthetics, and composites (Ahuja et al. 2000). Nayak et al. (1996) reported a method<br />

of rapid in vitro propagation in C. flexuosus through somatic embryogenesis. Somatic embryos<br />

were formed on callus cultures grown on MS medium with 5.0 mg/L 2,4-D, 0.1 mg/L NAA, and<br />

0.5 mg/L Kn. Embryo-to-plantlet conversion occurred on medium supplemented with 3.0 mg/L<br />

BAP, 1.0 mg/L GA 3, and 0.1 mg/L NAA. The regeneration potential of these lemongrass cultures<br />

was stably sustained for >2 years. These somatic embryo-derived plants, when examined under field<br />

conditions, did not show any significant variation over the base material.<br />

The scope and potential of tissue culture technology for widening the narrow genetic base in<br />

cymbopogons was first demonstrated from the author’s laboratory (Mathur et al. 1988a). Realizing


Biotechnological Studies in Cymbopogons 139<br />

the constraints of applying conventional breeding systems (sexual mating and recombination<br />

through meiosis) to create variability in C. winterianus, our group at CIMAP initiated an attempt<br />

to exploit the tools of somaclonal breeding in this grass. Conceptualized by Larkin and Scowcroft<br />

(1981) in wheat and sugarcane, somaclonal variations are defined as variations that arise in vitro<br />

during the culturing of a somatic cell or tissue. Such phenotypic variations generally result either<br />

from the preexisting genetic variability in the explants or through meiotically stable epigenetic<br />

DNA modifications induced by the culture environment (Philips et al. 1994; Bajaj 1990; Duncan<br />

1997). Cryptic chromosomal changes and change in degree of DNA methylation during culturing<br />

have been found to be the main factors behind such variations. In our study in citronella, more than<br />

700 callus-derived plantlets (regenerated through adventive shoot buds) were screened under field<br />

conditions, out of which 230 calli clones were advanced to plant to row stability assessment for<br />

several productivity-linked traits such as herb yield, tiller number, diameter of the bush, dry-matter<br />

production, and essential oil content. Variations were also recorded for six important constituents of<br />

the oil, namely, citronellal, citronellol, geraniol, citronellylacetate, geranylacetate, and elemol (the<br />

negative component in the oil). Correlation analysis between these agronomic characters indicated<br />

a strong negative linkage between herb yield and oil content.<br />

Interestingly, we could recover five somaclones with high herb and high oil content, showing<br />

thereby that even tight genetic linkages can be broken through somaclonal variation. Recurrent testing<br />

of five high herb and oil content bearing type clones at different locations ultimately led to the<br />

release of a somaclone (designated CIMAP/Bio-13) as India’s first plant variety improved through<br />

the somaclonal breeding approach. This variety has shown wide climatic adaptability in rainfed<br />

to subtropical to hot and moist conditions. Beside other agronomic traits of commercial importance,<br />

the variety registered a remarkable improvement in the initial establishment frequency of<br />

its vegetative slips (>85% in comparison to


140 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

2007; Jullien 2007; Shasany et al. 2007; Tyo et al. 2007; Xu and Crouch 2008). Such molecular<br />

markers today find extensive usage in settling taxonomic controversies and phylogenetic analysis of<br />

closely similar taxa. Besides, use of these DNA markers can also help plant breeders in their varietal<br />

development programs through marker-assisted selection of desired genotypes with a designed<br />

combination of genes that control the yield and quality traits (Khanuja et al. 1999; Sangwan and<br />

Sangwan 2000; Aharoni et al. 2006; Kumar et al. 2007a, 2007b; Xu and Crouch 2008). These DNA<br />

markers can, therefore, enhance the speed of any hybridization or transgenomic program aimed at<br />

producing designer specialty crops.<br />

Due to their large number and wider occurrence coupled with a prevalent introgressive hybridization<br />

mechanism, the precise phylogenetic identification of Cymbopogon species has always<br />

been a difficulty (Jagadishchandra 1975; Khanuja et al. 1999, 2005; Sangwan and Sangwan 2000;<br />

Sangwan et al. 2000; Shasany et al. 2000, 2007; Sharma and Ram 2000; Kumar et al. 2007a,<br />

2007b). Primary taxonomic delimitation is basically based on chemotaxonomic differences<br />

between the species. Isozyme patterns have also been employed to demarcate Cymbopogon species<br />

(Ganjewala and Luthra 2007). According to Shasany et al. (2007), the genus Cymbopogon is<br />

today known to have 140 species. Out of these, 52 are described from Africa, 45 from India, 6 each<br />

from Australia and South America, 4 from Europe, and the remaining species are distributed in<br />

different parts of Southeast Asia. A literature survey of taxonomy of the Cymbopogon species<br />

indicates that many species were classified on the basis of their morphological and oil quality<br />

profile alone. Information on their phylogenesis has been obscure. Distinct chemical features of<br />

otherwise morphologically similar C. martinii var. sofia and C. martinii var. motia on the one hand<br />

and morphologically distinct but chemically identical C. citratus and C. flexuosus on the other<br />

are typical examples of the complexity of such phenotypic/chemotaxonomic classification. Such<br />

an approach may lead to erroneous identity and evolutionary relationships between Cymbopogon<br />

species. Addressing this critical issue, Shasany et al. (2000, 2007) and Khanuja et al. (2005) conducted<br />

a detailed study to establish the ancestry of several Cymbopogon members. Phylogeny was<br />

traced by construction of RAPD/AFLP maps. These workers could successfully estimate the extent<br />

of molecular diversity among 19 Cymbopogon taxa belonging to 11 taxonomically classified species,<br />

2 released varieties, 1 hybrid taxon, and 4 unidentified wild collections (Khanuja et al., 2005).<br />

Citral content was employed as a base marker for chemotypic clustering. Based on their results,<br />

the workers have proposed the elevation of C. flexuosus var. microstachy to a separate species<br />

level. They also concluded that C. travancorensis, which was earlier merged with C. flexuosus,<br />

also deserves a separate species status. Their data also substantiated the independent species status<br />

for sofia and motia varieties of C. martinii. Several unidentified wild collections included in this<br />

molecular tagging work also required a separate taxonomic identity as intermediate forms in the<br />

evolution of new taxa.<br />

Advancing the scope of molecular markers for assigning phylogenetic relationships between different<br />

species of Cymbopogon, Kumar et al. (2007a, 2007b) have recently reported the development<br />

of a set of simple sequence repeat (SSR) markers. They used a genomic library of C. jwarancusa to<br />

develop these markers for the precise identification of Cymbopogon species up to the accession level.<br />

The SSRs containing genomic DNA clones of C. jwarancusa contained a total of 32 SSRs with a<br />

range of 1–3 SSRs per clone. About 68.8% of the 32 SSRs had dinucleotide repeat motifs, followed<br />

by 21.8% SSRs with trinucleotide and other higher-order repeat motifs. Eighteen of the 32 designed<br />

primers for the SSRs, amplified the products to anticipated sizes when tried with genomic DNA<br />

of source species C. jwarancusa. Thirteen of these 18 functional primers detected polymorphism<br />

among C. flexuosus, C. pendulus, and C. jwarancusa and amplified a total of 95 alleles with a PIC<br />

value of 0.44 to 0.96 per SSR. The workers have postulated that higher allelic range and high level<br />

of polymorphism depicted by these SSRs can be put to use in a variety of applications in genetic<br />

improvement of Cymbopogon species through genotype/species authentication and mapping or tagging<br />

the genes controlling the productivity-linked traits in marker-assisted breeding endeavors.


Biotechnological Studies in Cymbopogons 141<br />

4.3.3 Me ta b o l iC en G i n e e r i nG st u d i e s<br />

Metabolic engineering is a relatively new area of plant biotechnology research; the term was coined<br />

for the first time by Bailey (1991). The concepts in metabolic engineering normally revolved around<br />

the identification of the major limiting blocks in the synthesis and accumulation of a desired plant<br />

metabolite, followed by systematically removing these blocks with the help of gene manipulation<br />

options available today (Verpoorte and Alfermann 2000; Endt et al. 2002; Martin et al. 2003; Horn<br />

et al. 2004; Liao 2004; Koffas and Cardayre 2005). Three conditions must be met in a plant (cell)<br />

before a significant increment in the biogenesis of a metabolite can be expected to occur. They are<br />

(1) induction of genes/enzymes involved in the targeted metabolic pathway at right time and place;<br />

(2) availability and supply of the precursors has to be assured; and (3) there should be sufficient<br />

sink capacity for storage of the desired metabolite (Mathur et al. 2006). Since the primary goal in<br />

a metabolic engineering effort is to divert or channelize the pathway flux toward the synthesis and<br />

accumulation of a required phytomolecule, the experimental approaches that are the focus of such<br />

a program include the following:<br />

1. Downregulation of the flux in undesired route at branch points of a pathway by antisense<br />

or RNAi approach<br />

2. Overexpression of rate-limiting enzyme<br />

3. Decrease catabolism of end products<br />

4. Increase in metabolite-producing or metabolite-storing cell types<br />

5. Manipulation of regulatory elements acting as transcriptional activator/repressor of several<br />

genes associated with the target pathway<br />

6. Generation of heterologous expression systems capable of operating minipathways of<br />

a broad complex, multistep pathway to obtain intermediates that can be converted into a<br />

desired molecule using tools of semisynthetic chemistry or cell-free biotransformations.<br />

Metabolic engineering, therefore, is a sum of all the optimization efforts, culminating in the upgradation<br />

of a required biochemical/genetic expression of a phenotype in a defined genotypic background<br />

(Tyo et al. 2007). Evidence from the protein engineering and system biology approaches<br />

are further advancing the precision of metabolic engineering experiments in plants. Clearly, the<br />

pathway manipulation strategies are now evolving from single-gene perturbation to global transcription<br />

machinery engineering (gTME). The coming years will see more and more focus on<br />

those phytomolecules that are exceptionally costly, are exclusively of plant origin, and that otherwise<br />

defy the rules of synthetic chemistry due to complex chirality (Hall et al. 2002; Fiehn 2002;<br />

Brent 2004; Oksman-Caldentey and Saito 2005; Yonekura-Sakakibara and Saito 2006; Hall 2006;<br />

Gomez-Galera et al. 2007). It is, therefore, no surprise that plant terpenoids are attracting most<br />

attention from metabolic engineers (Verpoorte et al. 2000; Verpoorte and Memelink 2002; Trapp<br />

and Croteau 2001; Wallaart et al. 2001; Mahmoud and Croteau 2002; Schijlen et al. 2004; Akhila<br />

2007; Chang et al. 2007; Petersen 2007). Initial efforts have indicated that multiple-step terpene<br />

pathway engineering across several cellular compartments is feasible (Aharoni et al. 2005).<br />

An enormous wealth of information on chemistry, biological activity, synthesis, and regulation<br />

of plant terpenoid metabolism exists in the literature (Akhila 1985, 1986; Croteau and Gershenzon<br />

1994; Penuelas et al. 1995; McGarvey and Croteau 1995; Dixon et al. 1996; Lichtenthaler 1999;<br />

Verpoorte et al. 2000; Hallahan 2000; Bourgaud et al. 2001; Eisenreich et al. 2001; Trapp and<br />

Croteau 2001; Pichersky and Gershenzon 2002; Holopainen 2004; Cheng et al. 2007). Terpenoids,<br />

in general, are a structurally varied class of natural products that are commercially in demand<br />

as flavoring, perfumery, pharmaceuticals, insecticidal, and antimicrobial agents (Martin et al.<br />

2003). In plants, they play important roles in plant–plant, plant–environment, and plant–microbe<br />

or plant–insect interactions, and impart an ecological fitness to a plant (Aharoni et al. 2005, 2006).


142 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

A generalized scheme of biosynthesis of terpenoidal molecules starts with the condensation of isopentyl<br />

diphosphate (IDP) and its allylic isomer dimethyl-allyl diphosphate (DMADP). The sequential<br />

head-to-tail addition of IDP units with DMADP yields geranyl diphosphate (GDP), farnesyl<br />

diphosphate (FDP), and geranylgeranyl diphosphate (GGDP). Under the influence of synthases or<br />

cyclases, these three metabolites then serve as starting precursors for mono-, sesqui-, and diterpenoids,<br />

respectively. The terminal decoration of the basic terpenoidal skeleton is brought about by an<br />

array of substitution reactions such as hydroxylation, dehydrogenation, reduction, glucosylation, and<br />

methyl or acyl transfers to generate the wide spectrum of terpenoid diversity that we encounter in<br />

plants. Terpenoid biosynthesis occurs both in the cytosol and the plastids via mevalonate and MEP<br />

pathways, respectively (Lichtenthaler 1999; Eisenreich et al. 2001; Wallaart et al. 2001; Akhila<br />

2007; Cheng et al. 2007).<br />

A literature survey of metabolic engineering efforts in Cymbopogon species reflects a vacuum.<br />

This is partly because of the difficulty associated with the DNA isolation and genetic transformation<br />

protocols in these grasses (Khanuja et al. 1999). Since upstream steps of terpenoid biogenesis<br />

in Cymbopogon species are the same as those in other aromatic plants such as mints, and sufficient<br />

knowledge of their in vitro handling is available today, it may be anticipated that the void will be<br />

filled soon. According to Sangwan and Sangwan (2000), application of metabolic tools in cymbopogons<br />

would find increasing attention in the following three areas:<br />

1. Bringing alterations in magnitude of flux of pyruvate and glyceraldehydes-3 phosphate<br />

toward isopentenyl pyrophosphate (IPP) generation.<br />

2. Modulation of geranyl pyrophosphate synthase or farnesyl pyrophosphate synthase, or their<br />

overexpression, to divert the flux toward monoterpene or sesquiterpene accumulation.<br />

3. Secondary modification of parental terpene molecules via individual or a combination of<br />

isomerization, reduction, oxidation, lactonization, or epoxidation reaction to derive valueadded<br />

phytomolecules with enhanced organoleptic or pharmaceutical properties.<br />

Coining a new term terpenomics to refer to the ongoing metabolic engineering approaches in aroma<br />

crops, Khanuja (2003) also emphasized the need to focus on modulation of terpene synthases’ class<br />

of enzymes in Cymbopogon and Mentha species to facilitate unique regiochemical/stereochemical<br />

configuration in intermediates of terpene metabolism to derive newer activities. It is hoped that<br />

these new bioactives will equip our plants with improved disease resistance, weed/pest control, and<br />

more powerful pollination mechanism.<br />

4.4 Future oPtIons In bIotechnology oF cymboPogons<br />

Avenues for future research in the biotechnology of Cymbopogon species are enormous. Some of<br />

the areas that demand priority efforts are discussed below to ignite the excitement of, and the challenges<br />

for, future researchers:<br />

• Somaclonal breeding work in Cymbopogon species deserves more scrutiny and attention.<br />

With the documented success story of CIMAP/Bio-13 variety of citronella Java in the kitty,<br />

future efforts should be specifically focused on enhancing the volumetric yield of important<br />

essential oils and their industrially important pure constituents. Cultivars resistant<br />

to water stress or flooding stress are required to utilize Cymbopogon species in marginal<br />

lands or wasteland. Somaclones with better resistance to nutritional and microbial diseases<br />

are required through directed in vitro mutagenesis and cell line selections.<br />

• A sound understanding of handling Cymbopogon species in tissue cultures can also be<br />

put to use in understanding the influence of microbial associations on expression of the<br />

biogenetic pathways in these grasses. In vitro tissues maintained at different morphogenetic<br />

levels (callus, cell suspensions, shoots, roots, or whole plantlets) would not only


Biotechnological Studies in Cymbopogons 143<br />

permit a better understanding of the role of tissue differentiation in advancing the pathway<br />

but would also provide the ideal aseptic scenario to assess the influence of individual<br />

micro organisms (isolated from plants’ rhizosphere) on essential oil synthesis following<br />

their deliberate incorporation in the cultured tissue. One such effort has recently made<br />

in Vetiveria zizanioides (vetiver). This study by DelGuidice et al. (2008) has opened up<br />

an entirely new avenue for research in aromatic grasses. Using a tissue culture approach,<br />

these workers have reported that essential oil biosynthesis in vetiver is confined to the first<br />

cortical layer outside the root endodermis that also harbors a community of 10 endophytic<br />

bacteria. When these bacteria were cultured on medium containing vetiver oil as sole carbon<br />

source, they were able to metabolize the oil, and each endophyte was able to release<br />

into the medium a large number of derivatives that were absent in the raw oil fed to them.<br />

Besides opening a new line of investigation on in vitro bioconversions, this approach will<br />

be useful in understanding the microbial factors responsible for the final expression of<br />

essential oil quality in Cymbopogon plants.<br />

• Another potentially fruitful area of research in the biotechnology of cymbopogons would be<br />

the time course analysis of volatile terpene emission in the headspace of the culture vessels<br />

(Tholl et al. 2006; Predieri and Rapparini 2007). This can prove to be a powerful nondestructive<br />

method of monitoring the physiological status of the plants in relation to the synthesis of<br />

the oil constituents under a controlled environment (Alonzo et al. 2001). The biogenic emission<br />

of terpenes in field-grown plants has been found to be associated with oxidative chemistry<br />

of the surrounding environment and the plant itself. It will be interesting to trace these volatile<br />

emission changes following various stress and/or elicitation treatments given to a controlled<br />

culture in vitro (Maes and Debergh 2003; Loreto et al. 2006; Sharkey and Yeh 2001).<br />

• Most of the Cymbopogon species, similar to other aromatic crops, synthesize their essen-<br />

tial oils in the epidermal cells of the leaves and store the synthesized oil in glandular<br />

trichomes (Shankar et al. 1999; Hallahan 2000). It is probably this tight linkage between<br />

oil synthesis and tissue differentiation that has deterred researchers from exploiting cell<br />

culture approaches for in vitro metabolite production in these species. However, with recent<br />

advancements made in the molecular understanding of terpenoidal pathways in plant and<br />

refinements in cell culture technology, it is desirable to revisit this research area (Tisserat<br />

and Vaughan 2008). Since microshoots of cymbopogons reared in vitro also depict this tissue<br />

specialization, a modest beginning can be made by upscaling them in bioreactors for<br />

harvesting of desired metabolite. Such an in vitro production system will be particularly<br />

relevant for producing oil constituents that are otherwise synthesized in very low amounts<br />

in plants. Alternately, they can be used for the production of a single predominant terpene,<br />

as has been done for (−) carvone in Mentha spicata (Jullien 2007). Until this stage<br />

is reached with cymbopogons, the cell cultures of aromatic grasses must be vigorously<br />

screened as a source of common enzymes and intermediates of terpene metabolism to<br />

carry out useful biotransformation for value addition of aroma phytoceuticals.<br />

• In spite of major advances made in tissue culturing of Cymbopogon species, no effort<br />

•<br />

has so far been made to standardize transgenic protocols in these grasses. It seems to be<br />

more of a mental block rather than a technical hurdle. Development of an efficient somatic<br />

embryogenic-based transformation method in cymbopogons must be attempted to realize<br />

the full potential of “-omic” research.<br />

Anther and pollen culture is another research area that can complement conventional breeding<br />

work in cymbopogons. Haploid plants thus produced will be of immense utility for producing<br />

inbred lines and for inducing desired mutations in seed-propagated Cymbopogon species.<br />

• Chemomolecular prospecting of various essential oil constituents of Cymbopogon species<br />

to devise new therapeutics has enormous potential (Delespaul et al. 2000; Ojo et al. 2006;<br />

Adeneye and Agbaje 2007; Agbafor and Akubugnov 2007; Shen et al. 2007; Masuda et al.


144 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

2008). <strong>Essential</strong> oils of these grasses are already in use in aroma therapies (see Edris 2007).<br />

The bioactivities found to be associated with these oils ranged from antibacterial to antiviral,<br />

antioxidant, and antidiabetic to antineoplastic. Their capacity to enhance the permeability<br />

of stratum corneum of human skin makes them a good adjuvant in transdermal drug<br />

delivery formulations (Williams and Barry 2004). Citral, a major constituent of lemongrass<br />

oil, has been found to be a potent inducer of glutathione-s-transferase class of enzymes,<br />

which provide protection to healthy hepatocytes against apoptosis during chemo therapy<br />

of liver cancers (Nakamura et al. 2003). This protective capacity of citral was, however,<br />

confined to its trans configuration (geraniol) only. Its cis configuration (neral) had no such<br />

effect. Selection of plant cells capable of converting cis configuration to trans form should<br />

be an exciting goal for plant/microbial biotechnologists. Another constituent of lemongrass<br />

oil, namely, perillyl alcohol (POH), which is a hydroxylated analog of d-limonene, is also<br />

gaining a lot of attention as a powerful radio-sensitizer of tumor cells. It can drastically<br />

reduce the radiation doses to control breast, colon, and prostrate cancers (Duetz et al. 2003;<br />

Rajesh et al. 2003). Geraniol, the primary constituent of palmarosa grass, has been found<br />

to make tumor cells vulnerable to 5-flurouracil (5-FU) by reducing the activity of thymidylate<br />

synthase and thymidine kinase, which are responsible for 5-FU toxicity in mice.<br />

Bioprospection of essential oils of Cymbopogon species would yield agents against several<br />

plant and human pathogens such as Salmonella, Staphylococcus, Shigella, Aspergillus,<br />

Fusarium, Penicillium, etc. (Burt 2004; Wannissorn et al. 2005; Cimanga et al. 2002;<br />

Pandey et al. 2003). Bankole and Joda (2004) have reported that a very effective method<br />

of controlling storage deterioration of melon seeds due to Aspergillus species is mixing<br />

the seeds with powdered dry leaves of lemongrass (C. citratus). Complete inhibition of A.<br />

flavus and A. niger was realized, and the efficacy was found to be at par with fungicide<br />

(iprodione) treatment. Similarly, Duamkhanmanee (2008) has shown the total control of<br />

postharvest anthracnose disease of mango fruit caused by Colletotrichum gloeosporioides<br />

by C. citratus oil (4000 ppm in hot water). The anthracnose disease of cowpea crop in the<br />

field was demonstrated to be controlled by cold water leaf extract of lemongrass (Amadioha<br />

and Obi 1999). Lemongrass oil has also been shown to inhibit growth of Candida albicans<br />

at 2.0 µL/mL level in the culture medium (El-Khair 2007). <strong>Oil</strong> treatment kills the microbial<br />

cells due to enhanced K + depletion from the cells following a fall in membrane lipid<br />

content. A similar inhibitory mechanism was also observed in Saccharomyces cerevisiae<br />

challenged with geraniol of palmarosa oil.<br />

The mechanisms by which essential oils of Cymbopogon species inhibit microbial<br />

growth are far from completely understood. Their hydrophobicity has often been cited<br />

as a possible mode of action that helps them to get partitioned into the lipid bilayer of the<br />

microbial cell membrane, making it permeable enough to allow leakage of vital cellular<br />

content (Burt 2004; Pattnaik et al. 1997; Delaquis et al. 2002; Prashar et al. 2003). There<br />

is tremendous scope to advance this research at enzymatic and molecular levels to develop<br />

more ecofriendly and safe bioprotectants for plant and animal use.<br />

4.5 conclusIon<br />

Biotechnological studies so far carried out in Cymbopogon species should be integrated now for<br />

synergism, coordination, and upscaling to derive commercial gains. Availability of efficient tissue<br />

culture protocols for plant regeneration on the one hand and fairly explicit understanding of the biosynthetic<br />

pathways of their essential oils at the level of genes and enzymes on the other have set the<br />

stage to move toward the redesigning of these aroma crops with transgenic approaches. It is hoped<br />

that the vast variety of high-value terpenes that these grasses can harbor and the exclusivity of the<br />

chemical reactions that their enzymatic and genetic machinery is capable of carrying out, both in<br />

planta and in culture, would motivate chemists and biologists alike to convert them into efficient


Biotechnological Studies in Cymbopogons 145<br />

green bioreactors for aroma and pharmaceutical molecules in the near future. In vitro and in vivo<br />

bioprospecting of Cymbopogon species for deriving novel bioactives is another area that will keep<br />

the scientists intensely engaged with these aromatic grasses.<br />

reFerences<br />

Adeneye AA, Agbaje EO. 2007. Hypoglycemic and hypolipidemic effects of fresh leaf aqueous extract of<br />

Cymbopogon citratus Stapf. in rats. J Ethanopharmacol. 112: 440–444.<br />

Agbafor KN, Akubugwo EL. 2007. Hypocholesterelemic effect of ethanolic extract of fresh leaves of<br />

Cymbopogon citratus (lemongrass). African J. Biotechnol. 6: 596–598.<br />

Aharoni A, Jongsma MA, Bouwmeester HJ. 2005. Volatile science? Metabolic engineering of terpenoids in<br />

plants. Trends Plant Sci. 10: 594–602.<br />

Aharoni A, Jongsma MA, Kim T, Ri M, Giri AP, Verstappen FWA, Schwab W, Bouwmeester HJ. 2006.<br />

Metabolic engineering of terpenoid biosynthesis in plants. Phytochem. Review 5: 49–58.<br />

Ahuja PS, Mathur AK, Kukreja AK, Pandey B. 2000. Potential of plant tissue culture for the improvement of<br />

Cymbopogon species. In: Kumar S, Dwivedi S, Kukreja AK, Sharma JR, Bagchi, GD (Eds.), Cymbopogon:<br />

The Aromatic Grass: A Monograph, pp. 185–198. CIMAP (CSIR) Publication, Lucknow, India.<br />

Akhila A. 1985. Biosynthetic relationship of citral trans and citral cis in Cymbopogon flexuosus (lemongrass).<br />

Phytochemistry 24: 2585–2587.<br />

Akhila A. 1986. Biosynthesis of monoterpenes in Cymbopogon winterianus. Phytochemistry 25: 421–424.<br />

Akhila A. 2007. Metabolic engineering of biosynthetic pathways leading to isoprenoids: Mono- and sesquiterpenes<br />

in plastids and cytosol. J. Plant Interact. 2: 195–204.<br />

Alonzo G, Saiano F, Tusa N, Del Bosco SF. 2001. Analysis of volatile compounds released from embryogeneic<br />

cultures and somatic embryos of sweet oranges by head space SPME. Pl. Cell Tiss. Org. Cult.<br />

66: 31–34.<br />

Amadioha AC, Obi VI. 1999. Control of anthracnose disease of cowpea by Cymbopogon citrates and Ocimum<br />

gratissimum. Acta Phytopathologica et Entomologica Hungarica 34: 685–690.<br />

Bailey JE. 1991. Towards a science of metabolic engineering. Science 252: 1668–1675.<br />

Bajaj YPS. 1990. Somaclonal variation—Origin, induction, cryopreservation and implication in plant breeding.<br />

In: Bajaj YPS (Ed.) Biotechnology in Agriculture and Forestry, Vol 11: Somaclonal variation in crop<br />

improvement, pp. 3–48. Springer-Verlag, Berlin.<br />

Bankole SA, Joda AO. 2004. Effect of lemon grass (Cymbopogon citrates Stapf.) powder and essential oil on<br />

mould deterioration and aflatoxin contamination of melon seeds (Colocynthis citrullus L.). African J.<br />

Biotechnol. 3: 52–59.<br />

Baruah A, Bordoloi DN. 1989. High frequency plant regeneration of Cymbopogon martini (Roxb.) Wats. by<br />

somatic embryogenesis and organogenesis. Pl. Cell Rept. 8: 483–485.<br />

Bourgaud F, Gravot A, Milesi S, Gontier E. 2001. Production of plant secondary metabolites: A historical perspective.<br />

Pl. Sci. 161: 839–851.<br />

Brent R. 2004. A partnership between biology and engineering. Nature Biotechnol. 22: 1211–1214.<br />

Burt S. 2004. <strong>Essential</strong> oils: their antibacterial properties and potential applications in food—A review. Intl. J.<br />

Food Microbiol. 94: 223–253.<br />

Chang MCY, Eachus RA, Trieu W, Ro D, Keasling JD. 2007. Engineering Escherichia coli for production of<br />

functionalized terpenoids using plant P-450s. Nature Chem. Biol. 3: 274–277.<br />

Cheng A, Lou Y, Mao Y, Lu S, Wang L, Chen X. 2007. Plant terpenoids: Biosynthesis and ecological functions.<br />

J. Integrative Pl. Biol. 49: 179–186.<br />

Cimanga K, Kambu K, Tona L, Apers S, DeBruyne T, Hermans N, Totte J, Pieters L, Vlietinck AJ. 2002.<br />

Correlation between chemical composition and antibacterial activity of essential oils of some aromatic<br />

plants growing in Democratic Republic of Congo. J. Ethnopharmacol. 79: 213–222.<br />

Croteau R, Gershenzon J. 1994. Genetic control of monoterpene biosynthesis in mints (Mentha: Lamiaceae):<br />

Genetic engineering of plant secondary metabolism. Recent Adv. Phytochem. 28: 193–229.<br />

Croteau R, Kutchan TM, Lewis NG. 2000. Natural products (Secondary metabolites). In: Buchana B,<br />

Gruissem W, Jones R (Eds.), Biochemistry and Molecular Biology of Plants, pp. 1250–1318. Amer.<br />

Soc. Pl. Pathologists, Rockville, Maryland.<br />

Datta SK. 2007. Impact of plant biotechnology in agriculture. In: Pua EC, Davey MR (Eds.), Biotechnology in<br />

Agriculture and Forestry, Vol. 59. Transgenic Crops IV, pp. 3–34. Springer-Verlag, Berlin, Heidelberg.<br />

Delaquis PJ, Stanich K, Girard B, Mazza G. 2002. Antimicrobial activity of individual and mixed fraction of<br />

dill, cilantro, coriander and eucalyptus essential oils. Intl. J. Food Microbiol. 74: 101–109.


146 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Delespaul Q, de Billerbeck VG, Roques CG, Michel O. 2000. The antifungal activity of essential oils as determined<br />

by different screening methods. J. Essen. <strong>Oil</strong> Res. 12: 256–266.<br />

DelGiudice L, Massardo DR, Pontieri P, Bertea CM, Carata E, Tredici SM, Tala A, Destefamo M, Maffei M,<br />

Alifano P. 2008. The microbial community of vetiver root is necessary for essential oil biosynthesis.<br />

Proc. 52nd Italian Soc. Agric. Genet. Annual Cong., poster No. C.01, September 14–17, 2008, Italy.<br />

Dixon RA, Lamb CJ, Masoud S, Sewalt VJH, Paiva NL. 1996. Metabolic engineering: Prospects for crop<br />

improvement through the genetic manipulation of phenylpropanoid biosynthesis and defense responses—<br />

A review. Gene 179: 61–71.<br />

Duamkhanmanee R. 2008. Natural essential oils from lemongrass (Cymbopogon citrates) to control postharvest<br />

anthracnose of mango fruit. Intl. J. Biotechnol. 10: 104–108.<br />

Duetz W, Bouwmeester H, VanBeilent J, Witholt B. 2003. Biotransformation of limonene by bacteria, fungi,<br />

yeast and plants. Appl. Microbiol. Biotechnol. 61: 269–277.<br />

Duncan RR. 1997. Tissue culture induced variation and crop improvement. In: Sparks DL (Ed.), Advances in<br />

Agronomy, Vol. 58, pp. 201–210. Academic Press, London.<br />

Edris AE. 2007. Pharmaceutical and therapeutic potential of essential oils and their individual volatile constituents:<br />

A review. Phytother. Res. 21: 308–323.<br />

Eisenreich W, Rohdich F, Bacher A. 2001. Deoxyxylulose phosphate pathway to terpenoids. Trends Plant Sci.<br />

6: 78–84.<br />

El-Khair EK. 2007. Effect of Cymbopogon citrates L. essential oil on growth and morphogenesis of Candida<br />

albicans. Egyptian J. Biotechnol. 25: 145–158.<br />

Endt DV, Kijne JW, Memelink J. 2002. Transcription factors controlling plant secondary metabolism: What<br />

regulates the regulators?, Phytochemistry 61: 107–114.<br />

Fiehn O. 2002. Metabolomics—The link between genotypes and phenotypes. Pl. Mol. Biol. 48: 155–171.<br />

Galili G. 2002. Genetic, molecular and genomic approaches to improve the value of plant foods and feeds. Crit.<br />

Rev. Pl. Sci. 21: 167–204.<br />

Gamborg OL, Miller RA, Ojima K. 1968. Nutrient requirements of suspension cultures of soybean root cells.<br />

Exp. Cell Res. 50: 151.<br />

Ganjewala D, Luthra R. 2007. Identification of Cymbopogon species and C. flexuosus (Nees Ex. Steud.) Wats.<br />

cultivars based on polymorphism in esterases isoenzymes. J. Pl. Sci. 2: 552–557.<br />

Gomez-Galera S, Pelacho AM, Gene A, Capell T, Christon P. 2007. The genetic manipulation of medicinal<br />

and aromatic plants. Pl. Cell Rept. 26: 1689–1715.<br />

Hall R, Beale M, Fiehn O, Hardy N, Summer L, Bino R. 2002. Plant metabolomics: The missing link in functional<br />

genomics strategies. Plant Cell 14: 1437–1440.<br />

Hall RD. 2006. Plant metabolomics: From holistic hope to hype to hot topic. New Phytol. 169: 453–468.<br />

Hallahan DL. 2000. Monoterpenoid biosynthesis in glandular trichomes of Labiateae plants. Adv. Bot. Res. 31:<br />

77–120.<br />

Holopainen JK. 2004. Multiple functions of inducible plant volatiles. Trends Plant Sci. 9: 529–533.<br />

Horn ME, Woodard SL, Howard JA. 2004. Plant molecular farming: Systems and products. Pl. Cell Rept. 22:<br />

711–720.<br />

Jagadishchandra KS. 1975. Recent studies on Cymbopogon Spreng. (aromatic grasses) with special reference<br />

to Indian taxa: Taxonomy, cytogenetics, chemistry and scope. J. Plantation Crops 2: 43–57.<br />

Jagadishchandra KS. 1982. In vitro culture and morphogenetic studies in some species of Cymbopogon<br />

Spreng. (aromatic grasses). In: Fujiwara A (Ed.) Plant Tissue Culture, pp. 703, 704. Maruzen, Tokyo.<br />

Jagadishchandra KS, Sreenath HL. 1986. In vitro culture of rhizome in Cymbopogon Spreng and Vetiveria<br />

Bory. In: Reddy GM (Ed.) Proceedings of National Symposium on Recent Advances in Plant Cell and<br />

Tissue Culture of Economically Important Plants, pp. 199–208. Osmania University, Hyderabad, India.<br />

Jain N, Shasany AK, Sundaresan V, Rajkumar S, Darokar MP, Bagchi GD, Gupta AK, Khanuja SPS. 2003.<br />

Molecular diversity in Phyllanthus amarus assessed through RAPD analysis. Curr Sci. 85: 1454–1458.<br />

Jauhar PP. 2006. Modern biotechnology as an integral supplement to conventional plant breeding: The prospects<br />

and challenges. Crop Sci. 46: 1841–1859.<br />

Jullien F. 2007. Mints. In: Pua EC, Davey MR (Eds.), Biotechnology in Agriculture and Forestry vol 59.<br />

Transgenic Crops IV, pp. 435–466. Springer-Verlag, Berlin, Heidelberg.<br />

Khanuja SPS. 2003. Revisiting the fragrances—the terpenomics way. J. Med. Arom. Pl Sci. 25: 335.<br />

Khanuja SPS, Shasany AK, Darokar MP, Kumar S. 1999. Rapid isolation of DNA from dry and fresh samples of<br />

plants producing large amounts of secondary metabolites and essential oils. Pl. Mol. Biol. Rept. 17: 1–7.<br />

Khanuja SPS, Shasany AK, Pawar A, Lal RK, Darokar MP, Naqvi AA, Rajkumar S, Sundaresan V, Lal N,<br />

Kumar S. 2005. <strong>Essential</strong> oil constituents and RAPD markers to establish species relationship in Cymbopogon<br />

Spreng. (Poaceae). Biochem Syst. Ecol. 33: 171–186.


Biotechnological Studies in Cymbopogons 147<br />

Koffas M, Cardayre SD. 2005. Evolutionary metabolic engineering. Metabolic Eng. 7: 1–3.<br />

Kole C. 1985. Improvement of Cymbopogon winterianus Jowitt. through mutagenesis. Ind. Perfumer 29:<br />

129–138.<br />

Kole C. 2007. Genome Mapping and Molecular Breeding in Plants, Vol. 6—Technical Crops (Ed.). Springer-<br />

Verlag, Berlin.<br />

Kulkarni RN, Rajagopal K. 1986. Broad and narrow sense heritability estimates of leaf yield, leaf width, tiller<br />

number and oil content in East Indian lemongrass. J. Pl. Breed. 96: 135–139.<br />

Kulkarni RN. 1994. Phenotypic recurrent selection for oil content in East Indian lemongrass. Euphytica 78:<br />

103–107.<br />

Kumar J, Verma V, Qazi GN, Gupta PK. 2007a. Genetic diversity in Cymbopogon species using PCR based<br />

functional markers. J. Pl. Biochem. Biotechnol. 16: 119–122.<br />

Kumar J, Verma V, Sahai AK, Qazi GN, Balyan HS. 2007b. Development of simple sequence repeat markers in<br />

Cymbopogon species. Planta Med. 73: 262–266.<br />

Kuriakose KP. 1995. Genetic variability in East Indian lemongrass (Cymbopogon flexuosus Stapf.). Ind.<br />

Perfumer 39: 76–83.<br />

Larkin PJ, Scowcroft WR. 1981. Somaclonal variation—a novel source of variability from cell cultures for<br />

plant improvement. Theor. Appl. Genet. 60: 197–214.<br />

Liao JC. 2004. Custom design of metabolism. Nature Biotechnol. 22: 29–36.<br />

Lichtenthaler HK. 1999. The 1-deoxy-d-xylulose-5-phosphate pathway of isoprenoid biosynthesis in plants.<br />

Annu. Rev. Pl. Physiol. Pl. Mol. Biol. 50: 47–66.<br />

Linsmaier EM, Skoog, F. 1965. Organic growth factor requirement of tobacco tissue cultures. Physiol. Plant.<br />

18: 100–127.<br />

Loreto F, Batra C, Brilli F, Nogues I. 2006. On the induction of volatile organic compound emissions by plants<br />

as a consequence of wounding or fluctuation of light and temperature. Pl. Cell Environ. 29: 1820–1828.<br />

Maes K, Debergh PC. 2003. Volatiles emitted from in vitro grown tomato shoots during abiotic and biotic<br />

stress. Pl. Cell Tiss. Org. Cult. 75: 73–78.<br />

Maheshwari ML, Mohan J. 1985. Geranyl formate and other esters in palmarosa oil. Pafai J. 7: 21–26.<br />

Mahmoud SS, Croteau R. 2002. Startegies for transgenic manipulation of monoterpene biosynthesis in plants.<br />

Trends Plant Sci. 7: 366–373.<br />

Martin VJJ, Pitera DJ, Withers ST, Newman JD, Keasling JD. 2003. Engineering the mevalonate pathway in<br />

Escherichia coli for production of terpenoids. Nat. Biotechnol. 21: 1–7.<br />

Masuda T, Osaka Y, Ogawa N, Nakamoto K, Kuminaga H. 2008. Identification of geranic acid—a tyrosinase<br />

inhibitor in lemongrass. J. Agric. Food Chem. 56: 597–601.<br />

Mathur AK, Ahuja PS, Pandey B, Kukreja AK, Mandal S. 1988a. Screening and evaluation of somaclonal<br />

variations for quantitative and qualitative traits in an aromatic grass. Cymbopogon winterianus Jowitt.<br />

Pl. Breed. 101: 321–334.<br />

Mathur AK, Ahuja PS, Pandey B, Kukreja AK, Mathur A. 1988b. Development of superior strains of an aromatic<br />

grass Cymbopogon winterianus Jowitt, through somaclonal variation. In: Proc. International Conference on<br />

Research in Plant Sciences and its Relevance to the Future, p. 160. Delhi University Press, Delhi, India.<br />

Mathur AK, Ahuja PS, Pandey B, Kukreja A. 1989a. Potential of somaclonal variation in the genetic improvement<br />

of aromatic grasses. In: Kukreja AK, Mathur AK, Ahuja PS, Thakur RS (Eds.), Tissue Culture and<br />

Biotechnology of Medicinal and Aromatic Plants, pp. 79–90. Paramount Publishing House, New Delhi,<br />

India.<br />

Mathur AK, Ahuja PS, Kukreja AK, Pandey B, Mathur A. 1989b. Progress in the application of tissue culture<br />

for the improvement of aromatic grasses. In: Proc. XIII Annual Meeting of Plant Tissue Culture<br />

Association of India, pp. 31. North-Eastern Hill University, Shillong, India.<br />

Mathur AK, Mathur A, Seth R, Verma P, Vyas D. 2006. Biotechnological interventions in designing speciality<br />

medicinal herbs for 21st century: Some emerging trends in pathway modulation through metabolic engineering.<br />

In: Sharma RK, Arora R (Eds.), Herbal Drugs: A Twenty First Century Perspective, pp. 83–94.<br />

Jaypee Brothers Medical Publishers, New Delhi, India.<br />

McGarvey DJ, Croteau R. 1995. Terpenoid metabolism. Pl. Cell 7: 1015–1026.<br />

Murashige T, Skoog, F. 1962. A revised medium for rapid growth and bioassays with tobacco tissue cultures.<br />

Physiol. Plant. 15: 473–497.<br />

Nakamura Y, Miyamoto M, Murakami A, Ohigashi H, Osawa T, Uchida R. 2003. A phase II detoxification<br />

enzyme inducer from lemongrass: identification of citral and involvement of electrophilic reaction in the<br />

enzyme induction. Biochem. Biophys. Res. Commun. 302: 593–600.<br />

Nayak S, Debata, BK, Sahoo S. 1996. Rapid propagation of lemongrass [Cymbopogon flexuosus (Nees) Wats.]<br />

through somatic embryogenesis in vitro. Pl. Cell Rept. 15: 367–370.


148 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Ojo OO, Kabutu FR, Bello M, Babayo U. 2006. Inhibition of paracetamol-induced oxidative stress in rats<br />

by extracts of lemongrass (Cymbopogon citrates) and green tea (Camellia sinensis) in rats. African J.<br />

Biotechnol. 5: 1227–1232.<br />

Oksman-Caldentey KM, Saito S. 2005. Integrating genomics and metabolomics for engineering metabolic<br />

pathways. Curr. Opin. Biotechnol. 16: 174–179.<br />

Pandey AK, Rai MK, Acharya D. 2003. Chemical composition and antimycotic activity of the essential oils of<br />

corn mint (Mentha arvensis) and lemongrass (Cymbopogon flexuosus) against human pathogenic fungi.<br />

Pharmaceut. Biol. 41: 421–425.<br />

Patnaik J, Sahoo S, Debata BK. 1995. Somatic embryogenesis and cytological variations in long-term callus<br />

cultures of Cymbopogon martinii (Roxb.) Wats. An assessment. Pl. Sci. Res. 17: 1–5.<br />

Patnaik J, Sahoo S, Debata BK. 1996. Cytology of callus, somatic embryos and regenerated plants of palmarosa<br />

grass [Cymbopogon martinii (Roxb.) Wats.]. Cytobios. 87: 79–88.<br />

Patnaik J, Sahoo S, Debata BK. 1997. Somatic embryogenesis and plantlet regeneration from cell suspension<br />

cultures of Palmarosa grass (Cymbopogon martinii). Pl. Cell. Rept. 16: 430–434.<br />

Patnaik J, Sahoo S, Debata BK. 1999. Somaclonal variation in cell suspension culture-derived regenerants of<br />

Cymbopogon martinii (Roxb.) Wats. var. motia. Pl. Breed. 118: 351–354.<br />

Pattnaik S, Subramanyam VR, Bajpai M, Kole CR. 1997. Antibacterial and antifungal activity of aromatic<br />

constituents of essential oils. Microbios 89: 39–46.<br />

Pecchioni N, Faccioli P, Monetti A, Stanca AM, Terzi V. 1996. Molecular markers for genotype identification<br />

in small grain cereals. J. Genet. Breed. 50: 203–219.<br />

Penuelas J, Llusia J, Estiarte M. 1995. Terpenoids: A plant language. Trends Ecol. Evol. 10: 289.<br />

Petersen M. 2007. Current status of metabolic phytochemistry. Phytochemistry 68: 2847–2860.<br />

Philips RL, Kaepler SM, Olhoft P. 1994. Genetic instability of plant tissue cultures: breakdown of normal control.<br />

Proc. Natl. Acad. Sci. USA 91: 5222–5226.<br />

Pichersky E, Gershenzon J. 2002. The formation and function of plant volatiles: perfumes for pollinator attraction<br />

and defense. Curr. Opin. Pl. Biol. 5: 237–242.<br />

Prashar A, Hili P, Veness RG, Evans CE. 2003. Antimicrobial action of palmarosa oil (Cymbopogon martinii)<br />

on Saccharomyces cerevisiae. Phytochemistry 63: 569–575.<br />

Predieri S, Rapparini F. 2007. Terpene emission in tissue culture. Pl. Cell Tiss. Org. Cult. 91: 87–95.<br />

Quiala E, Barbon R, Jimenez E, DeFeria M, Chavez M, Capote A, Perez N. 2006. Biomass production of<br />

Cymbopogon citrates (D.C.) Stapf. A medicinal plant, in temporary immersion system. In Vitro Cell. Dev.<br />

Biol. Plant 42: 298–300.<br />

Rajesh D, Stenzel R, Howard S. 2003. Perillyl alcohol as a radio/chemo-sensitizer in malignant glioma. J. Biol.<br />

Chem. 278: 35968–35978.<br />

Robbins SRJ. 1983. Selected Markets for the <strong>Essential</strong> <strong>Oil</strong>s of Lemongrass, Citronella and Eucalyptus. Tropical<br />

Products. Institute Publication, London.<br />

Sangwan NS, Naqvi AA, Sangwan RS. 2000. A novel chemotype of aromatic grass Cymbopogon. In: Kumar S,<br />

Kukreja AK, Dwivedi S, Singh AK (Eds.), Proc. Of the National Seminar on Research and Development<br />

in Aromatic Plants: Current Trends in Biology, Uses, Production and Marketing of <strong>Essential</strong> <strong>Oil</strong>s,<br />

pp. 483, 484. J. Med. Pl. Sci. Vol. 22 (Suppl. 1B), CIMAP Publication, Lucknow, India.<br />

Sangwan RS, Sangwan NS. 2000. Metabolic and molecular analysis of chemotypic diversity in Cymbopogon.<br />

In: Kumar S, Dwivedi S, Kukreja AK, Sharma JR, Bagchi, GD (Eds.), Cymbopogon: The Aromatic<br />

Grass: A Monograph, pp. 223–247. CIMAP (CSIR) Publication, Lucknow, India.<br />

Schijlen EG, Ric de Vos CH, Van Tunen AJ, Bovy AG. 2004. Modification of flavonoid biosynthesis in crop<br />

plants. Phytochemistry 65: 2631–2648.<br />

Shanker S, Ajaykumar PV, Sangwan NS, Kumar S, Sangwan RS. 1999. <strong>Essential</strong> oil gland number and ultrastructure<br />

during Mentha arvensis leaf ontogeny. Biol. Plant. 42: 379–387.<br />

Sharkey T, Yeh S. 2001. Isoprene emission from plants. Annual Rev. Pl. Physiol. Pl. Mol. Biol. 52: 407–436.<br />

Sharma JR, Ram RS. 2000. Genetics and genotype improvement of Cymbopogon species. In: Kumar S,<br />

Dwivedi S, Kukreja AK, Sharma JR, Bagchi, GD (Eds.), Cymbopogon: The Aromatic Grass: A Monograph,<br />

pp. 85–128. CIMAP (CSIR) Publication, Lucknow, India.<br />

Shasany AK, Lal RK, Darokar MP, Patra NK, Garg A, Kumar S, Khanuja SPS. 2000. Phenotypic and RAPD<br />

diversity among Cymbopogon winterianus Jowitt. Accessions in relation to Cymbopogon nardus Rendle.<br />

Genet. Resour. Crop. Evol. 47: 553–559.<br />

Shasany AK, Shukla AK, Khanuja SPS. 2007. Medicinal and aromatic plants. In: Kole C (Ed.), Genome Mapping<br />

and Molecular Breeding in Plants, Vol. 6—Technical Crops, pp. 175–196. Springer-Verlag, Berlin.


Biotechnological Studies in Cymbopogons 149<br />

Shen T, Wan W, Yuan H. 2007. Secondary metabolites from Cymbopogon opobalsamum and their antiproliferative<br />

effect on human prostrate cancer cells. Phytochemistry 68: 1331–1337.<br />

Sobti SN, Verma V, Rao BL. 1982. Scope for development of new cultivars of Cymbopogon as a source of<br />

terpene chemicals. In: Atal CK, Kapur BM (Eds.), Cultivation and Utilization of Aromatic Plants,<br />

pp. 302–307. Regional Research Laboratory (CSIR) Jammu, India.<br />

Soenarko S. 1977. The genus Cymbopogon Sprengel (Gramineae). Reinwardtia 9: 225–375.<br />

Sreenath HL, Jagadishchandra KS. 1980. Callus induction and growth in two varieties of Cymbopogon nardus<br />

(L) Rendle. Curr. Sci. 49: 437, 438.<br />

Sreenath HL, Jagadishchandra KS. 1989. Somatic embryogenesis and plant regeneration from inflorescence<br />

culture of Java citronella (Cymbopogon winterianus Jowitt). Ann Bot. 64: 211–215.<br />

Sreenath HL, Jagadishchandra KS. 1991. Cymbopogon Spreng (Aromatic <strong>Grasses</strong>): In vitro culture, regeneration<br />

and the production of essential oils. In: Bajaj YPS (Ed.), Biotechnology in Agriculture and Forestry,<br />

Vol. 15 Medicinal and aromatic plants III, pp. 211–236. Springer-Verlag, Berlin, Heidelberg.<br />

Stapf O. 1906. The oil grasses of India and Ceylon. Kew Bull. 8: 297–463.<br />

Tapia A, Georgirev M, Bley T. 2007. Free radical scavengers from Cymbopogon citrates (D.C.) Stapf. Plants<br />

cultivated in bioreactor of the temporary immersion principle. Z. Naturforsch. Sect. C J Biosci. 62:<br />

447–457.<br />

Thapa RK, Bradu BL, Vashit VN, Atal CK. 1971. Screening of Cymbopogon species for useful constituents.<br />

Flavour India 2: 49–51.<br />

Tholl D, Boland W, Hansel A, Loreto F, Rose USR. 2006. Practical approaches to plant volatile analysis. The<br />

Plant J. 45: 540–560.<br />

Tisserat B, Vaughn SF. 2008. Growth, morphogenesis and essential oil production in Mentha spicata L. plantlets<br />

in vitro. In Vitro Cell. Dev. Biol.-Plant 44: 40–50.<br />

Trapp SC, Croteau R. 2001. Genomic organization of plant terpene synthases and molecular evolutionary<br />

implications. Genetics 158: 811–832.<br />

Tripathy SK, Kole C, Lenke PC. 2007. Isolation of some useful mutants in citronella. Orisa J. Hort. 35:<br />

100–102.<br />

Tyo ET, Alper HS, Stephanopoulos G. 2007. Expanding the metabolic engineering toolbox: more option to<br />

engineer cells. Trends Biotechnol. 25: 132–137.<br />

Vasil IK. 2005. The story of transgenic cereals: The challenge, the debate and the solution—A historical perspective.<br />

In Vitro Cell. Dev. Biol.- Plant 41: 577–583.<br />

Verpoorte R, Alfermann AW. 2000. Metabolic Engineering of Plant Secondary Metabolism. Kluwer,<br />

Dordrecht, Netherlands.<br />

Verpoorte R, Memelink J. 2002. Engineering secondary metabolite production in plants. Curr. Opin. Biotechnol.<br />

13: 181–187.<br />

Verpoorte R, van der Heijden R, Memelink J. 2000. Engineering the plant cell factory for secondary metabolite<br />

production. Transgenic Res. 9: 323–343.<br />

Wallaart TE, Bouwmeester HJ, Hille J, Poppinga L, Maijers NCA. 2001. Amorpha-4,11-diene synthase:<br />

Cloning and functional expression of a key enzyme in the biosynthetic pathway of the novel antimalarial<br />

drug artemisinin. Planta 212: 460–465.<br />

Wannissorn B, Jarikasem S, Siriwangchai T, Thubthimthed S. 2005. Antibacterial properties of essential oils<br />

from Thai medicinal plants. Fitoterapia 76: 233–236.<br />

Williams A, Barry B. 2004. Penetration enhancers. Adv. Drug Delivery Res. 56: 603–618.<br />

Xu Y, Crouch JH. 2008. Marker-assisted selection in plant breeding: From publication to practice. Crop Sci.<br />

48: 391–407.<br />

Yadav U, Mathur AK, Sangwan NS. 2000. In vitro calli-clone regeneration to enhance qualitative variability<br />

in essential oil of a wild Cymbopogon species. In: Kumar S, Kukreja AK, Dwivedi S, Singh AK<br />

(Eds.), Proc. of the National Seminar on Research and Development in Aromatic Plants: Current Trends<br />

in Biology, Uses, Production and Marketing of <strong>Essential</strong> <strong>Oil</strong>s, pp. 485–487. J. Med. Pl. Sci. Vol. 22<br />

(Suppl. 1B), CIMAP Publication, Lucknow, India.<br />

Yonekura-Sakakibara K, Saito K. 2006. Genetically modified plants for the promotion of human health.<br />

Biotechnol. Lett. 28: 1983–1991.<br />

Zheng Z, Zhou S, Zhang J, Qing GJ, Xu GZ, Guilin X. 2007. Study on the in vitro culture of African lemongrass.<br />

Acta Agric. Shanghai 23: 61–64.<br />

Zutchi NL. 1982. <strong>Essential</strong> oils—isolates and semi-synthetics. In: Atal CK, Kapur BM (Eds.), Cultivation and<br />

Utilization of Aromatic Plants, pp. 38, 39. Regional Research Laboratory (CSIR), Jammu Publication.


5<br />

contents<br />

The Trade in Commercially<br />

Important Cymbopogon <strong>Oil</strong>s<br />

Rakesh Tiwari<br />

5.1 Introduction .......................................................................................................................... 151<br />

5.2 Citronella <strong>Oil</strong> ........................................................................................................................ 151<br />

5.2.1 World Market, Demand, and Production .................................................................. 152<br />

5.2.2 World Demand .......................................................................................................... 153<br />

5.3 Lemongrass <strong>Oil</strong> .................................................................................................................... 153<br />

5.3.1 World Production and Demand ................................................................................ 159<br />

5.3.2 Production ................................................................................................................. 159<br />

5.4 Palmarosa <strong>Oil</strong> ....................................................................................................................... 160<br />

5.4.1 Production ................................................................................................................. 163<br />

References ...................................................................................................................................... 165<br />

5.1 IntroductIon<br />

Among the various species of cymbopogons numbering approximately 120, only four are of<br />

commercial importance, namely, Cymbopogon winterianus, C. nardus (citronella), C. flexuosus<br />

(lemongrass), and C. martinii var. motia (palmarosa). <strong>Essential</strong> oils extracted from mainly leaves<br />

and aboveground parts of these plants are valued commercially and traded globally (Anonymous<br />

1986). Accurate production, trade, and export and import figures from third world countries, and<br />

even some Western countries, are obscure (Lawrence 1985, 1986, 1993). It has been mentioned<br />

in some old references that reliable trade data are available for oils produced in the United States<br />

only (Simon 1990). However, export and import data from India are available up to country<br />

level. The statistics for export and import data for India are collected from available entry points,<br />

which include seaports, airports, and land routes. The export data, thus, is inclusive of both<br />

exports and re-exports.<br />

5.2 cItronella oIl<br />

Citronella oil is an industrially important essential oil obtained from leaves and stems of C. winterianus<br />

and C. nardus. The oil is regarded as one of the 20 most important essential oils (Lawrence<br />

1993) found in the world trade. Citronella oil is an important source of perfumery chemicals such as<br />

citronellol, citronellal, and geraniol, which are widely used in perfumery, soaps, detergents, industrial<br />

polishes, cleaning compounds, and other industrial products (Anonymous 1986; Lawless 1995). In the<br />

trade, citronella oil is classified into two chemotypes: Ceylon-type citronella oil and Java-type citronella<br />

oils. Ceylon-type citronella oil is extracted from C. nardus Rendle, and Java-type citronella oil<br />

is obtained from C. winterianus Jowitt (Torres and Tio [2003]). The name C. winterianus was given<br />

151


152 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

to this selected variety to commemorate Winter, an important oil distiller of Ceylon (now Sri Lanka),<br />

who first cultivated and distilled the Maha Pangeri type of citronella in Ceylon (Chang 2007).<br />

The main difference between the Ceylon and Java types of oil is the relative proportion of geraniol<br />

and citronellal. Java-type citronella oil is characterized by a high proportion of geraniol (11%–13%)<br />

and citronellal (32%–45%), making it an important source of derivatives such as citronellol and<br />

hydroxycitronellal, which are extensively used in compounding high-grade perfumes. Other major<br />

constituents of the oil are geranyl acetate (3%–8%) and limonene (1%–4%). The chemistry of the oil<br />

is discussed in another chapter 2 of this book.<br />

Ceylon-type citronella oil contains a relatively low proportion of geraniol (18%–20%) and citronellal<br />

(5%–15%), and is mainly used as such in lower-grade products. Unlike Java-type oil, it is<br />

rarely used for the extraction of derivatives. The other major constituents of Ceylon-type oil are<br />

limonene (9%–11%), methyl isoeugenol (7%–11%), and citronellol (6%–8%).<br />

Citronella oil (both Ceylon and Java types) is also a renowned plant-based insect repellent, and<br />

has been registered for use in the United States since 1948 as “McKesson’s <strong>Oil</strong> of Citronella” for<br />

human application to repel gnats and mosquitoes. The U.S. Environment Protection Agency (EPA)<br />

considers oil of citronella as a biopesticide with a nontoxic mode of action (Anonymous 1997, 2001,<br />

2004, 2006, 2007; Trongtokit et al. 2005). It is registered as an insect repellent, feeding depressant,<br />

and animal repellent. Research also shows that citronella oil has strong antifungal properties and is<br />

effective in calming barking dogs.<br />

5.2.1 wo r l d Ma r k e t, de M a n d, a n d Pr o d u C t i o n<br />

The data on world market size, market price, world production, global area under cultivation,<br />

number of cultivators engaged, etc., are not easily available due to lack of reliable documentation<br />

for such items. The problem is further aggravated because of the lack of scientific documentation<br />

in many of the major producer countries. The current world production of citronella<br />

oil is estimated at 5000 t, valued at about 20 million USD. The majority of the oil is Java type,<br />

with Indonesia as the major supplier in commercial quantities (Robbins 1983). Production of<br />

Ceylon-type oil is restricted to Sri Lanka, where it peaks out to approximately 200 t (Oyen<br />

and Nguyen 1999). Major producers of citronella oil are China and Indonesia, which account<br />

for more than 40% of the world production. The oil is also produced in Taiwan, Guatemala,<br />

Honduras, Brazil, Sri Lanka, India, Argentina, Ecuador, Jamaica, Madagascar, Mexico, and<br />

South Africa (Lawrence 1985). Citronella oil is used in a variety of products in India. India<br />

was a net importing country 60 years ago, but currently produces approximately 600 t of the oil<br />

(Singh et al. 2000).<br />

The leading exporters of citronella oil are China, Indonesia, Taiwan, Guatemala, Sri Lanka,<br />

Argentina, and Brazil. Officially, Indonesia is the leading exporter of oil, probably because statistics<br />

on China’s export volume are not available. European countries such as France and United Kingdom<br />

play an important role in the world’s export of citronella oil; presumably a significant proportion<br />

of their import is being re-exported. The same probably is true of Singapore, as this country has a<br />

considerable role as an entry port in the citronella oil trade.<br />

Citronella oil has been witnessing demand and price fluctuations due to proliferation of inexpensive<br />

synthetic isolates in the market (Robbins 1983). Earlier, Java citronella oil was the most<br />

important source of geraniol and citronellal, but the advent of pinene chemistry and production of<br />

these isolates in reasonable quantity reduced the demand and reliance for Java-type citronella oil.<br />

Further, due to its rich citronellal content, oil from Eucalyptus citriodora has become another major<br />

competitor of citronella oil in the world market (Anonymous 1986). Thus, the overall demand for citronella<br />

oil has been affected by synthetic isolates produced from turpentine oil and E. citriodora oil.<br />

These isolates and substitutes are generally cheaper than citronella oil, making them the preferred


The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s 153<br />

choice when price is the main criterion. However, natural citronella oil and its derivatives are still the<br />

preferred choice of the perfumery industry, mainly because of its unique olfactory and stable properties,<br />

which are vital in blending perfumes and compounding industrially important essences.<br />

The state of Assam leads in the production of citronella oil in India owing to climatic conditions<br />

suited for the cultivation of the crop. The area under cultivation is around 10,000 hectares (ha).<br />

Export and import of citronella oil in India are well documented by Ministry of Commerce and<br />

Industry publications (Anonymous 2003–2007a, 2003–2007b).<br />

5.2.2 wo r l d de M a n d<br />

Earlier, there were no official data available on the current world demand for citronella oil.<br />

International Trade Centre has published data on citronella-oil-importing countries in 1981, which<br />

was used in predicting possible trends for the market of citronella oil. However, the situation has<br />

improved significantly since then.<br />

The United States of America is the world’s largest importer of citronella oil, followed by<br />

European countries, namely, France, United Kingdom, Germany, and the Netherlands. The European<br />

countries are the major trading hub for Java-type oil because of the presence and proximity<br />

of the world-famous perfumery industry, notably in France and Germany. In the Asian region, the<br />

largest importers of citronella oil are Japan and Hong Kong. Hong Kong does a lot of re-export of<br />

citronella oil to the Philippines.<br />

The consolidated export–import data of citronella oil with respect to India is given under<br />

Table 5.1, and countrywise breakup of exports and imports is reproduced in Tables 5.2 and 5.3.<br />

5.3 lemongrass oIl<br />

table 5.1<br />

consolidated export and Import Figures of citronella<br />

oil for India<br />

year<br />

citronella oil<br />

exports Imports<br />

Quantity (t) Value (usd) Quantity (t) Value (usd)<br />

2007 617.502 7,490,027 270.787 1,669,257<br />

2006 440.065 5,322,190 189.462 1,013,885<br />

2005 308.043 2,325,885 115.888 863,690<br />

2004 100.179 1,418,954 122.765 724,823<br />

2003 72.200 616,819 42.558 502,509<br />

2002 33.477 174,241 76.087 574,743<br />

Source: Monthly Statistics of Foreign Trade of India, Vol. 1 and 2.<br />

Lemongrass oil is obtained by steam distillation of leaves and flowering tops of C. flexuosus and<br />

C. citratus. The oil is a commercially important commodity with antifungal properties. The name<br />

lemongrass is attributed to the lemon-like odor of the essential oil, which is due to high citral<br />

content. Two chemotypes of lemongrass oil are well recognized in the trade, namely, East Indian<br />

and West Indian lemongrass oil. East Indian lemongrass, obtained from C. flexuosus, is also called<br />

Cochin grass or Malabar grass. It is native to Cambodia, India, Sri Lanka, Burma, and Thailand.


154 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 5.2<br />

countrywise Import of citronella oil by India<br />

2002 2003 2004 2005 2006 2007<br />

Qty Value Qty Value Qty Value Qty Value Qty Value Qty Value<br />

country (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd)<br />

Australia — — — — 20 395 660 9,759 110 4,556 252 52,209<br />

Austria 125 3,191 — — — — — — — — 115 4,416<br />

Brazil — — 8,335 50,303 13 218 6,780 29,218 2 53 8,482 62,888<br />

Canada 1,000 3,704 — — 30 341 — — — — — —<br />

China 7,302 38,388 2,340 18,740 79,118 335,885 45,206 231,737 117,649 445,273 193,650 881,181<br />

France 1,473 27,250 6,523 64,726 1,315 63,728 4,482 97,189 3,646 55,110 678 27,290<br />

Germany 2,122 17,041 2,650 29,072 820 27,298 843 29,634 2,633 48,534 6,599 71,771<br />

Hong Kong — — — — — — — — 520 5,606 — —<br />

Hungary — — 25 1,205 — — — — — — — —<br />

Indonesia 2,090 67,914 105 2,636 1,650 6,120 20,540 83,200 27,520 103,900 1,650 20,409<br />

Ireland — — — — — — — 646 30 180<br />

Israel 165 1,691 20 258 5,352 22,455 2,118 21,682 120 3,887<br />

Italy 870 6,374 13 579 205 8,530 380 13,612 1,027 30,818 1,151 33,867<br />

Japan 596 16,888 111 507 — — — — — — 2 12<br />

Malaysia — — — — 10 75 — — 4 252 — —<br />

Nepal 24,498 33,624 4,000 20,592 17,200 47,655 7,730 87,136 2,400 12,121 1,445 15,493


The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s 155<br />

Netherlands 40 1,192 — — 1 48 25 504 — — 39 1,297<br />

Paraguay — — — — 2,090 21,280 — — 570 5,504 2,000 15,891<br />

Poland 815 1,189 138 1,678 713 4,561 40 2,204 2 374<br />

Singapore 2,291 10,510 8,865 120,551 2,300 14,072 175 1,343 4,307 35,610 795 24,131<br />

Slovenia — — — — 75 5087 — — — — — —<br />

Spain 545 1,690 1,506 25,302 1,100 31,386 4,944 21,130 11,401 46,841 21,189 98,421<br />

Sri Lanka — — — — 280 4,164 — — — — — —<br />

Swaziland — — — — — — — — — — 241 3,458<br />

Switzerland 840 7,840 — — 1,280 3,044 25 1,512 480 7,226 386 8,950<br />

Tanzania — — — — — — — — 200 13,087 1,249 77,109<br />

Thailand 42 1137 — — — — — — — — 95 3,171<br />

UAE — — — — 202 1,532 52 529 2 23 — —<br />

U.K. 2,776 39,089 1,247 32,498 1,066 35,243 1,789 93,782 1,496 36,255 3,931 52,479<br />

U.S. 25,712 275,506 9,500 130,291 7,201 84,528 15,826 123,247 13,625 153,526 24,716 196,726<br />

Uruguay — — — — 1,199 2,835 — — — — — —<br />

Unspecified — — — — 100 7,226 — — — — — —<br />

Vietnam 3,600 21,714 5,040 4,060 — — 3,600 13,269 1,800 7,206 2,000 13,827<br />

Total 76,087 574,743 42,558 502,509 122,765 724,823 115,888 863,690 189,462 1,013,885 270,787 1,669,257<br />

Source: Monthly Statistics of Foreign Trade of India, Vol. 1 and 2.


156 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 5.3<br />

countrywise export of citronella oil by India<br />

2002 2003 2004 2005 2006 2007<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

country<br />

Afghanistan — — — — — — — — — — 100 456<br />

Angola — — — — — — — — — — 2,000 42,912<br />

Australia 67 255 — — 300 1,507 951 9,193 630 11,054 1,840 24,376<br />

Austria 16 5,033 16 905 — — 700 1,740 — — 300 2,066<br />

Bahrain — — — — — — — — 178 15,844 1,450 55,539<br />

Bangladesh 2,500 5,574 — — — — 1,800 4,108 4,500 7,777 5,410 10,029<br />

Belgium — — 50 2,605 80 551 180 1,328 — — — —<br />

Bulgaria — — — — 20 1,065 1,530 15,913 — — — —<br />

Cambodia — — — — — — — — 2,080 12,737 25 5,978<br />

Canada 40 1,904 9,021 24,539 3,261 26,597 189 3,656 231 2,261 230 10,001<br />

Chile — — — — 700 5,753<br />

China — — 14,450 95,189 — — 11 147 45,200 430,406 76,874 1,088,379<br />

Chinese Taipei — — 107 11,569 — — — — — — — —<br />

Colombia — — — — — — — — 360 5,819 117 3,352<br />

Congo P REP — — — — 5,600 26,342 — — — — 26,773 79,226<br />

Costa Rica — — — — — — — — 40 867 — —<br />

Czech Republic — — — — 11 1,153 — — 1,005 17,847<br />

Denmark 720 3,841 — — 30 827 50 1,409 — — — —<br />

Djibouti — — — — — — — — 400 12,101 — —<br />

Egypt — — — — 45 998 75 1,458 70 1,342 — —<br />

Ethiopia — — — — 3,060 7,268 1,620 8,533 — — — —<br />

Fiji — — — — — — — — — — 17 746<br />

Finland — — — — — — — — — — 5,860 100,293<br />

France 1,407 22,365 1,257 22,476 1,180 13,597 8,225 152,677 5,376 117,235 2,908 39,223<br />

Gambia — — — — — — 56 3068 — — — —<br />

Germany 35 2,395 3,280 46,450 13,330 84,528 500 36,404 4,560 59,299 11,235 316,242<br />

Ghana — — — — — — 16,850 39,311 — — 11,600 108,785<br />

Greece — — — — 600 3,523 — — — — — —


The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s 157<br />

Guyana — — — — — — — — — — 150 2,023<br />

Hong Kong — — — — 145 4,066 7,500 251,457 40 1,499 28,895 347,037<br />

Hungary — — 100 795 — — — — 26 1,963 565 12,063<br />

Iceland — — — — — — — — — — 31 2,230<br />

Indonesia — — 1,000 10,937 10 358 — — 190 2,960<br />

Iran — — — — 50 1,163 50 1,317 — — — —<br />

Ireland — — — — — — 140 2,160 30 254 180 1,343<br />

Israel — — 1,000 5,863 20 213 — — — — 685 12,648<br />

Italy — — 1,854 26,273 2 177 14,543 98,830 30 1,587 24 772<br />

Jamaica — — — — — — — — — — 150 2,159<br />

Japan — — — — 5,102 438,600 4,320 36,859 4,430 64,159 4,071 124,168<br />

Jordan — — — — — — — — — — 50 652<br />

Kenya — — — — 15,40 76,777 34,783 199,318 31,819 263,200 40,750 337,873<br />

Korea DP RP — — — — — — 1,000 6,952 — — 228 4,907<br />

Korea RP — — — — 1,500 7,599 — — 1,566 23,138 215 6,015<br />

Kuwait — — 20 1,429 — — 298 9,933 299 10,420 118 11,164<br />

Latvia — — — — — — — — 333 5,169 101 4,836<br />

Malaysia 1,496 17,917 100 245 1,324 7,629 4,082 13,539 265 4,834 4,039 24,910<br />

Maldives — 40 1,192 430 4,285 100 511 — — — —<br />

Mauritius — 85 1,642 65 1,107 852 6,081 3,540 15,440 110 3,701<br />

Mexico 501 6,149 — — — — — — 50 375 — —<br />

Mozambique — — — — — — — — 720 1686 170 3503<br />

Myanmar 8,000 20,123 1,100 4,126 5,000 54,727 2,000 30,853 — — — —<br />

Nepal 359 5,612 10 61 1,061 6,672 4,421 30,561 5,474 36,963<br />

Netherlands 804 7,843 4,000 33,633 865 21,158 169 13,812 1,100 14,873 10,141 171,511<br />

New Caledonia — — — — — — — — — — 100 143<br />

New Zealand — — 9 395 — — 2 96 670 971 150 5,751<br />

Nigeria — — 24 570 4,182 52,693 — — 105,891 288,324 90,118 347,096<br />

Oman 2,000 2,953 — — 65 1,058 — — 200 1,489 55 1,247<br />

Pakistan — — — — 25 1,955 — — — — 145 7,607<br />

Panama Republic — — — — 42 3,034 171 9,072 — — — —<br />

Paraguay — — — — — — — — — — 4,000 86,736<br />

Peru — — — — — — 1,000 26,056 — — — —<br />

Philippines — — — — — — — — — — — —<br />

(continued on next page)


158 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 5.3 (continued)<br />

countrywise export of citronella oil by India<br />

2002 2003 2004 2005 2006 2007<br />

Qty Value Qty Value Qty Value Qty Value Qty Value Qty Value<br />

country<br />

(kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd)<br />

Poland — — — — — — — — 445 7,269 70 207<br />

Reunion — — — — — — — — — — — —<br />

Romania — — — — — — 110 7,843 — — — —<br />

Russia — — — — — — — — 1,100 5,714 1,400 13,883<br />

Rwanda — — — — — — 500 1,381 — — — —<br />

Saudi Arabia — — — — 1,925 6,036 1,170 24,321 2,605 66,550 1,287 45,409<br />

Singapore 6,250 13,724 1,408 24,902 3,358 30,928 1,3111 88,392 710 21,360 30,294 292,650<br />

South Africa 12 171 600 3,761 460 10,357 1,870 30,024 2,500 50,539 2,462 50,139<br />

Spain 600 17,784 16,346 112,975 193 2,443 1,654 118,038 4,570 60,387 9,400 128,492<br />

Sri Lanka 450 5,575 6,540 39,485 65 2,514 2,835 14,553 2,370 8,777 86 4,048<br />

Sudan — — — — — — 11,500 15,716 300 3,708 10 602<br />

Switzerland — — — — 190 980 1,252 15,471 750 63,819 2,000 49,887<br />

Syria — — — — — — — — — — 100 2,482<br />

Taiwan — — — — 120 3,597 151 4,028 11 381 6 778<br />

Tanzania — — — — 34,290 186,623 84,913 348,669 59,200 366,188 35,530 171,104<br />

Thailand 1,500 2,783 500 1,132 — — — — 2,700 37,477 43 5,749<br />

Tunisia — — — — — — — — — — 934 107,816<br />

Turkey — — — — — — — — 25 245 325 8,644<br />

Uganda 526 1,498 — — 460 2,500 600 5,215 20 419 1,080 12,586<br />

UAE 4,960 13,923 4,591 20,912 6,125 39,496 4,003 44,813 9,716 996,468 6,206 137,707<br />

U.K. 458 25,668 500 23,242 26,104 225,236 26,562 444,360 64,912 756,278<br />

U.S. 1,593 22,431 3,885 91,539 3,790 257,425 46,140 368,914 98,430 1,735,440 121,872 2,231,351<br />

Yemen — — — — — — — — 8,795 43,064 1,020 4,929<br />

Zambia — — — — — — 3,160 7,339 — — — —<br />

Unspecified — — — — — — 4,000 8,288 — — — —<br />

Total 33,477 174,241 72,200 616,819 100,179 1,418,954 308,043 2,325,885 440,065 5,322,190 617,502 7,490,027<br />

Source: Monthly Statistics of Foreign Trade of India, Vol. 1 and 2.


The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s 159<br />

West Indian lemongrass, obtained from C. citratus, is assumed to be native to Malaysia. While both<br />

can be used interchangeably, C. citratus is more suited for cooking. In India, C. citratus is used<br />

both as a medicinal herb and in perfumes. The main difference between these two types of oil is<br />

the relative percentage of citral. East Indian lemongrass oil is higher in citral content, which ranges<br />

up to 90%. West Indian lemongrass oil has lower citral content and lower solubility in alcohol. The<br />

lower solubility is attributed to the presence of myrcene, which polymerizes on exposure to air and<br />

light. In the trade, West Indian oil is considered inferior to East Indian oil and has meager trade. The<br />

name West Indian lemongrass oil is a misnomer in the sense that this grass is not indigenous to West<br />

Indies and, currently, the production of lemongrass in West Indies is very low (Husain 1993).<br />

The lemongrass oil finds widespread use in soap, perfumery, cosmetics, and the beverages industry.<br />

Additionally, it is an important natural source of citral, which is an important starting material<br />

for the synthesis of beta ionone. Beta ionone is further used for synthesis of a number of aroma<br />

chemicals widely used in perfumery and cosmetics. Beta ionone is also used to produce vitamin A.<br />

Due to its distinct lemony flavor, the herb itself finds use in imparting citrus flavor in fresh, chopped<br />

and sliced, or dried and powdered forms. Lemongrass is commonly used in teas, soups, and curries.<br />

In India, the record of medicinal use of lemongrass oil dates back to more than 2000 years, though<br />

its distillation started only in 1890.<br />

5.3.1 wo r l d Pr o d u C t i o n a n d de M a n d<br />

During the early 1950s, India produced over 1800 t/annum of lemongrass oil and held monopoly<br />

both in production and world trade. This has changed, as Guatemala, China, Mexico, Bangladesh,<br />

etc., have developed large-scale cultivation of lemongrass.<br />

Currently, the world production of oil of lemongrass ranges between 800 and 1300 t/annum<br />

(Singh et al. 2000). However, another 600 t of a substitute oil, that is, Litsea cubeba (rich in citral),<br />

is exported by China, which limits the scope for any faster growth in export trade of lemongrass oil<br />

(Lawrence 1985). Synthetic citral available at a relatively lower rate competes with lemongrass<br />

oil and natural citral in the market.<br />

5.3.2 Pr o d u C t i o n<br />

India is a major producer of the oil, and about 80% of the produce is exported, mainly to West<br />

Europe, United States, and Japan. The situation changed during the Second World War because of<br />

problems of production and supply logistics. Consequently, Guatemala, Haiti, Madagascar, Zaire,<br />

Cambodia, Vietnam, and Laos started producing lemongrass oil to meet the demand, and continue<br />

to do so. Guatemala, China, Mexico, Bangladesh, etc., have developed its cultivation over large<br />

areas. India has moved to systematic cultivation of lemongrass; earlier, it was mainly collected from<br />

wild forests. East Indian lemongrass oil is preferred due to its high citral content. Though exact<br />

information on production and trade from the producing countries is not available, the current world<br />

production of the oil is estimated at 1300 t/annum, with India contributing to the tune of 350 t. The<br />

consolidated export–import data of lemongrass oil with respect to India is given in Table 5.4, and<br />

the countrywise breakup of the export and import figures is presented in Tables 5.5 and 5.6.<br />

In India, the crop grows in an area of about 3000 ha, largely in the states of Kerala, Karnataka,<br />

Tamil Nadu, Maharashtra Uttar Pradesh and Assam. Supply of cheaper citral by China produced<br />

from alternative sources results in periodic fluctuation in demand for lemongrass oil. However, in<br />

general, the global demand for the oil is robust.<br />

Currently, other major lemongrass-oil-producing countries are Brazil, China, Guatemala, Haiti,<br />

Nepal, Russia, and Sri Lanka.


160 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 5.5<br />

countrywise Import of lemongrass by India<br />

country<br />

Qty<br />

(kg)<br />

5.4 Palmarosa oIl<br />

2002 2003 2004 2005 2006 2007<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Palmarosa oil of commerce is obtained by steam distillation of ground, freshly harvested or partially<br />

dried flowering shoot biomass leaves and stems of C. martinii var. motia. The oil is valued<br />

for its aroma chemical geraniol (75%–90%), which is separated through fractional distillation of<br />

the essential oil. Geraniol is widely used in flavor, fragrance, and pharmaceutical industries. The<br />

oil is extensively used as perfumery raw material in soaps, floral rose-like perfumes, cosmetics<br />

preparations, and in the manufacture of mosquito repellent products. It is used for flavoring tobacco<br />

products, foods, and nonalcoholic beverages. In medicine, the volatile oil is used as a remedy for<br />

lumbago, stiff joints, skin diseases, and for bilious complaints. Another variety, C. martinii var.<br />

sofia, which has lower geraniol content (


The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s 161<br />

table 5.6<br />

countrywise export of lemongrass oil by India<br />

2002 2003 2004 2005 2006 2007<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

country<br />

Australia 5,636 45,685 1,980 21,291 1,760 23,673 1,600 16,055 3,520 36,946 5,040 60,926<br />

Austria 20 336 — — — — — — — — — —<br />

Belgium — — — — — — — — 540 5,235 1,080 15,118<br />

Bosnia — — — — 22 1,162 — — — — — —<br />

Brazil — — — — — — — — 600 6,850 400 4,531<br />

Canada 400 4,031 415 4,323 400 3,986 35 2,149 800 4,935 600 6,514<br />

China — — 10 191 — — — — — — 95 1,464<br />

Denmark — — — — — — 420 9,623 100 679 — —<br />

Ecuador — — — — — — 2 34 10 161 — —<br />

Egypt — — 20 506 — — — — 100 796 100 791<br />

French South and — — — — — — — — — — 1 9<br />

Antarctic Territories<br />

Fiji — — — — — — 30 126 — — — —<br />

France 6,000 63,216 4,709 49,137 540 5,962 — — 1,982 20,855 5,640 70,816<br />

Germany 4,400 46,962 7,733 91,140 6,360 62,518 3,206 30,524 4,820 50,520 5,250 58,583<br />

Ghana — — 25 218 — — — — — — — —<br />

Greece 750 3,400 — — — — — — — —<br />

Guatemala — — — — — — — — 1,080 11,620 1,260 14,460<br />

Hong Kong — — 10 105 — — — — — — — —<br />

Indonesia — — 25 325 — — — — — — — —<br />

Israel 20 330 45 655 145 2031 20 185 940 8,220 — —<br />

Italy — — 56 4,243 — — — — — — 410 4,667<br />

Japan 1,620 17,380 — — 740 9,047 540 5,033 540 4,740 390 4,338<br />

Kenya 150 2,299 1,135 7,262 — — — — — — 100 1,360<br />

Korea RP 30 471 6 58 — — 1,000 2,110 — — 200 3,025<br />

Maldives — — — — — — — — 43 493 — —<br />

(continued on next page)


162 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 5.6 (continued)<br />

countrywise export of lemongrass oil by India<br />

2002 2003 2004 2005 2006 2007<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

country<br />

Mauritius — — — — 10 150 — — 50 375 — —<br />

Mexico — — — — 360 3,145 — — — — — —<br />

Myanmar — — — — 1,000 30,143 30 346 — — — —<br />

Nepal — — — — — — 100 663 — — — —<br />

Netherlands 180 1,942 — — 5,000 43,671 2,000 16,711 — —<br />

New Zealand — — — — — — — — — — 100 1,043<br />

Oman — — — — 210 1,406 — — — — — —<br />

Philippines — — — — — — — — 150 347 — —<br />

Singapore 1,200 13,542 1,880 20,873 367 4,741 1,960 17,871 3,965 35,172 2,760 29,741<br />

Slovenia 80 848 — — — — — — — — — —<br />

South Africa — — — — 50 632 — — — —<br />

Spain 900 8,890 — — 1,080 10,511 — — 1,190 13,089 — —<br />

Sri Lanka 1,140 10,828 210 2,726 20 176 1,315 10,929 584 6,411 975 11,325<br />

Surinam — — — — 420 5,135 — — — — — —<br />

Switzerland — — 3 114 — — — — 1,080 10,561 6,888 73,801<br />

Taiwan — — — — 30 522 50 485 — — 2,180 24,565<br />

Thailand 20 209 2,150 23,067 500 5,491 625 4,891 220 1,971 601 5,232<br />

Tanzania 50 466 — — — — — —<br />

U.K. 13,575 135,961 12,670 128,514 7,340 70,775 6,800 55,703 18,760 171,513 12,056 135,112<br />

U.S. 10,837 103,136 12,125 118,594 5,025 48,911 13,400 117,976 24,419 232,949 25,148 279,271<br />

Unspecified — — — — 700 122 — — — — — —<br />

Vietnam — — — — — — — — — — 100 438<br />

Yemen — — — — — — — — — — 66 1,392<br />

Total 46,778 457,523 45,387 475,284 27,079 290,072 36,183 319,004 67,493 641,147 71,440 808,521<br />

Source: Anonymous 2003–2007a.


The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s 163<br />

traditional medicine and household uses. The essential oil has a scent similar to that of rose oil, and<br />

hence the name palmarosa.<br />

5.4.1 Pr o d u C t i o n<br />

Historically, the first attempt at its cultivation dates back to 1924 in West Punjab, which now is in<br />

Pakistan. Later, its cultivation in India was started in the early 1950s in Dehra Dun (Uttarakhand).<br />

Currently, the crop is cultivated in several states of India, namely, Uttar Pradesh, Uttarakhand,<br />

Assam, Andhra Pradesh, and Karnataka. The grass has been introduced to the Central American<br />

countries of Guatemala, Honduras, Indonesia, and Brazil (Husain 1993).<br />

The current Indian production is around 100 t of oil per year, with consumption figures of<br />

40–60 t. India enjoyed a virtual monopoly in palmarosa oil production until the 1990s. Indian palmarosa<br />

oil was rated as premium quality oil with 80%–90% geraniol content. Brazil was the second<br />

largest producer of the oil (Lawrence 1985). Later, Brazil, Indonesia, Honduras, and Guatemala<br />

started producing and supplying better quality oil in the world market. With the appearance of<br />

synthetic geraniol, the world market witnessed a weakening in demand for natural palmarosa oil.<br />

The consolidated export–import data of palmarosa oil for India is given in Table 5.7, and countrywise<br />

breakup of the export and import is presented in Tables 5.8 and 5.9. The United States is the<br />

major importer of the oil, followed by Europe (France, Germany, the Netherlands, and the United<br />

Kingdom) and Australia.<br />

table 5.8<br />

country-wise breakup of the Import Figures of Palmarosa oil for India<br />

country<br />

table 5.7<br />

consolidated export and Import Figures of Palmarosa<br />

oil for India<br />

year<br />

Qty<br />

(kg)<br />

2003 2004 2005 2006<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Palmarosa oil<br />

exports Imports<br />

Quantity (t) Value (usd) Quantity (t) Value (usd)<br />

2007 20.684 365,323 0 0<br />

2006 17.883 212,466 0.56 24,852<br />

2005 7.257 79,313 0.027 226<br />

2004 14.795 183,779 0.11 7,949<br />

2003 22.828 318,365 0.5 24,405<br />

2002 11.620 166,449 — —<br />

Source: Anonymous 2003–2007a and b.<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Qty<br />

(kg)<br />

Value<br />

(usd)<br />

Germany 500 24,406 — — — — 200 11,116<br />

Indonesia — — — — — — 360 13,736<br />

UAE — — — — 3 118 — —<br />

U.K. — — 110 7,950 — — — —<br />

U.S. — — — — 24 108 — —<br />

Total 500 24,406 110 7,950 27 226 560 24,852


164 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 5.9<br />

countrywise export of Palmarosa oil by India<br />

2002 2003 2004 2005 2006 2007<br />

Qty Value Qty Value Qty Value Qty Value Qty Value Qty Value<br />

country (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd) (kg) (usd)<br />

Argentina — — — — 10 118 — — — — 20 351<br />

Australia — — 410 5,387 161 2,981 — — — — 25 386<br />

Bahrain — — — — — — — — — — 25 1,469<br />

Canada 1,000 14,358 — — — — — — — — 10 467<br />

France 3,860 55,877 2,338 33,458 2,080 22,070 540 5,838 1,105 16,394 3,740 67,284<br />

Germany 360 4,916 5,220 70,770 — — 485 5,062 1,880 24,615 2,365 37,042<br />

Hong Kong — — — — — — 100 1,798 300 4,963 145 8,408<br />

Ireland — — 360 4,694 — — — — — — 360 6,472<br />

Japan — — — — — — 1,080 10,754 20 506 4 107<br />

Mexico — — — — 50 589 — — — — — —<br />

Netherlands — — 1,800 25,422 4,680 63,232 3,060 32,600 2,880 44,090 2,700 54,114<br />

New Zealand 130 3,542 — — — — — — — — — —<br />

Singapore — — 260 3,372 240 3,419 — — — — 200 2,738<br />

South Africa — — — — — — 10 126 — — — —<br />

Spain 900 11,002 1,260 17,919 1,580 19,526 1,080 12,708 5,287 33,534 680 12,435<br />

Switzerland 540 8,451 720 12,992 — — 720 9,140 1,410 26,128 — —<br />

Taiwan — — — — 20 271 — — — — — —<br />

Tanzania — — — — — — — — — — 30 319<br />

Thailand — — — — — — 2 35 200 4,296 — —<br />

UAE — — — — 24 235 — — — — 500 9,191<br />

U.S. 4,830 68,303 10,460 144,351 5,950 71,338 180 1,252 4,801 57,940 9,880 164,540<br />

Total 11,620 166,449 22,828 318,365 14,795 183,779 7,257 79,313 17,883 212,466 20,684 365,323


The Trade in Commercially Important Cymbopogon <strong>Oil</strong>s 165<br />

reFerences<br />

Anonymous. 1997. U.S. Environmental Protection Agency Fact Sheet. 1997. Prevention, Pesticides and Toxic<br />

Substances (7508W), Re registration Eligibility Decision Sheet EPA-738-F-97-002 (February 1997).<br />

Anonymous. 1986. <strong>Essential</strong> oils and oleoresins: A study of selected producers and major markets. Int. Trade<br />

Centre, UNCTAD/GATT, Geneva.<br />

Anonymous. 2001. WHO International Programme on Chemical Safety: Guidance document for the use of<br />

chemical specific adjustment factors (CSAFs) for interspecies differences and human variability in dose<br />

concentration response assessment. 76 pp. World Health Organization, Geneva.<br />

Anonymous. 2007. Citronella (<strong>Oil</strong> of Citronella (021901) Fact Sheet), U.S. Environmental Protection Agency.<br />

Issued 11/99; Updated October 22, 2007.<br />

Anonymous. 2003–2007a. Monthly Statistics of Foreign Trade of India. Vol. I, Exports and Re-Exports March,<br />

DG CI&S, Ministry of Commerce and Industry, Govt. of India, Kolkata.<br />

Anonymous. 2003–2007b. Monthly Statistics of Foreign Trade of India. Vol. II, Imports, Directorate General of<br />

Commercial Intelligence and Statistics, Ministry of Commerce and Industry, Govt. of India, Kolkata.<br />

Anonymous. 2004. Re-evaluation of citronella oil and related active compounds for use as personal insect<br />

repellants, Proposed Acceptability for Continuing Registration PACR 3004–36, September 17, 2004,<br />

Pest Management Regulatory Agency, Ontario, Canada.<br />

Anonymous. 2006. Report of an independent science panel on citronella oil used as an insect repellent.<br />

March 16, 2006, Canada.<br />

Anonymous 2008. <strong>Essential</strong> oil dictionary, detailed reference guide and herbal encyclopaedia. http://www.<br />

deancoleman.com/essentialref.htm.<br />

Chang Yu S. 2007. Eight MAP species from Malaysia for ICS. Forest Research Institute Malaysia, Workshop<br />

on NFP, 28029 May 2007, Nanchang, China.<br />

Husain A. 1993. <strong>Essential</strong> <strong>Oil</strong> Plants and Their Cultivation. Central Institute of Medicinal and Aromatic Plants,<br />

Lucknow, India.<br />

Lawless J. 1995. The Illustrated Encyclopedia of <strong>Essential</strong> <strong>Oil</strong>s: Complete Guide to the Use of <strong>Oil</strong>s in Aromatherapy<br />

and Herbalism, Thorson, U.K.<br />

Lawrence B. M. 1985. A review of the world production of essential oils (1984). Perfum Flav. 10: 1–16.<br />

Lawrence B. M. 1986. <strong>Essential</strong> oil production: A discussion of influencing factors. In T. H. Parliament and<br />

R. Croteau (Eds.). Biogeneration of Aromas, pp. 363–369. ACS Symposium Series 317 Amer. Chem.<br />

Soc., Washington, DC.<br />

Lawrence B. M. 1993. A planning scheme to evaluate new aromatic plants for the flavour and fragrance industries,<br />

pp. 620–627. In J. Janick and J. E. Simon (Eds.). New Crops, Wiley, New York.<br />

Oyen L. P. A. and Nguyen X. D. 1999. Plant Resources of South East Asia No. 19: <strong>Essential</strong> <strong>Oil</strong> Plants.<br />

Backhuys Publishers, Leiden.<br />

Robbins S. R. J. 1983. Selected markets for essential oil of lemongrass, citronella and eucalyptus. Tropical<br />

Products Research Institute 6171, London.<br />

Rosalinda C. Torres and Barbara D. J. Tio. Citronella oil industry: challenges and breakthroughs.<br />

Simon J. E. 1990. <strong>Essential</strong> oils and culinary herbs. Advances in New Crops, pp. 472–483. In J. Janick and<br />

J. E. Simon (Eds.). Timber Press, Portland.<br />

Singh A. K., Gauniyal A. K., and Virmani O. P. 2000. <strong>Essential</strong> oil of important Cymbopogons: Production<br />

and trade. In Kumar S. et al. (Eds.). Cymbopogon: The Aromatic Grass Monograph. Central Institute of<br />

Medicinal and Aromatic Plants, Lucknow, India.<br />

Smith R. L., Cohen S. M., Doull J., Feron V. Y., Goodman J. I., Marnett L. J., Portoghese P. S., Waddel W. J.,<br />

Wagner B. M., Hall R. L., Higley N. A., Lucas-Gavin C., and Adams T. B. 2005. A procedure evaluation of<br />

natural flavour complexes used as ingredients in food: <strong>Essential</strong> oils. Food Chem. Toxicol. 43: 345–363.<br />

Torres R. C. and Tio B. D. 2001. Citronella (Cymbopogon winterianus) oil industry challenge.<br />

Trongtokit Y., Rongsriyam Y., Komalamisra N., and Apiwathnasorn C. 2005. Comparative repellency of 38<br />

essential oils against mosquito bites. Phytother Res. 19(4): 303–309.


6<br />

contents<br />

6.1 IntroductIon<br />

In Vitro Antimicrobial and<br />

Antioxidant Activities of<br />

Some Cymbopogon Species<br />

Watcharee Khunkitti<br />

6.1 Introduction .......................................................................................................................... 167<br />

6.2 Factors Affecting Antimicrobial Activity of <strong>Essential</strong> <strong>Oil</strong>s ................................................. 168<br />

6.2.1 Solubilizing Agents ................................................................................................... 168<br />

6.2.2 Type of Organism ..................................................................................................... 169<br />

6.2.2.1 Gram-Positive Bacteria .............................................................................. 169<br />

6.2.2.2 Gram-Negative Bacteria ............................................................................ 169<br />

6.2.2.3 Mould and Yeast ........................................................................................ 169<br />

6.2.3 The Correlation between <strong>Oil</strong> Components and Activity .......................................... 169<br />

6.2.4 Methods Commonly Used for Antimicrobial Assessment of <strong>Essential</strong> <strong>Oil</strong> ............. 170<br />

6.2.4.1 The Serial Broth Dilution Method ............................................................. 170<br />

6.2.4.2 The Agar Diffusion Method ...................................................................... 170<br />

6.2.4.3 Vapor Contact Assay .................................................................................. 170<br />

6.3 In Vitro Antioxidant Activity of Some Cymbopogon Species and Their Major<br />

Components .......................................................................................................................... 178<br />

6.3.1 Factors Affecting Antioxidant Activity .................................................................... 178<br />

6.3.1.1 Choice of Oxidizable Substrate and End-Product Evaluation ................... 178<br />

6.3.1.2 Media ......................................................................................................... 178<br />

6.3.1.3 Oxidation Conditions ................................................................................. 179<br />

6.3.1.4 Methods Commonly Used for Testing Antioxidant Activity of<br />

<strong>Essential</strong> <strong>Oil</strong>s ............................................................................................. 180<br />

6.4 Summary .............................................................................................................................. 180<br />

References ...................................................................................................................................... 181<br />

<strong>Essential</strong> oils have been widely used in antimicrobial, antiviral, antiparasitical, insecticidal, medical,<br />

and cosmetic applications (Bakkali et al. 2008). Cymbopogon species are well known as a source<br />

of commercially valuable compounds, such as geraniol, geranyl acetate, citral (neral and geranial),<br />

citronellal, piperitone, eugenol, etc. (Shahi and Tava 1993). Bioactivity of Cymbopogon species<br />

such as lemongrass (C. citratus), Indian lemongrass (C. flexuosus), Indian palmarosa (C. martinii),<br />

Java citronella (C. winterianus), and Ceylon citronella (C. nardus) has been reported. <strong>Essential</strong> oils<br />

from Cymbopogon species and their components are known for their antimicrobial (de Billerbeck<br />

et al. 2001; Pattnaik et al. 1995a; Pattnaik et al. 1995b) and antioxidant activities (Hierro et al. 2004;<br />

Ruberto and Baratta 2000; Lertsatittanakorn et al. 2006). Bioactivity of the same essential oils may<br />

be markedly different when using different strains of the same microorganism and different sources<br />

167


168 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

of essential oils. According to Oussalah et al. (2006), the major constituents of lemongrass oil<br />

are citral (77%) and limonene (8.5%); Indian lemongrass has citral (77%); palmarosa has geraniol<br />

(80%) and gerany1 acetate (8.6%) as main constituents; Java citronella oil contains citronellal (34%),<br />

geraniol (21.5%), and citronellol (11.5%) as major components; and Ceylon citronella oil contains<br />

geraniol (19.1%), limonene (9.9%), and camphene (9.0%). Since the oils are poorly soluble in water,<br />

many factors affect the results of their activities, for example, oil solubilizers, vehicles, method<br />

of testing, etc. Therefore, this chapter reviews the common methods of testing antimicrobial and<br />

antioxidant activities of some Cymbopogon species and their major components, as well as factors<br />

affecting their activities.<br />

6.2 Factors aFFectIng antImIcrobIal actIVIty oF essentIal oIls<br />

6.2.1 so l u b i l i z i nG aG e n t s<br />

The main problem in the study of essential oil bioactivities is that they are poorly soluble in water. In<br />

order to overcome this problem, many authors have used various solvents. Examples of solubilizing<br />

agents used are provided in Table 6.1.<br />

However, interaction between essential oil components and solubilizer has to be taken into consideration.<br />

For example, Tween 20 (polyoxyethylene (2) sorbitan monolaurate) and Tween 80 (polysorbate<br />

80), which are the most commonly used nonionic solublizers, may cause either enhancement<br />

or reduction of antimicrobial activity. Interaction of essential oil with nonionic surfactants could<br />

be the result of either micellar solubilization or complex formation between the two molecules. The<br />

activity of essential oil depends on an adequate concentration of free oil existing in aqueous phase<br />

outside the micelle. At concentrations above critical micelle concentration (CMC) of the surfactants,<br />

antimicrobials solubilized within the micelles do not contribute to the antimicrobial activity,<br />

whereas at low concentrations below CMC, antimicrobial activity increases because of an increase<br />

in bacterial cell permeability to the antimicrobial compound (Russell et al. 1992). Reduction of the<br />

bioactivity of tea tree oil and thyme oil has been reported. It is possible that Tween 80 might inactivate<br />

phenolic compounds in those essential oils (cited in Mann and Markham 1998; Manou et al.<br />

1998). In addition, Hili et al. (1997) reported that the reduction of antimicrobial activity of cinnamon<br />

oil in the presence of dimethylsulphoxide (DMSO) might be due to the partitioning of the oil<br />

between the aqueous phase and DMSO, distancing the oil from cells. This effect, however, did not<br />

table 6.1<br />

essential oil-solubilizing agents<br />

oil-solubilizing<br />

agent<br />

concentration<br />

used organism assay method references<br />

Tween 20 0.001% v/v Fungi Broth microdilution method Devkatte et al. (2005)<br />

Tween 20 0.1% w/v Fungi Agar dilution method Tampieri et al. (2005)<br />

Tween 20 5% v/v Bacteria Agar dilution method Hammer et al. (1999)<br />

Tween 80 1% w/v Bacteria<br />

Yeast<br />

Agar diffusion method<br />

(Paper disk 6 mm)<br />

Jirovetz et al. (2007)<br />

Lertsatithanakorn et al. (2006)<br />

DMSO 0.2% v/v Bacteria Broth microdilution method Ohno et al. (2003)<br />

DMSO 1% v/v Fungi Agar dilution method Inouye et al. (2001)<br />

Ethanol 2%v/v Fungi Broth microdilution method Tullio et al. (2007)<br />

Ethyl acetate Pure Fungi Vapor contact method Inouye et al. (2001)<br />

DMSO +<br />

Tween 80<br />

Ethanol +<br />

Tween 80<br />

10% v/v DMSO +<br />

0.5% v/v Tween 80<br />

5% Ethanol +<br />

5% Tween 80<br />

Bacteria Agar diffusion method<br />

(Paper disk 6 mm)<br />

Prabuseenivasan et al. (2006)<br />

Bacteria<br />

Yeast<br />

Broth microdilution method Unpublished data


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 169<br />

occur when low concentrations (0.15%–0.2% w/v) of bacteriological agar were used as a stabilizer<br />

of the oil–water mixture (Mann and Markham 1998; Remmal et al. 1993).<br />

6.2.2 ty Pe o f or G a n i s M<br />

The action of biocides depends on the type of microorganisms, which is mainly related to their<br />

cell wall structure and the outer membrane arrangement (Kalemba and Kunicka 2003; Russell<br />

et al. 1992).<br />

6.2.2.1 gram-Positive bacteria<br />

Gram-positive bacteria are more sensitive to biocides, particularly essential oil, than Gram-negative<br />

bacteria. Probably the main reason for this difference in sensitivity is the relative composition of the<br />

cell envelope. In general, the cell wall of Gram-positive bacteria is composed basically of peptidoglycan,<br />

which forms a thick, fibrous layer. Many antimicrobial agents much penetrate the outer and<br />

cytoplasm membranes to reach their site of action. The effects of various disinfectants, antiseptics,<br />

and preservatives on Gram-positive bacteria have been well documented. The action of essential<br />

oils against Gram-positive bacteria and fungi appears to be similar. The oil components destroy<br />

the bacterial and fungal cell wall and cytoplasmic membrane, causing a leakage of cytoplasm and<br />

coagulation. They also inhibit the synthesis of DNA, RNA, proteins, and polysaccharides in fungal<br />

and bacterial cells (Himejima and Kubo 1993; Zani et al. 1991).<br />

6.2.2.2 gram-negative bacteria<br />

Gram-negative bacteria, especially Escherichia coli, Klebsiella spp., Proteus spp., Pseudomonas<br />

aeruginosa, and Seratia macescens, appear to be increasingly implicated as hospital pathogens.<br />

Gram-positive bacteria are more sensitive to essential oils than Gram-negative bacteria. Pseudomonas<br />

aeruginosa, for example, is resistant to a wide variety of essential oils due to the hydrophilic<br />

surface of their outer membrane, which is rich in lipopolysaccharide molecules. Thus, essential oil<br />

constituents are unable to penetrate the membrane barrier (Nikaido 1994). <strong>Essential</strong> oils containing<br />

phenolic compounds such as carvarcrol and thymol cause the outer cell membrane damage<br />

(Helander et al. 1998). However, palmarosa, lemongrass, peppermint, and eucalyptus oils are found<br />

to be bactericidal to Escherichia coli strain SP-11. Only peppermint and palmarosa oils induced the<br />

formation of elongated filamentous forms of E. coli (Pattnaik et al. 1995b).<br />

6.2.2.3 mould and yeast<br />

Yeast and mould comprise important groups of microorganisms that are responsible for several<br />

infections and for causing spoilage of foods, pharmaceutical products, and cosmetic products. Some<br />

fungal species are of agricultural and industrial importance and many cause disease in plants.<br />

Antifungal activity of some Cymbopogon species has been reported. For example, the mycelium<br />

growth of Aspergillus niger is inhibited by C. nardus essential oil. It causes morphological changes<br />

such as hyphal diameter and hyphal wall thinning, plasma membrane disruption, and mitochondria<br />

structure disorganization (de Billerberk et al. 2001). Lemongrass oil is an effective postharvest<br />

fungitoxicant of higher-order plant origin, potentially suitable for protection of foodstuffs against<br />

storage fungi (Mishra and Dubey 1994). Palmarosa oil had some inhibitory activity against 12<br />

fungi. The response of this oil is dependent on the species of fungi (Pattnaik et al. 1996). Moreover,<br />

palmarosa oil also has antimicrobial against yeast cell, Saccharomyces cerevisiae, by passively<br />

entering the plasma membrane to initiate membrane disruptions followed by accumulation in the<br />

plasma membrane resulting in the inhibition of cell growth (Prashar et al. 2003).<br />

6.2.3 th e Co r r e l at i o n b e t w e e n oil Co M P o n e n t s a n d aCtiVity<br />

Biological activity of an essential oil is in strict direct relationship to its chemical composition.<br />

The relation between components and activity may be attributed both to their major components


170 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

and to the minor ones present in the oils. The antibacterial action of essential oils depends on the<br />

chemical structure of their components. Phenolic compounds such as eugenol and carvacrol have<br />

an aromatic nucleus and phenolic OH group that is known to be reactive and to form hydrogen<br />

bonds with active sites of target enzymes (Farag et al. 1989). Mahmoud (1994) reported that geraniol,<br />

nerol, and citronellol, which are aliphatic alcohols, are broad-spectrum antifungal agents. The<br />

effectiveness of essential oil is higher than the activity of each component. It is possible that major<br />

components and minor compounds may act together synergistically to contribute to the activity<br />

(Milos et al. 2000).<br />

6.2.4 Me t h o d s Co M M o n ly used f o r an t iM iC r o b i a l assessMent o f es s e n t i a l oil<br />

Although there are the nonconventional methods for determining the minimum inhibitory concentration<br />

of essential oils (Kalemba and Kunicka 2003; Mann and Markham 1998), only the basic<br />

techniques commonly used for assessment of antibacterial and antifungal activities of essential<br />

oil, which are serial broth dilution method and agar dilution method, are stated. In the case of estimation<br />

of essential oil activity on vapor contact assay, the agar diffusion method is slightly modified.<br />

Antimicrobial activity of some Cymbopogon species and their major components is shown in<br />

Tables 6.2 and 6.3, respectively. However, Janssen et al. (1987) noted that the antimicrobial activity<br />

from different methods is not necessarily comparable.<br />

6.2.4.1 the serial broth dilution method<br />

<strong>Essential</strong> oil is dissolved in a solubilizing agent, which is nontoxic to microorganisms, before<br />

being serially diluted in an appropriate broth or agar. The effectiveness of essential oil is generally<br />

expressed as minimum inhibitory concentration (MIC), which is defined as the lowest concentration<br />

of essential oil that inhibits the visible growth, and as minimum bactericidal concentration (MCC)<br />

and minimum fungicidal concentration (MFC), which are defined as the lowest concentration of<br />

essential oil that causes more than 99.9% reduction of microorganism number or 3 log reduction<br />

of microorganisms (Kalemba and Kunicka 2003).<br />

6.2.4.2 the agar diffusion method<br />

The agar diffusion method is generally recommended as a prescreening method for a large number<br />

of essential oils since it is easy to perform and requires a small amount of essential oil. An assay can<br />

be performed accordingly; an appropriate nutrient agar plate is inoculated with microorganisms,<br />

either by adding the organism to the agar before it is poured or by streaking the organisms across<br />

the surface of the plate. The amount of essential oil tested can be accurately incorporated either<br />

on paper disk or into the well. The effectiveness of essential oil is demonstrated by the size of the<br />

microorganism inhibition zone around the disk or well, and it is usually expressed as the diameter<br />

of the zone. However, the disadvantages of this method are that essential oils are likely to evaporate<br />

with solvent during the incubation period, and they may show limited agar diffusion. Moreover, the<br />

inhibition zone of the oils depends on the characteristics of the oil components partitioned through<br />

the agar (Kalemba and Kunicka 2003; Southwell et al. 1993).<br />

6.2.4.3 Vapor contact assay<br />

This method is used for evaluating the activity of essential oils that are to be employed as atmospheric<br />

biocides (Lopez 2005; Tullio et al. 2006). An appropriate nutrient agar plate is inoculated<br />

with microorganisms. <strong>Essential</strong> oil is diluted in organic solvent such as ethyl acetate, ethyl<br />

ether, and ethanol. The exact amount of diluted essential oil is added to a paper disk, attached to<br />

the lid of a petri dish, and completely sealed; the lid is then inverted and incubated. The results<br />

are expressed as minimum inhibitory dose (MID), which is defined as the lowest concentration<br />

of essential oil (mg/L in air) that inhibits visible growth. In some studies, the soaked paper disk<br />

is placed in the airtight box next to the agar plate (Inouye et al. 2001; Nakahara et al. 2003).


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 171<br />

table 6.2<br />

In Vitro antimicrobial activity of Cymbopogon essential oils<br />

Cymbopogon Cymbopogon Cymbopogon Cymbopogon Cymbopogon<br />

Cymbopogon citratus flexuosus (Indian giganteus martinii nardus (ceylon winterianeous<br />

organisms<br />

(lemongrass )<br />

lemongrass) (tsauri grass) (palmarosa)<br />

citronella) (Java citronella) references<br />

bacteria<br />

Acinetobacter MIC 0.03% v/v MIC 0.12% v/v MIC 0.25% v/v Hammer et al. (1999)<br />

baumaii<br />

Aeromonas MIC 0.12% v/v MIC 0.12% v/v nd Hammer et al. (1999)<br />

sobria<br />

Bacillus brevis MIC 0.16 µL/mL MICa 1.66 µL/mL Kalemba and Kunicka<br />

(2003)<br />

Campylobacter 15 µL in 6 mm paper disk:<br />

15 µL in 6 mm paper<br />

Wannissorn et al.<br />

jejuni<br />

Inhibition zone 90 mm<br />

disk: Inhibition<br />

(2005)<br />

zone 40 mm<br />

Clostridium 15 µL in 6 mm paper disk:<br />

15 µL in 6 mm paper<br />

Wannissorn et al.<br />

perfringens Inhibition zone 90 mm<br />

disk: Inhibition<br />

(2005)<br />

zone 39.5 mm<br />

Enterococcus MIC 0.12% v/v Leaf and stem MIC 0.25% v/v MIC 1.0% v/v Hammer et al. (1999)<br />

faecalis<br />

MIC 60 and<br />

600 ppm<br />

Escherichia coli MIC 0.06% v/v<br />

MIC > 0.8% v/v Leaf and stem MIC 0.06–0.12% v/v MIC 0.25–0.5% v/v MIC > 0.8% v/v Hammer et al. (1999)<br />

MIC > 0.8% v/v<br />

MIC 60 ppm MIC 0.12% v/v MIC > 0.8% v/v<br />

Inouye et al. (2001)<br />

MCC 0.12% v/v<br />

MIC 0.2% v/v MCC 0.25% v/v<br />

Oussalah et al. (2007)<br />

MCC 1.66 µL/mL<br />

MCC 1.66 µL/mL 15 µL in 6 mm paper<br />

Pattnaik (1995a)<br />

MID 100 mg/L, air<br />

disk: Inhibition<br />

Wannissorn et al.<br />

15 µL in 6 mm paper disk:<br />

zone 10.5 mm<br />

(2005)<br />

Inhibition zone 12 mm<br />

Haemophilus MID 1.56 mg/L, air Inouye et al. (2001)<br />

influenzae<br />

Helicobacter MIC


172 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 6.2 (continued)<br />

In Vitro antimicrobial activity of Cymbopogon essential oils<br />

Cymbopogon Cymbopogon Cymbopogon Cymbopogon Cymbopogon<br />

Cymbopogon citratus flexuosus (Indian giganteus martinii nardus (ceylon winterianeous<br />

(lemongrass )<br />

lemongrass) (tsauri grass) (palmarosa)<br />

citronella) (Java citronella) references<br />

MIC 0.25% v/v Leaf and stem MIC 0.25% v/v MIC 1.0% v/v Hammer et al. (1999)<br />

MIC 60 ppm<br />

MIC 0.4% v/v MIC 0.4% v/v MIC 0.2% v/v MIC 0.8% v/v MIC 0.4% v/v Oussalah et al. (2007)<br />

organisms<br />

Lertsatitthanakorn<br />

et al. (2006)<br />

MIC 0.6 µL/mL<br />

MCC 0.6 µL/mL<br />

Klebsiella<br />

pneumoniae<br />

Listeria<br />

monocytogenes<br />

Propionibacterium<br />

acnes<br />

MIC 0.005–<br />

0.3 µL/mL<br />

MCC 0.625–<br />

1.2 µL/mL<br />

MIC 1.0% v/v<br />

Leaf and stem MIC > 2.0% v/v MIC > 2.0% v/v Hammer et al. (1999)<br />

MIC 1.3 µL/mL<br />

MIC 60 ppm<br />

Kalemba and Kunicka<br />

(2003)<br />

MIC 0.8% w/v MIC 0.8% w/v MIC 0.2% w/v MIC 0.4% w/v MIC > 0.8% w/v Oussalah et al. (2006)<br />

Pseudomonas<br />

aeruginosa<br />

MIC 0.4% v/v MIC 0.5% v/v<br />

MIC 0.2% v/v<br />

MIC 0.80 µL/mL<br />

Pseudomonas<br />

putida<br />

Salmonella<br />

typhimurium<br />

MIC 0.4% v/v Hammer et al. (1999)<br />

Kalemba and Kunicka<br />

(2003)<br />

Oussalah et al. (2007)<br />

Wannissorn et al.<br />

MIC > 2.0% v/v<br />

MIC 0.8% v/v<br />

15 µL in 6 mm paper<br />

disk: Inhibition<br />

zone 24 mm<br />

MIC 0.25% v/v<br />

MIC 0.8% v/v<br />

MIC 1.66 µL/mL<br />

15 µL in 6 mm paper disk:<br />

Inhibition zone 24 mm<br />

(2005)<br />

MIC 0.25% v/v MIC 0.25% v/v MIC > 2.0% v/v Hammer et al. (1999)<br />

Serratia<br />

marcescens<br />

Staphylococcus<br />

aureus<br />

MIC 0.05% v/v Hammer et al. (1999)<br />

Inouye et al. (2001)<br />

Kalemba and Kunicka<br />

(2003)<br />

Oussalah et al. (2007)<br />

MIC 0.12–0.25% v/v<br />

MIC 0.4% v/v<br />

MCC 0.25% v/v<br />

MIC 0.12% v/v<br />

MIC 0.1% v/v<br />

MIC 0.66 µL/mL<br />

MCC 0.12% v/v<br />

MIC 0.1% v/v Leaf and stem<br />

MIC 60 ppm<br />

MIC 0.06% v/v<br />

MIC 0.1% v/v<br />

MIC 0.3 µL/mL<br />

MCC 0.06% v/v<br />

MID 12.5 mg/L, air


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 173<br />

MID 6.25 mg/L, air Inouye et al. (2001)<br />

MID 6.25 mg/L, air Inouye et al. (2001)<br />

Wannissorn et al.<br />

(2005)<br />

15 µL in 6 mm<br />

paper disk:<br />

Inhibition zone<br />

15 µL in 6 mm paper disk:<br />

Inhibition zone 11 mm<br />

Streptococcus<br />

pyogenes<br />

Streptococcus<br />

pneumoniae<br />

Streptococcus<br />

enteritidis<br />

12.8 mm<br />

Vibrio cholerae MIC 0.3 µL/mL MIC 0.66 µL/mL Kalemba and Kunicka<br />

(2003)<br />

MID 800 mg/L, air Pawar and Thaker<br />

(2006)<br />

yeast and Fungi<br />

5 µL in 5 mm paper<br />

disk: Hyphae<br />

inhibition zone<br />

8 mm<br />

Spore inhibition<br />

zone 12 mm<br />

Aspergillus niger 5 µL in 5 mm paper disk:<br />

Hyphae inhibition zone<br />

21 mm<br />

Spore inhibition zone<br />

33 mm<br />

MID 250 mg/L, air Paranagama et al.<br />

(2003)<br />

Nakahara et al. (2003)<br />

Aspergillus flavus MIC 0.6 mg/mL<br />

MFC 1.0 mg/mL<br />

MIC 0.6 mg/mL Hammer et al. (1999)<br />

Devkatte et al. (2005)<br />

Lertsatithanakorn<br />

et al. (2005)<br />

Tampieri et al. (2005)<br />

Duarte et al. (2005)<br />

Dutta et al. (2006)<br />

Jirovetz et al. (2007)<br />

De Billerbeck et al.<br />

(2001)<br />

MID 250 mg/L, air Nakahara et al. (2003)<br />

MIC 0.12% v/v;<br />

0.5–1.0% v/v<br />

MFCa 0.12% v/v;<br />

2.0% v/v<br />

Mycelium growth<br />

inhibited at<br />

800 mg/L<br />

MIC 0.06–0.12% v/v<br />

MFC 0.12% v/v<br />

Leaf and stem<br />

MIC 60 ppm<br />

MIC 500 ppm<br />

MIC > 0.2 mg/mL<br />

Candida albicans MIC 0.06% v/v<br />

MIC 322 µg/mL<br />

MFC 0.06–0.12% v/v<br />

Eurotium<br />

amstelodami<br />

(continued on next page)


174 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 6.2 (continued)<br />

In Vitro antimicrobial activity of Cymbopogon essential oils<br />

Cymbopogon<br />

winterianeous<br />

(Java citronella) references<br />

Cymbopogon<br />

nardus (ceylon<br />

citronella)<br />

Cymbopogon<br />

martinii<br />

(palmarosa)<br />

Cymbopogon<br />

giganteus<br />

(tsauri grass)<br />

Cymbopogon<br />

flexuosus (Indian<br />

lemongrass)<br />

Cymbopogon citratus<br />

(lemongrass )<br />

organisms<br />

MID 250 mg/L, air Nakahara et al. (2003)<br />

MID 250 mg/L, air Nakahara et al. (2003)<br />

MID 250 mg/L, air Nakahara et al. (2003)<br />

MID 250 mg/L, air Nakahara et al. (2003)<br />

MID 250 mg/L, air Nakahara et al. (2003)<br />

MIC 0.2 mg/mL MFC 0.1% v/v Prashar et al. (2003)<br />

Sacchetti et al. (2005)<br />

MID 1 µg/mL, air<br />

Inouye et al. (2001)<br />

MFD 5.2 µg/mL, air<br />

Inouye et al. (2006)<br />

MFD 1.56 µg/mL, air<br />

Kalemba and Kunicka<br />

MIC 50 µg/mL; 0.25 µL/mL<br />

(2003)<br />

Eurotium<br />

chevalieri<br />

Penicillium<br />

adametzii<br />

Penicillium<br />

citrinum<br />

Penicillium<br />

griseofulvum<br />

Penicillium<br />

islandicum<br />

Saccharamyces<br />

cerevisiae<br />

Tricophyton<br />

mentagrophytes<br />

Inouye et al. (2001)<br />

MFC 15.2 µg/mL<br />

MID 1 µg/mL, air<br />

MFD 5.2 µg/mL, air<br />

MIC 50 µg/mL<br />

MFC 15.2 µg/mL<br />

Tricophyton<br />

rubrum<br />

a Does not indicate species.


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 175<br />

table 6.3<br />

In Vitro antimicrobial activity of the major components of Cymbopogon essential oils<br />

alcohols aldehydes ester terpene hydrocarbons<br />

references<br />

Rosato et al. (2007)<br />

Van Zyl et al.<br />

(2006)<br />

geraniol citronellol citral citronellal geranyl acetate limonene mycene<br />

MIC 234.9 ±<br />

0.0 mM<br />

MIC 163.0 ±<br />

29.0 mM<br />

MIC 0.70 mg/mL MIC > 207.5 ±<br />

45.4 mM<br />

Bacillus cereus MIC 0.7 mg/mL<br />

MIC 51.9 mM<br />

Inouye et al. (2001)<br />

Rosato et al. (2007)<br />

Si et al. (2006)<br />

Van Zyl et al.<br />

(2006)<br />

MID > 800<br />

mg/L air<br />

MIC 176.2 ±<br />

40.4 mM<br />

MIC > 163.0 ±<br />

51.0 mM<br />

MIC 207.5 ±<br />

64.8 mM<br />

Bacillus subtilis MIC 0.8 mg/mL MIC 0.35 mg/mL<br />

Escherichia coli MID > 25 mg/L air MIC 1.4 mg/mL MID > 12.5 mg/L<br />

MIC 1.4 mg/mL<br />

air<br />

MBC 283–300 µg/mL<br />

MIC 25.9 ± 0.0 mM<br />

MID 200 mg/L<br />

air<br />

MID 6.25 mg/L air MID 3.13 mg/L<br />

air<br />

MIC 1000 µg/mL MIC 500 µg/mL Kalemba and<br />

Kunicka (2003)<br />

MIC 0.375 mg/mL Kalemba and<br />

Kunicka (2003)<br />

MID 400 mg/L<br />

air<br />

MID 200–400<br />

mg/L air<br />

MID 12.5 mg/L air MID 3.13 mg/L<br />

air<br />

MID 6.25 mg/L air MID 6.25 mg/L<br />

air<br />

Tampieri et al.<br />

(2005)<br />

Van Zyl et al.<br />

MIC 1000 ppm<br />

MIC 163.0 ±<br />

0.0 mM<br />

MIC 100 ppm MIC 77.8 ±<br />

0.0 mM<br />

MIC 100 ppm<br />

MIC 19.5 ± 0.0 mM<br />

Haemophilus<br />

influenzae<br />

Listeria<br />

monocytogenes<br />

Pseudomonas<br />

aeruginosa<br />

Streptococcus<br />

pyogenes<br />

Streptococcus<br />

pneumoniae<br />

Candida<br />

albicans<br />

MIC 73.4 ± 0.0<br />

mM<br />

(2006)<br />

MIC 500 µg/mL Kalemba and<br />

Kunicka (2003)<br />

Si et al. (2006)<br />

(continued on next page)<br />

MIC 500 µg/mL<br />

MBC 367 µg/mL<br />

Salmonella<br />

typhimurium


176 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 6.3 (continued)<br />

In Vitro antimicrobial activity of the major components of Cymbopogon essential oils<br />

alcohols aldehydes ester terpene hydrocarbons<br />

geraniol citronellol citral citronellal geranyl acetate limonene mycene references<br />

MID > 25 mg/L, air MIC 0.70 mg/mL MID 12.5 mg/L, MIC 129.7 ± MIC > 163.0 ± MID 800 mg/L,<br />

Inouye et al. (2001)<br />

MIC 0.7 mg/mL<br />

air<br />

55.1 mM<br />

35.7 mM air<br />

Rosato et al. (2007)<br />

MIC 38.9 ± 18.2 mM<br />

MIC 176.2 ±<br />

Van Zyl et al.<br />

40.4 mM<br />

(2006)<br />

MIC 0.125 mg/mL Kalemba and<br />

Kunicka (2003)<br />

No activity No activity No activity Potent activity No activity No activity No activity Nakahara et al.<br />

MID 28 mg/L, air<br />

(2003)<br />

MIC 500 ppm MIC 500 ppm No activity Potent activity No activity No activity No activity Kalemba and<br />

No activity<br />

No activity<br />

MID 56 mg/L, air<br />

Kunicka (2003)<br />

Nakahara et al.<br />

(2003)<br />

No activity No activity No activity Potent activity No activity No activity No activity Nakahara et al.<br />

MID 28 mg/L, air<br />

(2003)<br />

1.0 mM growth 1.0 mM growth 1.0 and 2.0 mM 2.0 mM growth<br />

2.0 mM growth 2.0 mM Kurita et al. 1981<br />

inhibition > 20 days inhibition > growth<br />

inhibition 0 days<br />

inhibition growth<br />

20 days<br />

inhibition 3 and<br />

0 days<br />

inhibition<br />

> 20 days,<br />

0 days<br />

respectively<br />

No activity No activity No activity Potent activity No activity No activity No activity Nakahara et al.<br />

MID 28 mg/L, air<br />

(2003)<br />

Staphylococcus<br />

aureus<br />

Staphylococcus<br />

epidermidis<br />

Aspergillus<br />

candidus<br />

Aspergillus<br />

flavus<br />

Aspergillus<br />

versicolor<br />

Aspergillus<br />

nidulans<br />

Eurotium<br />

amstelodami


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 177<br />

No activity No activity No activity Nakahara et al.<br />

(2003)<br />

No activity No activity No activity Nakahara et al.<br />

(2003)<br />

No activity No activity No activity Nakahara et al.<br />

(2003)<br />

No activity No activity No activity Nakahara et al.<br />

(2003)<br />

No activity No activity No activity Nakahara et al.<br />

(2003)<br />

Inouye et al. (2001)<br />

No activity Moderate activity Moderate activity Potent activity<br />

MID 14 mg/L, air<br />

No activity No activity No activity Potent activity<br />

MID 56 mg/L, air<br />

No activity No activity No activity Potent activity<br />

MID 28 mg/L, air<br />

No activity No activity No activity Potent activity<br />

MID 56 mg/L, air<br />

Moderate activity No activity Moderate activity Potent activity<br />

MID 14 mg/L, air<br />

MIC 25 µg/mL<br />

MFC 7.7 ±<br />

1.85 µg/mL<br />

MID 0.5 µg/L, air<br />

MFC 3.9 ±<br />

1.2 µg/L, air<br />

1.0 mM growth 1.0 mM growth 1.0 mM growth 1.0 mM growth<br />

inhibition > 20 days inhibition > inhibition > inhibition 0 days<br />

20 days<br />

20 days<br />

2.0 mM growth<br />

inhibition 2 days<br />

Eurotium<br />

chevalieri<br />

Penicillium<br />

adametzii<br />

Penicillium<br />

citrinum<br />

Penicillium<br />

griseofulvum<br />

Penicillium<br />

islandicum<br />

Tricophyton<br />

mentagrophytes<br />

Kurita et al. 1981<br />

2.0 mM<br />

growth<br />

inhibition<br />

0 days<br />

2.0 mM growth<br />

inhibition<br />

0 days<br />

Tricophyton<br />

rubrum


178 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

6.3 In VItro antIoxIdant actIVIty oF some Cymbopogon sPecIes<br />

and theIr maJor comPonents<br />

<strong>Essential</strong> oils and their components are gaining increasing interest because of their potential for<br />

multipurpose functional use. Apart from antimicrobial and antifungal properties, antioxidant and<br />

radical-scavenging properties are of great interest to health and food science researchers. An antioxidant<br />

may be defined as any substance that delays or inhibits oxidation of that substrate. In vitro methods<br />

provide a useful indication of antioxidant activity. Due to the differences between in vitro testing<br />

and biological environments, the data obtained from in vitro methods are difficult to apply to biological<br />

systems (Antolovich et al. 2002). Antioxidant activity can be divided into two classes: primary<br />

(chain-breaking or radical-scavenging antioxidants) and secondary (preventive antioxidants)<br />

(Laguerre et al. 2007). Frankel and Meyer (2000) pointed out that multifaceted testing of antioxidant<br />

activity is needed because antioxidants often act via mixed mechanisms. <strong>Essential</strong> oils are a<br />

very complex mixture that can contain two or three major components at fairly high concentrations<br />

and other components in trace amounts (Bakkali et al. 2008). The antioxidant activity of essential<br />

oils, therefore, would be associated with various mechanisms. In this review, only the common<br />

in vitro antioxidant testing methods for essential oils are mentioned.<br />

6.3.1 fa C t o r s affeC tinG an t i o x i d a n t aCtiVity<br />

The antioxidant activity of essential oils and their components shows the marked difference between<br />

the reported results (Table 6.4), which depend on many factors. The magnitudes of antioxidant<br />

activities, therefore, could only be compared for given process conditions. In this review, studies<br />

on the antioxidant activity of Cymbopogon species and their major components published in the<br />

literature are reported, and only factors affecting their antioxidant activity, using the three methods<br />

of testing mentioned earlier, are discussed.<br />

6.3.1.1 choice of oxidizable substrate and end-Product evaluation<br />

The choice of an oxidizable substrate depends on whether the antioxidant is for nutritional, therapeutic,<br />

or preservative purposes. It is important to select a substrate that is representative of in situ<br />

conditions. Since the aim of assessing antioxidant efficacy of essential oils is to preserve foods,<br />

the oxidizable substrate normally used to evaluate antioxidant activity is either linoleic acid or<br />

homogenized egg yolk (Laguerre et al. 2007). Therefore, the methods often used for this assessment<br />

are β-carotene bleaching assay and TBARS assay. However, a limitation of evaluating antioxidant<br />

activity by TBARS assay is that thiobarbituric acid not only reacts with MDA, which is a secondary<br />

oxidation product of lipid peroxidation but also can react with other aldehydes (Janero 1990).<br />

Thus, essential oils containing aldehyde compounds such as citral, citronellal, heptanal, octanal,<br />

etc., may cause an underestimation of their antioxidant activity. This evidence can be seen in the<br />

study of Ruberto and Baratta (2000), which found that antioxidant activity of aldehyde compounds<br />

can be assessed by determination of conjugate diene hydroperoxides, which are primary oxidation<br />

products of linoleic acid peroxidation but undetectable with TBARS assay.<br />

6.3.1.2 media<br />

There are two major types of media: homogeneous medium and heterogeneous medium. In a heterogeneous<br />

medium, the type of solvent could affect the antioxidant mechanism. Antioxidants<br />

behave differently in media with different polarities. For example, DPPH is a stable-free radical, but<br />

it is sensitive to some Lewis bases, solvent types, light, and oxygen. Polar solvents may decrease the<br />

odd electron density of the nitrogen atom in DPPH and increase the reactivity of DPPH. In general,<br />

DPPH in the methanol buffer system shows better stability than in the acetone buffer system (pH<br />

10) (Ozcelik et al. 2003). Porter (1993) described the “polar paradox”: lipophilic antioxidants are


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 179<br />

table 6.4<br />

In Vitro antioxidant activity of Cymbopogon species and their major components<br />

sample antioxidant activity method of testing references<br />

Citronellol AI at 1000 ppm = 27.5%<br />

AI at 0.001 M = 13.3%<br />

IC50 =>1000 µg/mL<br />

AI at 10 µL/mL = 15.9%<br />

Geraniol AI at 1000 ppm = 34.9%<br />

AI at .001 M = 26.5%<br />

IC50 => 1000 µg/mL<br />

AI at 10 µL/mL = 19.7%<br />

Limonene AI at 1000 ppm = 27.4%<br />

AI at 10 µL/mL = 22.2%<br />

AI at 0.001 M = 21.0%<br />

Alpha tocopherol AI at 1000 ppm = 93.5%<br />

AI at 0.001 M = 94.8%<br />

C. citratus AI at 5 µL/mL = 26.0%<br />

AI at 10 µL/mL = 31.6%<br />

IC50 = 27.0 µL/mL<br />

AI at 10 µL/mL = 63.8%<br />

AI at 2 µL/mL = approx 50%<br />

C. nardus IC50 = 2.0 µL/mL<br />

AI at 5 µL/mL = 89.0%<br />

AI at 10 µL/mL = 45.5%<br />

Citronellal AI at 1000 ppm = not detectable<br />

AI at 0.001 M = 21.9%<br />

AI at 10 µL/mL = 16.4%<br />

Citral AI at 1000 ppm = not detectable<br />

AI at 0.0001 M = 18.3%<br />

IC50 => 1000 µg/mL<br />

AI at 10 µL/mL = 21.1%<br />

TBARS method<br />

Conjugated diene formation<br />

DPPH test<br />

DPPH test<br />

TBARS method<br />

Conjugated diene formation<br />

DPPH test<br />

DPPH test<br />

TBARS method<br />

DPPH test<br />

Conjugated diene formation<br />

Conjugated diene formation<br />

TBARS method<br />

TBARS method<br />

TBARS method<br />

DPPH test<br />

DPPH test<br />

β-carotene bleaching assay<br />

DPPH test<br />

DPPH test<br />

TBARS method<br />

TBARS method<br />

Conjugated diene formation<br />

DPPH test<br />

TBARS method<br />

Conjugated diene formation<br />

DPPH test<br />

DPPH test<br />

Ruberto et al. (2000)<br />

Ruberto et al. (2000)<br />

Hierro et al. (2004)<br />

Unpublished data<br />

Ruberto et al. (2000)<br />

Ruberto et al. (2000)<br />

Hierro et al. (2004)<br />

Unpublished data<br />

Ruberto et al. (2000)<br />

Unpublished data<br />

Ruberto et al. (2000)<br />

Ruberto et al. (2000)<br />

Ruberto et al. (2000)<br />

Unpublished data<br />

Unpublished data<br />

Lertsatittanakorn et al. (2006)<br />

Sacchetti et al. (2005)<br />

Sacchetti et al. (2005)<br />

Lertsatittanakorn et al. (2006)<br />

Unpublished data<br />

Unpublished data<br />

Ruberto et al. (2000)<br />

Ruberto et al. (2000)<br />

Unpublished data<br />

Ruberto et al. (2000)<br />

Ruberto et al. (2000)<br />

Hierro et al. (2004)<br />

Unpublished data<br />

more active in a polar medium, whereas polar antioxidants are more active in a lipophilic medium.<br />

In a heterogeneous medium, either in oil in water emulsion or aqueous suspensions of liposome or<br />

LDLs, oxidation is affected by the type of interface and its viscosity, the size distribution of oil and<br />

liposome droplets, the partition and diffusion of oxygen toward reaction centers, and the antioxidant<br />

location. According to Porter’s polar paradox, in emulsion media, oxidation occurs at the oil–water<br />

interface, where lipophilic antioxidants are more efficient than hydrophilic antioxidants. In addition,<br />

lipid peroxidation is dependent on the pH of water in oil emulsion and in liposomes (Laguerre<br />

et al. 2007). Frankel (1998) reported that lipid oxidation is generally lower at high pH and, consequently,<br />

the oxidation rate increases as pH decreases.<br />

6.3.1.3 oxidation conditions<br />

Heat: Since essential oils are very volatile and undergo thermal degradation at high temperature,<br />

temperature is an important factor when assessing the extent of oxidation. It is,<br />

therefore, important to carry out tests at temperatures close to natural conditions, such as<br />

ambient or physiological temperatures (Laguerre et al. 2007).<br />

ROO peroxy radicals: ROO · peroxy radicals are a prime target for assessing antiradical activity.<br />

Azoinitiators are commonly used for ROO · generation. In β-carotene and TBARS


180 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

assays, a water-soluble 2,2′-Azobis(2-amidopropane)dihydrochloride (AAPH or ABAP) is<br />

most commonly used for ROO · generation. Hanlon and Seybert (1997) found that ABAP<br />

dramatically increased the rate of lipid peroxidation as pH increased from 5 to 7, and<br />

began to plateau at a pH of about 8. In addition, the rate of lipid peroxidation depends on<br />

temperature and the viscosity of the medium. It is likely that as the rate of ROO · diffusion<br />

increases, the rate of lipid peroxidation increases (Laguerre et al. 2007).<br />

6.3.1.4 methods commonly used for testing antioxidant activity of essential oils<br />

Three common methods used to evaluate antioxidant activity of essential oils are the β-carotene<br />

bleaching method, DPPH free radicals scavenging method, and thiobarbituric acid reactive species<br />

assay (TBARS). The principles of the assays are described in the following subsections.<br />

6.3.1.4.1 β-Carotene Bleaching Assay<br />

This method is based on the result of β-carotene oxidation by linoleic acid degradation products,<br />

which are catalyzed by heat. Tween is used for dispersion of linoleic acid and β-carotene in the aqueous<br />

phase. The addition of an antioxidant results in retarding β-carotene bleaching. Quantitative<br />

analysis of β-carotene is measured by UV spectrophotometry at 470 nm (Laguerre et al. 2007).<br />

Results are expressed as the percentage inhibition of β-carotene bleaching (Gachkar et al. 2007;<br />

Kulisic et al. 2004; Sacchetti et al. 2005; Wang et al. 2008), the relative antioxidant activity (RAA%)<br />

of the sample and butylated hydroxyltoluene (BHT) (Obame et al. 2007), or IC 50, which is a sample<br />

concentration providing 50% inhibition (Khadri et al. 2008).<br />

6.3.1.4.2 2,2-Diphenyl-1-Perylhydrazyl Free Radicals Scavenging Test (DPPH Test)<br />

The DPPH test is based on the reduction of 2,2-diphenyl-1-perylhydrazyl free radicals (DPPH) in<br />

the presence of hydrogen-donating antioxidant or free radical scavenging agent (Kulisic et al. 2004).<br />

The DPPH radical absorbs at 517 nm, and antioxidant activity can be determined by monitoring the<br />

decrease of DPPH radical. Results are expressed as EC 50, that is, the amount of antioxidant necessary<br />

to decrease the initial DPPH concentration by 50% (Antolovich et al. 2002; Lertsatitthanakorn<br />

et al. 2006; Demirci et al. 2007), or the percentage inhibition of DPPH radical (Gachkar et al. 2007;<br />

Khadri et al. 2008; Kulisic et al. 2004; Obame et al. 2007; Sacchetti et al. 2005; Singh et al. 2005;<br />

Wang et al. 2008).<br />

6.3.1.4.3 Thiobarbituric Acid Reactive Substances (TBARS) Assay<br />

This method is commonly used to detect lipid oxidation. This procedure measures malonaldehyde<br />

(MDA) formation, which is the split product and endoperoxide of unsaturated fatty acids resulting<br />

from oxidation of a lipid substrate. The MDA is reacted with thiobarbituric acid to form a pink pigment<br />

that is measured spectrophotometrically at 532–535 nm (Antolovich et al. 2002). Results may<br />

be expressed as the percentage inhibition of oxidation (AI%) (Ruberto and Baratta 2000; Kulisic<br />

et al. 2004) or thiobarbituric acid value (meq of malonaldehyde/g) (Singh et al. 2005).<br />

6.4 summary<br />

The results given in the literature suggest that essential oils from some Cymbopogon species show<br />

antimicrobial activity against a wide range of bacteria and fungal species, and also have a weakto-moderate<br />

antioxidant activity. Although the oils are used as fragrance in perfumery and in the<br />

food and beverage industry, they may also have great potential for protection of food and cosmetics<br />

from microbial spoilage and as a topical antiseptic. However, because it has weak-to-moderate<br />

antioxidant activity, using the oils to prevent oxidative deterioration of lipids in food might be helpful<br />

for a certain period of time.


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 181<br />

reFerences<br />

Antolovich, M., Prenzler, P., Patsalides, E., McDonald, S., Robards, K., 2002. Methods for testing antioxidant<br />

activity. Analyst 127, 183–198.<br />

Bakkali, F., Averbeck, A., Averbeck, V.D., Idaomar, M., 2008. Biological effects of essential oils—A review.<br />

Food and Chemistry Toxicology 46, 446–475.<br />

de Billerbeck, V., Roque, C., Bessiere, J., Fonvieille, J., Dargent, R., 2001. Effects of Cymbopogon nardus (L.)<br />

W. Watson essential oil on the growth and morphogenesis of Aspergillus niger. Canadian Journal of<br />

Microbiology 47, 9–17.<br />

Demirci, B., Kosar, M., Demirci, F., Dinc, M., Baser, K.H.C., 2007. Antimicrobial and antioxidant activities of<br />

the essential oil of Chaerophyllum libanoticum Boiss. et Kotschy. Food Chemistry 105, 1512–1517.<br />

Devkatte, A.N., Zore, G.B., Karuppayil, S.M., 2005. Potential of plant oils as inhibitors of Candida albicans<br />

growth. FEMS Yeast Research 5, 867–873.<br />

Duarte, M.C., Figueira, G.M., Sartoratta, A., Rehder, V.L., Delarmeliva, C., 2005. Anti-candida activity of<br />

Brazilian medicinal plants. Journal of Ethnopharmacology 97, 305–311.<br />

Dutta, B.K., Karmakar, S., Naglot, A., Aich, J.C., Begam, M., 2006. Anticandidial activity of some essential<br />

oils of a mega biodiversity hotspot in India. Mycoses 50, 121–124.<br />

Farag, R.S., Daw, Z.Y., Abo-Raya, S.H., 1989. Influence of some spice essential oils on Aspergillus parasiticus<br />

growth and production of aflatoxins in a synthetic medium. Journal of Food Science 54, 74–76.<br />

Frankel, E., 1998. Lipid Oxidation. The <strong>Oil</strong>y Press, Dundee, U.K.<br />

Frankel, E., Meyer, A., 2000. The problems of using one-dimensional methods to evaluate multifunctional food<br />

and biological antioxidants. Journal of Science Food Agriculture 80, 1925–1941.<br />

Gachkar, L., Yadegari, D., Rezaei, M.B., Taghizadeh, M., Astaneh, S.A., Rasooli, I., 2007. Chemical and biological<br />

characteristics of Cuminum cyminum and Rosmarinus officinalis essential oils. Food Chemistry<br />

102, 898–904.<br />

Hammer, K.A., Carson, C.F., Riley, T.V., 1999. Antimicrobial activity of essential oils and other plant extracts.<br />

Journal of Applied Microbiology 86, 985–990.<br />

Hanlon, M., Seybert, D., 1997. The pH dependence of lipid peroxidation using water-soluble azo initiators.<br />

Free Radical Biological Medicine 23, 712–719.<br />

Helander, I.M., Alakomi, H., Latva-Kala, K., Mattila-Sandholm, T., Pol, I., Smid, E., Gorris, L.G.M.,<br />

von Wright, A., 1998. Characterization of the action of selected essential oil components on Gramnegative<br />

bacteria. Journal of Agricultural and Food Chemistry 46, 3590–3595.<br />

Hierro, I., Valero, A., Perez, P., Gonzalez, P., Cabo, M.M., Montilla, M.P., Navarro, M.C., 2004. Action of different<br />

monoterpenic compounds against Anisakis simplex s.l. L3 larvae. Phytomedicine 11, 77–82.<br />

Hili, P., Evans, C.S., Veness, R.G., 1997. Antimicrobial action of essential oils: The effect of dimethylsulphoxide<br />

on the activity of cinnamon oil. Letters in Applied Microbiology 24, 269–275.<br />

Himejima, M., Kubo, I., 1993. Fungicidal activity of polygodial in combination with anethole and indole<br />

against Candida albicans. Journal of Agricultural and Food Chemistry 41, 1776–1779.<br />

Inouye, S., Takizawa , T., Yamaguchi, H., 2001. Antibacterial activity of essential oils and their major constituents<br />

against respiratory tract pathogens by gaseous contact. Journal of Antimicrobial Chemotherapy 47, 565–573.<br />

Inouye, S., Uchida, K., Abe, S., 2006. Vapor activity of 72 essential oils against a Trichophyton mentagrophytes.<br />

Journal of Infection and Chemotherapy 12, 210–216.<br />

Janero, D.R., 1990. Malondialdehyde and thiobarbituric acid-reactivity as diagnostic indices of lipid peroxidation<br />

and peroxidative tissue injury. Free Radical Biology and Medicine 9, 515–540.<br />

Janssen, A.M., Scheffer, J.J.C., Svendsen, A.B., 1987. Antimicrobial activity of essential oils: A 1976–1986<br />

literature review. Aspects of the test methods. Planta Medica 53, 395–398.<br />

Jirovetz, L., Buchbauer, G., Gernot, N., Benoit, M.M., Pierre, M., 2007. Composition and antimicrobial activity<br />

of Cymbopogon giganteus (Hochst.) Chiov. <strong>Essential</strong> flower, leaf and stem oils from Cameroon. Journal<br />

of <strong>Essential</strong> <strong>Oil</strong> Research 19, 485–489.<br />

Kalemba, D., Kunicka, A., 2003. Antibacterial and antifungal properties of essential oils. Current Medicinal<br />

Chemistry 10, 813–829.<br />

Khadri, A., Serralheiro, M.L.M., Nogueira, J.M.F., Neffati, M., Smiti, S., Araujo, M.E.M., 2008. Antioxidant<br />

and antiacetylcholinesterase activities of essential oils from Cymbopogon schoenanthus L. Spreng.<br />

Determination of chemical composition by GC-mass spectrometry and 13C NMR. Food Chemistry 109,<br />

630–637.<br />

Kulisic, T., Radonic, A., Katalinic, V., Milos, M., 2004. Use of different methods for testing antioxidative activity<br />

of oregano essential oil. Food Chemistry 85, 633–640.


182 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Kurita, N., Miyaji, M., Kurane, R., 1981. Antifungal activity of components of essential oils. Agricultural and<br />

Biological Chemistry 45, 945–952.<br />

Laguerre, M., Lecomte, J., Villeneuve, P., 2007. Evaluation of the ability of antioxidants to counteract lipid<br />

oxidation: Existing methods, new trends and challenges. Progress in Lipid Research 46, 244–282.<br />

Lertsatitthanakorn, P., Taweechaisupapong, S., Aromdee, C., Khunkitti, W., 2006. In vitro bioactivities of<br />

essential oils used for acne control. International Journal of Aromatherapy 16, 43–49.<br />

Lopez, P., Sanchez, C., Batlle, R., Nerin, C., 2005. Solid- and vapor-phase antimicrobial activities of six essential<br />

oils: Susceptibility of selected foodborne bacterial and fungal strains. Journal of Agricultural and<br />

Food Chemistry 53, 6939–6946.<br />

Mahmoud, A.L., 1994. Antifungal action and antiaflatoxigenic properties of some essential oil constituents.<br />

Letters in Applied Microbiology 19, 110–113.<br />

Mann, C.M., Markham, J.L., 1998. A new method for determining the minimum inhibitory concentration of<br />

essential oils. Journal of Applied Microbiology 84, 538–544.<br />

Manou, I., Bouillard, L., Devleeschouwer, M.J., Barel, A.O., 1998. Evaluation of the preservative properties of<br />

Thymus vulgaris essential oil in topically applied formulations under a challenge test. Journal of Applied<br />

Microbiology 84, 368–376.<br />

Milos, M., Mastelic, J., Jerkovic, I., 2000. Chemical composition and antioxidant effect of glycosidically bound<br />

volatile compounds from oregano (Origanum vulgare L. ssp. hirtum). Food Chemistry 71, 79–83.<br />

Mishra, A.K., Dubey, N.K., 1994. Evaluation of some essential oils for their toxicity against fungi causing<br />

deterioration of stored food commodities. Applied and Environmental Microbiology 60, 1101–1105.<br />

Nakahara, K., Alzoreky, N., Yoshihashi, T., Nguyen, H., Trakoontivakorn, G., 2003. Chemical composition<br />

and antifungal activity of essential oil from Cymbopogon nardus (Citronella grass). Japan Agricultural<br />

Research Quarterly 37, 249–252.<br />

Nikaido, H., 1994. Prevention of drug access to bacterial targets: Permeability barriers and active efflux.<br />

Science 264, 382–388.<br />

Obame, L., Koudou, J., Chalchat, J., Boassole, I., Edou, P., Ouattara, A., Traore, A., 2007. Volatile components,<br />

antioxidant and antibacterial activities of Dacryodes buettneri H.J. Lam. essential oil from Gabon.<br />

Scientific Research and Essay 2, 491–495.<br />

Ohno, T., Kita, M., Yamaoka, Y., Imamura, S., Yamamoto, T., Mitsufuji, S., Kodama, T., Kashima, K., Imanishi,<br />

J., 2003. Antimicrobial activity of essential oils against Helicobacter pylori. Helicobacter 8, 207–215.<br />

Oussalah, M., Caillet, S., Saucier, L., Lacroix, M., 2006. Antimicrobial effects of selected plant essential oils on<br />

the growth of a Pseudomonas putida strain isolated from meat. Meat Science 73, 236–244.<br />

Ozcelik, B., Lee, J.H., Min, D.B., 2003. Effects of light, oxygen, and pH on the absorbance of 2,2-diphenyl-1picrylhydrazyl.<br />

Journal of Food Science 68, 487–490.<br />

Paranagama, P.A., Abeysekera, K.H.T., Abeywickrama, K., Nugaliyadde, L., 2003. Fungicidal and anti-aflatoxigenic<br />

effects of the essential oil of Cymbopogon citratus (DC.) Stapf. (lemongrass) against Aspergillus<br />

flavus Link. isolated from stored rice. Letters in Applied Microbiology 37, 86–90.<br />

Pattnaik, S., Subramanyam, V.R., Kole, C., 1996. Antibacterial antifungal activity of ten essential oils in vitro.<br />

Microbios 86, 237–246.<br />

Pattnaik, S., Subramanyam, V.R., Kole, C.R., Sahoo, S., 1995a. Antibacterial activity of essential oils from<br />

Cymbopogon: Inter- and intra-specific differences. Microbios 84, 239–245.<br />

Pattnaik, S., Subramanyam, V.R., Rath, C.C., 1995b. Effect of essential oils on the viability and morphology of<br />

Escherichia coli. Microbios 84, 195–199.<br />

Pawar, V.C., Thaker, V.S., 2006. In vitro efficacy of 75 essential oils against Aspergillus niger. Mycoses 49,<br />

316–323.<br />

Porter, W., 1993. Paradoxical behavior of antioxidants in food and biological systems. In: Williams, G. (Ed.),<br />

Antioxidants: Chemical, Physiological, Nutritional and Toxicological Aspects, Princeton Scientific,<br />

Princeton, pp. 93–122.<br />

Prabuseenivasan, S., Jayakumar, M., Ignacimuthu, S. 2006. In vitro antibacterial activity of some plant essential<br />

oils. BMC Complementary and Alternative Medicine, 6, 39, doi:10.1186/1472-6882-1186-1139.<br />

Prashar, A., Hili, P., Veness, R.G., Evans, C.S., 2003. Antimicrobial action of palmarosa oil (Cymbopogon<br />

martinii) on Saccharomyces cerevisiae. Phytochemistry 63, 569–575.<br />

Remmal, A., Bouchikhi, T., Rhayour, K., Ettayebi, M., Tantaoui-Elaraki, A., 1993. Improved method for<br />

the determination of antimicrobial activity of essential oils in agar medium. Journal of <strong>Essential</strong> <strong>Oil</strong><br />

Research 5, 179–184.<br />

Rosato, A., Vitali, C., De Laurentis, N., Armenise, D., Antonietta Milillo, M., 2007. Antibacterial effect of some<br />

essential oils administered alone or in combination with Norfloxacin. Phytomedicine 14, 727–732.


In Vitro Antimicrobial and Antioxidant Activities of Some Cymbopogon Species 183<br />

Ruberto, G., Baratta, M.T., 2000. Antioxidant activity of selected essential oil components in two lipid model<br />

systems. Food Chemistry 69, 167–174.<br />

Russell, A.D., Hugo, W.B., Ayliffe, G., 1992. Principles and Practice of Disinfection, Preservation and Sterilization.<br />

2nd ed. Blackwell Science, Oxford.<br />

Sacchetti, G., Maietti, S., Muzzoli, M., Scaglianti, M., Manfredini, S., Radice, M., Bruni, R., 2005. Comparative<br />

evaluation of 11 essential oils of different origin as functional antioxidants, antiradicals and antimicrobials<br />

in foods. Food Chemistry 91, 621–632.<br />

Shahi, A.K., Tava, A., 1993. <strong>Essential</strong> oil composition of three cymbopogon species of Indian Thar Desert.<br />

Journal of <strong>Essential</strong> <strong>Oil</strong> Research 5, 639–643.<br />

Si, W., Gong, J., Tsao, R., Zhou, T., Yu, H., Poppe, C., Johnson, R., Du, Z., 2006. Antimicrobial activity of<br />

essential oils and structurally related synthetic food additives towards selected pathogenic and beneficial<br />

gut bacteria. Journal of Applied Microbiology 100, 296–305.<br />

Singh, G., Paimuthu, P., Murali, H., Bawa, A., 2005. Antioxidative and antibacterial potentials of essential oils<br />

and extracts isolated from various spice materials. Journal of Food Safety 25, 130–145.<br />

Southwell, I.A., Hayes, A.J., Marham, J., Leach, D.N., 1993. The search for optimally bioactive Australian tea<br />

tree oil. Acta Horticulturae 334, 256–265.<br />

Tampieri, M.P., Galuppi, R., Macchioni, F., Carelle, M., Falcioni, L., Cioni, P., Morelli, I., 2005. The inhibition<br />

of Candida albicans by selected essential oils and their major components. Mycopathologia 159,<br />

339–345.<br />

Tullio, V., Nostro, A., Mandras, N., Dugo, P., Banche, G., Cannatelli, M.A., Cuffini, A.M., Alonzo, V., Carlone,<br />

N.A., 2007. Antifungal activity of essential oils against filamentous fungi determined by broth microdilution<br />

and vapour contact methods. Journal of Applied Microbiology 102, 1544–1550.<br />

van Zyl, R., Seatlholo, S., van Vuuren, S., Viljoen, A.M., 2006. The biological activities of 20 nature identical<br />

essential oil constituents. Journal of <strong>Essential</strong> <strong>Oil</strong> Research 19–133.<br />

Wang, W., Wu, N., Zu, Y.G., Fu, Y.J., 2008. Antioxidative activity of Rosmarinus officinalis L. essential oil<br />

compared to its main components. Food Chemistry 108, 1019–1022.<br />

Wannissorn, B., Jarikasem, S., Siriwangchai, T., Thubthimthed, S., 2005. Antibacterial properties of essential<br />

oils from Thai medicinal plants. Fitoterapia 76, 233–236.<br />

Zani, F., Massimo, G., Benvenuti, S., Bianchi, A., Albasini, A., Melegari, M., Vampa, G., Bellotti, A., Mazza, P.,<br />

1991. Studies on the genotoxic properties of essential oils with Bacillus subtilis rec-assay and Salmonella/<br />

microsome reversion assay. Planta medica 57, 237–241.


7 Thrombolysis-Accelerating<br />

Activity of <strong>Essential</strong> <strong>Oil</strong>s<br />

contents<br />

7.1 IntroductIon<br />

Hiroyuki Sumi and Chieko Yatagai<br />

7.1 Introduction .......................................................................................................................... 185<br />

7.2 Test I ..................................................................................................................................... 186<br />

7.2.1 Results ....................................................................................................................... 186<br />

7.3 Test II .................................................................................................................................... 186<br />

7.3.1 Standard Fibrin Plate Method................................................................................... 186<br />

7.3.2 Euglobulin: Fibrin Plate Method .............................................................................. 186<br />

7.3.3 Results ....................................................................................................................... 187<br />

7.4 Test III ................................................................................................................................... 188<br />

7.4.1 Effects on Tissue Plasminogen Activator (tPA)-Producing Cells ............................. 188<br />

7.4.2 Platelet Aggregation Test .......................................................................................... 189<br />

7.4.3 Results ....................................................................................................................... 189<br />

7.5 Test IV .................................................................................................................................. 189<br />

7.5.1 In Vivo Fibrinolytic Activity of Euglobulin ............................................................. 189<br />

7.5.2 Effects on Blood Coagulation Activity ..................................................................... 189<br />

7.5.3 Results ....................................................................................................................... 189<br />

7.5.4 Discussion ................................................................................................................. 192<br />

7.6 Summary .............................................................................................................................. 193<br />

Acknowledgments .......................................................................................................................... 194<br />

References ...................................................................................................................................... 194<br />

We have presented reports that ingestion of various types of alcoholic drinks (Sumi et al. 1988,<br />

1998) and coffee (Sumi 1997) results in changes in blood coagulation–fibrinolysis systems; in particular,<br />

a rather lengthy period of promotion of fibrinolysis in the blood was observed with drinking<br />

of Oturui shochu liquors (distilled only once, retaining the character of the original ingredients and<br />

known as “authentic” shochu). Studies have been made on properties of the components bringing<br />

about such effects (Sumi 2003). It has also been generally recognized that essential oils have natural<br />

healing effects, and the psychological effect of inducing a feeling of relaxation. However, regarding<br />

the physiological effects of essential oils on blood circulation, especially their direct effects on<br />

blood coagulation–fibrinolysis systems, no report has yet been presented to our knowledge.<br />

Lemongrass being a well-known herb, we used the CLT method to check its effects on the<br />

blood coagulation system and the fibrinolysis system. Lemongrass has the effects of facilitating blood<br />

coagulation and inhibiting fibrinolysis, but it has been reported that the substantial citral and geraniol<br />

content in lemongrass contributes to strong inhibitory activity against platelet aggregation, producing<br />

fibrinolytic effects in vivo as a result. Imai et al. discovered that lemongrass oil, the essential oil<br />

component (Cymbopogon citrus) of lemongrass (oil yield: 0.2%), shows strong platelet-aggregation<br />

inhibitory activity (Imai et al. 1986). They reported that, when using platelet-rich plasma of humans,<br />

185


186 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

rabbits, and rats, aggregation through the application of ADP, collagen, and arachidonic acid was<br />

inhibited by adding lemongrass oil at an amount 1/12,000 of the amount of plasma. It was reported<br />

that the component of the oil was citral and that trans-citral was more effective than cis-citral.<br />

In the current study, we conducted tests regarding the effects of more than 100 types of essential<br />

oils (herbs) on the blood coagulation–fibrinolysis systems, using lemongrass as the control.<br />

7.2 test I<br />

Thirty types of essential oils purchased from Holistic Origin Pvt. Ltd. (Singapore) were used. The<br />

CLT test for essential oils was conducted in the following way:<br />

Solutions were mixed in a glass test tube (1.0 × 120 mm 2 ). They included 50 μL of a solution of<br />

each essential oil diluted with dimethyl sulfoxide (DMSO), 250 μL of 0.17 M borate-saline buffer<br />

with pH 7.8, 500 μL of bovine plasma fibrinogen manufactured by Sigma Co., Ltd. that was dissolved<br />

in 250 μL of 0.17 borate-saline buffer with pH 7.8 with a concentration of 0.6% and filtered<br />

with filter paper, and 100 μL of urokinase, a fibrinolytic enzyme in 20 IU/mL of saline. After incubation<br />

at 37°C for 1 min, 100 μL of bovine thrombin (2.25 IU/mL saline) was added and stirred,<br />

followed by another incubation. Measurements were then conducted for the time until coagulation<br />

and until complete lysis of the artificial thrombus (fibrin) were generated, in order to test the blood<br />

coagulation–fibrinolysis activity of the essential oils (Sumi 2000).<br />

7.2.1 re s u lt s<br />

As shown in the left part of Figure 7.1, lysis time was found to be longer for many samples in comparison<br />

with the control with no additions (broken line), which means that these samples had inhibitory<br />

effects on coagulation. Each value in the figure is the average value of five repetitions of the test.<br />

As shown on the right in Figure 7.1, lysis time was seen to be longer for many samples in comparison<br />

with the control with no additions (broken line), meaning these samples had inhibitory effects on<br />

fibrinolysis. Overall, there were considerable differences among the aroma essences with regard to<br />

the blood coagulation–fibrinolysis systems. As shown by the arrows, the results for lemongrass here<br />

were that it promoted coagulation (index 6/30 = 0.20) and reduced fibrinolysis (24/30 = 0.80).<br />

7.3 test II<br />

<strong>Essential</strong> oils (plant, animal, and synthetic aroma essences) supplied by Kanebo Cosmetics, Inc.,<br />

were used for the test.<br />

7.3.1 st a n d a r d fib r in Pl at e Me t h o d<br />

In a round, 90-mm-diameter petri dish, 30 μL of each of the samples, 30 μL of a solution with the<br />

sample diluted by an equal amount of 10 IU/mL urokinase, and 30 μL of 1% ethanol as the control<br />

were applied to the artificial thrombus (fibrin) plates, created with 10 mL of fibrinogen solution with<br />

a final concentration of 0.5% and 0.5 mL of 50 U/mL thrombin. After incubation at 37°C for 4 h,<br />

the area (mm 2 ) of the lysis area was measured.<br />

7.3.2 eu G l o b u l i n: fib r in Pl at e Me t h o d<br />

Citrate blood was drawn from the tail veins of etherized Wistar male rats, and the blood plasma<br />

obtained through centrifugal separation (3000 rpm, 10 min, room temperature) was diluted 20 times


Thrombolysis-Accelerating Activity of <strong>Essential</strong> <strong>Oil</strong>s 187<br />

2.5% Neroli in Jojoba<br />

2.5% Eucalyptus in Jojoba,<br />

Eucalyptusglobubus<br />

2.5% Chamomile in Jojoba, German<br />

2.5% Chamomile in Jojoba, Roman<br />

with 0.016% acetic acid. After being left standing for 30 min at 4°C, another centrifugal separation<br />

(3000 rpm, 5 min, room temperature) resulted in euglobulin fractions, which were dissolved with<br />

1/15 M phosphate buffer (pH 6.8) containing an amount of 0.9% sodium chloride equivalent to the<br />

amount of plasma. The euglobulin solution diluted with the same amount of each of the samples, 30<br />

μL of the mixture was applied to the fibrin plate and after incubation at 37°C for 4 h, the area (mm 2 )<br />

of the lysis area was measured.<br />

7.3.3 re s u lt s<br />

Citrus bergamia<br />

Pinus sylvestris<br />

Pogostemon patchouli<br />

Boswellia thurifera<br />

Santalum album<br />

Cananga odorata<br />

Vetiveria Aizanoides<br />

Rosmarinus officinalis<br />

Origanum marjorana<br />

Citrus aurantium<br />

Citratus Lemon<br />

Lavendula officinalis<br />

Cupressus sempervirens<br />

Juniperus virginiana<br />

Ocimum basiliicum<br />

Control<br />

Mentha piperita<br />

Salvia officinalis<br />

Thymus vulgaris<br />

Foeniculum Vulgare<br />

Commiphora myrrha<br />

Pelagonium graveolens<br />

25% Tea Tree in Jojoba,<br />

Melaleuca alternifolia<br />

Salvia sclarea<br />

Juniperus communis<br />

Cymbapogon Citratus<br />

2.5% Rose Otto in Jojoba<br />

0 10<br />

20 30 40 50 60<br />

Lysis Time (hr)<br />

FIgure 7.1 Effects of 30 types of aroma essences on the blood coagulation–fibrinolysis systems. Each<br />

sample dissolved to a concentration of 20% with DMSO, along with urokinase solution, was added to the<br />

fibrinogen solution. After adding thrombin solution, measurements were conducted for the time until coagulation<br />

(left, in minutes) and the time until the lysis of the coagulated fibrin (artificial thrombus).<br />

Lemongrass resulted in an index of 20/84 = 0.24 for the standard fibrin plate, 23/84 = 0.27 for<br />

ELT and for CLT, clotting promotion of 65/84 = 0.77 and a fibrinolytic reduction of 40/84 = 0.48.<br />

In comparison, results of the standard fibrin plate, ELT and CLT tests with the addition of 84<br />

types of diluted essential oils showed that elder, cashew, and grapefruit have strong fibrinolytic<br />

activity, whereas celery, fir tree, baca, olive, and rosemary have strong inhibitory effects against<br />

fibrinolysis.


188 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

7.4 test III<br />

Vetiveria Aizanoides<br />

Pogostemon patchouli<br />

Origanum marjorana<br />

Cymbapogon Citratus<br />

Rosmarinus officinalis<br />

Cananga odorata<br />

Commiphora myrrha<br />

Mentha piperita<br />

Lavendula officinalis<br />

Boswellia thurifera<br />

Salvia sclarea<br />

Juniperus virginiana<br />

Ocimum basiliicum<br />

Control<br />

25% Tea Tree in Jojoba,<br />

Melaleuca alternifolia<br />

Citratus Lemon<br />

Pelagonium graveolens<br />

Foeniculum Vulgare<br />

Cupressus sempervirens<br />

Juniperus communis<br />

Salvia officinalis<br />

Citrus aurantium<br />

Santalum album<br />

Citrus bergamia<br />

2.5% Chamomile in Jojoba, German<br />

2.5% Neroli in Jojoba<br />

2.5% Rose Otto in Jojoba<br />

2.5% Eucalyptus in Jojoba,<br />

Eucalyptusglobubus<br />

2.5% Chamomile in Jojoba, Roman<br />

FIgure 7.1 (continued).<br />

Thymus vulgaris<br />

Pinus sylvestris<br />

0 5 10 15 20 25<br />

Lysis Time (min)<br />

Similarly, 84 types of essential oils were used for this test. Pyrazin, its derivatives of 2-ethyl pyrazine,<br />

2,5-dimethyl pyrazine, 2,3,5-trimethyl pyrazine, and 2,6-dimethyl pyrazine and aspirin were<br />

purchased from Sigma–Aldrich, Inc.<br />

7.4.1 effeC ts o n ti s s u e Pl asMinoGen aC t iVat o r (tPa)-Pr o d u C i nG Cells<br />

Tissue plasminogen activator-free human cells were supplied by physiology Assistant Professor<br />

Yoshida of Miyazaki Medical University. After multiplying these cells using a 24-well microtest<br />

plate (Falcon) to confluence, the cultured solution inside the wells was removed using an aspirator.<br />

The cells were then washed twice with PBS(−), 450 μL of a new culture solution and 50 μL of each<br />

sample were added and incubated for 24 h, following which the culture solution was recovered (first<br />

medium). After again washing the cells twice with PBS(−), 500 μL of a new culture solution was<br />

added and incubated again for 24 h, and the culture solution was recovered (second medium). For<br />

the fibrinolytic activity, the area (mm 2 ) of the lysis area was measured through the standard fibrin<br />

plate method.


Thrombolysis-Accelerating Activity of <strong>Essential</strong> <strong>Oil</strong>s 189<br />

7.4.2 Pl at e l e t aG G r e G at i o n test<br />

As the aggregation-inducing substance, 22 μL of 300 μM ADP or of 1.2 mg/mL collagen (final<br />

concentration of 30 μM or 1.2 mg/mL) was added, and the platelet aggregation rate was measured<br />

for 5 min at 37°C using an aggregometer (PAT-4A: Mebanix). With 100% light transmittance set for<br />

PPP, platelet aggregation activity was measured.<br />

7.4.3 re s u lt s<br />

The effects of lemongrass on tPA-producing cells and platelet aggregation are shown with arrows.<br />

Regarding the comparative effects of the addition of other essential oils, promotion of tPA was<br />

observed for orange, basil, and clove (see Figure 7.2). In the case of orange in particular, adding<br />

450 μL of the cultured solution and 50 μL of the added sample and incubating for 24 h inside a 5%<br />

CO 2 incubator at 37°C (first medium) showed that, compared with the control, approximately eight<br />

times the fibrinolytic activity was seen.<br />

Regarding the effects on platelet aggregation, using ADP as the aggregation-inducing substance<br />

showed that the inhibitory activity of basil (77%) and tolu (64%) is fairly strong (see Figure 7.3).<br />

Platelet-aggregation inhibitory activity was also detected in coffee and white peach. The pyradine<br />

compounds contained in coffee had strong inhibitory effects on platelet aggregation, with strength<br />

equivalent to that of aspirin in many cases (Table 7.1).<br />

7.5 test IV<br />

As in the previous tests, 84 types of essential oils were used for this test.<br />

7.5.1 in Vi V o fi b r i n o ly t iC aCtiVity o f eu G l o b u l i n<br />

The aroma essences were diluted 200 times with 0.5% ethanol–0.9% sodium chloride solution, and<br />

5 mL per kg of body weight (aroma essence volume: 25 μL/kg) was orally administered to Wistar<br />

male rats using a feeding tube. Blood was drawn from the rat tail veins 1 h prior to oral administration<br />

and 1 and 2 h after administration.<br />

For the ELT, 0.1 mL of the blood plasma obtained from citrate blood through centrifugal separation<br />

(3000 rpm, 10 min, room temperature) was diluted 20 times with 0.016% acetic acid, and left<br />

standing for 30 min at 4°C. Centrifugal separation (3000 rpm, 5 min, room temperature) resulted<br />

in euglobulin fractions, which were dissolved with 0.1 mL of 0.1 M tris-hydrochloric acid buffer<br />

(pH 7.4) to obtain the sample. 90 μL of the euglobulin solution and 10 μL of 100 U/mL thrombin<br />

were mixed inside a 96-well microtest plate and incubated at 37°C, following which turbidity<br />

of the dissolving thrombi was measured every 10 min at 405 nm using a Well reader (SK601:<br />

Seikagaku Corp.).<br />

7.5.2 effeC ts o n bl o o d Co a G u l at i o n aCtiVity<br />

For the measurements, Data-Fi aPTT (Dade Behring) was used for the activated partial thromboplastin<br />

time (aPTT), and thromboplastin C plus (Dade Behring) was used for the prothrombin time (PT).<br />

All the animals used in the tests were healthy and all provisions of the Declaration of Helsinki<br />

(1964) were met.<br />

7.5.3 re s u lt s<br />

Finally, several essential oils were orally administered to rats and changes in their blood were<br />

tested for through ELT. For mainly coffee and lemongrass, the reduction of ELT was observed,


190 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Lemongrass<br />

Almond<br />

Angelica<br />

Artemisia 1<br />

Artemisia 2<br />

B.Megastigma<br />

Baka<br />

Balsamic<br />

Basil<br />

Cacao beans<br />

Cachute<br />

Cafe<br />

Camomile<br />

Cardamon<br />

Carnation<br />

Carrot<br />

Celery<br />

Citronellol<br />

Clove<br />

Coconuts<br />

Coffee<br />

Coriander<br />

Crocus<br />

Curry<br />

Elder<br />

Fir tree<br />

Fusel oil<br />

Galbanum<br />

Geraniol<br />

Ginger<br />

Grape fruits<br />

Helichrysum<br />

Herb 1<br />

Herb 2<br />

Hibiscus<br />

Immortelle<br />

Ionon A<br />

Ionon B<br />

Isopropanol<br />

Jatamansi<br />

Lemongrass<br />

a b c d e f g h<br />

a b c d e f g h<br />

Zymography of the conditioned media<br />

by HeLa S3 cells in the presence of<br />

aroma essences<br />

Loading sample was 20 µ l, and<br />

incubation time was 24 hrs. at 37°C.<br />

I: first medium, II: second medium.<br />

(a) Clove, (b) Violet-like synthetic<br />

aroma 1, (c) Resin 2, (d) Basil, (e) 0.1%<br />

EtOH, (f) Milli Q, (g) 2 IU/ml tPA,<br />

(h) 1 IU/ml UK.<br />

Jonquil<br />

Juniper<br />

Lacton synthetic aroma<br />

Lavender<br />

Linalool<br />

Lycoris<br />

Musklike synthetic aroma<br />

Nerol<br />

Niaouli<br />

Nigelle<br />

Olibanum<br />

Olive<br />

Opoponax<br />

Orange<br />

Parsley<br />

Pepper<br />

Pistacia<br />

Porto<br />

Reglis<br />

Resin 1<br />

Resin 2<br />

Resin 3<br />

Riz<br />

Rose<br />

Rosemary 1<br />

Rosemary 2<br />

Rovensara<br />

Rum<br />

Sage 1<br />

Sage D<br />

Sage S<br />

Shobu<br />

Synthetic aroma<br />

α-Terpineol<br />

Tolu<br />

Valerian<br />

Violetlike synthetic aroma 1<br />

Violetlike synthetic aroma 2<br />

Violetlike synthetic aroma 3<br />

Violetlike synthetic aroma 4<br />

Whisky<br />

White peach 1<br />

White peach 2<br />

Yuzu<br />

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0<br />

Ratio (sample/0.1% EtOH)<br />

FIgure 7.2 tPA activity generated through cultivation of tissue plasminogen activator-free cells. Further<br />

isolation of each activity through zymography possible (partial results shown). Average values (n = 3) shown;<br />

■: first medium, □: second medium.


Thrombolysis-Accelerating Activity of <strong>Essential</strong> <strong>Oil</strong>s 191<br />

Lemon grass<br />

Almond<br />

Angelica<br />

Artemisia 1<br />

Artemisia 2<br />

B.Megastigma<br />

Baka<br />

Balsamic<br />

Basil<br />

Cacao beans<br />

Cachute<br />

Cafe<br />

Camomile<br />

Cardamon<br />

Carnation<br />

Carrot<br />

Celery<br />

Citronellol<br />

Clove<br />

Coconuts<br />

Coffee<br />

Coriander<br />

Crocus<br />

Curry<br />

Elder<br />

Fir tree<br />

Fusel oil<br />

Galbanum<br />

Geraniol<br />

Ginger<br />

Grape fruits<br />

Helichrysum<br />

Herb 1<br />

Herb 2<br />

Hibiscus<br />

Immortelle<br />

Ionon A<br />

Ionon B<br />

Isopropanol<br />

Jatamansi<br />

Jonquil<br />

Juniper<br />

Lacton synthetic aroma<br />

Lavender<br />

Linalool<br />

Lycoris<br />

Musklike synthetic aroma<br />

Nerol<br />

Niaouli<br />

Nigelle<br />

Olibanum<br />

Olive<br />

Opoponax<br />

Orange<br />

Parsley<br />

Pepper<br />

Pistacia<br />

Porto<br />

Reglis<br />

Resin 1<br />

Resin 2<br />

Resin 3<br />

Riz<br />

Rose<br />

Rosemary 1<br />

Rosemary 2<br />

Rovensara<br />

Rum<br />

Sage 1<br />

Sage D<br />

Sage S<br />

Shobu<br />

Synthetic aroma<br />

α-Terpineol<br />

Tolu<br />

Valerian<br />

Violetlike synthetic aroma 1<br />

Violetlike synthetic aroma 2<br />

Violetlike synthetic aroma 3<br />

Violetlike synthetic aroma 4<br />

Whisky<br />

White peach 1<br />

White peach 2<br />

Yuzu<br />

Lemongrass<br />

Right Transmission(%)<br />

0 10<br />

20 30 40 50 60 70 80 90<br />

Inhibition (%)<br />

Incubation Time (min)<br />

4 3<br />

Inhibitory effect of aroma essences on ADP aducted<br />

platelet aggregation.<br />

Platelet aggregation was inhibited after preincubation<br />

of 50 µl rat-PRP, 100 µl Tyrode buffer and aroma<br />

essences for 5 min at 37°C. Final concentration of ADP<br />

in the reaction mixture was 30 µM. For determination,<br />

aggregometer PAT-4A was used with continuous stirring<br />

at 1000 rpm.<br />

1 : Basil, 2 : Whisky, 3 : Juniper, 4 : Control<br />

(0.225% EtOH)<br />

FIgure 7.3 Effects of essential oils on platelet aggregation activity. For rat blood platelets, 30 μM ADP<br />

used as aggregation-inducing substance. Aggregation patterns from aggregometer shown. Average values<br />

(n = 3) shown.<br />

1<br />

2


192 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 7.1<br />

antiplatelet aggregation effects of Pyradine compounds<br />

Pyridine<br />

2-ethyl<br />

pyridine<br />

2,5-dimethyl<br />

pyridine<br />

2,3,5-trimethyl<br />

pyridine<br />

2,6-dimethyl-<br />

3-methyl<br />

pyridine<br />

Aspirin<br />

H 3 C<br />

H3 C<br />

H 3 C<br />

meaning a tendency to promote fibrinolysis (Figure 7.4). The aPTT (38 ± 7 s) and PT (12 ± 3 s)<br />

values as a result of the effects of coffee and lemongrass were investigated, but the changes were<br />

not significant.<br />

7.5.4 di sC u s s i o n<br />

components aggregation 50 mg/ml 5 mg/ml 1 mg/ml 0.5 mg/ml<br />

N<br />

N<br />

COOH<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

C 2 H 5<br />

CH 3<br />

CH 3<br />

CH 3<br />

CH 3<br />

CH 3<br />

OOCCH 3<br />

Collagen ||| || | —<br />

ADP ||| || | —<br />

Collagen ||| ||| || —<br />

ADP ||| ||| ||| |<br />

Collagen ||| ||| | —<br />

ADP ||| ||| — —<br />

Collagen ||| ||| — —<br />

ADP ||| ||| ||| |<br />

Collagen ||| ||| — —<br />

ADP ||| || | —<br />

Collagen ||| ||| | —<br />

ADP ||| ||| || —<br />

Note: In the order of the strength of platelet aggregation, reactions as viewed with the naked eye are shown.<br />

Various tests related to the blood coagulation–fibrinolysis systems were conducted using essential<br />

oils I and II, but the results were not consistent. However, it seemed that there were effects on the<br />

promotion of tPA from vascular endothelial cells and on platelets, and a fibrinolytic tendency was<br />

observed in the test using lemongrass cells. In the case of coffee, it seemed that there were effects<br />

of the activity of pyridine compounds in promoting fibrinolysis. Yamamoto (2003) selected 6 out of


Thrombolysis-Accelerating Activity of <strong>Essential</strong> <strong>Oil</strong>s 193<br />

ELT(min)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

26 types of herbs for their blood fluidity, through a test in which each herb (aroma essence) was<br />

added to blood and fluidity of the blood was measured. The six types selected were echinacea,<br />

thyme, hip, marigold, lavender, and lemongrass.<br />

The primary purpose of aroma therapy using essential oils is deemed to be bringing about relaxation,<br />

but many essential oils have antibacterial activity and are effective in the sterilization of skin.<br />

Takarada et al. (2004) used manuka oil and rosemary oil to compare and study their antibacterial<br />

activities against bacteria in the mouth, which cause dental caries and periodontitis. Many applications<br />

have recently been filed for patents related to lemongrass. Examples include a stress reliever<br />

(Suntory Ltd. 2003) containing aromatic components of alcoholic drinks as effective components,<br />

bath tablets (Cosme Park 2005) containing herb ingredients, fragrances (Nitto Pharmaceutical<br />

2005) containing 0.1%–5% lemongrass oil per weight, and an aroma essence for increasing memory<br />

(Pola Chemical Industries 2001).<br />

If practical effects can be achieved merely by smelling aroma essences rather than ingesting<br />

them, it is possible that aromatic components may constitute a wholly new category of functional<br />

materials. In tests adding aromatic components using sweet potato-based shochu liquor, we have<br />

already shown that the amount of fibrinolytic enzymes produced from human vascular endothelial<br />

cells increases (Sumi et al. 2001). It has also been reported that many aroma essences are generated<br />

at the time of heating and distilling unprocessed shochu liquor for Otsurui shochu liquors, particularly<br />

sweet potato-based shochu liquors (Ota 1991). A more detailed analysis of each essential<br />

oil and identification of the in vivo effects of each essential oil would be issues to address from<br />

now on.<br />

7.6 summary<br />

Control<br />

(0.5% EtOH)<br />

Lemongrass Baca Clove Coffee Rosemary 1<br />

FIgure 7.4 Oral administration. Changes in the blood caused by oral administration of several essential<br />

oils as measured by plasma ELT. Values are means ± SE (n = 4); ■, 1 h before administration; □, 1 h after<br />

administration; , 2 h after administration.<br />

Tests conducted on more than 100 types of essential oils showed that some of them had fibrinolytic<br />

activity similar to lemongrass. However, no correlation was found among the results of the standard<br />

fibrin plate, ELT, and CLT methods. It is thought that promoting the release of tissue plasminogen<br />

activators from cells and inhibiting platelet aggregation are significant reactions, which were<br />

revealed to be the result of complex configurations.


194 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

acKnoWledgments<br />

We are grateful to Mr. S. Murakami for excellent technical assistance. <strong>Essential</strong> oils were kindly<br />

supplied by Kanebo Cosmetics, Inc. (Kanagawa).<br />

reFerences<br />

Cosme Park. 2005. Bath tablets, Published Patent Application 112794 (in Japanese).<br />

Imai H, Tamada T, Sawai Y, Yoshida M, Takao K, Endo E, Ohshiba S, Kojisawa Y. 1986. Blood platelet<br />

aggregation inhibitory activity of lemongrass oil. 28th Convention of the Japanese Society of Clinical<br />

Hematology, Extracts, 279 (in Japanese).<br />

Nitto Pharmaceutical Industries Ltd. 2005. Fragrances, Published Patent Application 023044 (in Japanese).<br />

Ota T. 1991. Aroma of sweet potato-based shochu liquor, Tokyo 86: 250–254 (in Japanese).<br />

Pola Chemical Industries Inc. 2001. Aroma essences and components containing aroma essences for increasing<br />

memory, Published Patent Application 288493 (in Japanese).<br />

Sumi H, Hamada H, Tsushima H, Mihara H. 1988. Urokinase-like plasminogen activator increased in plasma<br />

after alcohol drinking. Alcohol Alcoholism 23: 33–43.<br />

Sumi H, Iijima Y, Komine S, Sasahira T. 2001. Kagoshima Industry Support Center, Report on Commissioned<br />

Research and Development Project 2000: 1–6.<br />

Sumi H. 1997. Abnormalities in the blood coagulation and fibrinolysis systems. Nippon Rinsho 712: 233–239<br />

(in Japanese).<br />

Sumi H. 2000. Blood circulation: Analysis of foods and aroma essences related to blood coagulation–<br />

fibrinolysis systems. Food Technology 20: 64–70 (in Japanese).<br />

Sumi H. 2003. Improvements in blood circulation and disease prevention by food components, Food Style<br />

21(7): 47–53 (in Japanese).<br />

Sumi H, Kozaki Y, Yatagai C, Hamada H. 1998. Effects of wine on plasma fibrinolytic and coagulation systems.<br />

Japanese Journal of Alcohol Studies and Drug Dependence 33: 263–272.<br />

Suntory Ltd. 2003. Stress reliever, Published Patent Application 171286 (in Japanese).<br />

Takarada K, Kimizuka R, Takahashi N, Honma K, Okuda K, Kato T. 2004. A comparison of the antibacterial<br />

efficacies of essential oils against oral pathogens. Oral Microbiology and Immunology 19(1): 61–64.<br />

Yamamoto N. 2003. Effects of herbs on blood fluidity. Food Style 21: 61–63 (in Japanese).


8<br />

contents<br />

8.1 IntroductIon<br />

Analytical Methods<br />

for Cymbopogon <strong>Oil</strong>s<br />

Ange Bighelli and Joseph Casanova<br />

8.1 Introduction .......................................................................................................................... 195<br />

8.2 Methods for EO Analysis: A Summary ................................................................................ 196<br />

8.2.1 Chemical Methods and Fractional Distillation......................................................... 196<br />

8.2.2 Chromatographic Techniques ................................................................................... 197<br />

8.2.2.1 Thin-Layer Chromatography (TLC) .......................................................... 197<br />

8.2.2.2 Gas Chromatography ................................................................................. 197<br />

8.2.3 Hyphenated Techniques ............................................................................................ 198<br />

8.2.3.1 GC-MS and GC-MS Combined with GC(RI) ........................................... 198<br />

8.2.3.2 GC×GC-MS ............................................................................................... 199<br />

8.2.3.3 GC-MS-MS ................................................................................................ 199<br />

8.2.3.4 GC-FTIR, GC-MS-FTIR ...........................................................................200<br />

8.2.3.5 HPLC-MS, HPLC- 1 H NMR, and HPLC-GC-MS .....................................200<br />

8.2.3.6 <strong>Essential</strong> <strong>Oil</strong> Analysis by 13 C NMR ..........................................................200<br />

8.2.3.7 Enantiomeric Differentiation ..................................................................... 201<br />

8.2.3.8 Two-Step Procedure ...................................................................................202<br />

8.3 Analysis of Cymbopogon <strong>Oil</strong>s ..............................................................................................203<br />

8.3.1 Earliest Studies: Analysis by Chemical Methods (Isolation of Compounds and<br />

Miscellaneous Methods) ...........................................................................................204<br />

8.3.2 Analysis by Chromatographic Techniques [TLC, HPLC, GC(RT) or GC(RI)] ......205<br />

8.3.3 Analysis by Mass Spectrometry Coupled with Gas Chromatography .....................206<br />

8.3.4 Combined Analysis by Chromatographic and Spectroscopic Techniques: TLC,<br />

CC, GC(RT) or GC(RI) and GC-MS .......................................................................206<br />

8.3.4.1 Cymbopogon <strong>Oil</strong>s of Commercial Interest ................................................207<br />

8.3.5 Other Cymbopogon <strong>Oil</strong>s ...........................................................................................209<br />

8.3.6 Enantiomeric Differentiation by Chiral GC ............................................................. 210<br />

8.3.7 Combined Analysis of Cymbopogon <strong>Oil</strong>s by Various Techniques: TLC, CC,<br />

IR, GC(RI), GC-MS, and 13 C NMR ......................................................................... 211<br />

8.3.8 Analysis of C. Giganteus <strong>Oil</strong>: A Summary of Analytical Methods Involved in<br />

the Analysis of Cymbopogon <strong>Oil</strong>s ............................................................................ 212<br />

8.4 Conclusion ............................................................................................................................ 215<br />

Acknowledgments .......................................................................................................................... 215<br />

References ...................................................................................................................................... 215<br />

The Cymbopogon genus comprises 56 species, and most of them produce essential oil (EO) by<br />

hydrodistillation or steam distillation of aerial parts. Various Cymbopogon species, such as C. nardus<br />

(L.) Rendle, C. winterianus Jowitt, C. citratus Stapf, C. flexuosus Stapf and C. martinii (Roxb.)<br />

195


196 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Wats., are well known as sources of commercially valuable compounds such as geraniol, geranyl<br />

acetate, citral (neral and geranial), and citronellal (Surburg and Panten 2006). The other species<br />

contribute to the conservation of the biodiversity through the world.<br />

EOs, which are the raw materials used in the perfume, fragrance, and flavor industries, are, in<br />

general, complex mixtures of more than 100 individual components, containing mainly monoterpenes<br />

and sesquiterpenes that present various frameworks and functions. The identification of these<br />

compounds would be carried out using various techniques. Obviously, the choice of the analytical<br />

methods depends on whether the final objective is structural identification of an unknown substance<br />

or recognition of a substance that has already been identified.<br />

In this chapter, we first summarize the analytical methods developed for identification of individual<br />

components of EOs; then we apply these techniques and methods to the analysis of Cymbopogon<br />

EOs, including studies carried out in our laboratories using a combination of various analytical<br />

techniques.<br />

8.2 methods For eo analysIs: a summary<br />

Identification of the individual components of an EO is a difficult task that first requires an adequate<br />

choice of analytical technique and then a careful utilization of the instruments, software, spectral<br />

data libraries and, sometimes, statistical tools.<br />

Depending on the objective of analysis (routine, quality control, detailed analysis of an EO investigated<br />

for the first time, etc.) and the complexity of the mixture (number and structure of the components<br />

and heat sensitivity), various schemes can be drawn up.<br />

In routine analysis, identification of a component was achieved by combination of “on-line”<br />

chromatographic separation and spectroscopic identification. For instance, a fast-scanning mass<br />

spectrometer directly coupled with a gas chromatograph is the basic equipment of control in the<br />

field of EO analysis. Nevertheless, misuse—or abuse—of modern instrumentation is not unknown.<br />

Nowadays, a conventional analysis of an EO is carried out by matching against computerized mass<br />

spectral libraries and comparison of the retention indices with those of authentic samples (GC[RI]<br />

and GC-MS). However, more sophisticated techniques are needed for the analysis of EOs, and<br />

a wide variety of combinations of analytical techniques are being developed, including complex<br />

hybrids such as GC-MS-FTIR, HPLC-GC-MS, or HPLC- 1 H NMR. In a similar fashion, individual<br />

components of an EO may be identified by comparison of the chemical shift values in the 13 C NMR<br />

spectrum of the mixture with those of reference compounds.<br />

Despite the improvement of these on-line analytical techniques, the misidentification of some<br />

compounds, especially sesquiterpenes, still occurs. For this reason, several research groups developed<br />

a two-step procedure in which a small quantity of a substance is separated by a chromatographic<br />

technique and then identified by comparison of its spectral data, including 1 H NMR and<br />

sometimes 13 C NMR, with those of reference compounds. This “off-line” sequence is obviously<br />

more accurate but very time consuming.<br />

Nowadays, in numerous research laboratories, the chemical composition of complex EOs is analyzed<br />

by combination of various techniques, the oil being prefractionated by fractional distillation<br />

and/or column chromatography and, eventually, by HPLC or preparative gas chromatography (PGC).<br />

Prefractionation of the EO considerably reduces the number of coeluted compounds. Then, analysis of<br />

the fractions is achieved by combination of spectroscopic techniques, such as MS, IR, and/or NMR.<br />

8.2.1 Ch e M iC a l Me t h o d s a n d fr a C t i o n a l distill ation<br />

The procedures that were mostly used during the last century for the isolation of EO constituents<br />

were chemical methods and fractional distillation. Concerning chemical methods, a preparative<br />

separation of an EO into different groups of components was achieved. In this procedure, the<br />

oil was successively treated by aqueous sodium carbonate (5%), aqueous sodium hydroxide (5%),


Analytical Methods for Cymbopogon <strong>Oil</strong>s 197<br />

methanolic sodium hydroxide (0.5N), sodium hydrogen sulfite or Girard reagent and phthalic anhydride<br />

to separate various families of compounds: acids, phenols, esters and lactones, aldehydes<br />

and ketones, and alcohols (primary and secondary). Esters and lactones can be hydrolyzed, and the<br />

resulting acids separated as salts. The final residue, consisting of tertiary alcohols, ethers, and hydrocarbons,<br />

must be separated by physical methods such as fractional distillation, crystallization, or<br />

liquid–liquid extraction (on the basis of the difference of solubility of the components in various<br />

solvents) (Kubeczka 1985).<br />

Fractional distillation had an important role in the preparative isolation of EO constituents,<br />

although it is seldom possible to obtain chemically pure components by this technique. The development<br />

of efficient distillation columns (spinning-band columns) enables one to fractionate milliliter<br />

amounts of EOs under mild conditions. For instance, by using a so-called Spaltrohr column,<br />

amounts of a few milliliters may be distilled at reduced pressure, which allows the distillation<br />

of thermally unstable components. With this type of column, up to 100 theoretical plates may<br />

be achieved. The efficiency of this device is clearly demonstrated by the separation of citronellyl<br />

acetate and geranyl acetate (boiling points at 1 mbar: 73.5°C and 77.0°C, respectively). However,<br />

pure compounds from EOs are rarely obtained by fractional distillation since isomerization and<br />

decomposition of labile components occasionally take place (Kubeczka, 1985).<br />

8.2.2 Ch r o M at o G r a P h iC te C h n i q u e s<br />

8.2.2.1 thin-layer chromatography (tlc)<br />

Thin-layer chromatography is a type of liquid chromatography that can separate different chemical<br />

compounds based on the rate at which they move through a support under defined conditions.<br />

The support, known as the plate, is a layer of support material (silica gel or alumina) that has been<br />

spread out and dried on a sheet of material such as glass. The mobile phase is a solvent or a mixture<br />

of solvents. Individual components of a mixture are identified after separation by migration, by<br />

comparison of their Rf values with those of reference compounds obtained under the same experimental<br />

conditions. Due to its simplicity and speed, TLC was an important pioneering method in the<br />

analysis of EOs (Kubeczka 1985; Rouessac and Rouessac 1994).<br />

TLC can also be used for preparative separation. By means of plates up to 1 m long and a layer<br />

thickness up to 2 mm, several grams of a mixture of compounds can be applied as a line. The various<br />

bands are scraped off, collected, and the components eluted with appropriate solvents.<br />

8.2.2.2 gas chromatography<br />

Gas chromatography (GC), which allows the separation of volatile compounds from complex mixtures<br />

as well as their quantification, has become one of the most important tools in the analysis of<br />

EOs (Kubeczka, 1985). Nowadays, the development of efficient capillary columns allows the individualization<br />

of several hundreds of components. For instance, more than 200 compounds have been<br />

distinguished in the chromatogram of labdanum oil (Cistus ladanifer oil) or vetiver oil (Weyerstahl<br />

et al. 1998, 2000). Concerning the identification of individual components, the comparison of their<br />

retention times with those of reference compounds is generally not sufficient, even using two columns<br />

of different polarity. For that reason, utilization of retention indices is generally preferred.<br />

Retention indices are determined relative to the retention times of a series of n-alkanes with constant<br />

temperature (Kováts indices [KIs]) (Kováts 1965) or with linear temperature programmed<br />

(retention indices [RIs]) (Van Den Dool and Kratz 1963). Identification is then carried out by comparison<br />

of retention indices with those of authentic compounds or literature data (Figure 8.1).<br />

However, identifying sesquiterpenes (or diterpenes) by comparison of their RIs remains a difficult<br />

task. Indeed, as mentioned by Joulain (1994), more than 230 natural sesquiterpenes, which possess<br />

a molecular mass of 204, have been reported, and the values of their GC RIs differ by less than<br />

30 points. Moreover, RI values reported in the literature for the same compound can vary by more


198 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

GC-MS<br />

Chromatogram<br />

Mass spectrum<br />

Computer program<br />

Mass spectra libraries<br />

(Joulain, Nist, Wiley)<br />

• Comparison with<br />

reference spectra<br />

• Structure proposal<br />

than 10 points, particularly those concerning the polar column (Grundschober 1991). Consequently,<br />

identification of all the components of a complex EO using only GC appeared very difficult, if not<br />

impossible, for sesquiterpenes, even by taking into account their RIs on polar and apolar columns.<br />

8.2.3 hy P h e n at e d te C h n i q u e s<br />

ESSENTIAL OIL<br />

GC(RI)<br />

Chromatogram (apolar and<br />

polar columns)<br />

RI apolar<br />

RI polar<br />

Reference Indices<br />

RIs apolar<br />

RIs polar<br />

Comparison with<br />

RIs of reference<br />

compounds<br />

Identification of individual components<br />

13 C NMR<br />

<strong>Essential</strong> oil spectrum<br />

Computer program<br />

Spectral data libraries<br />

• «Laboratory» library<br />

• «Literature» library<br />

• Number of observed signals<br />

• Number of overlaped signals<br />

• Chemical shift variations<br />

FIgure 8.1 Combined analysis of essential oils by GC-MS, GC(RI) and 13 C NMR.<br />

8.2.3.1 gc-ms and gc-ms combined with gc(rI)<br />

A fast-scanning mass spectrometer using electronic impact (EI) mode, directly coupled with a gas<br />

chromatograph (GC-MS), is the basic equipment of control in the field of EO analysis. The two<br />

most-used analyzers in this field are the quadrupole and ion-trap types. Identification of the components<br />

in the oil is achieved by computer matching with laboratory-made or commercial mass spectral<br />

libraries containing several thousands spectra (Figure 8.1). Among the best-known commercial<br />

computerized libraries, we could mention the Wiley Registry of Mass Spectral Data (McLafferty<br />

and Stauffer 1994), the National Institute of Standards and Technology EPA/NIH Mass Spectral<br />

Library (NIST 1999), and the Terpenoids and Related Constituents of <strong>Essential</strong> <strong>Oil</strong>s library (König<br />

et al. 2001). Comparison of experimental spectra may be done with those of literature data compiled<br />

in various paper libraries published, among others, by McLafferty and Stauffer (1988), Joulain


Analytical Methods for Cymbopogon <strong>Oil</strong>s 199<br />

and König (1998), and Adams (2007). Better reliability is generally obtained when comparing the<br />

experimental spectra with those of reference compounds recorded in the laboratory with the same<br />

experimental conditions.<br />

Analysis of EOs by GC-MS began in the 1960s and was applied to common EOs of commercial<br />

interest. However, it is quite difficult to avoid misidentification of some compounds, especially those<br />

that possess insufficiently differentiated mass spectra. This problem concerns not only epimers<br />

such as α-cedrene and α-funebrene, stereoisomers such as (Z)-β-farnesene and (E)-β-farnesene, or<br />

diastereoisomers such as α-bisabolol and α-epi-bisabolol, but also compounds with different skeletons,<br />

such as (E,Z)-α-farnesene (linear sesquiterpene) and cis-α-bergamotene (bicyclic sesquiterpene)<br />

(Schultze et al. 1992), or 1-endo-bourbanol (tricyclic sesquiterpene) and 1,6- germacradien-5-ol<br />

(monocyclic sesquiterpene) (Joulain and Laurent 1989).<br />

For that reason, and although analyses by GC-MS are unfortunately still reported in the literature,<br />

identification of a component is much more accurate when GC-MS is used in combination with<br />

GC(RI) (Figure 8.1). Today, it is considered by specialized reviews that analysis of an EO by at least<br />

two techniques, for instance, MS in combination with RIs measured on two columns of different<br />

polarity, should be done to make the work publishable.<br />

Vernin et al. (1986, 1990) developed a computerized procedure based on the combination of<br />

mass spectral data and retention index on polar and apolar columns to identify the individual components<br />

of several EOs. Cavaleiro (2001) used the same procedure for the characterization of several<br />

Juniperus EOs.<br />

In some cases, positive or negative chemical ionization mass spectrometry [(PCI)MS or (NCI)<br />

MS] allows the identification of compounds that possess similar mass spectra. Chemical ionization<br />

produces mass spectra that exhibit “quasi-molecular” ions from which the molecular mass of the<br />

compounds can be easily deducted; that is not always possible using EI. As in the case of EI mode,<br />

the quadrupole or ion-trap analyzer can be used in the chemical ionization mode. Utilization of<br />

(CI)MS allowed the differentiation of alcohols, such as the four isomers of isopulegol (Lange and<br />

Schultze 1988a), or esters (Lange and Schultze 1988b), whereas the mass spectra of these compounds<br />

are generally similar when EI mode is used. In the same way, several isomers bearing the<br />

pinane skeleton (α-pinene, β-pinene, and cis-δ-pinene; pinocamphone and isopinocamphone; nopinone<br />

and isonopinone) were differentiated by (PCI)MS by using ammoniac as reactive gas (Badjah<br />

Hadj et al. 1988). The choice of the reactive gas (methane, isobutene, or ammoniac for PCI and a<br />

mixture of N 2O and methane or ammoniac for NCI) to differentiate some isomers depends on the<br />

structure and the functionalization of the molecules under consideration. For instance, GC-(CI)<br />

MS with isobutane and ammonia chemical ionization allowed the distinguishing of sesquiterpene<br />

hydrocarbons (Schultze et al. 1992). (PCI)MS and (NCI)MS proved to be useful for the analysis<br />

of complex EOs (Paolini et al. 2005). However, due to the difficulty of obtaining reproducible spectra,<br />

chemical ionization mass spectrometry is not usable as unique technique, but it is recommended<br />

in combination with GC-(EI)MS for the analysis of complex EOs.<br />

8.2.3.2 gc×gc-ms<br />

In some very complex mixtures in which the low peak resolution in GC prevents good characterization,<br />

the use of two-dimensional GC (GC×GC) can be useful to improve the resolution (Dugo et al.<br />

2000). For instance, using this two-dimensional technique, Mondello et al. (2005) eluted separately<br />

β-bisabolene, neryl acetate, and bicyclogermacrene in an EO of citrus, these three compounds being<br />

coeluted by conventional GC.<br />

8.2.3.3 gc-ms-ms<br />

Combination of GC and 2D mass spectrometry (GC-MS-MS) has been used for the analysis of complex<br />

EOs. For instance, Decouzon et al. (1990) clearly identified the four stereoisomers of dihydrocarveol<br />

using MS-MS (NCI) by observation of some characteristic fragments, which is not possible


200 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

by monodimensional mass spectrometry [(NCI)MS]. Cazaussus et al. (1988) identified khusimone,<br />

a particular sesquiterpene, by GC-MS-MS in EI mode, in the oxygenated fraction of a complex<br />

vetiver EO that contained more than 100 components.<br />

8.2.3.4 gc-FtIr, gc-ms-FtIr<br />

The improvement of the threshold of detection of FTIR in the last decade allowed the development<br />

of this on-line technique for the analysis of various families of organic compounds (Coleman et al.<br />

1989). Concerning EO components, GC-FTIR is particularly well suited to identifying compounds<br />

that possess insufficiently differentiated mass spectra. For instance, this technique has been used<br />

to identify two sesquiterpene alcohols, 1-endo-bourbanol (tricyclic) and 1,6-germacradien-5-ol<br />

(monocyclic) (Joulain and Laurent 1989), on the basis of their infrared mass spectra combined with<br />

their RIs. CG-FTIR is also well suited for analysis of compounds that suffer rearrangement in MS,<br />

such as germacrene-B and bicyclogermacrene, correctly identified by this method in citrus EOs<br />

(Chamblee et al. 1997). Conversely, it has been also reported that GC-FTIR was not appropriate to<br />

distinguish β-gurjunene and cis-1,3-dimethylcyclohexane, two compounds that possess quite different<br />

structure (Hedges and Wilkins 1991).<br />

It seems that GC-FTIR is not really appropriate for the analysis of EOs, but it provides useful<br />

information when utilized in combination with GC-MS. For that reason, hyphenated techniques<br />

such as GC-FTIR-MS have been developed and successfully used in the field of EOs. For example,<br />

Hedges and Wilkins (1991) identified, by GC-FTIR-MS, 26 minor components together with<br />

the main compound, 1,8-cineole, in an EO from Eucalyptus australiana. A better separation of the<br />

components was obtained when multidimensional GC is directly combined with FTIR and MS for<br />

analysis of eucalyptus EOs (Krock et al. 1994).<br />

8.2.3.5 hPlc-ms, hPlc- 1 h nmr, and hPlc-gc-ms<br />

Some EOs contain a variable amount of poor volatile compounds. Their analysis can be carried out<br />

by HPLC directly coupled with MS. Utilization of these combined techniques allowed the identification<br />

of oxygen heterocyclic compounds, coumarins, psoralenes, and flavones in citrus EOs (Dugo<br />

et al. 2000).<br />

More recently, HPLC has also been combined with 1 H NMR, which provides important structural<br />

information. Two methods can be used: (1) stopped flow, in which the elution is stopped to<br />

record the spectra of a component, and (2) continuous flow, in which the spectra of a component<br />

is recorded during the elution of the liquid. Today, the use of high-field spectrometers (up to 21 T)<br />

allows one to record spectra of pure compounds with a quantity near to 1 ng (Korhammer and<br />

Bernreuther 1996).<br />

Even though HPLC- 1 H NMR is mainly used in the pharmaceutical industry, it was also regularly<br />

used to identify some terpenes such as three sesquiterpene lactones present in a solvent extract of<br />

Zaluzania grayana (Spring et al. 1995).<br />

With the aim of limiting the number of coeluted compounds that can occur in complex EOs,<br />

HPLC was combined with another chromatographic technique such as GC before spectroscopic<br />

identification. This method (HPLC-GC-MS) has been used to identify very minor compounds in<br />

citrus EOs (Mondelo et al. 1995, 1996).<br />

8.2.3.6 essential oil analysis by 13 c nmr<br />

Since its discovery, NMR has been a powerful tool for structure elucidation of organic molecules.<br />

Even 1 H NMR spectra recorded with low field spectrometers provided structural information not<br />

available with other techniques at that time. The introduction of high-field spectrometers, combined<br />

with the discovery of Fourier Transform NMR and two- and three-dimensional NMR (2D and 3D<br />

NMR), made the use of that technique essential in structure elucidation of natural compounds isolated<br />

from plants (Derome 1987).


Analytical Methods for Cymbopogon <strong>Oil</strong>s 201<br />

Concerning the analysis of natural mixtures, the chief idea was to take benefit of the information<br />

provided by 13 C NMR, avoiding, or at least limiting as far as possible, the separation steps.<br />

Carbon-13 is preferred to proton, despite its low natural isotopic abundance (1.1%), for several<br />

reasons:<br />

1. Carbon constitutes the backbone of organic molecules, and the slightest structural modification<br />

induces measurable chemical shift variations.<br />

2. The sweep width, around 240 ppm, is much wider than that of protons, leading to a better<br />

spectral dispersion.<br />

3. 13 C NMR spectra can easily be simplified by complete decoupling of protons, so that each<br />

signal appears as a singlet.<br />

4. Transverse relaxation time, T 2, is higher than that of the proton, leading a better resolution.<br />

5. To avoid any thermal degradation, spectra were recorded at room temperature. Furthermore,<br />

the sample may be recovered and then submitted to other spectroscopic analyses.<br />

In their pioneering work, Formácek and Kubeczka (1982a,b) used 13 C NMR, in combination with<br />

GC, to confirm the occurrence of a compound previously identified (or suspected) by GC. Since that<br />

time, 13 C NMR has been regularly employed, in various laboratories, to contribute to the analysis of<br />

EOs (De Medici et al. 1992; Alencar et al. 1997; Ramidi et al. 1998; Ahmad and Jassbi 1999; Núñez<br />

and Roque 1999; Ferreira et al. 2001a, 2001b; Al-Burtamani et al. 2005). A computerized procedure<br />

has been proposed by Ferreira et al. (2001a, 2001b).<br />

In 1992, our group developed a computerized method, based on the analysis of the 13 C NMR<br />

spectrum, that allows the direct identification of the (major) components of a mixture (Corticchiato<br />

and Casanova 1992; Tomi et al. 1995).<br />

In that procedure, there is neither separation nor individualization of the components before<br />

their identification. The computer program, made up in our laboratories from Microsoft Access,<br />

compared the chemical shift of each carbon in the experimental spectrum with the spectra of pure<br />

compounds listed in our spectral libraries (Figure 8.1). Each compound is then identified, taking<br />

into account three parameters that are directly available from the computer program:<br />

1. The number of observed carbons with respect to the number of expected signals<br />

2. The number of overlapped signals of carbons that possess the same chemical shift<br />

3. The difference of the chemical shift of each signal in the mixture spectrum and in the<br />

reference spectra<br />

Two complementary spectral data libraries were created, the first one containing spectra of mono-, sesqui-,<br />

and diterpenes, and phenylpropanoids recorded in our laboratories under the same experimental<br />

conditions (solvent, concentration, pulse sequence). The reference compounds were commercially available<br />

or isolated from essential oils and extracts. The second library was constructed with literature data.<br />

Quantitative determination of some components may be carried out by NMR when necessary<br />

(nonvolatile compounds, thermolabile compounds) (Rezzi et al. 2002; Baldovini et al. 2001).<br />

Analysis of complex EOs may be achieved by a combination of CC and 13 C NMR (Gonny et al.<br />

2004; Blanc et al. 2006).<br />

Our results have been reviewed (Bradesi et al. 1996; Tomi and Casanova 2000) with a special<br />

insight into the application of NMR to the analysis of EOs from Labiatae (Tomi and Casanova<br />

2006).<br />

8.2.3.7 enantiomeric differentiation<br />

EOs contain mainly terpenes that are present in pure enantiomeric form or as a nonracemic mixture<br />

of both enantiomers. This point is quite important since olfactory characteristics of a terpene<br />

are generally different, depending on the considered enantiomer. For instance, (R)-(+)-limonene


202 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

smells like orange, while the (S)-(−)-enantiomer smells like turpentine. (−)-Menthol smells and<br />

tastes sweetish-minty and is fresh and strongly cooling, in contrast to the (+)-enantiomer, which<br />

smells and tastes herby-minty and is weakly cooling. (S)-(+)-Carvone possess the typical odor of<br />

caraway, and its (R)-(−)-enantiomer smells more like peppermint (Bretmaier 2006).<br />

Moreover, in an aromatic plant, the two enantiomers of a compound usually have different origin.<br />

For instance, (+)-limonene is the major component of orange zest oil, whereas (−)-limonene<br />

is characteristic of eucalyptus, mint, or pine oils (Mosandl 1988). Coriander oil contains mainly<br />

(+)-linalool, whereas (−)-linalool is present in basil oil (Boelens et al. 1993).<br />

Therefore, measurement of enantiomeric excess of selected terpenes in EOs is an important<br />

parameter for the characterization of a plant as well as for quality control. For example, the presence<br />

of pure (+)-fenchone in fennel EO indicates the natural origin of this oil. Conversely, (−)-fenchone<br />

is present only in Artemisia and cedar oils (Ravid et al. 1992).<br />

The enantiomeric differentiation of monoterpenes and monoterpenoids, and the quantitative<br />

determination of the enantiomers, is typically carried out using a chiral GC column. Permethylated<br />

cyclodextrins are the most commonly used chiral phases as they are efficient for many terpenoids<br />

present in various EOs (Bicchi et al. 1997, 1999).<br />

Using a single chiral column is not recommended since the enantiomeric differentiation of most<br />

components of the oil induces an increase of the number of peaks in the chromatogram and, consequently,<br />

the number of coelution. So, it is not easy to attribute the right peak to one enantiomer of a<br />

compound. In such a case, GC needs to be coupled with MS.<br />

Two-dimensional GC is preferred. The components of the oil are first separated on a nonchiral<br />

column contained in the first oven of the chromatograph and, then, only some compounds (selected<br />

with respect to their retention time and order of elution) are injected into the second chiral column<br />

where they are enantiomerically differentiated.<br />

Enantiomeric differentiation of oxygenated terpenes could be carried out using ( 1 H and 13 C)<br />

NMR spectroscopy and a chiral lanthanide shift reagent (CLSR). Although that technique was<br />

hardly used, some examples have been reported. The enantiomeric composition of linalool isolated<br />

from the oil of coriander was established by 1 H NMR, and the result was in agreement with that<br />

obtained by conventional polarimetry (Ravid et al. 1985). 13 C NMR and CLSR were successfully<br />

employed by Fraser et al. (1973) for the determination of the enantiomers of alcohols and acetates.<br />

In our laboratories, enantiomeric differentiation was performed for oxygenated monoterpenes in the<br />

pure form and in EOs: camphor and fenchone in Lavandula stoechas EO (Ristorcelli et al. 1998)<br />

and bornyl acetate in the EO of Inula graveolens (Baldovini et al. 2003).<br />

8.2.3.8 two-step Procedure<br />

In this procedure, a small quantity of a substance is separated by fractionated distillation or a chromatographic<br />

technique—liquid chromatography (LC), HPLC, preparative GC—and then identified<br />

by comparison of its spectral data—MS, FTIR, UV, 1 H NMR, and sometimes 13 C NMR—with<br />

those of reference compounds.<br />

Weyerstahl et al. (2000) have used this procedure for several years to fully characterize very<br />

complex EOs. For instance, they partitioned Haitian vetiver oil by fractionated distillation, column<br />

chromatography, and transformation of alcohols into their methyl esters. In total, they identified by<br />

GC-MS and NMR more than 150 components. Among them, several tens are new sesquiterpenes;<br />

they were fully characterized by 1 H and 13 C NMR. In the same way, these authors analyzed an<br />

EO of Cistus ladaniferus gum (Weyerstahl et al. 1998). They fractionated 400 g of a commercial<br />

sample by various methods (treatment by alkaline solution, fractionated distillation, CC, and TLC).<br />

Compounds present as mixtures were identified by GC-MS, whereas pure components were fully<br />

characterized by usual spectroscopic techniques: MS and 1 H, and 13 C NMR. The authors identified,<br />

in total, 186 components; among them, 26 are new compounds.<br />

Bicchi et al. (1998) used the same techniques to fractionate an essential oil from Artemisia roxburghiana.<br />

They identified 108 components, mainly by GC(RI) and GC-MS. However, 23 of these


Analytical Methods for Cymbopogon <strong>Oil</strong>s 203<br />

components were sesquiterpene isomers, and their identification required the use of 1 H NMR and,<br />

sometimes, 13 C NMR.<br />

This two-step procedure is obviously more accurate, but it is very time consuming and requires<br />

a large quantity of EO.<br />

8.3 analysIs oF Cymbopogon oIls<br />

The composition of Cymbopogon oils was investigated a long time ago, in connection with<br />

the properties of the oils. Most studies concerned oils of commercial importance: citronella,<br />

lemongrass, and palmarosa oils. Citronella oils are usually divided into two types: Ceylon type<br />

and Java type obtained from C. nardus (L.) Rendle and C. winterianus Jowitt. Lemongrass oil<br />

is found either in Cymbopogon citratus Stapf (sometimes known as West Indian lemongrass) or<br />

in C. flexuosus Stapf. Palmarosa oil is produced from C. martinii (Roxb.) Wats. var martinii.<br />

The compositions of these oils have been reviewed by Lawrence (1979, 1989, 1993, 1995, 2003,<br />

2006). However, the compositions of various oils obtained from other species of Cymbopogon<br />

are reported in the literature, among others: C. afronardus, C. distans, C. giganteus, C. jawarancusa,<br />

C. microstachys, C. olivieri, C. parkeri, C. travancorensis, and C. validus.<br />

Cymbopogon oils contain mainly oxygenated monoterpenes: geranial, neral (citral), citronellol,<br />

geraniol, etc. (Figure 8.2). However, a great variety of compounds was reported as minor components:<br />

monoterpene hydrocarbons, sesquiterpene hydrocarbons and oxygenated sesquiterpenes,<br />

phenylpropanoids, and nonterpenic acyclic compounds.<br />

Depending on the various techniques (chemical, chromatographic, spectroscopic) used to<br />

identify and quantify the individual components of Cymbopogon EOs, we first remember the analysis<br />

by chemical methods. Then, we shall develop the analysis by chromatographic techniques,<br />

analysis by mass spectrometry coupled with gas chromatography, and analysis by hyphenated<br />

techniques. We shall summarize the enantiomeric differentiation of terpenes by chiral GC and<br />

by NMR. Finally, the combined analysis of EOs by various techniques will be exemplified on<br />

C. giganteus oil.<br />

O<br />

1 2<br />

3<br />

OH<br />

O O<br />

OH<br />

4 5 6<br />

FIgure 8.2 1 = geranial, 2 = citronellal, 3 = neral, 4 = geraniol, 5 = citronellol, 6 = myrcene.


204 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

8.3.1 ea r l i e s t st u d i e s: an a ly s i s by Ch e M iC a l Me t h o d s (is o l at i o n o f Co M P o u n d s<br />

a n d Mi sC e l l a n e o u s Me t h o d s)<br />

According to Lawrence (1989c), the first studies date back to 1948 when Dasgupta and Narain<br />

used fractional distillation and chemical derivatization in combination with physicochemical data to<br />

identify geraniol and a few derivatives in a sample of palmarosa oil from India. Later, the content of<br />

citral in lemongrass oil was determined by condensation with barbituric acid (Lawrence 1995b).<br />

Several sesquiterpene hydrocarbons were isolated from citronella oil from Java: β-bourbonene,<br />

δ-cadinene, α-cubebene, β-elemene, and α-selinene (Lawrence 1989a).<br />

Various physical and chemical methods were developed to isolate the major constituents of<br />

Cymbopogon oils. For instance, the sodium method was used to isolate the alcoholic constituents<br />

(mostly geraniol and citronellol) of the EO of Java citronella, and it was suggested that this method<br />

was more efficient than removal of alcohols by fractionated distillation (Lawrence 1979a). Various<br />

chemical and chromatographic methods were compared to evaluate the total citral content in lemongrass<br />

oil: bisulfite method, oxime method, barbituric acid method, GC with internal standard,<br />

and the GC method after NaBH 4 reduction. The authors recommended that the GC method using<br />

phenylethyl alcohol as internal standard was the most accurate method for determination of citral<br />

(Lawrence 1995b).<br />

Only a few new compounds have been isolated from Cymbopogon oils and their structure<br />

elucidated.<br />

C. flexuosus oil contained iso-intermedeol, a sesquiterpene alcohol (Thappa et al. 1979; Huffman<br />

and Pinder 1980), and later, both the oil and the alcohol were investigated for their ability to induce<br />

apoptosis in human leukemia HL-60 cells because dysregulation of apoptosis is the hallmark of<br />

cancer cells (Kumar et al. 2008).<br />

Evans et al. (1982) reported that the sesquiterpene diol proximadiol (Figure 8.3), with antispasmodic<br />

properties, earlier isolated from Cymbopogon proximus, is in fact identical with cryptomeridiol.<br />

An eudesmandiol, [2R-(2α,4aβ,8α,8aα)]-decahydro-8a-hydroxy-α,α4a,8-tetramethyl-2-naphthalene<br />

methanol, was isolated from the EO of Cymbopogon distans. It was studied spectroscopically,<br />

and the absolute configuration was determined by means of x-ray diffraction (Mathela et al. 1989).<br />

HO<br />

OH<br />

H<br />

H OH<br />

OH<br />

OH OOH OH H OH<br />

7 8<br />

9<br />

H<br />

H<br />

OH<br />

10<br />

OH<br />

11 12<br />

FIgure 8.3 7 = proximadiol, 8 = 5α−hydroperoxy-β-eudesmol, 9 = 3-eudesmen-1β,11-diol, 10 = selin-<br />

4(15)-en-1β,11-diol, 11 = 7α,11-dihydroxycadin-10(14)ene, 12 = α-oxobisabolene.<br />

O


Analytical Methods for Cymbopogon <strong>Oil</strong>s 205<br />

Cymbodiacetal, a novel dihemiacetal bis-monoterpenoid, isolated from the EO of C. martinii,<br />

was identified by means of x-ray diffraction (Bottini et al. 1987).<br />

Recently, various sequiterpene alcohols, including two new compounds, 5α-hydroperoxy-βeudesmol<br />

and 7α,11-dihydroxycadin-10(14)-ene (Figure 8.3), have been isolated from a pentane<br />

extract of C. proximus (El-Askari et al. 2003).<br />

Finally, it should be noted that the presence of α-oxobisabolene (bisabolone) (12%) (Figure 8.3)<br />

was reported in a sample of Ethiopian C. citratus oil (Abegaz et al. 1983). Melkani et al. (1985) also<br />

reported the occurrence of that compound in concentrations ranging from 18% to 68% in one of the<br />

oils of two varieties of C. distans. Isolation of (+)-1-bisabolone from C. flexuosus and its utilization<br />

as antibacterial agent has been patented.<br />

However, the chemical techniques were supplanted by the introduction, first of chromatographic<br />

techniques (TLC and, obviously, GC) and later of spectroscopic techniques (MS, IR, NMR).<br />

8.3.2 an a ly s i s by Ch r o M at o G r a P h iC te C h n i q u e s [tlC, hPlC, GC(rt) o r GC(ri)]<br />

Due to the complexity of the EOs (number of constituents, diversity of the structures and functionalities,<br />

and also similarities resulting from the biosynthesis of terpenes from the same building block,<br />

the isoprene unit), TLC has not been considered a practical tool for analysis of Cymbopogon oils.<br />

Nevertheless, TLC has been used to characterize the quality of lemongrass oil from Bangladesh<br />

(Lawrence 1995). The authors concluded that the oil could be competitive on the market of citralrich<br />

oils.<br />

Following an improvement of the column for gas chromatography that allowed a good individualization<br />

of the components of essential oil, this technique became more and more useful for<br />

analysis of Cymbopogon EOs.<br />

The first experiments concerned the (tentative) identification of individual components, by comparison<br />

of their retention times with those of reference compounds:<br />

• As early as 1962, retention times were used to identify ten monoterpene hydrocarbon components<br />

in Ceylon citronella oil (Lawrence 1989a).<br />

• A few years later, eugenol and methyl eugenol were identified in the same way, besides<br />

monoterpenes, in a sample of Java citronella oil dominated by citronellal and geraniol<br />

(Lawrence 1989a);<br />

• In the meantime, Peyron (1973) was able to distinguish, among other monoterpenes, (Z)<br />

and (E) isomers of ocimene in palmarosa oil from Brazil, largely dominated by geraniol;<br />

• The same year, Wijesekera et al. (1973) reported the utilization of GLC profiles to distinguish<br />

both varieties of citronella oil, although both types contained comparable amounts of<br />

geraniol: Ceylon variety (Lenabatu) contained large amounts of monoterpene hydrocarbons,<br />

while the Java variety (Mahapengiri) contained only small amounts. In addition, the Ceylon<br />

type contained tricyclene, methyl eugenol, methyl isoeugenol, eugenol, and borneol;<br />

• Mohammad et al. (1981) determined the composition of two oil samples of lemongrass<br />

(C. flexuosus) and detected trace constituents using the peak enrichment technique (addition<br />

of a known compound in the EO and comparison of the new chromatogram with that<br />

of the pure oil). Besides neral and geranial, by far the main components, the authors identified<br />

various monoterpenes as well as decanal. The two samples also contained appreciable<br />

contents of farnesol and farnesal. However, the correct isomer of the two acyclic sesquiterpenes<br />

was not identified.<br />

More than 20 years later, Rauber et al. (2005) investigated a completely different approach to<br />

the quantitative determination of citral in C. citratus volatile oil. An HPLC method was developed<br />

(Spherisorb ® CN column, n-hexane: ethanol mobile phase, and UV detector) and validated


206 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

(precision, accuracy, linearity, specificity, quantification, and detection limits). The concentration<br />

of citral in C. citratus volatile oil obtained with this assay was 75%.<br />

8.3.3 an a ly s i s by Ma s s sPeC troM e try Co u P l e d w i t h Ga s Ch r o M at o G r a P h y<br />

In parallel, during the 1970s, owing to the improved sensitivity of mass spectrometers, identification<br />

of components became possible by comparison of their mass spectra with those of reference<br />

compounds. A gas chromatograph coupled with a mass spectrometer became the basic instrument<br />

for EO analysis. Using this technique, any volatile known compound, even present at the trace level,<br />

could be theoretically identified by comparison of its mass spectral data with those of reference<br />

compounds compiled in either paper or “computerized” libraries.<br />

A GC-MS study of the hydrocarbon fraction and the fraction containing oxygenated compounds<br />

showed the presence of 12 monoterpene hydrocarbons (28.4%), 13 sesquiterpene hydrocarbons<br />

(32.8%), 3 sesquiterpene alcohols (27.2%), 2 esters (7.2%), and 3 carbonyl compounds<br />

(4.4%) in the EO of C. distans. Among these, 27 compounds have been identified (Mathela and<br />

Joshi 1981).<br />

Analysis of an oil sample of C. flexuosus of Indian origin, by GC-MS, allowed the identification<br />

of monoterpenes (including perillene), various sesquiterpene hydrocarbons, as well as methyl<br />

eugenol and elemicin (Taskinen et al. 1983).<br />

In practice, the main drawback of that technique, particularly for the analysis of EOs, consists<br />

in the occurrence of insufficiently differentiated spectra for compounds built up with the same<br />

isoprene unit. Therefore, it was determined that sesquiterpenes bearing quite different skeletons<br />

exhibit similar (if not identical) mass spectra. Moreover, superimposable mass spectra are commonly<br />

observed for isomers (stereoisomers, diastereoisomers). Consequently, it is not surprising that<br />

the correct isomer of various monoterpenes and overall sesquiterpenes could not be identified by<br />

MS. In his excellent reviews, Lawrence pointed out the lack of specification of the correct isomer.<br />

For instance:<br />

• The correct isomer of various monoterpenes (menthone, isopulegols, and rose oxide) was<br />

not specified in the composition of citronella oils from Java and Sri Lanka investigated by<br />

GC-MS (Lawrence 1989a);<br />

• The isomer of verbenol contained in lemongrass ( C. citratus) grown in Zambia was not<br />

reported (Chisowa et al. 1998). Conversely, both isomers were found, as minor components,<br />

in oil samples from the Ivory Coast (Chalchat et al. 1998).<br />

• The correct isomer of elemene remained frequently unassigned in old papers and more<br />

surprisingly, in recent papers, as well as the stereoisomer of farnesol (Lawrence 2006d).<br />

Moreover, misidentification frequently occurred when the analysis was carried out only by MS.<br />

Concerning Cymbopogon oils, Lawrence, in his reviews, pointed out misidentified common compounds<br />

simply by the observation of the order of elution on apolar (or polar) column. Misidentifications<br />

concerned, in general, minor compounds but, sometimes, compounds present at appreciable<br />

contents.<br />

Anyway, year after year, MS became essential in EO analysis. Reliable results are obtained by combination<br />

of GC and MS. The combined use of both techniques will be developed in the next subsection.<br />

8.3.4 Co M b i n e d an a ly s i s by Ch r o M at o G r a P h iC a n d sP e C t r o s C o P iC<br />

te C h n i q u e s: tlC, CC, GC(rt) o r GC(ri) a n d GC-Ms<br />

In the previous text, we have seen the importance of both GC and MS for identification of individual<br />

components of EOs. Combination of the two techniques appeared really useful for analysis


Analytical Methods for Cymbopogon <strong>Oil</strong>s 207<br />

of volatiles. Identification of individual components resulted from comparison of mass spectral data<br />

in the case of MS and retention times or retention indices in the case of GC with those of reference<br />

compounds. More accurate and reproducible results are obtained by using retention indices<br />

(Van Den Dool and Kratz 1963; Kováts 1965) instead of retention times (see Part I).<br />

Reliable results are obtained when individual components of EO are identified by MS in combination<br />

with RIs measured on two columns of different polarity (apolar and polar). Nowadays, it is<br />

considered by specialized reviews that analysis of an EO by at least two techniques should be done,<br />

to keep the work publishable. To facilitate identification of components, prefractionation of the oil<br />

by liquid chromatography (LC) could be useful.<br />

8.3.4.1 Cymbopogon oils of commercial Interest<br />

The first combined analyses, by GC and GC-MS, of Cymbopogon oils of commercial interest (citronella,<br />

lemongrass, and palmarosa oils), were published in the 1970s and have been reviewed by<br />

Lawrence (1979, 1989, 1993, 1995, 2003, 2006). Although the oil composition was always dominated<br />

by oxygenated monoterpenes (geranial, neral, and geraniol), the number and the structure of<br />

identified compounds varied substantially from paper to paper.<br />

The composition of lemongrass oil (C. flexuosus) from Guatemala was reported in 1976. Among<br />

unusual monoterpenes, the nonterpenic aldehyde decanal, as well as germacrene D and γ-cadinene were<br />

identified. The correct isomer of ocimene and allo-ocimene was not specified (Lawrence 1979b).<br />

A few years later, 2-undecanone was detected, by GC and GC-MS, in an Indian lemongrass oil<br />

sample containing more than 72% of citral (Lawrence 1989b).<br />

In 1983, two samples of Turkish lemongrass oil (C. citratus) were investigated by GC and GC-MS,<br />

after prefractionation of the oil by liquid chromatography. Both oils were dominated by citral, and<br />

they contained appreciable amounts of myrcene (up to 19%) as well as nonanal and 2-undecanone<br />

(Lawrence 1989b).<br />

Retention times and GC-MS data were taken into account to investigate an oil sample of<br />

C. winterianus from Bangladesh. Usual monoterpene hydrocarbons and monoterpenols were identified,<br />

as well as elemol and α-eudesmol. The authors also found 2,3-dimethyl-5-heptenal probably<br />

misidentified, according to Lawrence (1995a), instead of methylheptenone.<br />

Twelve samples of palmarosa oil produced in Madagascar were investigated by GC(KI) and<br />

GC-MS. Besides geraniol (major component) and usual monoterpenes, numerous sesquiterpenes<br />

were identified, although the coelution of various sesquiterpene hydrocarbons was observed.<br />

Three isomers of farnesol, (Z,E), (E,Z), and (E,E), were found and differentiated (Figure 8.4)<br />

(Randriamiharisoa and Gaydou 1987). The same authors reported the detailed analysis of the hydrocarbon<br />

fraction of that oil (4.75%), separated by CC (SiO 2) and analyzed by GC-MS in combination<br />

13 (E,E)<br />

13 (Z,E)<br />

13 (E,Z)<br />

FIgure 8.4 Structure of farnesol stereoisomers 13.<br />

OH<br />

OH<br />

OH


208 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

with KI. Twenty-eight sesquiterpene hydrocarbons were identified. In addition, long-chain alkanes<br />

were characterized. We note also the occurrence of the three isomers of cymene (Gaydou and<br />

Randriamiharisoa 1987).<br />

Ntezurubanza et al. (1992) identified the individual components of lemongrass oil from Rwanda,<br />

using MS in combination with RIs measured on two columns of different polarity (apolar and<br />

polar). They found 2-tridecanone and lavandulyl acetate, besides the usually seen monoterpenes.<br />

The chemical composition of commercial citronella oils of various geographical origins has been<br />

examined, using MS data and RIs (ethyl esters). Fifty-five compounds were identified, including<br />

various sesquiterpenes and phenyl propanoids (eugenol, methyl eugenol, E and Z isomers of methyl<br />

isoeugenol) (Lawrence 2003a).<br />

A Brazilian citronella oil sample, examined using GC(KI) and GC-MS, was found to contain<br />

citronellal and neral as major components while the content of geranial was very low. The phenyl<br />

propanoid elemicin accounted for 7% of the oil, which also contained β-elemene (0.5%) and elemol<br />

(3.7%) (Lawrence 1995a).<br />

Despite the improvement of analytical techniques, the analysis of complex mixtures, such as<br />

EOs, remains a difficult task and needs careful work. Unfortunately, despite using two analytical<br />

techniques (GC and GC-MS), misidentification of individual components still occurs as illustrated<br />

by the analysis of C. nardus oil from Zimbabwe, where numerous sesquiterpenes appeared misidentified<br />

on the basis of their relative elution order (Lawrence 2003a).<br />

The EO from seeds of Indian Cymbopogon martinii (Roxb.) Wats. var. motia Burk. collected from<br />

three different geographical locations was analyzed by capillary GC (KI measured on apolar and<br />

polar columns and peak enrichment by co-injection of authentic samples) and GC-MS (Mallavarapu<br />

et al., 1998). The composition of the oil samples was compared with that of the oil of flowering<br />

palmarosa herb. Besides the main constituent, geraniol (74.5%–81.8%), 55 other constituents were<br />

identified in the seed EOs. Although the composition of the seed oils is similar to that of the herb<br />

oil, quantitative differences in the concentration of some constituents were observed. The seed oil<br />

was found to contain lower amounts of geranyl acetate and higher amounts of (E,Z)-farnesol than<br />

the herb oil (two isomers of farnesol were differentiated in the oils).<br />

The composition of the leaf oil of C. citratus (DC) Stapf, growing on the campus of Lagos State<br />

University (Nigeria), was determined by the use of GC (KI on polar column) and GC-MS (Kasali<br />

et al. 2001). Twenty-three (97.3%) constituents were identified. The main components were geranial<br />

(33.7%), neral (26.5%), and myrcene (25.3%). An unusual saturated monoterpene hydrocarbon<br />

(0.1%) was also detected and probably misidentified as 2,6-dimethyloctane.<br />

The steam-distilled volatile oil obtained from partially dried grass (citronella grass) C. nardus<br />

(Linn.) Rendle, cultivated in the Nilgiri Hills, India, was analyzed by capillary GC and GC-MS<br />

(Mahalwal and Ali 2003). The prominent monoterpenes were citronellal (29.7%), geraniol<br />

(24.2%), γ-terpineol (9.2%), and cis-sabinene hydrate (3.8%). The predominant sesquiterpenes<br />

were (E)-nerolidol (4.8%), β-caryophyllene (2.2%), and germacren-4-ol (1.5%). An irregular<br />

monoter pene compound eluted among sesquiterpenes was identified as 3,3,6-trimethylhepta-1,5diene,<br />

and two 1,2-benzenze dicarboxylic acid derivatives were tentatively identified (although they<br />

are probably not naturally occurring substances).<br />

The EOs of palmarosa (C. martinii (Roxb.) Wats. var. motia Burk) flowering herbs from three<br />

different geographical locations in India (Hyderabad, Lucknow, and Amarawati) were analyzed by<br />

high-resolution GC (apolar column) and GC–MS (Raina et al. 2003). In all three samples, geraniol<br />

(67.6%–83.6%) was the major constituent. However, quantitative differences in the concentration of<br />

some constituents were observed. Among the three oils analyzed, Amarawati oil was of the highest<br />

quality due to higher geraniol (83.6%) and lower geranyl acetate (2.3%) and geranial (1.0%) content.<br />

Isopropyl propionate and butyrate (as well as cyclohexanone!) were detected in that oil.<br />

The supercritical fluid extraction (SFE) of C. citratus was performed in sequential and dynamic<br />

extraction modes (Sargenti and Lanças 1997). Principal compounds in the SFE extract were analyzed<br />

by GC and GC-MS and identified as neral, geraniol, and geranial. Nerolic acid and geranic


Analytical Methods for Cymbopogon <strong>Oil</strong>s 209<br />

acid were identified as minor components in the essential oil. Different chromatographic profiles<br />

were obtained depending of the experimental conditions.<br />

The same technique was applied by Marongiu et al. (2006) who analyzed the influence of pressure<br />

on the supercritical extraction, in order to maximize citral content in the extract oil. Dried and<br />

ground leaves of lemongrass (C. citratus Stapf) were used as a matrix. The collected extracts were<br />

analyzed by GC-MS, and their composition was compared with that of the EO isolated by hydrodistillation<br />

and by steam distillation. At optimum conditions (90 bar and 50°C) citral represented<br />

more than 68% of the extract. At higher solvent density, the extract aspect changes because of the<br />

extraction of high-molecular-mass compounds.<br />

8.3.5 ot h e r Cy m b o p o g o n oils<br />

The composition of EO of other species has been also investigated:<br />

• The composition of the essential oil of C. jawarancusa (Khavi grass), was investigated by<br />

GC in combination with GC-MS (Saeed et al. 1978), and 64 compounds were identified.<br />

The oil exhibited a high content of piperitone (Figure 8.5; 60%–70%), which is mainly<br />

responsible for the smell of Khavi grass. The chemical variability of that oil was evidenced,<br />

and races rich in piperitone, phellandrene, and other chemical constituents have<br />

been identified (Dhar et al. 1981).<br />

• Analysis of the volatile oil of C. parkeri has been carried out by GC-MS (Rizk et al. 1983).<br />

To facilitate identification, prefractionation of the oil by adsorption high-performance liquid<br />

chromatography was performed. The major compounds are geraniol (33.5%), nerol<br />

(22.2%), geranyl acetate (8.9%), neryl acetate (3.8%), and farnesol (3.7%). A sesquiterpene<br />

alcohol that accounted for 4.8% remained unidentified. A quite different composition,<br />

determined by the use of GC and GC-MS, was reported for an oil sample obtained<br />

from aerial parts of C. parkeri Stapf, collected at flowering stage from Kerman province<br />

of Iran. The main constituents were piperitone (80.8%) and germacrene D (5.1%) (Bagheri<br />

et al. 2007).<br />

• The chemical composition of the essential oil of C. travancorensis Bor (Poaceae) was<br />

investigated by capillary GC and GC-MS. Thirty-five compounds were identified. The<br />

oil contains terpenes and phenyl propanoids (22.04%). The main constituents of the oil<br />

are limonene (18%), elemicin (17%), camphene (12%), elemol (11%), and borneol (10%)<br />

(Mallavarapu et al. 1992).<br />

• The steam-distilled oils from wild and cultivated C. validus (Stapf) Stapf ex Burtt Davy<br />

(Gramineae) were analyzed by GC and GC-MS (Chagonda et al. 2000). The major components<br />

from wild C. validus were the following: myrcene (23.1%–35.6%), (E)-β-ocimene<br />

(10.3%–11.5%), geraniol (3.4%–8.3%), linalool (3.2%–3.7%), and camphene (5.2%–6.0%).<br />

O<br />

14<br />

OH<br />

15<br />

OCH 3<br />

OCH 3<br />

OCH 3<br />

H 3 CO<br />

16 17<br />

FIgure 8.5 14 = piperitone, 15 = eugenol, 16 = (E)-methyl isoeugenol, 17 = elemicin.<br />

OCH 3<br />

OCH 3


210 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

•<br />

Cultivated mature plants contained myrcene (11.6%–20.2%), (E)-β-ocimene (6.0%–12.2%),<br />

borneol (3.9%–9.5%), geraniol (1.7%–5.0%), and camphene (3.3%–8.3%) as the major<br />

components. Young nursery crop/seedlings (20–30 cm high) contained oil with myrcene<br />

(20.6%), geraniol (17.1%), and germacrene-D-4-ol (8.3%) as main components. Geranyl<br />

acetate (4.5%), linalol (4.5%), and borneol (2.9%) were notable minor components.<br />

The constituents of the essential oil obtained by hydrodistillation of the aerial parts of<br />

C. olivieri (Boiss.) Bor, growing wild in Iran, was investigated by GC (RI on apolar column)<br />

and GC-MS (Norouzi-Arasi et al. 2002). Forty-two components, representing 97.6%<br />

of the oil, were identified, of which piperitone (53.3%), α-terpinene (13.6%), elemol (7.7%),<br />

β-eudesmol (4.4%), torreyol (3.3%), limonene (2.9%), and α-cadinol (2.1%) were the major<br />

components.<br />

C. microstachys (Hook.f.) Soenarko is reported to be found wild in some pockets of the Himalayan<br />

foothills of Assam, Uttar Pradesh, and West Bengal. Mathela et al. (1990) have studied the composition<br />

of its oil and identified 31 compounds. The oil was found to contain about 60% of phenyl propanoids<br />

with methyl eugenol (19.5%), methyl isoeugenol (4.2%), elemicin (25.3%), and isoelemicin<br />

(11.0%) (Figure 8.5). Another sample of C. microstachys had quite a different composition, with<br />

(E)-methyl isoeugenol being the main constituent (56.4%–60.7%) and myrcene (7.8%–12.2%) (Rout<br />

et al. 2005).<br />

C. afronardus oils, analyzed by GC and GC-MS, contained myrcene and intermedeol as major<br />

constituents. The oil is used in Uganda in the formulation of locally manufactured toothpaste (Baser<br />

et al. 2005).<br />

8.3.6 enantioMeriC differ entiation by Ch i r a l GC<br />

Ravid et al. (1992) carried out the enantioselective analysis of citronellol (10.2% in an oil sample<br />

of citronella) through a trifluoroacetyl derivative of the isolated compound; the (R)-(+)/(S)-(−) ratio<br />

was 74/26.<br />

Mosandl et al. (1990) used a combination of heart-cutting 2D GC and chiral GC to determine<br />

the enantiomeric distribution of monoterpene hydrocarbons in citronella oil: α-pinene [(+)/(−) =<br />

77/23] and limonene [(+)/(−) = 4/96] (Figure 8.6), and in lemongrass oil: α-pinene [(+)/(−) = 4/96],<br />

limonene [(+)/(−) = 0/100], and β-pinene [(+)/(−) = 0/100]. The enantiomeric composition of oxygenated<br />

monoterpenes was investigated: citronellol, linalool (Figure 8.6), terpinen-4-ol, cis and trans<br />

rose-oxides (Kreis and Mosandl 1994), linalool (Wang et al. 1995) lemongrass, and borneol (Ravid<br />

et al. 1996).<br />

The EO from aerial parts of C. winterianus Jowitt, cultivated in Southern Brazil, was analysed by<br />

GC-MS. Enantiomeric ratios of limonene [(R)-(+)/(S)-(−) = 13/87], linalool [(R)-(−)/(S)-(+) = 39/61]),<br />

citronellal [(R)-(+)/(S)-(−) = 91/09]), and β-citronellol [(R)-(+)/(S)-(−) = 85/15] were obtained by<br />

HO HO<br />

18(R) 18(S) 19(S) 19(R)<br />

FIgure 8.6 18(R) = (R)-(+)-limonene, 18(S) = (S)-(−)-limonene, 19(R) = (R)-(−)-linalool, 19(S) = (S)-(+)linalool.


Analytical Methods for Cymbopogon <strong>Oil</strong>s 211<br />

multidimensional gas chromatography, using a developmental model setup with two GC ovens. The<br />

enantiomeric distributions are discussed as indicators of origin authenticity and quality of this oil<br />

(Lorenzo et al. 2000).<br />

8.3.7 Co M b i n e d an a ly s i s o f Cy m b o p o g o n oils by Va r i o u s te C h n i q u e s:<br />

tlC, CC, ir, GC(ri), GC-Ms, a n d 13 C nMr<br />

In various examples, the analysis was carried out by combination of chromatographic (TLC and<br />

GC) and spectroscopic (IR) techniques. This type of work is illustrated by the analysis of lemongrass<br />

oil (C. citratus) from Egypt. About 28 compounds were clearly detected by GC, from which<br />

17 were identified, citral being mentioned as a major component. The components belonged mostly<br />

to the acyclic-oxygenated monoterpenes. δ-Cadinene and (E)-β-caryophyllene were also identified<br />

(Abdallah et al. 1975). The oil was also analyzed by TLC, with a gradient of solvents as eluent.<br />

Citral, citronellal, and various alcohols were detected (Zaki et al. 1975). Finally, the infrared spectrum<br />

of the lemongrass oil proved that the changes in the samples during storage could be detected,<br />

especially the changes of the carbonyl groups (Foda et al. 1975).<br />

A combined analysis of citronella oil (C. winterianus) by fractional distillation, TLC, GC, IR,<br />

and NMR was described: besides citronellal and geraniol, the main components, four acyclic oxygenated<br />

monoterpenes, a cyclic one, a sesquiterpenol (elemol), and a phenylpropanoid (eugenol)<br />

were identified (Lawrence 1989a).<br />

An elegant and unusual technique was used to analyze the nitrogenous fraction of palmarosa oil.<br />

The compounds were separated by preparative GC and characterized using a combination of GC<br />

(nitrogen and sulfur detectors), GC-MS, IR, and NMR. Eighteen pyrazines, as many pyridines, four<br />

thiazoles, and a few other nitrogen-containing compounds were identified (Lawrence 1993).<br />

Very recently, the composition of the sample of palmarosa oil that has proven antimicrobial<br />

properties against cells of Saccharomyces cerevisiae was determined as 65% geraniol and 20%<br />

geranyl acetate, as confirmed by GC–FTIR (Prashar et al. 2003).<br />

In the early 1980s, NMR began to be employed to analyze EO. In 1981, Chiang et al. used proton<br />

NMR to quantify citronellal in Taiwanese citronella oil (average: 38.4%) (Lawrence 1989a). One<br />

year later, Formácek and Kubeczka (1982a) used a combination of GC and 13 C NMR to identify<br />

various monoterpenes, as well as 2-nonanone and (E)-β-caryophyllene in East Indian lemongrass.<br />

More recently, the same authors again used a combination of techniques, capillary GC, and 13 C<br />

NMR to compare the composition of citronella oil (C. nardus) and Java-type oil (C. winterianus)<br />

(Kubeczka and Formácek 2002). Both oils differed overall by the content of monoterpene hydrocarbons<br />

(camphene, limonene), oxygenated monoterpenes (citronellal, citronellol, borneol), and<br />

(E)-methyl isoeugenol. Forty-eight compounds were identified, including various sesquiterpenols<br />

and phenylpropanoids.<br />

Seven oil samples of C. schoenanthus were recently analyzed by GC(RI), GC-MS, and 13 C<br />

NMR, without fractionation (Khadri et al. 2008). The composition of samples was dominated by<br />

monoterpene hydrocarbons, limonene, β−phellandrene, and δ-terpinene. The last samples contained<br />

higher amounts of sesquiterpenes, including β-eudesmol, valencene, and δ-cadinene. All<br />

the components were identified by GC-MS and GC(RI). The identification of 17 components was<br />

confirmed by 13 C NMR.<br />

In our laboratories, we developed a computerized procedure allowing the identification of individual<br />

components of EOs, as illustrated in the first part of this chapter. We applied this method to<br />

the analysis of two samples of C. winterianus and a sample of C. tortilis, all cultivated in Vietnam<br />

(Ottavioli et al. 2008a). In all, 47 components were identified by a combination of GC(RI) on two<br />

columns of different polarity, GC-MS and 13 C NMR, in C. winterianus leaf oil. The composition<br />

of the first sample was largely dominated by geraniol (43.3%), citronellal (15.0%), citronellol<br />

(9.2%), and geranial (8.9%). It is noteworthy that all the 12 compounds, accounting for 91.6% of


212 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

HO<br />

H<br />

O<br />

CHO<br />

HO<br />

20 21<br />

FIgure 8.7 Structure of 20 = trans-7-hydroxy-3,7-dimethyl-3,6-oxyoctanal, 21 = cis-7-hydroxy-3,7dimethyl-3,6-oxyoctanal.<br />

the composition of the oil, were identified by GC-MS and by NMR. The composition of the second<br />

sample was also dominated by geraniol (50.2%) and geranial (8.3%), but citronellal and citronellol<br />

were present at very low levels. In that sample, 46 components were identified, 13 of these<br />

(accounting for 0.6%–50.2% each) by NMR. The same procedure was applied to the analysis of an<br />

oil sample from C. tortilis, without fractionation. The composition of that sample was dominated<br />

by geranial (32.0%) and neral (19.1%). The oil contained epoxyneral and epoxygeranial as well as<br />

caryophyllene oxide at appreciable contents. It differed drastically from the methyl-eugenol-rich oil<br />

from China (Liu et al. 1981).<br />

A leaf oil sample of C. flexuosus was analyzed by GC(RI), GC-MS, and 13 C NMR (Ottavioli<br />

et al. 2008b). The composition of the investigated sample was dominated by geranial (39.2%)<br />

and neral (24.1%), and was somewhat similar to that of various samples reported in the literature<br />

(Surburg and Panten 2006). Most of the identified compounds were classical components found in<br />

EOs, and they were identified by GC-MS by computer matching against commercial mass spectral<br />

data libraries and by comparison of their retention indices with those of reference compounds on<br />

two columns of different polarity. All the compounds, present at appreciable contents, were also<br />

identified by 13 C NMR, following the computerized method described earlier. However, two compounds,<br />

accounting for 4.5% and 4.4%, respectively, remained unassigned. Their retention indices<br />

on apolar and polar columns (compound A: RI = 1252 and 1957; compound B: RI = 1257 and 1940)<br />

led us to suspect two oxygenated monoterpenes. Computer matching of the 13 C NMR data against<br />

a home-made library constructed with literature data suggested the occurrence of cis- and trans-7hydroxy-3,7-dimethyl-3,6-oxyoctanal<br />

(Figure 8.7).<br />

To ensure the occurrence of both isomers, the oil from C. flexuosus was partitioned by flash chromatography<br />

on silica gel. Indeed, in the most polar fraction, eluted with diethyl oxide, both compounds<br />

accounted for 21.5% and 21.3%, respectively. Their NMR data were in agreement with those reported<br />

by Yarovaya et al. (2002). In parallel, analysis of the other fractions of chromatography by 13 C NMR<br />

confirmed the presence of minor monoterpenes, borneol, nerol, as well as sesquiterpenes, trans-αbergamotene,<br />

1,5-diepi-aristolochene, δ-cadinene, cuparene, β-elemol, and (2E,6E)-farnesol.<br />

To conclude this chapter, we illustrate the various techniques or combination of techniques and methods<br />

developed for analyzing EOs on a unique Cymbopogon species. To do that, we chose C. giganteus.<br />

8.3.8 an a ly s i s o f C. gi g a n t e u s oil: a su M M a r y o f an a ly t i C a l Me t h o d s<br />

in V o lV e d in t h e an a ly s i s o f Cy m b o p o g o n oils<br />

C. giganteus (Hochst.) Chiovenda is a perennial and sweet-smelling grass that grows spontaneously in the<br />

savannahs of Asian and African tropical regions. Aerial parts (leaves, flowers, stems) produce by vapor<br />

distillation or water distillation an EO whose composition has been studied by various techniques.<br />

The physiochemical properties of an oil sample (leaf and flower oil) from Côte d’Ivoire have<br />

been measured: density and refraction index as well as viscosity. The authors concluded that the oil<br />

is not Newtonian (Kanko et al. 2004).<br />

The composition of essential oils, obtained from aerial parts of plants growing in different countries,<br />

has been investigated: Benin (Ayedoun et al. 1997; Alitonou et al. 2006), Burkina Faso (Menut<br />

H<br />

O<br />

CHO


Analytical Methods for Cymbopogon <strong>Oil</strong>s 213<br />

OH HO<br />

OH OH<br />

22 23 24<br />

25<br />

FIgure 8.8 Structure of 22 = trans-p-mentha-2,8-dien-1-ol, 23 = cis-p-mentha-2,8-dien-1-ol, 24 = trans-pmentha-1(7),8-dien-2-ol,<br />

25 = cis-p-mentha-1(7),8-dien-2-ol.<br />

et al. 2000), Cameroon (Ouamba 1991; Jirovetz et al. 2007), Côte d’Ivoire (Kanko et al. 2004;<br />

Boti et al. 2006), and Mali (Popielas et al. 1991; Keita 1993; Sidibe et al. (2001).<br />

Most analyses were carried out by a combination of GC(RI) and GC-MS, and the number of<br />

identified components varied substantially from 6 to 55. Whatever the origin, the composition<br />

of C. giganteus oils was dominated by monoterpenes, mostly oxygenated p-menthane derivatives,<br />

such as cis- and trans-p-mentha-2,8-dien-1-ols, cis- and trans-p-mentha-1(7),8-dien-2-ols, cis- and<br />

trans-isopiperitenols, limonene oxides, carvone, and limonene, the only monoterpene hydrocarbon<br />

found at appreciable content (Figure 8.8).<br />

Conversely, the oils differed from sample to sample by the occurrence of different minor components.<br />

Indeed, although several tens of compounds were identified in total in all the studies, only<br />

the six main compounds were identified in all the oil samples: cis- and trans-p-mentha-2,8-dien-1ols,<br />

cis- and trans-p-mentha-1(7),8-dien-2-ols, carvone, and limonene. Conversely, numerous compounds<br />

were identified only in one paper, including monoterpenes, acyclic nonterpenic compounds,<br />

and benzenoid derivatives.<br />

The first study on the chemical composition of C. giganteus essential oil was reported in 1991,<br />

for an oil sample from Bamako, Mali (flowering stage), investigated by GC(RT) and GC-MS (apolar<br />

column) (Popielas et al. 1991). Eighteen compounds were identified, the major ones being p-mentha-<br />

2,8-dien-1-ol (10.7%, isomer nonspecified), cis and trans-p-mentha-1(7),8-dien-2-ols (17.2% and<br />

20.4%, respectively). 1,2-Limonene oxide (13.7%, isomer nonspecified) was also among the major<br />

components. Isopiperitenol and trans-carveol coeluted.<br />

The next year, Keita (1993) published a short analysis of another oil sample of the plant of<br />

the same geographical origin, also carried out by GC(RT) and GC(MS). The menthadienols were<br />

not identified, although they were probably the main components (quoted as unidentified). Isopiperitenone<br />

accounted for 10.2% and carveol (isomer unspecified) for 8.0%.<br />

Finally, C. giganteus oil of Mali was also investigated by Sibidie et al. (2001) by GC (polar column)<br />

and GC-MS. Beside p-menthadienols, by far the main components, two p-menthatrienes and<br />

(Z)-2-phenylbut-2-ene were also identified.<br />

Ayedoun et al. (1997) investigated three oil samples hydrodistilled from plants harvested in<br />

different provinces of Benin. The three samples were characterized by high amounts of limonene<br />

(18%–24%) beside the four p-menthadienols. The essential oil of C. giganteus of Benin was<br />

also analyzed by GC and GC-MS. Once again, the major constituents were as follows: trans-p-<br />

1(7),8-menthadien-2-ol (22.3%), cis-p-1(7),8-menthadien-2-ol (19.9%), trans-p-2,8-menthadien-1-ol<br />

(14.3%), and cis-p-2,8-menthadien-1-ol (10.1%). This study reports the inhibitory effect produced by<br />

the chemical constituents of the essential oil, in vitro on 5-lipoxygenase (Alitonou et al. 2006).<br />

In the course of their studies on aromatic plants of tropical West Africa, Menut et al. (2000)<br />

reported the composition of an oil sample of C. giganteus from Burkina Faso. The analysis was


214 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

carried out by GC (two columns, apolar and polar) and GC-MS. The oil contained mainly limonene<br />

and a set of p-menthadienols. Two unusual esters, 3-methylbutyl hexanoate and octanoate, were<br />

also identified as minor components. The oil’s antioxidant and antiradical activities, which are low,<br />

were studied.<br />

The composition of C. giganteus oil from Cameroon has been first investigated by Ouemba<br />

(1991). A more detailed study was published very recently (Jirovetz et al. 2007). The EOs of fresh<br />

flowers, leaves, and stems of C. giganteus were investigated by GC and GC-MS. More than 55<br />

components have been identified in the three oil samples, the major ones being, as usual, the four<br />

p-menthadienol isomers. Additional components in higher concentrations, responsible for the characteristic<br />

aroma of these samples, are especially limonene, trans-verbenol, and carvone, as well as<br />

some other mono- and sesquiterpenes. Antimicrobial activities of the four oils were found against<br />

Gram-(+)- and Gram-(−)-bacteria, as well as the yeast Candida albicans, and these results were<br />

discussed with the compositions of each sample.<br />

The Cymbopogon giganteus oil from Côte d’Ivoire was only briefly investigated some time ago<br />

(nine components identified; Ouemba 1991). A more recent paper was no more informative (six<br />

compounds identified, method of analysis unspecified; Kanko et al. 2004).<br />

In order to get better insight into the composition of C. giganteus leaf oil from Côte d’Ivoire, a<br />

sample was analyzed using a combination of CC, GC(RI), GC-MS, and 13 C NMR (Boti et al. 2006).<br />

The bulk sample was analyzed by GC(RI), GC-MS, and 13 C NMR, and the fractions of chromatography<br />

were investigated by GC(RI) and 13 C NMR. In all, 46 constituents, which represented<br />

91.0% of the oil, were identified. As expected, cis- and trans-p-mentha-2,8-dien-1-ols (8.7% and<br />

18.4%), cis- and trans-p-mentha-1(7),8-dien-2-ols (16.0% and 15.7%), and limonene (12.5%) dominated<br />

largely the composition. Beside the main components, various compounds were present at<br />

appreciable contents: 3,9-oxy-mentha-1,8(10)-diene (2.2%), cis- and trans-isopiperitenols (2.2% and<br />

3.1%), carveol (2.2%), and carvone (2.7%).<br />

It should be pointed out that 22 compounds were identified by the three techniques. Obviously,<br />

all the major components were included in this group, which also contained some unusual compounds<br />

in C. giganteus oil, such as p-cymenene, 3-methylbutyl hexanaote and octanoate, safrole,<br />

precocene I, and caryophyllene oxide. A few other sesquiterpenes were identified for the first time in<br />

C. giganteus oil by GC-MS (β-elemene, (E)-β-caryophyllene, β-selinene, and δ-cadinene), besides<br />

nonterpenic esters (2-phenylethyl hexanoate and 2-phenylethyl octanoate).<br />

It is noteworthy that various compounds that were not identified by GC-MS of the whole oil<br />

sample were identified by 13 C NMR in the fractions of chromatography, in combination with RIs<br />

on apolar and polar columns: trans-dihydroperillaldehyde, geraniol, geranial, perillaldehyde,<br />

thymol, and ascaridole. The same association was useful for the identification of compounds<br />

that coeluted on apolar column, and consequently, although they were suggested by MS, they<br />

could not be accurately identified by this technique. For instance, trans- dihydrocarvone and cisdihydroperillaldehyde<br />

(RI = 1171) were identified by 13 C NMR in the fractions of chromatography,<br />

and then quantified by GC(FID) in the EO by comparison of their RIs on apolar and polar<br />

columns.<br />

A last interesting point is the identification of 3,9-oxy-p-mentha-1,8(10)-diene, not present in our<br />

MS and 13 C NMR data libraries. Its structure was determined by extensive NMR studies carried out<br />

on a fraction of chromatography where the oxide accounted for 80.0%, and then the NMR data were<br />

compared with those reported in the literature.<br />

In conclusion, the detailed analysis carried out by a combination of chromatographic and spectroscopic<br />

techniques, after fractionation of the bulk sample, gives better insight into the composition<br />

of the C. giganteus leaf oil from Côte d’Ivoire. Indeed, 25 compounds were reported for the first<br />

time in C. giganteus leaf oil, including oxygenated acyclic monoterpenes (citronellal, geraniol, geranial),<br />

oxygenated menthane derivatives (cis- and trans-dihydroperylaldehydes, thymol), phenylpropanoids<br />

(safrole), phenylethyl derivatives (hexanoate, octanoate), and oxides (ascaridole, precocene,


Analytical Methods for Cymbopogon <strong>Oil</strong>s 215<br />

caryophyllene oxyde), as well as sesquiterpene hydrocarbons (β-elemene, (E)-β-caryophyllene,<br />

β-selinene, δ-cadinene) and linear aldehydes (nonanal, decanal) and alkanes.<br />

8.4 conclusIon<br />

Nowadays, identification and quantitative determination of the individual components of EOs<br />

isolated from various species of Cymbopogon is mostly achieved by a combination of “on-line”<br />

chromatographic separation and spectroscopic identification. For that reason, a fast-scanning mass<br />

spectrometer directly coupled with a gas chromatograph is the basic equipment for such analyses.<br />

However, MS data should be considered in combination with retention indices (RIs). In summary, a<br />

component is identified (1) by comparison of its GC retention indices on polar and apolar columns,<br />

determined relative to the retention times of a series of n-alkanes with linear interpolation with those<br />

of authentic compounds or literature data; (2) on computer matching against laboratory-made and<br />

commercial mass spectral libraries and comparison of spectra with those of laboratory-made library<br />

or literature data. Identification by only one of the two techniques should be avoided, even for routine<br />

analysis.<br />

13 C NMR spectroscopy is an alternative method that has been used, at the beginning, to confirm<br />

the occurrence of a component, previously identified (or suggested) by MS or by RIs. Actually, the<br />

computerized comparison of the chemical shifts of signals in the 13 C NMR spectrum of the essential<br />

oil (EO) with those of reference spectra compiled in a library allows the identification of its<br />

main components, without previous separation.<br />

When the composition of an EO is investigated for the first time, more sophisticated techniques<br />

could be needed and a wide variety of combinations of analytical techniques are being developed,<br />

including complex hybrids such as GC-MS-MS, GC-MS-FTIR, HPLC-GC-MS, and HPLC- 1 H<br />

NMR-MS. Concerning Cymbopogon oils, to the best of our knowledge, none of these techniques<br />

have been used. Conversely, analyses of Cymbopogon oils have been carried out by a combination<br />

of various techniques, the oil being prefractionated by fractional distillation and/or column chromatography.<br />

Then, the fractions of chromatography have been analyzed by complementary techniques,<br />

GC(RI), GC-MS, 1 H, and 13 C NMR. Such types of analyses, which obviously are time consuming,<br />

provide unequivocal identification of the individual components present in EOs. Use of NMR spectroscopy<br />

for the analysis of natural mixtures is strongly encouraged by perfume, fragrance, and<br />

flavor manufacturers and quality control organizations.<br />

acKnoWledgments<br />

The authors are indebted to their coworkers and PhD students, who contributed substantially to this<br />

research. They appreciated the cooperation of colleagues of the universities of Corsica, Abidjan<br />

(Ivory Coast), and Hanoi (Vietnam). They acknowledge the Collectivité Territoriale de Corse and<br />

the European Community for partial financial support.<br />

reFerences<br />

Abd Allah M. A., Foda Y. H., Saleh M., Zaki M. S. A., and Mostafa M. M. (1975) Identification of the volatile<br />

constituents of the Egyptian lemongrass oil. Part I. Gas chromatographic analysis. Nahrung, 19,<br />

195–200.<br />

Abegaz B., Yohannes P. G., and Dieter R. K. (1983) Constituents of the essential oil of Ethiopian Cymbopogon<br />

citratus Stapf. J. Nat. Prod., 46, 424–426.<br />

Adams R. P. (2007) Identification of <strong>Essential</strong> <strong>Oil</strong> Components by Gas Chromatography/Mass Spectrometry,<br />

4th ed., Allured Publishing, Carol Stream, IL.<br />

Ahmad V. U. and Jassbi A. R. (1999) Analysis of the essential oil of Echinophora sibthorpiana Guss. by means<br />

of GC, GC/MS and 13C-NMR techniques. J. Essent. <strong>Oil</strong> Res., 11, 107–108.


216 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Al-Burtamani S. K. S., Fatope M. O., Marwah R. G., Onifade A. K., and Al-Saidi S. H. (2005) Chemical composition,<br />

antibacterial and antifungal activities of the essential oil of Haplophyllum tuberculatum from<br />

Oman. J. Ethnopharmacol., 96, 107–112.<br />

Alencar J. W., Vasconcelos Silva M. G., Lacerda Machado M. I., Craveiro A. A., Abreu Matos F. J., and<br />

de Abreu Magalhães R. (1997) Use of 13 C-NMR as complementary identification tool in essential oil<br />

analysis. Spectroscopy, 13, 265–273.<br />

Alitonou G. A., Avlessi F., Sohounhloue D. K., Agnaniet H., Bessiere J. -M., and Menut C. (2006) Investigations<br />

on the essential oil of Cymbopogon giganteus from Benin for its potential use as an anti-inflammatory<br />

agent. Internat. J. Aromatherapy, 16, 37–41.<br />

Ayedoun M. A., Moudachirou M., and Lamaty G. (1997) Composition chimique des huiles essentielles de deux<br />

espèces de Cymbopogon du Bénin exploitable industriellement. Bioressources, Energie, Développement,<br />

Environnement, 8, 4–6.<br />

Badjah Hadj Ahmed A. Y., Meklati B. Y., and Fraisse D. (1988) Mass spectrometry with positive chemical<br />

ionization of some α-pinene and β-pinene derivatives. Flavour Fragr. J., 3, 13–18.<br />

Bagheri R., Mohamadi S., Abkar A., and Fazlollahi A. (2007) <strong>Essential</strong> oil compositions of Cymbopogon parkeri<br />

STAPF from Iran. Pakistan J Biol. Sci., 10, 3485–3486.<br />

Baldovini N., Tomi F., and Casanova J. (2003) Enantiomeric differentiation of bornyl acetate by carbon-13<br />

NMR using a chiral lanthanide shift reagent. Phytochem. Anal., 14, 241–244.<br />

Baldovini N., Tomi F., and Casanova J. (2001) Identification and quantitative determination NMR of furanodiene,<br />

a heat-sensitive compound, in essential oils by 13 C-NMR. Phytochem. Anal., 12, 58–63.<br />

Baser K. H. C., Özek T., and Demirci B. (2005) Composition of the essential oil of Cymbopogon afronardus<br />

Stapf from Uganda. J. Essent. <strong>Oil</strong> Res., 17, 139–140.<br />

Bicchi C., D’Amato A., and Rubiolo P. (1999) Cyclodextrin derivatives as chiral selectors for direct gas chromatographic<br />

separation of enantiomers in the essential oil, aroma and flavour fields. J. Chromatogr. A,<br />

843, 99–121.<br />

Bicchi C., D’Amato A., Manzin V., and Rubiolo P. (1997) Cyclodextrin derivatives in GC separation of racemic<br />

mixtures of volatiles. Part XI. Some applications of cyclodextrin derivatives in GC enantioseparations<br />

of essential oil components. Flavour Fragr. J., 12, 55–61.<br />

Bicchi C., Rubiolo P., Marschall H., Weyerstahl P., and Laurent R. (1998) Constituents of Artemisia roxburghiana<br />

Besser essential oil. Flavour Fragr. J., 13, 40–46.<br />

Blanc M. C., Bradesi P., Gonçalves M. J., Salgueiro L., and Casanova J. (2006) <strong>Essential</strong> oil of Dittrichia viscosa<br />

ssp. viscosa: Analysis by 13 C-NMR and antimicrobial activity. Flavour Fragr. J., 21, 324–332.<br />

Boelens M. H., Boelens H., and Van Germert L. J. (1993) Sensory properties of optical isomers, Perf. Flav.,<br />

18, 1–16.<br />

Boti J. B., Muselli A., Tomi F., Koukoua G., N’Guessan T. Y., Costa J., and Casanova J. (2006) Combined analysis<br />

of Cymbopogon giganteus Chiov. Leaf oil from Côte d’Ivoire by GC/RI, GC-MS and 13C-NMR.<br />

C.R. Chimie, 9, 164–168.<br />

Bottini A. T., Dev V., Garfagnoli D. J., Hope H., Joshi P., Lohani H., Mathela C. S., and Nelson T. E. (1987)<br />

Isolation and crystal structure of a novel dihemiacetal bis-monoterpenoid from Cymbopogon martini.<br />

Phytochemistry, 26, 2301–2302.<br />

Bradesi P., Bighelli A., Tomi F., and Casanova J. (1996) L’analyse des mélanges complexes par RMN du<br />

carbone-13. Can. J. Appl. Spectrosc., 41, 15–24 and 41, 41–50.<br />

Bretmaier E. (2006) Terpenes, Flavors, Fragrances, Pharmaca, Pheromones. Wiley-VCH-Verlag GmbH and<br />

Co., Weinheim.<br />

Cavaleiro C. (2001) Öleos essenciais de Juniperus de Portugal. Thèse de doctorat, Université de Coïmbra.<br />

Cazaussus A., Pes R., Sellier N., and Tabet J. C. (1988) GC-MS and GC-MS-MS analysis of a complex essential<br />

oil. Chromatographia, 25, 865–869.<br />

Chagonda L. S., Makanda C., and Chalchat J. C. (2000) The essential oils of wild and cultivated Cymbopogon<br />

validus (Stapf) Stapf ex Burtt Davy and elionurus muticus (Spreng.) Kunth from Zimbabwe. Flavour<br />

Fragr. J., 15, 100–104.<br />

Chalchat J. C., Garry R. P., Harama M., and Sidibé L. (1998) A study of aromatic plants from Mali. Chemical<br />

composition of essential oils of two Cymbopogon varieties. Cymbopogon citratus and Cymbopogon<br />

giganteus. Riv. Ital. EPPOS, Spec. Num., 741–752.<br />

Chamblee T. S., Karelitz R. L., Radford T., and Clark B. C. (1997) Identification of sesquiterpenes in Citrus<br />

essential oils by cryofocusing GC/FT-IR. J. Essent. <strong>Oil</strong> Res., 9, 127–132.<br />

Chisowa E. H., Hall D. R., and Farman D. I. (1998) Volatile constituents of the essential oil of Cymbopogon<br />

citratus Stapf frown in Zambia. Flavour Fragr. J., 13, 29–30.


Analytical Methods for Cymbopogon <strong>Oil</strong>s 217<br />

Coleman III W. M., Gordon B. M., and Lawrence B. M. (1989) Examinations of the matrix isolation Fourier<br />

transform infrared spectra of organic compounds: part XII, Appl. Spectrosc., 43, 298–304.<br />

Corticchiato M. and Casanova J. (1992) Analyse des mélanges complexes par RMN du Carbone-13. Application<br />

aux huiles essentielles. Analusis, 20, M51–57.<br />

De Medeci D., Pieretti S., Salvatore G., Nicoletti M., and Rasoanaivo P. (1992) Chemical analysis of essential<br />

oils of Malagasy medicinal plants by gas chromatography and NMR spectroscopy. Flavour Fragr. J., 7,<br />

275–281.<br />

Decouzon M., Géribaldi S., Rouillard M., and Sturla J. -M. (1990) A new look at the spectroscopic properties<br />

of dihydrocarveol stereoisomers. Flavour Fragr. J., 5, 147–152.<br />

Derome A. E. (1987) Modern NMR Techniques for Chemistry Research, Vol. 6, Organic chemistry series,<br />

J. E. Baldwin Ed., Pergamon Press, Oxford.<br />

Dhar A. K., Thappa R. K., and Atal C. K. (1981) Variability in yield and composition of essential oil in<br />

Cymbopogon jawarancus. Planta Med., 41, 386–388.<br />

Dugo P., Mondello L., Dugo L., Stancanelli R., and Dugo G. (2000) LC-MS for the identification of oxygen<br />

heterocyclic compounds in citrus essential oils. J. Pharm. Biomed. Anal., 24, 147–154.<br />

El-Askari H. I., Meselhy M. R., and Galal A. M. (2003) Sesquiterpenes from Cymbopogon proximus.<br />

Molecules, 8, 670–677.<br />

Evans F. E., Miller D. W., Cairns T., Vernon Baddeley G., and Wenkert E. (1982) Structure analysis of proximadiol<br />

(cryptomeridiol) by 13 C NMR spectroscopy. Phytochemistry, 21, 937–938.<br />

Ferreira M. J. P., Costantin M. B., Sartorelli P., Rodrigues G. V., Limberger R., Henriques A. T., Kato M. J., and<br />

Emerenciano V. P. (2001a) Computer-aided method for identification of components in essential oils by<br />

13 C NMR spectroscopy. Anal. Chim. Acta, 447, 125–134.<br />

Ferreira M. J. P., Brant A. J. C., Rodrigues G. V., and Emerenciano V. P. (2001b) Automatic identification<br />

of terpenoids skeletons through 13 C nuclear magnetic resonance data disfunctionalization. Anal. Chim.<br />

Acta, 429, 151–170.<br />

Foda Y. H., Abdallah M. A., Zaki M. S., and Mostafa M. M. (1975) Identification of the volatile constituents of<br />

the Egyptian lemongrass oil. Part III. Infrared spectroscopy. Nahrung, 19, 395–400.<br />

Formácek V. and Kubeczka K. H. (1982a) <strong>Essential</strong> <strong>Oil</strong>s Analysis by Capillary Gas Chromatography and<br />

Carbon-13 NMR Spectroscopy, John Wiley & Sons, Chichester.<br />

Formácek V. and Kubeczka K. H. (1982b) 13C NMR Analysis of <strong>Essential</strong> <strong>Oil</strong>s, in Aromatic Plants: Basic and Applied<br />

Aspects, pp. 177–18, Margaris N., Koedam A., Vokou D., Eds., Martinus Nijhoff Publishers, La Haye.<br />

Fraser R. R., Stothers J. B., and Tan C. T. (1973) Determination of optical purity with chiral shift reagents and<br />

13 C magnetic resonance. J. Magn. Reson., 10, 95–97.<br />

Gaydou E. M. and Randriamiharisoa R. P. (1987) Gas chromatographic diastereomer separation of linalool<br />

derivatives. Application to the determination of the enantiomeric purity of linalool in essential oils.<br />

J. Chromatogr., 396, 378–381.<br />

Gonny M., Bradesi P., and Casanova J. (2004) Identification of the components of the essential oil from wild<br />

Corsican Daucus carota L. using 13 C-NMR spectroscopy. Flav. Fragr. J., 19, 424–433.<br />

Grundschober F. (1991) The identification of individual components in flavorings and flavoured foods.<br />

Z. Lebensm Unters Forsch., 192, 530–534.<br />

Hedges L. M. and Wilkins C. L. (1991) Components analysis of eucalyptus oil by gas chromatography–Fourier<br />

transform–infrared spectrometry–mass spectrometry. J. Chromatogr. Sci., 29, 345–350.<br />

Huffman J. W. and Pinder A. R. (1980) Comments on the structure of isointermedeol, a sesquiterpene reported<br />

from Cymbopogon flexuosus. Phytochemistry, 19, 2468–2469.<br />

Jirovetz L., Buchbauer G., Eller G., Ngassoum M. B., and Maponmetsem P. M. (2007) Composition and antimicrobial<br />

activity of Cymbopogon giganteus (Hochst.) Chiov. <strong>Essential</strong> flower, leaf and stem oils from<br />

Cameroon. J. Essent. <strong>Oil</strong> Res., 19, 485–489.<br />

Joulain D. (1994) Methods for analyzing essential oils. Modern analysis methodologies: use and abuse.<br />

Perfurmer and Flavorist, 19, 5–17.<br />

Joulain D. and Laurent R. (1989) Two closely related sesquiterpenols: 1-endo-bourbonanol and 1,6-germacradien-5-ol.<br />

J. Essent. <strong>Oil</strong> Res., 1, 299–301.<br />

Joulain D. and König W. A. (1998) The atlas of spectral data of sesquiterpene hydrocarbons, E.B. Verlag,<br />

Hamburg.<br />

Kanko C., Koukoua G., N’Guessan Y. T., Fournier J., Pradère J. P., and Toupet L. (2004) Etude des propriétés<br />

physico-chimiques des huiles essentielles de Lippia multiflora, Cymbopogon citratus, Cymbopogon nardus,<br />

Cymbopogon giganteus. C.R. Chimie, 7, 1039–1042.<br />

Kasali A. A., Oyedeji A. O., and Ashilokun A. O. (2001) Volatile leaf oil constituents of Cymbopogon citratus<br />

(DC) Stapf. Flavour Fragr. J., 16, 377–378.


218 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Keita A. (1993) Terpènes de l’huile essentielle des inflorescences de « ce kala » Cymbopogon anteus (Chiov.).<br />

Médecine Afr. Noire, 40, 229–233.<br />

Khadri A., Serralheiro M. L. M., Nogueira J. M. R., Neffati M., Smiti S., and Araujo M. E. M. (2008) Antioxidant<br />

and antiacetylcholinesterase activities of essential oils from Cymbopogon schoenanthus L. Spreng.<br />

Deter mi na tion of chemical composition by GC–mass spectrometry and 13 C NMR. Food Chemistry, 109,<br />

630–637.<br />

König W. A., Hochmuth D. H., and Joulain D. (2001) Terpenoids and related constituents of essential oils.<br />

Library of MassFinder 2.1, Institute of Organic Chemistry, Hamburg.<br />

Korhammer S. A. and Bernreuther A. (1996) Hyphenation of high-performance liquid chromatography (HPLC)<br />

and other chromatographic techniques (SFC, GPC, GC, CE) with nuclear magnetic resonance (NMR): A<br />

review. Fresenius J. Anal. Chem., 354, 131–135.<br />

Kováts E. (1965) Gas chromatographic characterization of organic substances in the retention index system, in<br />

Advances in Chromatography, vol. 1, pp. 229–247, Giddings J.C., and Keller R.A., Eds., Marcel Dekker,<br />

New York.<br />

Kreis P. and Mosandl A. (1994) Chiral compounds of essential oils. VVVII: Simultaneous stereoanalysis of<br />

Cymbopogon oil constituents. Flavour Fragr. J., 9, 257–260.<br />

Krock K. A., Ragunathan N., and Wilkins C. L. (1994) Multidimensional gas chromatography coupled with<br />

infrared and mass spectrometry for analysis of Eucalyptus essential oils, Anal. Chem., 66, 425–430.<br />

Kubeczka K. H. (1985) Progress in isolation techniques for essential oil constituents, in <strong>Essential</strong> <strong>Oil</strong>s and<br />

Aromatic Plants, pp. 107–126, Baerheim Svendsen and Scheffer Eds., Nijhoff Junk pub, Dordrecht.<br />

Kubeczka K. H. and Formácek V. (2002) <strong>Essential</strong> oils analysis by capillary gas chromatography and carbon-13<br />

NMR spectroscopy. Second ed. John Wiley & Sons, NY, pp. 67–80.<br />

Kumar A., Malik F., Bhushan S., Sethi V. K., Shahi A. K., Kaur J., Taneja S. C., Qazi G. N., and Singh J. (2008)<br />

An essential oil and its major constituent isointermedeol induce apoptosis by increased expression of<br />

mitochondrial cytochrome C and apical death receptors in human leukaemia HL-60 cells. Chem.-Biol.<br />

Interactions, 171, 332–347.<br />

Lange G. and Schultze W. (1988a) Differentiation of isopulegol isomers by chemical ionization mass spectrometry,<br />

in Bioflavour ‘87, p. 115–122, P. Schreier Ed., W. de Gruyter and Co., Berlin.<br />

Lange G. and Schultze W. (1988b), Studies on terpenoid and non-terpenoid esters using chemical ionization<br />

mass spectrometry in GC/MS coupling, in Bioflavour’87, pp. 105–114, Schreier P. Ed., de Gruyter W.<br />

and Co., Berlin.<br />

Lawrence B. M. (1979a) Citronella oil, in <strong>Essential</strong> <strong>Oil</strong>s 1976–1978, p. 22, Allured Publishing Co., Carol<br />

Stream, IL.<br />

Lawrence B. M. (1979b) Lemongrass oil, in <strong>Essential</strong> <strong>Oil</strong>s 1976–1978, p. 20 and pp. 30–31, Allured Publishing,<br />

Carol Stream, IL.<br />

Lawrence B. M. (1989a) Citronella oil, in <strong>Essential</strong> <strong>Oil</strong>s 1981–1987, pp. 29–31 and pp. 147–149, Allured Publishing,<br />

Carol Stream, IL.<br />

Lawrence B. M. (1989b) Lemongrass oil, in <strong>Essential</strong> <strong>Oil</strong>s 1981–1987, p. 111, Allured Publishing, Carol<br />

Stream, IL.<br />

Lawrence B. M. (1989c) Palmarosa oil, in <strong>Essential</strong> <strong>Oil</strong>s 1981–1987, pp. 237–239, Allured Publishing, Carol<br />

Stream, IL.<br />

Lawrence B. M. (1993) Palmarosa oil, in <strong>Essential</strong> <strong>Oil</strong>s 1988–1991, pp. 81–82, Allured Publishing, Carol<br />

Stream, IL.<br />

Lawrence B. M. (1995a) Citronella oil, in <strong>Essential</strong> <strong>Oil</strong>s 1992–1994, pp. 82–83 and p. 196, Allured Publishing,<br />

Carol Stream, IL.<br />

Lawrence B. M. (1995b) Lemongrass oil, in <strong>Essential</strong> <strong>Oil</strong>s 1992–1994, p. 60, Allured Publishing, Carol<br />

Stream, IL.<br />

Lawrence B.M. (2003a) Citronella oil, in <strong>Essential</strong> <strong>Oil</strong>s 1995–2000, pp. 52–53 and pp. 201–203, Allured Publish<br />

ing, Carol Stream, IL.<br />

Lawrence B. M. (2003b) Ginger grass oil, in <strong>Essential</strong> <strong>Oil</strong>s 1995–2000, pp. 345–346, Allured publishing, Carol<br />

Stream, IL.<br />

Lawrence B. M. (2003c) Lemongrass oil, in <strong>Essential</strong> <strong>Oil</strong>s 1995–2000, pp. 206–210, Allured Publishing, Carol<br />

Stream, IL.<br />

Lawrence B. M. (2003d) Palmarosa oil, in <strong>Essential</strong> <strong>Oil</strong>s 1995–2000, pp. 334–335, Allured Publishing, Carol<br />

Stream, IL.<br />

Lawrence B. M. (2006a) Citronella oil, in <strong>Essential</strong> <strong>Oil</strong>s 2001–2004, pp. 142–145, Allured Publishing, Carol<br />

Stream, IL.


Analytical Methods for Cymbopogon <strong>Oil</strong>s 219<br />

Lawrence B. M. (2006b) Lemongrass oil, in <strong>Essential</strong> <strong>Oil</strong>s 2001–2004, pp. 128–132, Allured Publishing, Carol<br />

Stream, IL.<br />

Lawrence B. M. (2006c) Palmarosa oil, in <strong>Essential</strong> <strong>Oil</strong>s 2001–2004, pp. 77–78 and pp. 313–314, Allured<br />

Publishing, Carol Stream, IL.<br />

Lawrence B. M. (2006d) Citronella oil. Perf. Flav., 31, 10, 46.<br />

Liu C., Zhang J., Yiao R., and Gan L. (1981) Chemical studies on the essential oils of Cymbopogon genus.<br />

Huaxue Kuebao 241–247.<br />

Lorenzo D., Dellacassa E., Atti-Serafini L., Santos A. C., Frizzo C., Paroul N., Moyna P., Mondello L., and<br />

Dugo G. (2000) Composition and stereoanalysis of Cymbopogon winterianus Jowitt oil from Southern<br />

Brazil. Flavour Fragr. J., 15, 177–181.<br />

Mahalwal V. S. and Ali M. (2003) Volatile constituents of Cymbopogon nardus (Linn.) Rendle. Flavour<br />

Fragr. J., 18, 73–76.<br />

Mallavarapu G. R., Rajeswara Rao B. R., Kaul P. N., Ramesh S., and Bhattacharya A. K. (1998) Volatile<br />

constituents of the essential oils of the seeds and the herb of palmarosa (Cymbopogon martinii (Roxb.)<br />

Wats. var. motia Burk. Flavour Fragr. J., 13, 167–169.<br />

Mallavarapu G. R., Ramesh S., Kulkarni R. N., and Syamasundar K. V. (1992) Composition of the essential oil<br />

of Cymbopogon travancorensis. Planta Med., 58, 219–220.<br />

Marongiu B., Piras A., Porcedda S., and Tuveri E. (2006) Comparative analysis of the oil and supercritical CO 2<br />

extract of Cymbopogon citratus Stapf. Nat. Prod. Res., 20, 455–459.<br />

Mathela C. S., Melkani A. B., Pant A., Dev V., Nelson T. E., Hope H., and Bottini A. T. (1989) A eudesmanediol<br />

from Cymbopogon distans. Phytochemistry, 28, 936–938.<br />

Mathela C. S. and Joshi P. (1981) Terpenes from the essential oil of Cymbopogon distans. Phytochemistry, 20,<br />

2770–2771.<br />

Mathela C. S., Pandey C., Pant A. K., and Singh A. K. (1990) Phenylpropanoid constituents of Cymbopogon<br />

microstachys. J. lnd. Chem. Soc., 67, 526–528.<br />

McLafferty F. W. and Stauffer D. B. (1988) Wiley/NBS Registry of Mass Spectral Data, 4th ed., Wiley-<br />

Interscience, New York.<br />

McLafferty F. W. and Stauffer D. B. (1994) Wiley Registry of Mass Spectral Data. 6th ed. Mass Spectrometry<br />

Library Search System Bench–Top/PBM, version 3.10d, Palisade, Newfield, UK.<br />

Melkani A. B., Joshi P., Pant A. K., Mathela C. S., and Dev V. (1985) Constituents of the essential oils from<br />

two varieties of Cymbopogon distans. J. Nat. Prod., 48, 995–997.<br />

Menut C., Bessiere J. M., Samate D., Djibo A. K., Buchbauer G., and Schopper B. (2000) Aromatic plants of<br />

tropical west Africa. XI. Chemical composition, antioxidant and antiradical properties of the essential<br />

oils of three Cymbopogon species from Burkina Faso. J. Essent. <strong>Oil</strong> Res., 12, 207–212.<br />

Mohammad F., Nigam M. C., and Rahman W. (1981) Detection of new trace constituents in the essential oils<br />

of Cymbopogon flexuosus. Perf. Flav., 6, 29–31.<br />

Mondello L., Casilli A., Tranchida P. Q., Dugo P., and Dugo G. (2005) Comprehensive two-dimensional GC for<br />

the analysis of citrus essential oils. Flavour Fragr. J., 20, 136–140.<br />

Mondello L., Dugo G., Dugo P., and Bartle K. D. (1996) On-line HPLC-HRGC in the analytical chemistry of<br />

Citrus essential oils. Perf. Flav., 21, 25–49.<br />

Mondello L., Dugo P., and Bartle K. D. (1995) Automated HPLC-HRGC: A powerful method for essential oils<br />

analysis. Part V. Identification of terpene hydrocarbons of bergamot, lemon, mandarin, sweet orange, bitter<br />

orange, grapefruit, clementine and Mexican lime oils by coupled HPLC-HRGC-MS (ITD). Flavour<br />

Fragr. J., 10, 33–42.<br />

Mosandl A. (1988) Chirality in flavour chemistry—recent developments in synthesis and analysis. Food Rev.<br />

Intern., 4, 1–43.<br />

Mosandl A., Hener O., Kreis P., and Schmorr H. G. (1990) Enantiomeric distribution of a-pinene, b-pinene and<br />

limonene in essential oils and extracts. Part I. Rutaceae and gramineae. Flavour Fragr. J., 5, 193–199.<br />

NIST (1999) National Institute of Standards and Technology, PC Version 1.7 of The NIST/EPA/NIH mass<br />

spectral library. Perkin–Elmer Corp., Norwalk, CT.<br />

Norouzi-Arasi H., Yavari I., Ghaffarzade F., and Mortazavi M. S. (2002) Volatile constituents of Cymbopogon<br />

olivieri (Boiss.) Bor. from Iran. Flavour Fragr. J., 17, 272–274.<br />

Ntezurubanza L., Collin G., Deslauriers G., and Nizeyimana J. B. (1992) Huiles essentielles de Geranium et<br />

de lemongrass cultivées au Rwanda. Rivista Ital. EPPOS (Num. Spec.), 631–639.<br />

Núñez C. V. and Roque N. F. (1999) Sesquiterpenes from the stem bark of Guarea guidonia (L.) Sleumer<br />

(Meliaceae). J. Essent. <strong>Oil</strong> Res., 11, 439–440.<br />

Ottavioli J., Bighelli A., Casanova J., Bui Thi Bang, and Pham Van Y. (2008a) unpublished results.


220 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

Ottavioli J., Bighelli A., Casanova J., Bui Thi Bang, and Pham Van Y. (2008b) submitted.<br />

Ouamba J.-M. (1991) Valorisation chimique des plantes aromatiques du Congo. Extraction et analyse des huiles<br />

essentielles, Oximation des aldéhydes naturels, Thèse de Doctorat d’Etat, Université Montpellier II.<br />

Paolini J., Costa J., and Bernardini A. F. (2005) Analysis of the essential oil from aerial parts of Eupatorium<br />

cannabinum subsp. corsicum (L.) by gas chromatography with electron impact and chemical ionisation<br />

mass spectrometry, J. Chromatogr. A., 1076, 170–178.<br />

Peyron L. (1973) Sur quelques essences en provenance du Mato Grosso. Parfum. Cosmet. Savon. France, 3,<br />

371–378.<br />

Popielas L., Moulis C., Kéita A., Fouraste I., and Bessière J. M. (1991) The essential oil of Cymbopogon giganteus,<br />

Planta Med., 57, 586–587.<br />

Prashar A., Hili P., Veness R. G., and Evans C. S. (2003) Antimicrobial action of palmarosa oil (Cymbopogon<br />

martinii) on Saccharomyces cerevisiae. Phytochemistry, 63, 569–575.<br />

Raina V. K., Srivastava S. K., Aggarwal K. K., Syamasundar K. V., and Khanuja S. P. S., (2003) <strong>Essential</strong> oil<br />

composition of Cymbopogon martinii from different places in India, Flavour Fragr. J., 18, 312–315.<br />

Ramidi R., Ali M., Velasco-Negueruela A., and Pérez-Alonso M. J. (1998) Chemical composition of the seed<br />

oil of Zanthoxylum alatum Roxb. J. Essent. <strong>Oil</strong> Res., 10, 127–130.<br />

Randriamiharisoa R. P. and Gaydou E. M. (1987) Composition of palmarosa (Cymbopogon martini) essential<br />

oil from Madagascar. J. Agr. Food Chem., 35, 62–66.<br />

Rauber C. S., da Guterres S. S., and Schapoval E. E. S. (2005) LC determination of citral in Cymbopogon citratus<br />

volatile oil. J. Pharm. Biomed. Anal., 37, 597–601.<br />

Ravid U., Putievsky E., and Katzir I. (1992) Chiral GC analysis of enantimerically pure fenchone in essential<br />

oils. Flavour Fragr. J., 7, 169–172.<br />

Ravid U., Putievsky E., and Katzir I. (1996) Stereochemical analysis of borneol in essential oils using permethylated<br />

beta-cyclodextrin as a chiral stationary phase. Flavour Fragr. J., 11, 191–195.<br />

Ravid U., Putievsky E., Weinstein V., and Ikan R. (1985) Determination of the enantiomeric composition of<br />

natural flavouring agents by 1H-NMR spectroscopy, pp. 135–138 in <strong>Essential</strong> <strong>Oil</strong>s and Aromatic Plants,<br />

Baerheim Svendsen and Scheffer Eds., Nijhoff Junk, Dordrecht.<br />

Rezzi S., Bighelli A., Castola V., and Casanova J. (2002) Direct identification and quantitative determination<br />

of acidic and neutral diterpenes using 13 C-NMR spectroscopy. Application to the analysis of oleoresin of<br />

Pinus Nigra. Appl. Spectrosc., 56, 312–317.<br />

Ristorcelli D., Tomi F., and Casanova J. (1998) Carbon-13 NMR as a tool for identification and enantiomeric<br />

differentiation of major terpenes exemplified by the essential oil of Lavandula Stoechas. Flavour<br />

Fragr. J., 13, 154–158.<br />

Rizk A. M., Heiba H. I., Sandra P., Mashaly M., and Bicchi C. (1983) Constituents of plants growing in qarar,<br />

constituents of the volatile oil of Cymbopogon Parkeri. J. Chromatogr., 279, 145–150.<br />

Rouessac F. and Rouessac A. (1994) Analyse chimique. Méthodes et Techniques Instrumentales Modernes.<br />

Masson, Ed., Paris.<br />

Rout P. K., Sahoo S., Rao Y. R. (2005) <strong>Essential</strong> oil composition of Cymbopogon microstachys (Hook.)<br />

Soenarke occurring in Manipur. J. Essent. <strong>Oil</strong> Res., 17, 358–360.<br />

Saeed T., Sandra P., and Verzele M. (1978) Constituents of essential oil of Cymbopogon jawarancusa.<br />

Phytochemistry, 17, 1433–1434.<br />

Sargenti S. R. and Lanças F. M. (1997) Supercritical fluid extraction of Cymbopogon citratus. Chromatographia,<br />

46, 285–290.<br />

Schultze W., Lange G., and Schmaus G. (1992) Isobutane and ammonia chemical ionization mass spectrometry<br />

of sesquiterpene hydrocarbons. Flavour Fragr. J., 7, 55–64.<br />

Sidibé L., Chalchat J. C., Garry R. P., and Lacombe L. (2001) Aromatic plants of Mali (IV): Chemical composition<br />

of essential oils of Cymbopogon citratus (DC) Stapf and Cymbopogon giganteus (Hochst) Chiov.<br />

J. Essent. <strong>Oil</strong> Res., 13, 110–112.<br />

Spring O., Buschmann H., Vogler B., Schilling E. E., Spraul M., and Hofmann M. (1995) Sesquiterpene lactone<br />

chemistry of Zaluzania grayana from on-line LC-NMR measurements. Phytochemistry, 39, 609–612.<br />

Surburg H. and Panten J. (2006) Common Fragrance and Flavor Materials, 5th ed., Wiley-VCH Verlag GmbH<br />

and Co., Weinheim.<br />

Taskinen Jyrki, Mathela D. K., and Mathela C. S. (1983) Composition of the essential oil of Cymbopogon<br />

flexuosus. J. Chromatogr. A, 262, 364–366.<br />

Thappa R. K., Dhar K. L., and Atal C. K. (1979) Isointermedeol, a new sesquiterpene alcohol from Cymbopogon<br />

flexuosus. Phytochemistry, 18, 671–672.<br />

Tomi F. and Casanova J. (2000) Contribution de la RMN du carbone-13 à l’analyse des huiles essentielles. Ann.<br />

Fals Expert. Chim., 93, 313–330.


Analytical Methods for Cymbopogon <strong>Oil</strong>s 221<br />

Tomi F. and Casanova J. (2006) 13 C NMR as a tool for identification of individual components of essential oils<br />

from Labiatae—A review. Acta Horticult., 723, 185–192.<br />

Tomi F., Bradesi P., Bighelli A., and Casanova J. (1995) Computer-aided identification of individual components<br />

of essential oil using carbon-13 NMR spectroscopy. J. Magn. Reson. Anal., 1, 25–34.<br />

Van Den Dool H. and Kratz P. D. (1963) A generalization of the retention index system including linear temperature<br />

programmed gas-liquid partition chromatography, J. Chromatogr., 11, 463–471.<br />

Vernin G., Metzger J., Suon K.N., Fraisse D., Ghiglione C., Hamoud A., and Párkányi C. (1990) GC–MS–<br />

SPECMA bank analysis of essential oils and aromas. GC-MS (EI-PCI) data bank analysis of sesquiterpenic<br />

compounds in juniper needle oil—Application of the mass fragmentometry SIM technique,<br />

Lebensmittel-Wissenschaft Technol., 23, 25–33.<br />

Vernin G., Petitjean M., Poite J. C., Metzger J., Fraisse D., and Suon K. N. (1986) Computer Aids to Chemistry,<br />

Chap. VII, pp. 294–333, Vernin G., Chanon M., Horwood E., Eds., Mass Spectra and Kováts’ Indices<br />

Databank of Volatile Aroma Compounds, Chichester.<br />

Wang X. H., Jia C. R., and Wan H. (1995) The direct chiral separation of some optically active compounds in<br />

essential oils by multidimensional gas chromatography. J. Chromatogr. Sci., 33, 22–25.<br />

Weyerstahl P., Marschall H., Splittgerber U., and Wolf D. (2000) 1,7-Cyclogermacra-1(10),4-dien-15-al, a<br />

sesquiterpene with a novel skeleton, and other sesquiterpenes from Haitian vetiver oil. Flavour Fragr. J.,<br />

15, 61–83.<br />

Weyerstahl P., Marschall H., Weirauch M., Thefeld K., and Surburg H. (1998) Constituents of commercial<br />

Labdanum oil. Flavour Fragr. J., 13, 295–318.<br />

Wijesekera R. O. B., Jayewardene A. L., and Fonseka B. D. (1973) Varietal differences in the constituents of<br />

citronella oil, Phytochemistry, 12, 2697–2704.<br />

Yarovaya O. I., Salomatina O. V., Korchagina D. V., Polovinka M. P., and Barkhash V. A. (2002) Transformations<br />

of 6,7-epoxy derivatives of citral and citronellal in various acidic media. Russian J. Org. Chem., 38,<br />

1649–1660.<br />

Zaki M. S. A., Foda Y. H., Mostafa M. M., and Abd Allah M. A. (1975) Identification of the volatile constituents<br />

of the Egyptian lemongrass oil. Part II. Thin Layer Chromatography. Food Nahrung, 19, 201–205.


9<br />

contents<br />

Citral from Lemongrass and<br />

Other Natural Sources<br />

Its Toxicology and Legislation<br />

David A. Moyler<br />

9.1 Introduction ..........................................................................................................................223<br />

9.2 Uses and Dose Rate ..............................................................................................................227<br />

9.3 Natural Sources ....................................................................................................................227<br />

9.3.1 Litsea cubeba <strong>Oil</strong> ......................................................................................................227<br />

9.3.2 Lemongrass <strong>Oil</strong>s .......................................................................................................229<br />

9.4 Manufacturing Methods .......................................................................................................230<br />

9.5 Analysis Methods .................................................................................................................230<br />

9.6 Toxicology ............................................................................................................................ 231<br />

9.6.1 RIFM ........................................................................................................................ 231<br />

9.7 Classification and Labeling ..................................................................................................234<br />

9.8 Safety Data Sheets ................................................................................................................ 235<br />

9.8.1 GHS ..........................................................................................................................236<br />

9.8.2 Reach ........................................................................................................................236<br />

9.9 Conclusion ............................................................................................................................236<br />

Acknowledgments ..........................................................................................................................237<br />

References ...................................................................................................................................... 237<br />

Other Sources .................................................................................................................................237<br />

9.1 IntroductIon<br />

Natural essential oils containing citral have been used in traditional and oriental folk medicines<br />

for millennia. The industrial manufacturing and chemical synthesis of citral has been known for<br />

more than a century, and even in the early stages of development, its reactivity and readily oxidizable<br />

characteristics were well known. It has long been considered best practice in laboratories and<br />

small-scale manufacturing units that spillages collected in absorbent paper or rags should be soaked<br />

in water before disposal to prevent combustion of citral and organic matter in the air. Citral is also<br />

active on human skin and can cause irritation, and even sensitization in some cases, although under<br />

“good manufacturing practice” and using personal protection equipment current in the industry,<br />

cases of reported problems are fortunately few.<br />

This chapter contains many acronyms used throughout this chapter and the trade, and these are<br />

listed in Table 9.1. More complete acronym lists are provided by some trade associations for their<br />

members, for example, BEOA, IFRA. Legislatively, citral is in the EU Annex I register of regulated<br />

substances as 605-019-00-3, in the ATP 19 (EFFA 2006). Citral consists of two isomers, neral<br />

[CAS 106-26-3] and geranial [CAS 141-27-5], usually occurring naturally in the ratio of 40:60.<br />

223


224 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 9.1<br />

acronyms listed in alphabetical order<br />

AISE Association Internationale de la Savonnerie de la Detergence & des Produits d’Entreti d’Entretien<br />

(European Soap and Detergent Assn.)<br />

ATP Adaptation of Technical Progress (EU Annex I), followed by revision no.<br />

BCF Bio Concentration Factor (REACH)<br />

BEMA British Essence Manufacturer’s Association<br />

BEOA British <strong>Essential</strong> <strong>Oil</strong>s Association<br />

BFA British Fragrance Association<br />

CA Competent Authority (REACH)<br />

CAS Chemical Abstracts Service<br />

C&L Classification and Labeling<br />

CHIP Chemical Hazard Information Packaging<br />

CMR Carcinogen, Mutagen or Reproductive toxin (REACH)<br />

COE Council of Europe<br />

COLIPA Cote de Liaison European de Industrie de la Parfumerie de Produits Cosmetiques et de Toilette<br />

(European Cosmetic, Toiletry, and Perfumery Association)<br />

CREOD Center for Research Expertise in Occupational Disease (U.S.)<br />

CSA Chemical Safety Assessment (REACH)<br />

CSR Chemical Safety Report (REACH)<br />

CWG Commission Working Group (of the EU)<br />

DIN Deutsches Institut fur Normung (German Institute for Standardization)<br />

DNEL Derived No Effect Level — human health (REACH)<br />

DPD Dangerous Preparations Directive (88/379/EEC as amended)<br />

DSD Dangerous Substances Directive (67/548/EEC as amended)<br />

DU Downstream User (REACH)<br />

EC European Community (now European Union)<br />

EC0 Environmental concentration 0, the minimal quantity of a substance administered orally or dermally that<br />

does not kill any target population within a specified time<br />

EC 3 Prioritize and refer individuals for further assessment and care (EUSC)<br />

EC10 Environmental concentration 10: quantity of a substance administered orally or dermally required to kill<br />

10% of a target population within a specified time<br />

EC50 Environmental concentration 50: quantity of a substance administered orally or dermally required to kill<br />

50% of a target population within a specified time<br />

ECB European Chemicals Bureau<br />

E Ch A European Chemicals Agency (REACH)<br />

EEIII Joint Associations of ‘EFFA / EFEO / IFEAT / IFRA / IOFI’<br />

EFEO European Federation for <strong>Essential</strong> <strong>Oil</strong>s<br />

EFFA European Flavour and Fragrance Association<br />

EINECS European Inventory Notified Existing Chemical Substances<br />

ELINCS European List of Notified Chemical Substances<br />

ES Exposure Scenario (REACH)<br />

e-SDS Extended Safety Data Sheet, SDS for healthcare professionals (REACH)<br />

ESIS European Chemical Substances Information System (REACH)<br />

ESR Existing Substances Regulation (793/93/EEC)<br />

ETF Environmental Task Force (of IFRA)<br />

EU European Union<br />

EUCLID European Chemicals Information Database (REACH)<br />

EUROTOX European Federation of Toxicological Societies (REACH)<br />

EUSC End User Support Centre


Citral from Lemongrass and Other Natural Sources 225<br />

table 9.1 (continued)<br />

acronyms listed in alphabetical order<br />

F&F Flavour & Fragrance (industry)<br />

FCC Food Chemical Codex (U.S.)<br />

FDA Food and Drug Administration (U.S.)<br />

FEMA Flavor Extract Manufacturers Association (U.S. Flavor Trade)<br />

FEXPAN FEMA Expert Panel<br />

FFIDS Flavor/Fragrance Ingredient Data Sheet<br />

FMA Fragrance Materials Association (U.S.)<br />

GC-MS Gas (liquid) Chromatography (combined) Mass Spectroscopy<br />

GHS Globally Harmonized System (REACH)<br />

GLC Gas Liquid Chromatography<br />

GLP Good Laboratory Practice<br />

GMP Good Manufacturing Practice<br />

HCWG Hazard Communication Working Group (of EFFA/IFRA/IOFI jointly)<br />

HPV High Production Volume Chemicals (REACH)<br />

HRIPT Human Repeat Insult Patch Test<br />

HSDS Health and Safety Data Sheet (EU)<br />

HSE Health and Safety Executive (U.K.)<br />

IC50 Infectious concentration 50: quantity of a substance administered orally or dermally required to kill 50%<br />

of a target population within a specified time<br />

IFEAT International Federation of <strong>Essential</strong> <strong>Oil</strong> and Aroma Trades<br />

IFRA International Fragrance Association<br />

IHCP Institute for Health and Consumer Protection<br />

INCI International Nomenclature of Cosmetic Ingredients<br />

In vitro Biological testing carried out “in glass,” for example, test tube, petri dish<br />

In vivo Biological testing carried out with humans or animals<br />

IOFI International Organisation of the Flavour Industry<br />

IOFI-CE IOFI Committee of Experts<br />

ISO International Standards Organization<br />

IUCLID International Uniform Chemicals Information Database (REACH)<br />

IUPAC International Union for Pure and Applied Chemistry<br />

JAG Joint Advisory Group of IFRA<br />

K o/c Coefficient of partition between <strong>Oil</strong> and Soil (log scale)<br />

K o/w Coefficient of partition between <strong>Oil</strong> and Water (log scale)<br />

LC50 Lethal concentration 50: quantity of a substance administered by inhalation, required to kill 50% of a<br />

target population within a specified time<br />

LD50 Lethal dose 50: quantity of a substance administered orally or dermally required to kill 50% of a target<br />

population within a specified time<br />

LLNA Local Lymph Node Assay<br />

LOAEL Lowest Observed Adverse Effect Level<br />

LOEL Lowest Observed Effect Level<br />

LOEC Lowest Observed Effect Concentration<br />

LPV Low Production Volume Chemicals (REACH)<br />

MAX Maximum Allowable Concentration<br />

MCS Multiple Component Substances (e.g., methyl ionone isomers)<br />

M/I Manufacturer/Importer (REACH)<br />

MITI Ministry of Trade and Industry (Japan)<br />

MSDS Material Safety Data Sheet (U.S.)<br />

(continued on next page)


226 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 9.1 (continued)<br />

acronyms listed in alphabetical order<br />

NCS Natural Complex Substance (e.g., essential oils and extracts)<br />

NESIL No Expected Sensitization Induction Level<br />

NI Nature Identical<br />

NLP No-Longer Polymers (REACH)<br />

NOAEL No Observed Adverse Effect Level<br />

NOEC No Observed Effect Concentration<br />

NOEL No Observed Effect Level<br />

NOEL-MAX No Observed Effect Level for Maximum Allowable Concentration<br />

NORMAN Network Of Reference Laboratories Monitoring Emerging Environmental Pollutants<br />

OECD Organisation for Economic Cooperation and Development<br />

ORATS Online European Assessment Tracking System, EEC 793/93<br />

PADI Possible Average Daily Intake<br />

PBT Persistent Bio-accumulative Toxic Substance (REACH)<br />

PNEC Predicted No-Effect Concentration — environmental (REACH)<br />

PPORD Product and Process Oriented Research & Development (REACH)<br />

Ph Eur European Pharmacopoeia<br />

PRODAROM Syndicate National des Fabricants de Produits Aromatiques<br />

PRODUCE Piloting Reach On Downstream Users Compliance Exercise (REACH)<br />

QMRF QSAR Model Reporting Formats (ECB - for fish toxicity testing) (REACH)<br />

QRA Quantitative Risk Assessment (REACH)<br />

QSAR Quantitative Structure Activity Relationship (REACH)<br />

REACH Registration Evaluation Authorisation of Chemicals<br />

REXPAN RIFM Expert Panel<br />

RIFM Research Institute for Fragrance Materials<br />

RIP’s Reach Implementation Projects (REACH)<br />

RRIx Relative Retention Index (GLC term)<br />

RIPT Repeat Insult Patch Test<br />

RSC Royal Society of Chemistry (London)<br />

SAR Structure Activity Relationship (REACH)<br />

SARA Structure Activity Relationship Assessment (REACH)<br />

SCCNFP European Scientific Committee Cosmetics and Non Food Products (now SCCP)<br />

SCCP Scientific Committee on Cosmetology (EC)<br />

SHE Committee for Occupational Safety, Health and Environment<br />

SIEF Substance Information Exchange Forum (REACH)<br />

SME Small- and Medium-Sized Enterprise (REACH)<br />

SVHC Substances of Very High Concern (REACH)<br />

TEAMSPACE A Web site for posting all of the Reach Implementation Projects (RIPs)<br />

TGD Technical Guidance Document (REACH)<br />

TNO Information and Communication Technology (Netherlands)<br />

T/yr Tonnes per year (REACH)<br />

UNECE United Nations Economic Commission for Europe<br />

UNITIS European Organisation for Cosmetic Ingredient Industry and Services<br />

UVCB Unknown Variable Complex Botanicals (REACH)<br />

VPVB Very Persistent Very Bioaccumulative (REACH)<br />

WAF Water Accommodated Fraction (for eco-toxicity testing) (REACH)<br />

WGK Water Pollution Class (Wassergefahrdungsklasse), Germany<br />

WHO World Health Organisation<br />

WOE Weight of Evidence


Citral from Lemongrass and Other Natural Sources 227<br />

table 9.2<br />

concentration of citral in Final Fragrance Product<br />

9.2 uses and dose rate<br />

Applications of citral in flavoring are widespread up to a reported average maximum level of<br />

430 ppm in chewing gum, a mean average daily consumption of 137 mg in baked goods, and a PADI<br />

of 25.27 ppm (RIFM 2006). It is reported as being used in beverages, baked goods, cheese, chewing<br />

gum, condiments, frozen dairy, puddings, gravies, candy, and meat products. Its use in fragranced<br />

products is limited by the IFRA standard, which restricts the level of use of citral in consumer goods<br />

(IFRA 2007). Hence, its characteristic perfumery notes are often substituted by alternative ingredients<br />

that have the added advantage of being more stable. This is particularly true in functional<br />

products, where stability is an issue and a use level in final product is reported (Food Cosmetics<br />

1979) in Table 9.2.<br />

9.3 natural sources<br />

Although citral occurs in many NCS (see Table 9.3), the first two on the list are the main commercial<br />

sources used for its extraction.<br />

Other more uncommon oils available commercially containing high levels of citral:<br />

9.3.1 litsea C u b e b a oil<br />

soap (%) detergent (%) creams, lotions (%) Perfume (%)<br />

Usual 0.02 0.002 0.005 0.2<br />

Maximum 0.2 0.02 0.02 0.8<br />

Lindera citriodora ~65%<br />

Backhousia citriodora ~95%<br />

Calypranthes parriculata ~62%<br />

Leptospermum liversidgei var. A ~75%<br />

Ocimum gratissimum ~66%<br />

The medium-sized tree Litsea cubeba, a member of the Lauraceae family, is commercially<br />

grown for essential oil distillation in the Yunnan and Guangxi Provinces of China, where it is<br />

known locally as May Chang. The fresh berries and leaves are cut and allowed to wilt before they<br />

are used for steam-distilled oil production. There are some differences in the composition of the oils<br />

derived from pure berries and pure leaves, and the collection of wood is best avoided because it can<br />

introduce low levels of toxic safrole into the oil.<br />

Litsea cubeba oil is mobile, pale yellow in color, with a powerful, fresh, intense herbal-lemon<br />

odor that is quite prominent. Virtually its only use is as a source of natural citral, which has several<br />

applications. In lemon flavors, it reinforces the main note, but it has limited stability in functional fragrance<br />

products. Citral is used for the synthesis of vitamin A and various ionone aroma chemicals.<br />

It has been featured in ISO standard 3214, last reviewed in 2000. The citral quoted therein was<br />

69.0% to 75.0%, measured by the ISO and FCC IV methods of GLC using a nonpolar column (Food<br />

Chemical Codex 1996). Classical wet analysis gives slightly higher results than this due to reaction<br />

of other minor components, leading to the usual commercial term of “70/75.”


228 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 9.3<br />

natural citral sources<br />

Citral (neral + geranial) Lemongrass oil


Citral from Lemongrass and Other Natural Sources 229<br />

Assay typical: citral 72%, limonene 12%, linalol 1.5%, citronellol 0.2%, nerol 0.5%, geraniol 1%,<br />

1,8 cineole 1%, 6-methyl hepten-2-one 2%, pinenes 2%, sabinene 1%, β-myrcene 1%, verbenols<br />

2%, citronellal 1%, α-terpineol 0.5%, β-caryophyllene 1.5%, other monoterpene hydrocarbons 0.5%,<br />

other sesquiterpene hydrocarbons 0.2% = 99.9% (RSC EOC 2007).<br />

9.3.2 le M o n G r a s s oils<br />

The plants of lemongrass Cymbopogon citratus and C. flexuosus are members of the Poaceae family;<br />

they are medium-sized grasses that are commercially grown for essential oil distillation in<br />

Guatemala (C. citratus) and Cochin, India (C. flexuosus) (Arctander 1962). The fresh grass is cut<br />

and allowed to wilt before it is used for steam-distilled oil production. <strong>Oil</strong>s from different regions<br />

have somewhat different compositions, Guatemalan West Indian (W.I.) having a slightly different<br />

citral content than East Indian (E.I.) Cochin origin.<br />

Lemongrass oil is mobile, pale yellow in color, with a powerful, fresh, floral-herbal odor that<br />

is quite prominent. Similar to Litsea cubeba oil, it is used as source of natural citral, but because<br />

this oil is more expensive and contains a higher level of geraniol (which is difficult to remove by<br />

distillation due to the closeness of its boiling point to that of the geranial isomer of citral), it is less<br />

popular; also, its flavor profile is slightly different compared to Litsea. Removal of the geraniol by<br />

absorption of the citral with alkaline bisulfite is a chemical treatment that results in the citral losing<br />

its natural status.<br />

These oils have been featured in the ISO standards 3217 (ISO 1974) and 4718 (ISO 2004); citral,<br />

quoted therein, was 70.0% to 75.0%.<br />

The physical constants recorded in ISO were ISO 3217:1974 and ISO 4718:2004<br />

Relative density at 20°C 0.872 to 0.897 0.885 to 0.905<br />

Refractive Index at 20°C 1.483 to 1.489 1.483 to 1.489<br />

Optical rotation at 20°C +1° to −3° +1° to – 4°<br />

The samples of oils analyzed of Guatemalan and Cochin origins conformed to these values and<br />

have the following registration and properties:<br />

CAS WI; 89998-14-1 [ISO, EFFA, EFEO, BEOA]<br />

EI; 91844-92-7 [ISO, EFEO, RIFM, BEOA]<br />

EINECS WI; 289-752-0 [ISO, EFEO, BEOA]<br />

EI; 295-161-9 [ISO, EFEO, BEOA, RIFM]<br />

TSCA 8007-02-1 [INCI, RIFM]<br />

FEMA 2624<br />

RIFM 5585<br />

EC 38n<br />

FDA 182.20<br />

INCI Cymbopogon schoenanthus oil; uses: tonic, masking<br />

Flash point: 71°C [EFEO, FMA]<br />

Log Kow calc. >3 calculated<br />

Assay typical W.I: citral 73%, geraniol 2.5%, nerol 1%, 6-methyl hepten-2-one 1%, limonene 8%,<br />

linalol 1%, eugenol 0.1%, citronellol 0.5%, citronellal 0.3%, neryl + geranyl acetates, 3%, verbenols<br />

3%, nonanone 1%, B-caryophyllene 2%, caryophyllene oxide 1%, other monoterpene hydrocarbons


230 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

2%, other sesquiterpene hydrocarbons 0.3% = 99.7% [7]. E.I: citral 78%, geraniol 3.5%, nerol 1.2%,<br />

6-methyl hepten-2-one 2.3%, limonene 0.3%, linalol 1%, eugenol 0.1%, citronellol 0.5%, citronellal<br />

0.3%, neryl + geranyl acetates 3%, verbenols 3%, nonanone 1%, B-caryophyllene 2%, caryophyllene<br />

oxide 1%, other monoterpene hydrocarbons 2%, other sesquiterpene hydrocarbons 0.3% = 99.5%<br />

(RSC 2007).<br />

9.4 manuFacturIng methods<br />

Ex-chemical synthesis: legislation: US—artificial, EU—nature identical. Assay 99%.<br />

The manufacturing methods are described in several books (Arctander 1969; Ohloff 1993).<br />

Synthesis can be based on acetylene chemistry (Kimel 1953) or isoprene (Leets et al.<br />

1957).<br />

Ex-NCS by absorption: legislation: US—artificial, EU—nature identical. Assay 99%.<br />

The essential oil is mixed with water, buffered to mildly alkaline conditions and a solution<br />

of sodium metabisulphite added, then thoroughly mixed by prolonged stirring. The citral<br />

dissolves in this aqueous phase and when stirring is stopped, the nonwater miscible phase<br />

floats to the surface and is separated. The aqueous phase is then neutralized with acid, and<br />

the citral separates as a floating layer. It is then washed with water and separated. It is then<br />

dried under vacuum or by filtering through a bed of anhydrous sodium sulfate.<br />

Ex-NCS by distillation: legislation: US—natural, EU—natural. Assay 92, 95, and 96%.<br />

Depending on the NCS source, some impurities such as geraniol are codistilled because<br />

their boiling points are close to that of citral. The current commercial preference is to use<br />

Litsea cubeba oil, with its lower geraniol content than lemongrass oil, as a higher assay<br />

citral can be obtained. However, some flavorists still prefer the flavor of citral ex lemongrass<br />

in lemonade applications, as the flavor is said to be closer to that of lemons. As can<br />

be seen, the production method affects the legislative status of the citral.<br />

9.5 analysIs methods<br />

GLC. The experienced chromatographer will often carefully examine GLC traces of the assay of<br />

citral using the FCC method (Food Chemical Codex 1996) on polymethylsiloxane phase columns,<br />

to look for the impurities that elute close to the citral peaks. The ISO standard method is to use an<br />

internal standard; it quotes a minimum 70% for the citral content of Litsea cubeba oil. On these<br />

nonpolar phases, the components elute in the order of their boiling points, as there is no interaction<br />

of the component with the stationary phase coating on the interior of the column.<br />

So, peaks that elute close to the citral isomers have similar boiling points and are particularly<br />

difficult to fractionate from citral, even under high vacuum on high theoretical plate Sulzer-packed<br />

fractionation columns. If these closely eluting peaks are present, it is difficult to achieve the higher<br />

assay, premium citral grades by distillation.<br />

GC-MS (Adams 1995) or GLC with relative retention indices on polarity-calibrated columns<br />

(RSC 1997) can be used to identify these components if necessary.<br />

Attempts to analyze citral using polar phase columns such as polyethylene glycol result in leading<br />

peaks and difficult quantification, as well as incomplete resolution of components. This is due to<br />

interaction of the aldehyde components with the alcohol radicals of the glycol phase. These effects<br />

can be overcome by reducing the citral to the corresponding alcohols with sodium borohydride<br />

before analysis (Jones et al. 1977). However, the widespread commercial acceptance of the FCC and<br />

ISO methods on nonpolar columns has made this effective but longer method all but obsolete.<br />

Titration. Assay of citral by titration of the aldehydes with standardized volumetric solutions<br />

gives good precision and accuracy, unless the oil under test contains citronellal or other aldehydes.


Citral from Lemongrass and Other Natural Sources 231<br />

Absorption. A simple method of analysis, particularly suited to field-testing of citral-containing<br />

oils, is absorption into alkaline bisulfite solution (see manufacturing methods) in a cassia flask. This<br />

is usually a 200 mL conical flask with a long, slim, accurately graduated 10 mL by 0.1 ml neck.<br />

This flask can also be used for the determination of cinnamaldehyde in cassia oil (hence the name<br />

of the flask) and eugenol in clove, bay, and other oils containing it, by absorption into strong alkali.<br />

For citral determination, the flask is filled with alkaline bisulfite solution to the base of the flask<br />

neck zero line; 10.0 mL of the test citral oil is added, and the flask capped (usually with the technician’s<br />

thumb!) and vigorously shaken for several minutes, while the technician’s other hand supports<br />

the main bulb of the flask. The flask is then set aside to allow the phases to separate, and any<br />

unabsorbed oil components float to the top of the flask. The meniscus of this oil layer, which usually<br />

consists of terpene hydrocarbons, is simply read from the scale, subtracted from 10.0 mL of the<br />

test sample, and multiplied by 10 to get the percentage of absorbed citral. So, under the conditions<br />

described, if 10.0 mL of oil leaves 2.5 mL after absorption:<br />

10.0 (minus) 2.5 mL = 7.5 × 10 = 75% citral in the test oil<br />

9.6 toxIcology<br />

In terms of its PBT properties, citral has been rated as follows:<br />

Persistence = low; bioaccumulation = low; toxicity = medium<br />

The tests that are used for classification and labeling (C&L) are described as endpoints at which an effect<br />

is calculated. A list of these is shown in Table 9.4, with the results for citral given where available.<br />

QSAR. As can be seen from Table 9.4, there are many data gaps for the endpoints for citral, as<br />

they are not yet in the public domain. This presents a problem for the consortia registrants under the<br />

REACH regulations. A potential solution is to use the principles of QSAR to predict or supplement<br />

insufficient test data with predictions based on the structural relationships of other molecules, if<br />

they are available.<br />

Sometimes called SAR or SARA, this can be useful for the prediction of the level of concern<br />

regarding a material or even a mixture, with commercial computer programs available to aid in the<br />

prediction of toxicological concerns regarding a particular chemical structure (Lawrence et al. 2007).<br />

To quote from the draft regulations—<br />

… to facilitate the considerations of a (Q)SAR model for regulatory purposes, it should be associated<br />

with the following information:<br />

1. A defined endpoint<br />

2. An unambiguous algorithm<br />

3. A defined domain of applicability<br />

4. Appropriate measures of goodness-of-fit, robustness, and predictivity<br />

5. A mechanistic interpretation, if possible<br />

9.6.1 rifM<br />

Citral. According to RIFM, the Fragrance Structure—Activity Group for citral is “Aldehydes,<br />

branch chain, unsaturated,” and they have summarized the sensitization potency of citral. These are<br />

shown in Table 9.5.<br />

The potency classification of citral is as a weak sensitizer.


232 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 9.4<br />

endpoints in the QmrF with Values for citral<br />

Physicochemical effects<br />

Melting point °C = 27 mg/mL (rabbit)<br />

Acute oral toxicity 6800 mg/kg (rat)<br />

Acute dermal toxicity >5000 mg/kg (rabbit)<br />

Skin irritation at 20%, NOEL 4%<br />

Skin corrosion N/A<br />

Acute photo-irritation N/A<br />

Skin sensitization NESIL 1400 µg/cm 2


Citral from Lemongrass and Other Natural Sources 233<br />

table 9.4 (continued)<br />

endpoints in the QmrF with Values for citral<br />

Respiratory sensitization N/A<br />

Photosensitization N/A<br />

Eye irritation/corrosion Irritation at 5%<br />

Mutagenicity E. coli; 0.1 mg/plate, no effect<br />

Photo mutagenicity N/A<br />

Carcinogenicity Strong suppressing activity, IC 50 < 0.00625%<br />

Photocarcinogenicity N/A<br />

Repeated dose toxicity N/A<br />

Developmental toxicity 22 d non-GLP, LOAEL = 60 mg/kg bw/d<br />

Fertility 20 d non-GLP, NOAEL = 0.43 mg/L<br />

Endocrine disruption: receptor binding Inhibited estrogen binding<br />

Endocrine disruption: gene expression N/A<br />

Toxicokinetics: skin penetration 63 min (mice)<br />

Toxicokinetics: ocular penetration N/A<br />

Toxicokinetics: gastrointestinal penetration N/A<br />

Toxicokinetics: blood–brain penetration N/A<br />

Toxicokinetics: placental penetration N/A<br />

Toxicokinetics: blood–testis penetration N/A<br />

Toxicokinetics: blood–lung penetration N/A<br />

Toxicokinetics: metabolism N/A<br />

Toxicokinetics: protein binding N/A<br />

* N/A = Not available.<br />

table 9.5<br />

sensitizing Potency for citral<br />

LLNA weighted mean EC 3 value In µg/cm2 = 1414 (11 studies)<br />

NOEL-HRIPT induction on human skin In µg/cm2 = 1400<br />

NOEL-MAX induction on human skin N/A<br />

LOEL induction on human skin In µg/cm2 = 3876<br />

WOE-NESIL In µg/cm2 = 1400<br />

The REXPAN conclusion is that the review of “the critical data for citral, based on the weight of<br />

evidence, established the No Expected Sensitization Induction Level as 1400 µg/cm 2 . They recommended<br />

the limits for the 11 different product categories, which derive from the application of the<br />

exposure-based quantitative risk assessment approach for fragrance ingredients, which is detailed<br />

in the QRA Expert Group technical dossier of March 15th 2006 (IFRA 2007).<br />

The 11 product categories can be found in Table 9.6 and replace the “wash-off” and “leave on”<br />

categories for consumer products that were used for many years.<br />

Cosmetics Directive EU. The EU 26th Cosmetics Directive, 7th amendment, lists citral as one<br />

of the 16 SCCNFP naturally occurring “alleged skin allergens” (there are 8 more aroma chemicals,<br />

plus oakmoss and treemoss for a total of 26). A paper was presented at the IFEAT conference 2001<br />

that explained these regulations, giving an XL spreadsheet of the maximum levels of these “16”<br />

in fragrance NCS (Moyler 2001). This XL was published by EFFA and distributed to its members<br />

at the same time, to enable the levels of the 16 in fragrance compounds to be calculated from<br />

the amounts added as such, plus the maximum contribution from the NCS, without the need to


234 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

table 9.6<br />

IFra classes of Product types (Qra)<br />

routinely analyze a complex fragrance. This is particularly helpful for SMEs, who then only need<br />

to use the IFRA GC-MS methods (Chaintreau et al. 2003; Leijs et al. 2005) for validation of the<br />

calculated result.<br />

9.7 classIFIcatIon and labelIng<br />

categories<br />

1. Lip products, insect repellents, toys<br />

2. Deodorants and antiperspirants<br />

3. Men’s facial balms and aftershaves, tampons<br />

4. Colognes, hair styling products, body lotions, foot care, strips for “scratch and sniff”<br />

5. Women’s makeup, hand cream, face masks<br />

6. Mouthwash, toothpaste, oral care<br />

7. Toilet paper, intimate wipes, baby wipes<br />

8. Make-up removers, nonaerosol hair styling products, nail care, talc<br />

9. Shampoo and conditioners, liquid soap, body and face cleansers, shaving creams, depilatory,<br />

shower gels, soap, feminine hygiene products, bath products, aerosols<br />

10. Hand-wash products, laundry cleaners, household cleaners, dish-wash cleaners, diaper cleaners,<br />

dry cleaning products, pet shampoos<br />

11. Candles, air fresheners, shoe polish, carpet cleaners, insecticides, toilet blocks, incense, machine<br />

dish wash, plastics, fuels, paints, cat litter, starch sprays, odorized water for steam irons<br />

Litsea cubeba<br />

Toxicology acute toxicity LD 50: oral, rat: >5000 mg/kg; dermal, rabbit: 4800 mg/kg<br />

Local effects LLNA, EC 3 8.4%; weak sensitizer [RIFM]<br />

Hydrocarbon content 15% (BEOA technical comm.)<br />

Kinematic viscosity 5000 mg/kg<br />

Local effects LLNA, EC 3 6.5% positive; weak sensitizer [RIFM]<br />

Hydrocarbon content WI 9 % (BEOA technical comm.)<br />

Kinematic viscosity


Citral from Lemongrass and Other Natural Sources 235<br />

This means that the C&L, for an essential oil that contains citral, is based on that of citral at the<br />

near maximum level (usually the 95 percentile) to be found in the oil, unless the actual test data for<br />

the oil is “robust.” Robust test data are considered to be results based on modern protocols such as<br />

LLNA and, if available, they can be used to overrule the calculated data based on the DPD.<br />

Examples of the C&L for citral, Litsea cubeba oil, and lemongrass oils are as follows:<br />

Citral<br />

Transport—Not Regulated<br />

Evaluation: Sensitizer—Xi; IRRITANT<br />

Conclusion: RISKS: R 38 irritating to skin<br />

R 43 May cause sensitization by skin contact<br />

SAFETY: S(2) Keep out of reach of children (when sold to general public)<br />

S 24/25 Avoid contact with skin and eyes<br />

S 37 Wear suitable gloves<br />

Litsea cubeba<br />

Evaluation: Sensitizer—Xi, Harmful by Aspiration—Xn, Toxic to Environment—N<br />

HARMFUL<br />

Conclusion: Xn+N; R 38–43–51/53-65; S 24/25–37–61-62<br />

The “Xn” label and its associated R 65—S 62 phrases apply because the >10% hydrocarbon<br />

content of this oil creates the potential for the oil to be an aspiration hazard. This means that<br />

if the oil is swallowed by accident and the patient is induced to vomit, the mobile low viscosity<br />

hydrocarbons have the potential to coat the inner surface of the lungs and prevent the entry of<br />

breathed oxygen into the bloodstream, the potential danger being asphyxiation (BEOA 2006).<br />

This “Xn label” overrides the “Xi label” of the citral when both hazards are present.<br />

Lemongrass<br />

Evaluation: Sensitizer—Xi, Toxic to the Environment—N<br />

IRRITANT<br />

Conclusion: Xi+N; R 38-43-51/53; S 24-37-61<br />

The “N” label and its associated R 51/53—S 61 phrases apply to lemongrass and Litsea<br />

cubeba oils because the terpene hydrocarbons that they contain are slow to biodegrade in the<br />

environment and represent a potential hazard.<br />

9.8 saFety data sheets<br />

The data from the C&L is incorporated into safety data sheets (ISO 2001), which also include the<br />

physical properties, safe-handling instructions, and transportation guidelines for the material. These<br />

SDS are the mechanism for the communication of data for safe handling throughout the supply chain<br />

from manufacturer to the user, and a useful start for data gathering under the REACH registration.


236 <strong>Essential</strong> <strong>Oil</strong>-<strong>Bearing</strong> <strong>Grasses</strong>: The Genus Cymbopogon<br />

9.8.1 Ghs<br />

In 1992, UNECE recognized that there were different regulatory standards being applied globally<br />

(EU: HSDS; U.S.: MSDS) and started work in a concerted way to bring these together in what has<br />

become known as the Globally Harmonized System.<br />

It is planned that an eSDS will be made available for professional users, which will contain additional<br />

toxicological and compositional data more detailed than that found on the HSDS.<br />

With the introduction of REACH (in Europe, the GHS is planned to be adopted at the same time),<br />

this has become an urgent matter, and there are technical specialist task forces operating in the U.S.<br />

and Europe that will then come together for the harmonization.<br />

This is to be achieved (simplistically) by enlarging the number of bands within each hazard<br />

group to accommodate the different classifications; for example, toxicity classes are expanded from<br />

the EU = 3 to GHS = 5, spanning the range up to LD 50 of 5000 mg/kg (based on rat oral tests). Some<br />

new categories are added as well, for example, respiratory sensitizers.<br />

There are some countries without their own system that had the option of adopting the GHS<br />

early; Brazil and New Zealand have done so, and Japan followed suit in December 2006.<br />

The rules of the GHS system do not allow for a “pick and mix” approach; either the CHIP or<br />

GHS system can be used until the introduction of the GHS, not some of each.<br />

One of the visible changes is the use of “diamond” pictograms for hazards, some of which are<br />

different from the “square” St. Andrews cross type, used under the CHIP regulations in the EU.<br />

9.8.2 re a C h<br />

Current CHIP GHS<br />

IRRITANT<br />

HARMFUL<br />

The European Parliament is introducing a registration system for chemicals that shifts the emphasis<br />

for the safe use of industrial materials from government to industry. All materials that are<br />

used at more than 1 tonne per annum are required to be registered, a registration that includes the<br />

modes of use. Depending on the volume of use and the hazard assessment, the amount of test data<br />

required to establish safe use increases, with the very high risk materials potentially substituted<br />

with safer alternatives.<br />

The mechanism for registration will be online by using IUCLID 5 [24] data, the latest version of<br />

the database that was first created in 1993 to meet the EU requirements of the ESR [25–28].<br />

9.9 conclusIon<br />

The toxicologic and legislative status of citral from all sources is covered by the EU legislation of<br />

Annex I. This is a listing that is binding under EU law, in recognition of the potentially harm to<br />

human health properties.


Citral from Lemongrass and Other Natural Sources 237<br />

This chapter has reviewed the status with respect to the latest guidelines and uses of citral within<br />

the flavor and fragrance industries.<br />

The sources, production, toxicology, “classification and labeling,” industrial uses, and dose rates<br />

have been covered for citral itself and its natural sources.<br />

Future requirements of the GHS and REACH systems are also briefly explained.<br />

Citral is an invaluable material within the F&F industries, with a long, safe history of use and an<br />

important future, when used under GMP guidelines to protect those who handle it.<br />

acKnoWledgments<br />

I gratefully acknowledge the advice given by Dr. C. Letizia on toxicology and M. Milchard on testing<br />

data.<br />

My thanks to the directors of Fuerst Day Lawson for their support and kind permission to publish<br />

this paper.<br />

reFerences<br />

Adams R. P. 1995. Identification of EO Components by GC/MS, Allured Publishing, Carol Stream, IL.<br />

Arctander S. 1969. Perfume and Flavor Chemicals, Allured Publishing, Carol Stream, IL.<br />

Arctander S. 1962. Perfume and Flavors of Natural Origin, Allured Publishing, Carol Stream, IL.<br />

BEOA 2006. CHIP List and Handbook BEOA for members (Aug.).<br />

Chaintreau, A. et al. 2003. J. Agric. Food Chem 51: 6398–6403.<br />

EFFA Code of Practice August 2008.<br />

Food Chemical Codex 1996. National Academy Press, Washington DC ISBN 0-309-05394-3 IV July.<br />

Food Cosmetics Toxicology 1979. 17(3): 259–266.<br />

IFRA Standard 43rd Draft 2009.<br />

ISO Standard 3214. 2000. Litsea cubeba oil.<br />

ISO Standard 11041. 2001. Format for Safety Data Sheets.<br />

ISO Standard 3217. 1974. Lemongrass oil West Indian.<br />

ISO Standard 4718. 2004. Lemongrass oil East Indian.<br />

Jones R. A., Neale M. E., Ridlington J. 1977. J. Chromatography 130: 368.<br />

Kimel S. 1953. US patent 2,661,368 Hoffmann-LaRoche.<br />

Lawrence B. M., Hayes J. R., Stavanja M. S. 2007. Biological and Toxicological Properties in Mint—The<br />

Genus Mentha, CRC Press, Boca Raton, FL.<br />

Leets et al. 1957. J. Gen. Chem. USSR 27 1584; 1959. Chem. Abst. 53: 4336e.<br />

Leijs H. et al. 2005. J. Agric. Food Chem 53: 5487–5491.<br />

Moyler D. A. November 2001. On behalf of EFFA, Proceedings IFEAT Conf. Argentina.<br />

Ohloff G. 1993. Scent and Fragrances, Springer-Verlag. ISBN 0-387-57108-6.<br />

RIFM-FEMA. 2009. Database monograph 116.<br />

RSC essential oils committee. 2009. Perfumer & Flavorist to be published.<br />

RSC essential oils committee. 1997. The Analyst (London) 122: 1167–1174.<br />

other sources<br />

http://ecb.jrc.it/iuclid5/ 2006.<br />

http://ec.europa.eu/enterprise/reach/prep_guidance_en.htm 2006.<br />

http://ecb.jrc.it/REACH.htm 2006.<br />

http://europa.eu.int/comm/environment/chemicals/index.htm 2006.<br />

http://europa.eu.int/comm/enterprise/chemicals/index.htm 2006.<br />

http://www.reachready.co.uk 2007.


Index<br />

a<br />

Acyclic monoterpene hydrocarbons, 35<br />

Acyclic oxygenated monoterpenes, 36–40<br />

Afghanistan, trade, 156<br />

Agar diffusion method, 170<br />

Aging process, potential to control, 92<br />

Alloaromadendrene, 57<br />

Alpha tocopherol, 179<br />

Analytical methods, 195–221<br />

chemical methods, 196–197, 204–205<br />

chromatographic, spectroscopic techniques, combined,<br />

206–209<br />

chromatographic techniques, 197–198, 205–206<br />

gas chromatography, 197–198<br />

thin-layer chromatography, 197<br />

Cymbopogon giganteus, 212–215<br />

enantiomeric differentiation by chiral GC, 210–211<br />

fractional distillation, 196–197<br />

hyphenated techniques, 198–203<br />

13 C NMR, 200–201<br />

enantiomeric differentiation, 201–202<br />

GC-FTIR, 200<br />

GC-MS, 198–199<br />

GC-MS combined with GC(RI), 198–199<br />

GC-MS-FTIR, 200<br />

GC-MS-MS, 199–200<br />

GcxGC-MS, 199<br />

HPLC-GC-MS, 200<br />

HPLC- 1 H NMR, 200<br />

HPLC-MS, 200<br />

two-step procedure, 202–203<br />

mass spectrometry with gas chromatography, analysis<br />

by, 206<br />

various techniques, combined analysis by, 211–212<br />

Anatomy, Cymbopogon genus, 1–24<br />

Cymbopogon citratus leaf ultrastructure, 7<br />

generic characters, 5–6<br />

leaf anatomy, 5–7<br />

rubisco immunolocalization, 7<br />

Angola, trade, 156<br />

Antimicrobia/antioxidant activities, 167–183<br />

agar diffusion method, 170<br />

alpha tocopherol, 179<br />

citral, 179<br />

citronellal, 179<br />

citronellol, 179<br />

Cymbopogon citratus, 179<br />

Cymbopogon nardus, 179<br />

essential oil-solubilizing agents, 168<br />

factors affecting antimicrobial activity, 168–177<br />

factors affecting antioxidant activity, 178–180<br />

geraniol, 179<br />

Gram-negative bacteria, 169<br />

Gram-positive bacteria, 169<br />

limonene, 179<br />

media, 178–179<br />

methods used for antimicrobial assessment, 170–177<br />

mould, 169<br />

oil components, activity, correlation between, 169–170<br />

oxidation conditions, 179–180<br />

oxidizable substrate, end-product evaluation, 178<br />

serial broth dilution method, 170<br />

solubilizing agents, 168–169<br />

testing antioxidant activity, 180<br />

carotene bleaching assay, 180<br />

diphenyl-1-perylhydrazyl free radicals scavenging<br />

test, 180<br />

thiobarbituric acid reactive substance assay, 180<br />

type of organism, 169<br />

vapor contact assay, 170–177<br />

yeast, 169<br />

Argentina, trade, 164<br />

Aromadendrene, 57<br />

Australia, trade, 154, 156, 161, 164<br />

Austria, trade, 154, 156, 161<br />

b<br />

Bahrain, trade, 156, 164<br />

Bangalore, 32–33<br />

Bangladesh, trade, 156<br />

Belgium, 31<br />

trade, 156, 161<br />

Bergamot oil<br />

distilled, 228<br />

expressed, 228<br />

Bergamotene, 58<br />

Bhutan, trade, 160<br />

Bicyclic monoterpene hydrocarbons, 42–43<br />

Bicyclic oxygenated monoterpenes, 54–56<br />

Biochemistry, Cymbopogon genus, 1–24<br />

characterization of photosynthetic variant, 8–9<br />

CO 2 assimilation, 10<br />

NADP + inhibition of NADP-MDH, 9<br />

pH and temperature dependence of NADPH-MDH,<br />

NADP-ME, 9–10<br />

Biogenesis, essential oil from genus Cymbopogon, 26–27<br />

Biological activities, 84–95<br />

Biology, Cymbopogon genus, 1–24<br />

Biosynthesis, terpenes in Cymbopogon species, 79–84<br />

Biosynthesis of, terpenes in Cymbopogon species, 79–84<br />

Biotechnological studies, 137–142<br />

Bisabolene, 58<br />

Bisabolol, 59<br />

Bitter orange oil, 228<br />

Blood coagulation activity, 189<br />

Blood sugar, lowering of, 91<br />

Borneol, 54<br />

Bornyl acetate, 54<br />

Bosnia, trade, 161<br />

Botany, Cymbopogon genus, 1–24<br />

239


240 Index<br />

Brazil, trade, 154, 161<br />

Bulgaria, trade, 156<br />

Butenol, 59<br />

c<br />

Cadinene, 60–61<br />

Cadinol, 61<br />

Calamenene, 60<br />

Cambodia, trade, 156<br />

Camphene, 43<br />

Camphor, 54<br />

Canada, trade, 154, 156, 161, 164<br />

Cardamom oil, 228<br />

Carene, 42<br />

Carotene bleaching assay, 180<br />

Carvacrol, 44<br />

Carveyl acetate, 45<br />

Carvone, 45<br />

Carvotanacetone, 45<br />

Caryophyllene, 61<br />

Caryophyllene alcohol, 62<br />

Caryophyllene oxide, 62<br />

Ceylon, 28, 32, 228<br />

Chamigrene, 62<br />

Characterization of photosynthetic variant, 8–9<br />

Chemical analysis methods, 196–197<br />

analysis by, 204–205<br />

Chemistry, 26–27, 68–95<br />

Chile, trade, 156<br />

China, 228<br />

trade, 154, 156, 161<br />

Chinese Taipei, trade, 156<br />

Chromatographic, spectroscopic analysis techniques,<br />

combined analysis by, 206–209<br />

Chromatographic analysis techniques, 197–198<br />

analysis by, 205–206<br />

gas chromatography, 197–198<br />

thin-layer chromatography, 197<br />

Cineole, 46<br />

Cis-carveol, 44<br />

Cis-isopiperitenol, 47<br />

Cis-ocimene, 35<br />

Cis-p-menth-1(7),8-dien-2-ol, 49<br />

Cis-piperitenol, 51<br />

Cis-piperitol, 52<br />

Citral, 179, 223–237<br />

analysis methods, 230–231<br />

classification, 234–235<br />

dose rate, 227<br />

IFRA classes, product types, 234<br />

labeling, 234–235<br />

lemongrass oils, 229–230<br />

Litsea cubeba oil, 227–229<br />

manufacturing methods, 230<br />

natural citral sources, 228<br />

natural sources, 227–230<br />

RIFM, 231–234<br />

safety data sheets, 235<br />

toxicology, 231–234<br />

uses, 227<br />

Citral-a(Z), 36<br />

Citral-b(E), 36<br />

Citronella oil, 71–73, 151–153, 156–158, 228<br />

Citronellal, 37, 179<br />

Citronellol, 37, 179<br />

Citronellyl acetate, 37<br />

Citronellyl butyrate, 37<br />

CO 2 assimilation, 10<br />

Cochin, 28<br />

Colombia, trade, 156<br />

Combined analysis by various techniques, 211–212<br />

Congo, trade, 33, 156<br />

Costa Rica, trade, 156<br />

Cubebene, 63<br />

Cuparene, 63<br />

Cymbopogon, aromatic genus, 136–137<br />

Cymbopogon asmastonii, 79<br />

Cymbopogon caesius, 33, 75<br />

Cymbopogon citratus, 30–31, 69–71, 179<br />

Cymbopogon citratus leaf ultrastructure, 7<br />

Cymbopogon coloratus, 33, 75<br />

Cymbopogon confertiflorus, 75<br />

Cymbopogon densiflorus, 33, 76<br />

Cymbopogon distans, 33–34, 76<br />

Cymbopogon flexuosus, 16–17, 68–69<br />

Cymbopogon genus, 11–13<br />

analysis, 203–215<br />

analytical methods, 195–221<br />

anatomy, 4–7<br />

Cymbopogon citratus leaf ultrastructure, 7<br />

generic characters, 5–6<br />

leaf anatomy, 5–7<br />

rubisco immunolocalization, 7<br />

antimicrobial activities, in vitro, 167–183<br />

antioxidant activity, in vitro, 178–180<br />

aromaticity, 136–137<br />

biogenesis of essential oil from, 26–27<br />

biosynthesis of terpenes in, 79–84<br />

botany, 1–24<br />

chemistry, 26–27<br />

combined analysis, 211–212<br />

essential oil chemistry, 25–106<br />

harvest, 107–133<br />

physicochemical characteristics of essential oils,<br />

26–67<br />

postharvest management, 107–133<br />

trade in, 151–165<br />

uses of, 68–95<br />

biochemistry, 7–10<br />

characterization of photosynthetic variant, 8–9<br />

CO 2 assimilation, 10<br />

NADP + inhibition of NADP-MDH, 9<br />

pH and temperature dependence of NADPH-MDH,<br />

NADP-ME, 9–10<br />

biochemistrygon genus, stomatal conductance, 10<br />

molecular biology, 11–13<br />

randomly amplified polymorphic DNA markers,<br />

11–13<br />

simple sequence repeat markers, 13<br />

physiology/ecophysiology, 14–20<br />

Cymbopogon flexuosus, 16–17<br />

Cymbopogon martinii, 14–16<br />

Cymbopogon winterianus, 17–18<br />

Cymbopogon giganteus, 79, 212–215<br />

Cymbopogon goeringii, 79<br />

Cymbopogon jwarancusa, 29, 74–75<br />

Cymbopogon khasianus, 77


Index 241<br />

Cymbopogon ladakhenis, 77<br />

Cymbopogon mardus, 127–128<br />

distillation, 128<br />

harvesting, 127–128<br />

uses, 127<br />

yield, 128<br />

Cymbopogon martinii, 14–16, 29–30, 73–74, 117–121<br />

distillation, 120<br />

harvesting, 118–119<br />

oil content, yield, 120–121<br />

oil storage, 120<br />

standard specifications, 121<br />

storage, 119<br />

uses, 118<br />

yield, 119–120<br />

Cymbopogon microstachys, 77<br />

Cymbopogon nardus, 32, 72–73, 179<br />

Cymbopogon nervatus, 34, 77<br />

Cymbopogon oils, 151–165<br />

citronella oil, 151–153, 156–158<br />

demand, 152–153<br />

export, import, 153<br />

lemongrass oil, 153–162<br />

palmarosa oil, 160–164<br />

production, 159–160, 163–164<br />

world demand, 153<br />

world production, demand, 159<br />

Cymbopogon olivieri, 77<br />

Cymbopogon parkeri, 78<br />

Cymbopogon pendulus, 31, 71<br />

Cymbopogon polyneuros, 78<br />

Cymbopogon procerus, 78<br />

Cymbopogon rectus, 78<br />

Cymbopogon schoenanthus, 33, 75<br />

Cymbopogon sennarensis, 78<br />

Cymbopogon stracheyi, 78<br />

Cymbopogon tortilis, 78<br />

Cymbopogon travancorenis, 78<br />

Cymbopogon winterianus, 17–18, 31–32, 71–72, 121–127<br />

distillation procedure, 126<br />

harvesting, 122–126<br />

hay storage, 126<br />

standard specifications, 126–127<br />

yield, 126<br />

Czech Republic, trade, 156<br />

d<br />

Denmark, trade, 156, 161<br />

Dihdrocarveol, 46<br />

Dihydro-alpha-copaene, 63<br />

Diphenyl-1-perylhydrazyl free radicals scavenging test,<br />

180<br />

Djibouti, trade, 156<br />

DPPH test. See Diphenyl-1-perylhydrazyl free radicals<br />

scavenging test; diphenyl-1-perylhydrazyl free<br />

radicals scavenging test<br />

e<br />

Ecophysiology, Cymbopogon genus, 14–20<br />

Cymbopogon flexuosus, 16–17<br />

Cymbopogon martinii, 14–16<br />

Cymbopogon winterianus, 17–18<br />

Ecuador, trade, 161<br />

Edema, activity to reduce, 92<br />

Egypt, trade, 156, 161<br />

Elemene, 64<br />

Enantiomeric differentiation by chiral GC, 210–211<br />

Epoxy-3,7-dimethyl-1,6-octadiene, 46<br />

<strong>Essential</strong> oil-solubilizing agents, 168<br />

Ethiopia, trade, 156<br />

Eucarvone, 45<br />

Eudesmol, 64<br />

Euglobulin, 186–187<br />

in vivo fibrinolytic activity, 189<br />

F<br />

Farnesene, 65<br />

Fenchone, 54<br />

Fibrin plate method, 186–187<br />

Fiji, trade, 156, 161<br />

Finland, trade, 156<br />

Fractional distillation, 196–197<br />

France, trade, 154, 156, 160–161, 164<br />

French South and Antarctic Territories, trade, 161<br />

Future developments, 135–149<br />

Cymbopogon, aromatic genus, 136–137<br />

biotechnological studies, 137–142<br />

future options, 142–144<br />

metabolic engineering studies, 141–142<br />

phylogenetic studies, 139–140<br />

tissue culture studies, 137–139<br />

g<br />

Gambia, trade, 156<br />

Gas chromatography, mass spectrometry with, analysis<br />

by, 206<br />

Generic characters, Cymbopogon genus, 5–6<br />

Geraniol, 38, 179<br />

Geranium oil, 228<br />

Geranyl acetate, 38<br />

Geranyl formate, 38<br />

Germacrene D, 66<br />

Germany, trade, 154, 156, 160–161, 164<br />

Ghana, trade, 156, 161<br />

Ginger oil, 228<br />

Gingergrass, 73–74<br />

Gram-negative bacteria, 169<br />

Gram-positive bacteria, 169<br />

Grapefruit oil, 228<br />

Greece, trade, 156, 161<br />

Guatemala, trade, 161<br />

Guyana, trade, 157<br />

h<br />

Haldwani, 31–32<br />

Harvest/postharvest management, Cymbopogons, 107–133<br />

citronella, 121–127<br />

distillation procedure, 126<br />

harvesting, 122–126<br />

hay storage, 126<br />

standard specifications, 126–127<br />

yield, 126<br />

jamrosa, 127–128


242 Index<br />

distillation, 128<br />

harvesting, 127–128<br />

uses, 127<br />

yield, 128<br />

lemongrass, 109–117<br />

condensers, 114<br />

distillation, 113–115<br />

drying, 113<br />

harvesting, 109–112<br />

oil separators, 114–115<br />

postharvest management, 112–116<br />

predistillation handling, 112–113<br />

purification of oil, 116<br />

quality analysis, 116–117<br />

steam boiler, 114<br />

stills, 114<br />

storage, 113<br />

storage and packing of oil, 116<br />

supercritical fluid extraction, 115–116<br />

treatment of oil prior to storage, 116<br />

yield, 112–116<br />

palmarosa, 117–121<br />

distillation, 120<br />

harvesting, 118–119<br />

oil content, yield, 120–121<br />

oil storage, 120<br />

standard specifications, 121<br />

storage, 119<br />

uses, 118<br />

yield, 119–120<br />

Hazara, 29<br />

Himachalene, 66<br />

Hong Kong, trade, 154, 157, 161, 164<br />

Hormones, activation of, 91<br />

Humulene, 66–67<br />

Hungary, trade, 154, 157<br />

Hyderabad, 30<br />

Hyphenated analysis techniques, 198–203<br />

13 C NMR, 200–201<br />

enantiomeric differentiation, 201–202<br />

GC-FTIR, 200<br />

GC-MS, 198–199<br />

GC-MS combined with GC(RI), 198–199<br />

GC-MS-FTIR, 200<br />

GC-MS-MS, 199–200<br />

GcxGC-MS, 199<br />

HPLC-GC-MS, 200<br />

HPLC- 1 H NMR, 200<br />

HPLC-MS, 200<br />

two-step procedure, 202–203<br />

I<br />

Iceland, trade, 157<br />

Immunolocalization, rubisco, 7<br />

In vitro antimicrobial, antioxidant activities, 167–183<br />

agar diffusion method, 170<br />

alpha tocopherol, 179<br />

citral, 179<br />

citronellal, 179<br />

citronellol, 179<br />

Cymbopogon citratus, 179<br />

Cymbopogon nardus, 179<br />

essential oil-solubilizing agents, 168<br />

factors affecting antimicrobial activity, 168–177<br />

factors affecting antioxidant activity, 178–180<br />

geraniol, 179<br />

gram-negative bacteria, 169<br />

gram-positive bacteria, 169<br />

limonene, 179<br />

media, 178–179<br />

methods used for antimicrobial assessment, 170–177<br />

mould, 169<br />

oil components, activity, correlation between, 169–170<br />

oxidation conditions, 179–180<br />

oxidizable substrate, end-product evaluation, 178<br />

serial broth dilution method, 170<br />

solubilizing agents, 168–169<br />

testing antioxidant activity, 180<br />

carotene bleaching assay, 180<br />

diphenyl-1-perylhydrazyl free radicals scavenging<br />

test, 180<br />

thiobarbituric acid reactive substance assay, 180<br />

type of organism, 169<br />

vapor contact assay, 170–177<br />

yeast, 169<br />

India, 30–31<br />

Indian Standards Institution, 29–30<br />

Indonesia, trade, 154, 157, 160–161<br />

The International Plant Names Index, 2–3<br />

Iran, trade, 157<br />

Ireland, trade, 154, 157, 164<br />

ISI. See Indian Standards Institution<br />

Isoborneol, 55<br />

Isopiperitenone, 47<br />

Isopulegol, 47<br />

Israel, trade, 154, 157, 161<br />

Italy, trade, 154, 157, 160–161<br />

J<br />

Jamaica, trade, 157<br />

Jammu, 30<br />

Japan, trade, 154, 157, 161, 164<br />

Java, 30–32<br />

Java, 228<br />

Jordan, trade, 157<br />

Jorhat, 32<br />

K<br />

Kenya, trade, 157, 161<br />

Kerala, 28<br />

Kordofan Sudan, 34<br />

Korea, trade, 157, 161<br />

Kumaon, 28, 32<br />

Kuwait, trade, 157<br />

l<br />

Latvia, trade, 157<br />

Lavandulol, 38<br />

Leaf anatomy, Cymbopogon genus, 5–7<br />

Lemon oil, 228


Index 243<br />

Lemongrass, 68–71, 109–117, 153–162, 228<br />

citral from, 223–237<br />

analysis methods, 230–231<br />

classification, 234–235<br />

dose rate, 227<br />

IFRA classes, product types, 234<br />

labeling, 234–235<br />

lemongrass oils, 229–230<br />

Litsea cubeba oil, 227–229<br />

manufacturing methods, 230<br />

natural citral sources, 228<br />

natural sources, 227–230<br />

RIFM, 231–234<br />

safety data sheets, 235<br />

toxicology, 231–234<br />

uses, 227<br />

distillation, 113–115<br />

condensers, 114<br />

oil separators, 114–115<br />

steam boiler, 114<br />

stills, 114<br />

drying, 113<br />

harvesting, 109–112<br />

postharvest management, 112–116<br />

predistillation handling, 112–113<br />

quality analysis, 116–117<br />

standard specifications, 117<br />

storage, 113<br />

supercritical fluid extraction, 115–116<br />

treatment of oil prior to storage, 116<br />

purification of oil, 116<br />

storage and packing of oil, 116<br />

yield, 112–116<br />

Leukemia, activity against, 91<br />

Lime oil, 228<br />

Limonene, 41, 179<br />

Limonene oxide, 48<br />

Linalool, 39<br />

Linalyl acetate, 39<br />

Lippia citriodora, 228<br />

List of Cymbopogon species, 2–3<br />

Litsea cubeba oil, 228<br />

Longifolene, 67<br />

Lowering blood sugar, 91<br />

Lucknow, 28, 30, 32<br />

m<br />

Madras, 30<br />

Malabar District, 33<br />

Malaysia, trade, 154, 157<br />

Maldives, trade, 157, 161<br />

Male hormones, activation of, 91<br />

Malignancy, activity against, 91<br />

Mass spectrometry with gas chromatography, analysis by,<br />

206<br />

Mauritius, trade, 157, 162<br />

Menthene-3-one, 48<br />

Menthol, 48<br />

Menthone, 49<br />

Menthyl acetate, 48<br />

Metabolic engineering studies, 141–142<br />

Methods used for antimicrobial assessment, 170–177<br />

Methyl thymyl ether, 50<br />

Mexico, trade, 157, 162, 164<br />

Microbes, activity against, 93–95<br />

Molecular biology, Cymbopogon genus, 1–24<br />

randomly amplified polymorphic DNA markers, 11–13<br />

simple sequence repeat markers, 13<br />

Monoterpenes in cymbogon essential oils, 35–56<br />

Mosquitoes, potential to repel, kill larvae, 91–92<br />

Mould, 169<br />

Mozambique, trade, 157<br />

Murrolene, 67<br />

Myanmar, trade, 157<br />

Myrcene, 41<br />

Myrtenol, 55<br />

n<br />

NADP + inhibition of NADP-MDH, 9<br />

Nainital, 32, 34<br />

Nepal, trade, 154, 157, 162<br />

Nerol, 39, 228<br />

Neryl acetate, 39<br />

Netherlands, trade, 155, 157, 162, 164<br />

New Caledonia, trade, 157<br />

New Zealand, trade, 157, 162, 164<br />

Nigeria, trade, 157<br />

Nilgiri Hills, 31<br />

o<br />

O-cymene, 40<br />

Odakali, 28, 31<br />

<strong>Oil</strong> components, activity, correlation between, 169–170<br />

Oman, trade, 157, 162<br />

Orange bitter oil, 228<br />

Orange sweet oil, 228<br />

Oxidation conditions, 179–180<br />

Oxidizable substrate, end-product evaluation, 178<br />

Oxygenated monoterpenes, 44–53<br />

P<br />

P-cymene, 40<br />

P-menth-2,8-dien-1-ol, 50<br />

P-menth-2-en-1-ol, 49<br />

P-menth-8-en-1-ol, 49<br />

Pain relievers, 89–91<br />

Pakistan, trade, 157<br />

Palmarosa, 160–164<br />

and gingergrass oils, 73–74<br />

harvesting, 118–119<br />

oil content, yield, 120–121<br />

oil storage, 120<br />

standard specifications, 121<br />

storage, 119<br />

uses, 118<br />

yield, 119–120<br />

Panama Republic, trade, 157<br />

Pantnagar, 28, 32<br />

Paraguay, trade, 155, 157<br />

Perillaldehyde, 50


244 Index<br />

Perillene, 51<br />

Perillyl alcohol, 51<br />

Peru, trade, 157<br />

Pests, activity against, 93<br />

Petitgrain bergamot oil, 228<br />

Petitgrain bitter orange oil, 228<br />

Petitgrain lemon oil, 228<br />

Petitgrain mandarin oil, 228<br />

Phellandrene, 35<br />

Philippines, trade, 157, 162<br />

Photosynthetic variant characterization, 8–9<br />

Phylogenetic studies, 139–140<br />

Physicochemical characteristics, 26–67<br />

Physiochemical properties, 28–34<br />

Physiology, Cymbopogon genus, 1–24<br />

Cymbopogon flexuosus, 16–17<br />

Cymbopogon martinii, 14–16<br />

Cymbopogon winterianus, 17–18<br />

Pinene, 43<br />

Piperitenone, 51<br />

Piperitone, 52<br />

Piperitone oxide, 52<br />

Platelet aggregation test, 189<br />

Poland, trade, 155, 158<br />

Postharvest management, Cymbopogons. See Harvest/<br />

postharvest management<br />

Pulegone, 53<br />

r<br />

Randomly amplified polymorphic DNA markers, 11–13<br />

RAPD markers. See Randomly amplified polymorphic<br />

DNA markers<br />

Reunion, trade, 158<br />

Romania, trade, 158<br />

Rose oil, 228<br />

Rubisco immunolocalization, 7<br />

Russia, trade, 158<br />

Rwanda, trade, 158<br />

s<br />

Sabienene, 43<br />

Saudi Arabia, trade, 158<br />

Serial broth dilution method, 170<br />

Sesquiterpenes in Cymbopogon oils, 57<br />

Simple sequence repeat markers, 13<br />

Sind, 29<br />

Singapore, trade, 155, 158, 160, 162, 164<br />

6,7-epoxy-3,7-dimethyl-1,3-octadiene, 46<br />

6-methylhept-5-en-2-one, 40<br />

Slovenia, trade, 155, 162<br />

South Africa, trade, 158, 162, 164<br />

Spain, trade, 155, 158, 160, 162, 164<br />

Sri Lanka, trade, 155, 158, 162<br />

SSRs. See Simple sequence repeat markers<br />

Standard fibrin plate method, 186<br />

Subspecies, Cymbopogon, 2–3<br />

Subvarieties, Cymbopogon, 2–3<br />

Sudan, trade, 158<br />

Sugar, blood, lowering of, 91<br />

Surinam, trade, 162<br />

Swaziland, trade, 155<br />

Switzerland, trade, 155, 158, 162, 164<br />

Syria, trade, 158<br />

t<br />

Taiwan, trade, 158, 162, 164<br />

Tanzania, trade, 155, 158, 162, 164<br />

TBARS assay. See Thiobarbituric acid reactive substance<br />

assay<br />

Temperature dependence, NADPH-MDH, NADP-ME,<br />

9–10<br />

Temperature dependence of NADPH-MDH, NADP-ME,<br />

9–10<br />

Terpenes in Cymbopogon species, 79–84<br />

Terpin-4-ol, 53<br />

Terpinene, 41–42<br />

Terpineol, 53<br />

Terpinolene, 42<br />

Testing antioxidant activity, 180<br />

carotene bleaching assay, 180<br />

diphenyl-1-perylhydrazyl free radicals scavenging test,<br />

180<br />

thiobarbituric acid reactive substance assay, 180<br />

Thailand, trade, 155, 158, 162, 164<br />

Thiobarbituric acid reactive substance assay, 180<br />

Thrombolysis acceleration, 185–194<br />

blood coagulation activity, 189<br />

euglobulin, 186–187<br />

in vivo fibrinolytic activity, 189<br />

fibrin plate method, 186–187<br />

platelet aggregation test, 189<br />

standard fibrin plate method, 186<br />

tissue plasminogen activator-producing cells, 188<br />

Thujene alcohol, 55<br />

Thujyl alcohol, 55<br />

Tissue culture studies, 137–139<br />

Tissue plasminogen activator-producing cells, 188<br />

Trade, 151–165<br />

Trans-isopiperitenol, 47<br />

Trans-ocimene, 35<br />

Trans-piperitol, 52<br />

Travancore, 29<br />

Tunisia, trade, 158<br />

Turkey, trade, 158<br />

u<br />

UAE, trade, 155, 158, 164<br />

Uganda, trade, 158<br />

U.K., trade, 155, 158, 160, 162<br />

Ultrastructure, Cymbopogon citratus leaf, 7<br />

Uruguay, trade, 155<br />

U.S., trade, 155, 158, 160, 162, 164<br />

Uses of Cymbopogon essential oils, 68–95<br />

V<br />

Vapor contact assay, 170–177<br />

Varieties, Cymbopogon, 2–3<br />

Verbena, 228


Index 245<br />

Verbenone, 56<br />

Vietnam, trade, 155, 162<br />

W<br />

West Bengal, 28, 30, 32<br />

World demand, 152–153<br />

Worms, activity against, 91<br />

y<br />

Yeast, 169<br />

Yemen, trade, 158, 162<br />

Z<br />

Zambia, trade, 158

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!