13.07.2013 Views

Allelochemicals Biologica... - Name

Allelochemicals Biologica... - Name

Allelochemicals Biologica... - Name

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>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases


Disease Management of Fruits and Vegetables<br />

VOLUME 2<br />

Series Editor:<br />

K.G. Mukerji, University of Delhi, Delhi, India


<strong>Allelochemicals</strong>: <strong>Biologica</strong>l<br />

Control of Plant<br />

Pathogens and Diseases<br />

Edited by<br />

INDERJIT<br />

University of Delhi, India<br />

and<br />

K.G. MUKERJI<br />

University of Delhi, India


A C.I.P. Catalogue record for this book is available from the Library of Congress.<br />

ISBN-10 1-4020-4445-3 (HB)<br />

ISBN-13 978-1-4020-4445-8 (HB)<br />

ISBN-10 1-4020-4447-X (e-book)<br />

ISBN-13 978-1-4020-4447-2 (e-book)<br />

Published by Springer,<br />

P.O. Box 17, 3300 AA Dordrecht, The Netherlands.<br />

www.springer.com<br />

Printed on acid-free paper<br />

Cover photo:<br />

Bio Control of powdery mildew pathogen Phyllactinia dalbergae on<br />

Dalbergia sisoo by hyperparasite Cladosporium spongiosum.<br />

(Microphotograph taken by Prof. K.G. Mukerji and Mr. S.K. Das)<br />

All Rights Reserved<br />

© 2006 Springer<br />

No part of this work may be reproduced, stored in a retrieval system, or transmitted<br />

in any form or by any means, electronic, mechanical, photocopying, microfilming,<br />

recording or otherwise, without written permission from the Publisher, with the<br />

exception of any material supplied specifically for the purpose of being entered<br />

and executed on a computer system, for exclusive use by the purchaser of the work.<br />

Printed in the Netherlands.


CONTENT<br />

Preface<br />

1. Discovery and Evaluation of Natural Product-based<br />

Fungicides for Disease Control of Small Fruits<br />

David E. Wedge and Barbara J. Smith<br />

1<br />

2. <strong>Allelochemicals</strong> as Biopesticides for Management of<br />

Plant-parasitic Nematodes<br />

Nancy Kokalis-Burelle and Rodrigo Rodríguez-Kábana<br />

15<br />

3. Allelopathic Organisms and Molecules: Promising<br />

Bioregulators for the Control of Plant Diseases,<br />

Weeds, and Other Pests<br />

Ana Luisa Anaya<br />

31<br />

4. The Impact of Pathogens on Plant Interference<br />

and Allelopathy<br />

Scott W. Mattner<br />

79<br />

5. Allelopathy for Weed Control in Aquatic and<br />

Wetland Systems<br />

Ramanathan Kathiresan, Clifford H. Koger , Krishna N. Reddy<br />

103<br />

6. Bacterial Root Zone Communities, Beneficial Allelopathies<br />

and Plant Disease Control<br />

Antony V. Sturz<br />

123<br />

7. The Role of Allelopathic Bacteria in Weed Management<br />

Robert J. Kremer<br />

143<br />

8. The Allelopathic Potential of Ginsenosides<br />

Mark A. Bernards, Lina F. Yousef, Robert W. Nicol<br />

157<br />

9. Antimicrobial and Nematicidal Substances from the<br />

Root of Chicory (Cichorium intybus)<br />

Hiroyuki Nishimura and Atsushi Satoh<br />

177<br />

10. Disease Resistance in Plants through Mycorrhizal<br />

Fungi Induced <strong>Allelochemicals</strong><br />

Ren-sen Zeng<br />

181<br />

11. <strong>Allelochemicals</strong> from Ageratum conyzoides L. and<br />

Oryza sativa L. and their Effects on Related Pathogens<br />

Chuihua Kong<br />

193<br />

Index 207<br />

List of Contributors 213<br />

v<br />

vii


Preface<br />

<strong>Biologica</strong>l control of plant diseases and plant pathogens is of great significance in<br />

forestry and agriculture. There is great incentive to discover biologically active natural<br />

products from higher plants that are better than synthetic agrochemicals and are much<br />

safer, from health and environmental considerations. The development of natural<br />

products as herbicides, fungicides, and their role in biological control of plant disease<br />

promises to reduce environmental and health hazards. Allelopathic techniques offer<br />

real promise in solving several problems linked with biological control of plant pests.<br />

The allelopathic effect of plants on microorganisms, and microorganisms on<br />

microorganisms is of great environmental and economic significance. This book is<br />

organized around the premise that allelochemicals can be employed for biological<br />

control of plant pathogens and plant diseases. Specifically, this volume focuses on (i)<br />

discovery and development of natural product based fungicides for agriculture, (ii)<br />

direct use of allelochemicals as well as indirect effects through cover crops and organic<br />

amendments for plant parasitic pest control and (iii) application of allelopathy in the<br />

pest management.<br />

In an effort to address above points, contributing authors provided up-to-date<br />

reviews and discussion on allelochemicals-related biological control of plant diseases<br />

and pathogens. Chapters 1 - 3 discuss discovery and development of allelochmicals<br />

and their role in the management of plant diseases. Chapter 4 discusses the effects of<br />

pathogens on the competitiveness and allelopathic ability of their hosts. Chapter 5<br />

highlights the importance of allelopathy for weed control in aquatic ecosystems.<br />

Chapters 6-7 deal with bacterial potential in weed management and plant disease<br />

control. Chapter 8 describes the role of organic compound ginsenosides from<br />

rhizosphere soil and root exudates of american ginseng plant in control of fungal<br />

diseases. Antimicrobial and nematicidal substances from the rhizome of chicory has<br />

been discussed in Chapter 9. The role of allelochemicals induced in mycorrhizal<br />

plants in imparting disease resistance is given in Chapter 10. The last chapter discusses<br />

the biocontrol of plant pathogens and diseases by allelochmicals from Ageratum<br />

conyzoides a weed and rice plants has been highlighted in Chapter 11.<br />

We are grateful to all authors for providing their valuable work to this volume.<br />

The articles are original and some have been written for the first time in any book. We<br />

are indebted to the following referees for their constructive comments and suggestions:<br />

Ana L. Anaya, Mark Bernards, Nancy Kokalis Burelle, Chester L. Foy, John M.<br />

Halbrendt, Robert Kremer, Azim Mallik, Susan Meyer, Reid J. Smeda, Tony Sturz,<br />

David Wedge and Jeff Weidenhamer. The editorial help of Ineke Ravesloot, Publishing<br />

Department, Springer is sincerely appreciated. It is our hope that this book will serve<br />

scientific community well, and equally hope that the book will stimulate young students<br />

to work on biological control of plant pathogens and diseases through natural<br />

allelochemicals.<br />

Inderjit and K.G. Mukerji<br />

October 2005<br />

vii


DAVID E. WEDGE 1 and BARBARA J. SMITH 2<br />

DISCOVERY AND EVALUATION OF NATURAL<br />

PRODUCT-BASED FUNGICIDES FOR DISEASE<br />

CONTROL OF SMALL FRUITS<br />

1<br />

United States Department of Agriculture, Agricultural Research Service,<br />

Natural Products Utilization Research Unit, The Thad Cochran National<br />

Center for Natural Products Research, University of Mississippi, University,<br />

MS 38677, USA; and 2<br />

Small Fruit Research Station, 306 S.<br />

High St., Poplarville, MS 39470, USA<br />

Email : dwedge@olemiss.edu<br />

Abstract. The continuing development of fungicide resistance in plant and human pathogens necessitates the<br />

discovery and development of new fungicides. Discovery and evaluation of natural product fungicides is largely<br />

dependent upon the availability of miniaturized antifungal bioassays. Essentials for natural product bioassays<br />

include sensitivity to microgram quantities, selectivity to determine optimum target pathogens, and adaptability<br />

to complex mixtures. Experimental accuracy and precision must be stable between assays over time. These<br />

assays should be relevant to potential pathogen target sites in the natural infection process of the host and<br />

applicable to the agrochemical industry. Bioassays should take advantage of current high-throughput technology<br />

available to evaluate dose-response relationships, commercial fungicides standards, modes of action, and structure<br />

activity studies. The focus of this chapter is the evaluation of natural product based fungicides for agriculture<br />

and we will provide a review of bioautography prescreens and microtiter assays (secondary assays). Also presented<br />

is more detailed information on newer techniques such as the detached leaf assays for evaluating fungicides<br />

against strawberry anthracnose (Colletotrichum spp.) and field plot trials for gray mold (Botrytis) and anthracnose<br />

control in strawberry.<br />

1. INTRODUCTION<br />

Since the early 1970s, agriculture worldwide has struggled with the evolution of<br />

pathogen resistance to disease control agents. Increased necessity for repeated chemical<br />

applications, development of pesticide cross-resistance, and disease resistance management<br />

strategies have characterized the use of agricultural chemicals to-date. As a<br />

consequence, producers are currently attempting to control agricultural pests with a<br />

decreasing arsenal of effective crop protection chemicals. In addition, the desire for<br />

safer pesticides with less environmental impact has become a major public concern.<br />

Particularly desirable is the discovery of novel pesticidal agents from new chemical<br />

classes that are able to operate using different modes of action and, consequently,<br />

against plant pathogens with resistance to currently used chemistries. In this regard,<br />

evaluating natural products and extracts as a source of new pesticides is one strategy<br />

for the discovery of new chemical moieties that have not previously been created by<br />

synthetic chemists.<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 1–14.<br />

© 2006 Springer. Printed in the Netherlands.<br />

1


2<br />

DAVID E. WEDGE<br />

AND BARBARA J. SMITH<br />

Antibiotics, antineoplastics, herbicides, and insecticides often originate from plant<br />

and microbial chemical defense mechanisms (Wedge and Camper, 2000). Secondary<br />

metabolites, once considered unimportant products, are now thought to mediate chemical<br />

defense mechanisms by providing chemical barriers against animal and microbial<br />

predators (Agrios, 1997; Wedge and Camper, 2000). Plants produce numerous chemicals<br />

for defense and communication, and can elicit their own form of offensive chemical<br />

warfare by targeting the proliferation of pathogens. These chemicals may have<br />

general or specific activity against key target sites in bacteria, fungi, and viruses.<br />

Exploiting the chemical warfare that occurs between plants and their pathogens shows<br />

promise in providing new natural products for new anti-infectives for human, plant<br />

and animal health. The successful development of strobilurin fungicides and spinosad<br />

insecticides has continued the interest in natural products as crop protectants. The<br />

importance and future of natural product agrochemistry is emphasized by the fact<br />

that 21 companies have filed 255 patent applications primarily for use of the strobilurin<br />

class of fungicides (Qo I MET complex 3 inhibitors).<br />

1.1. Direct Acting Defense Chemicals<br />

Since the discovery of the vinca alkaloids in 1963, many of the known antitubulin<br />

agents used in today’s cancer chemotherapy arsenal are products of plant and fungal<br />

secondary metabolism. Since 1991, 16 of 43 new pharmaceuticals were derived from<br />

natural products. In certain therapeutic areas 78% of the antibacterials and 74% of<br />

the anticancer compounds are natural products or have been derived from natural<br />

products (Roughi, 2003). These “natural products” are probably defense chemicals<br />

that target and inhibit cell division in invading pathogens (Wedge and Camper, 2000).<br />

Therefore, it is reasonable to hypothesize that plants and certain fungi can produce<br />

chemicals, such as resveratrol and strobilurin, that act directly in their defense by<br />

inhibiting pathogen proliferation, or indirectly by disrupting chemical signal processes<br />

related to growth and development of pathogens or herbivores (Wedge and<br />

Camper 2000).<br />

1.2. Indirect Acting Defense Chemicals<br />

Plant resistance to pathogens is considered to be systemically induced by some<br />

endogenous signal molecule produced at the infection site that is then translocated to<br />

other parts of the plant (Oku, 1994). The search for, and identification of, the putative<br />

signal molecule(s) is of great interest to many plant scientists because such moieties<br />

have possible uses as “natural product” disease control agents. However, research<br />

indicates that no single compound is involved, but rather a complex signal transduction<br />

pathway, which, in plants, can be mediated by a number of compounds that appears to<br />

influence octadecanoid metabolism. In response to wounding or pathogen attack,<br />

fatty acids of the jasmonate cascade are formed from membrane-bound α-linolenic<br />

acid by lipoxygenase-mediated peroxidation (Vick and Zimmerman, 1984). Analogous<br />

to the prostaglandin cascade in mammals, α-linolenic acid is thought to participate


NATURAL PRODUCT BASED FUNGICIDES 3<br />

in a lipid-based signaling system where jasmonates induce the synthesis of a family of<br />

wound-inducible defensive proteinase inhibitors (Farmer and Ryan, 1992) and low<br />

and high molecular weight phytoalexins such as flavonoids, alkaloids, and terpenoids<br />

(Gundlach et al., 1992; Mueller et al., 1993).<br />

Several plant and bacterial natural products have novel applications as plant<br />

protectants through the induction of systemic acquired resistance (SAR) processes.<br />

Commercial products that appear to induce SAR include Messenger® (EDEN<br />

Biosciences, Inc., Bothell, WA) and the bioprotectant fungicides Serenade®<br />

(AgraQuest, Davis, CA), Sonata® (AgraQuest, Davis, CA), and Milsana® (KHH<br />

BioSci, Inc., Raleigh, NC). Messenger is a harpin protein which switches on natural<br />

plant defenses in response to bacterial leaf spot, wilt, and blight and fungal diseases<br />

such as Botrytis brunch rot, and powdery mildew. Serenade is a microbial-protectant<br />

derived from Bacillus subtilis, with SAR activity that controls Botrytis, powdery and<br />

downy mildews, early blight, and bacterial spot. Sonata is also a microbial-biopesticide<br />

with activity against Botrytis, downy and powdery mildews, rusts, Sclerotinia blight,<br />

and rots. Milsana® is an extract from Reynoutria sachalinensis (giant knotweed)<br />

that induces phytoalexins able to confer resistance to powdery mildew and other<br />

diseases such as by Botrytis. However, elicitors with no innate antifungal activity<br />

will not appear active in most current screening high throughput screening systems.<br />

Many experimental approaches have been used to screen compounds and extracts<br />

from plants and microorganisms in order to discover new antifungal compounds.<br />

The focus of this paper is on laboratory methods and field procedures that we use to<br />

evaluate naturally occurring antifungal compounds produced by plants, pathogens,<br />

and other terrestrial and marine organisms. As part of a program to discover and<br />

develop naturally occurring fungicides, several new in vitro detection systems and a<br />

detached leaf assay were developed to evaluate small amounts of compound. Our<br />

fungicide field protocol used to test potential lead fungicides using commercial<br />

strawberry is also presented.<br />

2.1. Pathogen Production<br />

2. MATERIAL AND METHODS<br />

Isolates of Colletotrichum acutatum, C. fragariae and C. gloeosporioides, Phomopsis<br />

viticola and P. obscurans, Botrytis cinerea and Fusarium oxysporum are maintained<br />

on silica gel at 4-6 °C. The three Colletotrichum species and P. obscurans strain were<br />

isolated from strawberry (Fragaria x ananassa Duchesne). Phomopsis viticola and<br />

B. cinerea were isolated from commercial grape (Vitis vinifera L.) and F. oxysporum<br />

from orchid (Cynoches sp.). Fungal cultures were grown on potato-dextrose agar<br />

(PDA, Difco, Detroit, MI) in 9-cm petri dishes and incubated in a growth chamber at<br />

24 ± 2 °C under cool-white fluorescent lights (55 ± 5 µmols/m 2 /sec) with a 12h photoperiod.


4<br />

Conidia were harvested from 7-10 day-old cultures by flooding plates with 5 mL<br />

of sterile distilled water and softly brushing the colonies with an L-shaped glass rod.<br />

Aqueous conidial suspensions are filtered through sterile Miracloth (Calbiochem-<br />

Novabiochem Corp., La Jolla CA) to remove hyphae. Conidial concentrations were<br />

determined photometrically (Espinel-Ingrof and Kekering, 1991; Wedge and Kuhajek,<br />

1988) from a standard curve based on the absorbance at 625 nm, and suspensions are<br />

then adjusted with sterile distilled water to a concentration of 1.0x10 6 conidia/mL.<br />

Standard conidial concentrations are determined from a standard curve for each<br />

fungal species. Standard turbidity curves were periodically validated using both a<br />

Bechman/Coulter Z1 (Fullerton, CA) particle counter and hemocytometer counts.<br />

Conidial and mycelial growth are evaluated using a Packard Spectra Count<br />

(PerkinElmer Life and Analytical Sciences, Inc., Boston, MA). Conidial growth and<br />

germ tube development were evaluated using an Olympus IX 70 (Olympus Industrial<br />

America, Inc., Melville, New York) inverted microscope and recorded with a Olympus<br />

DP12 digital camera as appropriate for compounds that affect spore germination and<br />

early germ tube development.<br />

2.2. Direct Bioautography<br />

DAVID E. WEDGE<br />

AND BARBARA J. SMITH<br />

A number of bioautography techniques were used as primary screening systems to<br />

detect antifungal compounds. Matrix, one-dimensional, and two-dimensional bioautography<br />

protocols on silica gel thin layer chromatography (TLC) plates and<br />

Colletotrichum spp. as the test organisms were used to identify the antifungal activity<br />

according to published methods (Homan and Fuchs, 1970; Moore, 1996; Wedge and<br />

Nagle, 2000). Matrix bioautography is used to screen large numbers of crude extract<br />

at 20mg/mL. One-dimensional TLC (1D-TLC) and two-dimensional TLC (2D-TLC)<br />

are subsequently used to separate and identify the number of antifungal agents in<br />

extract. Modification of these procedures can be used to visually evaluate natural<br />

chemical defenses in disease resistant and susceptible plant cultivars (Vincent et al.<br />

1999).<br />

A 2D-TLC direct bioautography method was used to evaluate active crude or<br />

partially purified extracts. This protocol utilizes two sequential TLC runs in which<br />

the TLC plates are developed once with a polar solvent, turned 90 o , and then developed<br />

a second time with a non-polar solvent system (Wedge and Nagle, 2000). The<br />

method takes advantage of the resolving power of 2D-TLC to separate chemically<br />

diverse mixtures found in crude extracts. Two-dimensional TLC bioautography is<br />

well suited for resolving extracts containing lipophilic natural products that are difficult<br />

to separate by single elution TLC.<br />

Each plate was subsequently sprayed with a spore suspension (10 5 spores/mL) of<br />

the test fungus and incubated in a moisture chamber for 3 days at 26 ºC with a 12h<br />

photoperiod. Clear zones of fungal growth inhibition on the TLC plate indicate the<br />

Mention of trade names or commercial products in this article is solely for the purpose of providing specific<br />

information and does not imply recommendation or endorsement by the U. S. Department of Agriculture.


NATURAL PRODUCT BASED FUNGICIDES 5<br />

presence of antifungal constituents in each extract. Inhibition of fungal growth was<br />

evaluated 3-4 days after treatment. Antifungal metabolites can be readily located on<br />

the plates by visually observing clear zones where the active compounds inhibit fungal<br />

growth (Tellez, 2000,Vincent et al., 1999). The 2-D TLC method eliminates the<br />

need for the development of large numbers of plates in multiple solvent systems,<br />

reduces the amount of waste solvents for disposal, and substantially reduces the time<br />

required to identify active compounds.<br />

2.3. 96-Well Microbiassay<br />

The quick discovery and evaluation of natural product fungicides is heavily dependent<br />

upon miniaturized antifungal bioassay techniques. A reference method [M27-<br />

A from the National Committee for Clinical Laboratory Standards (NCCLS)] for<br />

broth-dilution antifungal susceptibility testing of yeast was adapted for evaluation of<br />

antifungal compounds against sporulating filamentous fungi (Wedge and Kuhajek,<br />

1998).<br />

This standardized 96-well microtiter plate assay was developed for the detection<br />

of natural product fungicidal agents, and can also be used to evaluate purified antifungal<br />

agents. In our study a 96-well microtiter assay was used to determine sensitivity of C.<br />

acutatum, C. fragariae, C. gloeosporioides, F. oxysporum, B. cinerea, P. obscurans,<br />

and P. viticola to the various antifungal agents in comparison with the commercial<br />

fungicides. Fungicides such as benomyl, azoxystrobin and captan with different<br />

modes of action were used as standards in these assays. Each fungal species was<br />

challenged in a dose-response format so that in the final test, compound concentrations<br />

of 0.3, 3.0, and 30.0 µM were achieved (in duplicate) in the different columns of the<br />

96-well plates.<br />

Fungal growth was evaluated by measuring absorbance of each well at 620 nm at<br />

0, 24, 48, and 72 hr, with the exception of tests with P. obscurans and P. viticola,<br />

where the data are recorded for up to 120 hr. Treatments are repeated so that mean<br />

absorbance values and standard errors can be calculated and are used to evaluate<br />

fungal growth. Differences in spore germination and mycelial growth in each of the<br />

wells in the 96-well plate demonstrate sensitivity to particular concentrations of<br />

compound and can indicate fungistatic or fungicidal effects. The microtiter assay<br />

can also be used to compare the sensitivity of fungal plant pathogens to natural and<br />

synthetic compounds with industry standards (Wedge et al., 2001).<br />

A novel application of the microbioassay was also developed for the discovery of<br />

compounds that inhibit Phytophthora spp. This protocol used the 96-well format for<br />

high-throughput capability and a standardized method for quantification of initial<br />

zoospore concentrations for maximum reproducibility. Zoospore suspensions were<br />

quantifiable between 0.7 and 1.5 × 10 5 zoospores/mL at an absorbance value of 620<br />

nm. Subsequent growth of mycelia was monitored by measuring absorbance (620<br />

nm) at 24-hour intervals for 96 hr. Full- and half-strength preparations of each of<br />

three media (V8 broth, Roswell Park Memorial Institute mycological broth, and mineral<br />

salts medium), and four zoospore concentrations (10, 100, 1000, and 10,000


6<br />

DAVID E. WEDGE<br />

AND BARBARA J. SMITH<br />

zoospores/mL) were evaluated. Both full- and half-strength Roswell Park Memorial<br />

Institute mycological broth were identified as suitable synthetic media for growing P.<br />

nicotianae, and 1000 zoospores/mL was established as the optimum initial concentration<br />

(Kuhajek et al., 2003).<br />

2.4. Detached Leaf Assay for Fungicide Evaluations In Planta<br />

Anthracnose susceptible ‘Chandler’ strawberry plants were grown in 10 x 10 cm<br />

plastic pots in a 1:1 (v/v) mixture of Jiffy-Mix (JPA, West Chicago, IL), and pasteurized<br />

sand in the greenhouse for a minimum of six weeks before inoculation. The<br />

plants were grown under standard conditions of a 16-hr day length and 24 o C temperature.<br />

Growth parameters are varied as needed to accommodate the needs of particular<br />

studies.<br />

Whole leaves (petiole and leaflets) were cut from plants no more than four hr<br />

before treatment or inoculation. Only the second or third youngest leaf on a plant was<br />

used for the fungicide assay, and only leaves with no visible signs of injury or symptoms<br />

of disease were collected for testing. Immediately after collection, the leaves<br />

were placed in a tray lined with moist paper towels and the tray is closed to retained<br />

near 100% RH and maintained at a cool temperature (~ 12 °C). To test for protective<br />

fungicide activity, treatment compounds were applied to the upper surface of each of<br />

the three leaflets on a leaf using a chromatography sprayer. After treatment, the base<br />

of each leaf stem was inserted into sterile distilled water in a 100 x 10 mm tissue<br />

culture tube. Each upper surface of each treated leaflet was inoculated with conidia<br />

from the test fungal isolate within 24 hr of treatment. Inoculated leaves were subsequently<br />

incubated in a dew chamber for 48 hour at 100% RH, 30 o C and then maintained<br />

at 25 o C in a moist chamber at 100% RH for 10 days. Sterile distilled water<br />

was added to each tube as needed to maintain the surface of the water above the base<br />

of the petiole. If a compound was to be tested for curative activity, the leaflets were<br />

inoculated 24 hr before the fungicidal compound is applied.<br />

Experimental compounds were evaluated in a dose-response format. Azoxystrobin<br />

or other fungicides that have both protective and curative activity were used as a<br />

standard, and a solvent control was used in every study. The number of disease<br />

lesions per leaf was used to determine the ability of the test compounds to prevent<br />

infections. The size of the lesions was used to determine the curative activity of<br />

compounds. Each fungicide concentration is replicated four times and the experiment<br />

is repeated at least once. This bioassay does not differentiate between direct<br />

effects on the fungus and indirect effects through induction of plant defenses. However,<br />

if a compound is much more active in this in vivo assay than than the in vitro<br />

microtiter assay, ‘induction’ activity is indicated.<br />

2.5. 2002-03 Experimental Field Studies at Hammond, Louisiana<br />

Strawberry plots were established and maintained following standard horticulture<br />

practices used by commercial strawberry farmers in Lousisana. The following fungi-


NATURAL PRODUCT BASED FUNGICIDES 7<br />

cides were applied to strawberries: fenhexamid (Elevate®, Arvesta, San Francisco,<br />

CA) at 1.6 kg/H, fenhexamid + captan (CaptEvate®, Arvesta, San Francisco, CA) at<br />

a low rate (3.9 kg/H) and a high rate (5.82 kg/H), azoxystrobin (Quadris®, Syngenta<br />

Crop Protection, Greensboro, NC) at 0.56 kg/H, cyprodinil + fludioxonil (Switch®,<br />

Syngenta Crop Protection, Greensboro, NC) at 0.8 kg/H, captan (Micro Flo Company,<br />

Memphis, TN) at 3.4 kg/H, mylclobutanil (Nova®, Dow AgroSciences, Indianapolis,<br />

IN) at 0.28 kg/H, pyrimethanil (Scala®, Bayer CropScience, Research Triangle<br />

Park, NC) at 1.9 kg/H, fenhexamid at 1.68 kg/H + captan at 3.4 kg/H,<br />

pyraclostrobin (Cabrio®, BASF, Research Triangle Park, NC) at 0.20 kg/H,<br />

pyraclostrobin + boscalid (Pristine, BASF, Research Triangle Park, NC) at 0.62 kg/<br />

H, boscalid (Emerald®, BASF, Research Triangle Park, NC) at 0..59 kg/H, and the<br />

experimental fungicide, CAY-1 at 0.37 and 0.74 kg/H. Fungicide treatments were<br />

applied every 7-10 days (or as soon after rainfall as possible) starting at or near fullbloom<br />

stage and continuing until 1 week before harvesting ceased.<br />

Berries were harvested twice a week throughout the entire season beginning on<br />

January 30 and continuing until April 24 in 2003. Total fruit count and weights were<br />

obtained for both marketable and cull fruits, along with average weight per berry and<br />

percentages of marketable, physical cull, and diseased cull fruit per acre. Fruit were<br />

separated into marketable and cull classes, with the culls being further divided into<br />

physical (deformed or small fruit) culls and diseased culls. Diseases were identified<br />

and counted on three occasions including berries with gray mold (B. cinerea), anthracnose<br />

(C. acutatum), leather rot (Phytophthora cactorum), stem end rot (Gnomonia<br />

comari), and other rots that occurred during the picking season. Plants were rated on<br />

April 24, 2003 for incidence of foliar disease, number of dead plants due to anthracnose,<br />

crown rot and plant vigor. Diseases were rated on a scale of 0-5; 0 plants<br />

showed no leaf symptoms and 5 plants were completely defoliated. Plant vigor was<br />

rated on a scale of 0-5; 0 were dead plants and 5 were extremely vigorous plants.<br />

3. RESULTS AND DISCUSSION<br />

Results from the bioautography studies suggest that chemically different populations<br />

of microorganisms and plants can easily be distinguished by their characteristic “2D-<br />

TLC fingerprint” and antifungal zone patterns (Wedge and Nagle, 2000). Thus, this<br />

system represents a very useful technique for the identification and selection of unique<br />

chemotypes from microbial isolates (Nagle and Paul, 1999) and plant extract(?)<br />

collections.<br />

Chromatographic properties such as relative polarity, UV absorbance, chemical<br />

reactivity associated with each active metabolite provide valuable information that<br />

allows for rapid dereplication of known or nuisance compounds. When strains of<br />

different phytopathogenic fungi with dissimilar fungicide resistance profiles are<br />

inoculated onto replicate bioautography plates prepared from any given extract<br />

containing active metabolites, it is possible to visually observe distinct differences in<br />

the sensitivity of each fungal pathogen to single metabolites. These differences in<br />

pathogen sensitivity (fungicide resistance) can be observed by direct comparison of


8<br />

DAVID E. WEDGE<br />

AND BARBARA J. SMITH<br />

inhibition zone dimensions produced by active metabolites and control standards<br />

against each pathogenic strain tested. Chemical profiles provide valuable information<br />

for the rapid selection of specific antifungal metabolites with unique activity against<br />

fungicide-resistant pathogens and identify new compounds with novel mechanisms<br />

of action.<br />

Using modifications of the methods described previously, bioautography<br />

techniques were successful in allowing us to track the path of naturally occurring<br />

fungitoxic compounds present in strawberry leaves. Our studies indicated that<br />

concentrations of fungitoxic compounds vary between anthracnose resistant and<br />

susceptible cultivars and are present in different amounts in vegetative tissues of<br />

different ages. Using leaves of the anthracnose-susceptible cultivar Chandler and the<br />

anthracnose-resistant cultivar Sweet Charlie, we isolated and demonstrated the presence<br />

of three antifungal compounds. While the mechanism of strawberry anthracnose<br />

resistance is unknown, results from this study indicate that anthracnose resistance in<br />

strawberry may depend on the concentration of two constitutive antifungal compounds<br />

and the elicitation of a third compound in younger leaves.<br />

These two constitutive antifungal compounds were exhibited in both ‘Chandler’<br />

and ‘Sweet Charlie’ plants but ‘Sweet Charlie’ plants produced approximately 15<br />

times more antifungal activity than ‘Chandler’ plants. Fungal growth inhibition<br />

associated with extracts from ‘Chandler’ plants appeared to be temporary. A third<br />

compound, detected exclusively in ‘Sweet Charlie’ plants, was produced only after<br />

young leaves were sprayed with a commercially available elicitor of antifungal compounds<br />

(Vincent et al., 1999).<br />

The antifungal activity of 32 naturally occurring quinones of four major classes:<br />

1,4-naphthoquinones, 1,2-naphthoquinones, 1,4-benzoquinones, anthraquinones, and<br />

other miscellaneous compounds from our natural products collection were tested for<br />

antifungal activity using bioautography. Bioautography allowed for the rapid evaluation<br />

of quinones which demonstrated good to moderate antifungal activity against<br />

Colletotrichum spp. Colletotrichum fragariae appeared to be the most sensitive species<br />

to quinone-based chemistry, C. gloeosporioides of intermediate sensitivity, and C.<br />

acutatum was the least sensitive species to these naturally occurring compounds<br />

(Meazza et al., 2003).<br />

Bioassay-directed isolation of antifungal compounds from an ethyl acetate extract<br />

of Ruta graveolens leaves yielded two furanocoumarins, one quinoline alkaloid, and<br />

four quinolone alkaloids, including a novel compound, 1-methyl-2-[6'-(3'’,4'’methylenedioxyphenyl)hexyl]-4-quinolone.<br />

Antifungal activities of the isolated<br />

compounds, together with 7-hydroxycoumarin, 4-hydroxycoumarin, and 7methoxycoumarin<br />

which are known to occur in Rutaceae species, were evaluated<br />

using bioautography and microbioassay procedures. Four of the alkaloids had moderate<br />

activity against Colletotrichum species, including a benomyl-resistant C. acutatum.<br />

These compounds and the furanocoumarins 5- and 8-methoxypsoralen had moderate<br />

activity against Fusarium oxysporum. The novel quinolone alkaloid was highly active<br />

against Botrytis cinerea. Phomopsis species were much more sensitive to most of the


NATURAL PRODUCT BASED FUNGICIDES 9<br />

compounds tested, with P. viticola being highly sensitive to all of the compounds<br />

screened (Oliva et al., 2003).<br />

Hexane and EtOAc phases of MeOH extract of Macaranga monandra demonstrated<br />

fungal growth inhibition in C. acutatum, C. fragariae, C. gloeosporioides, F.<br />

oxysporum, B. cinerea, Phomopsis obscurans and P. viticola. Bioassay-guided fractionation<br />

resulted in the isolation of two active clerodane-type diterpenes that were<br />

elucidated by spectral methods as kolavenic acid and 2-oxo-kolavenic acid. The 96well<br />

microbioassay revealed that kolavenic acid and 2-oxo-kolavenic acid produced<br />

moderate growth inhibition in P. viticola and B. cinerea (Salah et al., 2003).<br />

The application of the microbioassay to Phytophthora nicotianae was effectively<br />

used to determine EC 50 values (i.e., effective concentration for 50% growth reduction)<br />

for eight commercial antifungal compounds (azoxystrobin, fosetyl-aluminum,<br />

etridiazole, metalaxyl, pentachloronitrobenzene, pimaricin, and propamocarb). These<br />

EC 50 values were compared to those obtained using conventional plate methods by<br />

measuring linear growth of mycelia on fungicide-amended medium. The microbioassay<br />

proved to be a rapid, reproducible, and efficient method for testing the efficacy of<br />

compounds against P. nicotianae and should be effective for other species of<br />

Phytophthora as well. The assay requires relatively small amounts of a test compound<br />

and was suitable for the evaluation of natural product samples (Kuhajek et al.,<br />

2003).<br />

CAY-1 is a fungicidal steroidal saponin (Mol wt. 1243) isolated and identified<br />

by DeLucca et al. (2002) from the ground fruit of cayenne pepper (Capsicum<br />

frutescens). CAY-1 was lethal to germinating conidia of Aspergillus flavus, A.<br />

fumigatus, A. parasiticus and A. niger. It was also active against agricultural and<br />

medicinally important fungi and yeast. In vitro dose-response assays with CAY-1<br />

against plant pathogenic fungi showed that 3.0 µM inhibited growth of C.<br />

gloeosporioides and C. acutatum by 100% and C. fragariae, P. obscurans, and P.<br />

viticola by 90%. Sampangine and several alkaloid analogs were isolated from the<br />

root bark of Cleistopholis patens. These novel broad-spectrum compounds showed<br />

promising antifungal activity against several serious pathogenic fungi of plants including<br />

B. cinerea, C. fragariae, C. acutatum, C. gloeosporioides, and F. oxysporium<br />

(Wedge and Nagle, 2003).<br />

Detached leaf assays provide us with the opportunity to evaluate new fungicides<br />

directly on the leaf surface in a dose-response format (Table 1). This assay allowed us<br />

to benchmark potential lead compounds such as CAY-1 and sampangine with a commercial<br />

standard (azoxystrobin) of known mode of action (Q o I inhibitor). The number<br />

of diseased lesions was used to determine effective concentrations needed for<br />

disease control. Lesion size is used to determine the relative effectiveness of the systemic<br />

activity that produced curative activity 24 hrs after inoculation. The detached<br />

leaf assay was also used to establish experimental field rates for future studies. Study<br />

of ‘protectant’ activity indicated that 1250 ppm. CAY-1 or sampangine appeared to<br />

be an effective concentration for disease control of anthracnose on the leaf surface, or<br />

between 100-1000 times the concentration required for in vitro activity (Post, Table 1).


10<br />

DAVID E. WEDGE<br />

AND BARBARA J. SMITH<br />

Table 1. Anthracnose disease severity and phytotoxicity (Phyto) scores of detached<br />

leaves following inoculation with Colletotrichum fragariae isolate CF-75, pre and<br />

post treatment with commercial and experimental fungicides.<br />

Azoxystrobin CAY-1 Sampangine<br />

Level Disease 1 Phyto 1 Disease Phyto Disease Phyto<br />

Fungicide applied to the upper leaf surface. Leaves not inoculated.<br />

None 0.29 a 2 0.04 A 0.29 a 0.04 a 0.29 a 0.04 a<br />

Solvent 0.29 a 0.15 A 0.29 a 0.15 a 0.29 a 0.15 a<br />

625 0.17 a 0.00 A 0.10 a 0.00 a 0.08 a 0.08 a<br />

1250 0.19 a 0.08 A 0.21 a 0.04 a 0.13 a 0.04 a<br />

2500 0.08 a 0.04 A 0.21 a 0.00 a 0.17 a 0.08 a<br />

lsd 0.34 0.20 0.32 0.16 0.30 0.19<br />

Pr>F 0.70 0.67 0.75 0.52 0.50 0.80<br />

Fungicide applied to the upper leaf surface 24 hr before inoculation (Post).<br />

None 0.75 a 0.33 A 0.75 ab 0.33 a 0.75 a 0.33 a<br />

Solvent 0.83 a 0.00 B 0.83 a 0.00 b 0.83 a 0.00 b<br />

625 0.42 ab 0.08 B 0.33 bc 0.00 b 0.25 b 0.00 b<br />

1250 0.08 b 0.17 ab 0.21 c 0.00 b 0.17 b 0.00 b<br />

2500 0.42 ab 0.00 b 0.21 c 0.08 b 0.29 b 0.00 b<br />

lsd 0.54 0.23 0.42 0.17 0.41 0.17<br />

Pr>F 0.06 0.04 0.02 0.00 0.01 0.00<br />

Fungicide applied to the upper leaf surface 24 hr after inoculation (Pre).<br />

None 0.58 a 0.08 b 0.58 abc 0.08 b 0.58 a 0.08 b<br />

Solvent 0.25 a 0.33 a 0.25 c 0.33 ab 0.25 a 0.33 ab<br />

625 0.33 a 0.08 b 0.92 ab 0.58 a 0.75 a 0.08 b<br />

1250 0.83 a 0.08 b 1.00 a 0.25 ab 0.75 a 0.08 b<br />

2500 0.67 a 0.08 b 0.83 ab 0.08 b 0.58 a 0.50 a<br />

lsd 0.60 0.23 0.47 0.57 0.36<br />

Pr>F 0.27 0.14 0.04 0.28 0.36 0.08<br />

Fungicide applied to the lower leaf surface (translaminar) 24 hr before inoculation.<br />

625 0.33 a 0.04 a . . . . 0.13 a 0.08 a<br />

1250 0.08 a 0.04 a . . . . 0.25 a 0.08 a<br />

2500 0.17 a 0.00 a . . . . 0.29 a 0.08 a<br />

lsd 0.37 0.09 0.31 0.26<br />

Pr>F 0.34 0.51 0.46 0.63<br />

Fungicide applied to the lower leaf surface (translaminar) 24 hr after inoculation.<br />

625 0.33 a 0.00 a . . . . 0.38 a 0.17 a<br />

1250 1.21 a 0.25 a . . . . 0.58 a 0.17 a<br />

2500 0.58 a 0.17 a . . . . 0.67 a 0.17 a<br />

lsd 1.09 0.31 0.71 0.43<br />

Pr>F 0.22 0.23 0.64 1.00<br />

1Disease scored on scale of 0 = no lesions to 3 = severe; phytotoxicity scored on scale of 0 – 5 where, 0 = no<br />

signs of damage 5 = severe tissue damage.<br />

2Average values followed by the same letter were not significantly different at the 0.05 level using Least<br />

Significant Difference at P=0.05.


NATURAL PRODUCT BASED FUNGICIDES 11<br />

No systemic or translaminar activity was detected when azoxystrobin, CAY-1, or<br />

sampangine was applied 24 hrs after inoculation with C. fragariae (Pre, Table 1).<br />

Field studies indicated that significant differences occurred in marketable yield<br />

of strawberries as a result of fungicide treatments. The highest marketable yield (17,071<br />

kg/H) was recorded from plots receiving Pristine fungicide at the rate of 0.62 kg ai/H.<br />

The next highest yield (16,960 kg/H) was harvested from plants treated with Switch<br />

fungicide at the rate of 0.8 kg/H. Plants from the untreated control plots yielded more<br />

fruit than plants from four of the fungicide-treated plots. Plants treated with CAY-1<br />

produced the lowest yield of cull fruit weighing 1,390 kg/H. compared with the highest<br />

yield of cull fruit weighing 5,816 kg/H produced on the plots treated with Pristine.<br />

CAY-1 produced high yields of diseased fruit weighing over 9,400 kg/H. Five fungicide<br />

treatments resulted in the production of marketable fruit at 60%: Switch® (69.8%),<br />

Pristine (68.51%), Cabrio® (65.5%), and the two CaptEvate® treatments (5.8 kg.<br />

and 3.9 kg, 65.8% and 62.7% marketable fruit, respectively). Most treatments that<br />

were combinations of two fungicides produced the highest marketable yields and<br />

lowest disease percentage.<br />

The total number of berries with fruit rot symptoms and the number of berries<br />

with symptoms of anthracnose or stem end rot from the April 24 harvest time were<br />

significantly lower from plots treated with the fungicides Switch, Cabrio®,<br />

CaptEvate®, and Pristine® than from those receiving no fungicide treatment (Table<br />

2). The most prevalent diseases in the Louisiana field study were anthracnose caused<br />

Table 2. Fruit rot and field data from April 24, 2003 harvest of fungicide treated<br />

strawberry plots, Hammond, Louisiana.<br />

Fungicide Total Rots1 Anthracnose2 Stem end rot2 Plants Dead (%) 3 Foliar disease4 Switch 1.5 f 5 0.8 e 0.0 d 16.3 a 2.8 a<br />

Cabrio 5.8 f 3.8 e 0.5 d 8.8 cde 2.6 ab<br />

CaptEvate (High) 6.0 f 5.0 e 0.5 d 5.0 e 1.1 e<br />

Pristine 7.0 ef 5.0 e 0.3 d 5.0 e 2.5 abc<br />

CAY-1 (Half) 22.0 def 18.0 de 2.0 cd 15.6 ab 2.8 a<br />

Captan 27.8 cdef 21.3 cde 3.0 cd 7.5 cde 2.0 bcd<br />

Quadris 28.3 cdef 20.3 cde 2.8 cd 8.1 cde 1.9 cd<br />

CaptEvate (Low) 29.0 cdef 23.3 bcde 2.5 cd 7.5 cde 1.8 de<br />

Captan+Elevate 43.0 cdef 28.8 bcde 7.0 bcd 12.5 abc 2.0 bcd<br />

CAY-1 (Full* 50.5 bcde 37.8 bcd 7.5 bcd 10.6 abcde 3.1 a<br />

Scala 59.0 bcd 47.8 bcd 7.0 bcd 9.4 bcde 1.9 cd<br />

Control 63.5 bcd 45.3 bcd 9.8 abc 5.6 de 2.8 a<br />

Emerald 70.5 abc 50.3 Bc 7.8 abcd 5.6 de 2.9 a<br />

Nova 87.8 ab 54.5 Ab 16.0 a 7.5 cde 2.0 bcd<br />

Elevate 109.5a 84.5 A 14.3 ab 11.9 abcd 2.6 ab<br />

LSD (0.05) 44.1 31.6 8.4 6.8 0.7<br />

1Total number of fruit/20 plant plot with any disease symptom.<br />

2Number of fruit with anthracnose fruit rot symptomn (Colletotrichum acutatum) or Stem-End Rot<br />

(Gnomonia comari).<br />

3Percentage of plants/20 plant plot dead from anthracnose crown rot.<br />

4Foliar disease scored on scale of 0 – 5, where 0 = no disease and 5 = plant defoliated due to foliar disease.<br />

5Average values followed by the same letter were not significantly different at the 0.05 level using a Least<br />

Significant Difference.


12<br />

DAVID E. WEDGE<br />

AND BARBARA J. SMITH<br />

by the fungus C. acutatum and stem end rot caused by the fungus Gnomonia comari.<br />

Stem end rot lesions were often invaded by secondary pathogens such as Botrytis and<br />

Colletotrichum. The combination fungicides such as Elevate® + Captan®, Pristine®,<br />

and Switch® were effective in controlling this disease complex. The untreated control<br />

plants and those treated with Emerald® had the most berries affected with stem<br />

end rot. The incidence of gray mold and anthracnose fruit rot was extremely low.<br />

Gray mold was controlled by the fungicides Scala®, Elevate® + Captan®, Pristine®,<br />

Switch®, Emerald®, and Elevate®.<br />

4. CONCLUSIONS<br />

Information gained from the examination of thousands of extracts and their associated<br />

pure compounds have culminated in the development of a variety of standardized<br />

operating protocols for natural product discovery that are currently being used in our<br />

laboratory. Successful discovery, evaluation, and development of natural product<br />

fungicides are totally dependent upon the availability of high quality miniaturized<br />

antifungal bioassays. Bioassay-directed screening of compounds and extracts is the<br />

initial step in the discovery process for new agrochemicals and pesticides.<br />

Standardization of inoculum allows for meaningful comparison of growth<br />

inhibition between different fungal pathogens, test compounds, and experiments<br />

repeated in time. Bioautography provides a simple technique to visually follow<br />

antifungal components through the separation process. The 96-well microbioassay<br />

allows for the evaluation of microgram quantities, determination of dose-response<br />

relationships, and comparison of antifungal activity with fungicides with a known<br />

mode of action. Coupling bioautography techniques with the 96-well microbioassay<br />

provides us with a discovery protocol that combines the simple and visual nature of<br />

direct bioautography with the rapid, sensitive, and high throughput capabilities of a<br />

microtiter system.<br />

The 96-well microbioassay is accurate and sensitive; as little as 0.1 µM amounts<br />

of test compound permit discrimination between germination and mycelial growth<br />

inhibitors and identification of fungicide resistant pathogens. The microbioassay<br />

utilizes a chemically defined liquid medium with a zwitterion buffer that limits chemical<br />

interaction with test compounds and controls for pH variations. This new standardized<br />

method provides high-throughput capability and the capacity to study chemical<br />

compounds in detail, to perform mode of action studies, and to determine fungicide<br />

resistance profiles for specific fungal pathogens.<br />

Detached leaf assays are critical for establishing ‘real world’ activity prior to the<br />

field testing that agrochemical companies require before investing millions of dollars<br />

needed to develop a agrochemical. Subsequent efficacy testing in the greenhouse<br />

ultimately helps determine the potential usefulness of compounds as pest control agents.<br />

To maximize the detection of natural products, high-throughput bioassay techniques<br />

must target significant agricultural pests, include relevant commercial pesticide<br />

standards, and adhere to sound statistical principles.


NATURAL PRODUCT BASED FUNGICIDES 13<br />

While the use of SAR inducers is still in its infancy, these compounds will may<br />

be of use in specialized applications. However, our observations are such that the<br />

host plant must have a suitable ‘metabolic engine’ - found in many resistant cultivars<br />

- that is capable of being ‘revved up.’ Susceptible cultivars most often used in fungicide<br />

evaluations either produce a lower quantity of defense compound and or have a<br />

longer time lag between pathogen infection and plant symptom development than is<br />

the case in resistant cultivars. Systemic acquired resistance inducers have been marketed<br />

(Actigard, Messenger) with limited success and have never recouped their development<br />

costs. It thus appears unlikely that economically viable SAR products will<br />

be developed in the near future.<br />

Even so, allelochemicals and other natural products represent potentially rich<br />

and new sources of agrochemical plant protectants. The identification of suitable<br />

natural products coupled with traditional synthesis chemistry should be an effective<br />

approach for optimizing pesiticide activity and associated chemical properties. New<br />

requirements for fungicides are stringent, and require that new products must be<br />

environmentally safe and efficacious at low rates. In addition they must posses low<br />

mammalian toxicity, operate using novel modes of action, and have a low to moderate<br />

risk of developing resistance in target pathogens.<br />

5. REFERENCES<br />

Agrios, G. N. How Pathogens Attack Plants. In: Plant Pathology. Academic Press, San Diego, 1997; 63-82.<br />

Espinel-Ingrof, A., Kerkering, T. M. Spectrophotometric method of inoculum preparation for the in vitro<br />

susceptibility testing of filamentous fungi. Journal of Clinic Microbiology 1991; 29:393-394.<br />

De Lucca, A. J., Bland, J. M.,Vigo, C. B., Cushion, M., Selitrennikoff, C. P., Peter, J., Walsh, T. J. CAY-1, a<br />

fungicidal saponin from Capsicum species. Med Mycol 2002; 40:131-137.<br />

Farmer, E. E., Ryan, C. A. Octadecanoid precursors of jasmonic acid activate the synthesis of wound-inducible<br />

proteinase inhibitors. Plant Cell 1992; 4:129-134.<br />

Gundlach, H., Muller, M. J., Kutchan, T. M., Zenk, M. H. Jasmonic acid is a signal transducer in elicitorinduced<br />

plant cell cultures. Proc Natl Acad Sci USA 1992; 89:2389-2393.<br />

Homans, A. L., Fuchs, A. Direct bioautography on thin-layer chromatograms as a method for detecting fungitoxic<br />

substances. J Chromatogr 1970; 5(2):327-3299.<br />

Kuhajek, J. M., Jeffers, S. N., Slattery, M., Wedge, D. E. A Rapid Microbioassay for Discovery of Novel Fungicides<br />

for Phytophthora spp. Phytopathology 2003; 93:46-53.<br />

Nagle, D. G., Paul, V. J. Production of secondary metabolites by filamentous tropical marine cyanobacteria:<br />

ecological functions of the compounds. J Phycol 1999; 35 (Suppl. 607):1412-1421.<br />

Meazza, G., Dayan, F. E., Wedge, D. E. Activity of activity quinones on Colletotrichum spp. J Agric Food<br />

Chem 2003; 51:3824-3828.<br />

Moore, R. E. Cyclic peptides and depsipeptides from cyanobacteria: a review. J Ind Microbiol. 1996; 16:134-<br />

143.<br />

Mueller, M. J., Brodschelm, W., Spannagl, E., Zenk, M. H. Signaling in the elicitation process is mediated<br />

through the octadecanoid pathway leading to jasmonic acid. Proc Natl Acad Sci USA 1993; 90:7490-<br />

7494.<br />

Oku, H. In: Plant Pathogenesis and Disease Control. Lewis Publishers. Boca Raton, FL 1994; p.64.<br />

Oliva, A., Meepagala, K. M. Wedge, D. E. Hale, A. L., Harries, D., Aliotta, G., Duke, S.O. Natural Fungicides<br />

from Ruta graveolens L. leaves, including a new quinolone alkaloid. J Agric Food Chem 2003; 51: 890-<br />

896<br />

Pezzuto, J. M. Cancer chemopreventative agents: from plant materials to clinical intervention trials. In: A. D.<br />

Kinghorn, M. F. Balandrin (eds): Human Medicinal Agents from Plants. ACS Symposium Series No.<br />

534, American Chemical Society, Washington DC, 1993; 205-215.


14<br />

DAVID E. WEDGE<br />

AND BARBARA J. SMITH<br />

Roughi, A. M. Rediscovering natural products. Chemical and Engineering News. October 13, 2003; p. 77-91.<br />

Salah, A. M., Bedir, E., Ngeh, T. J., Kahn, I. A., Wedge, D.E. Antifungal clerodane diterpenes from Macaranga<br />

monandra (L.) Muell. Et Arg. (Euphorbiaceae). J Agric Food Chem 2003; 51:7607-7610.<br />

Tellez, M.R., Dayan, F.E., Schrader, K.K., Wedge, D.E., Duke, S.O. Composition and some biological activities<br />

of the essential oil of Callicarpa americana L. J Agric Food Chem 2000; 48:3008-3012.<br />

Vick, B. A., Zimmerman, D. C. Biosynthesis of jasmonic acid by several plant species. Plant Physiol 1984; 75:<br />

458-461.<br />

Vincent, A., Dayan, F.E., Maas, J.L., Wedge, D.E. Detection and isolation of antifungal compounds in strawberry<br />

inhibitory to Colletotrichum fragariae. Adv Strawberry Res 1999; 18:28-36.<br />

Wedge, D. E., Kuhajek, J. M. A microbioassay for fungicide discovery. SAAS Bull Biochem Biotech 1998; 11:<br />

1-7.<br />

Wedge, D. E., Camper, N. D. Connections between agrochemicals and pharmaceuticals. In: <strong>Biologica</strong>lly Active<br />

Natural Products: Agrochemicals and Pharmaceuticals. Cutler, H.G., Cutler, S. J. eds, ACS Symposium<br />

Series, American Chemical Society, Washington DC, 2000 .<br />

Wedge. D. E., Nagle, D.G. A new 2D-TLC bioautography method for the discovery of novel antifungal agents<br />

to control plant pathogens. J Nat Prod 2000; 63:1050-1054.<br />

Wedge, David E.; Nagle, Dale G. Preparation of sampangine and its analogs as fungicides. U.S. Pat. Appl.<br />

Publ. 12 pp. CODEN: USXXCO US 2004192721 A1 20040930 CAN 141:273006 AN 2004:803936,<br />

2004.<br />

Wedge, D.E., Curry, K.J., Boudreaux, J.E., Pace, P.F., Smith, B.J. In vitro sensitivity and resistance profiles of<br />

Botrytis cinerea isolates from Louisiana strawberry farms. Proceedings of the 5 th<br />

North American<br />

Strawberry Conference, Strawberry Research to 2001; p78-81.


ALLELOCHEMICALS AS BIOPESTICIDES FOR<br />

MANAGEMENT OF PLANT-PARASITIC<br />

NEMATODES<br />

1<br />

NANCY KOKALIS-BURELLE 1 and RODRIGO<br />

RODRÍGUEZ-KÁBANA 2<br />

Research Ecologist, USDA, Agricultural Research Service, U.S.<br />

Horticultural Research Lab, Fort Pierce, FL, 34945, USA, 2<br />

Distinguished<br />

University Professor, Auburn University, Auburn, AL, 36849, USA.<br />

Email:NBurelle@ushrl.ars.usda.gov<br />

Abstract. Many allelopathic compounds in their native or processed forms have potential for development as<br />

viable components of plant-parasitic nematode management strategies. <strong>Allelochemicals</strong> have been identified<br />

that possess differing levels of activity against a wide range of plant-parasitic nematodes. In general, these<br />

compounds are less toxic to nontarget species, and less persistent in soil than chemical nematicides. Operative<br />

mechanisms for plant-parasitic nematode control with allelopathic compounds include nematicidal activity,<br />

nematostatic activity, and nematode behavior modification. <strong>Allelochemicals</strong> are sometimes produced in large<br />

quantities in plant material or as agricultural waste, making the use of rotation crops, cover crops, and organic<br />

amendments effective means for production and/or distribution of the active compounds. A greater understanding<br />

of the effects of soil microbes and environmental conditions on allelopathic compounds is necessary to improve<br />

their efficacy for control of parasitic nematodes. Use of allelochemicals for nematode control will require that<br />

growers know specifically what types and population levels of nematodes are present in their production fields.<br />

Development of improved production and incorporation methods for rotation and green manure crops, and<br />

appropriate application methods for processed allelochemical compounds, will also enhance the efficacy and<br />

consistency of these compounds for nematode control.<br />

1. INTRODUCTION<br />

For the purpose of this chapter, allelochemicals will be defined as plant metabolites or<br />

their products that are released into the environment through volatilization, exudation<br />

from roots, leaching from plants or plant residues, and decomposition of residues<br />

(Waller, 1985; Putnam and Tang, 1986; Einhellig,1995; Halbrendt, 1996). As such,<br />

allelochemicals are classified as biopesticides, which are defined as being derived<br />

from natural materials such as plants and microorganisms and include both substances<br />

that control pests (biochemical pesticides) and microorganisms that control pests<br />

(microbial pesticides). Biochemical and microbial pesticides are considered inherently<br />

less toxic than conventional pesticides and generally affect only the target pest and<br />

closely related organisms, in contrast to broad spectrum, conventional pesticides that<br />

may affect organisms as different as birds, insects, and mammals. Often biopesticides<br />

are effective in small quantities and decompose quickly, resulting in lower exposure<br />

and fewer pollution problems than conventional pesticides. Also, when used as a<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 15–<br />

29.<br />

© 2006 Springer. Printed in the Netherlands.<br />

15


16<br />

NANCY KOKALIS-BURELLE AND RODRÍGO RODRIGUEZ-KÁBANA component of Integrated Pest Management (IPM) programs, biopesticides can decrease<br />

the use of conventional chemical pesticides (Halbrendt, 1996).<br />

Increased restrictions on use and phase-out of chemical fumigants such as methyl<br />

bromide and other chemical nematicides for control of plant-parasitic nematodes make<br />

the discovery of target-specific, environmentally safe, naturally occurring biopesticide<br />

compounds that suppress nematode populations or modify nematode behavior<br />

increasingly important. For instance, allelochemicals that affect nematode chemotaxis<br />

could be invaluable in many different scenarios for nematode control and represents<br />

an area of research with both great needs for identification of active compounds, and<br />

great possibilities for their use. The production of chemical cues by plants and<br />

behavioral responses to these cues by nematodes are critical to successful host location<br />

and reproduction in nematodes, which are highly dependent on these chemotactic<br />

stimuli during many stages of their life cycles (Bridge, 1996; Huettel, 1986; Yasuhira<br />

et al., 1982).<br />

In order to make effective use of nonchemical or biochemical pest control strategies<br />

such as allelochemicals for suppression of plant-parasitic nematodes, users need to be<br />

familiar with the types and quantity of nematodes present in their soil and determine<br />

the feasibility of growing a cover or rotation crop, using an organic amendment, or a<br />

formulated biopesticide. Identification of cash crops that produce compounds capable<br />

of reducing pathogenic nematode populations and that can be incorporated into existing<br />

production regimes is rare (Gardner et al., 1992; Mojtahedi et al., 1993a; Halbrendt,<br />

1996). Some notable exceptions to this include commercial production of Crotalaria,<br />

mustard, African marigold, asparagus, and sesame in India (Bridge, 1996). The level<br />

to which growers will employ the use of cover crops or rotation crops is ultimately<br />

dependent on the economic feasibility of this method for nematode control. When<br />

determining the economic feasibility of this approach, the additional benefits that<br />

cover crops provide should be considered. These benefits include nitrogen fixation,<br />

soil stabilization, and weed management (Halbrendt, 1996).<br />

Many plant constituents and metabolites have been investigated for activity against<br />

plant-parasitic nematodes. The conditions under which compounds are effective<br />

against nematodes vary with the compounds (Ferris and Zheng, 1999; Zasada and<br />

Ferris, 2004). These active compounds, or precursors of active compounds, can often<br />

be applied to soil as organic amendments, or refined and developed as biopesticide<br />

compounds.<br />

Most allelochemicals are short-lived in soil as they are often easily metabolized<br />

or hydrolyzed, and may require that plants are actively growing and secreting them<br />

into the rhizosphere in order to be effective (Cheng, 1992). In some cases, breakdown<br />

products of allelochemicals are the active components against nematodes (Borek<br />

et al., 1995). Many such compounds are produced upon decomposition of plant material<br />

that can be useful when incorporated into soil as green manures (Brown and Morra,<br />

1997; Mojtahedi et al., 1991; 1993a, b; Prot et al., 1992; Zasada and Ferris, 2004). In<br />

each of these instances, soil physical and chemical conditions, microbial populations,<br />

and environmental conditions influence the retention, transformation and transport<br />

of the allelochemicals (Cheng, 1992). Undoubtedly, these physical, microbiological,


ALLELOCHEMICALS : MANAGEMENT OF PLANT-PARASITIC NEMATODES 17<br />

and environmental factors contribute to the inconsistency of nematode control observed<br />

in research trials on crop rotation and cover crop systems, biofumigation, and<br />

biochemical pesticides. It is likely that synthetic chemical nematicides will continue<br />

to fill short-term nematode control needs while research continues to improve<br />

nonchemical and biopestide-based approaches, which will eventually become the<br />

management strategies of choice (Thomas, 1996). In this chapter we will review<br />

examples of plants known to release allelochemicals into soil while actively growing,<br />

when incorporated into soil as green manures or organic amendments, and the direct<br />

application of purified allelochemicals or formulations of biopesticides for plantparasitic<br />

nematode control.<br />

2. ROTATION AND COVER CROPS<br />

There are many examples of crop rotation sequences that passively suppress nematode<br />

populations which will not be reviewed here. Examples of active nematode suppression<br />

in crop rotation sequences are typically found with plant species that produce and<br />

excrete allelopathic compounds. These compounds then affect plant-parasitic<br />

nematodes in the rhizosphere either directly or indirectly by altering rhizosphere<br />

microbial populations (Halbrendt, 1996). For the purpose of this chapter, allelopathic<br />

Table 1. Rotation/cover crops that actively suppress<br />

parasitic nematode populations in soil.<br />

Common <strong>Name</strong> Scientific <strong>Name</strong> References<br />

American joint vetch Aeschynomene sp. Rodriguez-Kabana et al., 1991a<br />

Bahia grass Paspalum spp. Rodríguez-Kábana et al., 1994b<br />

Castor bean Ricinus communis Rodriguez-Kabana et al., 1991b<br />

Marigold Tagetes spp. Tyler, 1938<br />

Steiner, 1941<br />

Uhlenbroek and Bijloo, 1958, 1959<br />

Good et al., 1965<br />

Hairy indigo Indigofera hirsuta Rodriguez-Kabana et al., 1988b<br />

Horse bean Canavalia ensiformis Rodriguez-Kabana et al., 1992b<br />

Partridge pea Cassia fasciculata Rodriguez-Kabana et al., 1991a<br />

Rodriguez-Kabana et al., 1995<br />

Sesame Sesamum indicum Rodríguez-Kábana et al., 1994a<br />

Showy clotalaria Crotalaria spectabilis Rodriguez-Kabana et al., 1992b<br />

Sorghum-sudan grass S. bicolor X S. vulgare var. sudanense Kinloch and Dunavin, 1993<br />

Sudan grass Sorghum vulgare var. sudanense Mojtahedi et al., 1993a<br />

Sunn hemp Crotalaria juncea Sipes and Arakaki, 1997<br />

Robinson et al., 1998<br />

McSorley et al., 1999<br />

Wang et al., 2001<br />

Velvet bean Mucuna deeringiana Rodriguez-Kabana et al., 1992a<br />

Weaver et al., 1993<br />

Taylor and Rodriguez-Kabana, 1999<br />

Vargas-Ayala and Rodriguez-Kabana, 2001<br />

Vetch Vicia spp. Minton et al., 1966<br />

Minton and Donnelly, 1967


18<br />

rotation and cover crops are considered to be those that are not typically incorporated<br />

into soil in order to realize their allelopathic potential. Green manure crops considered<br />

to be crops that are most effective for nematode control when incorporated into soil as<br />

organic amendments. However, there are many examples of effective rotation crops<br />

that are also incorporated into soil as green manure at the end of the growing-season.<br />

Plants belonging to 57 families have been shown to possess nematicidal properties<br />

(Bridge, 1996). Crop rotation (crops planted in sequence) and intercropping (crops<br />

planted together) have both shown potential for reducing populations of parasitic<br />

nematodes. Cover crops are a type of rotation employed as an alternative to leaving<br />

land fallow, and are usually not cash crops. Many plants have been used as rotation<br />

or cover crops for plant-parasitic nematode control (Table 1).<br />

The active allelopathic compounds differ with respect to each crop and, in many<br />

cases, the mode of action for nematode suppression and the effects on nematode<br />

chemotaxis have not been established. Several examples of rotation crops effective in<br />

actively suppressing populations or controlling parasitic nematodes are reviewed.<br />

2.1. Marigold (Tagetes spp.)<br />

NANCY KOKALIS-BURELLE AND RODRÍGO RODRIGUEZ-KÁBANA Marigold was one of the first plants reported to be highly resistant to root-knot<br />

nematodes (Meloidogyne spp.) (Tyler, 1938). Soon after resistance was identified in<br />

marigold, it was reported that root-knot nematode larvae entered the roots but failed<br />

to develop to sexual maturity (Steiner, 1941). It was later discovered that Tagetes<br />

spp. produce nematotoxic compounds identified as alpha-terthienyls, which directly<br />

affect a wide range of nematodes (Uhlenbroek and Bijloo, 1958, 1959; Good et al.,<br />

1965). The evaluation of several marigold species as rotation or cover crops has<br />

shown Tagetes minuta to be highly resistant to both Meloidogyne incognita and M.<br />

javanica, and to perform well as a summer cover crop in southern Florida (McSorley,<br />

1999). The more commonly used T. erecta and T. patula are often used as winter or<br />

spring bedding plants in Florida, and are not as tolerant of high temperatures as T.<br />

minuta (McSorley and Frederick, 1994; McSorley, 1999). Because T. minuta is more<br />

heat- tolerant, it has potential for use as a nematode suppressive summer cover crop<br />

in Florida during months when fields are often left fallow after fall and spring vegetable<br />

crop production cycles. Ploeg (1999) found that T. erecta, T. patula, T. signata and a<br />

hybrid Tagetes reduced galling by M. incognita, M. javanica, M. arenaria, and M.<br />

hapla in a tomato rotation study. However, all four Meloidogyne species reproduced<br />

on T. signata ‘Tangerine Gem’. This indicates the potential for substantial variability<br />

in nematode suppression among similar cultivars of marigold.<br />

There has been little research addressing the extent of allelopathic activity and<br />

the mechanism involved in nematode suppression with marigold. Sasanelli and DiVito<br />

(1991) found that aqueous leaf and root extracts and root leachates of two Tagetes sp.<br />

were nematostatic rather than nematicidal to eggs of the golden nematode Globodera<br />

rostochiensis. Emergence of juveniles that was suppressed or reduced in the presence


ALLELOCHEMICALS : MANAGEMENT OF PLANT-PARASITIC NEMATODES 19<br />

of extracts and diffusates resumed when solutions were removed indicating that<br />

compounds found in solutions were nematostatic rather than nematicidal. More specific<br />

information is needed on the range of activity and mechanisms involved in nematode<br />

suppression with compounds found in marigold. Additional research in these areas<br />

would increase the potential for successful incorporation of marigold or its active<br />

allelopathic compounds into production systems for nematode control.<br />

2.2. Sorghum-sudangrass (Sorghum bicolor X S. sudanense)<br />

Sorghum has long been recognized for its allelopathic properties toward other plants<br />

(Guenzi and McCalla, 1966) and more recently to be suppressive to nematodes (Kinlock<br />

and Dunavin, 1993; Mojtahedi et al., 1993a). It was initially hypothesized that the<br />

nematicidal compound in sorghum-sudangrass green manures was hydrogen cyanide<br />

produced by hydrolysis of dhurrin in leaf tissue (Andewusi, 1990). However, Czarnota<br />

et al. (2003) examined root exudate production and composition of seven genetically<br />

diverse sorghum accessions, including two sorghum-sudangrass hybrids, and found<br />

that although variation occurred in exudate constituents among accessions, the<br />

predominant constituent in all exudates was the phenolic compound sorgoleone<br />

(Czarnota et al., 2003).<br />

Suppression of parasitic nematodes in the field with sorghum-sudangrass has<br />

been inconsistent (MacGuidwin and Layne, 1995). It has been demonstrated that this<br />

crop is not effective for reducing populations of lesion nematodes (Pratylenchus sp.),<br />

an important plant-parasitic genus (MacGuidwin and Layne, 1995). Many studies<br />

have confirmed that sorghum-sudangrass is effective in reducing field populations of<br />

Meloidogyne spp., but that it cannot be recommended if stubby root nematode,<br />

Paratrichodorus minor, is present and of concern due to the high reproductive rates<br />

of P. minor on sorghum-sudangrass (McSorley and Gallaher, 1991; McSorley et al.,<br />

1994a; McSorley and Dickson, 1995). The selective nature of parasitic nematode<br />

control with this allelopathic crop serves as an example of why it is critical to know<br />

what nematode species are present in a location when employing cover or rotation<br />

crops for key nematode pests. Threshold levels for secondary parasitic nematodes<br />

should be established and susceptibility of potential cover crop known when employing<br />

these strategies.<br />

2.3. Sesame (Sesamum indicum)<br />

Sesame is an important seed and oil crop worldwide. In addition to being a poor host<br />

for root-knot nematodes (Rodriguez-Kabana et al., 1988a; 1989), sesame is known to<br />

produce several lignin compounds including sesamin and sesamolin which are<br />

antioxidants that function as insecticides and insecticidal synergists and are<br />

hypothesized to be the active allelopathic compounds in sesame (Bedigian and Harlan,<br />

1986). Research has been conducted in the southern United States during the past 15<br />

years to evaluate the potential for use of sesame as a profitable rotation crop for root-


20<br />

NANCY KOKALIS-BURELLE AND RODRÍGO RODRIGUEZ-KÁBANA knot nematode control. Meloidogyne arenaria populations in peanut were reduced<br />

using a two-year sesame rotation followed by one year of peanut production (Rodriguez-<br />

Kabana et al., 1994a). Studies by Starr and Black (1995) confirm that sesame can be<br />

an effective rotation crop for control of M. arenaria and M. incognita but that it is not<br />

effective in controlling M. javanica. More work is needed to determine the identities<br />

of the active compounds found in sesame and the nature of their activity with regards<br />

to parasitic nematodes.<br />

2.4. Velvetbean (Mucuna deeringiana)<br />

Velvetbean (Mucuna deeringiana) is a legume commonly used as a cover crop, green<br />

manure, and forage in many subtropical and tropical regions (Allen and Allen, 1981).<br />

Velvetbean is known to produce L-DOPA (L-3,4-dihydroxyphenylalanine) which<br />

inhibits the growth of other plants and also has insecticidal properties (Fujii et al.,<br />

1991; Bell and Janzen, 1971). Vincente and Acosta (1987) demonstrated the<br />

antagonistic capabilities of velvetbean against plant-parasitic nematodes when used<br />

as a soil amendment. Vargas-Ayela and Rodriguez-Kabana (2001) found an increase<br />

in beneficial rhizosphere microorganisms including species of Paecilomyces and<br />

Burkholderia associated with reduction in disease caused by root-knot nematodes in<br />

velvetbean- soybean rotations compared to cowpea-soybean rotations. McSorley et al.<br />

(1994b) found velvetbean to be very successful in reducing populations of M. arenaria,<br />

M. javanica, and several races of M. incognita. Although the active compounds in<br />

velvetbean have been identified, little research has been performed to determine how<br />

to optimize the potential of this crop for nematode suppression.<br />

2.5. Sunn hemp (Crotalaria juncea)<br />

Sunn hemp is a legume that has many of the desirable qualities of a good rotation or<br />

cover crop. In addition to its ability to fix nitrogen, sunn hemp grows rapidly and<br />

produces a large amount of biomass, increases soil organic matter, sequesters carbon<br />

(Rotar and Joy, 1983), and suppresses many plant-parasitic nematodes (McSorley et<br />

al., 1999; Robinson et al., 1998; Wang et al., 2001). Sunn hemp also increases<br />

populations of important nematode-antagonistic fungi in soil (Quiroga-Madrigal et<br />

al., 1999; Rodríguez-Kábana and Kloepper, 1998; Wang et al., 2001). Populations of<br />

microbivorous nematodes, which are involved in soil nutrient cycling, have also been<br />

shown to increase in sunn hemp-planted soil (Venette et al., 1997; Wang et al., 2003).<br />

Sunn hemp is well suited for use as a cover crop in subtropical regions (McSorley,<br />

1999) and did not increase population levels of M. javanica during a test in Hawaii<br />

(Sipes and Arakaki, 1997). Sunn hemp ‘Tropic Sun’ was highly resistant, but not<br />

immune, to the Meloidogyne spp. tested (McSorley, 1999). It has been suggested that<br />

elevated numbers of free-living nematodes, including bacteriovores and fungivores,<br />

may improve plant tolerance to parasitic nematodes by increasing plant vigor through<br />

more efficient nutrient cycling, or by increasing numbers of predatory and omnivorous<br />

nematodes that then feed on plant-parasitic nematodes (Wang et al., 2003).


3.1. Brassicaceae<br />

ALLELOCHEMICALS : MANAGEMENT OF PLANT-PARASITIC NEMATODES 21<br />

3. GREEN MANURES AND ORGANIC AMENDMENTS<br />

Glucosinolates are compounds produced by many members of the Brassicaceae plant<br />

family including mustard, rape, canola, and cabbage (Fahey et al., 2001). These<br />

compounds are β-D-thioglucosides with differing organic side chains, and can be<br />

aliphatic, aromatic, or indole forms. Enzymatic degradation of glucosinolates produces<br />

several types of compounds including isothyocyanate (ITC), a well known nematicide<br />

(Brown and Morra, 1997).<br />

Glucosinolates are produced in various levels throughout the plant and to different<br />

degrees among plant species (Fahey et al., 2001). In studies by Potter et al. (1998) the<br />

nematicidal potential of Brassica spp. leaf and root tissue differed, with leaf tissue<br />

being far more toxic to P. neglectus than root tissue. However, leaf tissue toxicity was<br />

not correlated with either total glucosinolate content or with any individual<br />

glucosinolate, while the suppression achieved with the addition of root tissue was<br />

highly correlated with levels of 2-phenylethyl glucosinolate within roots (Potter et<br />

al., 1998). Zasada and Ferris (2004) performed studies where levels of glucosinolates<br />

in various Brassica spp. were determined and amendments containing equivalent<br />

amounts of active compounds were compared to determine their efficacy in controlling<br />

M. javanica and Tylenchulus semipenetrans. They determined that the degradation<br />

of brassicaceous amendments in soil resulted in a series of biological and chemical<br />

processes that differed among plant species. It was also determined that consistent<br />

nematode suppression could be achieved using brassicaceous amendments if the<br />

chemical composition of the organic material was considered and appropriate levels<br />

of biomass applied (Zasada and Ferris, 2004).<br />

3.2. Neem (Azadirachta indica)<br />

The neem tree is a tropical evergreen tree native to India whose biologically active<br />

properties have been recognized in Asia for centuries (Thakur et al., 1981). During<br />

the past 30 years, interest in the extensive variety of biologically active compounds<br />

produced by neem has increased. Various parts of the neem plant including leaves,<br />

seeds and bark, produce over 40 active diterpenoid, triterpenoid, limonoid, and<br />

flavonoid compounds, with a wide range of nematicidal activity (Bhatnagar and<br />

Goswami, 1987). The most well known and active compounds produced by neem are<br />

the azadirachtins (Thakur et al., 1981).<br />

Nematode control has been achieved following the incorporation of neem products<br />

into soil, and their subsequent decomposition and release of nematicidal compounds<br />

(Stirling, 1991). Numerous studies have shown that leaves and oilcake of neem are<br />

effective in controlling M. incognita and increasing growth and yield of vegetable<br />

crops when used as a soil amendment (Akhtar, 1998; Bhatnagar and Goswami, 1987).<br />

This effect has also been reported to persist in successive crops (Akhtar and Alam,<br />

1991; Akhtar and Mahmood, 1996). Neem based products have also shown nematicidal


22<br />

potential when applied as seed treatments (Akhtar and Mahmood, 1995; 1997) and<br />

bare-root treatments (Akhtar and Mahmood, 1993; 1994) leading some to conclude<br />

that compounds found in neem may act as inducers of resistance to some nematodes<br />

including M. incognita and Rotylenchulus reniformis (Siddiqui and Alam 1988). Other<br />

nematode-suppressive mechanisms of compounds derived from neem include<br />

antifeedent, repellent, deterrent, growth disruption, juvenile toxicant, and ovicidal<br />

properties (Akhtar, 1998). Testing of neem-based products and development of<br />

application techniques for plant-parasitic nematode control is increasing in western<br />

countries. There are currently several neem-based pesticides available in the United<br />

States for use on certain greenhouse and ornamental crops, with many more available<br />

for use in India as insecticides (Akhtar, 2000). A comprehensive review of the<br />

nematode suppressive potential of neem products is provided by Akhtar (2000).<br />

4. ALLELOCHEMICALS AS BIOPESTICIDES<br />

Extracts of many plants with anthelminthic or antimicrobial properties have been<br />

proven effective in reducing soil populations of plant-parasitic nematodes (Ferris and<br />

Zheng, 1999). Experiments which evaluated plant species documented in Chinese<br />

traditional medicine to be anthelminthic against plant-parasitic nematodes identified<br />

153 aqueous plant extracts with activity against nematodes (Ferris and Zheng, 1999).<br />

Within a 24-hour exposure period, seventy-three of the extracts killed either juveniles<br />

of M. javanica or mixed developmental stages of P. vulnus, or both (Ferris and Zheng,<br />

1999). Plants containing efficacious components included both annuals and perennials,<br />

which ranged in type from grasses and herbs to woody trees, representing 46 plant<br />

families (Ferris and Zheng, 1999). This research illustrates the tremendous potential<br />

for discovery of new active allelopathic compounds for plant-parasitic nematode<br />

control. In fact, many of the allelochemicals described below were isolated from<br />

crops observed to be nematode suppressive as rotation or green manure crops.<br />

4.1. Glucosinolates<br />

NANCY KOKALIS-BURELLE AND RODRÍGO RODRIGUEZ-KÁBANA Glucosinolates are compounds primarily found in plants in the family Brassicaceae<br />

and are described previously in this chapter with respect to green manure crops.<br />

Enzymatic decomposition of glucosinolates in plant tissue occurs rapidly and is<br />

primarily attributed to microorganisms in soil (Fenwick et al., 1983). Products of<br />

glucosinolate degradation include organic cyanides and isothiocyanates which in<br />

addition to being evaluated as active compounds in green manures and organic<br />

amendments, have been studied for their their direct toxicity to nematodes as<br />

biochemical pesticides. Lazzeri et al. (1993) studied the direct effects of purified<br />

glucosinolates on second-stage juveniles of the sugar beet cyst nematode Heterodera<br />

schachtii. Compounds were isolated from seeds and plant tissue of brassicaceous<br />

hosts of the nematode. None of the glucosinolates tested in their native form were<br />

nematicidal. However, when exposed to the enzyme myrosinase, several compounds<br />

including sinigrin, gluconapin, glucotropeolin, glucode-hydroerucin, and the entire


ALLELOCHEMICALS : MANAGEMENT OF PLANT-PARASITIC NEMATODES 23<br />

group of glucosinolate compounds extracted from rapeseed exhibited various levels<br />

of nematicidal activity depending on concentration and length of exposure (Lazzeri<br />

et al., 1993).<br />

Later studies by Borek et al. (1995) investigated the persistence of glucosinolatederived<br />

allyl isothiocyanate and allylnitrile in six soils. They found that the two<br />

compounds differed with respect to the temperature, moisture conditions, and soil<br />

physical conditions that effected their transformation in soil, and that both compounds<br />

dissipated from soil at relatively rapid rates. Donkin et al. (1995) studied the toxicity<br />

of glucosinolates and their enzymatic breakdown products to Caenorhabditis elegans.<br />

They found that allyl isothyocyanate, one of the decomposition products of the<br />

glucosinolate sinigrin, was three times more toxic to the nematode C. elegans than<br />

corresponding glucosinolate itself.<br />

4.2. Benzaldehyde (benzoic aldehyde)<br />

Benzaldehyde occurs in seeds of bitter almond (Prunus dulcis). It is found naturally<br />

in several cyanogenic glucosides and is used in food and fragrances for its almondlike<br />

aroma and flavor (Harborne and Baxter, 1993). The value of benzaldehyde as a<br />

fungicide is well established (Flor, 1926), with the nematicidal activity more recently<br />

demonstrated. Benzaldehyde reduced populations of M. incognita in field microplots<br />

with no phytotoxicity to cotton at 0.18 -2.14 ml/kg soil (Bauske et al., 1994). The<br />

combination of chitin and benzaldehyde added to peat-based potting mix improved<br />

tomato transplant growth and reduced galling by M. incognita in greenhouse trials<br />

(Kokalis-Burelle et al., 2002). When tested in vitro against M. incognita eggs,<br />

benzaldehyde was 100% effective as an ovicide for this species of root-knot nematode<br />

(Kokalis-Burelle et al., 2002).<br />

The direct effect of benzaldehyde on C. elegans chemotaxis kinetics was analyzed<br />

by Nuttley et al. (2001). An initial attractive response to 100% benzaldehyde was<br />

reported, followed by a strong aversion to the chemical. They determined this behavior<br />

to be mediated by two genetically separable response pathways. Initially, upon<br />

exposure, the attraction response dominates but eventually gives way to a repulsive<br />

response. Oka (2001) found that with juveniles of M. javanica, immobilization and<br />

hatching inhibition in vitro were greater with benzaldehyde and furfural than with<br />

several other essential oils. Benzaldehyde and furfural also reduced galling on tomato<br />

in pot experiments where other aldehydes were not effective (Oka, 2001).<br />

The effects of benzaldehyde combined with several organic amendments on soil<br />

microbial populations and plant-parasitic and nonparasitic nematodes were investigated<br />

by Chavarría-Carvajal et al. (2001). They found that benzaldehyde combined with<br />

most organic amendments reduced damage from parasitic nematodes and selected for<br />

predominantly gram-positive rhizosphere bacteria. When benzaldehyde was combined<br />

with root-knot nematode egg parasitizing isolates of the fungus Fusarium solani,<br />

increasing rates of benzaldehyde in soil reduced nematode penetration and infection<br />

of the host plant, and resulted in increased parasitism of M. javanica females by the<br />

fungus (Siddiqui and Shaukat, 2003). However, the increasing rates of benzaldehyde


24<br />

resulted in lower egg parasitism by the fungus (Siddiqui and Shaukat, 2003).<br />

4.3. Furfural<br />

The biological activity of furfural (also known as 2-furancarboxaldehyde, furaldehyde;<br />

2-furanaldehyde, 2-furfuraldehyde, fural, furfurol) has also been recognized for decades<br />

(Flor, 1926). Furfural is the aldehyde of pyromucic acid and has properties similar to<br />

those of benzaldehyde. Furfural is a derivative of furan and is prepared commercially<br />

by dehydration of pentose sugars obtained from sugarcane, cornstalks and corncobs,<br />

husks of oat and peanut, and other agricultural waste products (Harborne and Baxter,<br />

1993). Commercially available products for disease and nematode control are available<br />

in several countries including the United States. These products include Multiguard<br />

FFA (furfural (75%) + allyl isothiocynate (25%), Harborchem, Cranford, New Jersey),<br />

and Multiguard Protect (furfural (50%) + metham sodium (50%), Harborchem,<br />

Cranford, New Jersey).<br />

Rajendran et al. (2003) reported improved plant growth and reductions in soil<br />

populations of M. arenaria and R. reniformis in groundnut with formulations of furfural<br />

compared to an untreated control, with no effect on free-living nematodes in soil.<br />

Spaull (1997) also found that free-living nematodes were relatively unaffected by<br />

furfural application while parasitic nematodes differed in their susceptibility with<br />

species of Paratichodorus and Xiphinema being more susceptible than Helicotylenchus<br />

and Tylenchorhynchus. Sipes (1997) found furfural to be as effective as 1,3-D for<br />

reduction of preplant soil nematode populations in pineapple production. Rodríguez-<br />

Kábana et al. (1993) found that furfural was an effective nematicide against M.<br />

arenaria, M. incognita, Heterodera glycines, and Pratylenchus spp. on squash, okra,<br />

and soybean. Furfural reduced populations of M. incognita in field microplots with<br />

no phytotoxicity to cotton at 0.18 -2.14 ml/kg soil (Bauske et al., 1994).<br />

4.4. Thymol<br />

NANCY KOKALIS-BURELLE AND RODRÍGO RODRIGUEZ-KÁBANA Thymol (isopropyl-m-cresol) is a volatile, phenolic monoterpene produced by several<br />

plants including thyme (Thymus vulgaris L.) (Baerheim Svendsen and Scheffer, 1985).<br />

Thymol has well-known antiseptic, antifungal, and anthelminthic properties (Wilson<br />

et al., 1977) and is also used for food and fragrance applications (Bauer et al., 1990).<br />

Research by Soler-Serratosa et al. (1996) using combinations of thymol and<br />

benzaldehyde for root-knot and cyst nematode control on soybeans showed that both<br />

compounds exhibited wide spectrum nematicidal activity with Meloidogyne spp. and<br />

Dorylaimid nematodes being more sensitive than cyst nematode and nonparasitic<br />

nematodes (Soler-Serratosa et al., 1996). In addition to the direct toxicity of these<br />

compounds to nematodes, it was hyopothesized that stimulation of beneficial microflora<br />

by the compounds or their products, altered host response, and a deleterious physicochemical<br />

environment may all contribute to reduced gall formation (Soler-Seratosa et<br />

al., 1996).


4.5. Citral<br />

ALLELOCHEMICALS : MANAGEMENT OF PLANT-PARASITIC NEMATODES 25<br />

Citral, is the aldehyde of geraniol, and occurs in the volatile oils of lemon grass,<br />

lemon, orange, limetta, and pimento (Harborne and Baxter, 1993). The flavor of<br />

lemon oil is largely due to its citral content, and the pure aldehyde may be used to<br />

increase the flavoring power of commercial samples of that oil (Harborne and Baxter,<br />

1993). In research trials evaluating the nematicidal potential of citral compared to<br />

other allelopathic chemicals, citral was less nematicidal against M. incognita juveniles,<br />

and more phytotoxic to tomato than benzaldehyde in vitro, and when added to a peatbased<br />

potting mix (Kokalis-Burelle et al., 2002). When tested in vitro against rootknot<br />

nematode eggs, citral reduced egg viability by 80%, but also decreased tomato<br />

seed germination and growth in greenhouse trials (Kokalis-Burelle et al., 2002).<br />

However, when evaluated in soil, citral reduced populations of root-knot nematode<br />

juveniles, galling on roots, and increased cotton growth when applied at 0.1 -0.5 ml/<br />

kg soil in the greenhouse, and at 0.18 -2.14 ml/kg soil in field microplots with no<br />

phytotoxicity to cotton (Bauske et al., 1994). This difference in phytotoxicity may be<br />

due to the difference in host plants tested or to the presence of microorganisms in soil<br />

compared to the relatively uncontaminated conditions that occur in vitro and in potting<br />

media.<br />

5. CONCLUSIONS<br />

The use of allelopathic compounds in either their native, degraded, or processed forms<br />

for plant-parasitic nematode management is receiving increased attention as<br />

agricultural producers face increasing restrictions on chemical biocides and<br />

nematicides. <strong>Allelochemicals</strong> are classified as biochemical pesticides and have been<br />

shown to possess various levels of activity against a wide range of plant-parasitic<br />

nematodes, while exhibiting reduced toxicity to nontarget species and reduced<br />

persistence in soil. The fact that some allelochemicals can be produced in large<br />

quantities in plant material and incorporated into soil as green manures or organic<br />

amendments increases their potential for use as components of nematode management<br />

strategies. The wide array of plant families that produce nematicidal or nematistatic<br />

compounds provides almost unlimited research opportunities for discovery of novel<br />

compounds. Determination of the mechanisms responsible for nematode suppression<br />

with allelopathic compounds also represents a research area with exciting possibilities.<br />

In order to improve the nematicidal and nematistatic efficacy achieved with allelopathic<br />

compounds, a greater understanding of the effects of soil microbiology, soil properties<br />

and environmental conditions on the active compounds is necessary. Currently<br />

available application methods need to be refined and new and improved methods<br />

developed to enhance the performance of allelochemicals for nematode control. Further<br />

investigation is also needed to develop compatible companion applications of<br />

biochemical pesticides and biological control agents, to determine the effects of<br />

allelochemicals on nematode nervous systems, and how they act as nematode attractants<br />

and/or repellents. This research area has enormous potential for discovery of new


26<br />

NANCY KOKALIS-BURELLE AND RODRÍGO RODRIGUEZ-KÁBANA compounds, for elucidating the direct effects of known compounds on nematode<br />

behavior, and for the development of new products to serve as components in<br />

multifaceted approaches to nematode management in the post methyl bromide era.<br />

6. REFERENCES<br />

Adewusi, S.R.A. Turnover of dhurrin in green sorghum seedlings. Plant Physiol 1990; 94:1219-1224.<br />

Akhtar, M. <strong>Biologica</strong>l control of plant-parasitic nematodes by neem products in agricultural soils. Appl Soil<br />

Ecol 1998; 7:219-223.<br />

Akhtar, M. Nematicidal potential of the neem tree Azadirachta indica (A. Juss). Integ Pest Manag Rev 2000;<br />

5:57-66.<br />

Akhtar, M., Alam, M.M. Integrated control of plant-parasitic nematodes on potato with organic amendments,<br />

nematicide, and mixed cropping with mustard. Nematol Medit 1991; 19:169-171.<br />

Akhtar, M., Mahmood, I. Control of plant-parasitic nematodes with ‘Nimin’ and some plant oils by bare rootdip<br />

treatment. Nematol Medit 1993; 21:89-92.<br />

Akhtar, M., Mahmood, I. Control of root-knot nematode by bare-root dip in undecomposed and decomposed<br />

extracts of neem cake and leaf. Nematol Medit 1994; 22:55-57.<br />

Akhtar, M., Mahmood, I. Control of root-knot nematode Meloidogyne incognita in tomato plant by seedcoating<br />

with Achook and neem oil. Integ Pest Control 1995; 37: 86-87<br />

Akhtar, M., Mahmood, I. Organic soil amendments in relation to nematode management with particular reference<br />

to India. Integ Pest Manag Reviews 1996; 1:201-215.<br />

Akhtar, M., Mahmood, I. Control of root-knot nematode Meloidogyne incognita in tomato plants by seed<br />

coating with Suneem and neem oil. J Pesticide Sci 1997; 22:37-38.<br />

Allen, O.N., Allen, E.K. The Leguminosae, a source book of characteristics, uses and nodulation. The<br />

University of Wisconsin Press, Madison, WI, 1981; p. 446.<br />

Baerheim Svendsen, A., Scheffer, J.J.C. Essential oils and aromatic plants. Martinus Nijhoff/Dr. W. Junk<br />

Publishers, Dordrecht, The Netherlands, 1985.<br />

Bauer, K., Garbe, D., Surburg, H. Common fragrance and flavor materials. Preparation, properties, and<br />

Uses. VCH Verlagsgesellschraft, Weinheim, Germany, 1990.<br />

Bauske, E.M., Rodríguez-Kábana, R., Estaun, V., Kloepper, J.W., Robertson, D.G., Weaver, C.F., King, P.S.<br />

Management of Meloidogyne incognita on cotton by use of botanical aromatic compounds. Nematropica<br />

1994; 24:143-150.<br />

Bedigian, D., Harlan, J.R. Evidence for cultivation of sesame in the ancient world. Econ Bot 1986; 40:137-<br />

154.<br />

Bell, E.A., Janzen, D.H. Medical and ecological considerations of L-Dopa and 5 HTP in seeds. Nature 1971;<br />

229:136-137.<br />

Bhatnagar, A., Goswami, B.K. Comparative efficacy of neem and groundnut oilcakes with aldicarb against<br />

Meloidogyne incognita in tomato. Rev Nematol 1987; 10:461-471.<br />

Borek, V., Morra, M.J., Brown, P.D., McCaffrey, J.P. Transformation of the glucosinolate-derived allelochemicals<br />

allyl isothiocyanate and allylnitrile in soil. J of Agricult and Food Chem 1995; 43:1935-1940.<br />

Bridge, J. Nematode management in sustainable and subsistence agriculture. Ann Rev Phytopathol 1996; 34:201-<br />

225.<br />

Brown, P.D., Morra, M.J. Control of soilborne plant pests using glucosinolate-containing plants. Adv in Agron<br />

1997; 61:167-231.<br />

Chavarría-Carvajal, J.A., Rodríguez-Kábana, R., Kloepper, J.W., Morgan-Jones, G. Changes in populations of<br />

microorganisms associated with organic amendments and benzaldehyde to control plant parasitic nematodes.<br />

Nematropica 2001; 31:165-180.<br />

Cheng, H.H. A conceptual framework for assessing allelochemicals in the soil environment. In: Allelopathy,<br />

basic and applied aspects. S. J. H. Rizvi and V. Rizvi, eds. Chapman and Hall, NY, 1992; pp. 21-29.<br />

Czarnota, M.A., Rimando, A.M., Weston, L.A. Evaluation of root exudates of seven sorghum accessions. J<br />

Chem Ecol 2003; 29: 2073-2083.<br />

Donkin, S.G., Eiteman, M.A., Williams, P.L. Toxicity of glucosinolates and their enzymatic decomposition<br />

products to Caenorhabditis elegans. J Nematol 1995; 27:258-262.


ALLELOCHEMICALS : MANAGEMENT OF PLANT-PARASITIC NEMATODES 27<br />

Einhellig, F.A. Allelopathy: Current status and future goals. In: Allelopathy, organisms, processes, and<br />

applications. Inderjit, K.M.M. Dakshini, F.A. Einhellig. eds. Washington, DC: American Chemical Society.<br />

1995; pp. 1-24.<br />

Fahey, J.W., Zalemann, A.T., Talalay, P. The chemical diversity and distribution of glucosinolates and<br />

isothiocyanates among plants. Phytochemistry 2001; 56: 5-51.<br />

Fenwick, G.R., Heaney, R.K., Mullin, W.J. Glucosinolates and their breakdown products in food and plants.<br />

Crit Rev Food Sci Nutrit 1983; 18:123-201.<br />

Ferris H., Zheng, L. Plant sources of Chinese herbal remedies: Effects on Pratylenchus vulnus and Meloidogyne<br />

javanica. J Nematol 1999; 31:241-263.<br />

Flor, H.H. Fungicidal activity of furfural. Iowa State College J Science 1926; 1:199-227.<br />

Fujii, Y., Shibuya, T., Yasuda, T. L-3,4-Dihydroxyphenylalanine as an allelochemical candidate from Mucuna<br />

puriens (L.) DC. var. utilis. Agricult Biolog Chem 1991; 55:617-618.<br />

Gardner, J., Casswell-Chen, E.P., Westerdahl, B., Anderson, C., Lanini, T. The influence of Raphanus sativus,<br />

Sinapis alba, and Phacelia tanacetifolia on California populations of Heterodera schactii. J Nematol<br />

1992; 24:591.<br />

Good, J.M., Minton, N.A., Jaworski, C.A. Relative susceptibility of selected cover crops and coastal bermudagrass<br />

to plant nematodes. Phytopathol 1965; 55:1026-1030.<br />

http://www.epa.gov/pesticides/biopesticides/<br />

Guenzi, W.D., McCalla, I.N. Phenolic acids in oats, wheat, sorghum, and corn residues and their phytotoxicity.<br />

Agronomy J 1966; 58: 303-304.<br />

Halbrendt, J.M. Allelopathy in the management of plant parasitic nematodes. J Nematol 1996; 28:8-14.<br />

Harborne, J.B., Baxter, H. Phytochemical Dictionary. A handbook of bioactive compounds from plants.<br />

Taylor and Francis, London, 1993.<br />

Huettel, R.N. Chemical Communicators in Nematodes. J Nematol 1986; 18:3-8.<br />

Kinloch, R.A., Dunivan, L.S. Summer cropping effects on the abundance of Meloidogyne arenaria race 2 and<br />

subsequent soybean yield. Suppl J Nematol 1993; 25:806-808.<br />

Kokalis-Burelle, N., Martinez-Ochoa, N., Rodríguez-Kabana, R., Kloepper, J.W. Development of multicomponent<br />

transplant mixes for suppression of Meloidogyne incognita on tomato (Lycopersicon<br />

esculentum). J Nematol 2002; 34:362-369.<br />

Lazzeri, L., Tacconi, R., Palmieri, S. In vitro activity of some glucosinolates and their reaction products toward<br />

a population of the nematode Heterodera schachtii. J Agricult Food Chem 1993; 41:825-829.<br />

MacGuidwin, A.E., Layne, T.L. Response of nematode communities to sudangrass and sorghum-sudangrass<br />

hybrids grown as green manure crops. Suppl J Nematol 1995; 27:609-616.<br />

McSorley, R. Host suitability of potential cover crops for root-knot nematodes. Suppl J Nematol 1999; 31:619-<br />

623.<br />

McSorley, R., Dickson, D.W. Effect of tropical rotation crops on Meloidogyne incognita and other plant-parasitic<br />

nematodes. Suppl J Nematol 1995; 27:535-544.<br />

McSorley, R., Frederick, J.J. Response of some common annual bedding plants to three species of Meloidogyne.<br />

Suppl J Nematol 1994; 26:773–777.<br />

McSorley, R., Gallaher, R. Nematode changes and forage yields of six corn and sorghum cultivars. J Nematol<br />

1991; 23:673-677.<br />

McSorley, R., Dickson, D.W., de Brito, J.A. Host status of selected tropical rotation crops to four populations of<br />

root-knot nematodes. Nematropica 1994a; 24:45-53.<br />

McSorley, R., Dickson, D.W., de Brito, J.A., Hochmuth, R.C. Tropical Rotation Crops Influence Nematode<br />

Densities and Vegetable Yields. J Nematol 1994b; 26:308-314.<br />

McSorley, R., Ozores-Hampton, M., Stansly, P.A., Conner, J.M. Nematode management, soil fertility, and yield<br />

in organic vegetable production. Nematropica 1999; 29:205–213.<br />

Minton, N.A., Donnelly, E.D. Additional Vicia species resistant to root-knot nematodes. Plant Dis Report 1967;<br />

51:614-616.<br />

Minton, N.A., Donnelly, E.D., Shepherd, R.L. Reaction of Vicia species and F hybrids from V. sativa x V.<br />

5<br />

angustifolia to five root knot nematode species. Phytopathol 1966; 56:102-107.<br />

Mojtahedi, H., Santo, G.S., Hang, A.N., Wilson, J.W. Suppression of root-knot nematodes with selected rapeseed<br />

cultivars as green manures. J Nematol 1991; 23:170-174.<br />

Mojtahedi, H., Santo, G.S., Ingham, R.E. Suppression of Meloidogyne chitwoodi with sudangrass cultivars as<br />

green manure. J Nematol 1993a; 25: 303-311.


28<br />

NANCY KOKALIS-BURELLE AND RODRÍGO RODRIGUEZ-KÁBANA Mojtahedi, H., Santo, G.S., Wilson, J.H., Hang, A.N. Managing Meloidogyne chitwoodi on potato with rapeseed<br />

as green manure. Plant Dis 1993b; 77:42-46.<br />

Nuttley, W.M., Harbinder, S., van der Kooy, D. Regulation of distinct attractive and aversive mechanisms<br />

mediating benzaldehyde chemotaxis in Caenorhabditis elegans. Learning Memory 2001; 8:170-181.<br />

Oka, Y. Nematicidal activity of essential oil components against the root-knot nematode Meloidogyne javanica.<br />

Nematology 2001; 3: 159-164.<br />

Ploeg, A.T. Greenhouse studies on the effect of marigolds (Tagetes spp.) on four Meloidogyne species. J Nematol<br />

1999; 31:62-69.<br />

Potter, M.J., Davies, K., Rathjen, A.J. Suppressive impact of glucosinolates in Brassica vegetative tissues on<br />

root lesion nematode Pratylenchus neglectus. J Chem Ecol 1998; 24:67-80.<br />

Prot, J.C., Soriano, I.R.S., Mantias, D.M., Savary, S. Use of green manure crops in control of Hirschmanniella<br />

mucronata and H. oryzae in irrigated rice. J Nematol 1992; 24:127-132.<br />

Putnam, A.R., Tang, C. Allelopathy: State of the science. In The science of allelopathy. A.R. Putnam & C.<br />

Tang, eds. New York: John Wiley, 1986, pp. 1-19.<br />

Quiroga-Madrigal, R., Rodríquez-Kábana, R., Robertson, D.G., Weaver, C.F., King, P.S. Nematode populations<br />

and enzymatic activity in rhizospheres of tropical legumes in Auburn, Alabama. Nematropica 1999;<br />

29:129 (Abstr.).<br />

Rajendran, G., Ramakrishnan, S., Subramanian, S. Cropguard – a botanical nematicide for the management of<br />

Meloidogyne arenaria and Rotylenchulus reniformis in groundnut. In: Proceedings of National<br />

Symposium on Biodiversity and Management of Nematodes in Cropping Systems for Sustainable<br />

Agriculture, Jaipur, India, 2003, pp. 122-125.<br />

Reeves, D.W., Mansoer, Z., Wood, C.W. Suitability of sunn hemp as an alternative legume cover crop. In:<br />

Proceedings of the new technology and conservation tillage, vol. 96-07. Jackson, TN: University of<br />

Tennessee Agricultural Experiment Station, 1996, pp. 125-130.<br />

Robinson, A.F., Cook, C.G., Bridges, A.C. Comparative reproduction of Rotylenchulus reniformis and<br />

Meloidogyne incognita race 3 on kenaf and sunn hemp grown in rotation with cotton. Nematropica<br />

1998; 28:143 (Abstr.).<br />

Rodríguez-Kábana, R., King, P.S., Robertson, D.G., Weaver, C.F. Potential of crops uncommon to Alabama for<br />

management of root-knot and soybean cyst nematodes. Suppl J Nematol 1988a; 20:116-120.<br />

Rodríguez-Kábana, R., Robertson, D.G., Wells, L., Young, R.W. Hairy indigo for the management of<br />

Meloidogyne arenaria in peanut. Nematropica 1988b; 18:137-142.<br />

Rodríguez-Kábana, R., Robertson, D.G., Wells, L., King, P.S., Weaver, C.F. Crops uncommon to Alabama for<br />

the management of Meloidogyne arenaria in peanut. Suppl J Nematol 1989; 21:712-716.<br />

Rodríguez-Kábana, R., Robertson, D.G., King, P.S., Wells, L. American jointvetch and partridge pea for the<br />

management of Meloidogyne arenaria in peanut. Nematropica 1991a; 21:97-103.<br />

Rodríguez-Kábana, R., Robertson, D.G., Weaver, C.F., Wells, L. Rotations of bahiagrass and castorbean with<br />

peanut for the management of Meloidogyne arenaria. Suppl J Nematol 1991b; 23:658-661.<br />

Rodríguez-Kábana, R., Kloepper, J.W., Robertson, D.G., Wells, L.W. Velvetbean for the management of rootknot<br />

and southern blight in peanut. Nematropica 1992a; 22:75-80.<br />

Rodríguez-Kábana, R., Pinochet, J., Robertson, D.G., Weaver, C.F., King, P.S. Horsebean (Canavalia ensiformis)<br />

and crotalaria (Crotalaria spectabilis) for the management of Meloidogyne spp. Nematropica 1992b;<br />

22:29-35.<br />

Rodríguez-Kábana, R., Kloepper, J.W., Weaver, C.F., Robertson, D.G. Control of plant parasitic nematodes<br />

with furfural – a naturally occurring fumigant. Nematropica 1993; 26:63-73.<br />

Rodríguez-Kábana, R., Kokalis-Burelle, N., Robertson, D.G., Wells, L.W. Evaluation of sesame for control of<br />

Meloidogyne arenaria and Sclerotium rolfsii in peanut. Nematropica, 1994a; 24:55-61.<br />

Rodríguez-Kábana, R., Kokalis-Burelle, N., Robertson, D.G., Wells, L., King, P.S. Rotations with Coastal<br />

Bermudagrass, Cotton, and Bahiagrass for the Management of Root-Knot Nematode and Southern Blight<br />

in Peanut. Suppl J Nematol 1994b; 26:665-668.<br />

Rodríguez-Kábana, R., Kokalis-Burelle, N., Robertson, D.G., Weaver, C.F., Wells, L. Effects of partridge peapeanut<br />

rotations on populations of Meloidogyne arenaria, incidence of Sclerotium rolfsii, and yield of<br />

peanut. Nematropica 1995; 25:27-34.<br />

Rodríguez-Kábana, R., Kloepper, J.W. Cropping systems and the enhancement of microbial activities antagonistic<br />

to nematodes. Nematropica 1998; 28:144 (Abstr.).


ALLELOCHEMICALS : MANAGEMENT OF PLANT-PARASITIC NEMATODES 29<br />

Rotar, P.P., Joy, R.J. ‘Tropic Sun’ sunn hemp. Crotalaria juncea L. Research Extension Series 036, University<br />

of Hawaii, Honolulu, 1983.<br />

Sasanelli, N., M. DiVito. The effect of Tagetes spp. extracts on the hatching of an Italian population of Globodera<br />

rostochiensis. Nematol Medit 1991; 19:135-137.<br />

Siddiqui, M.A., Alam, M.M. Control of root-knot and reniform nematodes by bare-root dip in leaf extract or<br />

margosa and persian lilac. Zeitsch. Pflazenkh Pflazen Sch 1988; 95:138-142.<br />

Siddiqui, I.A., Shaukat, S.S. Factors influencing the effectiveness of non-pathogenic Fusarium solani strain Fs5<br />

in the suppression of root-knot nematode in tomato. Phytopathol. Mediterran, 2003; 42:17-26.<br />

Sipes, B.S. Pre-plant and post-plant pesticides for nematode control in pineapple. Acta Horticult 1997; 425:457-<br />

464.<br />

Sipes, B.S., Arakaki, A.S. Root-knot nematode management in dryland taro with tropical cover crops. Suppl J<br />

Nematol 1997; 29:721–724.<br />

Soler-Serratosa, A., Kokalis-Burelle, N., Rodriguez-Kabana, R., Weaver, C.F., King, P.S. <strong>Allelochemicals</strong> for<br />

control of plant-parasitic nematodes. 1. In vivo nematicidal efficacy of thymol and thymol/benzaldehyde<br />

combinations. Nematropica 1996; 26: 57-71.<br />

Spaull, V.W. On the use of furfural to control nematodes in sugarcane. Proceedings of the Annual Congress –<br />

South African Sugar Technologists Association 1997; 71:96.<br />

Starr, J.L., Black, M.C. Reproduction of Meloidogyne arenaria, M. incognita, and M. javanica on sesame.<br />

Suppl J Nematol 1995; 27: 624-627.<br />

Steiner, G. Proc. Proc Biol Soc Washington 1941; 54:31-34.<br />

Stirling, G.R. <strong>Biologica</strong>l Control of Plant-Parasitic Nematodes. Progress, Problems, and Prospects. CAB<br />

International, Wallingford, UK, 1991; pp. 212.<br />

Taylor, C.R., Rodriguez-Kabana, R. Population dynamics and crop yield effects of nematodes and white mold<br />

in peanut, cotton and velvetbean. Agricult Syst 1999; 59:177-191.<br />

Thakur, R.S., Singh, S.B., Goswami, A. Azadirachta indica A. Juss. – A Review. CROMAP 1981; 3:135-140.<br />

Thomas, W.B. Methyl bromide: Effective pest management tool and environmental threat. J Nematology 1996;<br />

28:286-289.<br />

Tyler, J. Proceeding of the root-knot nematode conference, Atlanta, GA, Feb. 4, 1938. Plant Dis Report Suppl<br />

1938; 109:133-151.<br />

Uhlenbroek, J.H., Bijloo, J.D. Investigations on nematicides. I. Isolation and structure of a nematicidal principle<br />

occurring in Tagetes roots. Recuceil des Travaux Chimiques des Pays-Bas 1958; 77:1004-1008.<br />

Uhlenbroek, J.H., Bijloo, J.D. Investigations on nematicides. II. Structure of a second nematicidal principle<br />

isolated from Tagetes roots. Recuceil des Travaux Chimiques des Pays-Bas 1959; 78:382-390.<br />

Vargas-Ayala, R., Rodriguez-Kabana, R. Bioremediative management of soybean nematode population densities<br />

in crop rotations with velvetbean, cowpea, and winter crops. Nematropica 2001; 31:37-46.<br />

Venette, R.C., Monstafa, F.A.M., Ferris, H. Trophic interactions between bacterial-feeding nematodes in plant<br />

rhizospheres and the nematophagus fungus Hirsutella rhossiliensis to suppress Heterodera schachtii.<br />

Plant Soil 1997; 191:213–223.<br />

Vincente, N.E., Acosta, N. Effects of Mucuna deeringiana on Meloidogyne incognita. Nematropica 1987;<br />

17:99-102.<br />

Waller, G.R. <strong>Allelochemicals</strong>: Role in agriculture and forestry. Washington, DC: American Chemical Society,<br />

1985.<br />

Wang, K.H., Sipes, B.S., Schmitt, D.P. Suppression of Rotylenchulus reniformis by Crotalaria juncea, Brassica<br />

napus, and Target erecta. Nematropica 2001; 31:237–251.<br />

Wang, K.H., McSorley, R., Gallaher, R.N. Effect of Crotalaria juncea amendment on nematode communities<br />

in soil with different agricultural histories J Nematol 2003; 35:294–301.<br />

Weaver, D.B., Rodríguez-Kábana, R., Carden, E.L. Velvetbean in rotation with soybean for management of<br />

Heterodera glycines and Meloidogyne arenaria. Suppl J Nematol 1993; 25:809-813.<br />

Wilson, C.O., Grisvold, O., Doerge, R.F. Textbook of organic medicinal and pharmaceutical chemistry. J.B.<br />

Lippincott Co., Philadelphia, PA, USA, 1977.<br />

Yasuhira, T., Nagase, A., Kuwahara, Y., Sugawara, R. Aggregation of Bursaphelenchus lignicolus (Nematoda:<br />

Aphelenchoididae) to several compounds containing oleyl group. Appl Entomol & Zool 1982; 17:46-51.<br />

Zasada, I.A., Ferris, H. Nematode suppression with brassicaceous amendments: application based upon<br />

glucosinolate profiles. Soil Biol Biochem 2004; 36:1017-1024.


ANA LUISA ANAYA<br />

ALLELOPATHIC ORGANISMS AND MOLECULES:<br />

PROMISING BIOREGULATORS FOR THE<br />

CONTROL OF PLANT DISEASES, WEEDS,<br />

AND OTHER PESTS<br />

Laboratorio de Alelopatía, Instituto de Ecología, Universidad Nacional Autónoma<br />

de México. Circuito Exterior, Ciudad Universitaria. 04510 México, D.F.<br />

Email: alanaya@miranda.ecologia.unam.mx<br />

Abstract. Increasing attention has been given to the role and potential of allelopathy as a management strategy<br />

for crop protection against weeds and other pests. Incorporating allelopathy into natural and agricultural management<br />

systems may reduce the use of herbicides, insecticides, and other pesticides, reducing environment/soil<br />

pollution and diminish autotoxicity hazards. There is a great demand for compounds with selective toxicity that<br />

can be readily degraded by either the plant or by the soil microorganisms. In addition, plant, microorganisms,<br />

other soil organisms and insects can produce allelochemicals which provide new strategies for maintaining and<br />

increasing agricultural production in the future.<br />

1. INTRODUCTION<br />

AGRICULTURE AND PEST MANAGEMENT SYSTEMS<br />

Agriculture is one of the world’s largest industries. On a worldwide basis, more people<br />

are in some way involved in agriculture than in all other occupations combined.<br />

Agriculture is also United States’ largest industry. This country produces more food<br />

than any other nation in the world and is the world’s largest exporter of agricultural<br />

products. According to the 2002 survey from the United States Department of<br />

Agriculture (USDA) National Agricultural Statistics Service, there are more than 941<br />

million acres used for farming in the U.S. with the average farm size being 436 acres.<br />

Agriculture in the United States is becoming more productive. In 1935, there<br />

were 6.8 million farms in the United States, and the average farmer produced enough<br />

food to feed 20 people. In 2002, the number of farms was estimated to be 2.16 million,<br />

and the average U.S. farmer produced enough food each year to feed more than<br />

100 people. In addition to providing an abundant food supply for domestic markets,<br />

crops from nearly one-third of U.S. farm acreage are exported to overseas customers<br />

(IFIC 2004).<br />

Pest problems and their management vary widely throughout the world, based on<br />

economical resources, cultural techniques, climate, soil types, and many other<br />

conditions. As a result, chemical pest control has won a central place in modern<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 31– 78.<br />

© 2006 Springer. Printed in the Netherlands.<br />

31


32<br />

ANA LUISA ANAYA<br />

agriculture, contributing to the dramatic increase in crop yields achieved in recent<br />

decades for most major field, fruit, and vegetable crops.<br />

Farmers must contend with approximately 80,000 plant diseases, 30,000 species<br />

of weeds, 1,000 species of nematodes, and more than 10,000 species of insects. Today,<br />

national and international agricultural organizations estimate that as much as 45<br />

percent of the world’s crops continue to be lost to these types of hazards. In the United<br />

States alone, about $20 billion worth of crops (one-tenth of production) are lost each<br />

year (IFIC 2004).<br />

The word “pesticides” refers to a broad class of crop protection chemicals,<br />

including four major groups: insecticides, rodenticides, herbicides, and fungicides.<br />

All pesticides must be toxic, or poisonous, to kill the pests they are intended to control.<br />

Because pesticides are toxic, they are potentially hazardous to humans and animals.<br />

Therefore, people who use pesticides or regularly come in contact with them must<br />

understand the relative toxicity and potential health effects of the products they use<br />

(Pesticide Education Program. Penn State’s College of Agricultural Sciences, 2004).<br />

Some pesticides (administered at extremely high dosages) have been found to cause<br />

cancer in laboratory animals (Agri 21FAO).<br />

Weeds represent the most serious threat to the sustainability and profitability of<br />

agricultural production around the world considering these facts:<br />

Weeds represent the most economically serious pest complex reducing world<br />

food and fiber production.<br />

Controlling weeds costs the United States economy more than $15 billion annually,<br />

surpassing the combined cost of controlling disease and insect pests. U.S. herbicide<br />

expenditures account for 85% of all pesticide purchases and 62% of the total<br />

amount of active ingredient applied.<br />

Nearly all (greater than 95%) of the corn and soybean acreage in the U.S. receives<br />

herbicide applications. In developing countries, the cost of weed control in terms<br />

of labor and loss of yields is even greater, proportionally, than in the U.S.<br />

Worldwide herbicide purchases ($16.9 billion in 1997) constitute 58% of all<br />

pesticides bought and 53% of the amount of active ingredient applied (1 billion<br />

kg a.i.) worldwide.<br />

Crop yield losses from weed competition can be substantial. The degree of loss<br />

depends on, crop and weed species present; timing and duration of competitive<br />

interactions; and resource availability (Agri 21 FAO*, IFIC 2004**).<br />

Worldwide, the competitive effect of weeds causes a 10% loss in agricultural<br />

production. Yield losses in rice and other grass crops in West Africa have been reported<br />

to range from 28-100% if weeds such as witchweed (Striga hermonthica) –a parasitic<br />

weed–are not controlled; the greatest reductions occur on nutrient-poor soils. Left<br />

unchecked, weeds cause dramatic reductions in food production that eventually can<br />

* Agri 21 FAO: Agriculture 21. IPM and Weed Management. Food and Agriculture Organization of the<br />

United Nations (FAO). Agriculture Department. 2004. http://www.fao.org/WAICENT/FAOINFO/AGRICULT/<br />

Default.htm.<br />

** IFIC 2004 : International Food Information Council. 2004. Agriculture & Food Production. Background<br />

on Agriculture & Food Production. http://www.ific.org/food/agriculture/index.cfm.


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

destabilize economic and social systems. Hence, there is an urgent need to develop<br />

and refine weed management strategies in crop production that are effective, safe,<br />

and economically viable.<br />

Regardless of cropping system or agricultural region of the world, effective weed<br />

management strategies are needed continually to maintain crop yields and crop quality<br />

as well as to reduce the negative impact of weeds in future years. This ongoing quest<br />

to optimize weed management strategies is largely because of the ability of agricultural<br />

weeds to adapt to many of our most important crop production systems (Agri 21 FAO,<br />

IFIC 2004).<br />

Integrated Pest Management (IPM) is a system that works in partnership with<br />

nature to produce foods efficiently (Upadhyay et al., 1996). The concept began in<br />

U.S.A. in the 1950s, and recently resurfaced in popularity. Although many definitions<br />

of IPM have been advanced, two elements are critical:<br />

using multiple control tactics;<br />

integrating a knowledge of pest biology into the management system.<br />

IPM involves the carefully managed use of an array of pest control techniques<br />

including biological, cultural, and appropriate chemical methods to achieve the best<br />

results with the least disruption of the environment. With IPM, growers are adopting<br />

less chemically intensive methods of farming, which may include pest-resistant plant<br />

varieties, adjustments in planting times, low tillage, and other non-chemical techniques.<br />

The objectives of IPM are:<br />

Appreciate the importance of controlling weeds within an integrated pest<br />

management program.<br />

Understand the key biological differences that make IPM for weeds more difficult<br />

than IPM for insects.<br />

Become familiar with integrated weed management (IWM) strategies.<br />

The United States Department of Agriculture (USDA) proposed national standards<br />

for organic farming and handling in 1997.The federal regulations for organic<br />

standards were finalized in 2001 and began full implementation in 2002. Generally,<br />

organic food is produced by farmers who emphasize the use of renewable resources<br />

and the conservation of soil and water to enhance environmental quality for future<br />

generations (IFIC 2004, Agri 21 FAO).<br />

Barberi (2002), assessed that despite the serious threat which weeds offer to organic<br />

crop production, relatively little attention has so far been paid to research on<br />

weed management in organic agriculture, an issue that is often approached from a<br />

reductionist perspective. Compared with conventional agriculture, in organic agriculture<br />

the effects of cultural practices (e.g. fertilization and direct weed control) on<br />

crop-weed interactions usually manifest themselves more slowly. Weed management<br />

should be tackled in an extended time domain and needs deep integration with the<br />

other cultural practices, aiming to optimize the whole cropping system rather than<br />

weed control per se. Many organic farmers are aware that successful weed management<br />

implies putting into practice the concept of maximum diversification of their<br />

cropping system. However, this task is often difficult to achieve, because practical<br />

solutions have to pass through local filters (soil and climate conditions, availability of<br />

33


34<br />

ANA LUISA ANAYA<br />

and accessibility to external inputs -seeds, crop cultivars, machinery, etc.) and socioeconomic<br />

constraints (market, tenure status, attitude towards entrepreneurial risk,<br />

etc.). The use of cover crops and organic amendments, via the promotion of diversity<br />

in insect, fungal, bacterial or mychorrhyzal communities, may alter antagonist or<br />

competitive effects to the benefit of crops and to the detriment of weeds. Once factors<br />

driving these effects are better understood, it might be possible to use this knowledge<br />

to improve organic weed management systems locally. It would also be helpful to find<br />

indicators of “functional biodiversity”, where weed species abundance is assessed on<br />

the role that they have in the agroecosystem (e.g. strong /weak competitors, promoters<br />

of the presence of beneficial arthropods, etc.). Management of allelopathy is another<br />

potential tool in the arsenal of the organic farmer (Barberi, 2002). In the United<br />

States, the rate of increase of organic growers was estimated at 12% in 2000. However,<br />

many producers are reluctant to undertake the organic transition because of<br />

uncertainty of how organic production will affect weed population dynamics and<br />

management. The organic transition has a profound impact on the agroecosystem.<br />

Changes in soil physical and chemical properties during the transition often impact<br />

indirectly insect, disease, and weed dynamics. Greater weed species richness is usually<br />

found in organic farms but total weed density and biomass are often smaller<br />

under the organic system compared with the conventional system. The improved weed<br />

suppression of organic agriculture is probably the result of combined effects of several<br />

factors including weed seed predation by soil microorganisms, seedling predation by<br />

phytophagus insects, and the physical and allelopathic effects of cover crops (Ngouajio<br />

and McGiffen, 2002).<br />

2. ALLELOPATHY<br />

Increasing attention has been given to the role and potential of allelopathy as a management<br />

strategy for crop protection against weeds and other pests. Incorporating<br />

allelopathy into natural and agricultural management systems may reduce the use of<br />

herbicides, fungicides, nematicides, and insecticides, cause less pollution and diminish<br />

autotoxicity hazards. There is a great demand for compounds with selective toxicity<br />

that can be readily degraded by either the plant or by the soil microorganisms.<br />

Plant, microorganisms, other soil organisms and insects can produce allelochemicals<br />

which provide new strategies for maintaining and increasing agricultural production<br />

in the future. Compounds with allelopathic activity may provide novel chemistry for<br />

the synthesis of herbicides, insecticides, nematicides, and fungicides that are not based<br />

on the persistent petroleum derived compounds which are such a public health concern<br />

(Waller and Chou, 1989; Waller, 1999).<br />

Several crops (some of which can be used as cover crops) have been proved to<br />

release allelopathic compounds in the soil (Jimenez-Osornio and Gliessman, 1987;<br />

Blum et al., 1997; Inderjit and Keating, 1999; Anaya, 1999), many of which have<br />

been chemically characterized (Pereda-Miranda et al., 1996; Inderjit, 1996; Seigler,<br />

1996; Waller et al., 1999). The idea of exploiting these compounds as natural herbicides<br />

is therefore very attractive (Putnam, 1988; Weston, 1996; Duke et al., 2000).


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

However, the large majority of the studies carried out on this topic have referred to<br />

reductionist trials carried out in controlled environments, often with the only aim to<br />

extract and characterize allelochemicals or, at the most, to test the effect of these<br />

compounds on the germination of selected sensitive species in bioassays. In the case<br />

of crop-weed interactions, absolute evidence of the occurrence of allelopathy in the<br />

field is difficult to obtain, mainly because allelopathic effects are difficult to disentangle<br />

from resource competition and other biotic effects (Weidenhamer, 1996; Inderjit<br />

and del Moral, 1997). Additionally, the production and release of allelochemicals<br />

depend largely upon environmental conditions, usually being higher when plants are<br />

under stress, e.g. extreme temperatures, drought, soil nutrient deficiency, high pest<br />

incidence (Einhellig, 1987); also, the range and concentration of chemicals that a<br />

given species can produce can vary with environment conditions (Anaya, 1999). Other<br />

effects that need to be examined are allelopathy-mediated weed-weed, weed-crop and<br />

crop-following (or companion) crop interactions. It is therefore questionable whether<br />

allelopathy management per se would ever represent a consistently effective weed<br />

management tool; however, a better understanding of allelopathic occurrence in field<br />

situations, and of how it is influenced by cultural practices, would make it possible to<br />

include allelopathic crops in organic cropping systems and use them as a complementary<br />

tactic in a weed management strategy (Barberi, 2002).<br />

The extracts of many dominant plants in Taiwan, such as Delonix regia, Digitaria<br />

decumbens, Leucaena leucocephala, and Vitex negundo, contain allelopathic<br />

compounds, including phenolic acids, alkaloids, and flavonoids that can be used as<br />

natural herbicides, fungicides, etc. which are less disruptive of the global ecosystem<br />

than are synthetic agrochemicals (Chou, 1995). Many important crops, such as rice,<br />

sugarcane, and mungbean, are affected by their own toxic exudates or by phytotoxins<br />

produced when their residues decompose in the soil. For example, in Taiwan the yield<br />

of the second annual rice crop is typically 25% lower than that of the first, due to<br />

phytotoxins produced during the fallowing period between crops. Autointoxication<br />

can be minimized by eliminating, or preventing the formation of the phytotoxins<br />

through field treatments such as crop rotation, water draining, water flooding, and<br />

the polymerization of phytotoxic phenolics into a humic complex (Chou, 1995).<br />

Wetland soils provide anoxia-tolerant plants with access to ample light, water,<br />

and nutrients. Intense competition, involving chemical strategies, ensues among the<br />

plants. The roots of wetland plants are prime targets for root-eating pests, and the<br />

wetland rhizosphere is an ideal environment for many other organisms and communities<br />

because it provides water, oxygen, organic food, and physical protection. Consequently,<br />

the rhizosphere of wetland plants is densely populated by many specialized<br />

organisms, which considerably influence its biogeochemical functioning. The roots<br />

protect themselves against pests and control their rhizosphere organisms by bioactive<br />

chemicals, which often also have medicinal properties. Anaerobic metabolites, alkaloids,<br />

phenolics, terpenoids, and steroids are bioactive chemicals abundant in roots<br />

and rhizospheres in wetlands. Bioactivities include allelopathy, growth regulation,<br />

extraorganismal enzymatic activities, metal manipulation by phytosiderophores and<br />

35


36<br />

ANA LUISA ANAYA<br />

phytochelatines, various pest-control effects, and poisoning. Complex biological-biochemical<br />

interactions among roots, rhizosphere organisms, and the rhizosphere solution<br />

determine the overall biogeochemical processes in the wetland rhizosphere and<br />

in the vegetated wetlands. In order to comprehend how wetlands really function and<br />

to understand these interactions it is necessary to implement long-term collaborative<br />

research (Neori et al., 2000). We can find promising allelochemicals and useful interactions<br />

in the rich biodiversity of these particular ecoystems, but without doubt, in all<br />

type of ecosystems.<br />

3. CHEMISTRY OF ALLELOPATHY<br />

As we know, plant and microbial compounds are continuously analyzed as potential<br />

sources of herbicides, pesticides, and pharmaceuticals because they provide a diversity<br />

of carbon skeletons and there has been success in that a number of compounds<br />

have shown biological activity. The same bioassays and techniques to reveal mechanisms<br />

of action apply to the search for herbicides as in the study of allelopathy. Certainly<br />

there is overlap in goals and compounds studied, but there also is a difference<br />

in that the starting point in the ‘herbicide search’ might be any natural product, as<br />

opposed to one identified with allelopathy. Most inhibitors of plants are secondary<br />

compounds that have their origin in either the shikimate or acetate pathways, or they<br />

are compounds having skeletal components from both of these origins (Einhellig,<br />

2002). Waller et al. (1999) listed over twenty classes of secondary metabolites that are<br />

produced, stored, and released into the rhizosphere where they have biological activity<br />

as well as undergo microbial transformation and degradation. Einhellig (2002)<br />

concluded that the 14 categories suggested by Rice (1984) are sufficiently broad to<br />

still retain validity: water-soluble organic acids, straight-chain alcohols, aliphatic<br />

aldehydes and ketones, unsatured lactones, long-chain fatty acids and polyacetylenes,<br />

naphthoquinones, anthraquinones and complex quinones, gallic acids and<br />

polyphloroglucinols, cinnamic acids, coumarins, flavonoids, tannins, terpenoids, and<br />

steroids, amino acids, and purines and nucleosides.<br />

In this chapter some of the main compounds and studies associated with allelopathy<br />

will be mentioned.<br />

Terpenoids and phenolics are the most common compounds involved in allelopathic<br />

interactions. Terpenoids are the largest group of plant chemicals (15,000-20,000)<br />

with a common biosynthethic origin. The terpenoid pathway generates great structural<br />

diversity and complexity of compounds, thus generating enormous potential for<br />

mediating ecological interactions (Duke, 1991; Langenheim, 1994). Terpenoids may<br />

produce effects on seeds and soil microbiota through volatilization, leaching from<br />

plants, or decomposition of plant debris. These interactions can significantly affect<br />

community and ecosystem properties, although studies of plant-plant chemical interactions<br />

have often been controversial because of difficulty in unambiguously demonstrating<br />

interference by chemical inhibition rather than through resource competition<br />

or other mechanisms (Harper, 1977).


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

3.1. Terpenoids and sesquiterpene lactones<br />

Vokou et al. (2003) compared the potential allelopathic activity of 47 monoterpenoids<br />

of different chemical groups, by estimating their effect on seed germination and<br />

subsequent growth of Lactuca sativa seedlings. Apart from individual compounds,<br />

eleven pairs at different proportions were also tested. As a group, the hydrocarbons,<br />

except for (+)-3-carene, were the least inhibitory. Of the oxygenated compounds, the<br />

least inhibitory were the acetates; whenever the free hydroxyl group of an alcohol<br />

turned into a carboxyl group, the activity of the resulting ester was markedly lower<br />

(against both germination and seedling growth). Twenty-four compounds were<br />

extremely active against seedling growth (inhibiting it by more than 85%), but only<br />

five against seed germination. The compounds that were most active against both<br />

processes belonged to the groups of ketones and alcohols; they were terpinen-4-ol,<br />

dihydrocarvone, and two carvone stereoisomers. These authors used a model to<br />

investigate whether compounds acted independently when applied in pairs. The<br />

combined effect varied. In half of the cases, it followed the pattern expected under the<br />

assumption of independence; in the rest, either synergistic or antagonistic interactions<br />

were found in both germination and elongation. However, even in cases of synergistic<br />

interactions, the level of inhibition was not comparable to that of a single extremely<br />

active compound, unless such a compound already participated in the combination.<br />

The effect of the sesquiterpene cacalol and extracts (water and petroleum ether)<br />

derived from the roots of Psacalium decompositum (Asteraceae) on the germination<br />

and radicle growth of two plants, Amaranthus hypochondriacus (Amaranthaceae)<br />

and Echinochloa crus-galli (Poaceae), and the radial growth of four phytopathogenic<br />

fungi was described (Anaya et al., 1996). The activity of two cacalol derivatives (methyl<br />

cacalol and cacalol acetate) was also investigated. Germination of A. hypochondriacus<br />

was inhibited by almost all the treatments. The extracts and cacalol produced a<br />

significant inhibition of radicle growth of A. hypochondriacus and E. crus-galli.<br />

Cacalol acetate showed a specific inhibition on E. crus-galli, and methyl cacalol<br />

inhibited significantly the growth of A. hypochondriacus. In general, antifungal activity<br />

depended upon the target fungi and the concentration of each treatment. Cacalol had<br />

also effects on the morphology and coloration of the fungal mycelium. The bioactivity<br />

shown by the extracts of Psacalium decompositum on the tested seeds and fungi is<br />

mainly due to their content in cacalol.<br />

The allelochemical potential of Callicarpa acuminata (Verbenaceae) was<br />

investigated using a biodirected fractionation study as part of a long-term project to<br />

search for bioactive compounds among the rich biodiversity of plant communities in<br />

the Ecological Reserve El Eden, Quintana Roo, Mexico. Aqueous leachate, chloroformmethanol<br />

extract, and chromatographic fractions of the leaves of the plant species<br />

inhibited the root growth of Amaranthus hypochondriacus, Echinochloa crus-galli,<br />

and tomato (23% , 59%, and 70% respectively). Some of these treatments caused a<br />

moderate inhibition of the radial growth of two phytopathogenic fungi,<br />

Helminthosporium longirostratum and Alternaria solani (18% to 31%). The<br />

chloroform-methanol (1:1) extract prepared from the leaves rendered five compounds:<br />

37


38<br />

ANA LUISA ANAYA<br />

isopimaric acid, a mixture of two diterpenols: sandaracopimaradien-19-ol and<br />

akhdarenol, α-amyrin, and the flavone salvigenin. The phytotoxicity exhibited by<br />

several fractions and the full extract almost disappeared when pure compounds were<br />

evaluated on the test plants, suggesting a synergistic or additive effect. Akhdarenol,<br />

α-amyrin and isopimaric acid methyl ether had antifeedant effects on Leptinotarsa<br />

decemlineata. Alpha-amyrin was most toxic to this insect. No correlation was found<br />

between antifeedant and toxic effects on this insect, suggesting that different modes<br />

of action were involved. All the test compounds were cytotoxic to insect Sf9 cells<br />

while salvigenin, akhdarenol, and isopimaric acid also affected mammalian Chinese<br />

Hamster Ovary (CHO) cells. Alpha-amyrin showed the strongest selectivity against<br />

insect cells (Anaya et al., 2003). In this study the authors emphasized that<br />

allelochemicals involved in allelopathic interactions often have multiple functionality.<br />

Sesquiterpene lactones (SL) occur in over 15 plant families, predominantly in<br />

the Asteraceae, and represent with about 3,500 naturally occurring compounds, one<br />

of the largest groups of natural products. It has been demonstrated that some<br />

sesquiterpene lactones exhibit a broad spectrum of biological activities including<br />

phytotoxic and plant growth regulatory properties, cytotoxicity and antitumor<br />

properties, antimicrobial, insecticidal, molluscicidal and antimalarial activity (Fischer,<br />

1986). Phytotoxic terpenoids and their possible involvement in allelopathy were covered<br />

in reviews on mono- and sesquiterpenes (Evenari, 1949; Fischer, 1986, 1991, 1994)<br />

and biological activities of SL were reviewed by Stevens and Merrill (1985), Picman<br />

(1986) and Elakovich (1988). Seedlings of Ambrosia cumanensis are inhibited by<br />

leachates of the adult plants and residues in soils. Some SL of this species have been<br />

implicated in this autototoxic mechanism (Anaya and del Amo, 1978). In the same<br />

way, parthenin and coronopilin of Parthenium hysterophorus also exhibited autotoxicity<br />

toward seedlings and older plants, this fact possibly reveal a mechanism of intraspecific<br />

population regulation (Picman and Picman, 1984). Axivalin and tomentosin from the<br />

seeds of Iva axillaris were inhibitory toward the germination and growth of Abutilon<br />

theophrasti (velvetleaf) (Spencer et al., 1984). The germacranolide-type SL represented<br />

by dihydrotartridin B significantly inhibited the root growth of Brassica rapa var.<br />

pervidis (Sashida et al., 1983). The α-methylene-γ-lactone group is present in many<br />

of the isolated natural sesquiterpene lactones, and has been proposed as one of the<br />

factors which can determine their allelopathic activity, in particular, as well as their<br />

biological activity in general. The different spatial arrangements that a molecule of<br />

SL can adopt is the other factor that has been related with the potential allelopathic<br />

activity of this type of secondary compounds (Macias et al., 1992). Data of several<br />

studies on the allelopathic potential of SL clearly demonstrated that they can selectively<br />

promote or inhibit germination or growth at concentrations as low as 1 µM. It is<br />

reasonable to assume that rain washes transport SL from the source plant or<br />

decomposing litter into the soil where they can reach significant concentration levels.<br />

In the case of isoalantolactone, it has been demonstrated that it can persist in mineral<br />

and organic soil for 3 months, supporting the assumption that SL play a significant<br />

role in allelopathic interactions in the environment (del Amo and Anaya, 1976; Stevens<br />

and Merril, 1985; Picman, 1986; Fischer, 1991).


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

Dehydrozaluzanin C, a natural SL, is a weak plant growth inhibitor with an I 50<br />

(or IC 50 , the concentration required to inhibit plant growth 50 %) of about 0.5 mM for<br />

lettuce root growth. It also causes rapid plasma membrane leakage in cucumber<br />

cotyledon discs. Dehydrozaluzanin C is more active at 50 µM than the same<br />

concentration of the herbicide acifluorfen. Symptoms include plasmolysis and the<br />

disruption of membrane integrity is not light dependent. Reversal of its effects on root<br />

growth was obtained with treatment by various amino acids, with histidine and glycine<br />

providing ca. 40% reversion. The strong reversal effect obtained with reduced<br />

glutathione is due to cross-reactivity with DHZ and the formation of mono- and diadducts.<br />

Photosynthetic, respiratory and mitotic processes, as well as NADH oxidase<br />

activity appear to be unaffected by this compound. Dehydrozaluzanin C exerts its<br />

effects on plants through two different mechanisms, only one of which is related to<br />

the disruption of plasma membrane function (Galindo et al., 1999).<br />

A structure-activity study to evaluate the effect of the trans,trans-germacranolide<br />

SL lactones costunolide, parthenolide, and their 1,10-epoxy and 11,13-dihydro<br />

derivatives (in a range of 100-0.001 µM) on the growth and germination of several<br />

mono and dicotyledon target species was carried out by Macias et al. (1999). These<br />

compounds appear to have more selective effects on the radicle growth of<br />

monocotyledons. Certain factors such as the presence of nucleophile-acceptor groups<br />

and their accessibility enhance the inhibitory activity. The levels of radicle inhibition<br />

obtained with some compounds on wheat are totally comparable to those of commercial<br />

herbicide Logran and allow proposing them as lead compounds. In addition, a structureactivity<br />

study to evaluate the effect of 17 guaianolide SL (in a range of 100-0.001<br />

µM) on the growth and germination of several mono- and dicotyledon target species<br />

was also performed by Macias et al. (2002a). These compounds appear to have deeper<br />

effects on the growth of either monocots or dicots than the previously tested<br />

germacranolides. Otherwise, the lactone group seems to be necessary for the activity,<br />

though it does not necessarily need to be unsaturated. However, the presence of a<br />

second and easily accessible unsaturated carbonyl system greatly enhances the<br />

inhibitory activity. Lipophilicity and the stereochemistry of the possible anchoring<br />

sites are also crucial factors for the activity.<br />

The dichloromethane extract of dried leaves of Helianthus annuus has yielded,<br />

in addition to the known SL annuolide E and leptocarpin, and the sesquiterpenes<br />

heliannuols A,C,D,F,G,H,1, the new bisnorsesquiterpene, annuionone E, and the new<br />

sesquiterpenes heliannuol L, helibisabonol A and helibisabonol B. Structural<br />

elucidation was based on extensive spectral (one and two-dimensional NMR<br />

experiments) and theoretical studies. The sesquiterpenes heliannuol A and<br />

helibisabonol A and the SL leptocarpin inhibited the growth of etiolated wheat<br />

coleoptiles (Macias et al., 2002b). In addition to (+)-, (-)- and (±)-heliannuol E, growthinhibitory<br />

activities of five synthetic chromanes and four tetrahydrobenzo[b]oxepins<br />

were examined against oat and cress. All heliannuol E isomers exhibited similar<br />

biological activities against cress, whereas when tested against oat roots, the unnatural<br />

optical isomer (+) showed no inhibitory activity. Four brominated chromans and two<br />

39


40<br />

ANA LUISA ANAYA<br />

tetrahydrobenzo[b]oxepins derivatives also showed apparent inhibition against both<br />

cress and oat (Doi et al., 2004).<br />

The tremendous impact of parasitic plants on world agriculture has prompted<br />

much research aimed at preventing infestation. Orobanche and Striga spp. are two<br />

examples of parasitic weeds that represent a serious threat to agriculture in large<br />

parts of the world. The life cycle of these parasitic weeds is closely regulated by the<br />

presence of their hosts, and secondary metabolites that are produced by host plants<br />

play an important role in this interaction. A special interest has been arising on those<br />

host-produced stimulants that induce the germination of parasite seeds. Three classes<br />

of compounds have been described that have germination-stimulating activity:<br />

dihydrosorgoleone, the strigolactones and SL. Keyes et al. (2001) suggest that<br />

dihydrosorgoleone is the active stimulant in the root exudates of sorghum and other<br />

monocotyledonous hosts. However, Butler et al. (1995) and Wigchert et al. (1999)<br />

suggest that dihydrosorgoleone is less likely to be the germination stimulant in vivo<br />

because of its low water solubility, and because no correlation between its production<br />

and the germination of Striga has been found. To date, there is no definite proof that<br />

the germination of parasitic weed seeds in the field is induced by one single signal<br />

compound or class of compounds (and indeed such proof will be hard to obtain)<br />

(Bouwmeester et al., 2003). The capacity of SL, which share some structural features<br />

with the strigolactones, to induce the germination of S. asiatica has been reported<br />

(Fischer et al., 1989, 1990). In addition, a decade after the results of Fischer studies,<br />

Francisco Macías and his group (Pérez de Luque et al., 2000; Galindo et al., 2002)<br />

performed some studies of the structure-activity relationship (SAR) directed to evaluate<br />

the effect of several SL as germination stimulants of three Orobanche spp. (O. cumana,<br />

O. crenata, and O. ramosa). Results are compared with those obtained in the same<br />

bioassay with an internal standard, the synthetic analogue of strigol GR-24. A high<br />

specificity in the germination activity of SL on the sunflower parasite O. cumana has<br />

been observed, and a relationship between such activity and the high sunflower SL<br />

content is postulated. Molecular properties of the natural and synthetic germination<br />

stimulants (GR-24, GR-7, and Nijmegen-1) and SL have been studied using MMX<br />

and PM3 calculations. Consequently, comparative studies among all of them and<br />

their activities have been made. SL tested present similarities in molecular properties<br />

such as the volume of the molecule and the spatial disposition of the carbon backbone<br />

to the natural germination stimulant orobanchol. These properties could be related to<br />

their biological activity. Considering that the sun-flower–O. cumana interaction is<br />

highly specific and that sunflower contains many SL, it is tempting to speculate that<br />

O. cumana has evolved to respond to sesquiterpene lactones (and not or less to<br />

strigolactones) (Bouwmeester et al., 2003).<br />

3.2. Phenolics<br />

In relation with phenolics, Inderjit et al. (1997) conducted a study to understand the<br />

effects of certain phenolics, terpenoides, and their equimolar mixture through agar<br />

gel and soil growth bioassays and their recovery from soils. The eight compounds


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

selected for this study were p-hydroxybenzoic acid, ferulic acid, umbelliferone, catechin,<br />

emodin, 1,8-cineole, carvone, and betulin. Lettuce (Lactuca sativa L.) was used as<br />

test species for agar gel and soil growth bioassays. Root and shoot growth of lettuce<br />

was inhibited for all the above except emodin and catechin. However, in soils treated<br />

with different phenolics and terpenoids, only root growth of lettuce was inhibited,<br />

whereas shoot growth was promoted. Recovery of p-hydroxybenzoic acid and<br />

umbelliferone was higher in unautoclaved soils, while that of catechin was lower.<br />

Nava-Rodriguez et al. (in press) observed the in vitro effects of aqueous leachates<br />

from fresh and dry, flowering and vegetative stage of Phaseolus species, faba bean,<br />

alfalfa, vetch, maize, and squash, and weed species on the root growth of selected<br />

crop and weeds, as well as on two strains of Rhizobium leguminosarum biovar phaseoli<br />

(CPMex1 and Tlaxcala). Most of the specimens were collected in a traditional<br />

agricultural drained field (“Camellon”) in Tlaxcala, Mexico where maize, beans,<br />

squash, alfalfa, faba-beans, and vetch are cultivated in mixed or rotation crops.<br />

Significant effects of leachates from fresh vegetative and flowering cultivated plants<br />

and weeds were predominantly stimulatory on the growth of tested crops, being the<br />

leachates from fresh aerial parts of alfalfa and pinto bean the most stimulatory.<br />

Nevertheless, aqueous leachates from fresh and dry cultivated legumes (vegetative<br />

and flowering) inhibited the growth of weeds. In contrast, the aqueous leachates from<br />

the dry aerial part of almost all plants resulted inhibitory on the root growth of the test<br />

crops, except maize. Aqueous leachates were also evaluated on the growth of two<br />

strains of Rhizobium leguminosarum biovar phaseoli. Leachates from some of the<br />

tested crops significantly stimulated the growth of both Rhizobium strains. The aqueous<br />

leachates from fresh aerial parts of the weeds Simsia amplexicaulis and Tradescantia<br />

crassifolia significantly inhibited the growth of CPMex1 Rhizobium strain. On the<br />

other hand, the aqueous leachates from fresh roots of these same weeds inhibited the<br />

growth of the Tlaxcala strain. In preliminary chemical tests using thin layer<br />

chromatography (TLC), phenolics were detected in dry aerial parts of vegetative alfalfa,<br />

pinto bean, and vetch, and dry aerial part of flowering faba bean suggesting the role<br />

of these compounds in the allelopathic effects of these legumes.<br />

Nilsson et al. (1998) reported on the temporal variation of phenolics and a<br />

dihydrostilbene, batatasin III, in Empetrum hermaphroditum leaves. These authors<br />

reported that first year shoots produced higher levels of phenolics than older tissues.<br />

High phenolic concentration was maintained through the second year, but it declined<br />

afterwards. However, the phytotoxicity of E. hermaphroditum extracts was related<br />

more to batatasin III than phenolics.<br />

Hyder et al. (2002) performed a study focused on the presence and distribution of<br />

secondary phenolic compounds found within creosotebush (Larrea tridentata). Total<br />

phenolics, condensed tannins and nordihydroguaiaretic acid (NDGA) were measured<br />

in nine categories of tissue within creosotebush. Total phenolic and condensed tannin<br />

concentrations were determined using colorimetric methods while NDGA content<br />

was determined with high performance liquid chromatography (HPLC). Phenolics<br />

were present throughout the plant with the highest concentrations in green stems<br />

(40.8 mg/g), leaves (36.2 mg/g), and roots (mean for all root categories=28.6 mg/g).<br />

41


42<br />

ANA LUISA ANAYA<br />

Condensed tannins were found in all tissues with highest concentrations in flowers<br />

(1.7 mg/g), seeds (1.1 mg/g), and roots less than 5 mm in diameter (1.1 mg/g).<br />

Flowers, leaves, green stems and small woody stems (


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

effects on germination and seedling growth of six weeds. Ferulic acid, gallic acid, pcoumaric<br />

acid, p-hydroxybenzoic acid, vanillic acid, and p-vanillin were bioassayed<br />

in concentrations of 10, 1, 0.1, and 0.01 mM. Equimolar mixtures containing all<br />

these phenolics were prepared at the final total concentration of 10, 1, 0.1, and 0.01<br />

mM to test for possible interactive effects. Chenopodium album, Plantago lanceolata,<br />

Amaranthus retroflexus, Solanum nigrum, Cirsium sp. and Rumex crispus were the<br />

selected target weeds. The highest concentration of the compounds inhibited the<br />

germination of all these weeds, but lower concentrations had no effect or were<br />

stimulatory. However, effects varied with the weed species, the concentration of the<br />

compound tested and the compound itself. In assays with the mixture of phenolics<br />

some additive effects were found (Reigosa et al., 1999).<br />

Reversible sorption of phenolic acids by soils may provide some protection to<br />

phenolic acids from microbial degradation. In the absence of microbes, reversible<br />

sorption 35 days after addition of 0.5-3 mu mol/g of ferulic acid or p-coumaric acid<br />

was 8-14% in Cecil A(p) horizon and 31-38% in Cecil B-t horizon soil materials.<br />

The reversibly sorbed/solution ratios (r/s) for ferulic acid or p-coumaric acid ranged<br />

from 0.12 to 0.25 in A(p) and 0.65 to 0.85 in B-t horizon soil materials. When microbes<br />

were introduced, the r/s ratio for both the A(p) and B-t horizon soil materials increased<br />

over time up to 5 and 2, respectively, thereby indicating a more rapid utilization of<br />

solution phenolic acids over reversibly sorbed phenolic acids. The increase in r/s ratio<br />

and the overall microbial utilization of ferulic acid and/or p-coumaric acid were much<br />

more rapid in A(p) than in B-t horizon soil materials. Reversible sorption, however,<br />

provided protection of phenolic acids from microbial utilization for only very short<br />

periods of time. Differential soil fixation, microbial production of benzoic acids (e.g.,<br />

vanillic acid and p-hydroxybenzoic acid) from cinnamic acids (e.g., ferulic acid and<br />

p-coumaric acid, respectively), and the subsequent differential utilization of cinnamic<br />

and benzoic acids by soil microbes indicated that these processes can substantially<br />

influence the magnitude and duration of the phytoxicity of individual phenolic acids<br />

(Blum, 1998).<br />

Soil solution concentrations of allelopathic agents (e.g., phenolic acids) estimated<br />

by soil extractions differ with extraction procedure and the activities of the various<br />

soil sinks (e.g., microbes, clays, organic matter). This led to the hypothesis that root<br />

uptake of phenolic acids is a better estimator of dose than soil solution concentrations<br />

based on soil extracts. This hypothesis was tested by determining the inhibition of net<br />

phosphorus uptake of cucumber seedlings treated for 5 hr with ferulic acid in wholeroot<br />

and split-root nutrient culture systems. Experiments were conducted with II ferulic<br />

acid concentrations ranging from 0 to 1 mM, phosphorus concentrations of 0.25, 0.5,<br />

or 1 mM, and solution pH values of 4.5, 5.5, or 6.5 applied when cucumber seedlings<br />

were 9, 12, or 15 days old. The uptake or initial solution concentration of ferulic acid<br />

was regressed on ferulic acid inhibition of net phosphorus uptake. Attempts were<br />

made to design experiments that would break the collinearity between ferulic acid<br />

uptake and phosphorus uptake. The original hypothesis was rejected because the initial<br />

ferulic acid solution concentrations surrounding seedling roots were more frequently<br />

and consistently related to the inhibition of net phosphorus uptake than to ferulic acid<br />

43


44<br />

ANA LUISA ANAYA<br />

uptake by these roots. The data suggest that root contact, not uptake, is responsible<br />

for the inhibitory activity of phenolic acids (Lehman and Blum, 1999).<br />

Bulk-soil and rhizosphere bacteria are thought to exert considerable influence<br />

over the types and concentrations of phytotoxins, including phenolic acids that reach<br />

a root surface. Induction and/or selection of phenolic acid-utilizing (PAU) bacteria<br />

within the bulk-soil and rhizosphere have been observed when soils are enriched with<br />

individual phenolic acids at concentrations greater than or equal to 0.25 µmol/g soil.<br />

However, since field soils frequently contain individual phenolic acids at concentrations<br />

well below 0.1 µmol/g soil, the actual importance of such induction and/or selection<br />

remains uncertain. Common bacteriological techniques (e.g., isolation on selective<br />

media, and plate dilution frequency technique) were used to demonstrate in Cecil Ap<br />

soil systems: (i) that PAU bacterial communities in the bulk soil and the rhizosphere<br />

of cucumber seedlings were induced and/or selected by mixtures composed of individual<br />

phenolic acids at concentrations well below 0.25 µmol/g soil; (ii) that readily available<br />

carbon sources other than phenolic acids, such as glucose, did not modify induction<br />

and/or selection of PAU bacteria; (iii) that the resulting bacterial communities readily<br />

utilize mixtures of phenolic acids as a carbon source; and (iv) that depending on<br />

conditions (e.g., initial PAU bacterial populations, and phenolic acid concentration)<br />

there were significant inverse relationships between PAU bacteria in the rhizosphere<br />

of cucumber seedlings and absolute rates of leaf expansion and/or shoot biomass. The<br />

decline in seedling growth could not be attributed to resource competition (e.g.,<br />

nitrogen) between the seedlings and the PAU bacteria in these studies. The induced<br />

and/or selected rhizosphere PAU bacteria, however, reduced the magnitude of growth<br />

inhibition by phenolic acid mixtures. For a 0.6 µmol/g soil equimolar phenolic acid<br />

mixture composed of ρ-coumaric acid, ferulic acid, ρ-hydroxybenzoic acid, and vanillic<br />

acid, modeling indicated that an increase of 500% in rhizosphere PAU bacteria would<br />

lead to an approximate 5% decrease (e.g., 20-25%) in inhibition of absolute rates of<br />

leaf expansion (Blum et al., 2000).<br />

Allelopathy due to humus phenolics is a cause of natural regeneration deficiency<br />

in subalpine Norway spruce (Picea abies) forests. If inhibition of spruce germination<br />

and seedling growth due to allelochemicals is generally accepted, in contrast there is<br />

a lack of knowledge about phenolic effects on mycorrhizal fungi. Thus, Souto et al.<br />

(2000) tested the effects of a humic solution and its naturally occurring phenolics on<br />

the growth and respiration of two mycorrhizal fungi: Hymenoscyphus ericae (symbiont<br />

of Vaccinium myrtillus, the main allelochemical-producing plant) and Hebeloma<br />

crustuliniforme (symbiont of P. abies, the target plant). Growth and respiration of H.<br />

crustuliniforme were inhibited by growth medium with the original humic solution (-<br />

6% and -30%), respectively, whereas the same humic solution did not affect growth<br />

but decreased respiration of H. ericae (-55%). When naturally occurring phenolics<br />

(same chemicals and concentrations in the original humic solution) were added to the<br />

growth medium, growth of H. crustuliniforme was not affected, whereas that of H.<br />

ericae significantly increased (+10%). These authors concluded that H. ericae is<br />

better adapted to the allelopathic constraints of this forest soil than H. crustuliniforme


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

and that the dominance of V. myrtillus among understory species could be explained<br />

in this way.<br />

Inderjit and Duke (2003) mentioned that the best evidence for allelopathy should<br />

include some understanding of natural concentrations and rates of allelochemicals.<br />

For example, (±)-catechin has been isolated from Centaurea maculosa (Bais et al.,<br />

2002), an invasive species in North America for which other lines of evidence suggest<br />

root allelopathy (Ridenour and Callaway, 2001). The more common enantiomer, (+)catechin,<br />

has anti-bacterial functions, whereas (–)-catechin has strong allelopathic<br />

effects on other plants. (±)-catechin is harmless to C. maculosa, but has negative<br />

effects on other species at concentrations of ˜100 mg L -1 . (±)-catechin is exuded from<br />

C. maculosa roots creating concentrations from 83.2 to 185 mg L -1 in aqueous solutions.<br />

Importantly, Bais et al. (2002) found (±)-catechin in extracts from natural soils in<br />

fields containing C. maculosa in concentrations as far higher than the minimum<br />

required dose, ranging from 291.6 to 389.8 µg cm -3 .<br />

In allelopathy studies a central goal is to isolate, identify, and characterize<br />

allelochemicals from the soil. However, since it is essentially impossible to simulate<br />

exact field conditions, experiments must be designed with conditions resembling those<br />

found in natural systems. Inderjit (1996) argued that allelopathic potential of phenolics<br />

can be appreciated only when we have a good understanding of i) species responses to<br />

phenolic allelochemicals, ii) methods for extraction and isolation of active phenolic<br />

allelochemicals, and iii) how abiotic and biotic factors affect phenolic toxicity.<br />

Duke et al. (2003) summarized the recent research of the Agricultural Research<br />

Service of United States Department of Agriculture on the use of natural products to<br />

manage pests. They discussed some studies on the use of both phytochemicals and<br />

diatomaceous earth to manage insect pests. Chemically characterized compounds,<br />

such as a saponin from pepper (Capsicum frutescens L), benzaldehyde, chitosan and<br />

2-deixy-D-glucose are being studied as natural fungicides. Resin glycosides for<br />

pathogen resistance in sweet potato and residues of semitropical leguminous plants<br />

for nematode control are also under investigation. Bioassay-guided isolation of<br />

compounds with potential use as herbicides or herbicide leads is underway at several<br />

locations. New natural phytotoxin molecular target sites (asparagine synthetase and<br />

fructose-1,6-bisphosphate aldolase) have been discovered. Weed control in sweet potato<br />

and rice by allelopathy is under investigation. Molecular approaches to enhance<br />

allelopathy in sorghum are also being undertaken. The genes for polyketide synthases<br />

involved in production of pesticidal polyketide compounds in fungi are found to provide<br />

clues for pesticide discovery. Gene expression profiles in response to fungicides and<br />

herbicides are being generated as tools to understand more fully the mode of action<br />

and to rapidly determine the molecular target site of new, natural fungicides and<br />

herbicides.<br />

Research on the chemical basis for allelopathy has often been hindered by the<br />

complexity of plant and soil matrices, making it difficult to track active compounds.<br />

Recent improvements in the cost and capabilities of bench-top chromatography-mass<br />

spectrometry instruments make these tools more powerful and more widely available<br />

45


46<br />

ANA LUISA ANAYA<br />

to assist with molecular studies conducted in today’s expanding field. Such instrumental<br />

techniques are herein recommended as economically efficient means of advancing<br />

the rigor of allelopathy research and assisting the development of a better understanding<br />

of the chemical basis for the allelopathy phenomenon (Haig, 2001).<br />

4. THE MODE OF ACTION OF ALLELOCHEMICALS<br />

Allelopathic chemicals alter plant growth and development by a multiplicity of actions<br />

on physiological processes because there are hundreds of different structures and many<br />

of the compounds have several phytotoxic effects. Whole plants bioassays and<br />

physiological tests designed to use a small quantity of compound are keys to strategies<br />

for elucidating mechanisms of action. Insight into the action of the responsible<br />

chemicals is critical to a more complete explanation of allelopathy and to its application<br />

for improving crop production. Unfortunately, we still find that linkages between<br />

inhibition of growth and the corresponding physiological mechanism are elusive. A<br />

major part of the problem is the array and diversity of allelochemicals. Several hundred<br />

different compounds have been identified, and we speculate that many others will be<br />

eventually being discovered. Most instances of allelopathic inhibition are the result of<br />

the simultaneous action of several compounds, and often these include compounds<br />

whose chemistry is divergent (Einhellig, 2002). The visible evidence in bioassays is<br />

that many phenolics, quinones, sesquiterpene lactones, alkaloids, and others alter<br />

root morphology. Although the cell membrane is an early interface with<br />

allelochemicals, relatively little attention has been given to membrane-related effects<br />

and their molecular targets.<br />

Effects of a single allelochemical or their mixtures on physiological processes<br />

include disruption of membrane permeability (Galindo, et al., 1999), ion uptake (Yu<br />

and Matsui, 1997), inhibition of electron transport in both the photosynthesis and<br />

respiratory chains (Calera et al., 1995a; Abrahim, et al., 2000), alteration of enzymatic<br />

activity (Politycka, 1999, Romagni, et al., 2000), and inhibition of cell division (Cruz-<br />

Ortega et al., 1988; Anaya and Pelayo-Benavides, 1997). In other examples, bioactivitydirected<br />

fractionation of the methanol extract of the roots of Ratibida mexicana resulted<br />

in the isolation of two bioactive sesquiterpene lactones, isoalloalantolactone and elema-<br />

1,3,11-trien-8,12-olide. Both compounds caused a significant inhibition of the radicle<br />

growth of Amaranthus hypochondriacus and Echinochloa crus-galli, exerted moderate<br />

cytotoxicity activity against three different solid tumour cell lines and inhibited the<br />

radial growth of three phytopathogenic fungi. Isoalloalantolactone also caused the<br />

inhibition of ATP synthesis, proton uptake, and electron transport (basal,<br />

phosphorylating and uncoupled) from water to methylviologen therefore acting as a<br />

Hill’s reaction inhibitor.The lactone inhibited only photosystem II (Calera et al., 1995b).<br />

Other compounds studied by Calera et al., (1995c) were the resin glycoside mixture<br />

from Ipomoea tricolor. They tested its effect on seedling growth and plasma membrane<br />

H + -ATPase activity in Echinochloa crus-galli. The resin glycoside mixture as well as<br />

Tricolorin A, the main compound in the mixture, inhibited the activity of the plasma<br />

membrane ATPase. In other studies, Cruz-Ortega et al. (1998) observed plasma


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

membrane disruption in root tips and plasmolized cells in the peripheral zone of<br />

beans and bottle gourd roots treated with the aqueous leachate of S. deppei suggesting<br />

that the allelopathics of this plant alter some membrane processes.<br />

4-phenyl coumarins isolated from Exostema caribaeum and Hintonia latiflora<br />

(Rubiaceae) and some semisynthetic derivatives acted as uncouplers in spinach<br />

chloroplasts. The glycoside 5-Ο-β-D-glucopyranosyl-7-methoxy-3’,4’-dihidroxy-4phenylcoumarin,<br />

5,7,3’,4’-tetrahydroxy-4-phenyl-coumarin, and 7-methoxy-5,3’,4’trihydroxy-4-phenylcoumarin<br />

inhibited ATP synthesis and proton uptake. On the other<br />

hand, basal and phosphorylating electron transport were enhanced by these compounds.<br />

The light-activated Mg 2+ -ATPase was slighted stimulated by the last two coumarins.<br />

In addition, at alkaline pH compound 5,7,3’,4’-tetrahydroxy-4-phenyl-coumarin<br />

stimulated the basal electron flow from water to methylviologen, but at the pH range<br />

from 6 to 7.5 the coumarin did not have any enhancing effect. This last compound,<br />

which possesses four free phenolic hydroxyl groups, was the most active uncoupler<br />

agent. Probably, the phenolate anions may be the active form responsible for the<br />

uncoupling behavior of 4-phenylcoumarins (Calera et al., 1996).<br />

Low molecular weight phenolic compounds were identified in two soils with<br />

different vegetative cover, Fagus sylvatica and Pinus laricio, and were tested at different<br />

concentrations on seed germination of Pinus laricio, and on respiratory and oxidative<br />

pentose phosphate pathway enzymes involved in the first steps of seed germination.<br />

There are marked differences in the phenolic acid composition of the two investigated<br />

soils. All the phenolic compounds bioassayed inhibited seed germination and those<br />

extracted from Pinus laricio soil were particularly inhibitory. Inhibition of germination<br />

of seeds is strongly correlated to the inhibition of the activities of enzymes of glycolysis<br />

and the oxidative pentose phosphate pathway (Muscolo et al., 2001).<br />

Seven-day-old seedlings of cucumber (Cucumis sativus cv. Wisconsin) were treated<br />

with 0.1 mM solutions of cinnamic acid (ferulic and p-coumaric acids) and benzoic<br />

acid (hydroxybenzoic and vanillic acids) derivatives as stressors. The content of free<br />

and glucosylated soluble phenols and the activity of phenylalanine ammonia-lyase<br />

(E.C.4.3.1.5), phenol-beta-glucosyltransferase (E.C.2.4.1.35.), and beta-glucosidase<br />

(E.C.3.2.1.21.) in seedling roots as well as their length and fresh weight were examined.<br />

Changes in glucosylated phenolic content and phenol-beta-glucosyltranspherase<br />

activity were observed under the influence of all phenolics applied. Treatment with<br />

ferulic and p-coumaric acids stimulated the increase of phenylalanine ammonia-lyase<br />

and beta-glucosidase activity and slightly inhibited cucumber root growth (Politycka,<br />

1998).<br />

Environmental stresses (biotic and abiotic), including allelochemicals have been<br />

shown to induce the synthesis of new proteins in plants. These proteins might have<br />

evolutionary value for survival under adverse environmental situations. Romero-<br />

Romero et al. (2002) tested the effect of the mixture of toxic allelochemicals from the<br />

aqueous leachates from Sicyos deppei, Acacia sedillense, Sebastiania adenophora,<br />

and Lantana camara on the radicle growth and cytoplasmic protein synthesis patterns<br />

of Zea mays (maize), Phaseolus vulgaris (bean), Cucurbita pepo (squash), and<br />

Lycopersicon esculentum (tomato). In general, high, medium and low molecular weight<br />

47


48<br />

ANA LUISA ANAYA<br />

cytoplasmic proteins were affected by the different aqueous leachates. Crop plant<br />

responses were diverse, but in general, an increase in protein synthesis was observed<br />

in the treated-roots. Maize was the least affected, but both the radicle growth and also<br />

the protein pattern of tomato were severely inhibited by all allelopathic plants. The<br />

changes observed on protein expression may indicate a biochemical alteration at the<br />

cellular level of the tested crop plants.<br />

Roshchina (2001) discussed the molecular-cellular basis of pollen allelopathy,<br />

related to possible chemosensory mechanisms. The phenomenon consists of a series<br />

of events, viz., a) excretion of signalling and regulatory substances from donor cell<br />

(pollens, pistil stigma); b) recognition of specific signal-stimulus from plant excretions<br />

by acceptor cell (pollen or pistil stigma); c) transmission of chemical information<br />

within the acceptor cell (pollen); and d) development of characteristic response in<br />

acceptor cell. The processes occur in growth, development and normal fertilization.<br />

In the first stage of interactions, allelochemicals are excreted, which act as chemical<br />

signals, growth regulators and modulators of cellular metabolism, etc. The<br />

allelochemicals, acting on fertilization may be, nitrogen-containing substances<br />

(acetylcholine, histamine, serotonin, dopamine, noradrenaline), phenols [(flavonoids:<br />

quercetin, kaempferol, rutin), aromatic acids (benzoic, gallic, vanillic)], terpenoids<br />

(monoterpenes: citral, linalool, cymol), sesquiterpene lactones: azulene and proazulenes<br />

(desacetylinulicine, inulicine, ledol, artemisinine, grosshemine, gaillardine and<br />

austricine), and polyacetylenes (capilline) found in flower excretions. These compounds<br />

were tested in vitro and in vivo on pollen germination of Hippeastrum hybridum.<br />

Nitrogenous compounds stimulate the growth of pollen tube, whereas, their antagonists<br />

blocked normal fertilization and thus fruits or seeds did not form. Terpenoids act on<br />

pollen germination and their stimulatory and inhibitory effects (block fruit formation)<br />

depend on their concentration. These effects of terpenoids on pollen germination are<br />

through chemosignalling and possible steps are: a) spreading of information in pollen<br />

secretions e.g. in olfactory slime; b) binding with special sensors or receptors in<br />

plasmalemma; and c) transfer of stimulus within the pollen cell to nucleus, where<br />

spermia appear and a pollen tube starts to grow. Moving from donor cell,<br />

allelochemicals penetrate the wall of acceptor cell either a) directly (without any<br />

changes in protoplasmic membrane); or b) after conversions [interaction with foreign<br />

substance of low or high-molecular weight (enzymes and protectory proteins) secreted<br />

from donor cells. or compounds of acceptor cell]. Often the second case includes free<br />

radical processes. The transmission of information within cell is third stage which<br />

includes participation of secondary messengers (cyclic AMP and GMP, inositol<br />

triphosphate, Ca ions) and some related enzymatic systems. The final transmission<br />

occurs in membranes of cellular organelles, which respond to information received<br />

through changes in enzymatic activity and metabolism. At cellular level, in pollen<br />

and pistil it may be active excretion, changes in the autofluorescence and membrane<br />

permeability, regulation of alternative pathways in respiration and photosynthesis<br />

and switching on free radical processes.<br />

The phenyl propanoid pathway (PPP) can be stimulated as demonstrated by<br />

Randhir et al. (2004) in mung bean sprouts through the pentose phosphate and


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

shikimate pathways, by natural elicitors such as fish protein hydrolysates (FPH),<br />

lactoferrin (LF) and oregano extract (OE). Elicitation significantly improved the<br />

phenolic, antioxidant and antimicrobial properties of mung bean sprouts. The optimal<br />

elicitor concentrations were 1 ml/l FPH, 250 ppm LF and 1 ml/l OE for the highest<br />

phenolic content that was approximately 20, 35 and 18% higher than control,<br />

respectively, on day 1 of dark germination. The antioxidant activity estimated by Pcarotene<br />

assay in mung bean sprouts was highest on day 1 of germination for all<br />

treatments and control. In general, higher antioxidant activity was observed in the<br />

elicited sprouts compared with control. In the case of 1,1-diphenyl-2-picrythydrazyl<br />

(DPPH) assay the antioxidant activity for all treatments and control was highest on<br />

day 2. Among the different elicitor treatments, OE elicited mung bean sprouts showed<br />

the highest antioxidant activity of 49% DPPH inhibition on day 2. This increased<br />

activity correlates with high guaiacol peroxidase (GPX) activity indicating that the<br />

polymerizing phenolics required during lignification with growth have antioxidant<br />

function. For all elicitor treatments a higher glucose-6-phosphate dehydrogenase<br />

(G6PDH) activity was observed during early germination following the high phenolic<br />

content. This is due to the general mobilization of carbohydrates to the growing<br />

sprouts in response to elicitation. In general the GPX activity steadily increased with<br />

germination for treatments and control. The higher phenolics produced on day 1<br />

was utilized for GPX-mediated polymerization to form polymeric phenolics and lignin<br />

required during germination. The late stage polymerization linked to GPX activity<br />

preceded stimulation of G6PDH. This indicated that as phenolics were polymerized<br />

by GPX in late stages, G6PDH linked precursors such as NADPH (2) and sugar<br />

phosphates were being made available. Antimicrobial activity against Helicobacter<br />

pylori was observed in the mung bean sprout extract from control, LF and OE<br />

treatments from the day 1 stage. Both the LF and OE elicited extracts showed high<br />

antimicrobial activity, which correlated to high antioxidant activity on day 1. The<br />

higher antimicrobial activity was also observed with the higher stimulation of G6PDH<br />

and GPX activity during early stages of germination. This leads to the hypothesis<br />

that enhanced mobilization of carbohydrates (as indicated by G6PDH activity on<br />

days 2) and 4), enhanced polymerization of simple phenols (as indicated by GPX<br />

activity on day 3) contributed to high antioxidant activity producing intermediary<br />

metabolites (day 2).<br />

5. WEED MANAGEMENT<br />

Bhowmik and Inderjit (2003) examined some considerable efforts in designing<br />

alternative weed management strategies due to increase in the number of herbicideresistant<br />

weeds and environmental concerns in the use of synthetic herbicides. The<br />

conventional synthetic herbicides are becoming less and less effective against the<br />

resistant weed biotypes. These authors discussed the role of allelopathic cover crops/<br />

crop residues, natural compounds, and allelopathic crop cultivars in natural weed<br />

management, and gave numerous examples of employing crop residues, cover crops,<br />

and allelopathic crop cultivars in weed management. They concluded that although<br />

49


50<br />

ANA LUISA ANAYA<br />

we cannot eliminate the use of herbicides, their use can be reduced by exploiting<br />

allelopathy as an alternate weed management tool for crop production against weeds<br />

and other pests.<br />

The use of allelopathy for controlling weeds could be either through directly<br />

utilizing natural allelopathic interactions, particularly of crop plants, or by using<br />

allelochemicals as natural herbicides. In the former case, a number of crop plants<br />

with allelopathic potential can be used as cover, smother, and green manure crops for<br />

managing weeds by making desired manipulations in the cultural practices and<br />

cropping patterns. These can be suitably rotated or intercropped with main crops to<br />

manage the target weeds (including parasitic ones) selectively. Even the crop mulch/<br />

residues can also give desirable benefits. The allelochemicals present in the higher<br />

plants as well as in the microbes can be directly used for weed management along<br />

with the management of some herbicides (Singh et al., 2003b).<br />

Singh et al. (2003b) also mentioned that the bioefficacy of allelochemicals can be<br />

enhanced by structural changes or the synthesis of chemical analogues based on them.<br />

Further, in order to enhance the potential of allelopathic crops, several improvements<br />

can be made with the use of biotechnology or genomics and proteomics. In this context<br />

either the production of allelochemicals can be enhanced or the transgenics with<br />

foreign genes encoding for a particular weed-suppressing allelochemical could be<br />

produced. These authors comment that in the former, both conventional breeding and<br />

molecular genetical techniques are useful. However, with conventional breeding being<br />

slow and difficult, more emphasis is laid on the use of modern techniques such as<br />

molecular markers and the selection aided by them. Although the progress in this<br />

regard is slow, nevertheless some promising results are coming and more are expected<br />

in future. In this sense, is important to point out that the potential use of transgenic<br />

plants and other genetically modified organisms (GMO’s) with such or other proposal,<br />

cause a strong controversial with the principles of organic agriculture defined and<br />

established by the International Federation of Organic Agriculture Movements<br />

(IFOAM) founded with the aim to promote an agriculture that is ecologically,<br />

economically, and socially sustainable. IFOAM is opposed to genetic engineering in<br />

agriculture, in view of the unprecedented danger it represents for the entire biosphere<br />

and the particular economic and environmental risks it poses for organic producers<br />

(IFOAM, 2002) * .<br />

Using a soil bioassay technique, Conkling et al. (2002) assessed seedling growth<br />

and incidence of disease of wild mustard (Brassica kaber) and sweet corn (Zea mays)<br />

in soil from field plots that received either of two treatments: incorporated red clover<br />

(Trifolium pratense) residue plus application of compost (‘amended soil’), or<br />

application of ammonium nitrate fertilizer (‘unamended soil’). Soils were analyzed<br />

for percent moisture, dissolved organic carbon, conductivity, phenolics, and nutrient<br />

content. A trend toward greater incidence of Pythium spp. infection of wild mustard<br />

seedlings grown in amended soil was observed during the first 40 days after<br />

* IFOAM, 2002: International Federation of Organic Agriculture Movements (IFOAM). Position on Genetic<br />

Engineering and Genetically Modified Organisms. http://www.ifoam.org/pospap/ge_position_0205.html 2002.


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

incorporation (DAI) of red clover and compost, with significant differences (α= 0.05)<br />

at two out of four sampling dates in 1997, and four out of four sampling dates in<br />

1998. Incidence of Pythium infection was 10-70% greater in the amended soil treatment<br />

during that period. Asymptomatic wild mustard seedlings grown in amended soil<br />

were also on average 2.5 cm shorter (α= 0.05) at 5 DAI than those grown in unamended<br />

soil in one year out of two. Concentration of phenolic compounds in soil solution was<br />

correlated with decreased shoot and root growth (r = 0.50, 0.28, respectively) and<br />

increased incidence of disease (r = 0.48) in wild mustard seedlings in one year out of<br />

two. Dissolved organic carbon concentration was correlated with increased disease in<br />

wild mustard seedlings in both years (r = 0.51, 0.33, respectively). Growth of corn<br />

seedlings did not differ between the two soil treatments, suggesting that red clover<br />

green manure and compost may selectively reduce density and competitive ability of<br />

wild mustard in the field. Bioassay results corresponded well with emergence and<br />

shoot weight results from a related field study, indicating that this technique may be<br />

useful for screening potential soil treatments prior to field studies.<br />

Anaya et al. (1987) applied leaves of Alnus firmifolia, Berula erecta and Juncus<br />

sp., as green manures in corn fields with corn, bean, and squash grown using traditional<br />

techniques. The growth of weeds during the crop period was decreased by the presence<br />

of the green manures. At the same time, stimulation of bean root nodulation by<br />

Rhizobium was obtained with these particular green manure species, nodulation was<br />

also increased in plots with abundant weed growth. These results suggest that the<br />

presence of different secondary metabolites liberated by these green manures and by<br />

some living weeds in the field plots increase the ability of Rhizobium sp. to infect<br />

bean roots.<br />

Weed control by rye, crimson clover, subterranean clover, and hairy vetch cover<br />

crops was evaluated in no-tillage corn during 1992 and 1993 at two North Carolina<br />

locations (Yenish, et al., 1996). Weed biomass reduction was similar with rye, crimson<br />

clover, and subterranean clover treatments, ranging between 19 and 95% less biomass<br />

than a conventional tillage treatment without cover. Weed biomass reduction using<br />

hairy vetch or no cover in a notillage system was similar averaging between 0 and<br />

49%, but less than other covers approximately 45 and 90 d after planting. Weed<br />

biomass was eliminated or nearly eliminated in all cover systems with pre- plus postherbicide<br />

treatments. Weed species present varied greatly between years and locations,<br />

but were predominantly common lambsquarters, smooth pigweed, redroot pigweed,<br />

and broadleaf signalgrass. Corn grain yield was greatest using pre-herbicides or preplus<br />

post-herbicides, averaging between 16 to 100% greater than the nontreated control<br />

across all cover treatments depending on the year and location.<br />

Studies were conducted by Burgos and Talbert (1996) at the Main Agricultural<br />

Experiment Station in Fayetteville and the Vegetable Substation in Kibler, Arkansas,<br />

in 1992 and 1993 on the same plots to evaluate weed suppression by winter cover<br />

crops alone or in combination with reduced herbicide rates in no-till sweet corn and<br />

to evaluate cover crop effects on growth and yield of sweet corn. Plots seeded to rye<br />

plus hairy vetch, rye, or wheat had at least 50% fewer early season weeds than hairy<br />

vetch alone or no cover crop. None of the cover crops reduced population of yellow<br />

51


52<br />

ANA LUISA ANAYA<br />

nutsedge. Without herbicides, hairy vetch did not suppress weeds 8 wk after cover<br />

crop desiccation. Half rates of atrazine and metolachlor (1.1 + 1.1 kg al ha (-1))<br />

reduced total weed density more effectively in no cover crop than in hairy vetch. Half<br />

rates of atrazine and metolachlor controlled redroot pigweed, Palmer amaranth, and<br />

goosegrass regardless of cover crop. Full rates of atrazine and metolachlor [2.2 + 2.2<br />

kg al ha (-1)] were needed to control large crabgrass in hairy vetch. Control of yellow<br />

nutsedge in hairy vetch was marginal even with full herbicide rates- Yellow nutsedge<br />

population increased and control with herbicides declined the second year, particularly<br />

with half rates of atrazine and metolachlor. All cover crops except hairy vetch alone<br />

reduced emergence, height, and yield of sweet corn. Sweet corn yields from half rates<br />

of atrazine and metolachlor equalled the full rates regardless of cover crops.<br />

It is presently not known what effect wheat root residues have in regulating<br />

dicotyledonous (dicot) weed emergence in no-till management systems. Past research<br />

has focused almost entirely on the role of shoot residues, while the role of root residues<br />

in weed control has been essentially ignored. A field study was designed by Blum et<br />

al. (2002) to determine the respective effects of wheat shoot and root residues in<br />

regulating the emergence of three dicotyledonous weed species (morning-glory, pigweed<br />

and prickly sida). Glyphosate-desiccated wheat plots and fallow plots were surface<br />

seeded with morning-glory, pigweed and prickly sida during the spring of 1996 and<br />

1997. Weed seedling emergence was determined for two months during each<br />

experimental period in plots with or without wheat shoot and/or root residues. The<br />

resulting data suggested that: a) the closer desiccation of the wheat cover crop occurred<br />

to the initial emergence of pigweed seedlings, the lower the emergence of that weed,<br />

b) the effects of wheat shoot and/or root residues on dicot weed seedling emergence<br />

vary considerably for the different weed species ranging from stimulation to inhibition<br />

and c) the role of root residues appear to be much more important to regulating weed<br />

emergence than that of surface shoot residues, Differences in soil moisture and<br />

temperature associated with the presence or absence of wheat residues could not be<br />

used to explain the observed treatment effects.<br />

The growth of four summer season crops, namely Cyamopsis tetragonoloba,<br />

Sorghum vulgare, Pennisetum americanum and Zea mays, in fields with or without<br />

residues of the preceding sunflower crop was poor. Crop density, weight of seed or<br />

grain and total yield were significantly lower in sunflower fields than in the control<br />

fields (i.e. those without previous sunflower crops). Growth in terms of plant height<br />

and biomass was drastically reduced after 60 days. The effect was more pronounced<br />

in the fields where sunflower residues were allowed to decompose than in those where<br />

residues were completely removed. The soil collected from sunflower fields (both<br />

with and without residues) was found to be rich in phenolics, which in a laboratory<br />

bioassay were found to be phytotoxic. The reduced growth and yield of crops can be<br />

attributed to the release of phytotoxic phenolics from decomposing sunflower residues<br />

(Batish, et al., 2002).<br />

John and Narwal (2003) assessed that Leucaena leucocephala is the most<br />

productive and versatile multipurpose legume tree in tropical agriculture and has<br />

several uses, thus called ‘miracle tree’. It is a popular choice for intercropping with


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

annuals in hedgerow or alley cropping systems. Its allelopathic effects on oil cereals,<br />

pulses (peas and beans), oilseeds, vegetables, fodder crops, weeds, trees etc. are reviewed<br />

in this paper. The foliage and pods of Leucaena contain the toxic amino acid mimosine<br />

[beta-N-(3-hydroxy-4-pyridone)-alpha-aminopropionic acid] and many other<br />

phytotoxic compounds. The toxic effects of mimosine oil plants and physiology of its<br />

action also are discussed. The future areas identified for research in Leucaena are: (a)<br />

studies on its allelopathic compatibility with different crops to identify sustainable<br />

agroforestry systems (b) investigations to overcome its adverse allelopathic effects<br />

and mimosine toxicity and (c) possibility of using the allelopathic compounds in<br />

Leucaena as natural herbicides.<br />

Marigold (Tagetes erecta) is another multipurpose crop with ceremonial,<br />

ornamental, medical and pharmaceutical uses, and reported antimicrobial properties.<br />

Gómez-Rodríguez et al. (2003) evaluated the effect of marigold intercropped with<br />

tomato (Lycopersicon esculentum) on Alternaria solani conidia germination in vitro,<br />

on conidial density and tomato leaf damage in vivo, as well as microclimatic changes,<br />

compared to tomato intercropped with pigweed (Amaranthus hypochondriacus) and<br />

monocropped tomato. They found that intercropping with marigold induced a<br />

significant (P < 0:05) reduction in tomato early blight caused by A. solani, by means<br />

of three different mechanisms. One was the allelopathic effect of marigold on A.<br />

solani conidia germination, as it was shown in vitro conditions; while pigweed did<br />

not have any of this inhibitory effect in conidia germination. The second way was by<br />

altering the microclimatic conditions around the canopy, particularly by reducing the<br />

number of hours per day with relative humidity 92%, thus diminishing conidial<br />

development. The third mechanism was to provide a physical barrier against conidia<br />

spreading. When intercroppped with tomato, pigweed plants worked also as a physical<br />

barrier and promoted reductions in the maximum relative humidity surrounding the<br />

canopy, but to a lesser extent than marigold.<br />

The allelopathic properties of unburnt (UR) and burnt (BR) residues of Parthenium<br />

hysterophorus towards the growth of two winter crops-radish and chickpea were<br />

investigated (Singh et al., 2003c). The extracts prepared from both UR and BR were<br />

toxic to the seedling length and dry weight of the test crops, those from BR in particular.<br />

The difference was attributed to the highly alkaline nature of the extracts prepared<br />

from BR. Growth studies conducted in soil amended with UR and BR extracts and<br />

residues also revealed phytotoxic effects towards test crops, UR being more active<br />

than BR unlike crude extracts. These effects were attributed to the presence of phenolics<br />

rather than to any significant change in pH or conductivity.<br />

Weston and Duke (2003) focused a review on a variety of weed and crop species<br />

that establish some form of potent allelopathic interference, either with other crops or<br />

weeds, in agricultural settings, in the managed landscape, or in naturalized settings.<br />

They remarked that recent research suggests that allelopathic properties can render<br />

one species more invasive to native species and thus potentially detrimental to both<br />

agricultural and naturalized settings. In contrast, allelopathic crops offer strong<br />

potential for the development of cultivars that are more highly weed suppressive in<br />

managed settings. Both environmental and genotypic effects impact allelochemical<br />

53


54<br />

ANA LUISA ANAYA<br />

production and release over time. A new challenge that exists for future plant scientists<br />

is to generate additional information on allelochemical mechanisms of release,<br />

selectivity and persistence, mode of action, and genetic regulation. In this manner, it<br />

is possible to further protect plant biodiversity and enhance weed management<br />

strategies in a variety of ecosystems.<br />

Ohno et al. (2000) based on previous studies that suggested phenolics from legume<br />

green manures may contribute to weed control through allelopathy, investigated if<br />

red clover (Trifolium pratense) residue amended field soils expressed phytotoxicity to<br />

a weed species, wild mustard (Sinapis arvensis). Field plots involving incorporation<br />

treatments of wheat (Triticum aestivum) stubble or wheat stubble plus 2530 kg ha(-1)<br />

red clover residue, were sampled at -12, 8, 21, 30, 41, 63, and 100 days after residue<br />

incorporation (DAI). Soil-water extracts (1:1, m:v) were analyzed for plant nutrients<br />

and phenolic content. Phytotoxicity of the extracts was measured using a laboratory<br />

wild mustard bioassay. There was a 20% reduction of radicle growth in the green<br />

manure treatment in comparison with the wheat stubble treatment, but only at the<br />

first sample date after residue incorporation (8 DAI). The radicle growth reduction<br />

had the highest correlation with the concentration of soluble phenolics in the soil,<br />

water extracts. Bioassays using aqueous extracts of the clover shoots and roots alone<br />

predicted a radicle growth reduction of 18% for the quantity of clover amendment<br />

rate used in the field plots. The close agreement of the predicted and observed root<br />

growth reduction at 8 DAI further supports clover residue as the source of the<br />

phytotoxicity.<br />

The allelopathic influence of sweet potato cultivar ‘Regal’ on purple nutsedge<br />

was compared to the influence on yellow nutsedge under controlled conditions. Purple<br />

nutsedge shoot dry weight, total shoot length and tuber numbers were significantly<br />

lower than the controls (47, 36, and 19% inhibition, respectively). The influence on<br />

the same parameters for yellow nutsedge (35, 21, and 43% inhibition, respectively)<br />

was not significantly different from purple nutsedge. Sweet potato shoot dry weight<br />

was inhibited by purple and yellow nutsedge by 42% and 45%, respectively. The<br />

major allelopathic substance from ‘Regal’ root periderm tissue was isolated and tested<br />

in vitro on the two sedges. The I 50 ’s for shoot growth, root number, and root length<br />

were 118, 62, and 44 µg/ml, respectively, for yellow nutsedge. The I 50 ’s for root number<br />

and root length were 91 and 85 µg/ml, respectively, for purple nutsedge and the I 50 for<br />

shoot growth could not be calculated (Peterson and Harrison, 1995). These allelopathic<br />

substances, the resin glycosides mixture extracted from the periderm tissue of storage<br />

roots from sweet potato, Ipomoea batatas, was bioassayed for effects on survival,<br />

development, and fecundity of the diamondback moth, Plutella xylostella. The resin<br />

glycoside was incorporated into an artificial diet and fed to P. xylostella larvae. First<br />

instars were placed individually into snap-top centrifuge vials containing artificial<br />

diet with one of six concentrations of resin glycoside material (0.00. 0.25, 0.50, 1.00,<br />

1.50, and 2.00 µg/ml). Each replication consisted of 10 individuals per concentration,<br />

and the experiment was repeated 13 times. Vials were incubated at 25 o C and a<br />

photoperiod of 14:10 (L:D) h in a growth chamber. After 6 d, surviving larvae were<br />

weighed and their sex determined, then returned to their vials. Later, surviving pupae


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

were weighed and incubated at 25 o C until moths emerged. Females were fed, mated<br />

with males from the laboratory colony, and allowed to lay eggs on aluminum foil<br />

strips. Lifetime fecundity (eggs/female) was measured. There were highly significant<br />

negative correlations between resin glycoside levels and survival and between glycoside<br />

levels and larval weight after 0 d of feeding. For larvae that lived at least 6 d, there<br />

was no additional mortality that could be attributed to the resin glycoside material.<br />

However, there was a significant positive correlation between glycoside dosages and<br />

developmental time of larvae (measured as days until pupation). Lifetime fecundity<br />

also was negatively affected at sublethal doses. Resin glycosides may contribute to the<br />

resistance in sweet potato breeding lines to soil insect pests (Jackson and Peterson,<br />

2000). It is important to consider that the use of allelopathic crops or plant residues in<br />

agricultural management will inevitably affect other crop pests, for example insect<br />

populations.<br />

The total resin glycoside content in the periderm of 37 sweetpotato cultivars and<br />

breeding clones was measured by HPLC and varied greatly among the clones, the<br />

highest content was 10.02 % of the periderm dry weight and the lowest was 0.05 %.<br />

Insect damage ratings of the clones and their periderm resin glycoside content were<br />

negatively correlated and all clones with high resin glycoside content exhibited<br />

moderate or low injury from insects. Resin glycosides extracted from ‘Regal’ periderm<br />

and incorporated into potato dextrose agar medium were inhibitory to the growth of<br />

four fungal species of sweetpotato roots; however, these fungi exhibited variable<br />

response. These observations provide evidence that sweetpotato resin glycosides<br />

contribute to the insect and disease resistance in the roots of some sweetpotato-clones<br />

(Harrison et al., 2003).<br />

Barazani and Friedman (2001) discussed the impact of allelopathic, nonpathogenic<br />

bacteria on plant growth in natural and agricultural ecosystems. In some natural<br />

ecosystems, evidence supports the view that in the vicinity of some allelopathically<br />

active perennials (e.g., Adenostoma fasciculatum, California), in addition to<br />

allelochemicals leached from the shrub’s canopy, accumulation of phytotoxic bacteria<br />

or other allelopathic microorganisms amplify retardation of annuals. In agricultural<br />

ecosystems allelopathic bacteria may evolve in areas where a single crop is grown<br />

successively, and the resulting yield decline cannot be restored by application of<br />

minerals. Transfer of soils from areas where crop suppression had been recorded into<br />

an unaffected area induced crop retardation without readily apparent symptoms of<br />

plant disease. Susceptibility of higher plants: to deleterious rhizobacteria is often<br />

manifested in sandy or so-called skeletal soils. The allelopathic effect may occur directly<br />

through the release of allelochemicals by a bacterium that affects susceptible plant(s)<br />

or indirectly through the suppression of an essential symbiont. The process is affected<br />

by nutritional and other environmental conditions; some may control bacterial density<br />

and the rate of production of allelochemicals. Allelopathic nonpathogenic bacteria<br />

include a wide range of genera and secrete a diverse group of plant growth-mediating<br />

allelochemicals. Although a limited number of plant growth-promoting bacterial<br />

allelochemicals have been identified, a considerable number of highly diversified<br />

growth-inhibiting allelochemicals have been isolated and characterized. Some species<br />

55


56<br />

ANA LUISA ANAYA<br />

may produce more than one allelochemical; for example, three different phyotoxins,<br />

geldanamycin, nigericin, and hydanthocidin, were isolated from Streptomyces<br />

hygroscopicus. Efforts to introduce naturally produced allelochemicals as plant growthregulating<br />

agents in agriculture have yielded two commercial herbicides,<br />

phosphinothricin, a product of Streptomyces viridochromogenes, and bialaphos from<br />

S. hygroscopicus. Both herbicides have the same mechanism of action. Many species<br />

of allelopathic bacteria that affect growth of higher plants are not plant specific, but<br />

some do exhibit specificity; for example, dicotyledonous plants were more susceptible<br />

to Pseudomonas putida than were monocotyledons. Differential susceptibility of higher<br />

plants to allelopathic bacteria was noted also in much lower taxonomical categories,<br />

at the subspecies level, in different cultivars of wheat, or of lettuce. Therefore, when<br />

test plants are employed to evaluate bacterial allelopathy, final evaluation must include<br />

those species that are assumed to be suppressed in nature. The release of allelochemicals<br />

from plant residues in plots of ‘continuous crop cultivation’ or from allelopathic living<br />

plants may induce the development of specific allelopathic bacteria.<br />

Striga hermonthica is an obligate root-parasitic flowering plant that severely<br />

threatens cereal production in sub-Saharan Africa. A potential biological control option<br />

for reduction of crop yield-loss within the season of application is the use of soilborne<br />

antagonists of S. hermonthica seed. A study was made (Ahonsi et al., 2002)<br />

with the aim to select soil-borne fluorescent pseudomonad strains capable of<br />

suppressing germination of S. hermonthica seeds and consequently reducing parasitism<br />

and damage to maize. An in vitro screening procedure was developed and was used to<br />

evaluate 460 fluorescent pseudomonad isolates from naturally suppressive soils. This<br />

resulted in the identification of 15 Pseudomonas fluorescens/P. putida isolates that<br />

significantly inhibited germination of S. hermonthica seeds. In a pot experiment using<br />

steam-sterilized soil, there was a significant reduction in the number of S. hermonthica<br />

plants on maize grown from seeds that were inoculated with any of the 15 bacterial<br />

isolates. Inoculation of maize seed with six of these isolates resulted not only in a<br />

reduced number of S. hermonthica plants, but also in an increased maize shoot biomass<br />

compared with the check. When soils inoculated with these bacterial isolates were<br />

left dried for 5 weeks after maize harvest and then planted with a second maize crop,<br />

no reduction in S. hermonthica parasitism was observed. This suggested that the<br />

bacteria did not persist in the soil after the first crop of maize. These results suggest<br />

that saprophytic fluorescent pseudomonads have potential for biological control of S.<br />

hermonthica in maize and that periodic application of bacteria, perhaps through seed<br />

treatment, may be necessary for sustained control.<br />

Chittapur et al. (2001) asserted that integrated weed management systems<br />

involving catch and trap crops are needed to reduce herbicide use in agriculture and<br />

to help to control parasitic weed growth. The effective catch crops viz., fodder millet<br />

(Panicum miliaceum), sorghum (Sorghum bicolor), corn (Zea mays), sudangrass<br />

(Sorghum sudanense) have been identified for the management of Striga asiatica,<br />

and the cowpea (Vigna catjang) for S. gesnerioides. Cotton (Gossypium spp.), soybean<br />

(Glycine max) and peanut (Arachis hypogaea) are important trap crops. Intercropping<br />

of soybean or peanut with sorghum effectively controls S. hermonthica. Flax


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

(Linum usitatissimum) is a useful trap crop for Orobanche ramosa, O. cernua, O.<br />

crenata, and O. aegyptica. In India, sunnhemp (Crotolaria juncea), blackgram<br />

(Phaseolus mungo), greengram (Phaseolus aureus) and sesame (Sesamum indicum)<br />

have shown good potential for Orobanche control. Rotation of trap crop reduces the<br />

population of Orobanche and 3 to 4 years long rotation of catch/trap crops provides<br />

its effective control. Sorghum/maize/paddy (Oryza sativa)-tobacco (Nicotiana tabacum)<br />

rotation reduces the infestation and weed biomass of Orobanche. Relay cropping of<br />

tobacco in capsicum (Capsicum annuum), onion (Allium cepa) and peanut also reduces<br />

the incidence of Orobanche.<br />

Nagabhushana et al. (2001) remarked that no matter how one may define<br />

sustainable agriculture, use of soil-conserving cropping practices, less synthetic<br />

herbicide inputs and better weed control would be compatible components. Previously,<br />

these components were considered incompatible, since it was widely believed that<br />

soil-conserving practices required increased pesticide use, including herbicides.<br />

However, it has been shown that environmental and ecological differences between<br />

the no-till and conventional tillage can enhance the control of certain weed species in<br />

no-till cropping systems. With proper choice and manipulation of cover crops and<br />

residues, it is often possible to reduce the herbicides use. Thus, in eliminating tillage,<br />

by utilizing the surface mulch and allelochemicals leached from a killed cover crop<br />

and using most effective herbicides when needed, weed management has become<br />

much more effective in no-till. In North Carolina, these authors have grown soybean<br />

(Glycine max), tobacco (Nicotiana tabaccum), corn (Zea mays), sorghum (Sorghum<br />

bicolor) and sunflower (Helianthus annuus) in killed heavy mulches of rye (Secale<br />

cereale) without herbicides, other than a non-selective one to kill the rye. Earlyseason<br />

control of broadleaf weeds such as sicklepod (Cassia obtusifolia), morningglory<br />

spp. (Ipomoea spp.), cocklebur (Xanthium strumarium), prickly sida (Sida spinosa),<br />

common purslane (Portulaca oleracea) and pigweed spp. (Amaranthus spp.) has been<br />

80 to 95%. Rye is the most weed suppressing cover crop among several small grains<br />

and subterranean clover (Trifolium subterraneum) and crimson clover (Trifolium<br />

incarnatum) the most suppressive legumes. This approach will still enhance<br />

agricultural sustainability because: (a) productive top-soil will be conserved, (b)<br />

herbicide use (especially preemergence herbicides) can be reduced and (c) herbicides<br />

for cover crop kill and postemergence selective herbicides, even if used, have little<br />

potential for environmental contamination.<br />

Staman et al. (2001) stated that in order to demonstrate that allelopathic<br />

interactions are occurring, one must, among other things, demonstrate that putative<br />

phytotoxins move from plant residues on or in the soil, the source, through the bulk<br />

soil to the root surface, a sink, by way of the rhizosphere. These authors hypothesized<br />

that the incorporation of phytotoxic plant residues into the soil would result in a<br />

simultaneous inhibition of seedling growth and a stimulation of the rhizosphere<br />

bacterial community that could utilize the putative phytotoxins as a carbon source. If<br />

true and consistently expressed, such a relationship would provide a means of<br />

establishing the transfer of phytotoxins from residue in the soil to the rhizosphere of<br />

a sensitive species under field conditions, presently, direct evidence for such transfer<br />

57


58<br />

ANA LUISA ANAYA<br />

is lacking. To test this hypothesis, cucumber seedlings were grown in soil containing<br />

various concentrations of wheat or sunflower tissue. Both tissue types contain phenolic<br />

acids, which have been implicated as allelopathic phytotoxins. The level of<br />

phytotoxicity of the plant tissues was determined by the inhibition of pigweed seedling<br />

emergence and cucumber seedling leaf area expansion. The stimulation of cucumber<br />

seedling rhizosphere bacterial communities was determined by the plate dilution<br />

frequency technique using a medium containing phenolic acids as the sole carbon<br />

source. When sunflower tissue was incorporated into autoclaved soil (to reduce the<br />

initial microbial populations), a simultaneous inhibition of cucumber seedling growth<br />

and stimulation of the community of phenolic acid utilizing rhizosphere bacteria<br />

occurred. Thus, it was possible to observe simultaneous inhibition of cucumber<br />

seedlings and stimulation of phenolic acid utilizing rhizosphere bacteria, and therefore<br />

provide indirect evidence of phenolic acid transfer from plant residues in the soil to<br />

the root surface. However, the simultaneous responses were not sufficiently consistent<br />

to be used as a field screening tool but were dependent upon the levels of phenolic<br />

acids and the bulk soil and rhizosphere microbial populations present in the soil. It is<br />

possible that this screening procedure may be useful for phytotoxins that are more<br />

unique than phenolic acids. Such an inverse relationship between phytotoxicity and<br />

the response of rhizosphere bacterial populations was also observed by Blum et al.<br />

(2000), and such interactions provide indirect evidence for the transfer of<br />

allelochemicals from the plant root to the rhizosphere.<br />

In relation with resistance of weeds to herbicides, Duke et al. (2000) mentioned<br />

that new mechanisms of action for herbicides are highly desirable to fight evolution<br />

of resistance in weeds, to create or exploit unique market niches, and to cope with<br />

new regulatory legislation. Comparison of the known molecular target sites of synthetic<br />

herbicides and natural phytotoxins reveals that there is little redundancy. Comparatively<br />

little effort has been expended on determination of the sites of action of phytotoxins<br />

from natural sources, suggesting that intensive study of these molecules will reveal<br />

many more novel mechanisms of action. These authors gave some examples of natural<br />

products that inhibit unexploited steps in the amino acid, nucleic acid, and other<br />

biosynthetic pathways: AAL-toxin, hydantocidin, and various plant-derived terpenoids.<br />

Natural products have not been utilized as extensively for weed management as<br />

they have been for insect and plant pathogen management, but there are several notable<br />

successes such as glufosinate and the natural product-derived triketone herbicides.<br />

The molecular target sites of these compounds are often unique. Strategies for the<br />

discovery of these materials and compounds are outlined by Duke et al. (2002a).<br />

Numerous examples of individual phytotoxins and crude preparations with weed<br />

management potential are provided by these authors. They described an example of<br />

research to find a natural product solution of a unique pest management problem<br />

(blue-green algae in aquaculture), and mentioned the two fundamental approaches to<br />

the use of natural products for weed management: i) as a herbicide or a lead for a<br />

synthetic herbicide and ii) use in allelopathic crops or cover crops (Duke et al., 2002b).<br />

As it was mentioned, crops may be genetically engineered for weed management<br />

purposes by making them more resistant to herbicides or by improving their ability to


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

interfere with competing weeds. Transgenes for bromoxynil, glyphosate, and<br />

glufosinate resistance are found in commercially available crops. Other herbicide<br />

resistance genes are in development. Glyphosate-resistant crops have had a profound<br />

effect on weed management practices in North America, reducing the cost of weed<br />

management, while improving flexibility and efficacy. In general, transgenic, herbicideresistant<br />

crops have reduced the environmental impact of weed management because<br />

the herbicides with which they are used are generally more environmentally benign<br />

and have increased the adoption of reduced-tillage agriculture. Crops could be given<br />

an advantage over weeds by making them more competitive or altering their capacity<br />

to produce phytotoxins (allelopathy). Strategies for producing allelopathic crops by<br />

biotechnology are relatively complex and usually involve multiple genes. One can<br />

choose to enhance production of allelochemicals already present in a crop or to impart<br />

the production of new compounds. The first strategy involves identification of the<br />

allelochemical(s), determination of their respective enzymes and the genes that encode<br />

them, and, the use of genetic engineering to enhance production of the compound(s).<br />

The latter strategy would alter existing biochemical pathways by inserting transgenes<br />

to produce new allelochemicals (Duke et al., 2002c).(See controversy between organic<br />

agriculture and biotechnology - Control of Weeds and Management of Agroecosystems,<br />

Pag. 18, first paragraph)<br />

More sophisticated techniques will be used to search for alternative to herbicides<br />

in agroecosystems. The use of winter cover crops is beneficial to agriculture. Stanislaus<br />

and Cheng (2002) tried to design a cover crop that self-destructs in response to an<br />

environmental cue, thereby eliminating the use of herbicides and tillage to remove<br />

the cover crop in late spring. Here, this novel concept is tested in a model system. The<br />

onset of summer brings with it elevated temperatures. Using this as the environmental<br />

cue, a self-destruction cassette was designed and tested in tobacco. A heat-shockresponsive<br />

promoter was used to direct expression of the ribonuclease Barnase. Because<br />

Barnase is extremely toxic to cells, it was necessary to coexpress its inhibitor, Barstar,<br />

whose expression was under the control of the CaMV 35S promoter. The wild-type<br />

and two mutated Barnase genes, one missense and one translation attenuated, were<br />

tested. The results indicated that the translation-attenuated version of the Barnase<br />

gene was most effective in causing heat-shock-regulated plant death. Analysis of the<br />

T-2 progeny of a transgenic plant carrying this Barnase mutant showed that the Barnase<br />

gene expression was sixfold higher in heat-shock-treated plants compared with<br />

untreated plants. This level of Barnase gene expression was sufficient to kill transgenic<br />

plants.<br />

Many advances in disciplines such as chemistry, biochemistry, plant breeding,<br />

genetics, engineering, and others have been applied in a positive manner to improve<br />

knowledge in weed science. The emerging field of genomics is likely to have a similar<br />

positive effect on our understanding of weeds and their management in various plant<br />

agriculture systems. Genomics involves the large-scale use of molecular techniques<br />

for identification and functional analysis of complete or nearly complete genomic<br />

complements of genes. Commercial application of genomics has already occurred for<br />

improvement in certain crop input and output traits, including improved quality<br />

59


60<br />

ANA LUISA ANAYA<br />

characteristics and herbicide and insect resistance. Additional commercial applications<br />

of genomics in weed science will be identification of genes involved in crop ability.<br />

Genes controlling early crop root emergence, rapid early-season leaf and root<br />

development for fast canopy closure, production of allelochemicals for natural weed<br />

control, identification of novel herbicide target sites, resistance mechanisms, and genes<br />

for protecting crops against specific herbicides can and will be identified. Successful<br />

crop improvement in these areas using the tools of genomics will dramatically affect<br />

weed-crop interactions and improve crop yields while reducing weed problems. In<br />

relation to improved basic knowledge of weeds and the resulting ability to improve<br />

our weed management techniques, genomics will offer the weed science community<br />

many new and exciting research opportunities. Scientists will be able to determine<br />

the genetic composition of weed populations and how it changes over time in relation<br />

to agricultural practices, Identification of genes contributing to weediness, perennial<br />

growth habit, herbicide resistance, seed and vegetative structure dormancy, plant<br />

architecture and morphology, plant reproductive characters (outcrossing and<br />

hybridization, introgression), and allelopathy will be identified and utilized with highthroughput<br />

DNA sequencing and other genomics-based technologies. Using genomics<br />

to improve our understanding of weed biology by determining which genes function<br />

to affect the fitness, competitiveness, and adaptation of weeds in agricultural<br />

environments will allow the development of improved management strategies.<br />

Information is provided concerning the current state of molecular research in various<br />

areas of weed science and specific genomic research currently being conducted at<br />

Purdue University using transfer DNA (TDNA) activation tagging to generate large<br />

populations of mutated plants that can be screened for genes of importance to weed<br />

science (Weller et al., 2001).<br />

6. SUPPRESSIVE SOILS<br />

When soils are characterized by a very low level of disease development even though<br />

a virulent pathogen and susceptible host are present, they are known as suppressive<br />

soils. Biotic and abiotic elements of the soil environment contribute to suppressiveness,<br />

however most defined systems have identified biological elements as primary factors<br />

in disease suppression. Many soils possess similarities with regard to microorganisms<br />

involved in disease suppression, while other attributes are unique to specific pathogensuppressive<br />

soil systems. The organisms’ operative in pathogen suppression does so<br />

via diverse mechanisms including competition for nutrients, antibiosis and induction<br />

of host resistance (Mazzola, 2002). Non-pathogenic Fusarium spp. and fluorescent<br />

Pseudomonas spp. play a critical role in naturally occurring soils that are suppressive<br />

to Fusarium wilt. Suppression of take-all of wheat, caused by Gaeumannomyces<br />

graminis var. tritici, is induced in soil after continuous wheat monoculture and is<br />

attributed, in part, to selection of fluorescent pseudomonads with capacity to produce<br />

the antibiotic 2,4-diacetylphloroglucinol. Cultivation of orchard soils with specific<br />

wheat varieties induces suppressiveness to Rhizoctonia root rot of apple caused by<br />

Rhizoctonia solani AG 5. Wheat cultivars that stimulate disease suppression enhance


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

populations of specific fluorescent pseudomonad genotypes with antagonistic activity<br />

toward this pathogen. Methods that transform resident microbial communities in a<br />

manner which induces natural soil suppressiveness have potential as components of<br />

environmentally sustainable systems for management of soilborne plant pathogens<br />

(Mazzola, 2002).<br />

Actually, agricultural soils suppressive to soilborne plant pathogens occur<br />

worldwide, and for several of these soils the biological basis of suppressiveness has<br />

been described. Two classical types of suppressiveness are known. General suppression<br />

owes its activity to the total microbial biomass in soil and is not transferable between<br />

soils. Specific suppression owes its activity to the effects of individual or select groups<br />

of microorganisms and is transferable. The microbial basis of specific suppression to<br />

four diseases, Fusarium wilts, potato scab, apple replant disease, and take-all, is<br />

discussed by Weller et al. (2002). One of the best-described examples occurs in takeall<br />

decline soils. In Washington State, take-all decline results from the buildup of<br />

fluorescent Pseudomonas spp. that produce the antifungal metabolite 2,4-diacetylphloroglucinol.<br />

Producers of this metabolite may have a broader role in diseasesuppressive<br />

soils worldwide. By coupling molecular technologies with traditional<br />

approaches used in plant pathology and microbiology, it is possible to dissect the<br />

microbial composition and complex interactions in suppressive soils.<br />

In three of 12 soils obtained from agricultural fields in California, population<br />

density development of Meloidogyne incognita under susceptible tomato was<br />

significantly suppressed when compared to identical but methyl iodide (MI)-fumigated,<br />

M. incognita re-infested soils. When the 12 soils were infested with second-stage<br />

juveniles (J2) of M. incognita and the juveniles were extracted after 3 days, significantly<br />

fewer J2 were recovered from 9 of the 12 non-treated soils than from the MI-fumigated<br />

equivalents. In one of the 12 soils, infestation 3 weeks before planting resulted in<br />

lower nematode population densities than infestation at planting in both MI-fumigated<br />

and non-treated soil. The combination of infestation 3 weeks before planting with<br />

infestation at planting did not alter the occurrence or degree of root-knot nematode<br />

suppressiveness (Pyrowolakis et al., 2002).<br />

Yin et al. (2004) established that for suppressive soils that have a biological<br />

nature, one of the first steps in understanding them is to identify the organisms<br />

contributing to this phenomenon. They presented a new approach for identifying<br />

microorganisms involved in soil suppressiveness. This strategy identifies<br />

microorganisms that fill a niche similar to that of the pathogen by utilizing substrate<br />

use assays in soil. To demonstrate this approach, they examined an avocado grove<br />

where a Phytophthora cinnamomi epidemic created soils in which the pathogen could<br />

not be detected with baiting techniques, a characteristic common to many soils with<br />

suppressiveness against P. cinnamomi. Substrate utilization assays were used to identify<br />

rRNA genes (rDNA) from bacteria that rapidly grew in response to amino acids known<br />

to attract P. cinnamomi zoospores. Six bacterial rDNA intergenic sequences were<br />

prevalent in the epidemic soils but uncommon in the non-epidemic soils. These<br />

sequences belonged to bacteria related to Bacillus mycoides, Renibacterium<br />

salmoninarum, and Streptococcus pneumoniae. We hypothesize that bacteria such as<br />

61


62<br />

ANA LUISA ANAYA<br />

these, which respond to the same environmental cues that trigger root infection by the<br />

pathogen, will occupy a niche similar to that of the pathogen and contribute to<br />

suppressiveness through mechanisms such as nutrient competition and antibiosis.<br />

A similar experimental approach was developed by Borneman et al. (2004) for<br />

identifying microorganisms involved in specified functions such as pathogen<br />

suppressiveness. In this approach, it was postulated that the microorganisms involved<br />

in pathogen suppressiveness could be discovered by identifying those organisms whose<br />

populations positively correlate with high levels of suppressiveness. The approach<br />

has three phases. The first phase is to identify bacterial and fungal rRNA genes (rDNA)<br />

from soils possessing various levels of suppressiveness. Ribosomal DNA sequences<br />

that are more abundant in the highly suppressive soils than in the less suppressive<br />

soils are considered candidate sequences. A method termed oligonucleotide<br />

fingerprinting of rRNA genes (OFRG) is used to obtain extensive analysis of microbial<br />

community composition. The second phase of this experimental approach is to verify<br />

the results obtained from phase one using quantitative PCR. Here, selective PCR<br />

primers for each of the candidate rDNA sequences are designed. These primers are<br />

then used to determine the relative amounts of the candidate sequences in soils<br />

possessing various levels of suppressiveness produced by several different methods<br />

such as mixing various quantities of suppressive and fumigation-induced nonsuppressive<br />

soil, biocidal treatments and temperature treatments. In phase three, the<br />

organisms that consistently correlate with suppressiveness are isolated and amended<br />

to non-suppressive soils to assess their abilities to produce suppressiveness. The utility<br />

of this experimental approach was demonstrated by using it to identify microorganisms<br />

involved in suppressiveness against the plant-parasitic nematode, Heterodera schachtii.<br />

This general experimental approach should also be useful for the identifying<br />

microorganisms involved in functions other then pathogen suppressiveness.<br />

Kloepper et al. (1999) discussed concepts and examples of how naturally occurring<br />

bacteria (plant-associated bacteria residing in the rhizosphere, phyllosphere, and inside<br />

tissues of healthy plants –endophytic), and introduced bacteria may contribute to<br />

management of soilborne and foliar diseases. Some introduced rhizobacteria have<br />

been found to enhance plant defences, leading to systemic protection against foliar<br />

pathogens upon seed or root-treatments with the rhizobacteria. In these cases,<br />

introduction of the rhizobacteria results in reduced damage to multiple pathogens,<br />

including viruses, fungi and bacteria. An alternative strategy to the introduction of<br />

specific antagonists is the augmentation of existing antagonists in the root environment.<br />

This augmentation may result from the use of specific organic, amendments, such as<br />

chitin, which stimulate populations of antagonists, thereby inducing suppressiveness.<br />

Intercropping or crop rotation with some tropical legumes, including velvetbean<br />

(Mucuna deeringiana), lead to management of phytoparasitic nematodes, partly<br />

through stimulation of antagonistic microorganisms. Some biorational nematicides,<br />

such as specific botanical aromatic compounds, also appear to induce suppressiveness<br />

through alterations in the soil microbial community.<br />

Single isolates of bacterial endophytes, obtained from the nematode antagonistic<br />

plant species African (Tagetes erecta) and French (T. patula) marigold, were introduced


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

into potatoes (Solanum tuberosum). Several bacterial species possessed activity against<br />

root-lesion nematodes (Pratylenchus penetrans) in soils around the root zone of<br />

potatoes, namely: Microbacterium esteraromaticum, Tsukamurella paurometabolum,<br />

isolate TP6, Pseudomonas chlororaphis, Kocuria varians and K. kristinae. Of these,<br />

M. esteraromaticum and K. varians depressed the population densities of root-lesion<br />

nematodes without incurring any yield penalty (tuber wet weight). No significant<br />

differences were found in the total numbers of P. penetrans nematodes, rhabditid<br />

nematodes or ‘other’ parasitic nematode species within the root tissues of bacterized<br />

potato plants compared to the unbacterized check. Overall, tuber fresh weights and<br />

tuber number were equal to or significantly lower (P < 0.05) in bacterized plants than<br />

their unbacterized counterpart (Sturz and Kimpinski, 2004). The authors of this study<br />

conclude that endoroot bacteria from Tagetes spp. can play a role in nematode<br />

suppression through the attenuation of nematode proliferation, and proposed that<br />

these nematode control properties are capable of transfer to other crops in a rotation<br />

as a beneficial ‘residual’ microflora – a form of beneficial microbial allelopathy.<br />

In relation with this same type of study, Hallman et al. (1998) performed a<br />

greenhouse experiments with cotton and cucumber to determine the effects of<br />

inoculation of the parasitic nematode Meloidogyne incognita on population dynamics<br />

of indigenous bacterial endophytes and introduced endophytic bacterial strains JM22<br />

(Enterobacter asburiae) and 89B-61 (Pseuedomonas fluorescens) applied as seed<br />

treatments. Internal communities of endophytic bacteria in roots were generally largest<br />

in the presence of M. incognita. Recovery of JM22 from cucumber roots was positively,<br />

but not significantly, associated with soilborne nematode inoculum size, except at 2<br />

weeks after inoculation. The internal populations of 89B-61 applied to seed also<br />

increased with nematode applications. The diversity of indigenous bacterial endophytes<br />

changed within 7 d after M. incognita inoculation. Species richness and diversity of<br />

endophytic bacteria were slightly, but not significantly, greater for nematode-infested<br />

plants than for non-infested plants. Alcaligenes piechaudii and Burkholderia pickettii<br />

occurred only in nematode-infested plants, whereas Bievundimonas vesicularis was<br />

mainly isolated from nematode-free plants. Agrobacterium radiobacter and<br />

Pseudomonas spp. were the most common taxa found in both treatments, accounting<br />

for a total of 41% and 37% of the community for non-inoculated and inoculated<br />

plants, respectively. JM22 colonized cotton roots internally and was also found in<br />

high numbers on the root surface around nematode penetration sites and on root galls<br />

where the root tissue had been disruptured due to gall enlargement. Single cells of<br />

JM22 were attached to the cuticle of M. incognita juveniles. Sturz et al. (2000) assesses<br />

that endophytic bacteria and M. incognita form complex associations and an<br />

understanding of these associations will aid efforts to develop and manage microbial<br />

communities of endophytic bacteria for practical use as biocontrol agents against<br />

plant-parasitic nematodes and soil-borne pests and pathogens.<br />

In addition, Postma et al. (2003) found that compost amended soil has also been<br />

found to be suppressive against plant diseases in various cropping systems. The level<br />

and reproducibility of disease suppressive properties of compost might be increased<br />

by the addition of antagonists. In this study, the establishment and suppressive activity<br />

63


64<br />

ANA LUISA ANAYA<br />

of two fungal antagonists of soil-borne diseases was evaluated after their inoculation<br />

in potting soil and in compost produced from different types of organic waste and at<br />

different maturation stages. The fungal antagonists Verticillium biguttatum, a<br />

mycoparasite of Rhizoctonia solani, and a non-pathogenic isolate of Fusarium<br />

oxysporum antagonistic to Fusarium wilt, survived at high levels (10 3 -10 5 CFU g -1 )<br />

after 3 months incubation at room temperature in green waste compost and in potting<br />

soil. Their populations faded-out in the organic household waste compost, especially<br />

in the matured product. In bioassays with R. solani on sugar beet and potato, the<br />

disease suppressiveness of compost increased or was similar after enrichment with V.<br />

biguttatum. The largest effects, however, were present in potting soil, which was very<br />

conducive for the disease as well as the antagonist. Similar results were found in the<br />

bioassay with F. oxysporum in carnation where enrichment with the antagonistic F.<br />

oxysporum had a positive or neutral effect. Postma et al. (2003) foresee great potential<br />

for the application of antagonists in agriculture and horticulture through enrichment<br />

of compost or potting soil with antagonists or other beneficial micro-organisms.<br />

All soils are suppressive to phytonematodes to some degree. The degree of<br />

suppressiveness to them or other soilborne pathogens in a soil can be enhanced not<br />

only by infesting soil with selected microorganisms, but by the use of appropriate<br />

cropping systems and the application to soil of specific organic amendments or chemical<br />

compounds. Conducive cropping systems such as monoculture can reduce soil<br />

suppressiveness to the point where the soil is not resistant to plant parasitic nematodes<br />

(Wang et al., 2002).<br />

Inorganic fertilizers containing ammoniacal nitrogen or formulations releasing<br />

this form of N in the soil are most effective for suppressing nematode populations.<br />

Anhydrous ammonia has been shown to reduce soil populations of Tylenchorhynchus<br />

claytoni, Helicotylenchus dihystera, and Heterodera glycines. The rates required to<br />

obtain significant suppression of nematode populations are generally in excess of 150<br />

kg N/ha. Urea also suppresses several nematode species, including Meloidogyne spp.,<br />

when applied at rates above 300 kg N/ha. Additional available carbon must be provided<br />

with urea to permit soil microorganisms to metabolize excess N and avoid phytotoxic<br />

effects. There is a direct relation between the amount of “protein” N in organic<br />

amendments and their effectiveness as nematode population suppressants. Most<br />

nematicidal amendments are oil cakes, or animal excrements containing 2-7% (w/w)<br />

N; these materials are effective at rates of 4-10 t/ha. Organic soil amendments<br />

containing mucopolysaccharides (e.g., mycelial wastes, chitinous matter) are also<br />

effective nematode suppressants (Rodriguez-Kábana, 1986).<br />

Vargas-Ayala and Rodriguez-Kábana (2001) established a field microplot trial to<br />

evaluate nematode population dynamics in a rotation program utilizing nematodesuppressive<br />

and non-suppressive legumes, and nematode-host and nonhost grass<br />

species. The rotation treatments consisted of velvetbean (Mucuna deeringiana) or<br />

cowpea (Vigna unguiculata) during the first year, followed in winter by oat (Avena<br />

sativa), wheat (Triticum aestivum), rye (Secale cereale), rye grass (Lolium sp.), clover<br />

(Trifolium sp.), hairy vetch (Vicia villosa), lupine (Lupinus sp.) or fallow. Rotation in


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

the second and third year consisted of soybean (Glycine max). Results showed that<br />

velvetbean had a generally suppressive effect on populations of root-knot (Meloidogyne<br />

incognita), cyst (Heterodera glycines), and stunt (Tylenchorhynchus claytoni)<br />

nematodes in soil and roots. It had little effect on populations of Helicotylenchus<br />

dihystera. Velvetbean rotations with winter grass species were also effective in reducing<br />

nematode population densities in soil. Soybean yields were positively correlated with<br />

velvetbean in rotations with winter grass species. High populations of M. incognita<br />

were negatively correlated with soybean yields. The use of velvetbean as a rotation<br />

crop assures reduction of important plant-parasitic nematodes in soil and an<br />

improvement in soybean yield.<br />

Wang et al. (2002) made an extensive review on the use of Crotalaria spp.<br />

(Fabaceae) as a suppressor of agricultural pests, particularly nematodes. These authors<br />

summarized the knowledge of the efficacy of Crotalaria spp. for plant-parasitic<br />

nematode management, described the mechanisms of nematode suppression, and<br />

outline prospects for using this crop effectively. They mentioned that Crotalaria is a<br />

poor host to many plant-parasitic nematodes including Meloidogyne spp.,<br />

Rotylenchulus reniformis, Radopholus similis, Belonolaimus longicaudatus, and<br />

Heterodera glycines. It is also a poor or non-host to a large group of other pests and<br />

pathogens. Besides, Crotalaria is competitive with weeds without becoming a weed,<br />

grows vigorously to provide good ground coverage for soil erosion control, fixes<br />

nitrogen, and is a green manure. However, most Crotalaria species are susceptible to<br />

Pratylenchus spp., Helicotylenchus sp., Scutellonema sp. and Criconemella spp.<br />

Crotalaria species are used as preplant cover crops, intercrops, or soil amendments.<br />

When used as cover crops, Crotalaria spp. reduces plant-parasitic nematode<br />

populations by: i) acting as a nonhost or a poor host, ii) producing allelochemicals<br />

that are toxic or inhibitory, iii) providing a niche for antagonistic flora and fauna, and<br />

iv) trapping the nematode.<br />

A non-host to a nematode species is a plant in which the nematode fails to<br />

reproduce. A plant is considered as resistant to nematodes when these fail to live<br />

inside the host or early dead in the host; they decreased the production of eggs; or<br />

their growth or development are inhibited by the plant (Wang et al., 2002).<br />

Allelopathic effects of the plant against nematodes were described as a mechanism<br />

of suppression of nematodes. Soler-Serratosa et al. (1996) evaluated the nematicidal<br />

activity of thymol, a phenolic monoterpene present in the essential oils of several<br />

plant families. Thymol was added to soil at rates 25-250 ppm. Initial and final<br />

population densities of Meloidogyne arenaria, Heterodera glycines, Paratrichodorus<br />

minor, and Dorylaimoid nematodes, as well as disease incidence, declined sharply<br />

with increased dosages of thymol. Thymol was also applied at 0, 50, 100, and 150<br />

ppm to soil in combination with 0, 50, and 100 ppm benzaldehyde, an aromatic<br />

aldehyde present in nature as a moiety of plant cyanogenic glycosides. Combinations<br />

in which benzaldehyde was applied at 100 ppm showed synergistic effects in<br />

suppressing initial and final soil populations of M. arenaria and H. glycines. Significant<br />

reductions in root galling and cyst formation in soybean were attributable to thymol<br />

at ≥ 50 ppm.<br />

65


66<br />

ANA LUISA ANAYA<br />

Hallmann and Sikora (1996) confirmed that endophytic fungi isolated from the<br />

cortical tissue of surface sterilized tomato roots collected from field plots produced<br />

secondary metabolites in nutrition broth that were highly toxic to Meloidogyne<br />

incognita. Especially strains of Fusarium oxysporum were highly active with 13 of 15<br />

strains producing culture filtrates toxic to nematodes. They investigated also the<br />

mechanism of action of the toxic metabolites produced by the non-pathogenic F.<br />

oxysporum strain 162 with proven biological control of M. incognita in pot experiments.<br />

These metabolites reduced M. incognita mobility within 10 min of exposure. After 60<br />

min, 98% of juveniles were inactivated. Fifty percent of juveniles with exposure of 5<br />

h were dead, and 24 h exposure resulted in 100% mortality. In a bioassay with lettuce<br />

seedlings metabolite concentrations >100 mg/l reduced the number of M. incognita<br />

juveniles on the roots comparing to the water control. The F. oxysporum toxins were<br />

highly effective towards sedentary parasites and less effective towards migratory<br />

endoparasites. Non-parasitic nematodes were not influenced at all. Metabolites of<br />

strain 162 also reduced significantly the growth of Phytophthora cactorum, Pythium<br />

ultimum and Rhizoctonia solani in vitro.<br />

Three out of 15 bacterial strains preselected for antagonistic activity in different<br />

pathosystems showed biocontrol activity towards Meloidogyne incognita on lettuce<br />

and tomato as described by Hoffmann-Hergarten et al. (1998). They found that seed<br />

treatment with the rhizobacteria Pseudomonas sp. W34 or Bacillus cereus S18 resulted<br />

in significant reductions in root galling and enhanced seedling biomass. The yield<br />

response of M. incognita-infested tomato was tested in a long-term pot experiment<br />

using three antagonistic bacteria, i.e., Pseudomonas sp. W34, Bacillus cereus S18<br />

and Bacillus subtilis VM1-32. Significant reduction in M. incognita gall index was<br />

observed within 18 weeks after inoculation with all three bacterial strains. B. cereus<br />

S18 caused a 9 % yield increase when compared with the nematode control and thereby<br />

compensated for the yield loss due to nematode infection. Early maturity of fruits on<br />

M. incognita-infested tomato plants after inoculation of B. cereus S18 was observed<br />

when compared with both the nematode and the untreated control.<br />

Vargas-Ayala et al. (2000) hypothesized that the induction of soil suppressiveness<br />

to plant parasitic nematodes that occurs following planting of velvetbean (Mucuna<br />

deeringiana) is associated with the development of an antagonistic microflora in soils<br />

and rhizospheres. They performed a crop rotation study in microplots, consisting of<br />

three crop cycles. Cycle 1 involved planting of either velvetbean or cowpea (Vigna<br />

unguiculata) in the first spring. Cycle 2 during the next fall and winter was fallow or<br />

cover-cropped with wheat (Triticum aestivum) or crimson clover (Trifolium<br />

incarnatum). Cycle 3 the next spring was soybean (Glycine max). Rhizosphere fungal<br />

populations were significantly smaller on velvetbean than on cowpea at the end of<br />

cycle 1. The use of velvetbean in cycle 1 significantly decreased rhizosphere bacterial<br />

populations on crops in cycle 2, compared to treatments which had cowpea in cycle 1.<br />

Velvetbean also influenced bacterial diversity, generally increasing frequency of bacilli,<br />

Arthrobacter spp. and Burkholderia cepacia, while reducing fluorescent<br />

pseudomonads. Some of these effects persisted through cycle 3. Fungal diversity was<br />

influenced in cycle 1 by velvetbean; however, effects generally did not persist through


cycles 2 and 3. The results indicate that the use of velvetbean in a cropping system<br />

alters the microbial communities of the rhizosphere and soil, and they are consistent<br />

with the hypothesis that the resulting control of nematodes results from induction of<br />

soil suppressiveness.<br />

6.1. Resistence through natural compounds<br />

Sometimes, natural compounds that confer resistance to a plant against nematode<br />

infestation could be of foreign origin (Reitz et al., 2000). Recent studies have shown<br />

that living and heat-killed cells of the rhizobacterium Rhizobium etli strain G12 induce<br />

in potato roots systemic resistance to infection by the potato cyst nematode Globodera<br />

pallida. To better understand the mechanisms of induced resistance, Reitz et al. (2000)<br />

focused on identifying the inducing agent. Since heat-stable bacterial surface<br />

carbohydrates such as exopolysaccharides (EPS) and lipopolysaccharides (LPS) are<br />

essential for recognition in the symbiotic interaction between Rhizobium and legumes,<br />

their role in the R. etli-potato interaction was studied. EPS and LPS were extracted<br />

from bacterial cultures, applied to potato roots, and tested for activity as an inducer of<br />

plant resistance to the plant-parasitic nematode. Whereas EPS did not affect G. pallida<br />

infection, LPS reduced nematode infection significantly in concentrations as low as 1<br />

and 0.1 mg ml(-1). Split-root experiments, guaranteeing a spatial separation of inducing<br />

agent and challenging pathogen, showed that soil treatments of one half of the root<br />

system with LPS resulted in a highly significant (up to 37%) systemic induced reduction<br />

of G. pallida infection of potato roots in the other half. The results clearly showed<br />

that LPS of R. etli G12 act as the inducing agent of systemic resistance in potato<br />

roots.<br />

In relation with Crotalaria spp. secondary metabolites, it is well known that<br />

these plants produce pyrrolizidine alkaloids and monocrotaline which have high<br />

vertebrate toxicity and could potentially be toxic to nematodes; but it is possible also<br />

that the low C/N ratio of Crotalaria may also contribute to its allelopathic effect<br />

against nematodes. Materials with very low C/N or high content of ammonia will<br />

either result in plasmolysis of nematodes, or proliferation of nematophagous fungi<br />

due to the release of NH4+-N (Rich and Rahi, 1995).<br />

6.2. Soil Amendments<br />

ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

A study was conducted to determine the effects of combinations of organic amendments<br />

and benzaldehyde on plant-parasitic and non-parasitic nematode populations, soil<br />

microbial activity, and plant growth (Chavarria-Carvajal et al., 2001). Pine<br />

bark, velvetbean and kudzu were applied to soil at rates of 30 g/kg and paper waste at<br />

40 g/kg alone and in combination with benzaldehyde (300 mul/kg), for control of<br />

plant-parasitic nematodes. Pre-plant and post-harvest soil and soybean root samples<br />

were analyzed, and the number of parasitic and non-parasitic nematodes associated<br />

with soil and roots were determined. Soil samples were taken at 0, 2, and 10 weeks<br />

after treatment to determine population densities of bacteria and fungi. Treatment<br />

67


68<br />

ANA LUISA ANAYA<br />

effects on microbial composition of the soybean rhizosphere were also determined by<br />

identifying microorganisms. Bacteria strains were identified rising fatty acid analysis,<br />

and fungus identification was done rising standard morphological measurements and<br />

appropriate taxonomic keys. Results showed that most amendments alone or in<br />

combination with benzaldehyde reduced damage from plant parasitic nematodes.<br />

Benzaldehyde applied alone or in combination with the amendments exerted a selective<br />

action on the activity and composition of microbial populations in the soybean<br />

rhizosphere. In control soils the bacterial flora was predominantly Gram-negative,<br />

while in soils amended with velvetbean or kudzu alone or with benzaldehyde. Grampositive<br />

bacteria were dominant. Mycoflora promoted by the different amendments or<br />

combinations with benzaldehyde included species of Aspergillus, Myrothecium,<br />

Penicillium, and Trichoderma.<br />

Calvet et al. (2001) evaluated the survival of two species of plant parasitic<br />

nematodes: the root-lesion nematode Pratylenchus brachyurus, and the root-knot<br />

nematode Meloidogyne javanica, in saturated atmospheres of 12 natural chemical<br />

compounds. The infectivity of two isolates of arbuscular mycorrhizal fungi: Glomus<br />

mosseae and Glomus intraradices, under identical experimental conditions, was also<br />

determined. All the compounds tested exerted a highly significant control against M.<br />

javanica and among them, benzaldehyde, salicilaldehyde, borneol, p-anisaldehyde<br />

and cinnamaldehyde caused a mortality rate above 50% over P. brachyurus. The<br />

infectivity of G. intraradices was inhibited by cinnamaldehyde, salicilaldehyde, thymol,<br />

carvacrol, p-anisaldehyde, and benzaldehyde, while only cinnamaldehyde and thymol<br />

significantly inhibited mycorrhizal colonization by G. mosseae.<br />

When soybean plant responses to Meloidogyne incognita infestation were<br />

compared to resistant (Bryan) and susceptible (Brim) cultivars at 0, 1, 3, 10, 20, and<br />

34 days after infestation, Qiu and collaborators (1997) observed that the resistant<br />

cultivar had higher chitinase activity than the susceptible cultivar at every sample<br />

time beginning at the third day. Results from isoelectric focusing gel electrophoresis<br />

analyses indicated that three acidic chitinase isozymes with isoelectric points (pIs) of<br />

4.8, 4.4, and 4.2 accumulated to a greater extent in the resistant compared to the<br />

susceptible cultivar following challenge. SDS-PAGE analysis of root proteins revealed<br />

that two proteins with molecular weights of approximately 31 and 46 kD accumulated<br />

more rapidly and to a higher level in the resistant than in the susceptible cultivar.<br />

Additionally, three major protein bands (33, 22, and 20 kD) with chitinase activity<br />

were detected with a modified SDS-PAGE analysis in which glycolchitin was added<br />

into the gel matrix. These results indicate that higher chitinase activity and early<br />

induction of specific chitinase isozymes may be associated with resistance to rootknot<br />

nematode in soybean.<br />

Antagonists, most likely favored by selected cover crops, include mainly fungal<br />

egg parasites, trapping fungi, endoparasitic fungi, fungal parasites of females,<br />

endomycorrhizal fungi, planthealth promoting rhizobacteria, and obligate bacterial<br />

parasites. There are several hypotheses on how cover crops can enhance nematodeantagonistic<br />

activities. A series of ecological events may be involved. The decomposing


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

organic material is a significant event because the bacteria which proliferate after<br />

organic matter incorporation become a food base for microbiovorous nematodes. In<br />

turn, these nematodes serve as a food source for nematophagous fungi (Wang et al.,<br />

2002). Leguminous crops enhance nematophagous fungi better than other crops.<br />

Rootknot symptoms were reduced more by alfalfa amendments in a 4-year microplot<br />

test than by chemical fertilization of plots (Mankau, 1968). Microplots amended with<br />

alfalfa meal increased nematode-trapping fungal activity of Drechmeria coniospora<br />

(Van Den Boogert et al., 1994). Pea enhanced the densities and species diversity of<br />

nematode-trapping fungi more than white mustard or barley. In addition, formation<br />

of conidial traps of nematode-trapping fungi was more prevalent in the pea rhizosphere<br />

than in root-free soil (Persmark and Nordbring-Hertz, 1997; Persmark and Jansson,<br />

1997).<br />

Being a legume, Crotalaria juncea has characteristics that may make the crop<br />

useful for nematode antagonism. Plant exudates from Crotalaria spp. were selective<br />

for microbial species antagonistic to phytopathogenic fungi and nematodes (Rodriguez-<br />

Kábana and Kloepper, 1998). The changes in soil enzymatic activity was investigated<br />

by Chavarría and Rodriguez-Kábana (1998) when they incorporated four organic<br />

amendments (velvetbean, kudzu, pine bark, and urea-N) to the soil to evaluate their<br />

effects on the root-knot nematode (Meloidogyne incognita). The amendments were<br />

applied to nematode-infested soil at rates of 0 to 5% and placed in pots planted with<br />

‘Davis’ soybean (Glycine max). The number of M. incognita juveniles and nonparasitic<br />

nematodes associated with the soil and root tissues were determined after 8 weeks.<br />

Soil samples were taken at 0, 2, and 10 weeks after amendment application for<br />

determination of soil enzyme activities. Most organic amendments were effective in<br />

reducing root galling and root-knot nematodes and increasing populations of nonparasitic<br />

nematodes. Catalase and esterase were sharply increased by most rates of<br />

velvetbean, kudzu, and pine bark. Application of velvetbean, kudzu, and urea to soil<br />

stimulated urease activity in proportion to the amendment rates. Results suggest that<br />

complex modes of action operating in amended soils are responsible for suppression<br />

of M. incognita.<br />

In relation with nematode-trapping fungi, a major group of nematode antagonists,<br />

they can be enhanced by incorporation of residues of C. juncea. These fungi have<br />

been categorized into two groups: parasitic and saprophytic. The saprophytic group<br />

consists of predators characterized by sticky three-dimensional networks and nonspontaneous<br />

trap formation. These fungi have a saprophytic and a predatory (trap<br />

formation) phase. In the presence of nematodes, or even exudates and homogenates<br />

of nematodes, trap formation is induced. The parasitic group consists of nematodetrapping<br />

fungi that form constricting rings, adhesive knobs, or adhesive branches.<br />

These fungi form traps spontaneously, and thus are more effective trappers (Wang,<br />

2000).<br />

Among these two groups of nematode trapping fungi, the population densities of<br />

parasitic fungi are more likely to be enhanced by organic matter due to the rich microbial<br />

flora and fauna. The nematode trapping by these fungi are not nematode species- or<br />

trophic group specific, therefore the enhancement of nematode-trapping fungi by<br />

69


70<br />

ANA LUISA ANAYA<br />

organic matter incorporation should lead to increased trapping of plant-parasitic<br />

nematodes (Wang, 2000).<br />

Soil amended with C. juncea to give a 1:100 (w:w) concentration, enhanced<br />

parasitic nematode-trapping fungi, nematode egg parasitic fungi, vermiform stage<br />

parasites, and bacterivorous nematode population densities more efficiently than soil<br />

amended with chopped pineapple tissues or non-amended soil. Crotalaria juncea<br />

amendment enhanced the population densities of nematode-trapping fungi and the<br />

percentage of eggs parasit-ized by the fungi. Enhancement of nematode-trapping fungi<br />

was most effective in soils that had not been treated with 1,3-dichloropropene for at<br />

least 5 months. Suppression of R. reniformis by C. juncea amendment was correlated<br />

with parasitic nematode-trapping fungi, fungal egg parasites, and bacterivorous<br />

nematodes. Nematode-trapping fungi population densities were higher in C. juncea<br />

planted plots than weed fallow plots. However, four months after removal of C. juncea,<br />

and replacement with pineapple plants, the population densities of nematode-trapping<br />

fungi greatly decreased (Wang, 2000).<br />

Suppressive cropping systems rely on the use of precisely defined sequences of<br />

crops to increase populations and activities of naturally occurring antagonistic<br />

microorganisms in soil. Some crops such as velvetbean (Mucuna deerengiana) produce<br />

compounds which are directly toxic to nematodes and stimulate microbial antagonism<br />

to plant parasitic nematodes. These ‘active’ crops when included in cropping systems<br />

can increase suppressiveness of the system against nematodes. There are a number of<br />

active crops throughout the world which can be used in a practical manner to enhance<br />

naturally occurring biological control of plant parasitic nematodes (Wang, 2000)<br />

Rich and Rahi (1995) conducted two greenhouse trials to determine the influence<br />

of ground seed of castor (Ricinus communis), crotalaria (Crotalaria spectabilis), hairy<br />

indigo (Indigofera hirsuta), and wheat (Triticum aestivum) on tomato (Lycopersicon<br />

esculentum) growth and egg mass production of Meloidogyne javanica (test 1) or M.<br />

incognita (test 2). Ground seed from each plant species was individually mixed with<br />

an air-dried, fine sandy soil at rates of 0, 0.5, 1.0, and 2.0% (w/w). The mixtures were<br />

placed in one-liter plastic pots, and water was added to bring soil to field capacity.<br />

After ten days, 0 or 10 000 M. javanica or M. incognita eggs and juveniles were<br />

added to each pot. A single ‘Homestead’ tomato seedling was transplanted into each<br />

pot and allowed to grow for 70 days in test 1 and 75 days in test 2. Compared to the<br />

non-amended control, egg mass production was significantly reduced by all treatments<br />

except the 0.5% levels of wheat and castor and the 1.0% castor treatment. The 2.0%<br />

levels of ground seed of Crotalaria and hairy indigo almost completely suppresses<br />

egg mass production of both M. javanica or M incognita. With the exception of the<br />

1% Crotalaria treatment in test 2, total plant weight did not differ between treatments<br />

and the control.<br />

Morris and Walker (2002) mixed dried ground plant tissues from 20 leguminous<br />

species with Meloidogyne incognita-infested soil at 1, 2 or 2.5, and 5% (w/w) and<br />

incubated for 1 week at room temperature (21 to 27 0 C). Tomato (‘Rutgers’) seedlings<br />

were transplanted into infested soil to determine nematode viability. Most tissues


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

reduced gall numbers below the non-amended controls. The tissue amendments that<br />

were most effective include: Canavalia ensiformis, Crotalaria retusa, Indigofera<br />

hirsuta, I. nummularifolia, I. spicata, I. suffruticosa, I. tinctoria, and Tephrosia adunca.<br />

Although certain tissues reduced the tomato dry weights, particularly at the higher<br />

amendment rates (5%), some tissues resulted in greater dry weights. These nontraditional<br />

legumes, known to contain bioactive phytochemicals, may offer considerable<br />

promise as soil amendments for control of plant-parasitic nematodes. Not only do<br />

these legumes reduce root-knot nematodes but some of them also enhance plant height<br />

and dry weight.<br />

Nematode management is rarely successful in the long term with unitactic<br />

approaches. It is important to integrate multiple-tactics into a strategy. Crotalaria<br />

offers the potential to be one of the tactics. Some Crotalaria species are potential<br />

cover crops for managing several important plant-parasitic nematodes including<br />

Meloidogyne spp. and R. reniformis. Unfortunately, the residual effects are short term<br />

(a few months). Crotalaria, a poor host, generally helps reduce nematode population<br />

densities, but the number of nematodes will resurge on subsequent host crops. The<br />

damage threshold level, especially on longer-term crops, will often be reached or<br />

exceeded. This scenario strongly suggests that integrating the Crotalaria rotation<br />

system with other nematode management strategies is necessary. Among the<br />

possibilities for integration are crop resistance, enhanced crop tolerance, selection for<br />

fast growing crop varieties, soil solarization, and biological control. Chemical<br />

nematicides should be avoided in a cropping system if the objective is to enhance<br />

nematode-antagonistic microorganisms in the cropping system. Several studies have<br />

demonstrated the destructive effect of fumigation treatments to nematode antagonistic<br />

microorganisms. Crotalaria juncea amendments failed to enhance nematode-trapping<br />

fungi populations in soils that were recently treated with 1,3-dichloropropane. Wang<br />

et al. (2002) concluded that the major impediment to using Crotalaria is its shortterm<br />

effect in agricultural production systems, and suggested that integrating other<br />

pest management strategies with Crotalaria could offer promising nematode<br />

management approaches.<br />

7. CONCLUSIONS AND FINAL REMARKS<br />

In this chapter the different roles that allelopathy can play as a bioregulator tool in<br />

agriculture are discussed. A wide spectrum of studies are given on allelopathic plants<br />

and other organisms, the chemistry involved in these studies, the mechanisms of<br />

action of some allelochemicals, and the use of allelopathy to control weeds, pests<br />

(nematodes) and diseases.<br />

Many arguments can be given in favor of organic and sustainable agricultural<br />

practices as new forms of resources management such as multiple cropping, cover<br />

crops, organic compost, and biological controls for pests. Allelopathy is an emerging<br />

tool for a more biorational management of natural resources. However, allelopathy is<br />

not a simple panacea for the solution of ecological problems in agroecosystems or in<br />

natural ecosystems. It has not been considered as a universal ecological phenomenon;<br />

71


72<br />

ANA LUISA ANAYA<br />

allelopathy is a challenging and exigent matter of study. At present we have proof<br />

that secondary metabolites are involved with biotic interactions, and that allelopathic<br />

effects may restrict or enhance, alone or in relation with other environmental factors<br />

(light, temperature, humidity and nutrients), the distribution, health and growth of<br />

species in natural, artificial or managed communities. In the search for application of<br />

allelopathy knowledge is crucial to understand other biotic interactions (competition,<br />

defense against herbivory) and also the actual and full significance of a mixture of<br />

secondary metabolites all together acting in the environment (Anaya, 1999).<br />

Allelopathy typically operates through the release, modification, and joint action<br />

of a number of allelochemicals in a particular situation, and transitions through the<br />

soil add to the complications for explaining the phenomenon. The frontiers in research<br />

on allelopathy include isolation of additional compounds that may be involved, and<br />

determining more precisely how allelochemicals production is regulated and how the<br />

compounds function to inhibit growth. Such information may allow modification of<br />

crop plants so they have enhanced capability for weed suppression. Alternatively,<br />

new herbicides, pesticides, and growth regulators may be developed from some of<br />

plant and microorganisms compounds (Einhellig, 1989).<br />

In the study of biological interactions mediated by secondary metabolites it is<br />

very important to perform multidisciplinary investigations in a long term approach in<br />

order to understand these interactions from an holistic point of view and make use of<br />

them for beneficial purposes in the management of natural resources in agroecosystems.<br />

8. REFERENCE<br />

Abrahim, D., Braguini, W.L., Kelmer-Bracht, A.M., Ishii-Iwamoto, E.L. Effects of four monoterpenes on<br />

germination, primary root growth, and mitochondrial respiration of maize. J Chem Ecol 2000; 26:611-<br />

624.<br />

Ahonsi, M.O., Berner, D.K. Emechebe, A.M., Lagoke, S.T. Selection of rhizobacterial strains for suppression of<br />

germination of Striga hermonthica (Del.) Benth. Seeds. Biol Con 2002; 24:143-152.<br />

Anaya, A.L., S. del Amo. Allelopathic potential of Ambrosia cumanensis H.B.K. (Compositae) in a tropical<br />

zone of México. J Chem Ecol 1978; 4:289-304.<br />

Anaya, A.L., Ramos, L., Cruz, R., Hernández, J., Nava, V. Allelopathy in Mexican Traditional Agroecosystems:<br />

A case study in Tlaxcala. J Chem Ecol 1987; 13:2083-2101.<br />

Anaya, A.L., Hernández-Bautista, B.E., Torres-Barragán, A., León-Cantero, J. Jiménez-Estrada, M. Phytotoxicity<br />

of cacalol and some derivatives obtained from the roots of Psacalium decompositum (A. Gray) H. Rob &<br />

Brettell (Asteraceae), “matarique” or “maturín”. J Chem Ecol 1996; 22:393-403.<br />

Anaya, A.L., Pelayo-Benavides, H.R. Allelopathic potential of Mirabilis jalapa L. (Nyctaginaceae): Effects on<br />

germination, growth and cell division of some plants. Allelopathy J 1997; 4:57-68.<br />

Anaya, A.L. Allelopathy as a Tool in the Management of Biotic Resources in Agroecosystems. Crit Rev Plant<br />

Sci 1999; 18: 697-739.<br />

Anaya, A.L., Mata, R., Sims, J., González-Coloma, A., Cruz-Ortega, R., Guadaño, A., Hernández-Bautista<br />

B.E., Ríos, G., Gómez-Pompa, A. Allelochemical potential of Callicarpa acuminata (Verbenaceae). J<br />

Chem Ecol 2003; 29:2725-2740.<br />

Bais, H.P., Walker, T.S., Stermitz, F.R., Hufbauer, R.A., Vivanco, J.M. 2002. Enantiomeric-dependent phytotoxic<br />

and antimicrobial activity of (±)-catechin. A rhizosecreted racemic mixture from spotted knapweed. Plant<br />

Physiol 2003; 128: 1173-1179.<br />

Barazani, O. Friedman J. Allelopathic bacteria and their impact on higher plants. Crit Rev Microbiol 2001;<br />

27:41-55.


Barberi, P. Weed management in organic agriculture: are we addressing the right issues? Weed Res 2002;<br />

42:177-193.<br />

Batish, D.R., Tung, P., Singh, H., Kohli, R.K. Phytotoxicity of sunflower residues against some summer season<br />

crops. J Agron Crop Sc 2002; 188:19-24.<br />

Bhowmik, P.C., Inderjit. Challenges and opportunities in implementing allelopathy for natural weed management.<br />

Crop Prot 2003; 22:661-671.<br />

Blum U., King, L.D., Gerig, T.M., Lehman, M.E., Worsham, A.D. Effects of clover and small grain cover crops<br />

and tillage techniques on seedling emergence of some dicotyledonous weed species. Amer J Altern Agricul<br />

1997; 4:146–161.<br />

Blum, U. Effects of microbial utilization of phenolic acids and their phenolic acid breakdown products on<br />

allelopathic interactions. J Chem Ecol 1998; 24:685-708.<br />

Blum, U., Staman, K.L., Flint, L.J., Shafer, S. Induction and/or selection of phenolic acid-utilizing bulk-soil and<br />

rhizosphere bacteria and their influence on phenolic acid phytotoxicity. J Chem Ecol, 2000; 26:2059-<br />

2078.<br />

Blum, U., King, L.D., Brownie, C. Effects of wheat residues on dicotyledonous weed emergence in a simulated<br />

no-till system. Allelopathy J 2002; 9:159-176.<br />

Borneman, J., Olatinwo, R., Yin, B., Becker J.O. An experimental approach for identifying microorganisms<br />

involved in specified functions: utilisation for understanding a nematode suppressive soil. Australasian<br />

Plant Pathol 2004; 33:151-155.<br />

Burgos, N.R., Talbert, R.E. Weed control and sweet corn (Zea mays var rugosa) response in a no-till system<br />

with cover crops. Weed Sci 1996; 44:355-361.<br />

Butler, L.G. Chemical communication between the parasitic weed Striga and its crop host. A new dimension of<br />

allelochemistry. In: Allelopathy: Organism, Processes and Application. Inderjit, Dakshini K.M.M.,<br />

Einhellig, F.A. eds. ACS Symposium Series Washington, DC 1995; 582:158-168.<br />

Bouwmeester, H.J., Matusova, R., Zhongkui, S., Beale, M.H. Secondary metabolite signalling in host-parasitic<br />

plant interactions.Curr Opin Plant Biol 2003; 6:358-364.<br />

Calera, M.R., Mata, R., Anaya A.L, Lotina, B. 5-O-b-D-galactopyranosyl-7-methoxy-3’, 4’-dihydro-4phenylcoumarin,<br />

an inhibitor of photophosphorylation in spinach chloroplasts. Photosynth Res 1995a;<br />

45:105-110.<br />

Calera, M.R., Soto, F., Sanchez, P., Bye, R., Hernandez-Bautista, B. Anaya, A.L., Lotina-Hennsen, Mata, R.<br />

Biochemically active sesquiterpene lactones from Ratibida mexicana. Phytochemistry 1995b; 40:419-<br />

425.<br />

Calera, M.R, Anaya, A.L., and Gavilanes-Ruiz, M. Effect of a phytotoxic resin glycoside on the activity of the<br />

H +<br />

ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

-ATPase from plasma membrane. J Chem Ecol 1995c; 21:289-297.<br />

Calera, M., Mata, R., Lotina-Hennsen, B., Anaya, A.L. Uncoupling behavior of the 4-phenylcoumarins in spinach<br />

chloroplasts- Structure-activity relationships. J Agr Food Chem 1996; 44:2966-2969.<br />

Calvet C, Pinochet J, Camprubi A, Estaun V, Rodriguez-Kabana R. Evaluation of natural chemical compounds<br />

against root-lesion and root-knot nematodes and side-effects on the infectivity of arbuscular mycorrhizal<br />

fungi. J Plant Pathol 2001; 107:601-605.<br />

Chavarria-Carvajal JA, Rodriguez-Kabana R, Kloepper JW, Morgan-Jones G. Changes in populations of<br />

microorganisms associated with organic amendments and benzaldehyde to control plant-parasitic<br />

nematodes. Nematropica 2001; 31:165-180.<br />

Chavarria-Carvajal, J.A., Rodriguez-Kabana, R. Changes in soil enzymatic activity and control of Meloidogyne<br />

incognita using four organic amendments. Nematropica 1998; 28:7-18.<br />

Chittapur, B.M., Hunshal, C.S., Shenoy H. Allelopathy in parasitic weed management: Role of catch and trap<br />

crops. Allelopathy J 2001; 8:147-159.<br />

Chou, C.H. Allelopathy and sustainable agriculture. ACS Symp. Ser. 1995; 582:211-223.<br />

Conklin, A.E., Erich, M.S., Liebman, M., Lambert, D., Gallandt, E.R., Halteman, W.A. Effects of red clover<br />

(Trifolium pratense) green manure and compost soil amendments on wild mustard (Brassica kaber)<br />

growth and incidence of disease. Plant Soil 2002; 238:245-256.<br />

Cruz-Ortega, R., Anaya, A.L., and Ramos, L. Effects of allelopathic compounds of maize pollen on respiration<br />

and cellular division of watermelon. J Chem Ecol 1988; 14:71-86.<br />

Cruz-Ortega, R., Anaya, A.L., Hernández, B.E., Laguna G. Effects of allelochemical stress produced by Sicyos<br />

deppei on seedling root ultrastructure of Phaseolus vulgaris and Cucurbita peppo. J Chem Ecol 1998;<br />

24: 2039-2057.<br />

73


74<br />

ANA LUISA ANAYA<br />

Del Amo, S., Anaya, A.L. Effect of some sesquiterpenic lactones on the growth of certain secondary tropical<br />

species. J Chem Ecol 1978; 4:305-313.<br />

Doi, F., Ohara, T., Ogamino, T., Sugai, T., Higashinakasu, K., Yamada, K., Shigemori, H., Hasegawa, K.,<br />

Nishiyama, S. Plant-growth inhibitory activity of heliannuol derivatives. Phytochemistry 2004; 65:1405-<br />

1411.<br />

Duke, S.O., Plant terpenoids as pesticides. In Keeler, R.F. and A.T. Tu (Eds.). Handbook of Natural Toxins. Vol.<br />

6. Marcel Dekker, New York, 1991.<br />

Duke, S.O., Romagni, J.G., Dayan, F.E. Natural products as sources for new mechanisms of herbicidal action.<br />

Crop Prot 2000; 19:583-589.<br />

Duke, S.O., Dayan, F.E., Rimando, A.M., Schrader, K.K., Aliotta, G., Oliva, A., Romagni, J.G. Chemicals from<br />

nature for weed management. Weed Sci 2002a; 50:138-151.<br />

Duke, S.O., Rimando, A,M., Baerson, S.R., Scheffler, B.E., Ota, E. Belz, R.G. Strategies for the use of natural<br />

products for weed management. J Pesticide Sci 2002b; 27:298-306.<br />

Duke, S.O., Scheffler, B.E., Dayan, F.E., Dyer, W.E. Genetic engineering crops for improved weed management<br />

traits. Crop Biotechnol. ACS Symp Ser 2002c ; 829: 52-66.<br />

Duke, S.O., Baerson, S.R., Dayan, F.E., Rimando, A.M., Scheffler, B.E., Tellez, M.R., Wedge, D.E., Schrader,<br />

K.K., Akey, D.H., Arthur, F.H., De Lucca, A.J., Gibson, D.M., Harrison, H.F., Peterson, J.K., Gealy,<br />

D.R., Tworkoski, T., Wilson, C.L., Morris, J.B. Research on natural products for pest management. United<br />

States Department of Agriculture, Agricultural Research Service. Pest Manag Sci 2003; 59:708-717.<br />

Einhellig, F.A. Interaction among allelochemicals and other stress factors of the plant environment. In:<br />

<strong>Allelochemicals</strong>, Role in Agriculture and Forestry, GR Waller ed., ACS Symposium Series, Vol. 330.<br />

American Chemical Society, Washington DC, USA 1987; pp. 343–357.<br />

Einhellig, F.A. The Physiology of Allelochemical Action: Clues and Views. In: Allelopathy. From Molecules to<br />

Ecosystems. Science Publishers, Reigosa, M.J. and Pedrol, N. eds. Inc. Enfield, NH, USA, 2002.<br />

Elakovich, S.D. Terpenoids as models for new agrochemicals. ACS 380. American Chemical Society, Washington,<br />

D.C. 1988; 250-261.<br />

Evenari, M. Germination inhibitors. Bot Rev 1949; 15:153-194.<br />

Fischer, N.H. The function of mono and sesquiterpenes as plant germination and growth regulators. In The<br />

Science of Allelopathy. Putnam, A.R. and Ch-Sh. Tang Eds. John Wiley & Sons. 1986; pp. 203-218.<br />

Fischer, N.H. Plant terpenoids as allelopathic agents. In: Ecological Chemistry and Biochemistry of Plant<br />

Terpenoids. Harborne, J.B. and F.A. Tomas-Barberan (Eds.). Proceedings of the Phytochemical Society<br />

of Europe. Clarendon Press-Oxford.: 1991; pp. 377-398.<br />

Fischer, N.H., Weidenhamer, J.D., Bradow, J.M. Dihydroparthenolide and other sesquiterpene lactones stimulate<br />

witchweed germination. Phytochemistry 1989; 28: 2315-2317.<br />

Fischer, N.H., Weidenhamer, J.D., Riopel, J.L., Quijano, L., Menelaou, M.A. Stimulation of witchweed<br />

germination by sesquiterpene lactones: a structure-activity study. Phytochemistry 1990; 29:2479-2483.<br />

Fischer, N.H., Williamson, G.B., Weidenhamer, J.D., Richardson, D.R. In search of allelopathy in the Florida<br />

scrub: The role of terpenoids. J Chem Ecol 1994; 20:1355-1380.<br />

Galindo, J.C.G., Hernández, A, Dayan F.E., Macías, F.A., Duke, S.O. Dehydrozaluzanin C, a natural<br />

sesquiterpenolide, causes rapid plasma membrane leakage. Phytochemistry 1999; 52:805-813.<br />

Galindo, J.C.G., de Luque, A.P., Jorrin, J., Macias, F.A. SAR studies of sesquiterpene lactones as Orobanche<br />

cumana seed germination stimulants. J Agric Food Chem 2002; 50:1911-1917.<br />

Gómez-Rodríguez, O, Zavaleta-Mejía, E., and González-Hernández, V.A. Livera-Muñoz M., and Cárdenas-<br />

Soriano, E. Allelopathy and microclimatic modification of intercropping with marigold on tomato early<br />

blight disease development. Field Crops Res 2003; 83:27-34.<br />

Haig, T. Application of hyphenated chromatography-mass spectrometry techniques to plant allelopathy research.<br />

J Chem Ecol 2001; 27:2363-2396.<br />

Hallmann, J. and Sikora, R.A.. Toxicity of fungal endophyte secondary metabolites to plant parasitic nematodes<br />

and soil-borne plant pathogenic fungi. Eurp J Plant Pathol 1996; 102:155-162.<br />

Hallmann, J., Quadt-Hallmann, A., Rodriguez-Kabana, R., and Kloepper, J.W. Interactions between Meloidogyne<br />

incognita and endophytic bacteria in cotton and cucumber. Soil Biol Biochem 1998; 30:925-937.<br />

Harper, J.L. Population Biology of Plants. Academic Press, New York, 1977.


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

Harrison, H.F., Peterson, J.K., Jackson, D.M., and Snook, M.E. Periderm resin glycoside contents of sweetpotato,<br />

Ipomoea batatas (L.) Lam. clones and their biological activity. Allelopathy J 2003; 12: 53-60.<br />

Hoffmann-Hergarten, S., Gulati, M.K., Sikora R.A.. Yield response and biological control of Meloidogyne<br />

incognita on lettuce and tomato with rhizobacteria. J Plant Diseas Protect 1998; 105:349-358.<br />

Hyder, P.W., Fredrickson, E.L., Estell, R.E., Tellez, M., Gibbens, R.P. Distribution and concentration of total<br />

phenolics, condensed tannins, and nordihydroguaiaretic acid (NDGA) in creosotebush (Larrea tridentata).<br />

Biochem Systemat Ecol 2002; 30: 905-912.<br />

Inderjit. Plant phenolics in allelopathy. Bot Rev 1996; 62:186-202.<br />

Inderjit and Dakshini, K.M.M. Allelopathic potential of the phenolics from the roots of Pluchea lanceolata.<br />

Physiol Plant 1994; 92:571-576.<br />

Inderjit and del Moral, R. Is separating resource competition from allelopathy realistic? Bot Rev 1997; 63:221-<br />

230.<br />

Inderjit and Mallik, A.U. Effect of phenolic compounds on selected soil properties. Forest Ecol Manag 1997;<br />

92:11-18.<br />

Inderjit, Muramatsu, M., Nishimura, H. On the allelopathic potential of certain terpenoids, phenolics, and their<br />

mixtures, and their recovery from soil. Can J Bot 1997; 75: 888-891.<br />

Inderjit, Keating, K.I. Allelopathy: Principles, procedures, processes, and promises for biological control. Adv<br />

Agron 1999; 67:141-232.<br />

Inderjit and Duke, S.O. Ecophysiological aspects of allelopathy. Planta 2003; 217:529-539.<br />

Jackson, D.M. and Peterson, J.K. Sublethal effects of resin glycosides from the periderm of sweetpotato storage<br />

roots on Plutella xylostella (Lepidoptera: Plutellidae). J Econom Entomol 2000; 93:388-393.<br />

Jimenez-Osornio J.J., S.R. Gliessman. Allelopathic interference in a wild mustard (Brassica campestris L.) and<br />

broccoli (Brassica oleracea L. var. italica) intercrop agroecosystem. In: <strong>Allelochemicals</strong>, Role in<br />

Agriculture and Forestry, Waller G.R. ed. ACS Symp. Ser. 330. American Chemical Society, Washington<br />

DC, USA, 1987; pp. 262–274.<br />

John, J. and Narwal, S.S. Allelopathic plants. 9. Leucaena leucocephala (Lam.) de Wit. Allelopathy J 2003;<br />

12:13-36.<br />

Keyes, W.J., Taylor, J.V., Apkarian, R.P., Lynn, D.G. Dancing together. Social controls in parasitic plant<br />

development. Plant Physiol 2001; 127:1508-1512.<br />

Kloepper, J.W., Rodriguez-Kabana, R., Zehnder, G.W., Murphy, J.F., Sikora, E., Fernandez. C. Plant rootbacterial<br />

interactions in biological control of soilborne diseases and potential extension to systemic and<br />

foliar diseases. Australas. Plant Pathol 1999; 28:21-26.<br />

Langenheim, J. Higher plant terpenoids: A phytocentric overview of their ecological roles. J Chem Ecol 1994;<br />

20:1223-1280.<br />

Lehman, M.E., Blum, U. Evaluation of ferulic acid uptake as a measurement of allelochemical dose: Effective<br />

concentration. J Chem Ecol 1999; 25:2585-2600.<br />

Macias, F.A., J.C.G. Galindo, G.M. Massanet. Potential allelopathic activity of several sesquiterpene lactone<br />

models. Phytochemistry 1992; 31:1969-1977.<br />

Macias, F.A., Galindo, J.C.G., Castellano, D, Velasco, R.F. Natural products as allelochemicals. 11. Sesquiterpene<br />

lactones with potential use as natural herbicide models (I): trans,trans-germacranolides. J Agric Food<br />

Chem 1999; 47:4407-4414.<br />

Macias, F.A., Galindo, J.C.G., Castellano, D., Velasco, R. Sesquiterpene lactones with potential use as natural<br />

herbicide models. 2. Guaianolides. J Agric Food Chem 2002a; 48:5288-5296.<br />

Macias, F.A., Torres, A., Galindo, J.L.G., Varela, R., Alvarez, J,A., Molinillo, J.M.G. Bioactive terpenoids from<br />

sunflower leaves cv. Peredovick (R). Phytochem 2002b; 61:687-692.<br />

Mankau, R. Reduction of root-knot disease with organic amendments under semifield conditions. Plant Disease<br />

Reporter 1968; 52:315-319.<br />

Mazzola, M. Mechanisms of natural soil suppressiveness to soilborne diseases. Antonie Van Leeuwenhoek<br />

Internat. J Gen Molec Microbiol 2002; 81:557-564.<br />

Morris, J.B., Walker, J.T. Non-traditional legumes as potential soil amendments for nematode control. J Nematol<br />

2002; 34:358-361.<br />

Muscolo, A., Panuccio, M.R., Sidari, M. The effect of phenols on respiratory enzymes in seed germination -<br />

Respiratory enzyme activities during germination of Pinus laricio seeds treated with phenols extracted<br />

from different forest soils. Plant Growth Regul. 2001; 35:31-35.<br />

75


76<br />

ANA LUISA ANAYA<br />

Nagabhushana, G.G., Worsham, A.D., Yenish, J.P. Allelopathic cover crops to reduce herbicide use in sustainable<br />

agricultural systems. Allelopathy J 2001; 8:133-146.<br />

Nava-Rodriguez, V. Hernandez-Bautista, B.E., Cruz-Ortega, R., Anaya, A.L. Allelopathic potential of beans<br />

(Phaseolus spp.), other crops, and weeds from Mexico. Allelopathy J<br />

Neori, A., Reddy, K.R., Ciskova-Koncalova, H., Agami, M. Bioactive chemicals and biological-biochemical<br />

activities and their functions in rhizospheres of wetland plants. Bot Rev 2000; 66:350-378.<br />

Ngouajio, M., McGiffen, M.E. Going organic changes weed population dynamics. Hort Technol 2002; 12:590-<br />

596.<br />

Nilsson, M.C., Gallet, C., Wallstedt, A. Temporal variability of phenolics and batatasin-III in Empetrum<br />

hermaphroditum leaves over an eight-year period: interpretations of ecological functions. Oikos 1998;<br />

81:6-16.<br />

Ohno, T., Doolan, K., Zibilske, L.M., Liebman, M., Gallandt, E., Berube, C. Phytotoxic effects of red clover<br />

amended soils on wild mustard seedling growth. Agricult Ecosyst Environm 2000; 78:187-192.<br />

Pereda-Miranda, R., R. Mata, A.L. Anaya, J.M. Pezzuto, D.B.M. Wickramaratne, A.D. Kinghorn. Tricolorin<br />

A, major phytogrowth-inhibitor from Ipomoea tricolor. J Natural Products 1993; 56:571-582.<br />

Pérez de Luque, A., Galindo, J.C.G., Macías, F.A., Jorrin, J. Sunflower sesquiterpene lactone models induce<br />

Orobanche cumana seed germination. Phytochemistry 2000, 2000; 53:45-50.<br />

Persmark, L., H.B. Jansson. Nematophagous fungi in the rhizosphere of agricultural crops. Federation of European<br />

Microbiological Societies. Microbiology Ecology 1997; 22:303-312.<br />

Persmark, L., B. Nordbring-Hertz. Conidial trap formation of nematode-trapping fungi in soil and soil extracts.<br />

Federation of European Microbiological Societies. Microbiology Ecology 1997; 22:313-323.<br />

Peterson, J.K., Harrison, H.F. Sweet-potato allelopathic substance inhibits growth of purple nutsedge (Cyperusrotundus).<br />

Weed Technol 1995; 9:277-280.<br />

Picman, A.K. <strong>Biologica</strong>l activities of sesquiterpene lactones. Biochem Syst Ecol 1986; 4: 255-281.<br />

Picman, J., Picman, A.K. Autotoxicity in Parthenium hysterophorus and its possible role in control of germination.<br />

Biochem Syst Ecol 1984; 12:287-292.<br />

Politycka, B. Phenolics and the activities of phenylalanine ammonia-lyase, phenol-beta-glucosyltransferase and<br />

beta-glucosidase in cucumber roots as affected by phenolic allelochemicals. Acta Physiol Plant 1998;<br />

20:405-410.<br />

Politycka, B. Ethylene-dependent activity of phenylalanine ammonia-lyase and lignin formation in cucumber<br />

roots exposed to phenolic allelochemicals. Acta Soc Bot Pol 1999; 68:123-127.<br />

Postma J., Montanari M., Van den Boogert P.H.J.F. Microbial enrichment to enhance the disease suppressive<br />

activity of compost. European J Soil Biol 2003; 39:157-163.<br />

Putnam, A.R. <strong>Allelochemicals</strong> from plants as herbicides. Weed Technol 1988; 2:510–518.<br />

Pyrowolakis, A., Westphal, A., Sikora, R.A,. Becker, J.O. Identification of root-knot nematode suppressive<br />

soils. Appl Soil Ecol 2002; 19:51-56.<br />

Qiu, J., Hallmann, J., Kokalis-Burelle, N., Weaver, D.B., Rodriguez-Kabana, R., Tuzan, S. Activity and<br />

Differential Induction of Chitinase Isozymes in Soybean Cultivars Resistant or Susceptible to Root-knot<br />

Nematodes. J Nematology 1997; 29:523-530.<br />

Randhir, R., Lin, Y.T., Shetty, K. Stimulation of phenolics, antioxidant and antimicrobial activities in dark<br />

germinated mung bean sprouts in response to peptide and phytochemical elicitors. Process Biochem 2004;<br />

39:637-646.<br />

Reigosa, M.J., Souto, X.C., Gonzalez, L. Effect of phenolic compounds on the germination of six weeds species.<br />

Plant Growth Regul 1999; 28:83-88.<br />

Reitz M., Rudolph K., Schroder I., Hoffmann-Hergarten S., Hallmann J., Sikora R.A. Lipopolysaccharides of<br />

Rhizobium etli strain G12 act in potato roots as an inducing agent of systemic resistance to infection by<br />

the cyst nematode Globodera pallida. Appl Environ Microbiol 2000; 66(8):3515-3518.<br />

Rice, E.L. Allelopathy. 2 nd.<br />

Ed., Academic Press, Inc. London 1984.<br />

Rich, J.R., Rahi, G.S. Suppression of Meloidogyne javanica and M. incognita on tomato with ground seed of<br />

castor, Crotalaria, hairy indigo, and wheat. Nematropica 1995; 25:159-164.<br />

Ridenour, W.M., Callaway, R.M. The relative importance of allelopathy in interference: the effect of an invasive<br />

weed on a native bunchgrass. Oecologia 2001; 126:444-450.<br />

Rodríguez-Kábana, R. Organic and inorganic nitrogen amendments to soil as nematode suppressants. J<br />

Nematology 1986; 18:129-134.


ALLEOPATHIC ORGANISMS AND<br />

MOLECULES<br />

Rodríguez-Kábana, R., Kloepper, J.W. Cropping systems and the enhancement of microbial activities antagonistic<br />

to nematodes. Nematropica 1998; 28:144.<br />

Romagni, J.G., Meazza, G., Nanayakkara, N.P.D., Dayan, F.E. The phytotoxic lichen metabolite, usnic acid, is<br />

a potent inhibitor of plant p-hydroxyphenylpyruvate dioxygenase. FEBS Letters 2000; 480:301-305.<br />

Romero-Romero Teresa, Anaya, A.L., Cruz-Ortega, R. Screening the effects of allelochemical stress produced<br />

by some plants on cytoplasmic protein synthesis pattern of crop plants. J Chem Ecol 2002; 28:617-629.<br />

Roshchina, V.V. Molecular-cellular mechanisms in pollen allelopathy. Allelopathy J 2001; 8:11-28.<br />

Sashida, Y., Nakata, H. Shimomura, H. and Kagaya, M.. Sesquiterpene lactones from Pyrethrum flowers.<br />

Phytochemistry 1983; 22:1219-1222.<br />

Seigler, D.S. Chemistry and mechanisms of allelopathic interactions. Agronomy J 1996; 88:876–885.<br />

Singh, H.P., Batish, D.R., Kaur, S., Kohli, R.K. Phytotoxic interference of Ageratum conyzoides with wheat<br />

(Triticum aestivum). J Agron Crop Sci 2003a; 189:341-346.<br />

Singh, H.P., D.R. Batish, R.K. Kohli. Allelopathic interactions and allelochemicals: New possibilities for<br />

sustainable weed management. Crit Rev Plant Sc 2003b; 22:239-311.<br />

Singh, H.P., Batish, D.R., Pandher, J.K., Kohli, R.K. Assessment of allelopathic properties of Parthenium<br />

hysterophorus residues. Agric Ecosyst Environ 2003c; 95:537-541.<br />

Soler-Serratosa, A., Kokalis-Burelle, N., Rodriguez-Kabana, R., Weaver, C.F., King P.S. <strong>Allelochemicals</strong> for<br />

control of plant parasitic nematodes. Nematropica 1996; 26:57-71.<br />

Souto, C., Pellissier, F., Chiapusio, G. Allelopathic effects of humus phenolics on growth and respiration of<br />

mycorrhizal fungi. J Chem Ecol 2000; 26:2015-2023.<br />

Spencer, G.F., R.B. Wolf, D. Weisleder. Germination and growth inhibitory sesquiterpenes from Iva Axillaris<br />

seeds. J Nat Prod 1984; 47:730-732.<br />

Staman, K, Blum, U., Louws, F., Robertson, D. Can simultaneous inhibition of seedling growth and stimulation<br />

of rhizosphere bacterial populations provide evidence for phytotoxin transfer from plant residues in the<br />

bulk soil to the rhizosphere of sensitive species? J Chem Ecol 2001; 27:807-829.<br />

Stanislaus, M.A., Cheng, C.L. Genetically engineered self-destruction: an alternative to herbicides for cover<br />

crop systems. Weed Sci 2002; 50:794-801.<br />

Stevens, K.L., G.B. Merrill. Sesquiterpene lactones and allelochemicals from Centaurea species. In: The<br />

Chemistry of Allelopathy: Biochemical Interactions Among Plants. Thompson, A.C. ed. ACS Symp.<br />

Ser. 268. American Chemical Society. Washington, D.C., 1985.<br />

Sturz, A.V., Christie, B.R., Nowak, J. Bacterial endophytes: Potential role in developing sustainable systems of<br />

crop production. Crit Rev Plant Sci 2000; 19:1-30.<br />

Sturz, A.V., Kimpinski, J. Endoroot bacteria derived from marigolds (Tagetes spp.) can decrease soil population<br />

densities of root-lesion nematodes in the potato root zone. Plant Soil 2004; 262:241–249.<br />

Toxicity of Pesticides. Pesticide Education Program in Penn State’s College of Agricultural Sciences. The<br />

Pennsylvania State University. http://pubs.cas.psu.edu/FreePubs/pdfs/uo222.pdf 2004<br />

Upadhyay, R.K., Mukerji, K.G., Rajak, R.L. Integrated Pest Management. In: IPM System in Agriculture. Vol.<br />

I. Principles and Perspectives. Upadhyaya R.K., Mukerji K.G., Rajak R.L., eds. Aditya Books Pvt. Ltd.,<br />

New Delhi, 1996; pp. 1-23.<br />

Van Den Boogert, P.H.J.F., Velvis, H., Ettema, C.H., Bouwman, L.A. The role of organic matter in the population<br />

dynamics of the endoparasitic nematophagus fungus Drechmeria coniospora in microcosms. Nematologica<br />

1994; 40:249-257.<br />

Vargas-Ayala, R., Rodriguez-Kabana, R., Morgan-Jones, G., McInroy, J.A., Kloepper J.W. Shifts in soil microflora<br />

induced by velvetbean (Mucuna deeringiana) in cropping systems to control root-knot nematodes. Biol<br />

Control 2000; 17:11-22.<br />

Vargas-Ayala, R., Rodriguez-Kabana, R. Bioremediative management of soybean nematode population densities<br />

in crop rotations with velvetbean, cowpea, and winter crops. Nematropica 2001; 31:37-46.<br />

Vokou, D., Douvli, P., Blionis, G.J., Halley, J.M. Effects of monoterpenoids, acting alone or in pairs, on seed<br />

germination and subsequent seedling growth. J Chem Ecol 2003; 29: 2281-2301.<br />

Waller G.R., C.H. Chou. Phytochemical Ecology: <strong>Allelochemicals</strong>, Mycotoxins, and Insect Pheromones and<br />

Allomones. Institute of Botany. Academia Sinica. Taipei, Taiwan 1989.<br />

Waller, G.R. Introduction. In: Recent Advances in Allelopathy. Macías, F.A., Galindo, J.C.G., Molinillo, J.M.G.,<br />

and Cutler, H.G. eds. Vol. I A science for the Future. International Allelopathy Society and Servicio de<br />

Publicaciones de la Universidad de Cádiz, España 1999<br />

77


78<br />

ANA LUISA ANAYA<br />

Waller, G.R., M.C. Feng, Y. Fujii. Biochemical analysis of allelopathic compounds: plants, microorganisms,<br />

and soil secondary metabolites. In: Principles and Practices in Plant Ecology : Allelochemical<br />

interactions, Inderjit, K.M.M., Dakshini and Foy C.L. eds. CRC Press, Boca Raton, FL, USA. 1999, pp.<br />

75-98.<br />

Wang, K.H., Sipes, B.S., Schmitt, D.P. Crotalaria as a cover crop for nematode management: A review.<br />

Nematropica 2002; 32 (1):35-57.<br />

Weidenhamer J.D. Distinguishing resource competition and chemical interference: overcoming the methodological<br />

impasse. Agron J 1996; 88:866–875.<br />

Weller, S.C., Bressan, R.A., Goldsbrough, P.B., Fredenburg, T.B., Hasegawa, P.M. The effect of genomics on<br />

weed management in the 21st century. Weed Sci 2001; 49: 282-289.<br />

Weller, D.M., J.M. Raaijmakers, B.B.M. Gardener, L.S. Thomashow. Microbial populations responsible for<br />

specific soil suppressiveness to plant pathogens. Ann Rev Phytopath 2002; 40:309-348.<br />

Weston L.A. Utilization of allelopathy for weed management in agroecosystems. Agron J 1996; 88:860–866.<br />

Weston, L.A., Duke, S.O. Weed and crop allelopathy. Crit Rev Plant Sci 2003; 22:367-389.<br />

Wigchert, S.C.M., Zwanenburg, B. A critical account on the inception of Striga seed germination. J Agric Food<br />

Chem 1999; 47:1320-1325.<br />

Yenish, J.P., Worsham, A.D., York, A.C. Cover crops for herbicide replacement in no-tillage corn (Zea mays).Weed<br />

Technol 1996; 10:815-821.<br />

Yin, B; Scupham, AJ; Menge, JA; Borneman, J. Identifying microorganisms which fill a niche similar to that of<br />

the pathogen: a new investigative approach for discovering biological control organisms. Plant Soil 2004;<br />

259:19-27.<br />

Yu, J.Q., Matsui, Y. Effects of root exudates of cucumber (Cucumis sativus) and allelochemicals on ion uptake<br />

by cucumber seedlings. J Chem Ecol 1997; 23: 817-827.


SCOTT W. MATTNER<br />

THE IMPACT OF PATHOGENS ON PLANT<br />

INTERFERENCE AND ALLELOPATHY<br />

Department of Primary Industries, Knoxfield Centre, Victoria, Private Bag<br />

15, Ferntree Gully Delivery Centre, 3156, VIC, Australia.<br />

E-mail : scott.mattner@dpi.vic.gov.au<br />

Abstract. Pathogenesis can have both detrimental and beneficial impacts on plant fitness. As such, pathogens<br />

are important forces that influence the structure and dynamics in natural and manipulated plant ecosystems.<br />

Plant production and numbers within a community are constrained by environmental limitations, which are<br />

often mediated through plant interference. Competition for resources and allelopathy (chemical interactions)<br />

are the two most important ways that plants interfere with each other. This chapter reviews the effects of pathogens<br />

on the competitiveness and allelopathic ability of their hosts. In most cases, pathogens reduce the competitive<br />

ability of their host, making the host prone to displacement by neighbouring, resistant plants. However, pathogens<br />

may simultaneously increase the allelopathic ability of their hosts, thereby offsetting their loss in competitiveness<br />

to varying degrees. Evidence for enhanced allelopathy by infected plants comes in two forms: (i) pathogens<br />

stimulate the production of secondary metabolites by plants, many of which are implicated in allelopathy (eg<br />

phenolics), and (ii) field, glasshouse and bioassay studies showing that infected plants may suppress their<br />

neighbours more than healthy plants, under conditions of low competition. By conferring the benefit of increased<br />

allelopathy on their hosts, pathogens may maintain a self-advantage through increasing the survival chances of<br />

their hosts and ultimately themselves. The enhanced allelopathy of infected plants supports the ‘new function’<br />

hypothesis, which suggests that pathogens evolve toward a mutualistic relationship with their host through the<br />

appearance of strains with beneficial effects on the host in addition to their detrimental effects.<br />

1. INTRODUCTION<br />

Plant pathogens are disease agents that live in or on their hosts and there obtain<br />

nutriment to the overall detriment of the plant. Fungal pathogens, which cause about<br />

70% of all major crop diseases (Deacon, 1997), are often characterised on the basis of<br />

two extremes in trophism, as necrotrophs or biotrophs (Lutrell, 1974; Parbery 1996).<br />

Necrotrophic organisms obtain nutriment from necrotic host tissues, which they kill<br />

prior to colonisation. Consequently, these pathogens although destructive to the host<br />

have little effect on the physiology of the rest of the plant. Biotrophic pathogens draw<br />

nutriment directly from living host tissue and can have a critical effect on host<br />

physiology. Many thorough reviews concerning the impact of infection on host<br />

physiology already occur in the literature (Goodman et al., 1986; Burdon, 1987; Ayres,<br />

1991; Sutic and Sinclair, 1991). By example though, pathogens can alter the<br />

partitioning of assimilates and dry matter; reduce net photosynthesis and increase<br />

respiration; increase water loss through transpiration and vulnerability to drought;<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 79– 101.<br />

© 2006 Springer. Printed in the Netherlands.<br />

79


80<br />

SCOTT W. MATTNER<br />

and reduce the overall growth, reproductive capacity and yield of their hosts.<br />

Consequently, pathogens are important factors that influence the composition and<br />

dynamics of natural and manipulated plant ecosystems.<br />

Despite pathogens having an overall detrimental effect on their host, some aspects<br />

of infection are beneficial (Burdon, 1991; Parbery, 1996). For example, infection<br />

may promote vegetative growth (Catherall, 1966; Wennström and Ericson, 1991),<br />

deter or limit grazing by herbivores (Morgan and Parbery, 1980), or increase the<br />

host’s scavenging ability for nutrients and water (Ahmad et al., 1982; Paul and Ayres,<br />

1988). By conferring some benefit on their host, these pathogens maintain a selfadvantage<br />

through increasing the chances of survival of their host and ultimately<br />

themselves. This is particularly important for obligate biotrophs (e.g. rusts, Uredinales;<br />

powdery mildews, Erysiphaceae) that require the host for self-preservation.<br />

In contrast to previous chapters that consider mechanisms for exploiting<br />

allelopathy for pathogen control, this chapter addresses the reverse situation – the<br />

impact of pathogens on plant interference and allelopathy. It reviews evidence that<br />

pathogens decrease the competitiveness and simultaneously increase the allelopathic<br />

ability of their hosts. The ability of pathogens to increase their host’s allelopathic<br />

ability may be an important means by which pathogens confer a benefit on their host.<br />

In so doing, these pathogens encourage the continuation of their host’s genotype into<br />

proceeding generations. Furthermore, the ability of pathogens to enhance allelopathy<br />

in their hosts may be one way that obligate biotrophic pathogens are evolving toward<br />

multualism.<br />

2. PLANT INTERFERENCE AND ALLELOPATHY<br />

2.1. Components of Interference<br />

Harper (1961) introduced the term ‘interference’ to encompass all effects placed on<br />

an organism by the proximity of neighbours. Plant interference is ‘the response of an<br />

individual plant to its total environment as this is modified by the presence and/or<br />

growth of other individuals’ (Begon et al., 1986). Interference may have a positive or<br />

negative effect on plant growth and may occur between plants of different species,<br />

between individuals within the same species or even between individual organs on a<br />

single plant. Some of the important ways plants modify the environment in each<br />

other’s presence are through: competition, allelopathy, protection (e.g. when an<br />

unpalatable plant protects a neighbouring palatable plant from grazing), direct<br />

stimulation (e.g. when nitrogen fixed by a legume becomes available to a non-legume),<br />

direct contact (e.g. when a thorny plant mechanically abrades neighbouring plants in<br />

the wind), and non-competitive inhibition (e.g. when a tree provides a rubbing post<br />

for grazers and so encourages local trampling). Of these mechanisms, competition is<br />

the most dominant in shaping plant populations (Jolliffe, 1988; Tilman, 1988), but<br />

research is increasingly demonstrating the importance of allelopathy (Siegler, 1996;<br />

Wardle et al., 1998). Comprehensive reviews of competition ( Milthorpe, 1961; Tilman,


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

81<br />

1988; Grace and Tilman, 1990; Casper and Jackson, 1997) and allelopathy (Putnam<br />

and Duke, 1978; Rice, 1984; Inderjit et al., 1995; Anaya, 1999) already occur in the<br />

literature.<br />

The view of competition in the present chapter is ‘the interaction between<br />

individuals brought about by the shared requirements for a resource in limited supply,<br />

and leading to a reduction in the survivorship, growth and/or reproduction of the<br />

individuals concerned’ (Begon et al., 1986). Plants mostly compete for the resources<br />

of light, water and nutrients (Donald, 1963), and less frequently for carbon dioxide,<br />

oxygen and space (Trenbath, 1974). In the exact sense of the definition, plants do not<br />

compete so long as these resources are in excess of the needs of both. When plants do<br />

compete, however, the outcome always reduces the growth of at least one of the<br />

competitors. The outcome of competition between two plants depends on the ability<br />

of each species to secure and utilise resources, i.e. their competitiveness. A highly<br />

competitive plant may be one that has a high rate of uptake of a particular resource or<br />

a low requirement for that resource (Grace and Tilman, 1990). Plants may differ in<br />

their ability to compete for individual resources, while environmental differences may<br />

also influence their overall competitive ability. Differences in competitive ability, in<br />

turn, help to structure the composition of mixed plant communities.<br />

2.2. Allelopathy<br />

Molisch first advanced the term allelopathy in 1937. It derives from the Greek words<br />

‘allelon’, meaning of each other, and ‘pathos’, meaning to suffer (Mandava, 1985).<br />

Despite the origin of its root words, Molisch used the term to refer to the chemical<br />

interactions between all plants (higher plants and microorganisms), including<br />

stimulatory as well as inhibitory effects. Some authors have considered that the concept<br />

covers only inhibitory effects (Rice, 1974; Putnam, 1985; Boyette and Abbas, 1995),<br />

yet others exclude microorganisms from the definition (Putnam and Duke, 1978;<br />

Putnam, 1985; Pratley, 1996). Many inhibitory chemicals produced by plants, however,<br />

stimulate growth at low concentrations (Liu and Lovett, 1993; Pratley, 1996), and<br />

microorganisms can mediate allelopathy (Rice, 1992; Bremner and McCarty, 1993).<br />

For these reasons, the definition of allelopathy used in this chapter closely follows<br />

that of Molisch’s, ‘the beneficial and detrimental chemical interaction among [plant]<br />

organisms including microorganisms’ (Rice, 1984). Although some authors have<br />

confused allelopathy with competition, the distinguishing feature is that allelopathy<br />

involves an addition of a chemical to the environment, whereas competition involves<br />

the shared utilisation of some limited factor required for growth (Muller, 1969; Rice<br />

1984).<br />

Any chemical produced by a plant (donor) that stimulates or inhibits the growth<br />

of a neighbour (receiver or receptor) is broadly termed an allelochemical. Typically,<br />

allelochemicals are secondary metabolites (Whittaker and Feeney 1971; Rice, 1984;<br />

Rizvi et al., 1992), produced as by-products of the acetate and shikimic acid pathways.<br />

They may also form as degradation products from the action of microbial enzymes


82<br />

SCOTT W. MATTNER<br />

(Putnam, 1985). Initially, the reason why plants devote resources to the production of<br />

these compounds was not understood as they were regarded as functionless waste<br />

products (Mothes, 1955). It is now increasingly accepted, however, that these<br />

compounds function as defensive agents against pathogens, insects and neighbouring<br />

plants (allelopathy). There is an enormous diversity of allelochemicals produced by<br />

plants (Bansal, 1994), with classification based on chemical structure (Whittaker and<br />

Feeney, 1971; Mandava, 1985; Putnam 1985) or on origin and chemical properties<br />

(Rice, 1984). For example, Rice (1984) recognised 14 main categories of<br />

allelochemicals. Of these groups, however, the phenolics are considered by many as<br />

the most important (Putnam and Duke, 1979; Mandava, 1985; Inderjit, 1996).<br />

Putative allelochemicals have been isolated from a variety of different plant organs,<br />

including shoots, roots, flowers, rhizomes, fruit and seed (Rice, 1984). They are<br />

stored in cell vacuoles so as not to interfere with the donor plant itself (Chou, 1989).<br />

Furthermore, secondary metabolites may be bound to sugars as glycosides or occur as<br />

polymers and crystals rendering them ineffective against the donor (Whittaker and<br />

Feeney, 1971). The release of allelochemicals into the environment may occur through<br />

volatilisation, root exudation, leaching or plant residue decomposition (Rice, 1984).<br />

<strong>Allelochemicals</strong> induce a wide range of symptoms in receiver plants, ranging<br />

from sudden wilting and death (e.g. tomato (Lycopersicon esculentum) grown in the<br />

vicinity of black walnut (Juglans nigra), Hale and Orcutt, 1987), to the more common<br />

subtle changes in growth. Determination of this reaction depends on the mode of<br />

action, the concentration, and the susceptibility of the receiver plant to the<br />

allelochemical. Allelopathic effects may be direct, such as affecting plant metabolism<br />

and growth, or indirect, such as altering of soil properties and nutrient status (Inderjit<br />

and Weiner, 2001). Rizvi et al. (1992) explained 12 plant functions that allelochemicals<br />

may affect, including membrane permeability, stomata function and photosynthesis,<br />

and cytology and ultrastructure.<br />

3. THE INFLUENCE OF PATHOGENS ON INTERFERENCE<br />

AND ALLELOPATHY<br />

Natural plant populations increase in production and number of individuals until<br />

constrained by environmental limitations (Burdon, 1987). The constraint of plant<br />

growth by the environment is often mediated through plant interference. Therefore,<br />

the ability of plants to interfere with their neighbours is important in determining<br />

their abundance in a community. Plant pathogens generally reduce the development,<br />

production and longevity of their hosts. In plant communities, this ‘burden of a<br />

parasite’ may result in a partly unoccupied niche that resistant plants re-inhabit.<br />

Overall, pathogens play a significant role in the ecology of plant communities by<br />

maintaining (Peters and Shaw, 1996) or reducing (Burdon, 1991) species diversity,<br />

driving succession (Van der Putten et al., 1993), ensuring plants do not establish<br />

under their parents (Augsberger, 1984) and help determine the long-term composition<br />

of plant communities (Dobson and Crawley, 1994).


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

3.1. The Impact of Pathogens on Competition<br />

Several pioneering publications on plant competition suggest that pathogens might<br />

alter the balance of competition in mixed communities in favour of the resistant<br />

components (de Wit, 1960; Harper, 1977). Since then, there have been numerous<br />

reviews and theoretical interpretations of the impact of pathogens in natural<br />

communities and on competition (Chilvers and Brittain 1971; Burdon, 1982; 1987;<br />

1991; Dinoor and Eshed, 1984; Gates et al., 1986; Alexander, 1990; Ayres and Paul,<br />

1990; Clay, 1990; Paul, 1990; Dobson and Crawley, 1994; Jarosz and Davelos, 1995;<br />

Alexander and Holt, 1998; Mattner 1998), but empirical experimentation is less<br />

extensive. Empirical studies have usually involved binary mixtures of a host and a<br />

non-host, grown under optimal conditions of nutrient and water availability, and<br />

inoculated with a single, copious and homogenous dose of inoculum. With few<br />

exceptions, these studies show that the influence of a host-specific pathogen reduces<br />

the vigour of the host, rendering it less able to compete with a neighbouring non-host<br />

species (Burdon, 1987; Ayres and Paul, 1990; Mattner, 1998). As such, the combination<br />

of the ‘burden of a parasite’ and competition can have a devastating effect on a plant.<br />

For example, Groves and Williams (1975) demonstrated that the combined effects of<br />

competition from subterranean clover (Trifolium subterraneum) and rust infection<br />

(Puccinia chondrillina) reduced the yield of skeleton weed (Chondrilla juncea) by<br />

94%. This combined effect was more marked than the action of either the competitor<br />

(yield was reduced by 70%) or the rust (yield was reduced by 51%) alone. Similarly,<br />

Friess and Maillet (1996) found that infection by cucumber mosaic virus reduced the<br />

vegetative yield of its host purslane (Portulaca oleracea), with this effect intensifying<br />

when infected plants were in competition with healthy plants. In mixtures of common<br />

lambsquarters (Chenopodium album) and corn (Zea mays) or beetroot (Beta vulgaris),<br />

foliar infection by Ascochyta caulina reduced the competitiveness of its host<br />

(lambsquarters) and increased the yield of non-host crops. In corn, infection negated<br />

the effects of competition from lambsquarters altogether, highlighting its potential as<br />

a biological control agent (Kempenaar et al., 1996). Despite pathogens reducing the<br />

competitiveness of their hosts, the effect of infection is often greater on a neighbouring<br />

non-host than on the host itself. For example, in mixtures of groundsel (Senecio<br />

vulgaris) and lettuce (Lactuca sativa), rust (Puccinia lagenophorae) increased the<br />

biomass of lettuce (the non-host) markedly, while only reducing that of its host<br />

groundsel marginally (Paul and Ayres, 1987a).<br />

Competitive stress for limited resources intensifies as plant density increases<br />

(Shinozaki and Kira, 1956). Consequently, the effect of a pathogen on the<br />

competitiveness of its host is most devastating at high densities. For example, rust<br />

reduced the competitive ability of groundsel in mixtures with healthy groundsel (Paul<br />

and Ayres, 1986) or with lettuce (Paul and Ayres, 1987a) more so at high densities<br />

than at low densities. Similarly, Ditommaso and Watson (1995) demonstrated that<br />

anthracnose (Colletotrichum coccodes) was more detrimental to the growth of<br />

velvetleaf (Abutilon theophrasti) in mixtures with soybean (Glycine max) at high<br />

plant densities. Conversely, the performance of the non-host, relative to the infected<br />

83


84<br />

SCOTT W. MATTNER<br />

host, generally increases as density increases. For example, the ability of lettuce to<br />

compete with rusted groundsel was greater at high densities than at low densities<br />

(Paul and Ayres, 1987a).<br />

Considering that competition occurs for shared resources that are in limited supply,<br />

it is not surprising that resource availability influences the interaction between<br />

competition and infection by pathogens. Paul and Ayres (1987b) examined the effect<br />

of rust on competition between infected and healthy groundsel under conditions of<br />

drought. They found that rust reduced the competitiveness of groundsel, however,<br />

this was more so in droughted plots than in well-watered plots. They concluded that<br />

water stress is important in determining the impact of rust in mixed populations in<br />

the field. Owing to the effects pathogens can have on nutrient uptake, Paul and Ayres<br />

(1990) also studied the effect of nutrient supply on the interaction between rust and<br />

competition. Under high nutrition, groundsel was more competitive than shepherd’s<br />

purse (Capsella bursa-pastoris). This superiority was lost, however, following the<br />

inoculation of groundsel with rust. In contrast, shepherd’s purse had a greater<br />

competitive ability than groundsel under nutrient-poor conditions. Despite this, it<br />

did not increase its advantage when groundsel was rusted.<br />

While most studies show that pathogens decrease the competitiveness of their<br />

hosts and increase that of neighbouring non-hosts, there are several exceptions. For<br />

example, Catherall (1966) found that for most of the year, barley yellow dwarf virus<br />

reduced the competitive ability of its host, perennial ryegrass (Lolium perenne), when<br />

grown with white clover (Trifolium repens). During spring, however, the virus<br />

stimulated tillering in ryegrass and increased its competitiveness. In another example,<br />

the rust Puccinia pulsatillae sterilises its host, Pulsatilla pratensis, by inhibiting<br />

flowering. Yet, infected plants are more vigorous and produce more leaves than<br />

healthy plants. In a natural community, Wennström and Ericson (1991) found that<br />

diseased plants had a greater survival rate than healthy plants, which they postulated<br />

was due to their greater competitiveness. The effects of pathogens on allelopathy may<br />

also mediate their influence on competition.<br />

3.2. The Impact of Pathogens on Allelopathy<br />

The term allelopathy is seldom used in plant pathology (Rice, 1984) even though, by<br />

definition, allelopathy includes chemical interactions involving microorganisms.<br />

Furthermore, the same or similar secondary metabolites implicated in allelopathy<br />

between plants are also important in plant pathology in: enhancing the germination<br />

of fungal spores (Odunfa, 1978); antibiosis between microorganisms (Di Pietro, 1995);<br />

regulating fungal growth and development (Calvo et al., 2002); pathogen/host<br />

recognition (Nicol et al., 2003); the development of disease symptoms (Daly and<br />

Deverall, 1983); the promotion of infection through the suppression of the host<br />

(Toussoun and Patrick, 1963); the breaking of fungistasis (Mol, 1995); and host<br />

resistance to pathogens. Similarly, some authors do not consider chemical interactions<br />

by microorganisms as part of allelopathy (Putnam and Duke, 1978; Putnam and Tang,<br />

1986; Pratley, 1996). Rice (1984) explained that this reluctance is due to the chemicals


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

involved not always escaping to the environment. This rationale seems invalid,<br />

however, because the chemicals involved do enter the environment of the pathogen or<br />

the plant. Moreover, microorganisms can mediate allelopathy between plants (Rice,<br />

1992; Bremner and McCartey, 1993). Such difficulties may have hindered the study<br />

of the impact of pathogens on plant allelopathy in the past.<br />

Both Rice (1984) and Einhellig (1995) hypothesised that pathogens enhance<br />

their host’s allelopathic ability, but few studies have observed such a relationship<br />

(Tang et al., 1995). This is despite several investigations on the effects of pathogens<br />

on plant competition (Burdon, 1987; Ayres and Paul, 1990; Section 3.1), all of which<br />

could potentially include allelopathic interactions. However, competition may have<br />

obscured allelopathy in these experiments (Trenbath, 1974), since competition is<br />

usually the dominant process of interference (Joliffe, 1988; Tilman, 1988). This is<br />

particularly so considering that many experiments studying the effect of pathogens<br />

on competition have utilised high plant densities, where competitive effects are most<br />

intense. Evidence supporting the hypothesis that pathogens can increase allelopathy<br />

between plants occurs in at least two forms: (i) pathogens can stimulate secondary<br />

metabolite production in their hosts, and (ii) field, glasshouse and bioassay experiments<br />

signifying an increased allelopathic ability of infected plants.<br />

3.2.1. Effect of Pathogens on Secondary Metabolite Production<br />

Numerous studies document the ability of plant pathogens to stimulate the metabolic<br />

activity (Daly, 1976) and increase the production of secondary metabolites (Stoessl,<br />

1982, 1983; Goodman et al., 1986; Kuc, 1997) by their hosts. Müller and Börger<br />

(1941) first proposed that plants produce defensive substances called phytoalexins in<br />

response to infection, which are important to host resistance. Phytoalexins are ‘low<br />

molecular weight; antimicrobial compounds that are both synthesised by and<br />

accumulated in plant cells after exposure to microorganisms’ (Paxton, 1981). As such,<br />

phytoalexins fit the definition of allelochemicals. Like allelochemicals that act against<br />

plants, phytoalexins are secondary compounds that belong to a wide range of different<br />

chemical classes (Stoessl, 1982), with more than 300 distinct phytoalexins already<br />

characterised (Smith, 1996). They are mostly synthesised via the acetate and shikimic<br />

acid pathways (Bailey, 1982; Bennett and Wallsgrove, 1994), as are allelochemicals<br />

that act against plants. Notwithstanding their similarities, allelopathy between plants<br />

and phytoalexin research has developed almost independently.<br />

Many compounds with phytoalexin activity are also implicated in allelopathy<br />

between plants (Rice, 1984). For example, isoflavonoids are important phytoalexins<br />

(Ingham, 1982; Paxton, 1981; Dakora and Phillips, 1996) and allelochemicals (Tamura<br />

et al. 1967; 1969) from the Leguminosae. Parbery et al. (1984) found that the<br />

isoflavonoids biochanin A, formononetin and genistein increased in subterranean<br />

clover by 62%, 123% and 75% respectively following infection by pepper spot<br />

(Leptosphaerulina trifolii). In comparison, Tamura et al. (1967; 1969) isolated a<br />

succession of isoflavonoids (including biochanin A, formononetin and genistein) from<br />

the shoots of red clover (Trifolium pratense) that inhibited its own germination by<br />

50% at concentrations of 50 ppm.<br />

85


86<br />

SCOTT W. MATTNER<br />

As expected, most studies concerned with the toxicology of phytoalexins<br />

concentrate on their effects on microorganisms. Increasingly, however, studies show<br />

that phytoalexins are also toxic to plants. Despite this, they are seldom referred to as<br />

allelochemicals. Pisatin and phaseollin, pterocarpans produced by pea (Pisum sativum)<br />

and bean (Phaseolus vulgaris) respectively, were the first phytoalexins characterised<br />

(Perrin and Bottomly, 1962; Cruickshank and Perrin 1963). Skipp et al. (1977) noted<br />

that phaseollin inhibited the respiration and growth of cell cultures of bean and tobacco<br />

(Nicotiana tabacum), eventually causing cell death. Similarly, pisatin reduced the<br />

growth of callus cultures of pea (Bailey, 1970) and inhibited the root growth of wheat<br />

(Cruickshank and Perrin, 1961). The phytoalexin rishitin (a sesquiterpene) accumulates<br />

in potato cells challenged by incompatible isolates of Phytophthora infestans<br />

(Tomiyama et al., 1968). Studies show that the compound also inhibits pollen<br />

germination in three Solanum species (Hodgkin and Lyon, 1979); causes lysis of<br />

potato and tomato protoplasts (Lyon and Mayo, 1978); and cell death in tubers and<br />

epidermal strips of potato (Ishiguri, et al., 1978; Lyon, 1980). Other studies show that<br />

phytoaxelins may inhibit seed germination (Chang et al., 1969), growth (Glazener<br />

and VanEtten, 1978) and cellular metabolism and function (Lyon, 1980; Boydston et<br />

al., 1983; Kurosaki et al., 1984; Giannini et al., 1990; Spessard et al., 1994) of plants.<br />

The ability of pathogens to stimulate secondary metabolite production in their hosts<br />

and for these to affect plant growth provides strong circumstantial evidence for the<br />

hypothesis that pathogens can increase plant allelopathy.<br />

Many studies have established that pathogens can increase the production of<br />

phenolic acids in their hosts, which are perhaps the most important group of plant<br />

allelochemicals. For example, when challenged by a range of different pathogens,<br />

concentrations of phenolics often associated in allelopathic interactions have increased<br />

in a variety plants, including carrot (Daucus carota) (Phan et al., 1991), chickpea<br />

(Cicer arietinum) (Singh et al., 2002); date palm (Phoenix dactylifera) (Daayf et al.,<br />

2003); potato (Solanum tuberosum) (Kuc et al., 1956); sorghum (Sorghum bicolor)<br />

(Woodhead, 1981); sugar cane (Saccharum officinarum) (Legaz et al., 1998); sunflower<br />

(Helianthus anuus) (Spring et al., 1991); and wheat (Triticum aestivum) (Siranidou<br />

et al., 2002), amongst many others. This increase in phenolics post-infection is one of<br />

the most important methods of disease resistance in plants. For example, inoculation<br />

with crown rust (Puccinia coronata f.sp. avenae) induced the production of three<br />

phenolic compounds (avenalumins I, II and III) by resistant oats (Avena sativa).<br />

Purified preparations of these chemicals inhibited the germination and germ tube<br />

growth of crown and stem rust (Puccinia graminis) at concentrations as low as 200<br />

µg/mL (Mayama, 1981; 1982). Similarly, Mandavia et al. (2000) found that varieties<br />

of cumin (Cuminum cyminum) tolerant of Fusarium wilt (Fusarium oxysporum f.sp.<br />

cumini) contained higher concentrations of salicylic acid, hydroquinone and<br />

umbelliferone in their root, stem and leaf tissues than susceptible varieties. These<br />

phenolics inhibited fungal spore germination and mycelial growth of F. oxysporum.<br />

However, to this author’s knowledge, the effect of the increased phenolic concentrations<br />

in infected plants on plant allelopathy has not been investigated empirically.


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

The production of glucosinolates by plants in the Capareles and their subsequent<br />

hydrolysis to toxic isothiocyanates, thiocyanates, and nitriles has been one of the<br />

most intensively studied systems in allelopathy, due partly to the similarity of these<br />

breakdown products to some synthetically produced soil fumigants (and therefore<br />

termed biofumigation; Kirkegaard et al., 2000). Glucosinolates are important in plant<br />

defence against insects, pathogens and nematodes. Jay et al. (1999) showed that<br />

infection of Brassica napus with beet western yellows virus increased glucosinolate<br />

concentration in tissues by 14%. Similarly, Li et al. (1999) found that infection of B.<br />

napus by Sclerotinia sclerotiorum increased glucosinolate content in resistant, but<br />

not in susceptible varieties. Exposure to the pathogenic bacterium, Erwinia carotovora,<br />

triggered the production of glucosinolates in Arabidopsis thaliana (Brader et al.,<br />

2001). The hydrolysis products from these glucosinolates inhibited the growth of E.<br />

carotovora in culture. Furthermore, Tierens et al. (2001) demonstrated that a range<br />

of pathogens were more aggressive in infecting an A. thaliana mutant that did not<br />

produce glucosinolates than the wild type, suggesting the importance of glucosinolates<br />

in protecting against infection. However, the role of glucosinolates in disease resistance<br />

may be species specific, since Andreasson et al. (2001) found that infection of B.<br />

napus by Leptosphaeria maculans had no effect on glucosinolate concentration in the<br />

resistant or susceptible host. Despite the ability for at least some pathogens to increase<br />

glucosinolate production in the Capareles, no one has investigated whether this directly<br />

translates to an increased allelopathic effect against neighbouring plants.<br />

3.2.2. The Effect of Rust on Ryegrass Allelopathy<br />

Perennial ryegrass (Lolium perenne) and white clover (Trifolium repens) are important<br />

components of improved pastures grown in temperate regions worldwide. The ability<br />

of ryegrass to become dominant in pastures has led to numerous investigations that<br />

demonstrate its potential to interfere with companion plants through allelopathy (Naqvi,<br />

1972; Naqvi and Muller, 1975; Newman and Rovira, 1975; Newman and Miller,<br />

1977; Gussin and Lynch, 1980; Buta et al., 1987; Quigley et al., 1990; Sutherland<br />

and Hoglund, 1990; Wardle et al., 1991; Prestidge et al., 1992; Chung and Miller,<br />

1995; Mattner, 1998; Mattner and Parbery, 2001). For example, in a study investigating<br />

the allelopathic ability of nine pasture species, Takahashi et al. (1988) found that<br />

leachate from soil surrounding ryegrass was the most inhibitory to the growth of the<br />

target species, including clover. In a subsequent experiment, they circulated nutrient<br />

solution between the roots of ryegrass and clover to eliminate competition effects.<br />

They found that the growth of clover declined in the system, particularly when the<br />

proportion of ryegrass was high. When they incorporated XAD-4 resin into the system<br />

(which selectively traps organic hydrophobic compounds) clover grew normally.<br />

Moreover, ryegrass became yellow and stunted when grown alone in the system, but<br />

this was prevented by the presence of the resin or clover (Takahashi et al., 1991). In<br />

a further experiment, root exudate from ryegrass not only inhibited the growth of<br />

clover, but also lettuce seedlings. The phytotoxic fraction of the extract contained pmethoxybenzoic,<br />

lauric, myristic, pentadecanoic, palmitoleic, palmitic, oleic and stearic<br />

87


88<br />

SCOTT W. MATTNER<br />

acids. Sodium salts of myristic, palmitic, oleic and stearic acids suppressed the growth<br />

of clover at concentrations as low as 5 ppm (Takahashi et al., 1993). Similarly, in a<br />

study examining the allelopathic effects of a number of crop and pasture species,<br />

Halsall et al. (1995) found that aqueous extract from the dried shoots of perennial<br />

ryegrass suppressed the germination, radicle elongation, nodulation and seedling root<br />

elongation of subterranean and white clover. The magnitude of this inhibition increased<br />

as the concentration of the extract increased.<br />

Crown rust, caused by Puccinia coronata f.sp. lolii, is the most devastating fungal<br />

disease of ryegrass, with epidemics regularly occurring between spring and autumn<br />

in temperate regions worldwide (Mattner and Parbery, 2001). Severe epidemics reduce<br />

ryegrass tillering by 20-38% (Lancashire and Latch, 1966; Mattner, 1998), leaf<br />

emergence by 60%, leaf area by 62%, root growth by 75% (Mattner, 1998), and increase<br />

the rate of leaf senescence by up to 184% (Lancashire and Latch, 1966; Trorey, 1979;<br />

Plummer et al., 1990; Mattner, 1998). Losses of herbage yield in ryegrass from rust<br />

have been as great as 94% (Critchett, 1991), with seed yield losses ranging from 12-<br />

36% (Hampton, 1986; Mattner 1998). Furthermore, rust infection reduces forage<br />

quality (Isawa et al., 1974; Trorey, 1979; Potter, 1987) and palatability to grazers<br />

(Cruickshank, 1957; Heard and Roberts, 1975).<br />

As would be expected by the devastating effect that rust has on ryegrass growth,<br />

most studies show that rust reduces the competitiveness of ryegrass with non-host<br />

plants such as clover. For example, in mixed swards of ryegrass and clover, Lancashire<br />

and Latch (1970) found that rust reduced ryegrass yield by 84% and increased the<br />

yield of clover by 87%. Furthermore, the proportion of clover in the rusted sward<br />

increased from 24% at the beginning to 80% at the termination of their experiment.<br />

Thus, their study pointed to a lowered competitiveness of rusted ryegrass. In mixtures<br />

of rust resistant and susceptible ryegrass, Potter (1987) found that rust reduced the<br />

yield of the susceptible cultivar and increased that of the resistant one, concluding<br />

that rust reduced ryegrass competitiveness. Similarly, crown rust infection in swards<br />

of ryegrass and cocksfoot (Dactylis glomerata) reduced ryegrass composition from<br />

30% to 15%, and was more marked in rust susceptible than resistant cultivars (Trorey,<br />

1979). However, in a series of experiments, Mattner (1998) reported an anomaly to<br />

the results of this previous research.<br />

In pot studies consisting of 50:50 mixtures of ryegrass and clover grown over a<br />

range of plant densities, Mattner (1998) found that rust reduced the yield of ryegrass<br />

by an average of 41%. However, interference from rusted ryegrass suppressed clover<br />

biomass by up to 47% compared with interference from the more productive, healthy<br />

ryegrass. The onset of the suppression of clover by rusted ryegrass was rapid, occurring<br />

as early as 6-13 days after inoculation, which according to some growth parameters<br />

was earlier than the effects of rust on ryegrass itself. The suppression of clover by<br />

rusted ryegrass was greatest at low plant densities and diminished or disappeared as<br />

density increased. In a separate trial, rusted ryegrass again suppressed clover growth,<br />

even after the removal of infected tissue by cutting and after the death of the ryegrass.<br />

In this instance, ryegrass killed by infection, with a competitive ability of virtually<br />

zero, inhibited the growth of clover more than living plants of healthy ryegrass. In


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

further trials, high rust severity and high soil moisture contents increased the<br />

suppression of clover by rusted ryegrass.<br />

The ability of rust to directly inhibit or infect clover could not explain the results<br />

from these studies, since clover is a non-host of P. coronata and inoculation with this<br />

rust did not reduce the yield of clover monocultures. Under some conditions, infection<br />

by rusts can increase the scavenging ability of some hosts for water and nutrient<br />

resources (Ahmad et al., 1982; Paul and Ayres, 1988), potentially increasing their<br />

competitive ability. For this to explain the results from Mattner’s studies, however,<br />

the suppression of clover by rusted ryegrass should have been greatest at high plant<br />

densities where competition for resources was most intense. Instead, the suppression<br />

of clover by rusted ryegrass was greatest at low densities where resources were plentiful<br />

and competition was low. Furthermore, the ability of rusted ryegrass to suppress<br />

clover continued even beyond the death of the plant, when it had no capacity to scavenge<br />

resources. For these reasons, Mattner (1998) posed the hypothesis that rust reduces<br />

the competitiveness of ryegrass, while simultaneously increasing its allelopathic ability.<br />

In this way, it was expected that the expression of allelopathy by rusted ryegrass was<br />

greatest at low densities, where there was little competition and the effects of rust in<br />

reducing ryegrass competitiveness did not obscure its effects on increasing allelopathy.<br />

Furthermore, high plant densities may detoxify or dilute the action of allelochemicals<br />

(Thijs et al., 1994).<br />

To test the validity of this hypothesis and to separate the effects of competition<br />

and allelopathy, four bioassays for allelopathy were conducted (Mattner, 1998; Mattner<br />

and Parbery, 2001). Each bioassay highlighted the potential for extracts, leachate, or<br />

residues from ryegrass to inhibit the yield of clover through allelopathy, and for rust<br />

to enhance this potential. For example, soil previously growing rusted ryegrass<br />

suppressed clover biomass by 36% compared with soil previously growing healthy<br />

ryegrass. Similarly, leachate from soil supporting rusted ryegrass suppressed clover<br />

biomass by 27% compared with that from healthy ryegrass (Mattner and Parbery,<br />

2001). Although some bioassays have confounding and interpretational problems<br />

(Stowe, 1979; Inderjit and Weston, 2000), the conformity of results between these<br />

different bioassays provides strong evidence for the hypothesis that rust infection<br />

increases the allelopathic ability of ryegrass. Furthermore, in the field, the proportion<br />

of prickly lettuce (Lactuca serriola) reduced in areas of a depleted pasture dominated<br />

by rusted ryegrass, compared with areas dominated by non-rusted ryegrass (Mattner,<br />

1998). Further bioassays provided evidence that this association potentially related<br />

to an enhanced allelopathic ability of ryegrass rather than to differences in soil<br />

chemistry. Thus, this study highlighted the potential for rust to increase ryegrass<br />

allelopathy in the field.<br />

Mattner (1998) suggested two mechanisms by which rust may increase ryegrass<br />

allelopathy. Firstly, rust may directly stimulate the production of allelochemicals by<br />

ryegrass in a defensive response to infection. Alternatively, or additionally, the increased<br />

rate of tissue senescence in ryegrass induced by rust may result in a higher concentration<br />

of plant residues reaching the soil. These residues may then form an allelochemical<br />

89


90<br />

SCOTT W. MATTNER<br />

source – either directly or following their decomposition. Most evidence gathered<br />

from his studies supported the first hypothesis. For example, when ryegrass residues<br />

were incorporated into soil, their ability to suppress clover growing in that soil depended<br />

on the residues being rusted and not their overall concentration. Furthermore, the<br />

knowledge that pathogens stimulate secondary metabolite formation in their hosts<br />

and the rapid effect rusted ryegrass had in suppressing clover, supported rust directly<br />

stimulating allelochemical production in ryegrass.<br />

Mattner’s findings appear to contradict those of Lancashire and Latch (1970)<br />

who studied the identical biological system in the field, and found that the proportion<br />

of clover in the sward more than doubled following infection of the ryegrass component<br />

by rust. However, Lancashire and Latch (1970) conducted their study under conditions<br />

that favoured the expression of the reduced competitiveness rusted ryegrass, i.e. at<br />

higher plant densities (2150 plants/m 2 ) than those of Mattner (1998) (57 plants/m 2 ).<br />

More importantly, however, their results only occurred in the highly susceptible ryegrass<br />

cultivar, Ruanui. In a more resistant cultivar, Ariki, rust reduced ryegrass yield by<br />

18%, but clover was unable to take advantage of this reduction, producing the same<br />

dry weight when grown with rusted ryegrass as when grown with healthy ryegrass.<br />

Furthermore, rusted Ariki ryegrass actually suppressed the growth of clover at some<br />

harvests. Perhaps rust infection increased the production of defensive chemicals by<br />

Ariki, thereby increasing its resistance to rust and its allelopathic ability against clover.<br />

Lancashire and Latch (1970) disregarded these results because, overall, the yield of<br />

Ariki was abnormally poor in their experiment compared with several previous studies.<br />

Nonetheless, the poor yield of Ariki ryegrass occurred in both rusted and non-rusted<br />

treatments and does not explain the inability of clover to compensate for the reduced<br />

yield of rusted ryegrass, or the suppression of clover by rusted ryegrass. Rather, the<br />

hypothesis that rust increases the allelopathic response of ryegrass while also reducing<br />

its competitiveness fits their results.<br />

Although there is strong circumstantial evidence from field, glasshouse and<br />

bioassay studies that rust increases the allelopathic ability of ryegrass, many questions<br />

remain. An important next step is to identify the allelochemicals concerned. Also,<br />

do rusted plants simply produce allelochemicals in higher concentrations or do they<br />

produce entirely different allelochemicals to healthy plants? Also, what is the<br />

allelochemical source in the plant and the mechanism of release to the environment?<br />

This information will provide a clearer understanding of the influence of pathogens<br />

on allelopathy.<br />

3.2.3. The Effect of Neotyphodium lolii on Ryegrass Allelopathy<br />

The endophytic fungus Neotyphodium lolii commonly infects perennial ryegrass<br />

forming a mutualistic relationship with its host. Apart from obtaining nutriment, the<br />

host provides the endophyte with a relatively exclusive niche and a vehicle for its<br />

transmittance through infected seed (Clay, 1987). The host benefits in several ways,<br />

including: (i) increased seed germination, dry matter production and tillering compared<br />

with non-infected ryegrass (Latch et al., 1985; Quigley, 2000); (ii) resistance to plant


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

diseases caused by nematodes (Stewart et al., 1993), fungi (Latch, 1993) and viruses<br />

(Lewis and Day, 1993); (iii) increased tolerance of drought stress (Ravel et al., 1997),<br />

(iv) increased nitrogen use efficiency (Arachevaleta et al., 1989); and (v) increased<br />

competitiveness (Sutherland and Hoglund, 1989). Additionally, the fungus produces<br />

various alkaloids (e.g. peramine, ergovaline, and lolitrem) that protect the host against<br />

herbivory (Clay, 1996).<br />

There is much speculation as to the role of the endophyte on the allelopathic<br />

ability of its host. This speculation originated with the observation that endophyteinfected<br />

ryegrass suppressed the growth of companion clovers to a greater extent than<br />

endophyte-free ryegrass (Stevens and Hickey, 1990). In order to explore several<br />

hypotheses on how this relationship might arise, Sutherland and Hoglund (1989)<br />

grew swards of endophyte-infected and endophyte-free perennial ryegrass in plots<br />

with white clover. Endophyte-infected ryegrass produced 16% more dry matter and<br />

suppressed the yield of clover by 72% compared with endophyte-free ryegrass. Neither<br />

mowing nor grazing by sheep affected the relationship. This demonstrated that selective<br />

grazing pressure on clover following a decline in the palatability of endophyte-infected<br />

ryegrass was not responsible for the reduction in clover yield. Rather, they suggested<br />

that the suppression was due to an increased competitiveness and allelopathic ability<br />

of endophyte-infected ryegrass. They implicated allelopathy in this effect because<br />

endophyte-infected ryegrass of a comparable yield to endophyte-free ryegrass,<br />

suppressed the yield of clover to a greater extent than endophyte-free ryegrass.<br />

In a subsequent experiment, Sutherland and Hoglund (1990) surrounded individual<br />

plants of white clover with zero to seven endophyte-infected or endophyte-free perennial<br />

ryegrass plants. This experiment failed to demonstrate that endophyte-infected ryegrass<br />

suppressed clover more than endophyte-free ryegrass. They explained that this might<br />

be due to the cutting regime imposed on ryegrass and the non-return of these clippings<br />

to the soil, which possibly contained allelochemicals. A bioassay experiment, which<br />

showed that aqueous extracts from the shoots of endophyte-infected ryegrass inhibited<br />

the shoot growth of clover seedlings by 8%, supported their explanation. In a similar<br />

experiment, Quigley et al. (1990) found that aqueous extracts from endophyte-infected<br />

ryegrass depressed the root length of four germinating legumes (including white<br />

clover), by an average of 10% compared with extract taken from endophyte-free<br />

ryegrass. In contrast, after conducting several bioassays for allelopathy, Prestidge et<br />

al. (1992) found little evidence to suggest that the presence of endophyte in ryegrass<br />

enhanced its allelopathic effect. Field trials also failed to show that clover yield<br />

declined when grown with endophyte-infected ryegrass. Similarly, Watson et al. (1993)<br />

and Mattner (1998) failed to substantiate an increased allelopathic effect of endophyteinfected<br />

ryegrass. Furthermore, there was no apparent interaction between endophyte<br />

and rust infection in increasing the allelopathic ability of ryegrass (Mattner, 1998).<br />

Applebee et al. (1999) found that the toxicity of tall fescue (Festuca arundinacea)<br />

infected with Neotyphodium coenophialum increased at elevated concentrations of<br />

CO 2 in the atmosphere. In a bioassay study conducted in sterile sand, Sutherland et<br />

al. (1999) applied aqueous extracts from ryegrass to potted clover seedlings. Extracts<br />

91


92<br />

SCOTT W. MATTNER<br />

from three ryegrass cultivars infected with three different strains of endophyte, all<br />

inhibited the growth of clover, by up to 27% compared with extracts from endophytefree<br />

ryegrass. Both ryegrass cultivar and endophyte strain influenced the degree that<br />

ryegrass extracts inhibited clover, but this did not relate to the type of alkaloids produced<br />

by the different endophyte strains. These studies suggest that both environmental<br />

and genetic influences may moderate the triggers for enhanced allelopathy by endophyte<br />

infected grasses, and this may explain the discrepancy in results between individual<br />

studies. Nonetheless, the majority of evidence suggests that endophyte infection has<br />

the capacity to increase ryegrass allelopathy, which is a further added benefit conferred<br />

by this mutualist to its host. Although the endophyte is a non-pathogenic organism,<br />

its ability to stimulate plant allelopathy adds further weight to the hypothesis that<br />

infection increases the allelopahtic ability of host plants.<br />

3.2.4. Other Systems<br />

The ability of rusts to stimulate allelopathy in their hosts may not be limited to ryegrass,<br />

as the rusts Puccinia hordei and Uromyces troflii-repentis increased the suppression<br />

of white clover by barley grass (Hordeum leporinum) and subterranean clover (Trifolium<br />

subterraneum), respectively (Mattner, 1998). Yet, the effect was not universal since<br />

Puccinia coronata in wild oat (Avena fatua), Puccinia graminis in cocksfoot (Dactylis<br />

glomerata) and Puccinia recondita in soft brome (Bromus mollis) all failed to increase<br />

their host’s allelopathic ability.<br />

Kong et al. (2002) studied the allelopathic potential of goatweed (Ageratum<br />

conyzoides) under different environmental stresses, including infection by powdery<br />

mildew (Erysiphe cichoracearum). Infection stimulated the production of 17 of the<br />

24 volatile chemicals produced by goatweed that they investigated, with total volatile<br />

production increasing by 50%. Exposure to the volatiles released by infected goatweed<br />

stimulated the growth of peanut (Arachis hypogaea), redroot amaranth (Amaranthus<br />

retroflexus), Italian ryegrass (Lolium multiflorum) and cucumber (Cucumis sativus)<br />

compared with volatiles from healthy goatweed. In contrast, volatiles from infected<br />

plants inhibited the growth of three fungal pathogens (Rhizoctonia solani, Botrytis<br />

cinerea, and Sclerotinia sclerotiorum). For this reason, they postulated that the<br />

allelochemicals stimulated by fungal infection in goatweed are more important in the<br />

defence of the plant against infection, rather than against competition from<br />

neighbouring plants. Nonetheless, this study currently provides the clearest<br />

demonstration of the ability of pathogens to influence the allelopathic ability of plants.<br />

4. IMPLICATIONS FOR PATHOGEN<br />

AGGRESSIVENESS AND EVOLUTION<br />

Plant/pathogen interactions in natural populations are often explained in terms of coevolution,<br />

which is ‘the joint evolution of two (or more) taxa that have close ecological<br />

relationships but do not exchange genes, and in which reciprocal selective pressures<br />

operate to make the evolution of either taxon partially dependent upon the evolution


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

of the other’ (Pianka, 1978). Burdon (1987) explained the co-evolution of plants and<br />

their pathogens in terms of the gene-for-gene concept of resistance and virulence,<br />

using the following model:<br />

(a) A uniform host population possessing a single gene for resistance is challenged<br />

by a uniform pathogen population with the complementary virulence gene,<br />

resulting in pathogenesis.<br />

(b) Under the above conditions, chance mutation favours the appearance of a novel<br />

resistance gene preventing pathogenesis through enhanced resistance. These<br />

resistant individuals have a greater fitness than their susceptible neighbours and,<br />

consequently, increase in frequency within the population.<br />

(c) As the frequency of the resistant genotype increases, selective pressure favours<br />

the appearance of a pathogen race with a novel virulence gene capable of attacking<br />

the resistant host.<br />

(d) As resistance is broken down, the advantage of the previously resistant host<br />

genotype in terms of plant fitness is lost, and its frequency within the population<br />

falls.<br />

(e) Under these conditions the appearance of yet another resistant gene is favoured<br />

and the cycle continues.<br />

Although Burdon’s model provides a good example of the concept of co-evolution,<br />

it is important to note that it does not account for the cost of resistance and virulence<br />

to the host and pathogen (Parker and Gilbert, 2004), host tolerance to disease (Roy<br />

and Kirchner, 2000), or the specificity of a pathogen to its host (Kirchner and Roy,<br />

2002).<br />

The importance of virulence (the degree or measure of pathogenicity) in<br />

determining the ability of a pathogen to infect a particular host is central to the concept<br />

of co-evolution of plants and pathogens. Despite this, its importance has often been<br />

emphasised to the exclusion of another component of pathogen fitness – aggressiveness<br />

(Burdon, 1987). In plant pathology, the term aggressiveness has been defined<br />

inconsistently (Jarosz and Davelos, 1995; Kirchner and Roy, 2002) or considered<br />

synonymous with virulence (Parker and Gilbert, 2004), but here is considered the<br />

negative effect of infection on plant fitness (Jarosz and Davelos, 1995). The role of<br />

aggressiveness in interactions between plants and biotrophic pathogens is represented<br />

by two seemingly opposing views. Harper (1977) suggested that biotrophic pathogens,<br />

which need living tissue for their survival, evolve towards minimising their<br />

aggressiveness, with evolutionary equilibrium occurring when a pathogen attains<br />

commensalistic relationship with its host. On the other hand, less aggressive pathogens<br />

are prone to displacement by more aggressive strains (Jarosz and Davelos, 1995),<br />

implying that pathogens should evolve towards increased aggressiveness. Despite<br />

these arguments portraying the role of aggressiveness differently, they need not be<br />

mutually exclusive.<br />

As an adjunct to the concept of co-evolution, the New Function Hypothesis<br />

proposed by Clay (1988) suggests that pathogens may evolve towards a mutualistic<br />

relationship with their hosts through the appearance of pathogen strains with beneficial<br />

93


94<br />

SCOTT W. MATTNER<br />

effects on the host in addition to their detrimental effects. In this way pathogens<br />

minimise their aggressiveness through the acquisition of ‘new functions’ that increase<br />

host fitness and, at the same time, are less prone to displacement by more aggressive<br />

strains by maintaining their original capacity for disease. A ‘new function’ suggested<br />

in this chapter is the potential for pathogens to increase the allelopathic ability of<br />

their host, which to varying degrees offsets the infected hosts’s loss in competitiveness.<br />

By conferring some benefit on its host, the pathogen maintains a self-advantage through<br />

increasing the survival chances of its host and ultimately itself.<br />

Since biotrophic parasites, such as the rusts, are heavily dependent on the<br />

continuity of their host’s genotype into succeeding generations, the evolution of<br />

interactions that enhance the chances of host survival are important. It is well<br />

established that infections of fodder species by biotrophic pathogens can create<br />

conditions that either limit the grazing of their hosts, limit the number of grazing<br />

animals, or both. Morgan and Parbery (1980) found that infection by Pseudopeziza<br />

medicagnis lowered protein content, digestibility and palatability of lucerne as well<br />

as increasing its oestrogenic activity. Similarly, rust reduces the digestibility and<br />

quality of ryegrass (Isawa et al., 1974; Trorey, 1979; Potter, 1987). For this reason<br />

ruminants preferentially graze healthy ryegrass rather than rusted ryegrass<br />

(Cruickshank, 1957; Heard and Roberts, 1975), indirectly benefiting rusted ryegrass.<br />

Furthermore, evidence presented in this chapter supports the ability of rust to add a<br />

further benefit to ryegrass, that of increased allelopathy with neighbouring plants. In<br />

a similar manner to crown rust, amongst other benefits, the mutualistic endophyte N.<br />

lolli reduces ryegrass palatability to ruminants (Fletcher and Sutherland, 1993) and<br />

increases its allelopathic ability (Sutherland and Hoglund, 1990; Quigley et al., 1990;<br />

Sutherland et al., 1999). The parallels between these two systems suggest that the<br />

pathogenic relationship between crown rust and ryegrass is evolving toward mutualism.<br />

5. CONCLUSIONS<br />

Pathogens are important forces that influence the structure and dynamics of plant<br />

communities. Although there are numerous interpretations and studies on the impact<br />

of pathogens on plant competition, few studies have considered their effect on<br />

allelopathy. Currently, however, most evidence suggests that pathogens may<br />

simultaneously decrease the competitiveness and increase the allelopathic ability of<br />

their hosts. By conferring the benefit of increased allelopathy on their hosts, pathogens<br />

may offset their host’s loss in competitiveness. In so doing, these pathogens make<br />

their hosts less prone to displacement by resistant components of the plant ecosystem<br />

and encourage the continuation of their host’s genotype into proceeding generations,<br />

and ultimately their own. Finally, the similarities in the ability of the ryegrass pathogen,<br />

P. coronata, and the ryegrass mutualist, N. lolii, to increase their host’s allelopathic<br />

capacity suggests that this ‘new function’ may be one way that the rust pathogen is<br />

evolving toward mutualism.


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

7. REFERENCES<br />

Ahmad, I., Owera, S.A.P., Farrar, J.F., Whitbread, R. The distribution of five major nutrients in barley plants<br />

infected with brown rust. Physiol Plant Pathol 1982; 21:335-346.<br />

Alexander, H.M., Holt, R.D. The interaction between plant competition and disease. Perspect Plant Ecol Evol<br />

System 1998; 1:206-220.<br />

Alexander, H.M. Dynamics of plant-pathogen interactions in natural plant communities. In: Pests, Pathogens<br />

and Plant Communities. Burdon J.J. and Leather S.R. eds., Oxford, UK: Blackwell Scientific Publications<br />

1990.<br />

Anaya, A.L. Allelopathy as a tool in the management of biotic resources in agroecosystems. Crit Rev Plant Sci<br />

1999; 18:697-739.<br />

Andreasson, E., Wretblad, S., Graner, G., Wu, X.M., Zhang, J.M, Dixelius, C., Rask, L., Meijer, J. The myrosinaseglucosinolate<br />

system in the interaction between Leptosphaeria maculans and Brassica napus. Mol Plant<br />

Pathol 2001; 21:281-286.<br />

Applebee, T.A., Gibson, D.J., Newman, J.A. Elevated atmospheric carbon dioxide alters the effects of<br />

allelochemicals produced by tall fescue on alfalfa seedlings. Trans Illinois State Acad Sci 1999; 92:23-<br />

31.<br />

Arachevaleta, M., Bacon, C.W., Hoveland, C.S., Radcliffe, D.E. Effect of tall fescue endophyte on plant response<br />

to environmental stress. Agron J 1989; 81:83-90.<br />

Augsberger, C.K. Seedling survival of tropical tree species: interactions of dispersal distance, light-gaps, and<br />

pathogens. Ecology 1984; 65:1705-1712.<br />

Ayres, P.G. Growth responses induced by pathogens and other stresses. In: Responses of Plants to Multiple<br />

Stresses. Mooney, H.A., Winner, W.E., Pell, E.J. eds. San Diego, USA: Academic Press 1991<br />

Ayres, P.G., Paul, N.D. The effects of disease on interspecific plant competition. Aspects of Appl Biol 1990;<br />

24:155-162.<br />

Bailey, J.A. Pisatin production by tissue cultures of Pisum sativum L. J Gen Microbiol 1970; 61:409-415.<br />

Bailey, J.A. Mechanisms of phytoalexin accumulation. In: Phytoalexins. Bailey, J.A., Mansfield, J.W., eds.,<br />

New York, USA: John Wiley and Sons. 1982.<br />

Bansal, G.L. The science of allelopathy: problems and prospects. In: Allelopathy in Agriculture and Forestry.<br />

Narwal S.S., Tauro, P. eds. Jodhpur, India: Scientific Publishers, 1994.<br />

Begon, M., Harper, J.L., Townsend, C.R. Ecology: Individuals, Populations and Communities. (1 st<br />

ed.). Oxford,<br />

UK: Blackwell Scientific Publications 1986.<br />

Bennett, R.N., Wallsgrove, R.M. Tansley Review No. 72: Secondary metabolites in plant defence mechanisms.<br />

New Phytol 1994; 127:617-633.<br />

Boydston, R., Paxton, J.D., Koeppe, D.E. Glyceollin: a site-specific inhibitor of electron transport in isolated<br />

soyabean mitochondria. Plant Physiol 1983; 72:151-155.<br />

Boyette, C.D., Abbas, H.K. Weed control with mycoherbicides and phytotoxins. In: Allelopathy: Organisms,<br />

Processes and Applications. Inderjit, K.M.M. Dakshini, Einhellig, F.A. eds., Washington DC, USA:<br />

American Chemical Society 1995.<br />

Brader, G., Tas, E., Palva, E.T. Jasmonate-dependent induction of indole glucosinolates in Arabidopsis by culture<br />

filtrates of the non-specific pathogen Erwinia caratovora. Plant Physiol 2001; 126:849-860.<br />

Bremner, J.M., McCarty, G.W. Inhibition of nitrification in soil by allelochemicals derived from plants and<br />

plant residues. In: Soil biochemistry. Bollag J. Stotzky, G. eds., Vol. 8. New York, USA: Marcel Dekker.<br />

1993.<br />

Burdon, J.J. The effect of fungal pathogens on plant communities. In: The Plant Community as a Working<br />

Mechanism. Newman E.I. ed., Oxford, UK: Blackwell Scientific Publications, 1982.<br />

Burdon, J.J. Diseases and plant population biology. Cambridge, UK: Cambridge University Press, 1987.<br />

Burdon, J.J. Fungal pathogens as selective forces in plant populations and communities. Aust J Ecol 1991;<br />

16:423-432.<br />

Buta, J.G., Spaulding, D.W., Reed, A.N. Differential growth responses of fractionated turfgrass seed leachates.<br />

Hort Sci 1987; 22:1317-1319.<br />

Calvo, A.M., Wilson, R.A., Bok, J.W., Keller, N.P. Relationship between secondary metabolism and fungal<br />

development. Microbiol Mol Biol Rev 2002; 66:447-459.<br />

95


96<br />

SCOTT W. MATTNER<br />

Casper, B.B., Jackson, R.B. Plant competition underground. Ann Rev Ecol System 1997; 28:545-570.<br />

Catherall, P.L. Effects of barley yellow dwarf virus on the growth and yield of single plants and simulated<br />

swards of perennial rye-grass. Ann Appl Biol 1966; 57:155-162.<br />

Chang, C.F., Suzuki, A., Kumai, S., Tamura, S. Chemical studies on ‘clover sickness.’ II <strong>Biologica</strong>l functions of<br />

isoflavonoids and their related compounds. Agri Biol Chem 1969; 33:398-408.<br />

Chilvers, G.A., Brittain, E.G. Plant competition mediated by host-specific parasites – a simple model. Aus J<br />

Biol Sci 1972; 25:749-756.<br />

Chou, C.H. The role of allelopathy in phytochemical ecology. In: Phytochemical Ecology: <strong>Allelochemicals</strong>,<br />

Mycotoxins and Insect Pheromones and Allomones. Chou C.H. and Waller, G.R. eds. Taipei, Taiwan:<br />

Institute Botanica Academia Senica, 1989.<br />

Chung, I.M., Miller, D.A. Allelopathic influence of nine forage grass extracts on germination and seedling<br />

growth of alfalfa. Agron J 1995; 87:767-772.<br />

Clay, K. Effects of fungal endophytes on the seed and seedling biology of Lolium perenne and Festuca<br />

arundinacea. Oecologia 1987; 73:358-362.<br />

Clay, K. Clavicipitaceous fungal endophytes of grasses: coevolution and the change from parasitism to mutualism.<br />

In: Coevolution of Fungi with Plants and Animals. Pirozynski K.A. and Hawksworth, D.L. eds., London,<br />

UK: Academic Press. 1988.<br />

Clay, K. The impact of parasitic and mutualistic fungi on competitive interactions among plants. In Perspectives<br />

on Plant Competition. Grace J.B., Tilman, D. eds. San Diego, USA: Academic Press. 1990.<br />

Clay, K. Interactions among fungal endophytes, grasses and herbivores. Res Plant Ecol 1996; 38:191-201.<br />

Critchett, C.I. Studies on ryegrass rusts in Victoria. PhD Thesis. Melbourne, Australia: La Trobe University,<br />

1991.<br />

Cruickshank, I.A.M. Crown rust of ryegrass. New Zealand J Sci Technol 1957; 38A:539-543.<br />

Cruickshank, I.A.M., Perrin, D.R. Studies on phytoalexins. III The isolation, assay, and general properties of a<br />

phytoalexin from Pisum sativum L. Aus J Biol Sci 1961; 14:336:348.<br />

Cruickshank, I.A.M., Perrin, D.R. Phytoalexins of the Leguminosae. Phaseollin from Phaseolus vulgaris L.<br />

Life Sci 1963; 2:680-682.<br />

Daayf, F., El-Bellaj, M., El-Hassni, M., J’Aiti, F., El-Hadrami, I. Elicitation of soluble phenolics in date palm<br />

(Phoenix dactylifera) callus by Fusarium oxysporum f.sp. albedinis culture medium. Environ Exp Bot<br />

2003; 49:41-47.<br />

Dakora, F.D., Phillips, D.A. Diverse functions of isoflavonoids in legumes transcend anti-microbial definitions<br />

of phytoalexins. Physiol Mol Plant Pathol 1996; 49:1-20.<br />

Daly, J.M., Deverall, B.T. Toxins and plant pathogenesis. Sydney, Australia: Academic Press. 1983.<br />

Deacon, J.W. Modern Mycology. (3 rd<br />

ed.). Oxford, UK: Blackwell Scientific Publications. 1997.<br />

Dinoor, A., Eshed, N. The role and importance of pathogens in natural plant communities. Ann Rev Phytopathol<br />

1984; 22:443-466.<br />

Di Pietro, A. Fungal antibiosis in biocontrol of plant disease. In: Allelopathy: Organisms, processes, and<br />

applications. Inderjit, K.M.M. Dakshini, Einhellig, F.A. eds. Washington DC, USA: American Chemical<br />

Society, 1995.<br />

Ditommaso, A., Watson, A.K. Impact of a fungal pathogen, Colletotrichum coccodes on growth and competitive<br />

ability of Abutilon theophrasti. New Phytol 1995; 131:51-60.<br />

Dobson, A., Crawley, M. Pathogens and the structure of plant communities. Trends Ecol Evol 1994; 9:393-398.<br />

Donald, C.M. Competition among crop and pasture plants. Adv Agron 1963; 15:1-118.<br />

Einhellig, F.A. Allelopathy: current status and future goals. In: Allelopathy: Organisms, processes, and<br />

applications. Inderjit, K.M.M. Dakshini, Einhellig, F.A. eds. Washington DC, USA: American Chemical<br />

Society 1995.<br />

Friess, N., Maillet, J. Influence of cucumber mosaic virus infection on the competitive ability and reproduction<br />

of chickweed (Stellaria media). New Phytol 1997; 135:667-674.<br />

Fletcher, L.R., Sutherland, B.L. Liveweight changes I lambs grazing perennial ryegrass with different endophyutes.<br />

In: Proceedings of the 2 nd<br />

International Symposium on Acremonium/Grass Interactions 1993.


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

Gates, D.J., Westcott, M., Burdon, J.J., Alexander, H.M. Competition and stability in plant mixtures in the<br />

presence of disease. Oecologia 1986; 68:559-566.<br />

Giannini, J.L., Holt, J.S., Briskin, D.P. The effect of glyceollin on proton leakage in Phytophthora megasperma<br />

f.sp. glycinea plasma membrane and red beet tonoplast vesicles. Plant Sci 1990; 68: 39-45.<br />

Glazener, J.A., VanEtten, H.D. Phytotoxicity of phaseollin to, and alteration of phaseollin by, cell suspension<br />

cultures of Phaseolus vulgaris. Phytopathology 1978; 68:111-117.<br />

Goodman, R.N., Király, Z., Wood, K.R. The Biochemistry and Physiology of Plant Disease. Columbia, USA:<br />

University of Missouri Press 1986.<br />

Grace, J.B., Tilman, D. Perspectives on Plant Competition. San Diego, USA: Academic Press 1990.<br />

Groves, R.H., Williams, J.D. Growth of skeleton weed (Chondrilla juncea L.) as affected by growth of<br />

subterranean clover (Trifolium subterraneum L.) and infection by Puccinia chondrillina Bubak and Syd.<br />

Aust J Agri Res 1975; 36:975-983.<br />

Gussin, E.J, Lynch, J.M. Microbial fermentation of grass residues to organic acids as a factor in the establishment<br />

of new grass swards. New Phytol 1981; 89:449-457.<br />

Halsall, D.M., Leigh, J.H., Gollasch, S.E., Holgate, M. The role of allelopathy in legume decline in pastures. II<br />

Comparative effects of pasture, crop and weed residues on germination, nodulation and root growth. Aus<br />

J Agri Res 1995; 46:189-207.<br />

Hale, M.G., Orcutt, D.M. Physiology of Plants Under Stress. New York, USA: Wiley Interscience 1987.<br />

Hampton, J.G. Fungicidal effects on stem rust, green leaf area, and seed yield in ‘Grasslands Nui’ perennial<br />

ryegrass. New Zealand J Exp Agri 1986; 14:7-12.<br />

Harper, J.L. Population Biology of Plants. London, UK: Academic Press. 1977.<br />

Heard, A.J., Roberts, E.T. Disorders of temporary ryegrass swards in south-east England. Ann Appl Biol 1975;<br />

81:235-279.<br />

Hodgkin, T., Lyon, G.D. Inhibition of Solanum pollen germination in vitro by the phytoalexin rishitin. Ann Bot<br />

1979; 44:253-255.<br />

Inderjit. Plant phenolics in allelopathy. Bot Rev 1996; 62:186-202.<br />

Inderjit, Dakshini, K.M.M., Einhellig, F.A. Allelopathy: Organisms, Processes, and Applications. Washington<br />

DC, USA: American Chemical Society 1995.<br />

Inderjit, Weiner, J. Plant allelochemical interference or soil chemical ecology? Persp Plant Ecol Evol System<br />

2001; 4:3-12.<br />

Inderjit, Weston, L.A. Are laboratory bioassays for allelopathy suitable for prediction of field responses?<br />

J Chem Ecol 2000; 26:2111-2118.<br />

Ingham, J.L. Phytoalexins from the Leguminosae. In: Phytoalexins. Bailey, J.A., Mansfield, J.W. eds., New<br />

York, USA: John Wiley and Sons. 1982.<br />

Isawa, K., Abe, A., Nishihara, N. Influence of diseases on the chemical composition and nutritive value of<br />

forage crops. I. Chemical composition of Italian ryegrass infected with crown rust. Ann Phytopathol Soc<br />

Japan 1974; 40:86-92.<br />

Ishiguri, Y., Tomiyama, A.K., Murai, A., Katsui, N., Masamune, N. Toxicity of rishitin, rishitin-M-1, and rishitin-<br />

M-2 to Phytophthora infestans and potato tissue. Ann Phytopathol Soc Japan 1978; 44:52-56.<br />

Jay, C.N., Rossall, S., Smith, H.G. Effects of beet western yellows virus on growth and yield of oilseed rape<br />

(Brassica napus). J Agri Sci 1999; 133:131-139.<br />

Jarosz, A.M., Davelos, A.L. Effects of disease in wild plant populations and the evolution of pathogens<br />

aggressiveness. New Phytol 1995; 129:371-387.<br />

Joliffe, P.A. Evaluating the effects of competitive interference on plant performance. J Theoret Biol 1988; 130:447-<br />

459.<br />

Kempenaar, C., Horsten, P.J.F.M., Scheepens, P.C. Growth and competitiveness of common lambsquarters<br />

(Chenopodium album) after foliar application of Ascochyta caulina as a mycoherbicide. Weed Sci 1996;<br />

44:609-614.<br />

Kirchner , J.W., Roy, B.A. Evolutionary implications of host-pathogen specificity: fitness consequences of pathogen<br />

virulence traits. Evolut Ecol Res 2002; 4:27-48.<br />

Kirkegaard, J.A., Sarwar, M., Wong, P.T.W., Mead, A., Howe, G., Newell, M. Field studies on the biofumigation<br />

of take-all by Brassica break crops. Aus J Agri Res 2000; 51:445-456.<br />

Kong, C., Hu, F., Xu, X. Allelopathic potential and chemical constituents of volatiles from Ageratum conyzoides<br />

under stress. J Chem Ecol 2002; 28:1173-1182.<br />

97


98<br />

SCOTT W. MATTNER<br />

Kuc, J. Molecular aspects of plant responses to pathogens. Acta Physiologiae Plant 1997; 19:551-559.<br />

Kuc, J., Henze, R.E., Ullstrup, A.J., Quackenbush, F.W. Chlorogenic and caffeic acids as fungistatic agents<br />

produced by potatoes in response to inoculation with Helminthosporium carbonum. J Amer Chem Soc1956;<br />

78:3123.<br />

Kurosaki, F., Matsui, K., Nishi, A. Production and metabolism of 6-methoxymellein in cultured carrot cells.<br />

Physiol Plant Pathol 1984; 25:313-321.<br />

Lancashire, J.A., Latch, G.C.M. Some effects of crown rust (Puccinia coronata Corda) on the growth of two<br />

ryegrass varieties in New Zealand. New Zealand J Agri Res 1966; 9:628-640.<br />

Lancashire, J.A., Latch, G.C.M. The effect of crown rust (Puccinia coronata Corda) on the yield and botanical<br />

composition of two ryegrass/white clover pastures. New Zealand J Agri Res 1970; 13:279-286.<br />

Latch, G.C.M. Physiological interactions of endophytic fungi and their hosts. Biotic stress tolerance imparted to<br />

grasses by endophytes. Agri Eco Sys Environ 1993; 44:143-156.<br />

Latch, G.C.M., Hunt, W.F., Musgrave, D.R. Endophytic fungi affect growth of perennial ryegrass. New Zealand<br />

J Agri Res 1985; 28:165-168.<br />

Legaz, M.E., de Armas, R., Piñón, D., Vicente, C. Relationships between phenolics, conjugated polyamines and<br />

sensitivity of sugar cane to smut (Ustilago scitaminea). J Exp Bot 1998; 49:1723-1728.<br />

Lewis, G.C., Day, R.C. Growth of a perennial ryegrass genotype with and without infection by ryegrass endophyte<br />

and virus diseases. Ann Appl Biol 1993; 122:144-145.<br />

Li, Y.C., Kiddle, G., Bennet, R.N., Wallsgrove, R.M. Local and systemic changes in glucosinolates in Chinese<br />

and European cultivars of oilseed rape (Brassica napus L.) after inoculation with Sclerotinia sclerotiorum<br />

(stem rot). Ann Appl Biol 1999; 134:45-58.<br />

Liu, D.L., Lovett, J.V. <strong>Biologica</strong>lly active secondary metabolites of barley. I. Developing techniques and assessing<br />

allelopathy in barley. J Chem Ecol 1993; 19:2217-2230.<br />

Lutrell, E.S. Parasitism of fungi on vascular plants. Mycologia 1974; 66:1-15.<br />

Lyon, G.D. Evidence that the toxic effect of rishitin may be due to membrane damage. J Exper Bot 1980;<br />

31:957-966.<br />

Lyon, G.D., Mayo, M.A. The phytoalexin rishitin affects the viability of isolated plant protoplasts. Phytopathol<br />

Z 1978; 92:294-304.<br />

Mandava, N.V. Chemistry and biology of allelopathic agents. In: The Chemistry of Allelopathy. A.C. Thompson<br />

ed., Washington DC, USA: American Chemical Society, 1985.<br />

Mandavia, M.K., Khan, N.A., Gajera, H.P., Andharia, J.H. Parameswaran, M. Inhibitory effects of phenolic<br />

compounds on fungal metabolism in host-pathogen interactions in Fusarium wilt of cumin. Allelopathy J<br />

2000; 7:85-92.<br />

Mattner, S.W. The impact of crown rust on interference between perennial ryegrass and white clover. PhD<br />

Thesis. Melbourne, Australia: The University of Melbourne 1998.<br />

Mattner, S.W., Parbery, D.G. Rust-enhanced allelopathy of perennial ryegrass against white clover. Agron J<br />

2001; 93:54-59.<br />

Mayama, S., Matsuura, Y., Iida, H., Tani, T. The role of avenalumin in the resistance of oat to crown rust,<br />

Puccinia coronata f.sp. avenae. Physiol Plant Pathol 1982; 20:189-199.<br />

Mayama, S., Tani, T., Matsuura, Y., Ueno, T., Fukami, H. The production of phytoalexins by oat in response to<br />

crown rust, Puccinia coronata f.sp. avenae. Physiol Plant Pathol 1981; 19:217-226.<br />

Milthorpe, F.L. Mechanisms in <strong>Biologica</strong>l Competition. Cambridge, UK: Cambridge University Press, 1961.<br />

Mol, L. Effect of plant roots on the germination of microsclerotia of Verticillium dahliae. II. Quantitative analysis<br />

of the luring effect of crops. Europ J Plant Pathol 1995; 101:679-685.<br />

Mothes, K. Physiology of alkaloids. Ann Rev Plant Physiol 1955; 6:393-432.<br />

Morgan, W.C., Parbery, D.G. Depressed fodder quality and increased oestrogenic activity of lucerne infected<br />

with Pseudopeziza medicagnis. Aust J Agri Res 1980; 31:1103-1110.<br />

Muller, C.H. Allelopathy as a factor in ecological process. Vegetatio 1969; 18:348-351.<br />

Müller, K.O., Börger, H. Experimentelle Untersuchungen über die Phytophthora-resistenz den Kartoffel. Arbeiten<br />

aus der Biologischen Bundensanstalt für Land-und Fortswirtschaft 1941; 23:189-231.<br />

Naqvi, H.H. Preliminary studies of interference exhibited by Italian ryegrass. Biologia (Lahore) 1972; 18: 201-<br />

210.


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

Naqvi, H.H., Muller, C.H. Biochemical inhibition (allelopathy) exhibited by Italian ryegrass (Lolium multiflorum<br />

L.). Pak J Bot 1975; 7:139-147.<br />

Newman, E.I., Miller, M.H. Allelopathy among some British grasslands species. II Influence of root exudates<br />

on phosphorus uptake. J Ecol 1977; 65:399-411.<br />

Newman, E.I., Rovira, A.D. Allelopathy among some British grasslands species. J Ecol 1975; 63:727-737.<br />

Nicol, R.W., Yousef, L., Traquair J.A., Bernards, M.A. Ginsenosides stimulate the growth of soilborne pathogens<br />

of American ginseng. Phytochemistry 2003; 64:257-264.<br />

Odunfa, V.S.A. Root exudation in cowpea and sorghum and the effect on spore germination and growth of some<br />

soil fusaria. New Phytol 1978; 80:607-612.<br />

Parbery, D.G., Gardner, W.K., Golebiowski, T. Stimulation of isoflavonoid content in subterranean clover by<br />

infection with a fungus. J Aus Inst Agri Sci 1984; 50:114-116.<br />

Parbery, D.G. Trophism and the ecology of fungi associated with plants. Biol Rev 1996; 71: 473-527.<br />

Parker, I.M., Gilbert, G.S. The evolutionary ecology of novel plant-pathogen interactions. Ann Rev Ecol Evol<br />

Syst 2004; 35:675-700.<br />

Paul, N.D. Modification of the effects of plant pathogens by other components of natural ecosystems. In: Pests,<br />

Pathogens and Plant Communities. Burdon J.J. and Leather S.R. eds., Oxford, UK: Blackwell Scientific<br />

Publications 1990.<br />

Paul, N.D., Ayres, P.G. Interference between healthy and rusted groundsel (Senecio vulgaris L.) within mixed<br />

populations of different densities and proportions. New Phytol 1986; 104:257-269.<br />

Paul, N.D., Ayres, P.G. Effects of rust infection of Senecio vulgaris on competition with lettuce. Weed Res<br />

1987a ; 27:431-441.<br />

Paul, N.D., Ayres, P.G. Water stress modifies intraspecific interference between rust (Puccinia lagenophorae<br />

Cooke)–infected and healthy groundsel (Senecio vulgaris L.). New Phytol 1987b ; 106:555-566.<br />

Paul, N.D., Ayres, P.G. Nutrient relations of groundsel (Senecio vulgaris L.) infected by rust (Puccinia<br />

lagenophorae) at a range of nutrient concentrations. I. Concentrations, content, and distribution of N, P<br />

and K. Ann Bot 1988; 61:489-498.<br />

Paul, N.D., Ayres, P.G. Effects of interactions between nutrient supply and rust infection of Senecio vulgaris L.<br />

on competition with Capsella bursa-pastoris (L.) Medic. New Phytol 1990; 114:667-674.<br />

Paxton, J.D. Phytoalexins – a working redefinition. Phytopathol Z 1981; 101:106-109.<br />

Peters, J.C., Shaw, M.W. Effect of artificial exclusion and augmentation of fungal plant pathogens on a regenerating<br />

grassland. New Phytol 1996; 134:295-307.<br />

Perrin, D.R., Bottomley, W. Studies on phytoalexins. V The structure of pisatin from Pisum sativum L. J Amer<br />

Chem Soc 1962; 84:1919-1922.<br />

Phan, C.T. Phenolics and polyketides in carrots. In: Mycotoxins and Phytoalexins. Sharma R.P., Salunkhe<br />

D.K., eds., CRC Press, Boca Raton, USA 1991.<br />

Pianka, E.R. Evolutionary Ecology. New York, USA: Harper and Row. 1978.<br />

Plummer, R.M., Hall, R.L., Watt, T.A. The influence of crown rust (Puccinia coronata) on tiller production<br />

and survival on perennial ryegrass (Lolium perenne) plants in simulated swards. Grass Forage Sci 1990;<br />

45:9-16.<br />

Potter, L.R., Cagas, B., Paul, V.H., Birckenstaedt, E. Pathogenicity of some European collections of crown rust<br />

(Puccinia coronata Corda) on cultivars of perennial ryegrass. J Phytopathol 1990; 130:119-126.<br />

Pratley, J.E. Allelopathy in annual grasses. Plant Protection Quart 1996; 11:213-214.<br />

Prestidge, R.A., Thom, E.R., Marshall, S.L., Taylor, M.J., Willoughby, B., Wildermoth, D.D. Influence of<br />

Acremonium lolii infection in perennial ryegrass on germination, emergence, survival and growth of<br />

white clover. New Zealand J Agri Res 1992; 35:225-234.<br />

Putnam, A.R. Allelopathic research in agriculture, past highlights and potential. In: The Chemistry of Allelopathy.<br />

Thompson, A.C. ed., Washington DC, USA: American Chemical Society. 1985.<br />

Putnam, A.R., and Duke, W.B. Allelopathy in agroecosystems. Ann Rev Phytopathol 1978; 16:431-451.<br />

Putnam, A.R., Tang, C. Allelopathy: state of the science. In: The Science of Allelopathy. Putnam A.R. Tang, C.<br />

eds., New York, USA: John Wiley and Sons. 1986.<br />

Quigley, P.E., Snell, F.J., Cunningham, P.J., Frost, W. The effects of endophyte infected ryegrass on the<br />

establishment, persistence and production of mixed pastures in Australia. In: Proceedings of the International<br />

Symposium on Acremonium/Grass Interactions: 1990; 49-51.<br />

99


100<br />

SCOTT W. MATTNER<br />

Quigley, P.E. Effects of Neotyphodium lolii infection and sowing rate of perennial ryegrass (Lolium perenne)<br />

on the dynamics of ryegrass/subterranean clover (Trifolium subterraneum) swards. Aust J Agri Res<br />

2000; 51:47-56.<br />

Ravel, C, Courty, C., Coudret, A., Charmet, C. Beneficial effects of Neotyphodium lolii on the growth and the<br />

water status in perennial ryegrass cultivated under nitrogen deficiency or drought stress. Agronomie 1997;<br />

17:173-181.<br />

Rice, E.L. Allelopathy. New York, USA: Academic Press, 1974.<br />

Rice, E.L. Allelopathy. (2 nd<br />

ed.). Orlando, USA: Academic Press, 1984.<br />

Rice, E.L. Allelopathic effects on nitrogen cycling. In: Allelopathy: Basic and applied aspects. Rizvi S.J.H.,<br />

Rizvi, V. eds., London, UK: Chapman and Hall, 1992.<br />

Rizvi, S.J.H., Haque, H., Singh, V.K., Rizvi, V. A discipline called Allelopathy. In: Allelopathy: Basic and<br />

Applied Aspects. Rizvi S.J.H., Rizvi, V. eds., London, UK: Chapman and Hall 1992.<br />

Roy, B.A., Kirchner, J.W. Evolutionary dynamics of pathogen resistance and tolerance. Evolution 2002; 54:51-<br />

63.<br />

Seigler, D.S. Chemistry and mechanisms of allelopathic interactions. Agron J 1996; 88:876-885.<br />

Skipp, R.A., Selby, C., Bailey, J.A. Toxic effects of phaseollin on plant cells. Physiol Plant Pathol 1977; 10:221-<br />

227.<br />

Shinozaki, K., Kira, T. Intraspecific competition among higher plants. VII Logistic theory of the C-D effect.<br />

Journal of the Institute of Polytechnics, Osaka City University 1956; 7:35-72.<br />

Singh, U.P., Sarma, B.K., Singh, D.P., Bahadur, A. Studies on exudates-deplete sclerotial development in<br />

Sclerotium rolfsii and the effect of oxalic acid, sclerotial exudates, and culture filtrate on phenolic acid<br />

induction in chickpea (Cicer arietinum). Can J Microbiol 2002; 48:443-448.<br />

Siranidou, E., Kang, Z., Buchenauer, H. Studies on symptom development, phenolic compounds and<br />

morphological defence responses in wheat cultivars differing in resistance to Fusarium head blight. J<br />

Phytopathol 2002; 150:200-208.<br />

Smith, C.J. Accumulation of phytoalexins: defence mechanism and stimulus response system. New Phytol 1996;<br />

132:1-45.<br />

Spessard, G.O., Hanson, C., Halvorson, J.S., Giannini, J.L. Effects of phaseollin on membrane leakage in red<br />

beet vacuoles and tonoplast vesicles. Phytochemistry 1994; 35:43-47.<br />

Spring, O., Benz, A., Faust, V. Impact of downy mildew (Plasmopara halstedii) infection on the development<br />

and metabolism of sunflower. Zeitschrift für Planzenkrankheiten und Planzenschutz 1991; 98:597-604.<br />

Stevens, D.R., Hickey, M.J. Effects of endophytic ryegrass on white clover. In: Proceedings of the International<br />

Symposium on Acremonium/Grass Interactions, 1990.<br />

Stewart, T.M., Mercer, C.F., Grant, J.L. Development of Meloidogyne naasi on endophyte-infected and endophytefree<br />

perennial ryegrass. Aus Plant Pathol 1993; 22:40-41.<br />

Stoessl, A. Biosynthesis of phytoalexins. In: Phytoalexins. Bailey, J.A., Mansfield J.W., eds., New York, USA:<br />

John Wiley and Sons. 1982.<br />

Stoessl, A. Secondary plant metabolites in preinfectional and postinfectional resistance. In: The Dynamics of<br />

Host Defence. Bailey, J.A., Deverall B.J. eds., Sydney, Australia: Academic Press 1983.<br />

Stowe, L.G. Allelopathy and its influence on the distribution of plants in an Illinois old-field. J Ecol 1979;<br />

67:1065-1085.<br />

Sutherland, B.L., Hoglund, J.H. Effect of ryegrass containing the endophyte (Acremonium lolii), on the<br />

performance of associated white clover and subsequent crops. Proc New Zealand Grassland Assoc 1989;<br />

50:265-269.<br />

Sutherland, B.L., Hoglund, J.H. Effect of ryegrass containing the endophyte Acremonium lolii on associated<br />

white clover. In: Proceedings of the International Symposium on Acremonium/Grass Interactions 1990.<br />

Sutherland, B.L., Hume, D.E., Tapper, B.A. Allelopathic effects of endophyte-infected perennial ryegrass extracts<br />

on white clover seedlings. New Zealand J Agri Res 1999; 42:19-26.<br />

Sutic, D.D., Sinclair, J.B. Anatomy and Physiology of Diseased Plants. Boca Raton, USA: CRC Press 1991.<br />

Takahashi, Y., Otani, T., Hagino, K. Studies on interactions among some grassland species. V Collection and<br />

isolation of allelopathic hydrophobic compounds from the root exudates of Lolium perenne L. J Jap Soc<br />

Grassland Sci 1993; 39:236-245.<br />

Takahashi, Y., Otani, T., Uozumi, S., Yoden, Y., Igarashi, R. Studies on the allelopathic interactions among<br />

some grassland species. I Effects of root exudates from some grass and legume species on the growth of<br />

their own species and other species. J Jap Soc Grassland Sci 1988; 33:334-338.


IMPACT OF PATHOGENS ON PLANT INTERFERENCE AND<br />

ALLELOPATHY<br />

Takahashi, Y., Otani, T., Uozumi, S., Yoden, Y., Igarashi, R. Studies on the allelopathic interactions among<br />

some grassland species. II Assessment of the allelopathic interactions between perennial ryegrass (Lolium<br />

perenne L.) and white clover (Trifolium repens L.) using a root exudate recirculating system. J Jap Soc<br />

Grassland Sci 1991; 37:274-282.<br />

Tang, C.S., Cai, W.F., Kohl, K., Nishimoto, R.K. Plant stress and allelopathy. In: Allelopathy: Organisms,<br />

processes, and applications. Inderjit, K.M.M. Dakshini, Einhellig, F.A. eds., Washington DC, USA:<br />

American Chemical Society 1995.<br />

Tamura, S., Chang, C., Suzuki, A., Kymai, S. Isolation and structure of a novel isoflavone derivative in red<br />

clover. Agri Biol Chem 1967; 31:1108-1109.<br />

Tamura, S., Chang, C., Suzuki, A., Kymai, S. Chemical studies on ‘clover sickness.’ I Isolation and structural<br />

elucidation of two new isoflavonoids in red clover. Agri Biol Chem 1969; 33:391-397.<br />

Thijs, H., Shann, J.R., Weidenhamer, J.D. The effect of phytotoxins on competitive outcome in a model system.<br />

Ecology 1994; 75:1959-1964.<br />

Tomiyama, K., Sakuma, T., Ishizaka, N., Sato, N., Katsui, N., Takasugi, M., Masamune, T. A new antifungal<br />

substance isolated from resistant potato tuber tissues infected by pathogens. Phytopathology 1968; 58:115-<br />

116.<br />

Toussoun, T.A., Patrick, Z.A. Effect of phytotoxic substances from decomposing plant residues on root rot of<br />

bean. Phytopathology 1963; 53:265-270.<br />

Tierens, K.F.M.J., Thomma, B.P.H.J., Brouwer, M., Schmidt, J., Kistner, K., Porzel, A., Mauch-Mani, B.,<br />

Cammue, B.P.A., Broekaert, W.F. Study of the role of antimicrobial glucosinolate-derived isothiocyantes<br />

in resistance of Arabidopsis to microbial pathogens. Plant Physiol 2001; 4:1688-1699.<br />

Tilman, D. Plant strategies and the dynamics and structure of plant communities. Princeton, USA: Princeton<br />

University Press. 1988.<br />

Trenbath, B.R. Biomass productivity of mixtures. Adv Agron 1974; 26:177-210.<br />

Trorey, G.M. The effect of crown rust (Puccinia coronata lolii) on the yield and competitive ability of perennial<br />

ryegrass. PhD Thesis. Huddersfield, United Kingdon: Huddersfield Polytechnic. 1979.<br />

Van der Putten, W.H., Van Dijk, C., Peters, B.A.M. Plant-specific soil-borne disease contribute to succession in<br />

fordune vegetation. Nature 1993; 362:53-56.<br />

Wardle, D.A., Nicholson, K.S., Ahmed, M. Allelopathic influence of nodding thistle (Carduus nutans L.) seeds<br />

on germination and radicle growth of pasture plants. New Zealand J Agri Res 1991; 34:185-191.<br />

Wardle, D.A., Nilsson, M.C., Gallet, C., Zackrisson, O. An ecosystem–level perspective of alleloapthy. Biol<br />

Rev Cambridge Philosoph Soc 1998; 73:305-319.<br />

Watson, R.N., Prestidge, R.A., Ball, O.J.P. Suppression of white clover by ryegrass infected with Acremonium<br />

endophyte. In: Proceedings of the 2 nd International Symposium on Acremonium/Grass Interactions, 1993.<br />

Wennström, A., Ericson, L. Variation in disease incidence in grazed and ungrazed sites for the system Pulsatilla<br />

pratensis – Puccinia pulsatillae. Oikos 1991; 60:35-39.<br />

Whittaker, R.H., Feeny, P.P. Allelochemics: chemical interactions between species. Science 1971; 171:757-<br />

770.<br />

de Wit, C.T. On competition. Verslagen van Landbouwkundige Onderzoekingen 1960; 66:1-82.<br />

Woodhead, S. Environmental and biotic factors affecting the phenolic content of different cultivars of Sorghum<br />

bicolor. J Chem Ecol 1981; 7:1035-1047.<br />

101


RAMANATHAN KATHIRESAN 1 , CLIFFORD H.KOGER 2<br />

AND KRISHNA N. REDDY 3<br />

ALLELOPATHY FOR WEED CONTROL IN<br />

AQUATIC AND WETLAND SYSTEMS<br />

1 Professor, Department of Agronomy, Annamalai University,<br />

Annamalainagar, Tamilnadu 608 002, India. 2 Weed Ecologist, Southern<br />

Weed Science Research Unit, USDA-ARS, P. O. Box 350, Stoneville,<br />

Mississippi 38776, USA. 3 Corresponding author. Plant Physiologist,<br />

Southern Weed Science Research Unit, USDA-ARS, P. O. Box 350,<br />

Stoneville, Mississippi 38776, USA<br />

E-mail : kreddy@ars.usda.gov<br />

Abstract. <strong>Allelochemicals</strong> offer ample scope for ecologically safe and effective weed control in aquatic and<br />

wetland systems. This could be attributed to the absence of soil interface in aquatic habitats that contributes<br />

largely for rapid degradation of allelochemicals. Simpler strategies involving allelopathy especially for small<br />

holder farms, low input agriculture and aquatic environments with appreciable results have been reported. Such<br />

strategies include use of allelopathic cultivars, organic manures and plant products. Though allelopathic<br />

suppression of weeds could not be construed as an alternate to replace synthetic herbicides, it can fit in an<br />

integrated weed management program very well as a prime component. Such strategies are reviewed. Further, a<br />

specific case study for the use of plant product along with insect agents for controlling water hyacinth in India<br />

and different steps involved in selecting allelopathic plant products for aquatic weed control are discussed.<br />

1. INTRODUCTION<br />

Many of the compounds produced by green plants that are not involved in primary<br />

plant metabolism are observed to function as chemical warfare agents against<br />

competing plants and pests. Many such natural compounds have the potential to be<br />

exploited as herbicides or as leads for discovery of new herbicides (Duke, 1986;<br />

Hoagland, 2001). Allelopathy simply refers to chemical warfare between different<br />

plants, in which the bio-chemicals from one plant impair germination, growth, survival,<br />

reproduction, and behavior of other plants. These allelopathic chemicals are produced<br />

by a ‘donor’ and transmitted to a ‘receiver’ that can either be ‘injured’ or ‘stimulated’.<br />

<strong>Allelochemicals</strong> act through direct interference with physiological functions of<br />

‘receiver’ such as seed germination, root growth, shoot growth, stem growth, symbiotic<br />

effectiveness or act indirectly through additive or synergistic impact along with<br />

pathological infections, insect injury and/or environmental stress. Though many of<br />

these allelochemicals exhibit inhibitory response on various morpho-physiological<br />

functions of receiver plants and such responses being observed to be dose dependant<br />

in a linear fashion, their concentrations required for control of weeds on a field scale<br />

are impracticably higher. Further, the degree of selectivity is often a factor limiting<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 103– 122.<br />

© 2006 Springer. Printed in the Netherlands.<br />

103


104<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

for their widespread commercial use. The soil-interface in agricultural field conditions<br />

also affects the effectiveness of these allelochemicals. However, allelochemicals provide<br />

ample opportunity for safe, selective and eco-friendly weed control in aquatic and<br />

wetland systems as resistance offered by soil-interface is largely circumvented by the<br />

water-interface. Under aquatic environment, the transport of allelochemical to the<br />

receiver plant is much more rapid, thus reducing the need for higher doses of<br />

allelochemicals for effective weed suppression.<br />

2. ALLELOPATHIC PLANT MATERIALS<br />

In spite of the widespread apprehension that applications of allelopathy in agriculture<br />

need to undergo an extensive refinement and techno-commercial perfections, simpler<br />

means especially for small holder farms with appreciable results have been reported.<br />

Strategies that involve the mechanism of allelopathy to an appreciable magnitude<br />

include use of allelopathic cultivars, mulches, organic manures, and plant products in<br />

wetlands. In aquatic systems, strategies include use of allelopathic plant products and<br />

plant species capable of aggressively replacing invasive undesirable weed species.<br />

2.1. Allelopathic Crop Cultivars in Wetlands<br />

Apparent allelopathic activity in rice accessions against the weed duck salad<br />

[Heteranthera limosa (Sw.) Willd] was first observed at USDA-ARS in Arkansas,<br />

during 1985-86. Within the next five years, 412 rice accessions that imparted<br />

allelopathic suppression of duck salad with in a radius of 10 cm were identified by<br />

USDA-ARS from their germplasm collection comprising 16,000 rice accessions from<br />

99 countries. The allelopathic suppression of the weed red stem (Ammania coccinea<br />

Rottb) over an area of 10 cm radius from the rice plant was observed with 145 accessions<br />

(Dilday et al., 1994). A hybrid between P1 338046 (allelopathic) and Katy (nonallelopathic)<br />

was shown to possess superior agronomic traits in green house studies<br />

and quantitative inheritance was indicated for the allelopathic activity in hybrid rice<br />

(Dilday et al., 1998). In Egypt, more than 30 rice varieties were shown to be allelopathic<br />

on barnyardgrass [Echinochloa crusgalli (L.) Beau.] and 10 rice varieties were observed<br />

to suppress smallflower umbrella sedge (Cyperus difformis L.). These allelopathic<br />

rice varieties inhibited the root development and emergence of first or second leaf of<br />

both weeds (Hassan et al., 1995). Chemical composition of ethyl acetate extracts from<br />

different rice cultivars allelopathic to barnyardgrass in South Korea consisted mainly<br />

fatty acid esters, unsaturated ketones, polycyclic aromatic compounds, and alkaloids<br />

(Kim and Shin, 1998). Gas chromatography/mass spectrometry (GC/MS) analysis of<br />

extracts of water from wetland with allelopathic rice plants showed significant levels<br />

of 3-hydroxybenzoic acid, 4-hydroxybenzoic acid, 4-hydroxy cinnamic acid and 3, 4dihydroxy<br />

hydrocinnamic acid (Mattice et al., 1998). Many studies have indicated<br />

that the active compounds involved in rice allelopathy are phenolic compounds (Rice,<br />

1987; Chou et al., 1991; Inderjit, 1996; Mattice et al., 1998). The GC/MS analysis of<br />

Kouketsumochi, a potential allelopathic rice variety identified several compounds


ALLELOPATHY FOR WEED CONTROL 105<br />

such as sterols, benzaldehydes, benzene derivatives, long chain fatty acids, esters,<br />

aldehydes, ketones, and amines as to have been biologically active (Kim et al., 2000).<br />

Kim and Shin (2003) observed that more than one allelochemical is likely to be involved<br />

in rice allelopathy and that rice allelopathy is not due to one specific phytotoxin. Such<br />

leads in rice allelopathy research offer scope for incorporation of this self defense<br />

mechanism in improved rice varieties, through breeding programs. A very important<br />

element needed for achieving this goal is understanding of the structure of the genetic<br />

background of the trait.<br />

However, today’s knowledge on allelopathy genetics is limited and this restricts<br />

the efforts for intentional breeding to genetically improve allelopathic potential of<br />

crops. A valuable milestone in this direction has been reached with the contributions<br />

of Jensen et al. (2001). Quantitative Trait Loci (QTL) mapping through 142 DNA<br />

markers were located in 142 recombinant inbred lines derived from a cross between a<br />

japonica upland strain with strong allelopathic traits and an indica irrigated strain<br />

with weak allelopathic traits. Three main loci, independently contributing for 10% of<br />

the promotion of allelochemical synthesis were localized to rice chromosomes 2 and<br />

3. The two QTL traits on chromosome 3 were closely linked, offering scope for easy<br />

manipulation. Three different approaches have been attempted recently to produce<br />

allelopathic crops. Traditional breeding methods are suggested to be a reasonable<br />

alternative by Courtois and Olofsdotter (1998). The first approach involves crossing<br />

of two parents with contrasting behavior and derivation of recombinant inbred lines<br />

(RILs) through single – seed descent (SSD). The second approach involves introduction<br />

of allelopathic traits in to an elite restorer line in developing three – line hybrid rice.<br />

Simultaneous back crossing and selfing methods of breeding were attempted to develop<br />

hybrid rice with allelopathic activity and its counterpart isogenic hybrid rice with a<br />

non-allelopathic effect on weeds. Three lines of rice, Kouketsumochi, Rexmont and<br />

IR-24 were used as allelopathic donor, non-allelopathic, and restoring genes,<br />

respectively (Lin et al., 2000). Analysis and monitoring of allelopotential and heterotic<br />

performance in laboratory and green house indicated a positive and significant<br />

allelopathy in this hybrid rice. A third approach suggested is molecular that could be<br />

achieved by the regulation of gene expression related to allelochemical bio-synthesis<br />

and insertion of allelochemical regulating genes into non-allelopathic crops to induce<br />

the synthesis of allelochemicals (Duke et al., 2001).<br />

2.2. Crop Residues and Organic Manures for Allelopathic Suppression of Weeds<br />

The residues of crops grown during preceding seasons or tree components of the farm<br />

are incorporated in the field before raising field crops to serve as manures. However,<br />

such crop residues also offer other environmental benefits that include protection<br />

from soil erosion, increase in biological diversity including beneficial organisms and<br />

suppression of pests and weeds (Sustainable Agriculture Network, 1998). The residues<br />

of such crops can suppress weeds by releasing allelochemicals (Teasdale, 2003). Living<br />

mulches, intercrops or smother crops may provide physical weed suppression but<br />

their effects in part depend on allelopathy (Bond, 2002). Cover crops suppress weeds


106<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

during early crop season, but they do not provide full-season weed suppression (Reddy,<br />

2001; Reddy et al., 2003; Teasdale, 1996). Selective activity of tree allelochemicals<br />

from Leucaena leucocephala (Lam.) Dewit and Eucalyptus species on different crops<br />

and weed species have been reported by Ferguson and Rathinasabapathi (2003). Field<br />

studies at the Experimental Farm, Annamalai University, India assessed the impact<br />

of different crop residues applied as green leaf manure to both rice nurseries and<br />

wetland (transplanted) rice fields (Parthiban and Kathiresan, 2002; Kathiresan, 2004).<br />

The green leaves of Eucalyptus globulus L., L. leucocephala and Calotropis gigantea<br />

L. were applied at 5 t ha -1 . Green leaves were soil incorporated at the time of final<br />

land preparation, flooded with water, and allowed to decompose. After 7 days, the<br />

land was leveled and used as nursery to raise rice seedlings. Similarly, in wetland<br />

rice fields, green leaves were incorporated and rice seedlings were transplanted into<br />

puddled soil. The results showed that Eucalyptus and Leucaena leaves significantly<br />

reduced the density of Echinochloa spp. and Cyperus rotundus L. in the nursery. In<br />

wetland rice fields, the densities of C. rotundus, Cyperus littoralis Gaud., C. difformis<br />

and Sphenoclea zeylanica Gaertn. were drastically reduced by these two crop residues.<br />

The weed suppression was even better than the rice herbicide, butachlor. However,<br />

use of these crop residues in the nursery reduced rice populations. Laboratory studies<br />

supported this observation and clearly showed a direct inhibitory response of higher<br />

magnitude on rice seed and Echinochloa seed by the leaf leachates whose<br />

concentrations corroborate with field doses. The crop management strategy of applying<br />

green leaf manures of Eucalyptus and Leucaena could compliment weed suppression<br />

(comparable to that of a single application of pre-emergence herbicide) in wetland<br />

transplanted rice. The risk of inhibition of rice seed germination restricts the application<br />

of this technique to rice nursery. However, it is possible to include these green leaves<br />

as component of an integrated weed management strategy in wetland rice production<br />

as rice seedlings are less sensitive to these crop residues. An array of compounds such<br />

as cineol, pinene, caffeic acid, gallic acid, eucalyptin, hyperoside, and rutin, which<br />

constitute primary allelochemicals in Eucalyptus, might have suppressed the<br />

germination of weed seed as well as rice in the nursery as observed in other studies<br />

(Rao et al., 1994; Sivagurunathan et al., 1997). Aqueous extracts of leaves, seeds and<br />

litter of Leucaena were shown to be significantly phytotoxic on many test species in<br />

South Korea (Chou, 1990) and the chemicals isolated include mimosine, quercetin<br />

and gallic, protocatechuic, p-hydroxybenzoic, vanillic, caffeic and p-coumaric acids.<br />

Recuperation of nutrients absorbed by the crops in small holder farms of developing<br />

countries has been traditionally taken care of through the incorporation of several<br />

kinds of bulky organic manures. Besides crop residues, other farm wastes such as<br />

cattle manure and some post harvest byproducts of crops are frequently used for<br />

manuring the crops. Filter pressmud, a bi-product of sugar factories, that crystallizes<br />

sugar from cane juice is the precipitated impurity from cane juice. Pressmud from<br />

cane juice accounts for about 3% and all sugar factories accumulate enormous quantities<br />

of pressmud. This bi-product is widely used in South Asian countries as organic<br />

manure. This pressmud upon incorporation in the soil, prior to transplanting of rice<br />

in wetlands, releases allelopathic metabolites that contribute to weed suppression.


ALLELOPATHY FOR WEED CONTROL 107<br />

Allelopathic suppression of weeds by the application of pressmud as organic manure<br />

in wetland agriculture has been included as a vital weed management component in<br />

Integrated Systems of Farming (Arulchezian and Kathiresan, 1990; Kathiresan, 2004a).<br />

2.3. Invasive Weeds with Allelopathic Potential<br />

Invasive species, particularly those non-indigenous to terrestrial, aquatic, wetland,<br />

and wildlife habitat, cause extensive environmental and economical damage to native<br />

ecosystems. Invasive plant species alone cause losses in excess of $35 billion annually<br />

in the U.S.A. (Pimentel et al., 2000). Non-indigenous weeds in the U.S.A. are estimated<br />

to invade 700,000 hectares of wildlife habitat per year (Babbitt, 1998). Invasive aquatic<br />

weeds are a significant problem in the U.S.A. Aquatic weeds choke waterways, reduce<br />

recreational use of lakes and rivers, and alter aquatic animal species (Pimentel et al.,<br />

2000). In the U.S.A. alone, an estimated $100 million is spent annually for managing<br />

aquatic weed species (OTA, 1993). An estimated $15 million is spent on managing<br />

the aquatic weed species hydrilla [Hyrdilla verticillata (L.F.) Royle] in the state of<br />

Florida, U.S.A. (Center et al., 1997). Invasive weeds in terrestrial ecosystems such as<br />

pastures cause estimated losses of $2 billion annually in the U.S.A. (Pimentel, 1991).<br />

Examples of invasive weeds in pastures are cogongrass [Imperata cylindrica (L.)<br />

Beauv.], yellow starthistle (Centaurea solstitialis L.), and purple loosestrife (Lythium<br />

salicaria L.). Specifically, cogongrass is a C 4 , rhizomatous, perennial monocot that<br />

has become an invasive weed in many gulf-states of the southeastern U.S.A. since its<br />

introduction to the U.S.A. in the late 19 th and early 20 th centuries (Byrd and Bryson,<br />

1999; Dickens, 1974; Dickens and Buchanan, 1971; Elmore, 1986). Cogongrass grows<br />

in tropical, subtropical, and some temperate regions of the world (Akobundu and<br />

Agyakwa, 1998; Bryson and Carter, 1993), and is found on all continents except<br />

Antarctica (Holm et al., 1977). Cogongrass is among the most troublesome weeds<br />

worldwide (Falvey, 1981; Holm et al., 1977). It is extremely competitive with crops<br />

and neighboring plant communities (Eussen and Wirjahardja, 1973). It has been<br />

reported to reduce corn (Zea mays L.) grain yield by 80 to 100% (Koch et al., 1990;<br />

Udensi et al., 1999). Koch et al. (1990) also reported > 90% yield reduction for<br />

intercropped corn and cassava (Manihot esculenta Crantz) grown in cogongrass<br />

infested fields.<br />

Cogongrass competes with other plant species commonly found in similar<br />

ecosystems (roadways, pastures, mining areas, parks, and other natural and recreational<br />

areas) via allelopathic-type mechanisms. Residues and liquid exudates of cogongrass<br />

foliage and root residues have been shown to reduce germination and shoot and root<br />

growth of monocot and dicot weed and crop species (Koger and Bryson 2004; Koger<br />

et al., 2004). Germination of bermudagrass [Cynodon dactylon (L.) Pers.] and Italian<br />

ryegrasses (Lolium multiflorum Lam.) was reduced by as much as 97%, and shoot and<br />

root growth by as much as 96% at the highest concentrations. However, cogongrass<br />

had no allelopathic effect on hemp sesbania [Sesbania exaltata (Raf.) Rydb. Ex A. W.<br />

Hill] (Koger and Bryson, 2004). Phenolic compounds present in foliage and roots of<br />

cogongrass may be responsible for the allelopathic inhibition of germination and


108<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

seedling development of other species. Inderjit and Dakshini (1991) reported several<br />

phenolic compounds extracted from leachates of cogongrass foliage and roots/rhizomes<br />

reduced germination and shoot and root length of mustard [Brassica juncea (L.)<br />

Czern and Coss.] and tomato (Lycopersicon esculentum Mill.). Inderjit and Dakshini<br />

(1991) also found phenolic compounds in leachates of soil collected near the<br />

rhizosphere of cogongrass as well as up to 3 m away that were not present in control<br />

soils. These reports suggest that allelopathic effect of cogongrass is species-specific.<br />

Thus, allelochemicals from cogongrass may serve as potential leads in discovery of<br />

new selective herbicides.<br />

2.4. Plant Products for Allelopathic Control of Weeds<br />

Presence of abundant moisture in wetlands allow faster transport of allelochemicals<br />

from the applied plant products to the target weeds. Aquatic systems have the advantage<br />

of no soil-interface that restricts the activity of allelochemicals on susceptible flora.<br />

As a result, use of plant products for allelopathic control of weeds has more potential<br />

in wetlands and aquatic systems than soil containing terrestrial systems. Carrot grass<br />

(Parthenium hysterophorus L.), an invasive obnoxious weed originated from Mexico<br />

has shown to be allelopathic on another introduced invasive aquatic weed water<br />

hyacinth [Eichhornia crassipes (Mart.) Solms. Laubach] in India. Dried residues of<br />

leaves and flower of carrot grass applied in to the water at 0.5% (w/v) killed water<br />

hyacinth within one month. Residues of leaves of carrot grass showed the highest<br />

biological activity followed by flowers, stem and roots. The flowers and leaves of<br />

carrot grass also had higher total phenolic acid levels in the medium which was<br />

responsible for the inhibition (Pandey et al., 1993). Dry powder of an epiphytic plant<br />

Cassytha sp dissolved in water infested with water hyacinth at 1-2% (w/v) resulted in<br />

complete desiccation of leaves of water hyacinth with in 15 days and drastic reduction<br />

in biomass (Kauraw and Bhan, 1994). Parthenin, a sesquiterpene lactone that is one<br />

of the major toxins in the weed, proved lethal to water lettuce (Pistia stratiotes Linn.)<br />

and duck weed (Lemna perpusilla Torr.) at 50 ppm concentration and to water hyacinth<br />

at 100 ppm concentration. The mechanism of allelopathic inhibition by parthenin<br />

was identified to be associated with decline in water use, root dysfunction, excessive<br />

leakage of solutes resulting in massive damage to cellular membranes, loss of<br />

dehydrogenase activity in the roots and destruction of chlorophyll in the leaves (Pandey,<br />

1996a).<br />

Among several allelochemicals (p-hydroxybenzoic acid, anisic acid, salicylic acid,<br />

coumaric acid, fumaric acid, tannic acid, gallic acid, chlorogenic acid, vanillic acid,<br />

caffeic acid and ferulic acid), p-hydroxybenzoic acid had highest phytotoxicity on<br />

aquatic weeds and was lethal at 50 ppm to all weeds tested, including floating aquatic<br />

weeds like salvinia (Salvinia molesta Mitchell), azolla (Azolla nilotica Decne. ex<br />

Mett.), spirodella (Spirodella polyrhiza (L.) Schieiden) and lemna (Lemna<br />

paucicostaha Hegelm.), and submerged weeds like hydrilla (H. verticilata),<br />

ceratophyllum (Ceratophyllum demersum L.) and najas (Najas graminea Del.).<br />

However, p-hydroxybenzoic acid was lethal for water hyacinth and water lettuce only


ALLELOPATHY FOR WEED CONTROL 109<br />

at 100 ppm. Water hyacinth (E. crassipes), water lettuce (P. stratiotes) and water fern<br />

(S. molesta Baker.) were relatively more tolerant to allelochemicals except for phydroxybenzoic<br />

acid compared to other floating or submerged weeds (Pandey, 1996b).<br />

An Indian medicinal herb Coleus amboinicus Lour., commonly used for curing<br />

cold, flu and other such ailments upon raw consumption, showed remarkable activity<br />

on water hyacinth among different weeds and herbs tried for their allelopathy on<br />

water hyacinth. Dried leaves of this medicinal herb C. amboinicus were ground to<br />

powder and applied to the water system as a suspension (30 g L -1 ). Death of water<br />

hyacinth occurred within 24 h and nearly 100% reduction in biomass was achieved<br />

within 9 days. However, spraying C. amboinicus over the foliage of water hyacinth at<br />

100 g L -1 proved ineffective due to lack of penetration and absorption of allelochemicals<br />

from dried powder into the plant system. Applied as a suspension in water, the natural<br />

product was active, killing the weed even at lower dosages of 12.5 g L -1 within 14<br />

days (Kathiresan and Kannan, 1998). In the Philippines, 57 different compounds,<br />

including α humulene, carvacrol, thymol, α-pinene and α-terpine were identified<br />

from C. amboinicus. Some of these compounds showed a high degree of biological<br />

activity, proving lethal to several micro organisms, insects and snails (Vasquez et al.,<br />

1999). C. amboinicus powder was shown to inhibit algal growth in static-water systems<br />

(Kathiresan, 1998).<br />

Graded dosages of the natural products of C. amboinicus were evaluated for their<br />

inhibitory effect on water hyacinth (Kathiresan, 1999) through specific bio-assay<br />

methods, involving whole plants and cut leaves, separately. The study indicated that<br />

dry powder of C. amboinicus is a candidate plant product for the control of floating<br />

aquatic weeds water hyacinth, water lettuce and duck weed with rapid biological<br />

activity. The biological activity is due not only to mere water loss but also physiological<br />

effects (as compared to sodium chloride at 40 g L -1 , the standard desiccant with 25%<br />

weed biomass reduction after 9 days). The data also implies a higher magnitude of<br />

inhibition on root system (Kathiresan, 2000).<br />

The lowest dosage of 1% concentration (10 g L -1 ) may seem high, but is similar<br />

to that widely used for therapeutic purposes. Earlier studies have shown that water<br />

hyacinth was relatively tolerant to allelochemicals, as 100 ppm of p-hydroxybenzoic<br />

acid was required to cause death, whereas 50 ppm concentration was sufficient for<br />

other weeds (Pandey, 1996b). Similarly, fungal oxalates and pure oxalic acid that<br />

caused considerable and severe chlorosis in other aquatic weeds induced only slight<br />

chlorosis in water hyacinth (Charudattan and Lin, 1974). Under these circumstances,<br />

a crude extract of a safe medicinal herb C. amboinicus, causing death of water hyacinth<br />

within 9 days, might offer an effective control strategy. Further, a stable inhibitory<br />

response caused by C. amboinicus powder applied to cut leaves of water hyacinth<br />

under controlled conditions in static water at dosages ranging from 30 g L -1 down to<br />

1.0 g L -1 reduced fresh weight from 61 to 49% respectively, suggesting that C.<br />

amboinicus dry powder could exert adequate allelopathy at low dosages. The lowest<br />

dosage of 0.1 g L -1 (100 ppm) of C. amboinicus caused 24% fresh weight reduction.<br />

Apparently, the dosage of C. amboinicus powder lethal to water hyacinth could be<br />

reduced drastically, if either the natural product or the active principle is formulated


110<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

to enhance absorption through foliage. In India, when the sources of irrigation water<br />

recede during the summer, the smaller volume of water is more accessible for treating<br />

with C. amboinicus at 10 g L -1 (1% w/v). C. amboinicus also hold promise for biocontrol<br />

of water hyacinth on small farms. In another aquatic habitat, Myriophyllum<br />

spicatum L. has been shown to be allelopathic on submerged macrophytes and the<br />

compound that was identified as major active allelochemical Tellimagrandin II (Gross<br />

et al., 1996) also interferes with photosystem II and photosynthetic oxygen evolution<br />

in Anabaena sp. (Leu et al., 2002).<br />

Barnyardgrass is a problem weed in wetland transplanted rice. The weed<br />

infestation begins with use of contaminated rice seeds. Morphology of barnyardgrass<br />

closely resembles rice during seedling stage. The seedlings are pulled along with rice<br />

seedlings from the nursery and transplanted in the main field. Transplanted seedlings<br />

of barnyardgrass may escape control with pre-emergence herbicides in the main field.<br />

Allelopathy of rice seeds is shown to be stimulatory to the germination of barnyardgrass.<br />

Farmers in developing countries normally soak the gunny bags with rice seeds harvested<br />

in the previous season, in water canals overnight prior to seeding. The allelochemiclas<br />

similar to gibberellin released from the rice seeds help break the dormancy of<br />

barnyardgrass and stimulate germination (Kathiresan and Gurusamy, 1995). To<br />

overcome this allelopathic stimulation of barnyardgrass, it is recommended that small<br />

farmers soak the rice seeds in open containers, skim the floating barnyardgrass seeds<br />

(due to differential specific gravity) as a preventive weed control measure. However,<br />

in other studies rice hull extracts have been shown to exert allelopathic inhibition of<br />

germination and growth in barnyardgrass (Ahn and Chung, 2000).<br />

3. ALLELOPATHIC MICROORGANISMS<br />

Microbes play a vital role in designing and implementing allelopathic control of aquatic<br />

weeds and certain weeds of wetland eco-systems. They serve as agents for classical<br />

bio-control in several parts of the world with a simple idea of locating highly host<br />

specific agents from the native range of the weed and introducing them in new regions<br />

requiring control where the weed has established, after rigorous experimental<br />

evaluation. The microbial toxins either in their parental form or as metabolites also<br />

offer scope to serve as effective herbicides for weed control. Some of the microbial<br />

toxins exhibit a new mode of action with novel target site that could stimulate discovery<br />

of new herbicidal chemistry.<br />

3.1. Pathogens for Weed Control<br />

Bio-control using pathogens holds promise mostly in non-cropland situations because<br />

of the slow pace of control of weeds and the wider window for control as compared<br />

with the shorter window of the cropping season and associated disturbances under<br />

cropland situations. However, in aquatic systems they appear to be more realistic, due<br />

to the absence of such cropping barriers and enhanced mode of dispersal in water.<br />

Among the wetland weeds, Echinochloa crusgalli and Cyperus rotundus are being


ALLELOPATHY FOR WEED CONTROL 111<br />

evaluated for biological control using the fungi Exserohilum monocerus and Dactylaria<br />

higginsi (Luttrell) MB Ellis, respectively (Charudattan and Dinoor, 2000). Fungal<br />

pathogens of aquatic weeds have been identified from Neotropical regions since the<br />

late 1970s. Alligator weed (Alternanthera philoxeroides (Mart.) Griseb) introduced<br />

from Brazil, became a serious invader of aquatic systems in Australia, Asia and North<br />

America. A preliminary survey of fungal pathogens on alligator weed in Brazil has<br />

yielded two species Nimbya alternantherae (Holocomb and Antanopoulos) Simmons<br />

and Alcorn and Cercospora alternantherae Ellis and Langois (Barreto and Torres,<br />

1999). Both caused leaf spots on the alligator weed and their virulence varied with<br />

geographic range and altitude. Aquatic ferns such as Salvinia molesta and Salvinia<br />

auriculata JB Aublet., are regarded as invasive aquatic weeds in many of the water<br />

habitats in Neotropics, Africa and Asia. Necrotic spots caused by the fungus Drecshlera<br />

sp. on S. auriculata and the inoculum was easily produced and pathogenic (Muchovej<br />

and Kushalappa, 1979). Cattail (Typha domingensis Pers.,) native to Neotropics has<br />

also evolved to be invasive in aquatic systems. The fungal pathogen Phoma typhae –<br />

domingensis is regarded as a potential candidate for the development of mycoherbicide<br />

for cattail (Barreto and Evans, 1996). Egeria densa Planch, a worse submerged aquatic<br />

weed interrupts hydro electricity generation, damages grids, and causes substantial<br />

economic loss to hydroelectric companies. Among eight fungal isolates that have<br />

shown promise, Fusarium graminearum Schw. was pathogenic to E. densa causing<br />

chlorosis followed by necrosis and complete tissue disintegration (Barreto et al., 2000).<br />

Three different fungi (Chaetomella raphigera Swift, Cercospora sp. and<br />

Mycosphaerella sp.) caused severe blight in parrot’s feather (Myriophyllum aquaticum<br />

(Vell.) Verdc), an invasive aquatic weed (Barreto et al., 2000). Water hyacinth,<br />

recognized as the world’s worst aquatic weed (Holm et al., 1977), has been the prime<br />

target weed for bio-control. One of the first pathogens to be patented as a mycoherbicide<br />

was Cercospora piaropi Tharp. (Charudattan, 1996). Other promising pathogens are<br />

Uredo eichhorniae Fragosco and Ciferri, Myrothecium roridum Tode ex Fr. Alternaria<br />

eichhorniae NagRaj and Ponnappa (Barreto et al., 2000).<br />

3.2. Microbial Toxins as Leads for Herbicides<br />

The unique relationship between plants and their pathogens suggest that<br />

microorganisms may be a better source of future herbicides than allelochemicals<br />

produced by higher plants (Duke, 1986). The major constraint with most<br />

allelochemicals from higher plants is that their range of selectivity is narrow and they<br />

are often autotoxic. Perhaps, this may be one reason that breeding for crop plants to<br />

produce higher levels of allelochemicals has not been aggressively pursued. In contrast,<br />

many microbial phytotoxins are both selective and efficacious at low rates. Bialaphos<br />

[L-2-amino-4-(hydroxyl) (methyl) o-phosphinyl)-butyryl-L-alanyl-L-alanine], a<br />

tripeptide extract from Streptomyces viridichromogens Schulz. Freiburg. marketed<br />

as Herbiace®, tentoxin, a cyclic tetrapeptide produced by Alternaria alternata (Fr.)<br />

Keissl., rhizobitoxine produced by Rhizobium japonicum Kirchner and an analogue<br />

of cystathionine are a few examples. Several non-host selective phytotoxins produced<br />

by microorganisms have been reported (Duke, 1986; Hoagland, 2001).


112<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

4. ALLELOPATHY – LEAD FOR NOVEL HERBICIDAL CHEMISTRY<br />

<strong>Allelochemicals</strong> can provide potential leads in discovery of new herbicides with novel<br />

molecular target sites of action. Hydantocidin from Streptomyces hygroscopicus B.<br />

Straubinger is an analogue of the allelochemical phosphinothricin and inhibits<br />

adenylosuccinate synthetase. This analogue mimics the substrate inositol<br />

monophosphate (IMP) and binds the enzyme 1000 fold tighter than IMP, thereby<br />

forming a dead-end complex. Hadacidin and alanosine, both microbial products are<br />

also inhibitors of this enzyme (Duke et al., 2000). To date, bialaphos and glufosinate<br />

derived from microbial compounds are the two most successfully commercialized<br />

herbicides (Hoagland, 2001). Glufosinate is the synthetic version of phosphinothricin,<br />

a breakdown product of bialaphos (Hoagland, 2001; Lydon and Duke, 1999). Bialaphos<br />

sold as herbicide in Japan is derived from Streptomyces species is a proherbicide that<br />

is metabolically degraded to phosphinohricin by target plant in order to be herbicidally<br />

active. Glufosinate is the only commercial herbicide that inhibits glutamine synthetase<br />

despite plethora of natural and synthetic products known to inhibit glutamine synthetase<br />

(Lydon and Duke, 1999). A gene coding for an enzyme that acetylates the active acid<br />

of phosphinothricin rendering it non-phytotoxic, phosphinothricin acetyl transferase<br />

(Pat) gene has been isolated from S. hygroscopius. This gene has been used in truncated<br />

form to transform crops such as maize to impart tolerance to glufosinate (Copping,<br />

2002). Other examples of herbicides derived from natural product chemistry are<br />

triketones from Syngenta and cinmethlin from BASF. Another potential herbicide of<br />

microbial origin is AAL-toxin, a natural metabolite from Alternaria alternata f sp.<br />

lycopersici. Monocots were generally immune to the effects of AAL-toxin, whereas<br />

several broad leaved species are susceptible. Abbas et al. (1995) proposed that the<br />

selective weed control could be possible through AAL-toxin.<br />

5. SCREENING OF ALLELOPATHIC PLANT MATERIALS<br />

Allelopathy is essentially a chemical defense mechanism used by plants to keep other<br />

plants out of their space. Though allelopathy forms a part of network of chemical<br />

communication between plants, it is part of plant interference along with competition<br />

for resources. Competition involves the removal or a diminution of shared resources,<br />

while allelopathy involves the addition of a chemical compound to the environment<br />

through different processes (Rice, 1984; Putnam, 1985). Allelopathic chemicals in<br />

general affect seed germination, root growth, shoot growth and/or seedling vigor in<br />

the early stages of the receiver’s growth and may interfere with metabolic functions<br />

like photosynthesis, membrane permeability, biosynthesis of enzymes, lipids, protein,<br />

etc. as the receiver progress in growth processes. In aquatic systems and wetlands,<br />

screening of allelopathic plant materials for biological efficiency is relatively easier<br />

as the allelochemicals are frequently absorbed through the roots of the receiver and<br />

transported from the donor directly through water without much of resistance from<br />

the soil-interface. However, screening processes have different phases involving either<br />

larger size or larger population of target weeds in different environments. Screening


ALLELOPATHY FOR WEED CONTROL 113<br />

is used mainly to confirm the biological efficiency of the substances to a magnitude<br />

which would at least serve as component of integrated weed management though not<br />

as a holistic weed control measure. Such a rigorous or repeated experimentation with<br />

different sets or modes of bio-assay becomes imperative as many of the allelochemicals<br />

exhibit inhibitory response on seedling germination and establishment but seldom<br />

lethal on large sized receiver plants. Screening techniques for allelopathy in aquatic<br />

systems or allelochemicals which are transported mainly through water body should<br />

include some critical points raised by Willis (1985), Putnam and Tang (1986), and<br />

Cheng (1992). They are:<br />

• Pattern of inhibition of one species by another must be shown using suitable<br />

controls, describing the symptoms and quantitative growth reduction;<br />

• The putative aggressor plant must produce a toxin;<br />

• There must be a mode of toxin release from the plant to the environment and thus<br />

the target plant;<br />

• The afflicted plant must have some means of toxin uptake, be exposed to the<br />

chemical in sufficient quantities and time to cause damage, and to notice similar<br />

symptoms;<br />

• The observed pattern of inhibition should not be explained solely by physical<br />

factors or other biotic factors, especially competition.<br />

The first phase or initial stage of screening include bio-assays. Those plant<br />

materials that are confirmed to possess biological activity through bio-assay need to<br />

be further studied for their dose response pattern under controlled environment. Plant<br />

materials elucidating appreciable response even at minimal doses could be further<br />

evaluated for their allelopathic potential under natural habitats.<br />

5.1. Bio-Assays<br />

Bioassays are an integral part in all studies of allelopathy. They have multiple uses<br />

such as evaluating allelopathic potential of different plant material, tracing activity<br />

during extraction, purification, and identification of bioactive compounds. The<br />

techniques used vary greatly and one has to standardize the procedure independently,<br />

and modify to suit the occasion and conditions. According to Rice (1984) and Putnam<br />

and Tang (1986), seed germination is used as a test in most bioassays. Though different<br />

types of bioassays are used, all of them in general include seed placed on substrate<br />

saturated with the test solution. Germination, as indicated by the emergence of the<br />

radical 2 mm beyond the seed coat is scored over a period of time. The factors that<br />

need to be kept constant are oxygen availability, osmotic potential of the test solution,<br />

pH, and temperature. The elongation of the hypocotyls or coleoptiles are often observed<br />

along with germination. Dry biomass which is easier to measure could be used as a<br />

measure of growth (Bhowmik and Doll, 1984). Sensitivity is normally higher in growth<br />

bioassays than in germination bioassays. When the quantity of allelochemicals is<br />

limited, agar cultures can be used. Pre-germinated seeds can be exposed to the agar<br />

cultures containing test solution.


114<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

Bioassay for searching of allelopathy in aquatic weeds is comparatively easier<br />

and this could be designed under both laboratory or greenhouse conditions using<br />

either part or whole plant of the aquatic weed (Kathiresan, 2000; Kannan, 2002).<br />

Whole plants of floating or submerged aquatic weeds targeted for allelopathic control<br />

can be grown in suitable containers with water containing standardized nutrients.<br />

The powder of candidate allelopathic substances or plant products are added to water<br />

either on w/v or v/v basis with appropriate untreated controls. Periodic observations<br />

at designated intervals could include reduction in root length and mass, reduction in<br />

shoot length and mass, desiccation, scorching and bleaching or mortality score similar<br />

to herbicide injury. Based on the screening data, plant products could be classified in<br />

to highly allelopathic, moderately allelopathic, and less allelopathic. For example, 55<br />

different plant products were screened for allelopathic inhibition of water hyacinth<br />

involving whole plants as well as cut leaves at Annamalai University. In bio-assays<br />

involving whole plants, ten of them including C. amboinicus, P. hysterophorus and<br />

L. leucocephala were highly allelopathic based on fresh weight reduction (>30%) of<br />

water hyacinth within 48 hr after treatment. Another 12 including Acalypha indica<br />

Linn., Trianthema portulacastum L. and Sesbania grandiflora (L.) Pers. showed<br />

moderate allelopathy, fresh weight reduction of water hyacinth was 15-30% within<br />

48 hr after treatment. Twelve other plant products including Croton sparsiflorus<br />

Morong, Cleome viscosa L. and Eclipta alba L. showed less allelopathy, fresh weight<br />

reduction of water hyacinth was less than 15% within 48 hr after treatment (Table 1).<br />

The remaining 21 plant species including Leucas aspera Spreng. Curcuma longa L.<br />

and Euphorbia hirta L. did not show any allelopathic effect on water hyacinth (Kannan,<br />

2002). To measure dose dependant responses of allelopathic substances more precisely<br />

Table 1 : Percentage reduction in fresh weight of water hyacinth (Eichhornia<br />

crassipes.) due to various plant products. (Kannan, 2002) a .<br />

Treatments Days After Treatment<br />

plant products @ 30 g L-1 2 4 6 8<br />

Coleus amboinicus 56.57 (69.66) 90.00 (100.00) 90.00 (100.00) 90.00 (100.00)<br />

Acalypha indica 29.98 (24.97) 44.24 (48.68) 90.00 (100.00) 90.00 (100.00)<br />

Leucas aspera -b - - -<br />

Croton sparsiflorus 20.53 (12.30) 34.51 (32.11) 45.37 (50.66) 90.00 (100.00)<br />

Curcuma longa - - - -<br />

Trianthema portulacastum 27.06 (20.70) 45.09 (50.17) 90.00 (100.00) 90.00 (100.00)<br />

Cleome viscosa 21.10 (12.96) 35.77 (34.17) 45.66 (51.16) 90.00 (100.00)<br />

Leucaena leucocephala 34.84 (32.65) 90.00 (100.00) 90.00 (100.00) 90.00 (100.00)<br />

Sesbania grandiflora 27.17 (20.86) 44.51 (49.16) 90.00 (100.00) 90.00 (100.00)<br />

Parthenium hysterophorous 36.92 (36.10) 90.00 (100.00) 90.00 (100.00) 90.00 (100.00)<br />

Euphorbia hirta - - - -<br />

Eclipta alba 21.27 (13.17) 33.93 (31.17) 44.94 (49.91) 90.00 (100.00)<br />

Control - - - -<br />

SED CD (P = 0.05)<br />

2.40<br />

4.80<br />

2.34<br />

4.69<br />

1.25<br />

2.51<br />

-<br />

-<br />

a Figures in parenthesis are original values before arc-sine transformation.<br />

b No allelopathic effect.


ALLELOPATHY FOR WEED CONTROL 115<br />

and to detect the sensitivity of the target weed to minute quantities of allelopathic<br />

substances, incised plant parts of aquatic weeds would serve better than whole plants.<br />

For screening of different allelopathic substances for inhibition of water hyacinth, a<br />

specific bio-assay method was designed with cut leaves of water hyacinth (Kathiresan,<br />

2000; Kannan, 2002). This bioassay involved exposure of cut leaves of water hyacinth<br />

to graded doses of plant materials to be tested (dissolved in water). The leaves of<br />

water hyacinth plants (with healthy leaves submerged in water) were detached by<br />

cutting the petiole with a razor blade with care to retain the incision point below<br />

water level. The detached portions of leaf with a part of petiole intact was kept<br />

submerged in water for 90 seconds to ensure that no air was trapped internally. Then<br />

these leaves were transferred to scintillation vials with water where in different plant<br />

products were dissolved and individually compared with an untreated control. The<br />

percent fresh weight reduction of the cut leaves was calculated using the formula<br />

Initial weight of the cut leaves – weight after 24 hr of treatment<br />

X 100<br />

Initial weight of cut leaves<br />

5.2. Dose Response Studies<br />

Even if a plant product proves appreciably allelopathic on aquatic weeds in bio-assay,<br />

tracing the pattern of allelopathic inhibition with graded doses of the plant product<br />

becomes vital as aquatic systems requires enormous quantities of plant product due to<br />

the quantum of water body (dilution), which in many cases is not practically feasible.<br />

Furthermore, this dose response has to be plotted for differing morpho-physiological<br />

states of the weed that occurs in common, as the quantity of allelochemical required<br />

for a knock down effect is less with a small statured weed compared to that of larger<br />

sized weed. For this purpose, before screening of allelopathic plant products on water<br />

hyacinth, the latter was classified in to different morphological stages (large, medium<br />

and small) prevalent in the state of Tamilnadu, India using discriminant analysis<br />

(Kannan and Kathiresan, 1998). Both whole plant and cut leaf bio-assays were done<br />

on each of the three stages of water hyacinth. Higher (10 to 30 g L -1 ) doses were used<br />

in whole plants to impart a near lethal effect on the whole plant and lower doses were<br />

used in cut leaf bio-assays to cause some allelopathic injury, if not lethal. The doses<br />

used for water hyacinth cut leaf bio-assay were 30, 25, 20, 15 10, 5 2.5, 1, 0.5, 0.25,<br />

and 0.1 g L -1 . The dose response data revealed that cut leaf-bioassay was superior to<br />

whole plant assay. For example, C. amboinicus at as low as 0.1 g L -1 dose caused 24%<br />

fresh weight reduction in water hyacinth within a week of exposure (Kathiresan,<br />

2000). Therefore, cut leaf bio-assay could be useful to detect allelopathic potential of<br />

plant products which otherwise may have been missed if whole plant assay was used.<br />

In contrast, the dose response study data revealed that lethal doses for large, medium,<br />

and small plants of water hyacinth were relatively closer. C. amboinicus at low (10 g<br />

L -1 ) dose caused death of the water hyacinth after 20 days whereas at high (25 g L -1 )<br />

dose caused death within one week.


116<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

5.3. Field Testing of Allelopathic Substances<br />

In many instances, the materials or plant substances that prove to be allelopathic in<br />

laboratory or pot culture experiments may not elucidate similar magnitude of<br />

allelopathic response on aquatic weeds in aquatic environments, watersheds, or<br />

wetlands. Hence, it is imperative to confirm plant products for their allelopathic<br />

potential on weeds in their own natural habitat. A search was made on allelopathic<br />

plant products for use in water hyacinth control programs at Department of Agronomy,<br />

Annamalai University. Ten of 55 different plant products that showed allelopathic<br />

suppression of water hyacinth within 48 h of treatment were selected and tested for<br />

their efficacy in natural habitats. The field testing was done in a two tier model. First,<br />

the ten plant products were tested in microponds (simulated natural habitat). Second,<br />

the plant products that confirmed to be allelopathic in microponds were further<br />

evaluated in natural watersheds.<br />

5.4. Preparation of Microponds<br />

Small (microponds) ponds of 1m x 1m x 1m dimension were dug in the field and<br />

canals with running water on both sides of the pond were maintained to minimize<br />

seepage loss of water from the ponds. Initially, the microponds were conditioned by<br />

flooding and maintaining water as needed for 20 days. After 20 days, healthy water<br />

hyacinth plants were introduced and allowed to grow for 2-3 days. The plant products<br />

to be tested were dissolved in the water of microponds, separately. Allelopathic potential<br />

of the plant products was evaluated in comparison with untreated control using data<br />

on fresh weight, number of healthy leaves, and chlorophyll content of the water<br />

hyacinth.<br />

5.5. Evaluation in Natural Habitats<br />

Top three best performing plant products in the micropond study were selected for<br />

realistic confirmation of their allelopathic potential in natural habitats. To this end,<br />

three watersheds of 168, 123, and 492 m 2 area were selected at three different locations<br />

within a radius of 50 km from Annamalai University. Water hyacinth was allowed to<br />

establish in watersheds and the surface area of water was completely covered by weed.<br />

Each watershed was divided in to three equal strips using polyethylene sheets stretched<br />

between bamboo poles running down the entire depth of water, with the poles anchored<br />

to the bottom of the watershed. The plant products were applied to water individually<br />

to one of the strips in each watershed. The data were collected on fresh weight, number<br />

of healthy leaves, and chlorophyll content at five days interval. Results proved that<br />

the plant products showing higher magnitude of allelopathic inhibition on water<br />

hyacinth in initial bio-assays like C. amboinicus continued to retain and exhibit the<br />

same magnitude of inhibition on water hyacinth with out any dissipation in natural<br />

watersheds or aquatic environments.


ALLELOPATHY FOR WEED CONTROL 117<br />

6. ALLELOPATHY AND INTEGRATED WEED MANAGEMENT<br />

Because of limited resources, an average farmer in a developing country can neither<br />

afford to take big economic risks nor opt for technologies associated with a lot of<br />

external inputs. As a result, research on vegetation management strategies capable of<br />

minimizing weed infestation and simultaneously favoring sustainable crop production<br />

that are economical and eco-friendly needs attention (Akobundu, 2000). Allelopathy<br />

fits in to this approach as one of the integral principles in any such cropping systems<br />

involving crop rotation, inter-cropping, cover crop, and off-season land management<br />

(such as raising green manures and ploughing in situ). Linking similar integrated<br />

farming approaches to integrated weed management and integrated pest management<br />

helps to address bio-diversity concerns with a simultaneous reduction in agrochemical<br />

use especially in low input agriculture and small hold farms. A 3-yr study of weed<br />

management in wetland transplanted rice, rice - mung bean cropping sequence with<br />

treatments assigned to the same plots every season at Annamalai University revealed<br />

that lowland weeds like C. difformis was drastically reduced by the introduction of a<br />

relay crop of mung bean in the sequence (Kathiresan, 2002). Raising a green manure<br />

crop of Sesbania aculeata Poir in the off-season (May - July) and ploughing it in situ<br />

at the age of 45 days, before the cultivation of rice in the first (August - January) as<br />

well as second (January - April) season, helped in reducing weed competition in both<br />

the rice crops (Gnanavel and Kathiresan, 2002). Off-season land management such<br />

as raising green manure crop significantly reduced the weed seed reserves in the soil<br />

through allelopathic interference, whereas rotation of an upland crop like mung bean<br />

with rice interrupted the weed flora in lowland through mung bean residues.<br />

Integrated weed management assumes significance in managing aquatic systems.<br />

Use of herbicides are constrained with drastic reduction in water quality and ultimate<br />

ill effect on associated non-target organisms. In countries like India, herbicides are<br />

yet to get registered for use in aquatic systems. Under these conditions managing<br />

infestations of water hyacinth, water fern, and water lettuce is challenging. In one of<br />

the recreational lakes with tourist attraction in a hill resort in Ooty, in the state of<br />

Tamilnadu, India, the public authority has spent heavily (Indian Rupees 1.25 crores,<br />

about US $200,000) for manual clearing of water hyacinth for one time. Similarly,<br />

thousands of army personnel were used for clearing water hyacinth in a lake in<br />

Bangalore, the capital city of Karnataka State, India. Classical biological control is<br />

the only option available and that too is difficult in situations where the water body<br />

dries off in the peak summer, leaving the released insects to starve and die due to<br />

interrupted host range. Accordingly, integration of short term control measures with<br />

classical biocontrol might offer excellent results. Allelopathy reinforced classical biocontrol<br />

research has been targeted and taken up at Department of Agronomy,<br />

Annamalai University through National Agricultural Technology Project funded by<br />

Indian Council of Agricultural Research. This project originated from the basic concept<br />

of allelopathic inhibition of water hyacinth by C.amboinicus as mentioned earlier.<br />

However, the requirement of plant product for treating larger watersheds might pose<br />

practical difficulties. Previous results also indicated that if absorbed in to plant through


118<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

foliage, the plant product could be very effective even under very low doses (0.1 g<br />

L -1 ). The only hurdle faced for application of the plant product on the foliage is retention<br />

of plant product due to the repulsion by leaf cuticle. Any rupture and/or damage to<br />

leaf cuticle could potentially enhance absorption of plant product. To this end, the<br />

well established insect bio-control agents in India, Neochetina bruchii Hustache /<br />

eichhorniae Warner were chosen for the study to serve as a component of integrated<br />

weed management. These weevils normally scrape on the leaves of water hyacinth.<br />

An attempt was made to integrate both these bio-control tools viz. classical biocontrol<br />

using N. bruchii / eichhorniae and application of the plant product C.<br />

amboinicus. Integrating both the tools are possible with two different sequences.<br />

Treating the water body first with plant product at a lesser dose with the expectation<br />

that it will reduce the vigor of the weed, predisposing it for faster and rapid destruction<br />

by the insect agents that are to be released later is one possibility. Whereas releasing<br />

the insect agents first on the weed host, allowing them to make leaf scrapings that<br />

might help foliar uptake of plant product that could be sprayed later is another. Both<br />

these sequences were compared in the study. It was observed that treating the water<br />

body first with plant product followed by the release of insect agents on the weed<br />

showed an antagonistic interaction, as the insects migrated from treated, partially<br />

killed plants to healthy plants. The second sequence of releasing the insect agents<br />

first followed by spraying of the plant product on the weed foliage produced an additive<br />

or synergistic response with rapid and complete weed control with in a single season.<br />

The optimum inoculation loads of insect agents, concentration of the spray fluid of<br />

plant product required, length of interlude between the release of insect agents and<br />

spraying of plant product were standardized for all three different growth stages of<br />

the weed, and the success of this integrated approach was demonstrated at three different<br />

watershed environments in the state of Tamilnadu, India. The plant product was also<br />

shown to be safe for the insect agents with out inducing migratory behavior and<br />

without causing any histo-pathological injury on different tissues of the insects like<br />

salivary gland, gut, cutin, testis, and brain. Further, the integrated approach also<br />

proved safe for non-target organisms and water quality (Kathiresan, 2004b).<br />

7. CONCLUSIONS<br />

Plants can interfere with each other through allelopathy or competition for resources.<br />

Allelopathy can be used in weed management in several ways including cover crops,<br />

smother crops, green manure crops, breeding for allelopathic crop cultivars, mulching<br />

and crop residue management. Allelopathic suppression of weeds will not replace<br />

synthetic herbicides which are the dominant method of weed control in many countries<br />

nor will it be economically competitive with herbicides. However, allelopathy can fit<br />

in an integrated weed management strategy very well as a vital component. This<br />

approach could reduce the sole dependence on synthetic herbicides for solving many<br />

complex weed problems. The examples discussed herein include an aggressive rice<br />

cultivar for complimenting weed control in direct seeded rice and plant productreinforced<br />

classical bio-control (through weevils) of water hyacinth.


Disclaimer<br />

ALLELOPATHY FOR WEED CONTROL 119<br />

Mention of trade names or commercial products in this publication is solely for the<br />

purpose of providing specific information and does not imply recommendation or<br />

endorsement by the United States Department of Agriculture.<br />

8. REFERENCES<br />

Abbas H.K., Tanaka T., Duke S.O., Boyette C.D. Susceptibility of various crop and weed species to AAL-toxin,<br />

a natural herbicide. Weed Technol 1995; 9:125-130.<br />

Ahn J.K., Chung I.M. Allelopathic potential of rice hulls on germination and seedling growth of barnyardgrass.<br />

Agron J 2000; 92:1162-1167.<br />

Akobundu I.O. Getting weed management technologies to farmers in the developing world. Proceedings of III<br />

International Weed Science Congress, Brazil, CD-ROM. International Weed Science Society, Oxford,<br />

MS, USA 2000; 2 p.<br />

Akobundu I.O., Agyakwa C.W. A Handbook of West African Weeds. 2 nd<br />

ed. Ibadan, Nigeria: International<br />

Institute of Tropical Agriculture 1998; 496 p.<br />

Arulchezian M.P., Kathiresan R.M. Effect of organic manures and herbicides on transplanted rice cv. ADT.37.<br />

In Proceedings 3 rd<br />

Tropical Weed Science Conference, Kuala Lumpur, Malaysia, 1990; pp. 313-317.<br />

Babbitt B. Statement by Secretary of the Interior on invasive alien species. National Weed Symposium. Bureau<br />

of Land Management Denver, CO. 1998; pp. 8-10.<br />

Barreto R., Charudattan R., Pomella A., Hanada R. <strong>Biologica</strong>l control of neotropical aquatic weeds with fungi<br />

Crop Prot 2000; 19:697-703.<br />

Barreto R.W., Evans H.C. Fungal pathogens of some Brazilian aquatic weeds and their potential use in biocontrol.<br />

In Proceedings of the IX International Symposium on <strong>Biologica</strong>l Control of Weeds. V.C. Moran, J. H.<br />

Hoffmann, eds. University of Cape Town, Cape Town, South Africa, 1996; pp. 121-126.<br />

Barreto R.W., Torres A.N.L. Nimbya alternantherae and Cercospora alternantherae: two new records of fungal<br />

pathogens on Alternanthera philoxeroides (alligator weed) in Brazil. Aust Plant Pathol 1999; 28:103-<br />

107.<br />

Bhowmik P.C., Doll J.D. Allelopathic effects of annual weed residues on growth and nutrient uptake of corn and<br />

soybeans. Agron J 1984; 76:383-388.<br />

Bond W. Non-chemical weed management. In Weed Management Handbook. E.L.Robert Nayler, ed. BCPC<br />

Black Well Publishers, UK. 2002.<br />

Bryson C.T., Carter R. Cogongrass, Imperata cylindrica, in the United States. Weed Technol 1993; 7:1005-<br />

1009.<br />

Byrd J.D., Bryson C.T. Biology, ecology, and control of cogongrass [Imperata cylindrica (L.) Beauv.]. Mississippi<br />

Dept. Agric. and Commerce, Bureau of Plant Industry, Mississippi State, Mississippi, USA. Fact Sheet<br />

1999-01. 2 p.<br />

Center T.D., Frank H.J., Dray F.A. <strong>Biologica</strong>l control. In Strangers in Paradise. D. Simberloff, D.C. Schmitz,<br />

T.C. Brown, eds. Island Press, Washington DC, 1997. pp 245-266.<br />

Charudattan R. Pathogens of <strong>Biologica</strong>l control of water hyacinth In Strategies for water hyacinth control. R.E.<br />

Charudattan, R. Labrada, T.D. Center, C. Kelly–Begazo. eds. FAO, Rome, 1996; pp. 189-196.<br />

Charudattan R., Dinoor A. <strong>Biologica</strong>l control of weeds using plant pathogens: accomplishments and limitations.<br />

Crop Prot 2000; 19:691-695.<br />

Charudattan R., Lin Y. Isolates of Penicillium, Aspergillus and Trichoderema toxic to aquatic plants. Hyacinth<br />

Control J 1974; 12:70-73.<br />

Cheng H.H. A conceptual framework for assessing allelochemicals in the soil environment. In Allelopathy:<br />

Basic and Applied Aspects. S.J.H. Rizvi, V. Rizvi, eds. Chapman and Hall Publishers, 1992; pp. 21-29.<br />

Chou C.H. The role of allelopathy in agro-ecosystem: studies from tropical Taiwan. In: Gliessman, S.R. (ed).<br />

Agroecology: Researching the ecological basis for sustainable agriculture. Ecological studies #78. Springer<br />

– Verlag. Berlin., 1990; pp. 105-121.<br />

Chou C.H., Chang F.J., Oka H.I. Allelopathic potential of wild rice Oryza perennis. Taiwania 1991; 36:201-<br />

210.


120<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

Copping G.L. Herbicides discovery. In Weed Management Handbook. BCPC. Blackwell Publishers, UK. 2002.<br />

Courtois B., Olofsdotter M. Incorporating the allelopathy trait in upland rice breeding programs. In Allelopathy<br />

in Rice. M. Olofsdotter, ed. International Rice Research Institute, Manila, Philippines: 1998; pp. 57-68.<br />

Dickens R. Cogongrass in Alabama after sixty years. Weed Sci 1974: 22:177-179.<br />

Dickens R, Buchanan G.A. Old weed in a new home-that’s cogongrass. Highland Agri Res 1971; 18:2.<br />

Dilday R.H., Lin J., Yan W. Identification of allelopathy in the USDA-ARS rice germplasm collection. Aust J<br />

Exp Agric 1994; 34:907-910.<br />

Dilday R.H., Yan W.G., Moldenhauer K.A.K., Gravois K.A.. Allelopathic activity in rice for controlling major<br />

aquatic weeds. In Allelopathy in Rice. M. Olofsdotter, ed. International Rice Research Institute, Manila,<br />

Philippines, 1998; pp. 7-26.<br />

Duke S.O. Naturally occurring chemical compounds as Herbicides. Reviews of Weeds Science 1986; 2:15-44.<br />

Duke S.O., Romagni J.G., Dayan F.E. Natural products as source of new mechanisms of herbicidal action. Crop<br />

Prot 2000; 19:583-589.<br />

Duke S.O., Scheffler B.E., Dayan F.E., Weston L.A., Ota E. Strategies for using transgenes to produce allelopathic<br />

crops. Weed Technol 2001; 15:826-834.<br />

Elmore C.D. Weed survey-southern states. Research Report from Southern Weed Science Society 1986; 39:36-<br />

158.<br />

Eussen J.H.H, Wirjahardja S. Studies of an alang-alang (Imperata cylindrica (L.) Beauv.) vegetation. Biotrop<br />

Bull 1973; 6:1-24.<br />

Falvey J.L. Imperata cylindrica and animal production in southeastern Asia: A review of Tropical Grasslands<br />

1981; 15:52-56.<br />

Ferguson J.J., Rathinasabapathi B. Allelopathy : How plants suppress other plants. HS944 series Institute of<br />

food and agricultural sciences, University of Florida, online publication 2003; http://edis.ifas.ufl.edu<br />

Gnanavel I., Kathiresan R.M. Sustainable weed management in rice-rice cropping system. Indian J Weed Sci<br />

2002; 34:192-196.<br />

Gross E.M., Meyer H., Schilling G. Release and ecological impact of algicidal hydrolysable polyphenols in<br />

Myriophyllum spicatum. Phytochemistry 1996; 41:133-138.<br />

Hassan S.M., Aidy I.R., Bastawisi A.O. Allelopathic potential of rice varieties against major weeds in Egypt.<br />

Paper presented at annual meeting of Weed Science Society of America. Seattle, Washington, USA. 1995.<br />

Hoagland, R.E. Microbial allelochemiclas and pathogens as bioherbicidal agents. Weed Technol 2001; 15:835-<br />

857.<br />

Holm L.G., Plucknett D.L., Pancho J.V., Herberger J.P. The World’s Worst Weeds – Distribution and Biology.<br />

University Press of Hawaii, Honolulu. 1977.<br />

Inderjit. Plant phenolics in allelopathy. Bot Rev 1996; 62:186-202.<br />

Inderjit, Dakshini K.M.M. Investigations on some aspects of chemical ecology of cogongrass. Imperata cylindrica<br />

(L.) Beauv. J Chem Ecol 1991; 17:343-352.<br />

Jensen L.B., Courtois B., Shen L., Li Z., Olofsdotter M., Mauleon R.P. Location genes controlling rice allelopathic<br />

effects against barnyardgrass in upland rice. Agron J 2001; 93:21-26.<br />

Kannan C. Integrated control of water hyacinth (Eichhornia crassipes (Mart.) Solms). Ph.D. Dissertation.<br />

Annamalai University, Tamilnadu, India 2002.<br />

Kannan C., Kathiresan R.M. <strong>Biologica</strong>l control at different growth stages of water hyacinth. Proceedings of the<br />

1 st<br />

IOBC global working group meeting for the biological and integrated control of water hyacinth. 16-19<br />

November, Harare, Zimbabwe. M.P. Hill, M.H. Julien, T.D. Center, eds. 1998; pp. 87-90.<br />

Kathiresan R.M. Allelopathic control of water hyacinth. In Annual Report 1997 of Centre of Aquatic Plant<br />

Management, J.R. Newman, ed. UK. 1998; pp. 54-56.<br />

Kathiresan R.M. Allelopathic potential of naïve plants on water hyacinth. XIV International Plant Production<br />

Congress, Jerusalem, Israel 1999; Abstract, pp. 146.<br />

Kathiresan R.M. Allelopathic potential of native plants against water hyacinth. Crop Prot 2000; 19:705-708.<br />

Kathiresan, R.M. Weed Management in Rice-Blackgram Cropping System. Indian J Weed Sci 2002; 34:220-<br />

226.<br />

Kathiresan R.M. Integration of elements of farming system for sustainable weed and pests management in the<br />

tropics. IV International Weed Science Congress, Durban, South Africa. 2004a; Abstract, pp. 6.


ALLELOPATHY FOR WEED CONTROL 121<br />

Kathiresan, R.M. Integration of botanical herbicide Coleus amboinicus / aromaticus with insect biological<br />

control of water hyacinth. Completion report on research project funded from National Agricultural<br />

Technology Project – Indian Council of Agricultural Research, 2004b.<br />

Kathiresan R.M., Gurusamy. Allelopathic potential of paddy seeds and some plant products on Barnyardgrass.<br />

Biennial Conference of Indian Society of Weed Science. Annamalai University, Annamalainagar 1995;<br />

Abstract, pp.110.<br />

Kathiresan R.M., Kannan C. Allelopathy for the control of water hyacinth. First Meeting of Global Working<br />

Group on Integrated and <strong>Biologica</strong>l Control of Water Hyacinth. IOBC, Harare, Zimbabwe 1998; Abstract,<br />

pp. 26.<br />

Kauraw L.P., Bhan V.M. Efficacy of cassytha powder to water hyacinth and marigold to Parthenium population.<br />

Weed News 1994; 1:3-6.<br />

Kim K.U., Shin D.H. Rice allelopathy research in Korea. In Allelopathy in Rice. M. Olofsdotter, ed. International<br />

Rice Research Institute, Manila, Philippines, 1998; pp. 39-43.<br />

Kim K.U., Shin D.H. The importance of allelopathy in breeding new cultivars. In Weed Management for<br />

Developing Countries. Addendum I. Labrada, ed., FAO Plant Production and Protection Paper 2003;<br />

pp.195-210.<br />

Kim K.W., Kim K.U., Shin D.H., Lee I.J., Kim H.Y., Koh J.C., Nam S.H. Searching for allelochemicals from the<br />

allelopathic rice cultivar, Kouketsumochi. Korean J Weed Sci 2000; 20:197-207.<br />

Koch W, Grobbmann F, Weber A, Lutzeyer H.J, Akobundu I.O. Weeds as components of maize/cassava cropping<br />

systems. In Standortgemaesse landwirtschaft in West Africa, M. von Oppen, ed. Stuttgart, Germany:<br />

Universitaet Hohenheim,1990; pp. 219-244<br />

Koger C.H., Bryson C.T. Effect of cogongrass (Imperata cylindrica) extracts on germination and seedling<br />

growth of selected grass and broadleaf species. Weed Technol 2004; 18:236-242.<br />

Koger C.H., Bryson C.T, Byrd J.D. Response of selected grass and broadleaf species to cogongrass (Imperata<br />

cylindrica) residues. Weed Technol 2004; 18:353-357.<br />

Leu E., Liszkay A.K., Goussias C., Gross E.M. Polyphenolic allelochemicals from the aquatic angiosperm<br />

myriophyllum spicatum inhibit photosystem II. Plant Physiol 2002; 130:2011-2018.<br />

Lin W., Kim K.U., Liang K., Guo Y. Hybrid rice with allelopathy. In Proceedings of the International Workshop<br />

in Rice allelopathy, K.U. Kim, D.H. Shin, eds., Kyungpook National University, Taegu, Korea, 17-19<br />

August 2000. Institute of Agricultural Science and Technology, Kyungpook National University, Taegu,<br />

2000; pp. 49-56.<br />

Lydon, J., Duke, S.O. Inhibitors of glutamine biosynthesis. In Plant Amino Acids: Biochemistry and<br />

Biotechnolgy, B.K. Singh, ed. Marcel Dekker, New York. 1999. pp. 445-464.<br />

Mattice J., Lavy T., Skulman B., Dilday R. Searching for allelochemicals in rice that control duck salad. In<br />

Allelopathy in Rice. M. Olofsdotter, ed. International Rice Research Institute, Manila, Philippines, 1998;<br />

pp. 81-98.<br />

Muchovej J.J., Kushalappa A.C. Dreschslera leaf spot of Salvinia auriculata. Plant Dis Rep 1979; 63:154.<br />

[OTA] Office of Technology Assessment. Harmful non-indigenous species in the United States. Washington<br />

DC: Office of Technology Assessment, US Congress 1993.<br />

Pandey D.K. Phytotoxicity of sesquiterpene lactone parthenin on aquatic weeds. J Chem Ecol 1996a; 22:151-<br />

160.<br />

Pandey D.K. Relative toxicity on allelochemicals to aquatic weeds. Allelopathy J 1996b; 3:241-246.<br />

Pandey D.K., Kauraw L.P., Bhan V.M. Inhibitory effect of parthenium (Parthenium hysterophorus L.) residue<br />

on growth of water hyacinth (Eichhornia crassipes Mart. Solms.). J Chem Ecol 1993; 19:2663-2670.<br />

Parthiban C., Kathiresan R.M. Use of certain plant materials for weed management in transplanted rice. Indian<br />

J Weed Sci 2002; 34:187-191.<br />

Pimentel D., Lach L., Zuniga R., Morrison D. Environmental and economic costs of nonindigenous species in<br />

the United States. BioScience 2000. 50:53-65.<br />

Pimentel D. Handbook on pest management in Agriculture. CRC Press, Boca Raton FL 1991.<br />

Putnam A.R. Weed allelopathy. In Weed Physiology, Volume 1: Reproduction and Ecophysiology, S.O. Duke,<br />

ed., CRC Press Boca Raton FL 1985; pp.131-155.<br />

Putnam A.R., Tang C.S. Allelopathy: State of Science. In The Science of Allelopathy, A.R. Putnam, C.S. Tang,<br />

eds., Wiley, New York, 1986; pp. 1-19.


122<br />

RAMANATHAN KATHIRESAN, CLIFFORD H. KOGER , KRISHNA N. REDDY<br />

Rao O.P., Saxena A.K., Singh B.P. Allelopathic effects of certain agroforestry tree species on the germination of<br />

wheat, paddy and gram. Ann Forestry. 1994; 2:1.<br />

Reddy K.N. Effects of cereal and legume cover crop residues on weeds, yield, and net return in soybean (Glycine<br />

max). Weed Technol 2001; 15:660-668.<br />

Reddy K.N., Zablotowicz R.M., Locke M.A., Koger, C.H. Cover crop, tillage, and herbicide effects on weeds,<br />

soil properties, microbial populations, and soybean yield. Weed Sci 2003; 51:987-994.<br />

Rice E.L. Allelopathy: An overview. Allelochemical: Role in agriculture and forestry. American Chemical<br />

Society Symposium Series 330, 1987; pp. 8-22.<br />

Rice E.L., Allelopathy. 2 nd<br />

edition. Academic Press. 1984.<br />

Sivagurunathan, M., Sumithradevi G., Ramasamy K. Allelopathic compounds in Eucalyptus spp. Allelopathy<br />

J 1997; 4:313-320.<br />

Sustainable Agriculture Network. 1998. Managing cover crops profitably. Second edition. Handbook Series<br />

Book 3. Beltsville, MD.<br />

Teasdale J.R. Contribution of cover crops to weed management in sustainable agricultural systems. J Prod<br />

Agric 1996; 9:475-199.<br />

Teasdale J.R. Principles and practices of using cover crops in weed management systems. In Weed Management<br />

for Developing Countries. Addendum I. Labrada, ed., FAO Plant Production and Protection Paper 2003;<br />

pp. 169-178.<br />

Udensi E.U., Akobundu I.O, Ayeni A.O, Chikoye D. Management of cogongrass (Imperata cylindrica) with<br />

velvet bean (Mucuna pruriens var. utilis) and herbicides. Weed Technol 1999; 13:201-208.<br />

Vasquez E.A., Morallo-Rejesus B., Punzalan E.G., Kraus W. <strong>Biologica</strong>l studies on the essential oil of Coleus<br />

amboinicus. IXV International Plant Production Congress, Jerusalem, Israel 1999; Abstract. pp. 29.<br />

Willis R.J. The historical bases of the concept of allelopathy. J Hist of Bio 1985; 18:71-102.


ANTONY V. STURZ<br />

BACTERIAL ROOT ZONE COMMUNITIES,<br />

BENEFICIAL ALLELOPATHIES AND<br />

PLANT DISEASE CONTROL<br />

Prince Edward Island Department of Agriculture, Fisheries, Aquaculture<br />

and Forestry, P.O. Box 1600, Charlottetown, P.E.I., C1A 7N3, Canada.<br />

Tel: 902-368-5664, FAX: 902-368-5661<br />

E-mail:avsturz@gov.pe.ca<br />

Abstract. The release of root exudates from plants encourages the formation of beneficial bacterial communities<br />

in the root zone capable of generating secondary metabolites that improve plant health and crop yield. Metabolites<br />

with antibiotic or lytic action have been identified , while others are known to induce systemic disease resistance<br />

in the host plant, or interfere with the nutritional requirements of phytopathogens. However, despite existing<br />

positive relationships between bacterial communities and their plant hosts, man-made attempts at applying<br />

bacteria for biocontrol purposes have met with limited success. Inconsistent performance of biocontrol bacteria<br />

in the field may be due to the variable expression of genes involved in the biocontrol action, or simply the<br />

resistance of established soil communities to a sudden and inundative influx of adventive bacterial species or<br />

strains. Regardless of the inherent capacity of ‘naturally occurring’ soil microbial ecosystems to buffer<br />

anthropogenic interference, crop management systems are regularly used to distort agro-ecosystems through,<br />

for example, the use of tillage operations, alternate cropping systems, monoculture, crop rotation length, fertilizer<br />

and organic amendments, and various crop protection chemistries. The management of soil microbial communities<br />

for disease control appears to involve, in part, the creation of short term chaos in the microbial community<br />

through the application of such perturbation stresses. While hope remains that bacterial communities with<br />

biocontrol activity will one day be used as an adjunct to, or replacement for, agrichemical crop protectants,<br />

reliable biological controls that moderate pathogen attack remain elusive. In the interim, disease suppressive<br />

soils may be encouraged to form through the use of modest perturbation stresses that promote microflora<br />

species’ diversity and functionalities underpinning natural bioantagonism.<br />

1. INTRODUCTION<br />

In its broadest sense allelopathy has been defined as “...any process involving secondary<br />

metabolites produced by plants, microorganisms, viruses and fungi that influence the<br />

growth and development of biological systems...” (IAS, 1998). In the bacterial<br />

kingdom, the production of secondary metabolites (allelochemicals) can result in the<br />

development of mutualistic, beneficial or antagonistic relationships (Smith and<br />

Goodman, 1999; Boller, 1995) amongst bacteria, or between bacteria and other living<br />

organisms. As such, bacterial secondary metabolites can act at the trophic level, directly<br />

affecting nutrient uptake or metabolism, or at the informational level, being recognized<br />

as signals by appropriate chemoperception systems (Dusenbery, 1992; Boller, 1995).<br />

Consequently, allelopathy may be viewed as an ecological phenomenon (Romeo, 2000)<br />

123<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 123– 142.<br />

© 2006 Springer. Printed in the Netherlands.


124<br />

ANTONY V. STURZ<br />

capable of regulating ecosystem health and biodiversity (Wardle et al., 1997; Mallik,<br />

2000).<br />

Many secondary metabolites can act at both the trophic and the informational<br />

level, becoming attractants or repellents, toxins or growth stimulants, depending upon<br />

the microbial partnerships involved (Dusenbery, 1992). Accordingly, microbially<br />

produced allelochemicals have been reported to express themselves as lytic agents or<br />

enzymes (Fridlender et al., 1993; Jacobson et al., 1994; Glick et al., 1999), antibiotics<br />

(Lynch, 1976; Bender et al., 1999), siderophores (Buysens et al., 1994; Marschener<br />

and Crowley, 1997) auxins (Patten and Glick, 1996; Glickman et al., 1998; Glick et<br />

al., 1999), volatile compounds (Claydon et al., 1987; Bakker and Schippers, 1987)<br />

and phytotoxic substances (Hoagland and Cutler, 2000).<br />

An intimacy is seen to extend between plants and microbes, in as much as plants<br />

appear able to influence the composition of the microbial community around their<br />

root systems by leaking specific carbohydrates, carboxylic and amino acids (Grayston<br />

et al., 1998) into the root zone, as well as through the ‘carbon-loading’ that occurs as<br />

root cell material is sloughed off during root growth (Hawes and Brigham, 1992).<br />

Hawes et al., 1998). In turn, rhizobacteria appear able to induce root exudation<br />

responses in plants (Bolton et al., 1993; Merharg and Killham, 1995). The result is a<br />

circular allelopathic cascade initiated by plant root exudates that trigger a positive<br />

microbial ‘allelopathic feedback’ in which the final receptor organism is also the<br />

initiator.<br />

It is this allelochemical interaction amongst soil microbial communities and the<br />

way in which their relationships subsequently influence plant health and disease<br />

development that will form the emphasis of the present chapter.<br />

2. MICROBE-MICROBE INTERACTIONS<br />

2.1. Commensal Relationships, Protocooperative Assemblages and Plant Pathogens<br />

A large proportion of successful biocontrol events in the infection court can be attributed<br />

to the positive outcome of multiple allelopathic episodes governed by the interplay<br />

among bacterial populations. In the root zone - that region characterized as occurring<br />

outside the host plant, but within the sphere of influence of the root system - bacterial<br />

populations often form protocooperative assemblages (loose associations) that are<br />

mutually beneficial (though not obligatory). By contrast, commensal relationships<br />

where the microorganism (commensal) benefits, while the host (plant) is neither<br />

harmed nor helped, are a feature of the closer physical juxtaposition of bacteria and<br />

host plant, often marked by the sharing of the same food resource. This latter<br />

relationship can be a feature of bacteria inhabiting the root surface (rhizoplane), but<br />

can also be extended to include those bacterial colonists found within plants<br />

(endophytes)-specifically in the endoroot tissues. However, not withstanding the above,<br />

distinctions between commensal or protocooperative assemblages become equivocal


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

in the face of community antagonism to pathogen invasion, and should, perhaps, be<br />

viewed independently of any plant health benefits.<br />

While pathogen antagonism has been collectively termed biological control and<br />

defined by Baker (1987) as ‘... a decrease of inoculum or the disease-producing activity<br />

of a pathogen accomplished through one or more organisms, including the host plant...’,<br />

it can also be argued that antagonism, at the communal level, is not necessarily directed<br />

at phytopathogens specifically, but at any group of organisms ‘invading’ the trophic<br />

level of an established community. In this sense, interactions amongst autochthonous<br />

(established) consortia (functional groupings) of microflora can be classified as<br />

beneficial where they promote plant health or inhibit phytopathogen attack (Mukerji<br />

et al., 1999).<br />

Accordingly, any exochthonous latecomer (colonist or pathogen) is likely to<br />

provoke a negative response from an established community, since its arrival will<br />

upset any balance amongst the community members (Atlas, 1986). In this respect,<br />

the resilience of a soil microbial community, when expressed in terms of its ability to<br />

inhibit invasions (colonization) by ‘non-community’ species will, in part, define its<br />

stability.<br />

As such, exochthonous species, enter the niche environment by chance but cannot<br />

maintain themselves in an active condition (Cooke and Rayner 1984). By contrast,<br />

indigenous communities may be subdivided into the slow growing autochthonous<br />

groups and the fast growing transient zymogenous groups - the former surviving on<br />

refractory substrates, and so, for the most part, remaining constantly active, while the<br />

latter only become active when a suitable food resource presents itself, and so are<br />

otherwise quiescent (Cooke and Rayner 1984).<br />

2.2. ‘Self-awareness’ in Bacterial Communities<br />

It is believed that population density of a consortium component species mediates<br />

population function through ‘self-awareness’ mechanisms such as ‘quorum sensing’,<br />

that enable bacteria to communicate among and between species in a consortium<br />

(Miller and Bassler, 2001).<br />

The degree to which bacterial consortia behave as commensal, protocooperative<br />

or pathogenic assemblages is dictated (in part) by the component populations’ density,<br />

which will vary according to the prevailing abiotic conditions affecting secondary<br />

metabolite production - including those with antibiotic properties (Grimwood et al.,<br />

1989; Tateda et al., 2001). Key abiotic factors include, among others, pH, temperature,<br />

moisture, salinity, oxygen concentration and carbon availability (Duffy and Défago,<br />

1999; Gaballa et al., 1997; Gutterson et al., 1988; Nakata et al., 1999; Shanahan et<br />

al., 1992; Slininger and Jackson, 1992; Slininger and Sheawilbur, 1995)<br />

A population’s ability to identify ‘itself’ through the recognition of diffusible<br />

signaling molecules (autoinducers) - generally acylated homoserine lactones (acyl-<br />

HSLs) for gram-negative bacteria, and oligopeptides for gram-positive bacteria - elicits<br />

the modulation of gene expression, that can alter bacterial function in ways that may<br />

125


126<br />

ANTONY V. STURZ<br />

ultimately destabilize cooperative behaviour (Salmond et al., 1995; Albus et al., 1997;<br />

Surette and Bassler, 1998).<br />

Paradoxically, the ability to exchange low molecular weight diffusible signal<br />

molecules suggests that certain species specific traits are repressed to allow individual<br />

bacteria to form consortia, so enabling them to survive within a specific habitat until<br />

such time as it benefits that species population to dissolve the co-operative group.<br />

This event is often catalyzed by the build up of sufficient amounts of signal molecule<br />

to activate receptor proteins that trigger changes in gene expression.<br />

Quorum sensors in bacteria have been implicated in regulating a range of<br />

physiological activities including conjugal transfer (Piper et al., 1993), swarming<br />

responses (Givskov et al., 1997), sporulation (Dworkin and Kaiser, 1985; Hoch, 1995),<br />

biofilm formation (Davies et al., 1998), and, most importantly, the regulation of<br />

virulence genes that initiate extracellular polysaccharides, enzymes, surfactants and<br />

antibiotics prior to and during pathogen attack (Beck von Bodman and Farrand, 1995;<br />

Fuqua, et al., 2001; Parsek et al., 1999; Whithers et al., 2001).<br />

Thus, for example, the causal agent of potato soft rot, Erwinia carotovora, will<br />

only initiate exoenzyme secretion - involving cellulases and various pectinases - when<br />

cell density levels reach a certain threshold (Pirhonen et al., 1993); presumably to<br />

ensure that invading bacteria do not prematurely elicit a host-defence response prior<br />

to achieving sufficient bacterial cell numbers to mount a successful infection.<br />

Interestingly, E. carotovora exoenzymes are produced in tandem with the antibiotic<br />

carbapenem, which it is believed inhibits other competitor bacteria in the infection<br />

court (Bainton et al., 1992). Consequently, it appears that some pathogen’s attack<br />

and self-defence measures are co-ordinated during attempts at plant infection.<br />

2.3. Community Self-regulation<br />

The biological basis of community homeostasis is generally believed to result from<br />

the dynamic balance which member populations in the community must exert to<br />

inhibit the recognition of ‘population self’ at the expense of the ‘community self’.<br />

Where complex regulatory cascades control gene expression of colonization and<br />

pathogenicity, gene induction is generally modulated by the rate at which a population’s<br />

signal molecule strength is diluted away. Accordingly, any situation that allows for<br />

signal molecule build-up - such as the close proximity of cells, or an environment<br />

limited diffusion rate - will result in the greater the likelihood that population selfrecognition<br />

and gene expression will be triggered.<br />

To-date, compounds reported to regulate population density dependent behaviours<br />

include N-acyl-homoserine lactones, among plant associated Proteobacteria (Eberl,<br />

1999; Fuqua et al., 2001), ã-butryolactone in the Streptomyces (Yamada and Nihira,<br />

1998), oligopeptides in a variety of gram-positive species (Dunny and Winans, 1999;<br />

Kleerebezem and Quadri, 2001), cyclic dipeptides in some gram-negative species<br />

(Holden et al., 1999) and fatty acid and butyrolactone derivatives in the plant pathogenic<br />

bacteria Xanthomonas campestris and Ralstonia solanacearum (Barber et al., 1997).


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

2.4. Microbe-microbe Disruption of Quorum Sensing Mechanisms<br />

It should now be apparent that quorum sensing plays a significant role in the biology<br />

and regulation of both plant-microbe and microbe-microbe interactions. And while<br />

pathogenic bacteria depend significantly on quorum sensing regulation to coordinate<br />

the saprophytic and parasitic phases of their life cycles, plants and their adherent root<br />

zone microbial communities have evolved mechanisms by which to disrupt this strategy.<br />

For example, Variovorax paradoxus, a relatively common soil organism, is able<br />

to utilize (degrade) acyl-HSL signaling molecules as an energy source (Leadbetter<br />

and Greenberg, 2000) with the effect that those classes of bacteria relying on acyl-<br />

HSLs for cell-to-cell signaling molecules will be kept ‘unaware’ of their own presence<br />

and population density. Pierson et al. (1998) demonstrated that approximately 8% of<br />

rhizobacteria recovered at random from the surfaces of wheat roots could specifically<br />

stimulate quorum sensing gene regulated expression in adjacent P. aureofaciens<br />

bacteria. Thus different bacteria appear able to exchange quorum sensing signals,<br />

with the possibility of forming functional mixed communities (Bauer and Robinson,<br />

2002).<br />

Several instances have been reported of soil bacteria in possession of enzymes<br />

designed to degrade or inactivate acyl-HSLs (Bauer and Robinson, 2002; Dong et al.,<br />

2001, 2002; Leadbetter, 2001; Whitehead et al., 2001). A case in point is the lactonase<br />

enzyme, AiiA, from Bacillus cereus, which, it is believed, opens the lactone ring in<br />

acyl-HSLs, thereby reducing signal strength in the order of a 1000 fold (Dong et al.,<br />

2000). In circumstances where B. cereus and E. carotovora co-exist as commensals<br />

in field soils, AiiA is able to inactivate the acyl-HSL autoinducer in E. carotovora<br />

rendering the pathogen avirulent.<br />

Clearly, understanding the mechanism of signal synthesis, and being able to<br />

identify signal synthesis inhibitors, has implications for developing quorum sensingtargeted<br />

antivirulence molecules, or engineering beneficial communities which utilize<br />

acyl-HSL signals as an energy source and so inhibit pathogenic trait expression. In<br />

the former case, AHL signal-inactivating molecules (the so-called ‘quorum quashing’<br />

moieties) - namely AHL-lactonases and AHL-acylases - have already been identified<br />

in a Ralstonia sp. isolated from a mixed-species biofilm (Lin et al., 2003).<br />

However, while selection pressures upon component bacterial populations in a<br />

community will include biological pressures created by microbially mediated<br />

moderation of habitats, external abiotic pressures created by environmental<br />

perturbations - such as climate or crop management practices - also exist. As a result,<br />

even though “... microorganisms are potentially everywhere, [the] environment<br />

selects...” (Alexander, 1997).<br />

3. MICROBIAL ANTAGONISM AND DISEASE CONTROL<br />

3.1. Microbially Induced <strong>Biologica</strong>l Control in Soils<br />

Since every living soil sample will yield organisms with antagonistic activity to some<br />

other organism, or group of organisms, it has almost become axiomatic that<br />

127


128<br />

ANTONY V. STURZ<br />

“...antagonistic potential resides in every soil microorganism...” (Baker and Cook,<br />

1974). Consequently, it is generally held that most soils possess the biological<br />

propensity to inhibit or reduce their soil microflora’s tendency toward disease, and so<br />

can be considered disease suppressive to some extent (Hornby, 1983). As a result,<br />

there exists in the literature a vast array of a terms used to describe soils that are<br />

inhospitable to plant pathogens. For example, i) soils where plant pathogens fail to<br />

become established have been referred to as resistant (Walker and Snyder, 1933),<br />

long-life, immune, intolerant, or antagonistic (Baker and Cook, 1974; Huber and<br />

Schneider, 1982), ii) soils where pathogens become established but fail to produce<br />

disease have also been termed suppressive (Schroth and Hancock, 1982); while iii)<br />

soils where disease incidence diminishes with continued monoculture have been termed<br />

decline soils (Shipton, 1975; Hornby, 1979, 1983).<br />

Attempts to simplify the biological basis for disease suppression in agricultural<br />

soils have reduced this concept to two broad mechanisms; namely that of i) a “general<br />

suppression” based upon the activity of the total microbial biomass that is not<br />

transferable between soils, and ii) a “specific suppression” that depends upon the<br />

activity of specific groups of microorganisms (Weller et al., 2002).<br />

Whether a bacterial population behaves pathogenically or not will be a function<br />

of that component species’ genetics, the restraints which other members in the<br />

community are able to impose, and the result of any overriding selection pressures<br />

dictated by environmental factors governing habitat type and host predisposition to<br />

disease.<br />

Environmentally mediated host predisposition to disease have been linked with<br />

obligate pathogen performance, and included exposure to cold (Schulz and Bateman,<br />

1969), low light intensity, or short day lengths (Foster and Walker, 1947), salinity<br />

stress (MacDonald, 1982), high temperature (Edmunds, 1964), and drought or moisture<br />

stress (Boyer, 1995; Duniway, 1977).<br />

In contrast, factors predisposing host plants to attack by rogue members of a<br />

commensal community, or protocooperative assemblage, are less well understood,<br />

though it appears likely that any dramatic change in the niche environment can provide<br />

an ecological advantage that benefits some community members at the expense of<br />

others. During the resulting population increase of the favoured community population,<br />

cell density dependent pathogenesis is triggered.<br />

3.2. Disease Suppression and Pathogen Evasion<br />

The wide array of nomenclature used to describe disease suppression in agricultural<br />

soils, is matched by an equally wide variety of individual microbial mechanisms<br />

postulated to explain these phenomena. However, it should be noted that these<br />

mechanisms are fairly presumptive, and, if they occur in vivo, are likely to operate in<br />

parallel with each other (Figure 1).<br />

In general, microbial biocontrol mechanisms have been classified according to<br />

effect (Baker, 1968) and have included such actions as parasitism/predation, niche<br />

competition, antibiosis and systemic induced resistance - the latter three falling


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

Induced Systemic<br />

Resistance<br />

ROOT – SOIL<br />

INTERFACE<br />

Niche competition/<br />

exclusion<br />

Lytic enzyme<br />

production<br />

Microbe-microbe disruption<br />

Quorum sensing and<br />

community self-awareness<br />

Root camouflage<br />

Direct antibiosis action<br />

ENDOROOT EXOROOT<br />

Figure 1. Parallel action of disease suppression mechanisms operating within the host plant<br />

(endoroot) and in the surrounding soil (exoroot).<br />

reasonably comfortably within the ambit of allelopathy (see Keel and Défago, 1997;<br />

Mukerji et al., 1999; Weller, 1988).<br />

Five common mechanisms are usually cited, namely:<br />

i) Resource(niche) competition (or exploitation) (O’Sullivan and O’Gara, 1992;<br />

Stephens et al., 1993; Whipps, 2001) - for example, siderophore (chelator)producing<br />

bacteria with high affinities for, and capable of sequestering, specific<br />

mineral elements, can inhibit phytopathogens with the same requirement if that<br />

mineral is limited in the soil (Schroth and Hancock, 1982; Dowling et al, 1996;<br />

Loper and Henkels, 1997, 1999).<br />

ii) Antibiosis action - through the production of specific or non-specific microbial<br />

metabolites with antibacterial, antifungal and anti-nematode activity (Levy and<br />

Carmelli, 1995; Fujimoto et al., 1995; Raaijmakers et al., 2002; Thomashow et<br />

al., 1997). To-date, several antibiotic substances have been identified of which<br />

those produced by the pseudomonads have been particularly well characterized.<br />

Of these, antibiotics identified with biocontrol properties include the<br />

phloroglucinols, phenazine derivatives, pyoluteorin, pyrrolnitrin, cyclic<br />

129<br />

c


130<br />

ANTONY V. STURZ<br />

lipopeptides and hydrogen cyanide (Haas and Keel, 2003). Among the other<br />

antibiotics characterized are agrocin 84 (Agrobacterium sp.), herbicolin A<br />

(Erwinia sp.), iturin A, surfactin, and zwittermicin A (Bacillus sp.) and<br />

xanthobacin (Stenotrophomonas sp.) (Hashidoko et al., 1999; He et al., 1994;<br />

Sayre and Starr, 1988; Thomashow et al., 1997; Silo-Suh et al., 1994). A comprehensive<br />

account of bacterially produced antibiotics may be found in Raaijmakers<br />

et al. (2002).<br />

iii) Lytic enzyme action - a feature of several bacteria with proven biocontrol ability,<br />

and generally involves the direct degradation of pathogen cell wall material, or<br />

the disruption of a particular developmental stage. Thus, for example, chitinase<br />

production by Serratia plymuthica has been reported to inhibit spore germination<br />

and germ-tube elongation in Botrytis cinerea (Frankowski et al., 2001), while ß-<br />

1,3-glucanase synthesized by Paenibacillus sp. and Streptomyces sp. can lyse<br />

fungal cell walls of Fusarium oxysporum f. sp. cucumerinum (Singh et al., 1999).<br />

Other enzymes produced by bacteria with biocontrol activity include hydrolase<br />

(Chernin and Chet, 2002), laminarinase (Lim et al., 1991) and protease (Kamensky<br />

et al., 2003).<br />

iv) Induced systemic resistance (ISR) in plants (Wei et al., 1991; Tuzun and Kloepper,<br />

1994) - whereby non-pathogenic rhizobacterial stimulation of defence-related<br />

genes is elicited through the encoded production of jasmonate (van Wees et al.,<br />

1999), peroxidase (Jetiyanon et al., 1997) or enzymes involved in the synthesis<br />

of phytoalexins (van Peer et al., 1991). Though no specific ISR-eliciting signal<br />

has been identified, thus far, evidence for the involvement of lipopolysaccharides,<br />

siderophores and phloroglucinols has been submitted (Hoffland, et al., 1995;<br />

Leeman et al., 1995, 1996; Maurhofer et al., 1994; van Wees et al., 1997), and,<br />

v) Root camouflage (Gilbert et al., 1994) - proposed as a mechanism to explain the<br />

observation that certain rhizobacterial populations in disease resistant cultivars<br />

are able to minimize the ‘attractive’ nature of the host’s root system so masking<br />

its presence to potential plant pathogens by restricting local population density<br />

development. Such microbial systems may operate in tandem with those that<br />

desensitize the chemoperception systems of microorganisms in the root zone,<br />

through the over production of chemical stimulii (Armitage, 1992; Dusenbery,<br />

1992).<br />

The parallel operation of all these biocontrol mechanisms in a four dimensional<br />

soil space makes their action and interaction difficult to follow. Biocontrol strains<br />

only occupy a small fraction of the root surface, in microcolonies spread out unevenly<br />

along the root surface (Bowen and Rovira, 1976; Normander et al., 1999). Disease<br />

suppression, when it occurs through antibiosis, is most likely restricted to local action<br />

only, and most probably at sub-inhibitory levels. Even so, antibiotics can cause intense<br />

physiological effects upon neighbouring organisms at subinhibitory concentrations.<br />

Quinolone and macrolide antibiotics have been reported to block cell- to cell signaling,<br />

and the production of virulence factors in P. aeruginosa (Grimwood et al., 1989;<br />

Tateda et al., 2001). Similarly, subinhibitory concentrations of antibiotics can suppress<br />

adherence mechanisms in bacteria (Breines and Burnham, 1994), and the production


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

of extracellular virulence factors in bacteria (Herbert et al., 2001). Accordingly,<br />

secondary metabolites can impact soil microbial ecosystems in a variety of ways, and<br />

at a variety of levels (Haas and Keel, 2003).<br />

3.3. Partitioning of disease suppressive bacteria in the endo and exoroot<br />

Most research into soil bacterial communities has been restricted to the exoroot - that<br />

fraction of the microfloral community found in the rhizoplane, plant rhizosphere or<br />

root zone soil (Haas and Keel, 2003). Endoroot bacteria have largely been ignored,<br />

despite plant-bacteria interactions extending into the endoroot of all plants (Conn et<br />

al., 1997; Bensalim et al., 1998). The frequent recovery of communities of endophytic<br />

bacteria, in the absence of any pathological condition (Chanway, 1996) and the finding<br />

that bacterial endoplant communities are capable of mediating against phytopathogen<br />

invasion (Benhamou et al., 1996) has led to the suggestion that plants may have coopted<br />

bacteria as part of a disease suppressive response to phytopathogen attack.<br />

Several instances have been reported of endophytic bacteria as effective biocontrol<br />

agents (van Buren et al., 1993; Brooks et al., 1994). van Peer et al. (1990) found that<br />

endo- and exoroot bacteria from the same genera formed discrete sub-populations,<br />

each suited to colonizing their respective environmental niches. Tissue-specific<br />

relationships can form between communities of bacterial endophytes and their host<br />

plant, and endobacteria have been shown to adapt functionally to certain tissue sites<br />

and among certain tissue types (Sturz et al., 1999). Unfortunately, the population<br />

densities of endophytic bacteria tend to be highly variable among plant tissues and so<br />

may be of little practical value in terms of affording plants a comprehensive first line<br />

of defence to pathogen attack.<br />

4. ‘ENGINEERING’ BENEFICIAL MICROBIAL ALLELOPATHIES<br />

4.1. Anthropogenic intrusions<br />

The premise underlying most anthropogenic biocontrol systems is the notion that it is<br />

possible to encourage the occurrence and development of beneficial rhizobacterial<br />

allelopathies in the root zone, primarily through the direct application of specific<br />

biocontrol agents, and or soil conditioning amendments (Sturz and Christie, 2003).<br />

It must be acknowledged at the outset that such systems have had varying degrees of<br />

success.<br />

4.2. Inundative approaches in biological control<br />

In general, anthropogenic attempts to import biological control agents into the field<br />

have been through the inundative application of super quantities of a few key biocontrol<br />

or plant growth promoting agents. These have been applied directly as drenches or<br />

sprays, or alternatively as a cell suspensions incorporated in mulches or compost<br />

131


132<br />

ANTONY V. STURZ<br />

material. Carriers such as granular peat formulations, mineral soils (Chao and<br />

Alexander, 1984), bacterial encapsulations within polymer gels (Bashan, 1986) or in<br />

natural gum or talc mixtures (Kloepper and Schroth, 1981) have also been tried.<br />

Perhaps, in hindsight, it is not surprising that such inundative approaches have<br />

been relatively unsuccessful, with most biocontrol agents failing to fulfill their initial<br />

promise. Such failures have usually been attributed to poor competence of biocontrol<br />

agents in the infection court and the difficulties associated with the instability of<br />

biocontrol agents in culture (Schroth et al., 1984; Weller, 1988); not least because in<br />

the case of bacterial agents, the accumulation of extracellular signalling molecules<br />

within large population densities of individual strains, can modulate a diverse range<br />

of metabolic processes, some of which are incompatible with the goals of<br />

phytoprotection (see above).<br />

However, the use of single antagonists may itself be an inappropriate strategy,<br />

arising from the belief that a given plant disease can be attributed to a single pathogen<br />

only (Baker, 1987). Such one-on-one syndrome concepts follow from the belief that<br />

successful control of a pathogen is achievable with a single fungicide, or by singlefactor<br />

resistance, coupled with the observation that single antagonists have often<br />

provided effective control in presumptive tests for antagonists in in vitro studies, or as<br />

biocontrol agents applied to sterilized soil (Baker, 1987)<br />

Needless to say, the very practice of applying massive quantities of a single bacterial<br />

species to the infection court will not only alter the putative biocontrol agent’s<br />

physiology, but also its niche behaviour. Perhaps, more intriguingly, it may also<br />

garner a general and antagonistic response from the resident population.<br />

Several attempts at engineering bacterial strains with more reliable biocontrol<br />

performance have been tried, whereby biosynthetic genes for various antibiotics have<br />

been designed to be constitutively over-expressed. On occasion, such engineered<br />

strains have provided improved plant protection in the soil microcosm (Delany et al.,<br />

2001; Ligon, et al., 2000; Timms-Wilson et al., 2000). However, in long term field<br />

evaluations, engineered derivatives have also lacked consistency; loss of stable<br />

performance and lack of superiority to wild-type strains being cited as the principal<br />

reasons for failure (Bakker et al., 2002).<br />

Although it would be premature to generalize the findings of such studies, it<br />

appears that the engineering of a single trait (antibiotic production) in a single<br />

biocontrol strain can not overcome the problem of inconsistent performance in the<br />

field, given the multi-factor nature of biocontrol mechanisms and the potential for<br />

interaction with wild-type species in the soil microbial community.<br />

While one-on-one antagonism may indeed be the sole operating mechanism<br />

involved in microbial disease suppression, an equally valid interpretation might be<br />

that the inundative addition of biocontrol agents can stimulate a general antagonistic<br />

response from the autochthonous microbial population to the ‘invader’ (inundating)<br />

species. In this circumstance and irrespective of any inconsistencies in the field<br />

performance of biocontrol agents attributed to unfavourable edaphic factors - such as<br />

temperature, soil moisture, pH, clay content, soil type - biocontrol success or failure<br />

may simply be due to the resident community’s antagonistic response following the


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

inundative insertion of a non-indigenous species. Consequently, both pathogen and<br />

biocontrol agent are inhibited in collateral fashion and to various degrees; a scenario<br />

that is congruent with the defensive mutualism theory proposed by Clay (1988).<br />

4.3. Modifying Soil Agro-ecosystems<br />

The extent to which producers can develop beneficial root zone allelopathies amongst<br />

microbial communities will depend largely upon the resilience of the soil in question<br />

(Szabolcs, 1994) and the type of crop management and tillage systems being practised<br />

(Sturz and Christie, 2003). Plant species are known to apply a selective and specific<br />

influence on microbial diversity in the rhizosphere through their differential root<br />

exudate spectra (Grayston et al., 1998), and the plastic nature of the relationship<br />

between resident microbial communities. Thus the level of disease suppressiveness in<br />

a soil is eminently amenable to deformation through the use of selected cultural<br />

practices. This regardless of the inherent capacity of ‘natural’ soil microbial ecosystems<br />

to buffer anthropogenic interference.<br />

Crop management systems are regularly used to distort agro-ecosystems through,<br />

for example, the use of tillage operations, alternate cropping systems, monoculture,<br />

crop rotation length, fertilizer and organic amendments, and various crop protection<br />

chemistries. The management of soil microbial communities for crop yield<br />

maximization appears to involve, in part, the creation of short term chaos in the<br />

microbial community through the application of a plethora of perturbation stresses<br />

(Odum et al., 1979). Moderate levels of ‘input perturbation’ are considered to improve<br />

ecosystem performance, while higher levels of perturbation stress result in performance<br />

loss.<br />

Input perturbations have commonly been used to modify soil microbial agroecosystems<br />

at the expense of pathogen populations. The subsequent variation in<br />

habitat and increase in niche heterogeneity - though on a microscale and at multiple<br />

sites along the root - is believed to encourage microbial biodiversity and consequently<br />

increase the potential for root zone competition (Smucker, 1993; Andrews and Harris,<br />

2000). Thus, for example, increasing soil acidity (Davis and Callihan, 1974, Sturz et<br />

al., 2003), applying irrigation soon after tuber initiation (Lapwood et al., 1973;<br />

Oestergaard and Nielsen, 1979) and the addition of soil amendments, green manures<br />

and mulches (Tremblay and Beauchamp, 1998) have all been relatively successful in<br />

reducing the development common scab on potatoes.<br />

Disease suppression has also been achieved against a wide range of pathogens by<br />

incorporating green manures (plough-down crops) (Tu and Findlay, 1986), animal<br />

manures (Gorodecki and Hadar, 1990) and composts (including organic solid wastes)<br />

(Nelson and Hoitink, 1983; Cohen et al., 1998) into field soils. All these amendments<br />

can encourage aggressive competition among microbial communities (Hoitink and<br />

Boehm, 1999; Hoitink and Fahy, 1986; Hoitink et al., 1997), with the added effect<br />

that manure and compost decomposition can release both volatile and non-volatile<br />

toxic compounds that inhibit phytopathogenic nematodes (Sayre et al., 1965; Abawi<br />

133


134<br />

ANTONY V. STURZ<br />

and Widmer, 2000) and reduce the survival rates of pathogenic microbes (De Brito et<br />

al., 1995; Chen et al., 1987 a, b).<br />

Though time consuming, unfashionable and often slow to show effect, traditional<br />

crop production practices that involve environmentally sustainable practices, such as<br />

conservation tillage (Sturz et al., 1997; Bockhus and Shroyer, 1998), ‘creative’<br />

fallowing options (Sturz et al., 2001), manuring (Hoitink and Boehm, 1999), long<br />

term crop rotations (Peters et al., 2003) and compatible cropping systems (Sturz et<br />

al., 2003), can yield plant health and crop yield benefits. Whether such knowledgebased,<br />

time-intensive management practices can be made more popular remains the<br />

challenge.<br />

5. CONCLUSION<br />

The close relationships formed between plant root systems and their respective<br />

rhizobacterial communities can lead to profoundly positive allelopathies that improve<br />

plant health and crop yield. The selective release of root exudates and plant leachates<br />

activates and sustains specific beneficial rhizobacterial communities in the root zone<br />

(endo- and exoroot). In turn these bacterial communities are able to generate a wide<br />

array of secondary metabolites which can improve plant health, either directly through<br />

biological control mechanisms, or by the modulation of phytopathogen populations<br />

in root zone. Simultaneously, certain root zone bacteria can induce systemic disease<br />

resistance in their plant host. Certain microbial communities are also able to interact<br />

functionally with each other through a variety of sensing mechanisms and gene<br />

expression triggers that are prompted by bacterial secondary metabolites with multiple<br />

activity.<br />

Despite the natural occurrence of strong positive relationships that can develop<br />

between bacteria and plants, anthropocentric attempts at applying bacteria for<br />

biocontrol purposes have had limited commercial success, notwithstanding advances<br />

in our understanding of the molecular mechanisms governing biocontrol interactions<br />

in the rhizosphere. The inconsistent performance of biocontrol bacteria in the field<br />

could be due to variable expression of genes involved in biocontrol, or merely the<br />

resistance of established soil communities to a sudden and inundative influx of<br />

adventive bacterial species or strains.<br />

While the hope remains that bacteria with biocontrol activity will one day be<br />

used as an adjunct to, or replacement for, commercial chemical fungicides, it will be<br />

necessary to better understand the influence on, within and between those microbial<br />

communities resident in the soil agro-ecosystems. In the interim, more traditional<br />

‘hit and miss’ methods for encouraging serendipitous beneficial allelopathies to arise<br />

in root zone communities - through, for example, the application of perturbation<br />

stresses - still have a role to play in modern systems of crop management, since they<br />

encourage the development of microflora species’ diversity and functionalities that<br />

underpin the antagonism systems promoting biological control in disease suppressive<br />

soils.


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

6. REFERENCES<br />

Abawi, G.S., Widmer, T.L. Impact of soil health management practices on soilborne pathogens, nematodes and<br />

root diseases of vegetable crops. Appl Soil Ecol 2000; 15:47-47.<br />

Albus, A.M., Pesci, E. C., Runyen-Janecky, L. J., West, S. E. H., Iglewski, B.H. Vfr controls quorum sensing in<br />

Pseudomonas aeruginosa. J Bacteriol 1997; 179: 3928-3935.<br />

Alexander, M. Microbial communities and interactions: a prelude. In Manual of Environmental Microbiology.<br />

Hurst, C.J., Knudsen, G.R., McInerney, M.J., Stetzenbach, L.D., Walter M.V., eds., ASM Press: Washington,<br />

1997; pp. 5-13.<br />

Andrews, J.H., Harris, R.F. The ecology and biogeography of microorganisms on plant surfaces. Annu Rev<br />

Phytopathol 2000; 38:145-180.<br />

Atlas, R.M. Applicability of general ecological principles to microbial ecology. In, Bacteria in Nature. Poindexter,<br />

J.S., Leadbetter, E.R., eds. Volume 2, Plenum Press: New York, 1986; pp. 339-370.<br />

Armitage, J.P. Behavioral responses in bacteria. Annu Rev Physiol 1992; 54:683-714<br />

Bainton, N.J., Stead, P., Chhabra, S.R., Bycroft, B.W., Salmond, G.P., Stewart G.S.A.B, Williams, P. N–(3oxohexanoyl)-L-homoserine<br />

lactone regulates carbapenem antibiotic production in Erwinia carotovora.<br />

Biochem J 1992; 288:997-1004.<br />

Baker, R. Mechanisms of biological control of soil-borne pathogens. Annu Rev Phytopathol 1968; 6:263-94.<br />

Baker, K.F., Cook, R.J. Examples of biological control. In, <strong>Biologica</strong>l Control of Plant Pathogens, Baker K.F,<br />

Cook, R.J. Freeman: San Francisco, 1974; pp. 61-106.<br />

Baker, K. Evolving Concepts of <strong>Biologica</strong>l Control of Plant Pathogens. Ann Rev Phytopathol 1987; 25: 67-85.<br />

Bakker, P.A.H.M., Glandorf, D.C.M., Viebahn, M., Ouwens, T.W.M., Smit, E., Leeflang, P., Wernars, K.,<br />

Thomashow, L.S., Laureijs, E., Thomas-Oates, J.E., Bakker, P.A.H.M.,Leendert C. van Loon, L.C. Effects<br />

of Pseudomonas putida modified to produce phenazine-1-carboxylic acid and 2,4-diacetylphloroglucinol<br />

on the microflora of field grown wheat. Antonie van Leeuwenhoek Int. J Gen Mol Microbiol 2002; 81:617-<br />

24.<br />

Bakker, A.W., Schippers, B. Microbial cyanide production in the rhizosphere in relation to potato yield reduction<br />

and Pseudomonas spp. - mediated plant growth-stimulation. Soil Biol Biochem 1987; 19: 451-57.<br />

Barber, C.E. Tang, J.L., Feng, J.X., Pan M.Q., Wilson, T.J. et al. A novel regulatory system for the pathogenicity<br />

of Xanthomonas campestris is mediated by a small diffusible signal molecule. Mol Microbiol 1997;<br />

24:555-566.<br />

Bashan, Y. Field dispersal of Pseudomonas syringae pv. tomato, Xanthomonas campestris pv. vesicatoria,<br />

and Alternaria macrospora by animals, people, birds, insects, mites, agricultural tools, aircraft, soil<br />

particles, and water sources. Can J Bot 1986; 64:76-281.<br />

Bauer, W.D., Robinson, J.B. Disruption of bacterial quorum sensing by other organisms. Curr Opin Biotechol<br />

2002; 13:234-237.<br />

Beck von Bodman, S., Farrand, S.K. Capsular polysaccharide biosynthesis and pathogenicity in Erwinia stewartii<br />

require induction by an N-acylhomoserine lactone autoinducer. J Bacteriol 1995; 177: 5000-5008.<br />

Bender, C.L., Rangaswamy, V., Loper J. Polyketide production by plant associated pseudomonads. Annu Rev<br />

Phytopathol 1999; 37:175-196.<br />

Benhamou, N., Kloepper, J.W., Quadt-Hallman, A., Tuzun, S. Induction of defence-related ultrastructural<br />

modifications in pea root tissues inoculated with endophytic bacteria. Plant Physiol 1996; 112: 919-929.<br />

Bensalim, S., Nowak, J., Asiedu, S. A plant growth promoting rhizobacterium and temperature effects on<br />

performance of 18 clones of potato. Am Pot J 1998; 75:145-152.<br />

Bockhus, W.W., Shroyer, J.P. The impact of reduced tillage on soilborne plant pathogens. Annu Rev Phytopathol<br />

1998; 36:485-500.<br />

Boller, T. Chemoperception of Microbial Signals in Plant Cells. Annu. Rev. Plant Physiol. Plant Mol. Biol<br />

1995; 46:189-214.<br />

Bolton, H.J., Frederickson, J.K. and Elliott, L.F. Microbial ecology of the rhizosphere. In Soil Microbial<br />

Ecology, Metting F.B.J. ed. Marcel Dekker: New York, 1993; pp. 27-63.<br />

Bowen, G.D., Rovira, A.D. Microbial colonization of plant roots. Annu Rev Phytopathol 1976; 14:121-44.<br />

Boyer, J.S. Biochemical and Biophysical Aspects of Water Deficits and the Predisposition to Disease. Annu Rev<br />

Phytopathol 1995; 33:251-74.<br />

135


136<br />

ANTONY V. STURZ<br />

Breines, D., Burnham, J. Modulation of Escherichia coli type 1 fimbrial expression and adherence to uroepithelial<br />

cells following exposure of logarithmic phase cells to quinolones at subinhibitory concentrations. J<br />

Antimicrob Chemother 1994; 34:205-21.<br />

Brooks, D.S., Gonzalez, C.F., Appel, D.N., Filer, T.H. Evaluation of endophytic bacteria as potential biological<br />

control agents for oak wilt. Biol Cont 1994; 4:373-381.<br />

Buysens, S., Poppe, J., Hofte, M. Role of siderophores in plant growth stimulation and antagonism by<br />

Pseudomonas aeruginosa 7NSK2. In, Improving plant productivity with rhizosphere bacteria. Ryder,<br />

M.H., Stephens P.M., Bowens G.D., eds. CSIRO: Adelaide, 1994; pp. 139-141.<br />

Chanway, C.P. Endophytes: they’re not just fungi! Can J Microbiol 1996; 74:321-322.<br />

Chao, W.L., Alexander, M., Mineral soils as carriers for Rhizobium inoculants. Appl Environ Microbiol 1984;<br />

47:94-97.<br />

Chen, W., Hoitink, H.A J., Schmittenner, A.F. Factors affecting suppression of Pythium damping-off in container<br />

media amended with composts. Phytopathology 1987a; 77:755-60.<br />

Chen, W., Hoitink, H.A.J., Tuovinen, O.H. The role of microbial activity in suppression of damping-off caused<br />

by Pythium ultimum. Phytopathology 1987; 78:314-22.<br />

Chernin, L., Chet I. Microbial enzymes in biocontrol of plant pathogens and pests. In, Enzymes in the<br />

Environment: Activity, Ecology, and Applications. Burns, R.G., Dick, R.P., eds. Marcel Dekker: New<br />

York, 2002; pp. 171-225.<br />

Clay, K. Fungal endophytes of grasses: a defensive mutualism between plants and fungi. Ecology 1988; 69:10-<br />

16.<br />

Claydon, N., Allan, M., Hanson, J. R., Avent, A.G., Antifungal alkyl pyrones of Trichoderma harzianum.<br />

Trans Br Mycol Soc 1987; 88:503-13.<br />

Cohen, R., Chefetz, B., Hadar, Y., Suppression of soil-borne pathogens by composted municipal solid waste. In,<br />

Beneficial co-utilization of agricultural, municipal and industrial bi-products. Brown, S., Angle, J.S.,<br />

Jacobs, L., eds. Kluwer Academic Publishers: Dordrecht, Netherlands, 1998, pp. 113-130.<br />

Conn, K.L., Nowak, J., Lazarovits, G. A gnotobiotic bioassay for studying interactions between potatoes and<br />

plant growth-promoting rhizobacteria. Can J Microbiol 1997; 43:801-808.<br />

Cooke, R.C., Rayner, A.D.M. Ecology of Saprotrophic Fungi. Longman Group Limited: London, 1984.<br />

Davies, D.G., Parsek, M.R., Pearson, J.P., Iglewski, B.H., Costerton, J.W., Greenberg, E.P. The involvement of<br />

cell-to cell signals in the development of a bacterial biofilm. Science 1998; 280:295-298.<br />

Davis, J.G., Callihan, R.H., Effects of gypsum, sulfur, terractor, and terractor super-x for potato scab control.<br />

Am Pot J 1974; 51:35-43.<br />

De Brito Alvarez, M.A., Gagne S., Antoun., H. Effect of compost on rhizosphere microflora of the tomato and<br />

on the incidence of plant growth-promoting rhizobacteria. App Envir Microbiol 1995; 61(1):194-199.<br />

Delany, I.R., Walsh, U.F., Ross, I., Fenton, A.M,, Corkery, D.M., O’Gara, F. Enhancing the biocontrol efficacy<br />

of Pseudomonas fluorescens F113 by altering the regulation and production of 2,4-diacetylphloroglucinol.<br />

Plant Soil 2001; 232:195-205.<br />

Dong, Y.H, Xu, J.L., Li, X.Z., Zhang, L.H. AiiA, an enzyme that inactivates the acyl-homoserine lactone quorum<br />

sensing signal and attenuates the virulence of Erwinia carotovora. Proc Natl Sci 2000; 97: 3526-3531.<br />

Dong, Y.H., Wang, L.H., Xu, J.L., Zhang, H.B., Zhang, X.F., Zhang, L.H. Quenching quorum sensing-dependent<br />

bacterial infection by an N-acyl homoserine lactonase. Nature 2001; 411:813-817.<br />

Dong, Y.H., Gusti, AR, Zhang, Q., Xu, J.L., Zhang, L.H. Identification of quorum-quenching N-acyl homoserine<br />

lactonases from Bacillus species. Appl Environ Microbiol 2002; 68:1754-1759.<br />

Dowling, D.N., Sexton, R., Fenton, A., Delany, I., Fedi, S., McHugh, B., Callanan, M., Moënne-Loccoz, Y.,<br />

O’Gara, F. Iron regulation in plant-associated Pseudomonas fluorescens M114: implications for biological<br />

control. In Molecular Biology of Pseudomonads. Nakazawa, K., Furukawa, K., Haas D. and Silver S.<br />

eds. ASM Press: Washington, 1996; pp. 502-511.<br />

Duffy, BK., Défago, G. Environmental factors modulating antibiotic and siderophore biosynthesis by<br />

Pseudomonas fluorescens biocontrol strains. Appl Environ Microbiol 1999; 65:2429-2438.<br />

Duniway, J.M. Predisposing effect of water stress on the severity of Phytophthora root rot of safflower.<br />

Phytopathology 1977; 67:884-89.<br />

Dunny, G.M., Winans, S.C. Bacterial life: neither dull nor boring. In, Cell-Cell Signaling in Bacteria. Dunny<br />

G.M., Winans S.C., eds., ASM Press: Washington, 1999; pp. 1-5.<br />

Dusenbery, D.B. Sensory Ecology: How Organisms Acquire and Respond to Information. Freeman:New York,<br />

1992.


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

Dworkin, M., Kaiser, D. Cell interactions in myxobacterial growth and development. Science 1985; 230: 18-<br />

24.<br />

Eberl, L. N-acyl homoserine lactone mediated gene regulation in gram-negative bacteria. Syst Appl Microbiol<br />

1999; 22:493-506.<br />

Edmunds, L.K. Combined relation of plant maturity, temperature and soil moisture to charcoal rot development<br />

in grain sorghum. Phytopathology 1964; 54:514-17.<br />

Foster, R.E., Walker, J.C. Predisposition of tomato to Fusarium wilt. J Agr Res 1947; 74:165-85.<br />

Fox, R.T., Manners, J.G., Myers, A. Ultrastructure of entry and spread of Erwinia carotovora var atroseptica<br />

into potato tubers. Potato Res 1971; 14:61-73.<br />

Frankowski, J., Lorito, M., Scala, F., Schmidt, R., Berg, G., Bahl, H. Purification and properties of two chitinolytic<br />

enzymes of Serratia plymuthica HRO-C48. Arch Microbiol 2001; 176:421-426.<br />

Fridlender, M., Inbar, J., Chet, I. <strong>Biologica</strong>l control of soilborne plant pathogens by a â-1,3 glucanase-producing<br />

Pseudomonas cepacia. Soil Biol Biochem 1993; 25:1211-1221.<br />

Fujimoto, D.K., Weller, D.M., Thomashow, L.S. Role of secondary metabolites in root disease suppression. In,<br />

Allelopathy: organisms, processes and applications. Inderjit, K.M.M. Dakshini, Einhellig F.A., eds.,<br />

ACS Symposium Series: Washington, D.C. 1995; pp. 330 - 347.<br />

Fuqua, C., Parsek, M.R., Greenberg, E.P.Regulation of gene expression by cell-to-cell communications: acylhomoserine<br />

lactone quorum sensing. Annu Rev Genet 2001; 35:439-468.<br />

Gaballa, A., Abeysinghe, P.D., Urich, G., Matthijs, S., Degreve, H. Trehalose induces antagonism towards<br />

Pythium debaryanum in Pseudomonas fluorescens ATCC 17400. Appl Environ Microbiol 1997;<br />

63:4340-4345<br />

Gilbert, G.S., Handelsman, J., Parke, J.L. Root camouflage and disease control. Phytopathology 1994; 84:222-<br />

225.<br />

Givskov, M., Eberl, L., Molin, S. Control of exoenzyme production, motility and cell differentiation in Serratia<br />

liquefaciens. FEMS Microbiol Lett 1997; 148:115-122.<br />

Glick, B.R., Patten, C.L., Holguin, G., Penrose, D.M. Auxin Production. In, Biochemical and genetic<br />

mechanisms used by plant growth promoting bacteria. Glick, B.R., Patten, C.L., Holguin, G., Penrose,<br />

D.M. eds. Imperial College Press: London, 1999; pp. 86 -133.<br />

Glickman, E., Gardan, L., Jaquet, S., Hussain, S., Elasri, M., Petit, A., Dessaux, Y. Auxin production is a<br />

common feature of most pathovars of Pseudomonas syringae. Mol Plant-Microbe Interact 1998; 11:156-<br />

162.<br />

Grayston, S.J., Wang, S., Campbell, C.D., Edwards, A.C. Selective influence of plant species on microbial<br />

diversity in the rhizosphere. Soil Biol Biochem 1998; 30:369-378.<br />

Grimwood, K., To, M., Rabin, H.R., Woods, D.E. Inhibition of Pseudomonas aeruginosa exoenzyme expression<br />

by subinhibitory antibiotic concentrations. Antimicrob Agents Chemother 1989; 33:41-47.<br />

Gorodecki, B., Hadar, Y. Suppression of Rhizoctonia solani and Sclerotium rolfsii in container media containing<br />

composted separated cattle manure and composted grape marc. Crop Prot 1990; 9:271-274.<br />

Gutterson, N., Ziegle, J.S., Warren, G.J., Layton, T.J. Genetic determinants for catabolite induction of antibiotic<br />

biosynthesis in Pseudomonas fluorescens HV37a. J Bacteriol 1988; 170:380-385.<br />

Haas, D., Keel, C. Regulation of antibiotic production in root-colonization Pseudomonas spp. and relevance for<br />

biocontrol of plant disease. Annu Rev Phytopathol 2003; 41:117-153.<br />

Hashidoko, Y., Nakayama, T., Homma, Y., Tahara, S. Structure elucidation of xanthobaccin A, a new antibiotic<br />

produced from Stenotrophomonas sp. strain SB-K88. Tetrahedron Lett 1999; 40:2957-2960.<br />

Hawes, M.C., Brigham, L.A., Wen, F., Woo, H.H., Zhu, Y. Function of root border cells in plant health: pioneers<br />

in the rhizosphere. Annu Rev Phytopathol 1998; 36:311-327.<br />

Hawes, M.C., Brigham, L.A. Impacts of root border cells on microbial populations in the rhizosphere. Adv<br />

Plant Pathol 1992; 8:119-148.<br />

He, H., Silo-Suh, L.A., Clardy, J., Handelsman, J. Zwittermycin A, an antifungal and plant protection agent<br />

from Bacillus cereus. Tetrahedron Lett 1994; 35:2499-502.<br />

Herbert, S., Barry, P., Novick, R.P. Subinhibitory clindamycin differentially inhibits transcription of exoprotein<br />

genes in Staphylococcus aureus. Infect Immun 2001; 69:2996-3003.<br />

Hinton, D.M., Bacon, C.W. Enterobacter cloacae is an endophytic symbiont of corn. Mycopathologia 1995;<br />

129:117-125.<br />

Hoagland, R.E., Cutler, S.J. Plant microbial compounds as herbicides. In, Allelopathy in Ecological Agriculture<br />

and Forestry, Narwal, S.S., Hoagland, R.E., Dilday, R.H., Reigosa, M.J. eds., (Proceedings of the III<br />

137


138<br />

ANTONY V. STURZ<br />

International Congress on Allelopathy in Ecological Agriculture and Forestry, Dharwad, India. 18-21<br />

August 1998). Kluwer Academic Publications: London, 2000; pp.73-99.<br />

Hoch, J.A. Control of cellular development in sporulating bacteria by the phosphorelay two-component signal<br />

transduction system. In, Two component signal transduction. Hoch, J.A., Silvay, T.J. eds., ASM Press:<br />

Washington, D.C. 1995; pp. 129-144.<br />

Hoffland, E., Pieterse, C., Bik, L., van Pelt, J.A. Induced systemic resistance in radish is not associated with<br />

accumulation of pathogenesis-related proteins. Physiol Mol Plant Pathol 1995; 46:309-320.<br />

Hoitink, H.A.J, Boehm, M.J. Biocontrol within the context of soil microbial communities: a<br />

substrate-dependent phenomenon. Annu Rev Phytopathol 1999; 37:427-446.<br />

Hoitink, H.A.J., Fahy, P.C. Basis for the control of soilborne plant pathogens with composts. Ann Rev Phytopathol<br />

1986; 24:93-114.<br />

Hoitink, H.A.J., Stone, A.G., Han, D.Y. Suppression of plant disease by composts. HortScience 1997; 32:184-<br />

197.<br />

Holden, M.T.G., Chhabra, S.R. de Nys, R., Stead, P., Bainton, N.J. , Hill, P.J., Manefield, M., Kumar, N.,<br />

Labatte, M., England, D., Rice, S., Givskov, M., Salmond, G.P.C., Stewart, G.S.A.B., Bycroft, B.W.,<br />

Williams, P. Quorum sensing cross-talk: isolation and chemical characterization of cyclic dipeptides from<br />

Pseudomonas aeruginosa and other gram-negative bacteria. Mol Microbiol 1999; 33:1254-1266.<br />

Hornby, D. Suppressive soils. Ann Rev Phytopathol 1983; 21:65-85.<br />

Hornby, D. Take-all decline: a theorist’s paradise. In, Soilborne plant pathogens. Schippers B., Gams W., eds.,<br />

Academic Press: New York, 1979; pp. 133-156.<br />

Huber, D.M., Schneider, R.W. The description and occurrence of suppressive soils. In, Suppressive Soils and<br />

Plant Disease. Schneider, R.W. ed. APS: St. Paul, Minnesota, 1982; pp. 1-7.<br />

International Allelopathy Society (IAS). Article 2, IAS Constitution (draft), Second Newsletter 1998, http://<br />

www2.uca.es/dept/quimica_organica/ias/newslett2.htm#draft International Allelopathy Society.<br />

Department of Biochemistry, Oklahoma State University, Stillwater, Oklahoma, USA, 1998.<br />

Jacobson, C.B., Pasternak, J.J., Glick, B.R. Partial purification and characterization of ACC deaminase from<br />

plant-growth promoting rhizobacterium Pseudomonas putida GR12-2. Can J Microbiol 1994; 40:1019-<br />

1025.<br />

Jetiyanon, K., Tuzun, S., Kloepper J. Lignification, peroxidase and superoxide dismutases as early plant defence<br />

reactions associated with PGPR-mediated induced systemic resistance. In, Plant growth-promoting<br />

rhizobacteria: present status and future prospects. Kobayashi, K., Homma, Y., Kodama, F., Kondo N.,<br />

Akino, S. eds. OECD: Paris, France, 1997; pp. 265-268.<br />

Kamensky , M., Ovadis, M., Chet, I., and Chernin, L. Soil-borne strain IC14 of Serratia plymuthica with multiple<br />

mechanisms of antifungal activity provides biocontrol of Botrytis cinerea and Sclerotinia sclerotiorum<br />

diseases. Soil Biol Biochem 2003; 35:323-331.<br />

Keel, C., Défago, G. Interactions between beneficial soil bacteria and root pathogens: mechanisms and ecological<br />

impact. In, Multitrophic Interactions in Terrestrial Systems. Gange A.C., Brown, V.K. eds. Blackwell<br />

Science Ltd.: London, UK, 1997; pp. 27-46.<br />

Kleerebezem, M., Quadri, L.E. Peptide pheromone-dependent regulation of antimicrobial peptide production in<br />

Gram-positive bacteria: a case of multicellular behaviour. Peptides 2001; 22:1579-1596.<br />

Kloepper, J.W., Schroth, M.N. Development of a powder formulation of rhizobacteria for inoculation of potato<br />

seed pieces. Phytopathology 1981; 71:590-92.<br />

Lapwood, D.H., Wellings, L.W., Hawkins, J.H. Irrigation as a practical means to control potato common scab<br />

(Streptomyces scabies): final experiment and conclusions. Plant Pathol 1973; 22:35-41.<br />

Leadbetter, J.R. Plant microbiology. Quieting the raucous crowd. Nature 2001; 411:748-749.<br />

Leadbetter, J.R., Greenberg, E.P. Metabolism of acyl-homoserine lactone quorum-sensing signals by Variovorax<br />

paradoxus. J Bacteriol 2000; 182:6921-6926.<br />

Leeman, M., Denouden, E.M., Vanpelt, J.A., Dirkx, F., Steijl, H., Bakker, P.A.H.M., Schippers, B. Iron availability<br />

affects induction of systemic resistance to fusarium wilt of radish by Pseudomonas fluorescens.<br />

Phytopathology 1996; 86:49-55.<br />

Leeman, M., Vanpelt, J.A., Denouden, F.M., Heinsbroek, M., Bakker, P.A.H.M, Schippers, B. Induction of<br />

systemic resistance against fusarium wilt of radish by lipopolysaccharides of Pseudomonas fluorescens.<br />

Phytopathology 1995; 85:1021-27.


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

Levy, E., Carmelli, S. <strong>Biologica</strong>l control of plant pathogens by antibiotic-producing bacteria. In, Allelopathy:<br />

organisms, processes and applications. Inderjit, K.M.M. Dakshini, Einhellig, F.A. eds., ACS Symposium<br />

Series: Washington, D.C., 1995, pp. 300-309.<br />

Ligon, J.M., Hill, D.S., Hammer, P.E., Torkewitz, N.R., Hofmann, D., Kempf, H.J., van Pée, K.H. Natural<br />

products with antifungal activity from Pseudomonas biocontrol bacteria. Pest Manage Sci 2000; 56:<br />

688-695.<br />

Lim, H.S., Kim, Y.S., Kim, S.D. Pseudomonas stutzeri YPL-1 genetic transformation and antifungal mechanism<br />

against Fusarium solani, an agent of plant root rot. Appl Environ Microbiol 1991; 57:510-516.<br />

Lin, Y.H., Xu, J.L., Hu, J., Wang, L.H., Ong, S.L., Leadbetter, J.R., Zhang, L.H. Acyl- homoserine lactone<br />

acylase from Ralstonia strain XJ12B represents a novel and potent class of quorum-quenching enzymes.<br />

Mol Microbiol 2003; 47:849-860.<br />

Lynch, J.M. Products of soil microorganisms in relation to plant growth. CRC Crit Rev Microbiol 1976; 5: 67-<br />

107.<br />

Loper, J.E., Henkels, M.D. Availability of iron to Pseudomonas fluorescens in rhizosphere and bulk soil evaluated<br />

with an ice nucleation reporter gene. Appl Environ Microbiol 1997; 63:99-105.<br />

Loper, J.E., Henkels, M.D. Utilisation of heterologous siderophores enhances levels of iron available to<br />

Pseudomonas putida in the rhizosphere. Appl Environ Microbiol 1999; 65:5357-5363.<br />

MacDonald, J.D. Effect of salinity stress on the development of Phytophthora root rot of chrysanthemum.<br />

Phytopathology 1982; 72:214-19.<br />

Mallik, A.U. Challenges and opportunities in allelopathy research: a brief overview. J Chem Ecol 2000; 26:2007-<br />

2009.<br />

Marschener, P., Crowley, D.E. Iron stress and pyoverdin production by a fluorescent pseudomonad in the<br />

rhizosphere of white lupin (Lupinus albus L.) and barley (Hordeum vulgare L.) Appl Environ Microbiol<br />

1997; 63:277-281.<br />

Maurhofer, M., Hase, C., Meuwley, P., Métraux, J.P., Défago, G. Induction of systemic resistance of tobacco to<br />

tobacco necrosis virus by the root-colonizing Pseudomonas fluorescens strain CHA0: influence of the<br />

gacA gene and of pyoverdine production. Phytopathology 1994; 84:139-46.<br />

Merharg, A.A., Killham, K. Loss of exudates from roots of perennial rye-grass inoculated with a range of<br />

microorganisms. Plant Soil 1995; 170:345-349.<br />

Miller, M.B., Bassler, B.L. Quorum sensing in bacteria. Annu Rev Microbiol 2001; 55:165-199.<br />

Misaghi, I.J., Donndelinger, C.R. Endophytic bacteria in symptom-free cotton plants. Phytopathology 1990;<br />

80:808-811.<br />

Mukerji, K.G., Chamola, B.P., Upadhyay, R.K. eds. Biotechnolocal approaches in Biocontrol of Plant Pathogens.<br />

Kluwer Academic/Plenum Publishers, New York, Dordrecht, London, 1999; 255 p.<br />

Nakata, K., Yoshimoto, A., Yamada, Y. Promotion of antibiotic production by high ethanol, high NaCl<br />

concentration, or heat shock in Pseudomonas fluorescens S272. Biosci Biotechnol Biochem 1999; 63:<br />

293-297.<br />

Nelson, E.B., Hoitink, H.A.J. The role of microorganisms in the suppression of Rhizoctonia solani in container<br />

media amended with composted hardwood bark. Phytopathology 1983; 73:274-78.<br />

Normander, B., Hendriksen, N.B., Nybroe, O. Green fluorescent protein-marked Pseudomonas fluorescens:<br />

localization, viability, and activity in the natural barley rhizosphere. Appl Environ Microbiol 1999; 65:4646-<br />

51.<br />

O’Sullivan, D.J., O’Gara, F. Traits of fluorescent Pseudomonas spp. involved in suppression of plant root<br />

pathogens. Microbiol Rev 1992; 56:662-676.<br />

Odum, E.P., Finn, J.T., Franz, E.H. Perturbation theory and subsidy-stress gradient. BioScience 1979; 29:349-<br />

352.<br />

Oestergaard, S.P., Nielsen, S. Control of potato scab (Streptomyces scabies) by irrigation. Tidsskrift for Planteavl<br />

1979; 83:201-204.<br />

Parmar, N. Dardarwal, K.R. Stimulation of nitrogen fixation and induction of flavonoid like compounds by<br />

rhizobacteria. J Appl Microbiol 1999; 86:36-44.<br />

Parsek, M.R., Val, D.L., Hanzelka, B.L., Cronan, J.E. Jr., Greenberg E.P. Acyl homoserine-lactone quorumsensing<br />

signal generation. Proc Natl Acad Sci USA. 1999; 13:4360-4365.<br />

Patten, C.L., Glick, B.R. Bacterial biosynthesis of indole-3-acetic acid. Can J Microbiol 1996; 42:207-220.<br />

139


140<br />

ANTONY V. STURZ<br />

Peters, R.D., Sturz, A.V., Carter, M.R., Sanderson, J.B. Developing disease-suppressive soils through crop rotation<br />

and tillage management practices. Soil Till Res 2003; 72:139-152.<br />

Petersen, C.A., Emanuel, M.E., Humphreys, G.B. Pathway of movement of apoplastic fluorescent dye tracers<br />

through the endodermis at the site of secondary root formation in corn (Zea mays) and broad bean (Vicia<br />

faba). Can J Bot 1981; 59:618-625.<br />

Pierson, E.A., Wood, D.W., Cannon, J.A., Blachere, F.M. III. LSP: interpopulation signaling via homoserine<br />

lactones among bacteria in the wheat rhizosphere. Mol Plant Interact 1998; 11:1078-1084.<br />

Piper, K.R., Beck von Bodman, S., Farrand S.K. Conjugation factor of Agrobacterium tumefaciens regulates<br />

Ti plasmid transfer by autoinduction. Nature 1993; 362:448-450.<br />

Pirhonen, M., Flego, D., Heikinheimo, R., and Palva, E.T. A small diffusible signal molecule is responsible for<br />

the global control of virulence and exoenzyme production in the plant pathogen Erwinia carotovora.<br />

EMBO J 1993; 12:2467-2476.<br />

Raaijmakers, J.M., Vlami, M., de Souza, J.T. Antibiotic production by bacterial biocontrol agents. Antonie van<br />

Leeuwenhoek Int Genet Mol Microbiol 2002; 81:537-547.<br />

Romeo, J.T. Raising the beam: moving beyond phytotoxicity. J Chem Ecol 2000; 26:2011-2014.<br />

Roos, I.M.M., Hattingh, M.J. Scanning electron microscopy Pseudomonas syringae pv. morsprunorum on<br />

sweet cherry leaves. Phytopathol Z 1983; 108:18-25.<br />

Salmond, G.P.C., Bycroft, B.W., Stewart, G.S.A.B., Williams, P., The bacterial ‘enigma’: cracking the code of<br />

cell-cell communication. Mol Microbiol 1995; 16:615-624.<br />

Sayre, R.M., Patrick, Z.A., Thorpe. H.J. Identification of a selective nematicidal component in extracts of plant<br />

residues decomposing in soil. Nematologica 1965; 11:263-68.<br />

Sayre, R.M., and Starr, M.P. Bacterial diseases and antagonisms of nematodes. In Diseases of nematodes.<br />

Poinar, G.O., Jansson H.B. eds. CRC Press: Florida, U.S.A, 1988; pp. 69-101.<br />

Schroth, M.N., Loper, J.E., Hildebrand, D.C. Bacteria as agents of plant edisease. In Current Perspectives in<br />

Microbial Ecology. Klug, M.J., Reddy, C.C. eds. Am Soc Microbiol: Washington, D.C., U.S.A. 1984;<br />

pp. 362-369.<br />

Schroth, M.N., Hancock, J.G. Disease-suppressive soil and root-colonizing bacteria. Science 1982; 216:1376-<br />

81.<br />

Schulz, F.A., Bateman, D.F. Temperature response of seeds during the early stages of germination and its relation<br />

to injury by Rhizoctonia solani. Phytopathology 1969; 59:352-55.<br />

Scott, R.I., Chard, J.M., Hocart, M.J., Lennard, J.H., Graham, D.C. Penetration of potato tuber lenticels by<br />

bacteria in relation to biological control of blackleg disease. Potato Res 1996; 39:333-344.<br />

Shanahan, P., O’Sullivan, D.J., Simpson, P., Glennon, J.D., O’Gara, F. Isolation of 2,4-diacetylphloroglucinol<br />

from a fluorescent pseudomonad and investigation of physiological parameters influencing its production.<br />

Appl Environ Microbiol 1992; 58:353-358.<br />

Shipton, P.J. Take-all decline during cereal monoculture. In Biology and Control of Plant Pathogens, Bruehl<br />

G.W. ed., St. Paul, MN: Am. Phytopathol. Soc. 1975; pp. 137-144.<br />

Siciliano, S.D., Theoret, C.M., de Freitas, J.R., Hucl, P.J., Germida, J.J. Differences in the microbial communities<br />

associated with the roots of different cultivars of canola and wheat. Can J Microbiol 1998; 44:844-851.<br />

Silo-Suh, L.A., Lethbridge, B.J., Raffel, S.J., He, H., Clardy, J., Handelsman, J. <strong>Biologica</strong>l activities of two<br />

fungistatic antibiotics produced by Bacillus cereus UW85. Appl Environ Microbiol 1994; 60:2023-30.<br />

Singh, P.P., Shin, Y.C., Park, C.S., Chung, Y.R. <strong>Biologica</strong>l control of Fusarium wilt of cucumber by chitinolytic<br />

bacteria. Phytopathology 1999; 89:92-99.<br />

Slininger, P.J., Jackson, M.A. Nutritional factors regulating growth and accumulation of phenazine 1-carboxylic<br />

acid by Pseudomonas fluorescens 2-79. Appl Microbiol Biotechnol 1992; 37:388-392.<br />

Slininger, P.J., Sheawilbur, M.A. Liquid culture pH, temperature, and carbon (not nitrogen) source regulate<br />

phenazine productivity of the take-all biocontrol agent Pseudomonas fluorescens 2-79. Appl Microbiol<br />

Biotechnol 1995; 43:794-800.<br />

Smith, K.P., Goodman, R.M. Host-variation for interaction with beneficial plant associated microbes. Annu<br />

Rev Phytopathol 1999; 37:473-491.<br />

Smucker, A.J.M. Soil Environmental Modifications of Root Dynamics and Measurement. Annu Rev Phytopathol<br />

1993; 31:191-216.<br />

Stephens, P.M., Crowley, J.J., O’Connel, C. Selection of pseudomonad strains inhibiting Pythium ultimum on<br />

sugarbeet seed in soil. Soil Biol Biochem 1993; 25:1283-1288.


BACTERIAL ROOT ZONE COMMUNITIES, BENEFICIAL ALLELOPATHIES AND PLANT DISEASE CONTROL<br />

Sturz, A.V., Carter, M.R., Johnston, H.R. A review of plant disease, pathogen interactions and microbial<br />

antagonism under conservation tillage in temperate humid agriculture. Soil Till Res 1997; 41:169-189.<br />

Sturz A.V., Christie, B.R. Beneficial microbial allelopathies in the root zone: the management of soil quality and<br />

plant disease with rhizobacteria. Soil Agro-ecosystems: Impacts of Management on Soil Health and Crop<br />

Diseases - Special Issue. Soil Till Res 2003; 72:107-123.<br />

Sturz, A.V., Christie, B.R., Arsenault, W.J. Red clover-potato cultivar combinations for improved potato yield.<br />

Agron J 2003; 95:1089-1092.<br />

Sturz, A.V., Christie, B.R., Nowak, J. Bacterial Endophytes: A Critical Component of Sustainable Crop<br />

Production, CRC Press, Crit Rev Plant Sci 2000; 19:1-30.<br />

Sturz, A.V., Christie, B.R., Matheson, B.G., Arsenault, W.J., Buchanan, N.A. Endophytic bacteria, antibiosis<br />

and community ecology in potato tubers, and implications for induced resistance against plant pathogens.<br />

Plant Pathol 1999; 48:360-370.<br />

Sturz, A.V., Matheson, B.G., Arsenault, W.J., Kimpinski, J., Christie, B.R. Weeds as a source of beneficial<br />

rhizobacteria in agricultural soils. Can J Microbiol 2001; 47:1013-1024.<br />

Sturz, A.V., Nowak, J. Endophytic communities of rhizobacteria and the strategies required to create yield<br />

enhancing associations with crops. Appl Soil Ecol 2000; 15:183-190.<br />

Sturz, A.V., Ryan, D.A.J., Matheson, B.G., Arsenault, W.J., Kimpinski, J., Christie, B.R. Stimulating disease<br />

suppression in soils: sulphate fertilizers can increase biodiversity and antibiosis ability of root zone bacteria<br />

against Streptomyces scabies. Soil Biol Biochem 2004; 36:343-352.<br />

Surette, M.G., Bassler, B.L. Quorum sensing in Escherichia coli and Salmonella typhimurium. Proc Natl<br />

Acad Sci USA. 1998; 95:7046-7050.<br />

Szabolcs, I. The concept of soil resilience. In, Soil resilience and sustainable land use. Greenland, D.J. and<br />

Szabolcs, I. eds. CAB International: Wallingford, U.K. 1994; pp. 33-39.<br />

Tateda, K., Comte, R., Pechère, J.C., Köhler, T., Yamaguchi, K., Van Delden, C. Azithromycin inhibits quorum<br />

sensing in Pseudomonas aeruginosa. Antimicrob. Agents Chemother 2001; 45:1930-33.<br />

Thomashow, L.S., Bonsall, R.F., Weller, D.M. Antibiotic production by soil rhizosphere microbes in situ. In<br />

Manual of environmental microbiology. Hurst, C.J., Knudsen, G.R., McInerney, M.J., Stetzenbach L.D.,<br />

Walter, M.V. eds. ASM Press: Washington, 1997; pp. 493-499.<br />

Timms Wilson, T.M., Ellis, R.J., Renwick ,A., Rhodes, D.J., Mavrodi, D.V., Weller, D.M., Thomashow, L.S.,<br />

Bailey, M.J. Chromosomal insertion of phenazine-1-carboxylic acid biosynthetic pathway enhances efficacy<br />

of damping-off disease control by Pseudomonas fluorescens. Mol Plant-Microbe Interact 2000; 13:1293-<br />

300.<br />

Tremblay, J., Beauchamp, C.J. Split applications of supplementary N fertilizer following incorporation of chipped<br />

ramical wood: changes in selected biological and chemical properties of a soil cropped to potatoes. Can J<br />

Soil Sci 1998; 78:275-282.<br />

Tu, J.C., Findlay, W.I. The effects of different green manure crops and tillage practices on pea root rots. British<br />

Crop Protection Conference - Pest and Diseases. British Crop Protection Council Publ.: Surrey, U.K.<br />

1986; pp. 229-236<br />

Tuzun S., Kloepper J.W. Induced systemic resistance by plant growth promoting rhizobacteria. In Improving<br />

Plant Productivity with Rhizosphere Bacteria. Ryder, M.H., Stephens, P.M., Bowen, G.D. eds. Proceedings<br />

of 3rd International Workshop on PGPR. CSIRO, Division of Soils: Adelaide. 1994; pp. 104-109.<br />

Van Buren, A.M., Andre, C., Ishimaru, C.A. <strong>Biologica</strong>l control of the bacterial ring rot pathogen by endophytic<br />

bacteria isolated from potato. Phytopathology 1993; 83:1406.<br />

van Peer, R., Punte, H.L.M., De Weger, L.A., Schippers, B. Characterization of root surface and endorhizosphere<br />

pseudomonads in relation to their colonization of roots. Appl Environ Microbiol 1990; 56:2462-2470.<br />

van Peer R., Niemann, G.J., Schippers B. Induced resistance and phytoalexin accumulation in biological control<br />

of Fusarium wilt of carnation by Pseudomonas sp. strain WCS417r. Phytopathology 1991; 81:728-34.<br />

van Wees, S.C.M., Luijendijk, M., Smoorenburg, I., van Loon, L.C., Pieterse, C.M.J. Rhizobacteria-mediated<br />

induced systemic resistance (ISR) in Arabidopsis is not associated with a direct effect on expression of<br />

known defense-related genes but stimulates the expression of the jasmonate-inducible gene Atvsp upon<br />

challenge. Plant Mol Biol 1999; 41:537-549.<br />

van Wees, S., Pieterse, C., Trijssenaar, A., Van’t Westende, Y., Hartog, F., Van Loon, L.C. Differential induction<br />

of systemic resistance in Arabidopsis by biocontrol bacteria. Mol Plant-Microbe Interact 1997; 10:716-<br />

24.<br />

141


142<br />

ANTONY V. STURZ<br />

Walker, J.C., Snyder, W.C. Pea wilt and root rots. Wisc Agr Exp Sta Bull 1933; 424:1-16.<br />

Wardle, D.A., Bonner, K.I., Nicholson, K.S. Biodiversity and plant litter: experimental evidence which does not<br />

support the view that enhanced species richness improves ecosystem function. Oikos 1997; 79:247-258.<br />

Wei, G., Kloepper, J.W., Tuzun, S. Induction of systemic resistance of cucumber to Colletotrichum orbiculare<br />

by select strains of plant-growth promoting rhizobacteria. Phytopathology 1991; 81:1508-1512.<br />

Weller, D.M., Raaijmakers, J.M., McSpadden Gardener, B.B., Thomashow, L. Microbial populations responsible<br />

for specific soil suppressiveness to plant pathogens Ann. Rev. Phytopathol 2002; 40:309-348.<br />

Weller, D.M. <strong>Biologica</strong>l control of soilborne plant pathogens in the rhizosphere with bacteria. Ann Rev Phytopathol<br />

1988; 26:379-407.<br />

Whipps, J.M. Microbial interactions and biocontrol in the rhizosphere. J Exp Bot 2001; 52:487-511.<br />

Whitehead, N.A., Welch, M., Salmond, G.P.C. Silencing the majority. Nat. Biotechnology. 2001; 19:735-736.<br />

Withers, H., Swift, S., Williams, P. Quorum sensing as an integral component of gene regulatory networks in<br />

Gram-negative bacteria. Curr Opin Microbiol 2001; 4:186-193.<br />

Yamada, Y., Nihira, T. Microbial hormones and microbial chemical ecology. In Comprehensive Natural Products<br />

Chemistry Barton, D.H.R., Nakanishi, K. eds.. Elsevier: Oxford, 1998; pp. 377-413.


ROBERT J. KREMER<br />

THE ROLE OF ALLELOPATHIC BACTERIA IN<br />

WEED MANAGEMENT<br />

U.S.D.A., Agricultural Research Service, Cropping Systems & Water Quality<br />

Unit, University of Missouri, Columbia, Missouri, U.S.A.<br />

Email: KremerR@missouri.edu<br />

Abstract. Allelopathic bacteria encompass those rhizobacteria that colonize the surfaces of plant roots, produce,<br />

and release phytotoxic metabolites, similar to allelochemicals, that detrimentally affect growth of plants. Practical<br />

application of this group of bacteria to agriculture could contribute to biological weed management systems that<br />

have less impact on the environment than conventional systems by reducing inputs of herbicides. Allelopathic<br />

bacteria have been investigated for potential as inundative-type biological control agents on several weeds.<br />

Because allelopathic bacteria generally do not attack specific biochemical sites within the plant, unlike<br />

conventional herbicides, they offer a means to control weeds without causing direct selective pressure on the<br />

weed population, therefore, development of resistance is not a major consideration. Additionally, the use of<br />

allelopathic bacteria appears to be environmentally benign relative to herbicides. These characteristics make<br />

allelopathic bacteria an attractive approach for managing crop weeds in a sustainable manner, even within the<br />

boundaries of conventional agriculture systems. However, recent evidence suggests that indigenous allelopathic<br />

bacteria might be exploited under certain crop and soil management practices that are inherently part of sustainable<br />

agricultural systems. The development of “weed-suppressive” soils in diverse sustainable systems is encouraging<br />

because indigenous populations of allelopathic bacteria may develop in several soils and environments using<br />

similar practices. The recent demonstrations of apparent weed-suppressive soils may lead to development of<br />

specific management strategies for the establishment and persistence of native allelopathic bacteria directly in<br />

soils conducive to annual weed infestations.<br />

1. INTRODUCTION<br />

<strong>Biologica</strong>l weed management is a system that incorporates the use of diverse biological<br />

organisms and biologically-based approaches including allelopathy, crop competition,<br />

and other cultural practices to significantly reduce weed densities in a manner that is<br />

similar to use of chemical herbicides alone (Cardina, 1995). Interest in developing<br />

effective biological weed management systems continues to increase because of a<br />

growing awareness of problems associated with the constant and intensive use of<br />

chemical herbicides, which include surface- and groundwater contamination,<br />

detrimental impacts on nontarget organisms, development of weeds resistant to<br />

herbicides (including those that are used in transgenic herbicide-resistant crops), and<br />

consumer concerns for residues on food (Gliessman, 2002). A component of biological<br />

weed management involves biological control, the intentional use of living organisms<br />

(insects, nematodes, fungi, and bacteria) to reduce the vigor, reproductive capacity,<br />

density, or impact of weeds (Quimby and Birdsall, 1995). A number of reviews are<br />

143<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 143– 155.<br />

© 2006 Springer. Printed in the Netherlands.


144<br />

ROBERT J. KREMER<br />

available that discuss the various strategies of biological control and the numerous<br />

organisms that are involved as biological agents within these strategies (TeBeest,<br />

1991; Harley and Forno, 1992; Boland and Kuykendall, 1998). The focus of the present<br />

paper will be a broad group of bacteria that are associated with seeds and seedlings,<br />

which can be developed to suppress the establishment and growth of weeds in<br />

agroecosystems.<br />

Unlike many herbicides and biological agents that have been developed to attack<br />

growing weeds established in crops, most bacterial agents target weed seeds residing<br />

in soil and the roots of developing weed seedlings. The microenvironments consisting<br />

of the zones of soil surrounding the seed (spermosphere) and root (rhizosphere) provide<br />

organic substrates that are readily available for soil microorganisms in contrast with<br />

the nutrient-limiting condition of the bulk soil (Kennedy, 2005). The spermosphere<br />

and rhizosphere, therefore, are ideal sites for establishing selected bacteria able to<br />

suppress weed seed germination and seedling growth because of the continuous supply<br />

of carbon and energy sources released from germinating seeds and developing root<br />

systems (Zahir et al., 2004). Soil borne bacteria adapted to competitive colonization<br />

of the spermosphere, rhizosphere, and the root can be grouped under the general term<br />

rhizobacteria (Schroth and Hancock, 1982). Although the spermosphere is defined<br />

with reference to a seed prior to root emergence, the zone of soil contains substances<br />

exuded from the seed that rapidly attracts and regulates a microbial community, which<br />

often, if the seed survives, establishes the dominant microbial communities of the<br />

longer-lived rhizosphere environment (Nelson, 2004). A component of the<br />

rhizobacteria group in the spermosphere and rhizosphere that inhibit plant growth<br />

without causing obvious disease symptoms are known as deleterious rhizobacteria.<br />

Deleterious rhizobacteria are predominantly saprophytic bacteria that live on or in<br />

plant seeds and roots, surviving on organic compounds released by plant seed root<br />

cells (Schippers et al., 1987). These bacteria do not parasitize the plant or penetrate<br />

the stele like major or true pathogens, but may colonize seed tissues or root hairs and<br />

the root tip as well as in the intercellular spaces of root cortical cells (Schippers et al.,<br />

1987), in which case they may be characterized as “endorhizal bacteria.” Because<br />

many of these deleterious rhizobacteria release phytotoxic metabolites that are also<br />

considered allelochemicals that influence the growth of plants, it has been suggested<br />

that term “allelopathic bacteria” may more accurately describe these bacteria (Barazani<br />

and Friedman, 1999; Sturz and Christie, 2003). Therefore, rhizobacteria that<br />

detrimentally affect seed germination and seedling development of weeds through the<br />

production of allelochemical substances will be referred to as allelopathic bacteria<br />

(AB) in this paper.<br />

The use of rhizobacteria in weed management has typically involved an inoculative<br />

approach whereby selected AB are applied at high rates to establish critical population<br />

densities in soil or on vegetative residues to achieve rapid initiation of growth-inhibitory<br />

activity (Kremer and Kennedy, 1996). The goal is not complete kill or eradication of<br />

the weed population but the reduction of the competitive ability of the weeds growing<br />

with the crop. Recent investigations of AB suggest that sustainable agricultural


ALLELOPATHIC BACTERIA IN WEED MANAGEMENT 145<br />

practices for many crops may be linked to management of the indigenous bacterial<br />

communities for biological control of weeds, thus reducing dependence on selected<br />

AB as inoculative biocontrol agents. Development of agroecosystems with the capacity<br />

to suppress weeds using naturally-occurring soil bacteria-weed interactions has received<br />

very little attention (Gallandt et al., 1999). Therefore, weed management strategies<br />

involving AB should also consider the development of beneficial, indigenous soil<br />

bacteria similar to a conservation biological control approach (Newman et al., 1998),<br />

which conceivably would address sustainable weed management with reduced or no<br />

herbicide inputs (Parr et al., 1992).<br />

The objectives of this chapter are to demonstrate how both introduced and<br />

indigenous AB contribute to weed management in cropping systems and to identify<br />

crop and soil management strategies that promote the proliferation of AB.<br />

2. ALLELOPATHIC BACTERIA AND INTERACTIONS IN<br />

AGROECOSYSTEMS<br />

<strong>Biologica</strong>l weed control using introduced agents including selected AB has often<br />

been limited by inconsistency of efficacy when placed into practical field use. This<br />

suggests that standard screening bioassays conducted under laboratory conditions<br />

poorly represent actual environmental situations where biological activity must be<br />

expressed. For individual AB agents, therefore, selection, bioassays, and inoculum<br />

development should be based on ecological principles, accounting for characteristics<br />

of the environment in which the agents are to be introduced. The principles should<br />

further be applied to potential strategies for enhancing indigenous AB in the soil<br />

environment. The ecological interactions associated with biological control of weed<br />

Soil Environment<br />

Allelopathic Bacteria<br />

Introduced | Indigenous<br />

Cultural<br />

Practices<br />

Weed Crop<br />

Figure 1. Environmental-interaction diagram associated with allelopathic<br />

bacteria in agroecosystems.


146<br />

ROBERT J. KREMER<br />

seeds and seedlings with AB can be expressed graphically (Figure 1) to emphasize<br />

the following key areas for consideration: a) biology and ecology of weeds; b) growth<br />

of the crop within the established cultivation system; c) characterization of AB,<br />

including those that are selected and introduced as biological control agents and those<br />

that are indigenous to the soil environment; and d) the wide range of cultural practices<br />

available for implementation of biological control within the agroecosystems.<br />

After recognizing the need for consideration of ecological interactions, a program<br />

for development of selected AB or management of the indigenous bacterial<br />

population can be undertaken<br />

3. CHARACTERIZATION OF ALLELOPATHIC BACTERIA<br />

Kremer et al. (1990) identified colonizing ability, chemotactic response, and mode of<br />

action to be vital characteristics for the successful development of rhizobacteria as<br />

weed biocontrol agents. Bacteria that can rapidly colonize the root will likely be a<br />

successful biocontrol agent. Migration towards the seed or root is the first step in<br />

colonization, illustrated by movement of rhizobacterial isolates through 2-cm of soil<br />

towards velvetleaf seeds (Begonia and Kremer, 1994). As the seedling develops,<br />

movement of bacteria along roots and within the rhizosphere is influenced by root<br />

binding sites, amounts of organic material present, type of root (i.e., seminal vs.<br />

nodal roots), water movement through soil and along roots, and soil texture (Bolton<br />

et al., 1993). Compared to other bacterial groups present in non-rhizosphere soil,<br />

gram-negative bacteria readily colonize the rhizosphere, partly due to their metabolic<br />

diversity (Nehl et al., 1997). Pseudomonads are particularly adapted for rhizosphere<br />

colonization because of the ability to utilize diverse carbon sources present in root<br />

exudates. Observations with scanning electron microscopy reveal the intimate<br />

relationship of rhizoplane colonization by selected AB (Figure 2).<br />

Figure 2. Root surface of a two-week old velvetleaf root colonized by Pseudomonas<br />

fluorescens strain 239 cells (arrow) aligned in intercellular spaces of root epidermal cells.<br />

Magnification is X 6,000. Scanning electron micrograph from Begonia et al. (1990).


ALLELOPATHIC BACTERIA IN WEED MANAGEMENT 147<br />

Successful competition of bacteria living in the rhizosphere depends on several<br />

factors, including rapid growth on multiple substrates, antibiotic production, and<br />

downward growth with the root. A major factor contributing to successful competition<br />

of rhizobacteria over other microorganisms is the growth stimulation by exuded organic<br />

compounds and sloughed-off root hair and epidermal cell materials (De Weger et al.,<br />

1995). The ability to efficiently compete for these available resources and to produce<br />

siderophores for obtaining iron is important in establishment, colonization, and<br />

persistence of rhizobacteria in the rhizosphere.<br />

A characteristic of many AB is the high specificity toward their weed host(s)<br />

with no detrimental effects on growth of nonweedy plant species (Cherrington and<br />

Elliott, 1987; Elliott and Lynch, 1985; Kennedy et al., 1991; 2001). Although effects<br />

on plants are subtle (Kremer and Kennedy, 1996), AB may be as significant as<br />

traditional bacterial pathogens in affecting plant growth (Schroth and Hancock, 1982;<br />

Suslow and Schroth, 1982). Because AB attack the seed and/or seedling rather than<br />

the growing plant, weed seed or vegetative propagule production is suppressed, a key<br />

to any weed management program, which reduces the need for repeated postemergence<br />

herbicide applications and increases the chances of success for control of a growing,<br />

competitive weed (Aldrich and Kremer, 1997).<br />

4. MODES OF ACTION OF ALLELOPATHIC BACTERIA<br />

Many AB strains produce secondary metabolites that are inhibitory to plants, including<br />

phytotoxins and antibiotics, which can be considered allelopathic. Phytotoxins from<br />

fluorescent Pseudomonas spp., a diverse group of plant pathogenic bacteria abundant<br />

in the soil and rhizosphere, have been well studied (Mitchell, 1991). There are fewer<br />

reports on phytotoxins from AB and many have not been extensively studied.<br />

A phytotoxin from Pseudomonas fluorescens strain D7 was shown to be<br />

responsible for root growth inhibition of downy brome (Bromus tectorum) (Tranel et<br />

al., 1993). Further characterization revealed that the active fraction was a complex of<br />

chromopeptides, other peptides and fatty acid esters in a lipopolysaccharide matrix<br />

(Gurusiddaiah et al., 1994). Secondary metabolites isolated from Pseudomonas<br />

syringae strain 3366 inhibitory to downy brome consisted of phenazine-1-carboxylic<br />

acid, 2-aminophenoxazone and 2-aminophenol (Gealy et al., 1996). Gealy et al.<br />

(1996) showed that phenazine-type antibiotics of Pseudomonas fluorescens also<br />

inhibited downy brome root growth. Electron microscopy of AB colonizing the<br />

rhizoplane and endorhizal cells of leafy spurge (Euphorbia esula) revealed disruption<br />

of plant cell walls and membranes apparently due to production of phytotoxins and/or<br />

enzymes by the bacteria, which consequently inhibited seedling growth (Souissi et<br />

al., 1997). AB may also produce “phytotoxic antibiotics” that affect plant growth<br />

such as the broad-spectrum antibiotic, 2,4-diacetylphloroglucinol, released by P.<br />

fluorescens strain CHA0, which suppressed soilborne fungal plant pathogens but was<br />

also highly phytotoxic to seedlings of several plant species (Keel et al., 1992).<br />

Plant-inhibitory effects of some AB are auxin-mediated, illustrated by direct uptake<br />

of bacterially produced indoleacetic acid (IAA). Plant response to microbially


148<br />

ROBERT J. KREMER<br />

synthesized auxins is related to the concentration released into the rhizosphere. Growth<br />

inhibition of several weed and crop species was correlated with elevated IAA levels<br />

produced by AB (Sarwar and Kremer, 1995). Similar responses were observed when<br />

tryptophan, an IAA precursor, was added to soil (Sarwar and Frankenberger, 1994;<br />

Sarwar and Kremer, 1995). Presumably, the presence of extra tryptophan in the soil<br />

provided rhizobacteria with additional substrates for auxin biosynthesis.<br />

The production of hydrogen cyanide (HCN), a volatile metabolite that negatively<br />

affects root metabolism and root growth by inhibiting cytochrome oxidase respiration,<br />

is common among rhizosphere pseudomonads (Schippers et al., 1990). The rate of<br />

HCN synthesis is affected by the availability of precursors such as glycine, methionine,<br />

proline and cyanogenic glucosides (Knowles and Bunch, 1986; Schippers et al., 1990).<br />

The amino acid composition of root exudates as well as environmental factors affecting<br />

root exudation (i.e, light intensity, soil water potential, nutrients) may be important<br />

as well (Schippers et al., 1990). Two strains of cyanogenic pseudomonads,<br />

Pseudomonas putida and Acidovorax delafieldii, significantly inhibited velvetleaf<br />

(Abutilon theophrasti) growth, but did not reduce corn growth in the presence of<br />

supplemental glycine (Owen and Zdor, 2001). Cyanide production by several<br />

rhizobacterial strains was a major factor in the inhibition of seedling growth of several<br />

weed species and was suggested as a trait for consideration in selecting AB as potential<br />

weed biological control agents (Kremer and Souissi, 2001).<br />

5. ROLES OF ALLELOPATHIC BACTERIA IN WEED<br />

MANAGEMENT STRATEGIES<br />

Strategies for the potential use of allelopathic bacteria for weed management include<br />

application of selected cultures as bioherbicides, integration of bioherbicides with<br />

other crop and soil management practices, use of AB in sustainable agricultural systems<br />

with other non-chemical means of weed control, and management of soils to enhance<br />

populations and activity of indigenous AB as a conservation biological control strategy.<br />

5.1. Bioherbicides<br />

The high host specificity of AB is a disadvantage for their use as a biocontrol strategy<br />

in most agroecosystems that are typically infested with multiple weed species. However,<br />

AB may have the greatest impact for management of specific weed problems in certain<br />

cropping systems. Isolates of AB specific for management of downy brome in wheat<br />

in the Pacific Northwest are under development as bioherbicides (Kennedy et al.,<br />

1991; 2001). In Kansas, selected AB suppressed early seedling growth of downy<br />

brome and Japanese brome (Bromus japonicus) in soils under winter wheat production<br />

(Harris and Stahlman, 1996). Several strains of rhizobacteria from over 2000 accessions<br />

isolated from prairie soils in Canada have been selected for inhibition of downy brome<br />

and green foxtail (Setaria viridis) seed germination and seedling growth (Boyetchko,<br />

1999). AB that were formulated in a starch-based granular carrier and applied to soil


ALLELOPATHIC BACTERIA IN WEED MANAGEMENT 149<br />

in leafy spurge-infested sites in South Dakota suppressed growth by decreasing root<br />

weight and root carbohydrate content (Brinkman et al., 1999).<br />

Most of the weeds targeted for biocontrol by AB infest cereal and row crops, but<br />

a few are perennial weeds of rangeland and forest ecosystems (Kremer, 2002). Selected<br />

AB are intended for soil application, however, some cultures might be effective when<br />

applied directly to growing weeds in a postemergence control strategy. Selected AB<br />

might also be applied directly to growing weeds as a postemergence control strategy.<br />

For example, cultures and cell-free supernatants of AB strains sprayed on common<br />

chickweed (Stellaria media), common lambsquarters (Chenopodium album), and field<br />

pennycress (Thlaspi arvense) in the greenhouse and field reduced plant biomass and<br />

survival (Weissmann and Gerhardson, 2001). Preliminary results suggest AB cultures<br />

might be used for selective weed control in growing crops through a one-time foliar<br />

application or as a follow-up to soil-incorporation of the same cultures.<br />

5.2. Integrated Weed Management<br />

Integrated weed management systems rely on a number of available strategies including<br />

tillage, cultural practices, herbicides, allelopathy, and biological control to reduce the<br />

weed seedbank, prevent weed emergence, and minimize competition from weeds<br />

growing with the crop (Aldrich and Kremer, 1997). These systems may be most suitable<br />

for implementing bioherbicides based on AB to counteract their limited weed host<br />

specificity. Like chemical herbicides, such bioherbicides may be most effective as a<br />

component in a multi-faceted management program rather than as a single tactic<br />

approach (Hatcher and Melander, 2003). Effective weed management offers several<br />

opportunities for integration of selected AB at the critical stages of weed development:<br />

as seeds in soil, as growing and competitive plants, and during seed production (Aldrich<br />

and Kremer, 1997). This may be the most promising situation for AB to be considered<br />

as practical management options in cropping systems.<br />

To broaden the limited spectrum of activity of AB, several tactics have been<br />

proposed for integration of these organisms with other weed management methods.<br />

Weed growth suppression by AB combined with herbicides applied at sublethal rates<br />

has met with some success (Greaves and Sargent, 1986). AB inhibitory to downy<br />

brome and jointed goatgrass suppressed growth to a greater extent when combined<br />

with metribuzin and/or diclofop at less than label rates (Kremer and Kennedy, 1996).<br />

An understanding of the mechanisms of herbicide-AB interactions will lead to<br />

strategies where AB selected for activity toward a weed can be paired with a specific<br />

chemical that increases the susceptibility of that weed to the AB (Kremer, 1998). Use<br />

of AB in this manner may develop into a systems management approach that involves<br />

integration of bioherbicides and herbicides on a physiological basis to control<br />

economically important weeds in corn and other crops. This is currently under intensive<br />

evaluation as a potential integrated management system in Europe (Müller-Schärer<br />

et al., 2000). Successful development of these integrated strategies will increase efficacy<br />

of AB agents, reduce herbicide inputs for weed control, and decrease potential<br />

environmental contamination.


150<br />

ROBERT J. KREMER<br />

Application of selected AB with cultural practices such as tillage may be effective<br />

in integrated weed management. Greater proportions of indigenous rhizobacteria<br />

inhibitory to downy brome and jointed goatgrass were detected under either<br />

conventional or reduced tillage compared to no-till (Kremer and Kennedy, 1996).<br />

Vegetative residues at or near the soil surface may serve as substrates for production<br />

of weed-suppressive chemicals by AB applied as bioherbicides directly to the residues.<br />

Kremer (1998) suggested that application of AB to surface vegetative residues to<br />

promote phytotoxin production might suppress weed growth prior to planting the<br />

crop, similar to a preemergence herbicide tactic. Crop rotation may also be manipulated<br />

to encourage development of specific inhibitory bacteria on weed roots. A “rotation<br />

effect” in corn, due partly to certain rhizobacteria specifically associated with corn<br />

roots, illustrates the potential for using AB to achieve suppression of weeds in crop<br />

rotation systems (Turco et al., 1990). Crop rotation may be manipulated to encourage<br />

development of specific inhibitory bacteria on weed roots. Thus, weed seeds and<br />

seedlings might be attacked by selected AB directly inoculated onto crop seeds or by<br />

promoting colonization of crop roots by microbial agents applied at planting (Skipper<br />

et al., 1996). Crop roots may deliver AB to adjacent roots of weeds and also maintain<br />

or even enhance AB numbers for attack of seedlings emerging later in the season.<br />

These observations suggest that combination of biocontrol agents, including selected<br />

AB, with cultural practices enhance weed growth suppression (Kremer and Kennedy,<br />

1996) and may be effective in controlling weeds that escape cultural control methods<br />

(Hatcher and Melander, 2003).<br />

5.3. Sustainable Agricultural Systems<br />

Sustainable agricultural systems involve a range of technological and management<br />

options to reduce costs, protect health and environmental quality, and enhance<br />

beneficial biological interactions and natural processes. Sustainable agricultural<br />

systems offer the greatest opportunities to study and refine nonchemical weed<br />

management (Liebman and Gallandt, 1997) and yield valuable information useful in<br />

developing improved bioherbicides, including AB, and advancing their use in broader<br />

biologically based weed management systems.<br />

High inputs of organic amendments and green manure (cover crop residues) in<br />

sustainable agricultural systems promotes the ability of crops to compete more<br />

vigorously with weeds, which intuitively suggests that the efficacy of bioherbicides<br />

would also be enhanced when used with these amendments (Gallandt et al., 1999).<br />

Addition of organic matter to soil is one of the most effective ways to change soil<br />

environment and favor increases in populations of beneficial soil organisms. Organic<br />

amendments subjected to decomposition in soils release compounds that suppress<br />

pathogens and provide substrates for other organisms, indigenous or those added to<br />

the amendments, that may also produce compounds that inhibit pathogens and/or<br />

weed seedling growth. Inoculation of organic materials (manures, composts, mulches)<br />

has been suggested as a means for assuring establishment and efficacy of the selected


ALLELOPATHIC BACTERIA IN WEED MANAGEMENT 151<br />

biocontrol agents, however, field performance to date has been inconsistent (Sturz<br />

and Christie, 2003).<br />

Cover crops and mulches as components of sustainable management systems<br />

may be used for integrating bioherbicides by delivering the agents on seeds and<br />

promoting their establishment in soils for attack of weeds and seedlings prior to<br />

planting. Recent research demonstrated that several cover crop species inoculated<br />

with selected AB at planting established and maintained the selected bacterial<br />

populations on their roots and in adjacent soil. When giant foxtail (Setaria faberi)<br />

emerged later in the season, the selected bacteria colonized seedling roots after the<br />

cover crop was terminated (Kremer, 2000). The selected AB and allelopathic cover<br />

crop residues acted synergistically to suppress the growth of the weeds.<br />

5.4. Suppressive Soils and Conservation <strong>Biologica</strong>l Control<br />

Soils under sustainable management may develop antagonistic microbial populations<br />

in rhizospheres of selected weeds that suppress their growth. This occurrence is similar<br />

to natural disease-suppressive soils in which indigenous soil microorganisms effectively<br />

protect crop plants from soilborne plant pathogens (Weller et al., 2002). Diseasesuppressive<br />

soils may be defined as soils in which a pathogen does not establish or<br />

persist, establishes but causes little or no damage, or establishes and causes disease<br />

for a short time but thereafter the disease is less important even though the pathogen<br />

may persist in the soil (Baker and Cook, 1974). Suppression is due primarily to<br />

antagonistic microorganisms, however, soil physical and chemical factors also may<br />

be involved (Weller et al., 2002). Similarly, weed-suppressive soils may be defined as<br />

soils in which certain weeds do not establish or persist, or establish and grow with the<br />

crop but cause little interference due to suppressed growth and vigor caused by native<br />

AB. Native and desirable plants may also in stimulate high populations of AB in their<br />

rhizospheres that reduce growth of invasive weed species, suggesting that plant-soil<br />

interactions are also involved in development of weed-suppressive soils (Kulmatiski<br />

et al., 2004).<br />

Evidence of apparent weed-suppressive soils has been reported for a variety of<br />

sustainable cropping systems. A study of crop management practices on claypan soils<br />

(Epiaqualfs) that involved reduced tillage, maintenance of high soil organic matter,<br />

and limited inputs of agrichemicals found increased levels of AB associated with<br />

weed seedlings that likely contributed to natural weed suppression (Li and Kremer,<br />

2000). It was reported that agronomic practices that resulted in relatively high organic<br />

matter, such as uncultivated prairie, organic farming and integrated cropping systems,<br />

supported higher proportions of weed AB. Compost-amended soils planted to winter<br />

wheat showed 29 and 78% reductions in broadleaf and grassy weed densities,<br />

respectively, compared to soils amended with inorganic fertilizers only (Carpenter-<br />

Boggs et al., 2000). Organic amendments (composts and cover crops) increased soil<br />

microbial biomass and decreased the seedbank density and emergence of shepherd’s


152<br />

ROBERT J. KREMER<br />

purse (Capsella bursa-pastoris) and burning nettle (Urtica urens) in a California<br />

vegetable production field (Fennimore and Jackson, 2003). The natural soil<br />

suppressiveness of the parasitic weed Striga hermonthica in Nigeria appears to be<br />

related to soils under rotations of cereal and leguminous crops that promote antagonist<br />

microbial populations that destroys Striga seeds before germination or kills the<br />

germinated seedlings (Berner et al., 1996; Dashiel et al., 1991). Each of the above<br />

systems strongly suggests that AB growth can be exploited as a sustainable weed<br />

control strategy using relatively simple management practices.<br />

Tactics and approaches for manipulating the field environment to enhance survival,<br />

physiological behavior, and performance of AB might easily be incorporated into<br />

diverse sustainable crop production systems. Such a strategy for natural weed<br />

suppression, also known as conservation biological control (Newman et al., 1998) or<br />

endemic soil-based control (Kulmatiski et al., 2004) relies on establishment of<br />

populations of indigenous or endemic, weed-suppressive microorganisms in soil. As<br />

demonstrated previously, many of these indigenous microorganisms are AB (Kremer<br />

and Li, 2003; Li and Kremer, 2000). Management practices including tillage, crop<br />

rotation, residue manipulation, and organic amendments enhance or induce favorable<br />

factors in the habitat for sustaining effective populations of natural AB. Crop<br />

management practices that involve reduced tillage, maintain high soil organic matter,<br />

and limit inputs of agrichemicals increased levels of deleterious rhizobacteria associated<br />

with weed seedlings and contribute to natural weed suppression (Li and Kremer,<br />

2000). Deliberate use of management practices that benefit natural weed-antagonistic<br />

AB can adversely affect weed population dynamics in production fields through seed<br />

and seedling mortality and growth suppression.<br />

6. SUMMARY<br />

The future use of AB to manage weeds in both conventional and sustainable agriculture<br />

seems promising. Because AB generally do not attack specific biochemical sites within<br />

the plant, unlike conventional herbicides, they offer a means to control weeds without<br />

causing direct selective pressure on the weed population, therefore, development of<br />

resistance is not a major consideration. Additionally, the use of AB appears to be<br />

environmentally benign relative to herbicides. These characteristics make AB an<br />

attractive approach for managing crop weeds in a sustainable manner, even within<br />

the boundaries of conventional agriculture systems. The recent demonstrations of<br />

apparent weed-suppressive soils may lead to development of specific management<br />

strategies for the establishment and persistence of native AB directly in soils conducive<br />

to annual weed infestations.<br />

7. REFERENCES<br />

Aldrich, R.J., Kremer, R.J. Principles in Weed Management. Iowa State University Press: Ames, IA, 1997<br />

Baker, K.F., Cook, R.J. <strong>Biologica</strong>l Control of Plant Pathogens. Freeman: San Francisco, 1974.


ALLELOPATHIC BACTERIA IN WEED MANAGEMENT 153<br />

Barazani, O., Friedman, J. Allelopathic bacteria and their impact on higher plants. Crit Rev Plant Sci 1999;<br />

18:741-755.<br />

Begonia, M.F.T., Kremer, R.J. Chemotaxis of deleterious rhizobacteria to velvetleaf (Abutilon theophrasti)<br />

seeds and seedlings. FEMS Microbiol Ecol 1994; 15:227-236.<br />

Begonia, M.F.T., Kremer, R.J., Stanley, L., Jamshedi, A. Association of bacteria with velvetleaf roots. Trans Mo<br />

Acad Sci 1990; 24:17-26.<br />

Berner, D., Carsky, R., Dashiell, K., Kling, J., Manyong, V. A land management based approach to integrated<br />

Striga hermonthica control in sub-Saharan Africa. Outlook Agric 1996; 25:157-164.<br />

Boland, G.J., Kuykendall, L.D. Plant-Microbe Interactions and <strong>Biologica</strong>l Control. Marcel Dekker: New York,<br />

1998.<br />

Bolton, Jr., H., Fredrickson, J.K., Elliott, L.F. Microbial ecology of the rhizosphere. In: Soil Microbial Ecology.<br />

Blaine Metting, Jr., F. ed., Marcel Dekker, Inc.: New York., 1993; pp. 27-63.<br />

Boyetchko, S.M. Innovative applications of microbial agents for biological weed control. In: Biotechnological<br />

Approaches in Biocontrol of Plant Pathogens. Mukerji, K.G., Chamola, B.G., Upadhyay, R.K., eds.,<br />

Plenum Press: London, 1999; pp. 73-97.<br />

Brinkman, M.A., Clay, S.A., Kremer, R.J. Influence of deleterious rhizobacteria on leafy spurge (Euphorbia<br />

esula) roots. Weed Technol 1999; 13:835-839.<br />

Cardina, J. <strong>Biologica</strong>l weed management. In, Handbook of Weed Management Systems, Smith, A.E. ed. Marcel<br />

Dekker: New York, 1995; pp. 279-341.<br />

Carpenter-Boggs, L., Reganold, J.P., Kennedy, A.C. Biodynamic preparations: short-term effects on crops, soils,<br />

and weed populations. Am J Altern Agric 2000; 15:110-118.<br />

Cherrington, C.A., Elliott, L.F. Incidence of inhibitory pseudomonads in the Pacific Northwest. Plant Soil 1987;<br />

101:159-165.<br />

Dashiell, K.E., Jackai, L.E.N., Hartman, G.L., Ogundipe, H.O., Asafo-Adjei, B. Soybean germplasm diversity,<br />

uses and prospects for crop improvement in Africa. In, Crop Genetic Resources of Africa, Ng, N.O.,<br />

Perrino, P., Attere, F., Zedan, S. eds. Vol II. Proc. International Conference on Crop Genetic Resources of<br />

Africa (pp. 203-212), IITA: Ibadan, Nigeria, 1991; pp. 203-212.<br />

De Weger, L.A., van der Bij, A.J., Dekkers, L.C., Simons, M., Wijffelman, C.A., Lugtenberg, B.J. Colonization<br />

of the rhizosphere of crop plants by plant-beneficial pseudomonads. FEMS Microbiol Ecol 1995; 17:221-<br />

228.<br />

Elliott, L.F., Lynch, J.M. Plant growth-inhibitory pseudomonads colonizing winter wheat (Triticum aestivum<br />

L.) roots. Plant Soil 1985; 84:57-65.<br />

Fennimore, S.A., Jackson, L.E. Organic amendment and tillage effects on vegetable field weed emergence and<br />

seedbanks. Weed Technol 2003; 17:42-50.<br />

Gallandt, E.R., Liebman, M., Huggins, D.R. Improving soil quality: implications for weed management. J Crop<br />

Product 1999; 2:95-121.<br />

Gealy, D. R., Gurusiddaiah, S., Ogg, Jr., A.G. Isolation and characterization of metabolites from Pseudomonas<br />

syringae strain 3366 and their phytotoxicity against certain weed and crop species. Weed Sci 1996;<br />

44:383-392.<br />

Gliessman, S.R. Allelopathy and agroecology. In Chemical Ecology of Plants: Allelopathy in Aquatic and<br />

Terrestial Ecosystems, Inderjit, A.U. Mallik eds., Birkhauser Verlag: Zurich, 2002; pp. 173-185.<br />

Greaves, M. P., Sargent, J.A. Herbicide-induced microbial invasion of plant roots. Weed Sci 1986; 34:50-53.<br />

Gurusiddaiah, S., Gealy, D.R., Kennedy, A.C., Ogg, Jr., A.G. Isolation and characterization of metabolites<br />

from Pseudomonas fluorescens-D7 for control of downy brome (Bromus tectorum). Weed Sci 1994;<br />

42:492-501.<br />

Harley, K.L.S., Forno, I.W. <strong>Biologica</strong>l Control of Weeds: A Handbook for Practitioners and Students. Inkata<br />

Press: Melbourne, 1992.<br />

Harris, P.A., Stahlman, P.W. Soil bacteria as selective biological control agents of winter annual grass weeds in<br />

winter wheat. Appl Soil Ecol 1996; 3:275-281.<br />

Hatcher, P.E., Melander, B. Combining physical, cultural and biological methods: prospects for integrated nonchemical<br />

weed management strategies. Weed Res 2003; 43:303-322.<br />

Keel, C., Schnider, U., Maurhofer, M. Voisard, C., Laville, J., Burger, U., Wirthner, P., Haas, D., DeFago, G.<br />

Suppression of root diseases by Pseudomonas fluorescens CHA0: importance of the bacterial secondary<br />

metabolite 2,4-diacetylphloroglucinol. Molec Plant Microbe Interact 1992; 5:4-13.


154<br />

ROBERT J. KREMER<br />

Kennedy, A.C. Rhizosphere. In, Principles and Applications of Soil Microbiology, Sylvia, D.M., Fuhrmann,<br />

J.J., Hartel, P.G., Zuberer D.A., eds. Pearson Prentice Hall: Upper Saddle River, NJ, 2005; pp. 242-262.<br />

Kennedy, A.C., Johnson, B.N., Stubbs, T.L. Host range of a deleterious rhizobacterium for biological control of<br />

downy brome. Weed Sci 2001; 49:792-797.<br />

Kennedy, A.C., Elliott, L.F., Young, F.L., Douglas, C.L. Rhizobacteria suppressive to the weed downy brome.<br />

Soil Sci Soc Am J 1991; 55:722-727.<br />

Knowles, C. J., Bunch, A.W. Microbial cyanide metabolism. Adv Microbiol Physiol 1986; 27:73-111.<br />

Kremer, R.J., Li, J. Developing weed-suppressive soils through improved soil quality management. Soil Till<br />

Res 2003; 72:193-202.<br />

Kremer, R.J. Bioherbicides: potential successful strategies for weed control. In, Microbial Biopesticides, Koul<br />

O., Dhaliwal B., eds., Taylor & Francis: London, 2002, pp. 307-323.<br />

Kremer, R.J., Souissi, T. Cyanide production by rhizobacteria and potential for suppression of weed seedling<br />

growth. Curr Microbiol 2001; 43:182-186.<br />

Kremer, R.J. Growth suppression of annual weeds by deleterious rhizobacteria integrated with cover crops. In,<br />

Proceedings of the Xth International Symposium on <strong>Biologica</strong>l Control of Weeds, Spencer, N.R. ed.<br />

USDA-ARS and Montana State University: Bozeman, MT, 2000; pp. 931-940.<br />

Kremer, R.J. Microbial interactions with weed seeds and seedlings and its potential for weed management. In,<br />

Integrated Weed and Soil Management, Hatfield, J.L., Buhler, D.D., Stewart, B.L. eds., Ann Arbor<br />

Press: Chelsea, MI, 1998; pp. 161-179.<br />

Kremer, R.J., Kennedy, A.C. Rhizobacteria as biocontrol agents of weeds. Weed Technol 1996; 10:601-609.<br />

Kremer, R.J., Begonia, M.F.T., Stanley, L., Lanham, E.T. Characterization of rhizobacteria associated with<br />

weed seedlings. Appl Environ Microbiol 1990; 56:1649-1655.<br />

Kulmatiski, A., Beard, K.H., Stark, J.M. Finding endemic soil-based controls for weed growth. Weed Technol<br />

2004; 18:1353-1358.<br />

Li, J., Kremer, R.J. Rhizobacteria associated with weed seedlings in different cropping systems. Weed Sci 2000;<br />

48:734-741.<br />

Liebman, M., Gallandt, E.R. Many little hammers: ecological management of crop-weed interactions. In, Ecology<br />

in Agriculture, Jackson, L.E. ed. Academic Press: San Diego, 1997; pp. 291-343.<br />

Müller-Schärer, H., Scheepens, P.C., Greaves, M.P. <strong>Biologica</strong>l control of weeds in European crops: recent<br />

achievements and future work. Weed Res 2000; 40:83-98.<br />

Nehl, D.B., Allen, S.J., Brown, J.F. Deleterious rhizosphere bacteria: an integrating perspective. Appl Soil Ecol<br />

1997; 5:1-20.<br />

Nelson, E.B. Microbial dynamics and interactions in the spermosphere. Annu Rev Phytopathol 2004; 42:271-<br />

309.<br />

Newman, R.M., Thompson, D.C., Richman, D.B. Conservation strategies for the biological control of weeds.<br />

In, Conservation <strong>Biologica</strong>l Control, Barbosa P. ed. Academic Press: San Diego, 1998; pp. 371-396.<br />

Owen, A., Zdor, R. Effect of cyanogenic bacteria on the growth of velvetleaf (Abutilon theophrasti) and corn<br />

(Zea mays) in autoclaved soil and the influence of supplemental glycine. Soil Biol Biochem 2001; 33:801-<br />

809.<br />

Parr, J.R., Papendick, R.I., Hornick, S.B., Meyer, R.B. Soil quality: attributes and relationship to alternative<br />

and sustainable agriculture. Amer J Alt Agric 1992; 7:5-11.<br />

Quimby, P.C., Birdsall, J.L. Fungal agents for biological control of weeds: classical and augmentative approaches.<br />

In, Novel Approaches to Integrated Pest Management. Reuveni R. ed., CRC Press: Boca Raton, FL,<br />

1995, pp. 293-308.<br />

Sarwar, M., Kremer, R.J. Enhanced suppression of plant growth through the production of L-tryptophan derived<br />

compounds by deleterious rhizobacteria. Plant Soil 1995; 172:261-269.<br />

Sarwar, M., Frankenberger, Jr., W.T Influence of L-tryptophan and auxins applied to the rhizosphere on the<br />

vegetative growth of Zea mays L. Plant Soil 1994; 160:97-104.<br />

Schippers, B., Bakker, A.W., Bakker, P.A.H.M., van Peer, R. Beneficial and deleterious effects of HCN-producing<br />

pseudomonads on rhizosphere interactions. Plant Soil 1990; 129:75-83.<br />

Schippers, B., Bakker, A.W., Bakker, P.A. Interactions of deleterious and beneficial rhizosphere microorganisms<br />

and the effect of cropping practices. Annu. Rev. Phytopathol 1987; 25:339-358.<br />

Schroth, M.N., Hancock, J.G. Disease-suppressive soil and root-colonizing bacteria. Science 1982; 216:1376-<br />

1381.


ALLELOPATHIC BACTERIA IN WEED MANAGEMENT 155<br />

Skipper, H.D., Ogg, Jr., A.G., Kennedy, A.C. Root biology of grasses and ecology of rhizobacteria for biological<br />

control. Weed Technol 1996; 10:610-620.<br />

Souissi, T., Kremer, R.J., White, J.A. Scanning and transmission electron microscopy of root colonization of<br />

leafy spurge (Euphorbia esula L.) seedlings by rhizobacteria. Phytomorphology 1997; 47:177-193.<br />

Sturz, A.V., Christie, B.R. Beneficial microbial allelopathies in the root zone: the management of soil quality<br />

and plant disease with rhizobacteria. Soil Till Res 2003; 72:107-123.<br />

Suslow, T.V. Role of root-colonizing bacteria in plant growth. In, Phytopathogenic Prokaryotes, Mount, M.S.,<br />

Lacy, G.H. eds. Volume 1. Academic Press: New York, 1982, pp. 187-223.<br />

Suslow, T.V., Schroth, M.N. Role of deleterious rhizobacteria as minor pathogens in reducing crop growth.<br />

Phytopathology 1982; 72:111-115.<br />

TeBeest, D.O. Microbial Control of Weeds. Chapman and Hall: New York, 1991.<br />

Tranel, P.J., Gealy, D.R., Kennedy, A.C. Inhibition of downy brome (Bromus tectorum) root growth by a<br />

phytotoxin from Pseudomonas fluorescens strain D7. Weed Technol. 1993; 7:134-139.<br />

Turco, R.F., Bischoff, M., Breakwell, D.P., Griffith, G.R. Contribution of soil-borne bacteria to the rotation<br />

effect in corn. Plant Soil 1990; 122:115-120.<br />

Weissmann, R., Gerhardson, B. Selective plant growth suppression by shoot application of soil bacteria. Plant<br />

Soil 2001; 234:159-170.<br />

Weller, D.M., Raaijmakers, J.M., McSpadden-Gardner, B.B., Thomashow, L.S. Microbial populations responsible<br />

for specific soil suppressiveness to plant pathogens. Annu Rev Phytopathol 2002; 40:309-348.<br />

Zahir, Z.A., Arshad, M., Frankenberger, Jr., W.T. Plant growth promoting rhizobacteria: applications and<br />

perspectives in agriculture. Adv Agron 2004; 81:97-168.


MARK A. BERNARDS 1 , LINA F. YOUSEF 1 , ROBERT W. NICOL 2<br />

THE ALLELOPATHIC POTENTIAL<br />

OF GINSENOSIDES<br />

1<br />

Department of Biology, University of Western Ontario, London,<br />

2<br />

ON, Canada, N6A 5B7<br />

NovoBiotic Pharmaceuticals, Cambridge MA, 02138, USA<br />

Email: bernards@uwo.ca<br />

Abstract. American ginseng (Panax quinquefolius L.) is a perennial herb valued for the medicinal properties<br />

of its large, fleshy tap root. These medicinal properties are purported to be due to the triterpenoid saponins, or<br />

ginsenosides, that accumulate to 3-6% of the root dry weight. We asked the question: what is the ecological role<br />

of ginsenosides in Panax species? In addressing this question, we have determined that ginsenosides, like other<br />

saponins, possess fungitoxic properties, although different fungi and oomycotan organisms appear to be<br />

differentially affected by them in vitro. In order to play an allelopathic role, however, ginsenosides must be<br />

present in the soil at biologically active (i.e., ecologically relevant) concentrations. Results to date support the<br />

hypothesis that ginsenosides are phytoanticipins and serve as host resistance factors. The success of certain<br />

pathogens (e.g., Pythium cactorum, Pythium irregulare, Cylindrocarpon destructans) on ginseng may arise<br />

from their ability to detoxify or otherwise utilize ginsenosides as a nutritive source or growth stimulating factor,<br />

while other soil borne organisms appear susceptible to their fungitoxic properties. Ginsenosides have been<br />

isolated from rhizosphere soil and root exudates suggesting that these compounds are involved in allelopathic<br />

interactions between the host plant and soil fungi. Ultimately this allelopathic interaction may influence the<br />

fungal diseases of ginseng.<br />

1. INTRODUCTION<br />

American ginseng (Panax quinquefolius L.) produces triterpenoid saponins called<br />

ginsenosides, which are slowly released into the rhizosphere. Ginsenosides, like other<br />

saponins, are fungitoxic, and as such may modify the balance of microorganisms in<br />

the soil; i.e., they are potentially allelopathic. From a disease management point of<br />

view, the extent to which ginsenosides alter the soil microbial community may have<br />

profound consequences, especially when disease causing organisms are favoured in<br />

the new balance. Understanding how ginsenosides affect soilborne microbes is important<br />

in understanding disease cycles in this crop.<br />

2.1. Saponins<br />

2. SAPONINS<br />

Saponins, are glycosylated natural products with surfactant and soap-like properties<br />

that tend to froth in aqueous solution, even at low concentration (Dewick, 1997).<br />

157<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 157– 175.<br />

© 2006 Springer. Printed in the Netherlands.


158<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

They also cause haemolysis, though are generally non-toxic when taken orally. Saponins<br />

are broadly categorised as being either triterpenoid- or steroid-derived, despite the<br />

common triterpenoid origin of both. Triterpenoid saponins are widely distributed in<br />

many dicot families, and are generally based on pentacyclic triterpene parent carbon<br />

skeletons (e.g., α- and β-amyrin), or tetracyclic parent compounds (e.g., the<br />

dammaranes). By contrast, steroidal saponins are commonly found in monocots,<br />

and are characterised by a spiroketal moiety at C-22. Plants of the Solanaceae<br />

contain steroidal glycoalkaloid saponins, which are nitrogen analogues of steroidal<br />

saponins.<br />

In general, saponins are glycosylated at the C-3 hydroxyl group, typically with a<br />

glucose molecule. However, there are other potential glycosylation sites on the parent<br />

carbon skeletons, and multi-site glycosylation can occur. Accordingly, saponins are<br />

subdivided into monodesmosidic (one glycosylation site) and bisdesmosidic<br />

(glycosylation sites at both ends of the compound) categories. The majority of saponins,<br />

however, are monodesmosidic. While glucose is the most common sugar in<br />

saponins, arabinose, galactose, rhamnose and xylose, as well as sugar acids are also<br />

present. The glycosylation patterns of saponins can lead to small families of compounds<br />

based on the same parental carbon skeleton (i.e., sapogenin or aglycone),<br />

differing only in the composition and arrangement of sugars.<br />

2.2. Saponins as Fungitoxic and Plant Defense Compounds<br />

There are numerous reports in the literature that describe the anti-fungal activity of<br />

saponins (Zimmer et al., 1967; Levy et al., 1986; 1989; Marston et al., 1988; Takechi<br />

and Tanaka, 1990; Ohtani et al., 1993; Ouf et al., 1994; Favel et al., 1994; Escalante<br />

et al., 2002; Nicol et al., 2002). These reports coupled with molecular data (Bowyer et<br />

al., 1995; Papadopoulou et al., 1999) suggest that these phytochemicals are constitutive<br />

plant defenses against fungi (Osbourn, 1996; 2003), that is phytoanticipins (Van<br />

Etten et al., 1994). For example, the role that the triterpenoid saponin avenacin A-1<br />

plays in conferring resistance to “take-all disease”, caused by Gaeumannomyces<br />

graminis in oats, is well established (Turner, 1956; Burkhardt et al., 1964; Maizel et<br />

al., 1964; Bowyer et al., 1995; Papadopoulou et al., 1999).<br />

While the molecular mechanism of saponin fungitoxicity is not known, avenacin<br />

A-1 as well as steroidal glycoalkaloid saponins such as α-tomatine and solanine have<br />

been demonstrated to form complexes with membrane sterols, thereby causing a reduction<br />

in membrane integrity (Roddick, 1979; Steel and Drysdale, 1988; Keukens et<br />

al., 1992; 1995; Armah et al., 1999). Two models have been proposed to explain the<br />

consequences of saponin-sterol aggregation in target membranes (Morrissey and<br />

Osbourn, 1999). One suggests that the deleterious effects of saponins are related to<br />

the formation of transmembrane pores (Armah et al., 1999), whereas the other suggests<br />

that membrane integrity is compromised due to the extraction of sterols (Keukens<br />

et al., 1992; 1995). Regardless of the exact mechanism, saponins probably disrupts<br />

fungal membranes through complexation with ergosterol, the major membrane sterol<br />

in higher fungi (Evans and Gealt 1985; Weete 1989; Griffiths et al., 2003).


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 159<br />

Saponins may act as host chemical defenses, but because fungi successfully attack<br />

plants containing these defenses, there must be means of tolerating or avoiding<br />

saponin toxicity. Fungal resistance to defensive chemicals in general can either involve<br />

enzymatic detoxification of antifungal compounds or the more ambiguous “innate<br />

resistance” (Morrissey and Osbourn, 1999). Some fungi are able to detoxify<br />

phytoalexins (Van Etten et al., 1995) as well as phytoanticipins such as saponins<br />

(Osbourn et al., 1995a; Van Etten et al., 1995; Weltring et al., 1997). Although<br />

saponin detoxification generally occurs via enzymatic cleavage of the saccharides to<br />

form the aglycone parent compound, different fungi employ different strategies and<br />

enzymes. For example, fungal detoxification of α-tomatine by Botrytis cinerea (Quidde<br />

et al., 1998) Septoria lycopersici (Arneson and Durbin, 1967) and Verticillium alboatrum<br />

(Pegg and Woodward, 1986) involves cleavage of one terminal monosaccharide<br />

from the tetrasaccharide chain at carbon three of the sapogenin, whereas Fusarium<br />

oxysporum f. sp. lycopersici removes the entire tetrasaccharide (Ford et al., 1977).<br />

The saponin-detoxifying enzymes tomatinase from S. lycopersici and avenacinase<br />

from G. graminis exhibit high sequence similarity (Osbourn et al., 1995b), whereas<br />

the tomatinase from F. oxysporum f. sp. lycopersici appears to be more related to a<br />

different family of glycosyl hydrolases (Roldán-Arjona et al., 1999). The ability to<br />

efficiently detoxify host chemical defenses determines virulence and host range in<br />

fungal pathogens such as G. graminis, Rhizoctonia solani and Phoma lingam (Bowyer<br />

et al., 1995; Pedras et al., 2000a,b). Efficient detoxification is taken an extra step by<br />

S. lycopersici as the hydrolysis products of the saponin-based defense subsequently<br />

inhibit inducible defenses of the host by interfering with signal transduction pathways<br />

(Bouarab et al., 2002).<br />

Transformation of secondary metabolites (e.g., cleaving sugars from saponins)<br />

has been suggested as a way of providing fungi with a carbon source in addition to<br />

achieving detoxification (Van Etten et al., 1995; Roldán-Arjona et al., 1999). Since<br />

some fungal detoxification enzymes are repressed by glucose (Straney and Van Etten,<br />

1994; Roldán-Arjona et al., 1999) they may have a dual function. This would be<br />

analogous to non-pathogenic fungi that obtain carbon from phenolic monomers produced<br />

during lignin decomposition (Henderson and Farmer, 1955; Rahouti et al.,<br />

1989). Presumably it would be advantageous for pathogens to circumvent host defenses<br />

and simultaneously derive nutrition.<br />

Pathogens in the Pythiaceae (Oomycota) appear to possess an innate resistance to<br />

the toxic effects of saponins. Members in this family are reported to be relatively<br />

unaffected by aescin (Olsen, 1971) and other saponins (Assa et al., 1972) in vitro, and<br />

probably the lack of ergosterol in these species (Olsen, 1973a; Weete, 1989) allows<br />

them to avoid saponin toxicity (Arneson and Durbin 1968; Olsen, 1971; Weltring,<br />

1997). Although Pythium spp. appear to be unaffected by saponins, saponin toxicity<br />

can be induced through the addition of sterols to the growth medium (Olsen, 1973b;<br />

Steel and Drysdale, 1988). Members of the Pythiaceae can incorporate sterols by the<br />

binding and transport action of small protein carriers called elicitins (Mikes et al.,<br />

1997; Panabières et al., 1997; Capasso et al., 2001) and this uptake of exogenous<br />

sterols could be the cause of the observed increase in the deleterious effects of saponins<br />

on these organisms.


160<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

2.3. Criteria for Saponins as <strong>Allelochemicals</strong><br />

Allelopathy is the study of those interactions between and among plants and microbes<br />

that are mediated by secondary compounds i.e., allelochemicals (Rice, 1984). The<br />

concept of allelopathy is most frequently applied within the context of plant-plant<br />

chemical interactions. <strong>Allelochemicals</strong> can act as a form of chemical interference<br />

between competing species (Rice, 1984) as well as conspecific plants (Singh et al.,<br />

1999). However, it has recently been argued that allelochemicals are unlikely to<br />

reach phytotoxic levels in the soil and therefore plant-microbe allelopathy may be<br />

more likely to occur (Schmidt and Ley, 1999). It has been suggested that slow diffusion<br />

rates and complexation reactions in soil, coupled with degradation/utilization by<br />

microbes would generally prevent allelochemicals from accumulating to phytotoxic<br />

concentrations (Schmidt and Ley, 1999). Microbes are expected to metabolize<br />

allelochemicals because soil is generally considered to be low in available carbon<br />

(Sparling et al., 1981; Scow, 1997) and these organisms are known to utilize a wide<br />

range of molecules as carbon sources (Henderson and Farmer, 1955; Black and Dix,<br />

1976; Campbell et al., 1997).<br />

Two related but unintegrated lines of evidence suggest that plants do in fact<br />

influence specific soil microbes via secondary chemicals. First, both plant species<br />

and genotype are known to affect the rhizosphere species composition of mycorrhizal<br />

fungi (Johnson et al., 1992) and actinomycetes and some bacteria (Azad et al., 1987;<br />

Miller et al., 1989; Larkin et al., 1993). Recently, results obtained using molecular<br />

(Miethling et al., 2000) and physiological (Grayston et al., 2001) methods have confirmed<br />

the primary importance of plant species on the composition of soil microbial<br />

communities. Second, of the carbon fixed by photosynthesis in plants, an estimated<br />

10 to 20% is released into the rhizosphere (Bowen and Rovira, 1991) and in some<br />

instances this amount may exceed 20% (Shepherd, 1994). The potential effects of<br />

soil-deposited carbon may extend a distance from the root, as carbon fixed by the<br />

aerial portions of maize plants has been found over 3 cm away from the roots (Helal<br />

and Sauerbeck, 1984). A diverse array of organic secondary compounds from plants<br />

(e.g., alkaloids, phenolics, quinones, saponins, stilbenes) are potent antifungal agents<br />

(Grayer and Harborne, 1994), and various species/pathovars of fungi can be differently<br />

susceptible to these chemicals (Zimmer et al., 1967; Arneson and Durbin, 1968;<br />

Suleman et al., 1996; Sandrock and Van Etten, 1998). It then follows that if secondary<br />

compounds are present in the rhizosphere, they could influence the growth and/or<br />

species composition of the soil microbial community and therefore have to be considered<br />

allelochemicals. However, with the exception of the flavonoids identified in treemycorrhizal<br />

interactions (Bécard et al., 1992; Lagrange et al., 2001) and legumenodulating<br />

bacteria interactions (D’Arcy Lameta and Jay, 1987; Peters and Long,<br />

1988) and the determination of chemicals in Arabidopsis thaliana root exudates<br />

(Narasimhan et al., 2003), these “root exudates” are not well characterized. In order<br />

to establish whether specific compounds or groups of compounds such as saponins<br />

are allelochemicals, therefore, they first must be shown to be present in the rhizosphere<br />

(i.e., to determine their ecologically relevant concentration), and second, be


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 161<br />

shown to be biologically active at their ecologically relevant concentration. Lastly,<br />

the allelopathic role of the compounds has to be demonstrated at the field level.<br />

3.1. The Host Plant<br />

3. GINSENG AND GINSENG SAPONINS<br />

American ginseng (Panax quinquefolius L.) is a native North American member of<br />

the Araliaceae, a family whose more than 800 species are found mostly in the tropics<br />

(Lawrence, 1951). Panax quinquefolius is a perennial understory herb that is associated<br />

with deciduous forests (Fountain, 1986; Anderson et al., 1993) and ranges from<br />

Ontario and Quebec, south to northern Florida and west to Minnesota (Small and<br />

Catling, 1999). The aboveground tissues senesce at the end of each growing season<br />

and estimates of the maximum age of this plant are 23-30 yr (Anderson et al., 1993)<br />

to >50 yr (Lewis and Zenger, 1982). Ginseng typically has one aerial stem, with<br />

three to five palmately compound leaves and an umbelliferous inflorescence. The<br />

main pollinators of the small greenish-white to greenish-yellow flowers are bees<br />

(Catling and Spicer, 1995) and the mature red fruits usually contain two seeds, but<br />

range from one to three seeds (Anderson et al., 1993; Schlessman, 1985). Before<br />

germinating, the seeds of American ginseng require an after-ripening period of one<br />

to two winters (Lewis and Zenger, 1982; Anderson et al., 1993). Panax quinquefolius<br />

is threatened in Canada (Small and Catling, 1999), probably due to over-collection of<br />

the root for economic gain.<br />

The main commercial product from ginseng is the taproot, which is sought for its<br />

purported medicinal properties. Commercially, the roots are harvested after 3-5 yr of<br />

intensive cultivation. American ginseng has been commercially cultivated in Canada<br />

since the late nineteenth century (Proctor and Bailey, 1987), under artificial shade or<br />

in interplanted woodlands. In 2002 Ontario, which is one of Canada’s largest provincial<br />

producers, exported over one thousand metric tons of ginseng root representing a<br />

value of almost $ 40 million (OMAF, 2003). However, commercial productivity is<br />

hindered by susceptibility to several fungal diseases of the leaves, stem, fruit and<br />

roots.<br />

3.2. Ginsenoside Structure and Biosynthetic Origins<br />

Ginsenosides are triterpenoid saponins primarily based on two tetracyclic dammarane<br />

ring structures: (20S)-protopanaxadiol and (20S)-protopanaxatriol (Figure 1), with<br />

one representative (Ro) derived from oleanoic acid. While more than 25 ginsenoside<br />

structures have been identified in Panax spp. (Fuzzati et al., 1999), six are considered<br />

major: the (20S)-protopanaxadiol-derived Rb 1 , Rb 2 , Rc and Rd and the (20S)protopanaxatriol-derived<br />

Re and Rg 1 . With the exception of a few monodesmosidic<br />

compounds (e.g., Rh 2 , Rg 3 , Rh 1 , Rf and Rg 2 ), ginsenosides are generally bisdesmosidic.<br />

(20S)-Protopanaxadiols are 3-O and 20-O diglycosides, while (20S)-protopanaxatriols


162<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

are 6-O and 20-O diglycosides (Figure 1). Ginsenosides can represent approximately<br />

3-7% dry weight of the root (Court et al., 1996; Li et al., 1996; Nicol et al., 2002) and<br />

are found in the foliage at lower levels (Chen and Staba, 1978). In Asian ginseng,<br />

Panax ginseng C.A. Meyer, ginsenosides are found in the outer cortex and periderm<br />

of roots (Tani et al., 1981).<br />

Figure 1. The Major Ginsenosides of Panax spp. The three main triterpenoid aglycones<br />

(20S)-protopanaxadiol, (20S)-protopanaxatriol, oleanoic acid), as well as the six major<br />

ginsenosides of Panax spp. are shown. Ginsenosides are named according to their migration<br />

properties on TLC, with no regard for whether they are (20S)-protopanaxadiol-, (20S)protopanaxatriol-<br />

or oleanoic acid-derived. Ginsenoside Rf (not shown) is a major<br />

ginsenoside of P. ginseng but is not found in P. quinquefolius.


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 163<br />

Triterpenoids are derived from the cytosolic mevalonate pathway via the common<br />

30-carbon intermediate squalene (Dewick, 1997; Haralampidis et al., 2002). Briefly,<br />

oxidation of squalene to 2,3-oxidosqualene generates an intermediate that can be<br />

folded and cyclised (by 2,3-oxidosqualene cyclases) into any one of the different classes<br />

of triterpenes found in plants. That is, if the 2,3-oxidosqualene is folded into a chairboat-chair-boat<br />

configuration, protonation of the 2,3-epoxide (which ultimately<br />

generates the 3-OH common to all triterpenoids and steroids) results in the concerted<br />

cyclisation of 2,3-oxidosqualene into a protosteryl cation which upon protonation<br />

yields either cycloartenol (plants) or lanosterol (animals, fungi), both of which are<br />

precursors to the steroidal saponins and steroids. Alternative folding into a chairchair-chair-boat<br />

configuration gives rise to a dammarenyl cation (Figure 2), which<br />

Figure 2. Hypothetical Biosynthesis of Ginsenoside Aglycones. The biosynthesis of ginsenoside<br />

aglycones begins with the formation of squalene-2,3-oxide by a flavoprotein-containing epoxide<br />

synthase (1). A putative dammarane synthase (2a-c) is thought to bind squalene-2,3oxide<br />

and act as a template for folding it into a chair-chair-chair-boat configuration (2a).<br />

Protonation of the epoxide initiates a concerted series of ring closures (2b), terminating in a<br />

dammarenyl cation with the appropriate stereochemistry. In Panax spp., the dammarenyl cation<br />

is expected to be quenched by water (2c) yielding 3,20-dihydroxydammarene. Subsequent<br />

hydroxylation reactions would yield the protopanaxadiol and protopanaxatriol parent carbon<br />

skeletons of the common ginsenosides. Two potential hydroxylation pathways are shown (see<br />

text for details). All enzymes involved in the production of the tetracyclic ginsenosides are<br />

hypothetical for Panax spp. (based on Dewick, 1997; Haralampidis et al., 2002).


164<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

can undergo subsequent 1,2-hydride and/or 1,2 alkyl shifts and further annulation.<br />

Loss of a proton yields pentacyclic saponin precursors such as α- or β- amyrin or<br />

lupeol.<br />

In ginseng, however, only one relatively low abundant ginsenoside (Ro) has a<br />

pentacyclic parent structure similar to most other non-ginseng triterpenoid saponins.<br />

Instead, most ginsenosides are based on the (20S)-protopanaxadiol and (20S)protopanaxatriol<br />

parent carbon skeletons derived from the dammarenediol structure<br />

that results from the quenching of the dammarenyl cation by water. This reaction<br />

(i.e., chair-chair-chair-boat template folding, concerted annulation and final cation<br />

quenching) is hypothesized to be catalysed by dammarenediol synthase (Dewick, 1997;<br />

Haralampidis et al., 2002), although this has not been proven. Other enzymes involved<br />

in triterpenoid biosynthesis have been identified in Asian ginseng (P. ginseng) including<br />

an oxidosqualene cyclase (â-amyrin synthase, Kushiro et al., 1998) as well as other<br />

candidate oxidosqualene cyclases and enzymes involved in modification of the<br />

sapogenin (Jung et al., 2003). Further hydroxylation reactions to yield the (20S)protopanaxadiol<br />

and (20S)-protopanaxatriol parent carbon skeletons from 3,20dihydroxydammarene<br />

have not been described. However, two routes are theoretically<br />

possible (Figure 2). In the first, 3,20-dihydroxydammarene is hydroxylated at C-12 to<br />

yield the (20S)-protopanaxadiol skeleton directly. Subsequent hydroxylation at C-6<br />

yields the (20S)-protopanaxatriol skeleton. Alternatively, the (20S)-protopanaxatriol<br />

skeleton could arise independent of the (20S)-protopanaxadiol skeleton via<br />

hydroxylation at C-6 first, followed by C-12 hydroxylation. In either case, the<br />

assumption is made that the parent carbon skeletons are assembled first, before<br />

glycosylation. No details are yet available with respect to the glycosylation of (20S)protopanaxadiol<br />

and (20S)-protopanaxatriol parent carbon skeletons. Clearly<br />

glycosylation of the unique C-20 hydroxyl group, which is glycosylated in all<br />

ginsenosides except two (20S)-protopanaxadiols (Rh 2 , Rg 3 ) and Ro, represents a novel<br />

glycosyl transferase activity, as does glycosylation of C-6 of the (20S)-protopanaxatriols.<br />

4. GINSENOSIDE FUNGITOXICITY IN VITRO<br />

4.1. Differential Response of Fungi and Oomycetes to the Addition of Ginsenosides<br />

to Culture Media<br />

In commercial gardens, ginseng is attacked by the foliar pathogens Alternaria panax,<br />

A. alternata, and Botrytis cinerea, and the root pathogens Cylindrocarpon destructans,<br />

Rhizoctonia solani, P. cactorum, Py. irregulare, Py. ultimum and several species of<br />

Fusarium including F. oxysporum and F. solani (Brammall, 1994a, b; Reeleder and<br />

Brammall, 1995; Punja, 1997). Although the pathogens in the Pythiaceae (i.e.,<br />

Phytophthora cactorum, Pythium irregulare, Py. ultimum) are superficially similar,<br />

and share the same nutritional mode as the other pathogens, they are not true fungi,<br />

but rather belong in the kingdom Straminipila (Burnett, 2003). Most other ginseng


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 165<br />

pathogens, are ascomycetes. The activity of ginsenosides against these pathogens, as<br />

well as saprotrophic Trichoderma spp., was evaluated in vitro (Nicol et al., 2002;<br />

unpublished data). The Trichoderma spp. are potential antagonists towards soilborne<br />

pathogens, evident by the greater levels of Trichoderma spp. found in healthy ginseng<br />

fields than in replanted ginseng fields with high disease incidence (Shin and Lee,<br />

1986). In vitro anti-fungal bioassays were completed by adding ginsenosides to growth<br />

media and comparing the relative growth to that of controls (i.e., with no ginsenoside<br />

addition to the growth media). Under these conditions, the growth of six of the nine<br />

tested organisms was inhibited. The highest growth inhibition was observed in the<br />

saprotrophs T. harzianum and T. hamatum followed by the leaf pathogen A. panax<br />

and T. viride (Table 1).<br />

Table 1. Growth Response of Selected Fungal and Oomycotan Species to<br />

Ginsenosides. The relative growth of the microorganisms was compared in vitro with<br />

and without the addition of ginsenosides isolated from Panax quinquefolius roots.<br />

Growth was standardized to that of controls (i.e., no ginsenosides added) to allow<br />

comparisons across species, even though different assay conditions were used for<br />

different organisms. Growth data is shown as a mean ± standard deviation. With the<br />

exception of F. oxysporum, the growth of all organisms in the presence of ginsenosides<br />

was significantly different from the mean growth of the respective controls (data not<br />

shown). Data compiled from Nicol et al. 2002, 2003 and unpublished data.<br />

Fungal Species % Growth relative to control<br />

Trichoderma harzianum -26.4 ± 0.9<br />

Trichoderma hamatum -22.2 ± 2.8<br />

Trichoderma viride -9.4 ± 1.1<br />

Alternaria panax -17.1 ± 3.8<br />

Fusarium solani -3.3 ± 0.6<br />

Fusarium oxysporum -3.0 ± 0.8<br />

Cylindrocarpon destructans +7.6 ± 2.9<br />

Phytophthora cactorum +324.9 ± 1.0<br />

Pythium irregulare +392.8 ± 0.5<br />

The growth of the two Fusarium species, F. oxysporum and F. solani, was consistently<br />

found to be slightly lower, but not always significantly different from control<br />

(Table 1). Conversely, the growth of the causal organisms of some of the most devastating<br />

diseases in the North American ginseng industry (i.e., C. destructans, P.<br />

cactorum and Py. irregulare) was significantly stimulated over that of control. When<br />

analysed as a group, significantly different growth responses to the ginsenosides were<br />

observed across the fungal and oomycotan species tested (Table 1). That is, the organisms<br />

tested were generally inhibited (Trichoderma spp. and A. panax), unaffected<br />

(Fusarium spp.) or stimulated (C. destructans, P. cactorum and Py. irregulare) by<br />

ginseng saponins. By comparison, greater antifungal activity was found with aescin,<br />

a mixture of saponins from the horse chestnut tree (Nicol et al. 2002) and consequently,<br />

ginsenosides can only be considered to be mildly antifungal.


166<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

4.2. “Detoxification” of Ginsenosides by Oomycetes<br />

As noted in Section 1.2., two mechanisms of resistance have been proposed to explain<br />

the insensitivity of Oomycetes to saponins, (i) innate resistance, due to little or no<br />

sterols in the membranes of Oomycete species and (ii) detoxification, via an enzymatic<br />

degradation of saponins into less bioactive derivatives. With respect to the<br />

observed growth stimulation of the Oomycetes Py. irregulare and P. cactorum in the<br />

presence of ginsenosides in vitro, it is not clear which of these mechanisms may be<br />

involved. However, preliminary results suggest that the “detoxification” of ginsenosides<br />

(exemplified by Py. irregulare) is the likely mechanism. For example, we initially<br />

hypothesized that if the fungitoxicity of ginsenosides resulted from their interaction<br />

with sterols, then the addition of ergosterol to the growth medium of Pythium spp.<br />

and Phytophthora spp. would render these organism more susceptible (as has been<br />

demonstrated with aescin, e.g., Olsen, 1973b). However, we have found that the<br />

addition of ergosterol and ginsenosides to the growth medium of the two ginseng<br />

pathogens, Py. irregulare and P. cactorum resulted in a cumulative biomass increase<br />

(~ 200% and 150% compared to control colony weights) relative to cultures grown<br />

with either ergosterol or ginsenosides alone (Nicol et al., 2003). This observation<br />

suggests that the resistance of Pythium spp. and Phytophthora spp. to ginsenosides<br />

cannot simply be due to a lack of sterol content in their membranes.<br />

Recently, a glycoside hydrolase (BGX1) was described for P. infestans (Brunner<br />

et al., 2002), the first glycosidase of its kind for an oomycotan organism. With the<br />

assumptions that (i) the pathogenesis mechanisms of Oomycetes are similar to those<br />

of fungi (Latijnhouwers et al., 2003), and (ii) ginsenosides act as preformed defense<br />

compounds (Nicol et al., 2002), then the production of glycosidases by members of<br />

the Pythiaceae could be an important factor in explaining their pathogenicity to American<br />

ginseng. Current attempts to rationalize the effect of ginsenosides on the growth<br />

of Py. irregulare involve exploring the ability of this organism to modify ginsenosides<br />

in vitro. Thus, when ginsenosides were re-extracted from spent broth that had been<br />

supplemented with 0.1% ginsenosides prior to inoculation with Py. irregulare, only<br />

the (20S)-protopanaxatriol-derived ginsenosides Re and Rg 1 were recovered (Figure<br />

3, lower trace). That is, the (20S)-protopanaxadiol ginsenosides (Rb 1 , Rb 2 , Rc, Rd)<br />

were all substantially, if not completely depleted from the medium compared to the<br />

initial supplement (compare Figure 3 upper trace with Figure 3 lower trace).<br />

Moreover, the decline in (20S)-protopanaxadiol ginsenosides in the spent broth<br />

of Py. irregulare was linear over the five day culture period (data not shown).<br />

The disappearance of protopanaxadiol ginsenoside peaks from HPLC chromatograms<br />

was coincident with the appearance of an unidentified peak eluting at approximately<br />

45 minutes (Figure 3, lower trace). It is tempting to speculate that the 45<br />

minute peak is the (20S)-protopanaxadiol aglycone resulting from the de-glycosylation<br />

of ginsenosides by Py. irregulare, but this requires experimental verification. By contrast,<br />

the profile of ginsenosides recovered from broth inoculated with T. hamatum, (a<br />

control organism sensitive to ginsenosides; Table 1) was unaltered relative to that of<br />

broth in which no organism had been cultured (compare Figure 3 upper trace with


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 167<br />

Figure 3. HPLC Analysis of Ginsenosides. Ginsenosides were isolated from spent<br />

broth in which either no organism (upper trace), Trichoderma hamatum (middle trace)<br />

or Pythium irregulare (lower trace) had been cultured for five days at 25 o C in the<br />

dark. Ginsenosides were chromatographed on a Microsorb-MV C-18 column (150 x<br />

4.6 mm, 5 mm) using an acetonitrile:H 2 O gradient (Nicol et al., 2002), and detected<br />

at 203 nm. The * in the lower trace indicates the unknown metabolite found in the<br />

spent broth of Py. irregulare.


168<br />

Figure 3 middle trace). Therefore, the differential response of these two organisms to<br />

the presence of ginsenosides in their growth medium is coincident with differences in<br />

the profile of ginsenosides that can be recovered from their spent medium.<br />

Interestingly, preliminary observations also show that the recovery of ginsenosides<br />

from the spent broth of Py. irregulare is dependent on the presence of sucrose in the<br />

original culture medium, since significantly smaller amounts of ginsenosides were<br />

recovered from a spent broth lacking sucrose (Yousef and Bernards, unpublished<br />

data). This observation implies that Py. irregulare is using ginsenosides as a source<br />

for carbon. However, the additional observation that mineral broth supplemented with<br />

a concentration of glucose equimolar to that expected to be released into the medium<br />

by ginsenoside de-glycosylation, does not support the same degree of growth increase<br />

in Py. irregulare biomass observed when the same broth is supplemented with<br />

ginsenosides (Yousef and Bernards, unpublished data), argues against a simple carbon<br />

source mechanism. Since Pythium spp. have been reported to incorporate sterols into<br />

their membranes (Olsen, 1973a) and because sterols are known to mediate growth<br />

(Nes, 1987), we now hypothesize that Py. irregulare secretes saponinases to (partially)<br />

deglycosylate ginsenosides, the latter of which are then incorporated as “sterol”<br />

triterpenoids into its membrane. It is further assumed that under the experimental<br />

conditions employed, this incorporation favours growth.<br />

5.1. RETS and Soil Analysis<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

5. GINSENOSIDES IN THE RHIZOSPHERE<br />

When the soil associated with three-year-old ginseng roots was extracted and analysed<br />

by HPLC, six major ginsenosides were tentatively identified by co-elution with<br />

standards. While much of the soil chemical HPLC profile remains unidentified, HPLC-<br />

MS analysis confirmed the presence of the six major ginsenosides (Rb 1 , Rb 2 , Rc, Rd,<br />

Re and Rg 1 ) plus pseudoginsenoside F 11 and another protopanaxadiol ginsenoside in<br />

the soil extracts (Nicol et al., 2003). The amount of ginsenosides as percent weight of<br />

dry soil was calculated to range from 0.02% to 0.098% (average 0.06%).<br />

In order to confirm that ginsenosides were present in the exudate of intact ginseng<br />

roots, (and not isolated from residual root tissue in our soil preparations) root exudates<br />

were collected from pot-grown ginseng plants using a root exudate trapping system,<br />

or RETS (Tang and Young, 1982). HPLC analysis of the trapped exudate revealed<br />

the presence of peaks that had the same retention times as the ginsenoside standards.<br />

These peaks were not present in the exudate collected from control pots (no ginseng<br />

plants) and were taken as evidence of ginsenosides in the exudate of pots containing<br />

ginseng plants. HPLC-MS analysis of the trapped exudate confirmed the presence of<br />

the same suite of ginsenosides as found in the soil (Nicol et al., 2003). After<br />

quantification of the ginsenoside content of the root exudates, the individual ginseng<br />

plants were determined to be losing approximately 25 µg of ginsenosides per day<br />

(i.e., amount of recovered ginsenoside / number of plants / number of days the<br />

experiment ran).


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 169<br />

5.2. Bioactivity of Ginsenosides at Ecologically Relevant Concentrations<br />

As emphasized in Section 1.3, before ginsenosides can be considered allelochemicals,<br />

it has to be demonstrated that they are in fact present in the rhizosphere soil at biologically<br />

active concentrations. Therefore, after demonstrating the presence of<br />

ginsenosides in the soil, in vitro bioassays were conducted using ginsenosides at an<br />

ecologically relevant concentration of 0.06%. At this level, ginsenosides were shown<br />

to remain bioactive. That is, the growth of Py. irregulare was significantly greater<br />

than control, while that of T. hamatum remained unchanged (Nicol et al., 2003).<br />

5.3. Plant-Fungal Allelopathy in Ginseng Gardens and Implications for Disease<br />

The discovery of ginsenosides in ginseng root exudates and rhizosphere soil at biologically<br />

active concentrations suggests that plant-fungal allelopathy could occur within<br />

this crop. This in turn could influence disease levels as the ginseng secondary chemicals<br />

potentially, and differentially, influence the growth of different groups of soilborne<br />

organisms (i.e., pathogens and antagonistic saprotrophs).<br />

Naturally-occurring soilborne antagonists can play a role in preventing or reducing<br />

disease levels in crops (Azad et al., 1987; Miller et al., 1989; Larkin et al., 1993).<br />

However, the applied use of Trichoderma spp. as a biocontrol agent often does not<br />

exert the desired level of disease control. One reason is that isolates of these biocontrol<br />

fungi often suffer from poor rhizosphere competence (Papavizas, 1982; Chao et al.,<br />

1986). Consequently, poor (natural) Trichoderma spp. competence in the rhizosphere<br />

of ginseng could lead to increased disease levels in the plant. Interestingly, in an<br />

attempt to address the issue of the lack of Trichoderma spp. rhizosphere competence,<br />

it was found that exudates from healthy plant roots did not support the growth of T.<br />

harzianum, but did in fact support that of Py. ultimum (Green et al., 2001). More<br />

evidence for the involvement of a soil factor in preventing successful biocontrol by<br />

Trichoderma spp. is found in the work of Hong et al. (2000) where several Trichoderma<br />

isolates were shown to inhibit C. destructans in vitro, but this biocontrol effect<br />

failed to materialize when applied to potted ginseng plants. Again, the specific mechanism<br />

involved in both of these reports was not identified, but our results suggest that<br />

plant-fungal allelopathy is one possible explanation.<br />

Ginsenoside-mediated stimulation of several major pathogens is likely to be<br />

involved in the eventual establishment of ginseng diseases. Three important root<br />

pathogens were repeatedly observed to grow better after the addition of ginsenosides<br />

to their growth medium (i.e., C. destructans, P. cactorum and Py. irregulare).<br />

Therefore, it can be hypothesized that the chemical environment of the ginseng<br />

rhizosphere favours fungal root pathogens over potential antagonists of these pathogens<br />

and that this has a direct consequence on the year-to-year disease levels in ginseng<br />

gardens. We are still working on the exact mechanism of growth stimulation in<br />

pathogens, but our preliminary results lead us to believe that in one case (i.e., Py.<br />

irregulare) the activity is extra-nutritional. For Py. irregulare, increased growth may<br />

be due to the ginsenosides or transformed ginsenosides acting as sterol analogs. Also,


170<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

rhizosphere secondary chemicals in general may be serving as signal molecules for<br />

plant pathogens in a manner analogous to Rhizobium-legume and tree-mycorrhizae<br />

symbioses (Peters and Long, 1988; Larange et al., 2001). The ability of rhizosphere<br />

phytochemicals to fulfill a molecular need for the pathogen or to induce cellular<br />

pathways important for pathogenesis or virulence is a relatively unexplored avenue of<br />

plant-microbe allelopathy, even though this has obvious implications for diseases in<br />

economically important plants.<br />

Anecdotal evidence suggests that ginseng crops suffer from a replant syndrome<br />

and a common farming practice is to avoid planting consecutive ginseng gardens in<br />

the same area. Although there is no data on the cause of the replant problem in<br />

ginseng, it has been suggested that pathogens could be involved (Reeleder et al.,<br />

1999). In fact the ginseng replant problem bears a similarity to the situation in apple<br />

orchards. Using our results, a replant syndrome in ginseng gardens could be postulated<br />

to have the same basic ingredients (i.e., phytochemicals, pathogens and antagonists)<br />

present in the model of the apple replant syndrome. For example, it is well<br />

established that the “dominant causal agents” (Mazzola, 1998) of the apple replant<br />

syndrome are a complex of soilborne fungal pathogens, especially species in the genera<br />

Cylindrocarpon, Fusarium, Phytophthora and Pythium (Mazzola 1999; Isutsa<br />

and Merwin, 2000) including C. destructans, P. cactorum and Py. irregulare (Braun,<br />

1991; 1995; Mazzola, 1998). Although pathogens play a primary role, antagonists<br />

have also been implicated through results obtained from their application as biocontrol<br />

microbes (Utkhede et al., 2001) or surveys of orchard soil (Mazzola, 1999). Finally,<br />

it has been suggested that an important component of apple replant syndrome is the<br />

induction of virulence in microbes by root exudates (Szabo and Wittenmayer, 2000).<br />

Similarly, unidentified allelochemicals may also be involved in the asparagus replant<br />

syndrome (Peirce and Colby, 1987; Hartung and Stephens, 1983; Pedersen et al.,<br />

1991). Therefore, a ginseng replant syndrome, like that of apple and asparagus,<br />

could be due to a phytochemical-mediated alteration in the soil microbial population<br />

resulting in the stimulation of a complex of pathogens coupled with an inhibition of<br />

beneficial microbes. In other words, allelopathy may play a role in year-to-year soilborne<br />

crop disease cycles as well as replant syndromes.<br />

6. CONCLUDING REMARKS<br />

Plant disease researchers often focus on causal agents in relative isolation. Our research<br />

further demonstrates that other host-derived factors may also play an important role<br />

in the etiology and severity of plant diseases and replant syndromes. That is, allelopathy<br />

may contribute to plant diseases through the ability of certain microbes to grow and/<br />

or survive preferentially in the phytochemical profile of the rhizosphere surrounding<br />

specific crop roots. Because pathogenicity is a common nutritional mode in microbes,<br />

it is not surprising that pathogens are adapted to the chemical environment of their<br />

host plants. Manipulation of rhizosphere chemistry, therefore, either through genetic<br />

modification of plants, identification and use of microbe-influencing genotypes or the


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 171<br />

planting of mixed crops, represents a potential, but as yet underdeveloped, opportunity<br />

to reduce disease levels in economically important plants.<br />

7. REFERENCES<br />

Anderson, R.C., Fralish, J.S., Armstrong, J.E., Benjamin, P.K. The ecology and biology of Panax quinquefolium<br />

L. (Araliaceae) in Illinois. Am Midl Nat 1993; 129:357-372.<br />

Armah, C.N., Mackie, A.R., Roy, C., Price, K., Osbourn, A.E., Bowyer, P., Ladha, S. The membranepermeabilizing<br />

effect of avenacin A-1 involves the reorganization of bilayer cholesterol. Biophys J 1999;<br />

76:281-290.<br />

Arneson, P.A., Durbin, R.D. The sensitivity of fungi to α-tomatine. Phytopathology 1968; 58:536-537<br />

Assa, Y., Gestetner, B., Chet, I., Henis, Y. Fungistatic activity of lucerne saponins and digitonin as related to<br />

sterols. Life Sci 1972; 11:637-647.<br />

Azad, H.R., Davis, J.R., Schnathorst, W.C., Kado, C.I. Influence of Verticillium wilt resistant and susceptible<br />

potato genotypes on populations of antagonistic rhizosphere and rhizoplane bacteria and free nitrogen<br />

fixers. Appl Microbiol Biotechnol 1987; 26:99-104.<br />

Bécard, G., Douds, D.D., Pfeffer, P.E. Extensive in vitro hyphal growth of VAM fungi in the presence of CO and<br />

2<br />

flavonols. Appl Environ Microbiol 1992; 58:821-825.<br />

Black, R.L.B., Dix, N.J. Utilization of ferulic acid by microfungi from litter and soil. Trans Br Mycol Soc 1976;<br />

66:313-317.<br />

Bouarab, K., Melton, R., Peart, J., Baulcombe, D., Osbourn, A. A saponin-detoxifying enzyme mediates<br />

suppression of plant defences. Nature 2002; 418:889-892.<br />

Bowen, G.D., Rovira, A.D. The rhizosphere. In, Plant Roots, the Hidden Half. Waisel, Y., Eshel, A., Kafkafi,<br />

U., eds. New York: Marcel Dekker Inc. 1991; pp. 641-670.<br />

Bowyer, P., Clarke, B.R., Lunness, P., Daniels, M.J., Osbourn, A.E. Host range of a plant pathogenic fungus<br />

determined by a saponin detoxifying enzyme. Science 1995; 267:371-374.<br />

Brammall, R.A. Alternaria blight. In Diseases and Pests of Vegetable Crops in Canada. Howard, R.J, Garland,<br />

J.A., Seaman, W.L. ed., The Canadian Phytopathological Society and the Entomological Society of Canada,<br />

Ottawa, ON. 1994a; pp. 294-295.<br />

Brammall, R.A. Disappearing root rot. In, Diseases and Pests of Vegetable Crops in Canada. Howard, R.J.,<br />

Garland, J.A., Seaman, W.L., ed., The Canadian Phytopathological Society and the Entomological Society<br />

of Canada, Ottawa, ON. 1994b; pp. 296-297.<br />

Braun, P.G. The combination of Cylindrocarpon lucidum and Pythium irregulare as a possible cause of apple<br />

replant disease in Nova Scotia. Can J Plant Pathol 1991; 13:291-297.<br />

Braun, P.G. Effects of Cylindrocarpon and Pythium species on apple seedlings and potential role in apple<br />

replant disease. Can J Plant Pathol 1995; 17:336-341.<br />

Brunner, F., Wirtz, W., Rose, J.K.C., Darvill, A.G., Govers, F., Scheel, D., Nuernberger, T. A beta-glucosidase/<br />

xylosidase from the phytopathogenic oomycete, Phytophthora infestans. Phytochemistry 2002; 59:689-<br />

696.<br />

Burkhardt, H.J., Maizel, J.V., Mitchell, H.K. Avenacin, an antimicrobial substance isolated from Avena sativa.<br />

II. Structure. Biochemistry 1964; 3:426-431.<br />

Campbell, C.D., Grayston, S.J., Hirst, D.J. Use of rhizosphere carbon sources in sole carbon utilisation tests to<br />

discriminate soil microbial communities. J Microbiol Methods 1997; 30:33-41<br />

Capasso, R., Cristinzio, G., Di Maro, A., Ferranti, P., Parente, A. Syringicin, a new alpha-elicitin from an isolate<br />

of Phytophthora syringae, pathogenic to citrus fruit. Phytochemistry 2001; 58:257-262.<br />

Catling, P.M., Spicer, K.W. Pollen vectors in an American ginseng crop. Econ Bot, 1995; 49:99-102.<br />

Chao, W.L., Nelson, E.B., Harman, G.E., Hoch, H.C. Colonization of the rhizosphere by biological control<br />

agents applied to seeds. Phytopathology 1986; 76:60-65.<br />

Chen, S.E., Staba, E.J. American ginseng I. Large scale isolation of ginsenosides from leaves and stems. Lloydia<br />

1978; 41:361-366.<br />

Court, W.A., Reynolds, L.B., Hendel, J.G. Influence of root age on the concentration of ginsenosides of American<br />

ginseng (Panax quinquefolium). Can J Plant Sci 1996; 76:853-855.<br />

D’Arcy Lameta, A., Jay, M. Study of soybean and lentil root exudates: 3. Influence of soybean isoflavonoids on<br />

the growth of rhizobia and some rhizosphere microorganisms. Plant Soil 1987; 101:269-272.


172<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

Dewick, P.M. Medicinal natural products: A biosynthetic approach. Chichester: John Wiley and Sons, 1997.<br />

Escalante, A.M., Santecchia, C.B., López, S.N., Gattuso, M.A., Ravelo, A.G., Monache, F.D., Sierra, M.G.,<br />

Zacchino, S.A. Isolation of antifungal saponins from Phytolacca tetramera, an Argentinean species in<br />

critic risk. J Ethnopharm 2002; 82:29-34.<br />

Evans, J.L., Gealt, M.A. The sterols of growth and stationary phases of Aspergillus nidulans cultures. J Gen<br />

Microbiol 1985; 131:279-284.<br />

Favel, A., Steinmetz, M.D., Regli, P., Vidal-Olivier, E., Elias, R., Balansard, G. In vitro antifungal activity of<br />

triterpenoid saponins. Planta Med 1994; 60:50-53.<br />

Ford, J.E., McChance, D.J., Drysdale, R.B. The detoxification of α-tomatine by Fusarium oxysporum f. sp.<br />

lycopersici. Phytochemistry 1977; 16:544-546.<br />

Fountain, M.S. Vegetation associated with natural populations of ginseng Panax quinquefolium in Arkansas<br />

USA. Castanea 1986; 51:42-48.<br />

Fuzzati, N., Gabetta, B., Jayakar, K., Pace, R. Peterlongo, F. Liquid chromatography-electrospray mass<br />

spectrometric identification of ginsenosides in Panax ginseng roots. J Chromatography A, 1999; 854:69-<br />

79.<br />

Grayer, R.J., Harborne, J.B. A survey of antifungal compounds from higher plants, 1982-1993. Phytochemistry<br />

1994; 37:19-42.<br />

Grayston, S.J., Griffith, G.S., Mawdsley, J.L., Campbell, C.D., Bardgett, R.D. Accounting for variability in soil<br />

microbial communities of temperate upland grassland ecosystems. Soil Biol Biochem 2001; 33:533-551.<br />

Green, H. Heiberg, N., Lejbolle, K., Jensen, D.F. The use of a GUS transformant of Trichoderma harzianum,<br />

strain T3a, to study metabolic activity in the spermosphere and rhizosphere related to biocontrol of Pythium<br />

damping-off and root rot. Eur J Plant Pathol 2001; 107:349-359.<br />

Griffiths, K.M., Bacic, A., Howlett, B.J. Sterol composition of mycelia of the plant pathogenic ascomycete<br />

Leptosphaeria maculans. Phytochemistry 2003; 62:147-153.<br />

Haralampidis, K., Trojanowska, M., Osbourn, A.E. Biosynthesis of triterpenoid saponins in plants. In Advances<br />

in Biochemical Engineering/Biotechnology. Scheper Th. ed., Berlin: Springer-Verlag, 2002; pp.31-49.<br />

Hartung, A.C., Stephens, C.T. Effects of allelopathic substances produced by asparagus on incidence and severity<br />

of asparagus decline due to Fusarium crown rot. J Chem Ecol 1983; 9:1163-1174.<br />

Helal, H.M., Sauerbeck, D.R. Influence of plant roots on C and P metabolism in soil. Plant Soil 1984; 76:<br />

175-182.<br />

Henderson, M.E.K., Farmer, V.C. Utilization by soil fungi of p-hydroxybenzaldehyde, ferulic acid, syringaldehyde<br />

and vanillin. J Gen Microbiol 1955; 12:37-46.<br />

Hong, S.C., Gray, A.B., Asiedu, S.K., Ju, H.Y. The evaluation of Trichoderma isolates, benomyl, and<br />

propiconazole against Cylindrocarpon destructans. Can J Plant Sci 2000; 80:231.<br />

Isutsa, D.K., Merwin, I.A. Malus germplasm varies in resistance or tolerance to apple replant disease in a<br />

mixture of New York orchard soils. Hortscience 2000; 35:262-268.<br />

Johnson, N.C., Tillman, D., Wedin, D. Plant and soil controls on mycorrhizal fungal communities. Ecology<br />

1992; 73:2034-2042.<br />

Jung, J.D., Park, H.W., Hahn, Y., Hur, C.G., In, D.S., Chung, H.J., Liu, J.R., Choi, D.W. Discovery of genes for<br />

ginsenoside biosynthesis by analysis of ginseng expressed sequence tags. Plant Cell Reports 2003; 22:224-<br />

230.<br />

Keukens, E.A.J., De Vrije, T., Fabrie, C.H.J.P., Demel, R.A., Jongen, W.M.F., De Kruijff, B. Dual specificity of<br />

sterol-mediated glycoalkaloid induced membrane disruption. Biochim Biophys Acta 1992; 1110:127-<br />

136.<br />

Keukens, E.A.J., De Vrije, T., Van Den Boom, C., De Waard, P., Plasman, H.H., Thiel, F., Chupin, V., Jongen,<br />

W.M.F., De Kruijff, B. Molecular basis of glycoalkaloid induced membrane disruption. Biochim Biophys<br />

Acta 1995; 1240:216-228.<br />

Kushiro, T., Shibuya, M. Ebizuka, Y. beta-Amyrin synthase: Cloning of oxidosqualene cyclase that catalyzes<br />

the formation of the most popular triterpene among higher plants. Eur J Biochem 1998; 256:238-244.<br />

Lagrange, H., Jay-Allemand, C., Lapeyrie, F. Rutin, the phenolglycoside from Eucalyptus root exudates,<br />

stimulates Pisolithus hyphal growth at picomolar concentrations. New Phytol 2001; 150:349-355.<br />

Larkin, R.P., Hopkins, D.L., Martin, F.N. Effect of successive watermelon plantings on Fusarium oxysporum<br />

and other microorganisms in soils suppressive and conducive to Fusarium wilt of watermelon.<br />

Phytopathology 1993; 83:1097-1105.


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 173<br />

Latijnhouwers, M., de Wit, P.J.G.M., Govers, F. Oomycetes and fungi: similar weaponry to attack plants. Trends<br />

Microbiol, 2003; 11:462-469.<br />

Lawrence, G.H.M. Taxonomy of Vascular Plants. Macmillan Press, New York, 1951.<br />

Levy, M., Zehavi, U., Naim, M., Itzhack, P. An improved procedure for the isolation of medicagenic acid 3-Oß-D-glucopyranoside<br />

from alfalfa roots and its antifungal activity on plant pathogens. J Agric Food Chem<br />

1986; 34:960-963.<br />

Levy, M., Zehavi, U., Naim, M., Polacheck, I. Isolation, structure determination, synthesis, and anti-fungal<br />

activity of a new native alfalfa-root saponin. Carbohydr Res 1989; 193:115-123.<br />

Lewis, W.H., Zenger, V.E. Population dynamics of the American ginseng Panax quinquefolium (Araliaceae).<br />

Am J Bot 1982; 69:1483-1490.<br />

Li, T.S.C., Mazza, G., Cottrell, A.C., Gao, L. Ginsenosides in roots and leaves of American ginseng. J Agric<br />

Food Chem 1996; 44:717-720.<br />

Maizel, J.V., Burkhardt, H.J., Mitchell, H.K. Avenacin, an antimicrobial substance isolated from Avena sativa.<br />

I. Isolation and antimicrobial activity. Biochemistry 1964; 3:424-426.<br />

Marston, A., Gafner, F., Dossaji, S.F., Hostettmann, K. Fungicidal and molluscicidal saponins from Dolichos<br />

kilimandscharicus. Phytochemistry 1988; 27:1325-1326.<br />

Mazzola, M. Elucidation of the microbial complex having a causal role in the development of apple replant<br />

disease in Washington. Phytopathology 1998; 88:930-938.<br />

Mazzola, M. Transformation of soil microbial community structure and Rhizoctonia-suppressive potential in<br />

response to apple roots. Phytopathology 1999 89:920-927.<br />

Miethling, R., Wieland, G., Backhaus, H., Tebbe, C.C. Variation of microbial rhizosphere communities in response<br />

to crop species, soil origin, and inoculation with Sinorhizobium meliloti L33. Microbial Ecology 2000;<br />

40:43-56.<br />

Mikes, V., Milat, M.L., Ponchet, M., Ricci, P., Blein, J.P. The fungal elicitor cryptogein is a sterol carrier protein.<br />

FEBS Lett 1997; 416:190-192.<br />

Miller, H.J., Henken, G., van Veen, J.A. Variation and composition of bacterial populations in the rhizosphere of<br />

maize, wheat and grass cultivars. Can J Microbiol 1989; 35:656-660.<br />

Morrissey, J.P., Osbourn, A.E. Fungal resistance to plant antibiotics as a mechanism of pathogenesis. Microbiol.<br />

Mol Biol Rev 1999; 63:708-724.<br />

Narasimhan, K., Basheer, C., Bajic, V.B., Swarup, S. Enhancement of plant-microbe interactions using a<br />

rhizosphere metabolomics-driven approach and its application in the removal of polychlorinated biphenyls.<br />

Plant Physiol 2003; 132:146-153.<br />

Nes, D.W. Biosynthesis and requirement for sterols in the growth and reproduction of Oomycetes. In Ecology<br />

and Metabolism of Plant Lipids. Fuller, G., Nes, W.D. ed., American Chemical Society, Washington,<br />

DC. 1987; pp. 304-328.<br />

Nicol, R.W., Yousef, L., Traquair, J.A., Bernards, M.A. Ginsenosides stimulate the growth of soilborne pathogens<br />

of American ginseng. Phytochemistry 2003; 64:257-264.<br />

Nicol, R.W., Traquair, J.A., Bernards, M.A. Ginsenosides as host resistance factors in American Ginseng (Panax<br />

quinquefolius). Can J Bot 2002; 80:557-562.<br />

Ohtani, K., Mavi, S., Hostettmann, K. Molluscicidal and antifungal triterpenoid saponins from Rapanea<br />

melanophloeos leaves. Phytochemistry 1993; 33:83-86.<br />

Olsen, R.A. Triterpene glycosides as inhibitors of fungal growth and metabolism. 1. effect on growth, endogenous<br />

respiration and leakage of UV-absorbing material from various fungi. Physiol Plant 1971; 24:534-543.<br />

Olsen, R.A. Triterpene glycosides as inhibitors of fungal growth and metabolism. 5. Role of sterol contents of<br />

some fungi. Physiol Plant 1973a ; 28:507-515.<br />

Olsen, R.A. Triterpene glycosides as inhibitors of fungal growth and metabolism. 6. The effect of aescin on<br />

fungi with reduced sterol contents. Physiol Plant 1973b ; 29:145-149.<br />

OMAF. Ontario ginseng exports, 1996-2002. Queen’s Printer for Ontario, 2003.<br />

Osbourn, A.E., Wubben, J.P., Daniels, M.J. Saponin detoxification by phytopathogenic fungi. In Plant-Microbe<br />

Interactions Volume 2. Stacey, G., Keen, N.T. ed., New York: Chapman and Hall, 1995a, pp. 99-124.<br />

Osbourn, A. Bowyer, P., Lunness, P., Clarke, B., Daniels, M. Fungal pathogens of oat roots and tomato<br />

leaves employ closely related enzymes to detoxify different host plant saponins. Mol Plant Microb Interact<br />

1995b ; 8:971-978.<br />

Osbourn, A.E. Preformed antimicrobial compounds and plant defense against fungal attack. Plant Cell 1996;<br />

8:1821-1831.


174<br />

MARK A. BERNARDS, LINA F. YOUSEF, ROBERT W. NICOL<br />

Osbourn, A.E. Saponins in cereals. Phytochemistry 2003; 62:1-4.<br />

Ouf, S.A., Hady, F.K., Abdel, E.M.H., Shaker, K.H. Isolation of antifungal compounds from some Zygophyllum<br />

species and their bioassay against two soil-borne plant pathogens. Folia Microbiol 1994; 39:215-221.<br />

Panabières, F., Ponchet, M., Allasia, V., Cardin, L., Ricci, P. Characterization of border species among Pythiaceae:<br />

several Pythium isolates produce elicitins, typical proteins from Phytophthora spp. Mycol Res 1997;<br />

101:1459-1468.<br />

Papadopoulou, K., Melton, R.E., Leggett, M. Daniels, M.J., Osbourn, A.E. Compromised disease resistance in<br />

saponin-deficient plants. Proc Natl Acad Sci USA 1999; 96:12923-12928.<br />

Papavizas, G.C. Survival of Trichoderma harzianum in soil and in pea and bean rhizosphere. Phytopathology<br />

1982; 71:121-125.<br />

Pedersen, C.T., Safir, G.R., Siqueira, J.O., Parent, S. Effect of phenolic compounds on asparagus mycorrhizae.<br />

Soil Biol Biochem 1991; 23:491-494.<br />

Pedras, M.S.C., Khan, A.Q. Biotransformation of the phytoalexin camalexin by the phytopathogen Rhizoctonia<br />

solani. Phytochemistry 2000a ; 53:59-69.<br />

Pedras, M.S.C., Okanga, F.I., Zaharia, I.L., Khan, A.Q. Phytoalexins from crucifers: synthesis, biosynthesis,<br />

and biotransformation. Phytochemistry 2000b; 53:161-176.<br />

Pegg, G.F., Woodward, S. Synthesis and metabolism of alpha tomatine in tomato Lycopersicon esculentum<br />

isolines in relation to resistance to Verticillium albo-atrum. Physiol Mol Plant Pathol 1986; 28:187-202.<br />

Peirce, L.C., Colby, L.W. Interaction of asparagus root filtrate with Fusarium oxysporum f.sp. asparagi. J Am<br />

Soc Hort Sci 1987; 112:35-40.<br />

Peters, N.K., Long, S.R. Alfalfa root exudates and compounds which promote or inhibit induction of Rhizobium<br />

meliloti nodulation genes. Plant Physiol 1988; 88:396-400.<br />

Proctor J.T.A., Bailey, W.G. Ginseng, industry botany and culture. Hort Rev 1987; 9:187-236.<br />

Punja, Z.K. Fungal pathogens of American ginseng (Panax quinquefolium) in British Columbia. Can J Plant<br />

Pathol 1997; 19:301-306.<br />

Quidde, T., Osbourn, A.E., Tudzynski, P. Detoxification of alpha-tomatine by Botrytis cinerea. Physiol Mol<br />

Plant Pathol 1998; 52:151-165.<br />

Rahouti, M. Seigle-Murandi, F., Steiman, R., Eriksson, K.E. Metabolism of ferulic acid by Paecilomyces variotii<br />

and Pestalotia palmarum. Appl Env Microbiol 1989; 55:2391-2398.<br />

Reeleder, R.D., Brammall, R.A. Pathogenicity of Pythium species, Cylindrocarpon destructans, and Rhizoctonia<br />

solani to ginseng seedlings in Ontario. Can J Plant Pathol 1995; 16:311-316.<br />

Reeleder, R.D., Roy, R., Capell, B. Seed and root rots of ginseng (Panax quinquefolius) caused by<br />

Cylindrocarpon destructans and Fusarium spp. Phytopathology 1999; 89:S65.<br />

Rice, E.L. Allelopathy. Academic Press Inc., New York, 1984.<br />

Roddick, J.G. Complex formation between solanaceous steroidal glycoalkaloids and free sterols in vitro.<br />

Phytochemistry 1979; 18:1467-1470.<br />

Roldán-Arjona, T., Pérez-Espinosa, A., Ruiz-Rubio, M. Tomatinase from Fusarium oxysporum f. sp. lycopersici<br />

defines a new class of saponinases. Mol Plant Microb Interact 1999; 12:852-861.<br />

Sandrock, R.W., Van Etten, H.D. Fungal sensitivity to and enzymatic degradation of the phytoanticipin a-tomatine.<br />

Phytopathology 1998; 88:137-143.<br />

Schlessman, M.A. Floral biology of American ginseng (Panax quinquefolium). Bull Torrey Bot Club 1985;<br />

112:129-133.<br />

Schmidt, S.K., Ley, R.E. Microbial competition and soil structure limit the expression of allelochemicals in<br />

nature. In Principles and Practices in Plant Ecology: Allelochemical Interactions, Inderjit, K.M.M.,<br />

Dakshini, Foy, C.L. ed. Boca Raton: CRC Press, 1999, pp. 339-351.<br />

Scow, K.M. Soil microbial communities and carbon flow in agroecosystems. In Ecology in Agriculture. Jackson,<br />

L.E. ed. Academic Press, N.Y. 1997, pp. 367-413.<br />

Shepherd, T. The microchemistry of plant/microorganism interactions. In Ecology of Plant Pathogens. Blakeman,<br />

J.P., Williamson, B. ed.,Wallingford: CAB International, 1994, pp. 39-62.<br />

Shin, H.S., Lee, M.W. Environmental factors and the distribution of soil microorganisms in ginseng field. Kor J<br />

Microbiol 1986; 24:184-193.<br />

Singh, H.P., Batish, D.R., Kohli, R.K. Autotoxicity: concepts, organisms, and ecological significance. Crit Rev<br />

Plant Sci 1999; 18:757-772.<br />

Small, E., Catling, P.M. Canadian medicinal crops. NRC Research Press, Ottawa, ON, 1999.


THE ALLELOPATHIC POTENTIAL OF GINSENOSIDES 175<br />

Sparling, G.P., Ord, B.G., Vaughan, D. Microbial biomass and activity in soils amended with glucose. Soil Biol<br />

Biochem 1981; 13:99-104.<br />

Steel, C.C., Drysdale, R.B. Electrolyte leakage from plant and fungal tissues and disruption of liposome<br />

membranes by a-tomatine. Phytochemistry 1988; 27:1025-1030.<br />

Straney, D.C., Van Etten, H.D. Characterization of the PDA1 promoter of Nectria haematococca and<br />

identification of a region that binds a pisatin-responsive DNA binding factor. Mol Plant Microb Interact<br />

1994; 7:256-266.<br />

Suleman, P., Tohamy, A.M., Saleh, A.A., Madkour, M.A., Straney, D.C. Variation in sensitivity to tomatine and<br />

rishitin among isolates of Fusarium oxysporum f. sp. lycopersici, and strains not pathogenic on tomato.<br />

Physiol Mol Plant Pathol 1996; 48:131-144.<br />

Szabo, K., Wittenmayer, L. Plant specific root exudations as possible cause for specific replant diseases in<br />

Rosaceen. J Appl Bot 2000; 74:191-197.<br />

Takechi, M., Tanaka, Y. Structure-activity relationships of the saponin α-hederin. Phytochemistry 1990; 29:451-<br />

452.<br />

Tang, C.S., Young, C.S. Collection and identification of allelopathic compounds from the undisturbed root<br />

system of bigalta limpograss (Hemarthria altissima). Plant Physiol 1982; 69:155-160.<br />

Tani, T., Kubo, M., Katsuki, T., Higashino, M., Hayashi, T., Arichi, S. Histochemistry II. Ginsenosides in<br />

ginseng (Panax ginseng root). J Nat Prod 1981; 44:401-407.<br />

Turner, E.M.C. The nature of the resistance of oats to the take-all fungus. II. Inhibition of growth and respiration<br />

of Ophiobolus graminis Sacc. and other fungi by a constituent of oat sap. J Exp Bot 1956; 7:80-92.<br />

Utkhede, R.S., Sholberg, P.L., Smirle, M.J. Effects of chemical and biological treatments on growth and yield of<br />

apple trees planted in Phytophthora cactorum infested soil. Can J Plant Pathol 2001; 23:163-167.<br />

Van Etten, H.D., Mansfield, J.W., Bailey, J.A., Farmer, E.E. Two classes of plant antibiotics: phytoalexins<br />

versus “phytoanticipins”. Plant Cell 1994; 6:1191-1192.<br />

Van Etten, H.D., Sandrock, R.W., Wasmann, C.C., Soby, S.D., McCluskey, K., Wang, P. Detoxification of<br />

phytoanticipins and phytoalexins by phytopathogenic fungi. Can J Bot 1995; 73:518-525.<br />

Wardle, D.A. A comparative assessment of factors which influence microbial biomass carbon and nitrogen<br />

levels in soil. Biol Rev 1992; 67:321-358.<br />

Weete, J.D. Structure and function of sterols in fungi. Adv Lipid Res 1989; 23:115-167.<br />

Weltring, K.M., Wessels, J., Geyer, R. Metabolism of the potato saponins α-chaconine and α-solanine by<br />

Gibberella pulicaris. Phytochemistry 1997; 46:1005-1009.<br />

Zimmer, D.E., Pedersen, M.W., McGuire, D.F. A bioassay for alfalfa saponins using the fungus Trichoderma<br />

viride Pers. ex Fr. Crop Sci 1967; 7:223-224.


HIROYUKI NISHIMURA and ATSUSHI SATOH<br />

ANTIMICROBIAL AND NEMATICIDAL<br />

SUBSTANCES FROM THE ROOT OF CHICORY<br />

(Cichorium intybus)<br />

Department of Biosciences and Technology, School of Engineering,<br />

Hokkaido Tokai University, Sapporo 005-8601, Japan<br />

Email: nishimura@db.htokai.ac.jp<br />

Abstract. Antimicrobial sesquiterpenoids, 8a-angeloyloxycichoralexin and guaianolides such as cichoralexin<br />

and 10a-hydroxycichopumilide were isolated and identified from the root of chicory (Cichorium intybus). These<br />

sesquiterpenoids exhibited antifungal activities against Pyricularia oryzae, Pellicularia sasaki and Alternaria<br />

kikuchiana. Ether soluble phenolics from the chicory root were found to exhibit nematicidal activity. The<br />

addition of dry chicory root powder to noodles, a boiled fish paste, and cocoa- and coffeecakes, during food<br />

processing, provided some protection from food spoilage organisms, and this product may have value as a<br />

natural food preservative.<br />

1. INTRODUCTION<br />

Green plants produce numerous secondary compounds, that are not involved in primary<br />

metabolism (Kaur et al., 2000). The essential role that secondary metabolites,<br />

such as terpenoid and phenolic compounds, play in complex interactions among living<br />

organisms in the natural environment is gradually being determined. Although<br />

some of these natural products have been found to be pollinator or feeding attractants,<br />

many of them seem to function as chemical weapons against insects, pathogenic organisms,<br />

and competing plants (Whittaker and Feeny, 1971; Seister, 1977; Berenhaum,<br />

1995; Inderjit and Duke, 2003). Such secondary metabolites may take part in plantanimal,<br />

plant-microorganism and plant-plant interactions, and are generally termed<br />

allelochemicals.<br />

Chicory (Cichorium intybus) is cultivated in cool regions such as Northern Europe.<br />

Recently, this vegetable has arisen out of claims that it is able to promote “good<br />

health” since no pesticides are used to cultivate chicory in the field, while the plant<br />

remains noticeably free from herbivore and microbial attack. The bitter substances,<br />

lactupicrin, 8-deoxylactucin and some phenolics had previously been shown to pos-<br />

sess insect antifeedant properties in chicory (Rees and Harborne, 1985). Specifically,<br />

sesquiterpenoid lactones from chicory leaves, such as 8-deoxylactucin and lactupicrin<br />

(Figure 1), were identified as insect antifeedants against desert locust, Schistocerca<br />

gregaria. Similarly, we found some biologically active secondary metabolites in the<br />

177<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 177– 180.<br />

© 2006 Springer. Printed in the Netherlands.


178<br />

HIROYUKI NISHIMURA AND ATSUSHI SATOH<br />

chicory roots and reported their activity against microorganisms and small animals<br />

living in soil (Nishimura et al., 2000). In addition, chicory dry-powder has a potential<br />

application in preserving processed foods.<br />

HO<br />

a : R = H, 8-deoxylactucin<br />

b : R = OCOCH 2 C 6 H 4 OH(p), lactupicrin<br />

2. ANTIPATHOGENIC SUBSTANCES<br />

2.1. Antimicrobial Substances in Chicory Root<br />

O<br />

H<br />

O<br />

Figure 1. Insect antifeedants in chicory leaves. Source: Nishimura et al. (2000). Reproduced<br />

after permission from Kluwer Academic Publisher (Springer).<br />

White shoots germinated from cultivated chicory roots have been used for French and<br />

Italian dishes for a long time. We noticed that fungi were not observed on the chicory<br />

root in spite of being placed in moist conditions favorable for the development of rots.<br />

We found that chicory root extracts, recovered with hexane or ether, exhibited antifungal<br />

activities against Cladosporium herbarum, Pyricularia oryzae, Pellicularia sasaki<br />

and Alternaria kikuchiana. The ether extract was fractionated by SiO 2 column<br />

chromatography. Rechromatography in a SiO 2 column and HPLC with a CHCl 3 -MeOH<br />

solvent system revealed three active compounds. Two of these were sesquiterpenoids<br />

(C-1 and C-2) identified as cichoralexin (Monde et al., 1990) and 10αhydroxycichopumilide,<br />

respectively, by the interpretation of spectral data. From the<br />

interpretation of its spectral data, the third compound (C-3) was identified as 8αangeloyloxy-cichoralexin<br />

(Figure 2) (Nishimura et al., 2000). These three<br />

sesquiterpenoids, cichoralexin, 10α-hydroxycichopumilide and 8αangeloyloxycichoralexin<br />

from the chicory root were found to have significant<br />

antimicrobial properties.<br />

O<br />

H<br />

R


ANTIMICROBIAL AND NEMATICIDAL SUBSTANCES FROM CHICORY ROOT 179<br />

1.97 (dd)<br />

O<br />

1.11 (d)<br />

2.15 (m)<br />

H3C 1.79 (m)<br />

H H<br />

H<br />

2.16 (m)<br />

H<br />

5.20 (dt)<br />

O<br />

5.95 (s) H<br />

4.00 (t)<br />

H<br />

H<br />

O<br />

H<br />

H 2.18 (ddd)<br />

2.28 (s) H3C 2.87 (dd)<br />

O<br />

H 2.60 (dq)<br />

1.40 (d)<br />

2.2. Nematicidal Substances in Chicory Root<br />

O<br />

We found that the ether and ethyl acetate extracts of chicory roots exhibited the best<br />

nematicidal activities. The extracts were separated according to their acidity to give<br />

organo-acidic, phenolic, basic and neutral fractions. The phenolic fraction was found<br />

to have the highest activity.<br />

Figure 3. The relative survival ratios (%) of the golden nematode (Globodera rostochiensis)<br />

following exposure to chicory plant extracts recovered from the skin, inner part, wounded part<br />

and hairy roots of chicory and from the acidic fraction of neighboring soil.<br />

Source: Nishimura et al. (2000). Reproduced with permission from Kluwer Academic Publisher<br />

(Springer).<br />

CH 3<br />

CH 3<br />

CH 3<br />

H<br />

1.91 (m)<br />

2.01 (dq)<br />

6.17 (qq)<br />

Figure 2. Chemical structure of 8α-angeloyloxycichoralexin. Source: Nishimura et al. (2000).<br />

Reproduced after permission from Kluwer Academic Publisher (Springer).<br />

Hairy roots<br />

R. S. R. =91.0%<br />

Acidic fraction<br />

from rhizoplane<br />

soil<br />

R. S. R. =53.5%<br />

Acidic fraction from<br />

bulk field soil<br />

R. S. R. =100.0%<br />

R. S. R. = Relative Survival Ratio<br />

Inner part<br />

R. S. R. =41.7%<br />

Wounded part<br />

R. S. R. =78.6%<br />

Rhizome skin<br />

R. S. R. =37.5%


180<br />

HIROYUKI NISHIMURA AND ATSUSHI SATOH<br />

The nematicidal activities of extracts from different parts of a chicory root were<br />

examined. The relative survival ratios (R. S. R.) at the root skin, inner root tissue,<br />

wounded root tissue and hairy roots were measured (Figure 3). The extract from the<br />

root skin exhibited the highest nematicidal activity. Interestingly, the nematicidal<br />

activity of an acidic fraction from rhizoplane soil was much higher than that from<br />

bulk field soil. We found that some phenolics from the root also exhibited nematicidal<br />

activity. Thus, it seems that secondary metabolites such as terpenoids and phenolics<br />

can play an important role in chicory defense.<br />

3. UTILIZATION OF ALLELOCHEMICALS IN PROCESSED FOODS<br />

Traditional Japanese foods - noodles, a boiled fish paste, and cocoa- and coffee cakes<br />

- were cooked without chicory dry powder. After five days, food spoilage fungi such<br />

as Mucor and Rhizopus species were observed growing on the test foods. Interestingly,<br />

when 0.1 – 5% (w/w) of the dried chicory powder was added to the processed<br />

foods, no fungi were observed even after ten days storage. It seems that antimicrobial<br />

sesquiterpenoids in chicory root can inhibit the growth of spoilage fungi.<br />

4. REFERENCES<br />

Berenhaum M.R. Turnabout is fair play: secondary roles for primary compounds. J Chem Ecol 1995; 21:<br />

925-940.<br />

Inderjit, Duke S.O. Ecophysiological aspects of allelopathy. Planta 2003; 217:529-539.<br />

Kaur H., Inderjit, Keating K.I. Do allelochemicals operate independent of substratum factors? In: Chemical<br />

ecology of plants: allelopathy in aquatic and terrestrial ecosystems. Inderjit, Mallik AU eds. Birkhauser-<br />

Verlag AG, Basal, Switzerland, 2002; pp 99-107.<br />

Monde, K., Oya, T., Shirata, A., Takasugi, M. A guaianolide phytoalexin, cichoralexin, from Cichorium intybus.,<br />

Phytochemistry 1990; 29:3449-3451.<br />

Nishimura, H., Kondo, Y., Nagasaka, T., Satoh, A. <strong>Allelochemicals</strong> in chicory and utilization in processed<br />

foods, J Chem Ecol 2000; 26:2233-2241.<br />

Rees, S.B., Harborne, J.B. The role of sesquiterpene lactones and phenolics in the chemical defense of the<br />

chicory plant. Phytochemistry 1985; 24:2225-2231.<br />

Seigler D.S. Primary roles for secondary compounds. Biochem Syst Ecol 1977; 5:195-199<br />

Whittaker R.H., Feeny P.P.Allelochemics: chemical interactions between species. Science 1971; 171: 757-770.


REN-SEN ZENG<br />

DISEASE RESISTANCE IN PLANTS THROUGH<br />

MYCORRHIZAL FUNGI INDUCED<br />

ALLELOCHEMICALS<br />

Chemical Ecology Lab., Institute of Tropical & Subtropical Ecology,<br />

Agricultural College, South China Agricultural University Wushan, Tianhe<br />

District, Guangzhou, 510642 P.R. China<br />

Email: rszeng@scau.edu.cn; zengrs8@hotmail.com<br />

Abstract. Allelochemals induced in mycorrhizal plants play an important role in disease resistance. Mycorrhizal<br />

associations are the most important symbiosis systems in terrestrial ecosystems and offer many benefits to the<br />

host plant. Arbuscular mycorrhizal associations can reduce damage caused by soil and root - borne pathogens.<br />

1. INTRODUCTION<br />

As plants are non-motile orgnaisms and lack the sophisticated immune system that<br />

animals have, they had to develop their own defence system against pathogens and<br />

predators, and systems to lure motile creatures for fertilization and dissemination.<br />

Plants are capable of synthesizing an overwhelming variety of low-molecular-weight<br />

organic compounds, called secondary methbolities, many with very unique and complex<br />

structures. More than 100,000 different secondary metabolites have been found from<br />

plants (Buckingham, 1998; Dixon, 2001), with many more yet to be discovered;<br />

estimated of the total number in plants exceed 500,000 (Mendelsohn and Balick,<br />

1995). Many of these compounds are differentially distributed among limited taxonomic<br />

groups within the plant kingdom and, conversely, each plant species has a distinct<br />

profile of secondary metabolites.<br />

Many of these secondary products are bactericidal, fungicidal, repellent, or even<br />

poisonous to herbivores. Plant secondary metabolites have for thousands of years<br />

played an important role in medicine, crop disease and weed control. Nowadays, they<br />

are used either directly or after chemical modification. Examples of these compounds<br />

include flavonoids, phenols and phenolic glycosides, unsaturated lactones, sulphur<br />

compounds, saponins, cyanogenic glycosides and glucosinolates (Go’mez Garibay et<br />

al., 1990; Bennett and Wallsgrove, 1994; Grayer and Harborne, 1994).<br />

Plant diseases cause serious losses in crop production. Pesticide application is<br />

currently the primary way to control crop disease, but it has raised an array of<br />

environmental problems. Achieving sustainable agriculture will require avoiding a<br />

181<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 181– 192.<br />

© 2006 Springer. Printed in the Netherlands.


182<br />

REN-SEN ZENG<br />

heavy reliance on pesticides. Enhancing crop resistance against diseases and herbivores<br />

is an ideal approach to reduce pesticide application.<br />

Plants possess both constitutive and inducible chemical defense mechanisms.<br />

Before pathogen infection, plants may contain significant amounts of constitutive<br />

secondary metabolites including phenolics, terpenoids, and steroids, which are toxic<br />

to invading organisms (Levin, 1976; Mauch-Mani and Metraux, 1998). Plants may<br />

also activate their production of certain defensive chemicals after pathogen infection.<br />

These inducible defense compounds are usually only produced and accumulated after<br />

specific recognition of the invading organism, and are known as phytoalexins (Dixon,<br />

1986; Hammerschmidt, 1999). Plants can acquire induced resistance to pathogens<br />

after infection with necrotrophic attackers, nonpathogenic root-colonizing<br />

pseudomonads, salicylic acid, beta-aminobutyric acid and many other natural or<br />

synthetic compounds (Conrath et al., 2002; Benhamou, 1996).<br />

Mycorrhizal fungi provide an effective alternative method of disease control<br />

especially for those pathogens which effect the below ground plant parts. In mycorrhizal<br />

fungi lies enormous potential for use as biological control agent for soil-borne diseases<br />

as the root diseases are of the most difficult to manage and lead to losses in disturbing<br />

proportions.<br />

The mycorrhizal symbiosis substantially influence plant growth under a variety<br />

of stressful conditions and their role in biological control of soil/root - borne pathogens<br />

is of immense importance both in the agricultural system as well as in the forestry<br />

(Linderman, 1994).<br />

Mycorrhizal associations increase growth and yield of many crop plants by<br />

enhancing nutrient uptake, resistance to drought and salinity and increases tolerance<br />

to pathogens (Gianinazzi-Pearson, 1996; Mukerji, 1999; Singh et al., 2000; Ludwig-<br />

Müller, 2004). Of the seven types of mycorrhiza known (Srivastava et al., 1996; Mukerji<br />

et al., 1997; Raina et al., 2000; Redecker et al., 2000), ectomycorrhiza and vesiculararbuscular<br />

mycorrhiza (VAM) are more important in agriculture and forestry.<br />

Ecotmyocrrhizal associations are more prevalent in temperate and sub-temperate<br />

regions, while VAM/AM associations are common features of tropical and substropical<br />

regions of the world. During colonization, distinct structures are formed by the<br />

arbuscular mycorrhizal (AM) fungi, with in the host roots - internal hyphae, arbuscules<br />

and vesicles (Walker, 1992). The complex cellular relationship between host roots<br />

and AM fungi requires a continuous exchange of signals, for proper development of<br />

mycorrhiza in the roots of a host plant (Gianinazzi-Pearson, 1996). Plant hormones<br />

may be a suitable candidate for the regulation of such a symbiosis. There is little<br />

information about the function of plant harmones during the colonization process<br />

although there is evidence that they are involved in signaling events between AM<br />

fungi and host plants (Barker and Tagu, 2000; Ludwig - Müller, 2000a,b). In addition,<br />

it has been suggested that phytohormones, such as IAA and cytokinins, released by<br />

mycorrhizal fungi may also contribute to the enhancement of plant growth.<br />

(Frankenberger Jr. and Arsad, 1995).<br />

This review describes the role of mycorrhizal associations in disease control.


DISEASE RESISTANCE IN PLANTS THROUGH<br />

ALLELOCHEMICALS 183<br />

2. MYCORRIZA IN DISEASE RESISTANCE<br />

Safir (1968) found that inoculation of onion with Glomus mosseae could significantly<br />

reduce pink root disease due to Pyrenochaeta terrestris. Later studies indicate that<br />

AM fungi can induce resistance or increase tolerance to some root-borne pathogens<br />

(Azcon - Aguilar and Barea, 1996, 1997; Caron et al., 1986; Caron, 1989; Cordier et<br />

al., 1997; Dehne, 1982; Hooker et al., 1994; Trotta et al., 1996). Glomus mosseae<br />

protected peanut plants from infection by pod rot fungal pathogens Fusarium solani<br />

and Rhizoctonia solani (Abdalla and Abdel-Fateh, 2000).<br />

The Glomus intraradices increased P uptake and reduced disease development<br />

of Aphanomyces euteiches in pea roots (Bodker et al., 1998) Mycorrhization with<br />

Glomus mosseae and G. intraradices induced local or systemic resistance to<br />

Phytophthora parasitica in tomato roots (Cordier et al., 1996, 1998; Pozo et al.,<br />

2002). Decreased pathogen development in mycorrhizal and non-mycorrhizal parts<br />

of inoculated roots is associated with accumulation of phenolics and plant cell defense<br />

response. The protective effects induced by AM fungi against a phytoplasma is reported<br />

in tomato (Lingua et al., 2002).<br />

AM protects an annual grass from root pathogenic fungi in the field (Newsham<br />

et al., 1995). Inoculation of onion with Glomus sp. Zac-19 delayed onion white rot<br />

epidemics caused by Sclerotium cepivorum Berk by two weeks and increased the<br />

yield by 22% under field conditions (Torres-Barragan et al., 1996). Diospyros lotus<br />

inoculated with Glomus mosseae, Glomus intraradices, Glomus versiforme<br />

significantly increased the plant growth and decreased the disease caused by Cercospora<br />

kaki under field conditions. The AM fungal inoculum even suppressed postharvest<br />

disease of potato dry rot (Fusarium sambucinum) in prenuclear minitubers (Niemira<br />

et al., 1996).<br />

Root rot caused by Fusarium solani significantly contributes to crop yield decline,<br />

up to 50%. The inoculation of common bean (Phaseolus vulgaris) with Glomus<br />

mosseae, besides decreasing propagule number of F. solani in the rhizosphere,<br />

decreased root rot by 34 to 77% (Dar et al., 1997). In the presence of the root nodulating<br />

symbiont Rhizobium leguminosarum, mycorrhizal inoculated plants were more tolerant<br />

to the fungal root pathogen. This indicates that interactions between mycorrhizal<br />

fungi and other rhizosphere microbes might have greater effects on soil-borne<br />

pathogens than mycorrhizal fungi alone. Davis and Menge (1980) found that Glomus<br />

fasciculatum reduced Phytophthora root rot of citrus at low level of soil phosphorus<br />

but had no effect in high phosphorus soil. The VAM fungi has also been employed as<br />

biocontrol agents for Macrophomina root rot of cowpea and Fusarium wilt of tomato<br />

(Ramaraj et al., 1988). The understanding of the mechanisms of plant disease resistance<br />

in mycorrhizal plants would provide better directions towards the development of<br />

efficient crop production and sustainable agriculture.<br />

3. MECHANISM OF DISEASE CONTROL BY MYCORRHIZAL FUNGI<br />

The mycorrhizal symbiosis involves several mechanisms in control of plant diseases.<br />

(i) Creating a mechanical barrier for the pathogen penetration and subsequent spread


184<br />

REN-SEN ZENG<br />

as in the case of sheathing mycorrhiza. In ectomycorrhizal associations forming highly<br />

interwoven mycelial network cover on the root (fungal mantle) and internally with<br />

cortical cells whose cell walls are surrounded by fungal hyphae i.e., hartigs’ net<br />

(Ingham, 1991; Maronek, 1981).<br />

(ii) Thickening of cell wall through lignifications and production of other<br />

polysaccharides which in turn hinder the entry of root pathogen (Dehne and Schonbeck,<br />

1979a,b).<br />

(iii) Stimulating the host roots to produce and accumulate sufficient concentrations<br />

of metabolities (terpenes, phenols etc.) which impart resistance to the host tissue<br />

against pathogen invasion (Krupa et al., 1973; Sampangi, 1989).<br />

(iv) Stimulating flavonolic wall infusions as in the case of Laccaria bicolor which<br />

prevented lesion formation by the pathogen Fusarium oxgsporum in roots of Douglas<br />

fir (Strobel and Sinclair, 1991).<br />

(v) Increasing the concentraton of orthodihydroxy phenols in roots which deter<br />

the activity of pathogens (Krishna and Bagyarj, 1984).<br />

(vi) Producing antifungal and antibacterial antibiotics and toxins that act against<br />

pathogenic organisms (Marx, 1972).<br />

(vii) Competing with the pathogens for the uptake of essential nutrients in the<br />

rhizosphere and the root surface (Reid, 1990).<br />

(viii) Stimulating the microbial activity and competitions in the root zone<br />

(rhizosphere, rhizoplane) and thus preventing the pathogen to get access to the roots<br />

(Rambelli, 1973; Singh and Mukerji, 2005). Roots colonized by VAM/AM fungi may<br />

also harbour more actinomycetes antagonistic to root pathogens (Secilia and Bagyaraj,<br />

1987; Dixon, 2000; Bansal et al., 2000; Singh et al., 2000; Mukerji, 2002).<br />

(ix) Compensating the nutrient absorption system from damage to roots by<br />

pathogens (Simth and Reid, 1997).<br />

(x) Changing the amount and type of plant root exudates. Pathogens dependent<br />

on certain exudates will be at a disadvantage as the exudates change (Matsubara et<br />

al., 1995; Newsham et al., 1995).<br />

4. PLANT DEFERENCE REACTIONS<br />

There is considerable evidence for the role of VAM/AM fungi in control of root<br />

pathogens (St-Arnaud et al., 1995, 1997; Mukerji, 1999). In AM associations some<br />

plant resistance marker molecules are formed in the tissue which interfere with the<br />

pathogens to attack (Gianinazzi-Pearson et al., 1996). These are phytoalexins, callose,<br />

peroxidase, chitinase, β-1,3 glucanase, PR-1 protein. Plants which constitutively overexpress<br />

defence-related genes provide interesting material for determing how AM<br />

fungi contend with plant defence, or their modifications may occur in the expression<br />

of other genes in such plants is still unclear.<br />

4.1. Mycorrhiza induced secondary Metabolites<br />

In AM plants, accumulation of secondary metabolites has been reviewed (Azcon-<br />

Aguilar and Barea, 1996; Morandi, 1996; Vierheilig et al., 1998; Mukerji, 1999).


DISEASE RESISTANCE IN PLANTS THROUGH<br />

ALLELOCHEMICALS<br />

Phytoalexins are produced de novo in plants in response to infection or attempted<br />

invasion by microbial pathogens or treatment with abiotic agents (Harborne and<br />

Ingham, 1978). They are low molecular weight, anti-microbial compounds that are<br />

both synthesized by and accumlated in plants after contact (exposure) with microbial<br />

pathogens (Paxton, 1981). Enhancement of disease resistance in mycorrhizal plants<br />

is due to accumulation of soluble antimicrobial phytoalexins. Rishitin and solavetivone<br />

are well-known phytoalexins produced by potato plants in response to pathogen<br />

infection (Engström et al., 1999). Mycorrhizal colonization of potato by Glomus<br />

etunicatum stimulated the accumulation of the phytoalexins rishitin and solavetivone<br />

in the roots of plantlets challenged with Rhizoctonia solani but did not change<br />

phytoalexin levels in non-challenged plantlet roots (Yao et al., 2002, 2003). The<br />

phytoalexin concentrations in mycorrhizal roots were significantly higher than those<br />

roots only challenged with Rhizoctonia solani. The result shows that mycorrhization<br />

can amplify phytoalexin production. Accumulation of phenolpropanoid compounds<br />

such as phytoalexins, and hydroxyproline-rich glycoproteins, which play important<br />

roles in plant defense, have been shown to increase either temporarily during<br />

mycorrhiza formation (Lambais, 2000; Garcia- Garrido and Ocampo, 2002). The<br />

major phytoalexin in alfalfa (Medicago truncatula) is the isoflavonoid (-)-medicarpin.<br />

Mycorrhizal colonization resulted in medicarpin accumulation between 7 and l3 days<br />

(Harrison and Dixon, 1993). The isoflavonoids glyceollin and coumestrol which are<br />

fungitoxic and nematostatic, respectively, accumulated in the mycorrhizal roots<br />

(Morandi and Le Querre, 1991). Isoflavonoids also has been showed to accumulate in<br />

mycorrhizal soybean roots (Morandi, 1984).<br />

Consistent increases in formononetin levels and transient increases in medicarpin-<br />

3-0-glycoside and formononetin conjugates were found in inoculated alfafa roots when<br />

mycorrrizal colonization began (Volpin et al., 1995). It is interesting that concentrations<br />

of formononetin increases when an AM fungus was in the rhizosphere, even when<br />

the plants growing there were not yet colonized by the fungus (Volpin and Kapulnik,<br />

1994). Formononetin has not been shown to have any antimicrobial activities, but is<br />

a precursor of isoflavonoid phytoalexin produced in alfalfa in response to microbial<br />

infections.<br />

Mycorrhizal roots of many plants develop so-called ‘yellow pigment’ , which<br />

has been used as an indicator to estimate the degree of mycorrhizal colonization<br />

(Fyson and Oaks, 1992). The chromophore of the ‘yellow pigment’ is an acyclic C 14<br />

carotenoid-derived polyene called mycorradicin (Klingner et al., 1995). Other<br />

components of the highly complex mixture of apocarotenoids accumulating in<br />

mycorrhizal roots are glycosylated C 13 cyclohexenone derivatives (Peipp et al., 1997).<br />

Mycorrhizal roots of barley, wheat, rye and oat were found to accumulate the blumenina<br />

terpenoid glycoside (Maier et al., 1995). The level of the compound was directly<br />

related to the degree of root colonization. Upon colonization by AM fungi, roots of<br />

many plant families accumulate certain apocarotenoids (Fester et al., 1999, 2002;<br />

Maier et al., 1995; 1998, 1999, 2000; Walter et al., 2000). Colonization of barley,<br />

wheat and maize roots by different arbuscular mycorrhizal fungi, i.e.<br />

185


186<br />

REN-SEN ZENG<br />

Glomus intraradices, Glomus mosseae, and Gigaspora rosea leads to the accumulation<br />

of similar cyclohexenone derivatives (Vierheilig et al., 2000). However, no fungusspecific<br />

induction of these compounds are known. Pathogens and endophyte did not<br />

induce the formation of cyclohexenone derivatives in barley roots (Maier et al., 1997).<br />

The role of cyclohexenone derivatives in disease resistance is unknown.<br />

In response to pathogen attack, plants activate an array of inducible defense<br />

reactions, many of which involve the transcriptional activation of the corresponding<br />

defense genes, including genes that encode enzymes involved in the synthesis of lignin<br />

and phytoalexins (Dixon et al., 1984; Dixon and Harrison 1990). Transcript levels of<br />

some pivotal enzymes of defense response of plants significantly increase after<br />

mycorrhizal fungal infection (Harrison and Dixon, 1993; 1994). Several inducible<br />

defence-related genes, including those encoding isoflavonoid phytoalexins such as<br />

phenylalanine ammonia lyase (PAL), chalcone synthase (CHS), chalcone isomerase<br />

(CHI) and for the cell wall structural protein HRGP, have been reported to be induced<br />

during mycorrhiz establishment (Tagu and Martin, 1996). Mycorrhization resulted<br />

in a local and systemic induction of plant defence-related enzymes chitinase,<br />

chitosanase and beta-1,3-glucanase, as well as superoxide dismutase in tomato plants<br />

(Pozo et al., 2002). Chalcone isomerase (CHI) and chitinase activities increased in<br />

inoculated roots prior to mycorrhizal colonization (Volpin et al., l994, Kapulink et<br />

al., 1996), whereas the increase in PAL activity coincided with colonization. Production<br />

of some new compounds, and increase in the activity of the enzymes peroxidase and<br />

polyphenol oxidase, was observed following inoculation with AM fungi (Charitha<br />

Devi and Reddy, 2002). Dumas-Gaudot et al. (1992a,b) found new chitinase isoforms<br />

that were specifically induced in several AM associations and were different from<br />

those elicited by root fungal pathogens, indicating a different pattern of plant response<br />

to pathogenic and mutualistic fungi. Glomus mosseae induced new chitinase isoforms<br />

in tomato roots (Pozo et al., 1996). Expression of genes encoding enzymes that<br />

synthesize phenolpropanoid compounds has been detected in mycorrhizal roots (Garcia-<br />

Garrido and Ocampo, 2002). Other defense related genes shown to be upregulated in<br />

mycorrhizal symbioses include: genes involved in metabolism of reactive oxygen<br />

species, chitinase and beta I ,3-glucanase, and genes involved in senescence, including<br />

glutathione-S-transferase. Mycorrhiza also induced changes in PR protein expression<br />

in tobacco leaves (Shaul et al., 1999).<br />

Colonization of barley, wheat and maize and rice roots by Glomus intraradices<br />

resulted in strong induction of transcript levels of the pivotal enzymes of<br />

methylerythritol phosphate pathway of isoprenoid biosynthes i.e., 1 -deoxy-D-xylulose<br />

5-phosphate synthase (DXS) and 1 -deoxy-D-xylulose 5-phosphate reductoisomerase<br />

(DXR) (Walter et al., 2000). At the same time six cyclohexenone derivatives were<br />

characterized from mycorrhizal wheat and maize roots. DXS2 transcript levels are<br />

low in most tissues but are strongly stimulated in roots upon colonization by<br />

mycorrhizal fungi, correlated with accumulation of carotenoids and apocarotenoids<br />

(Walter et al., 2002). Some reports show that the AM symbiosis may cause an increase,<br />

decrease, or no change in the plant defense reactions (Guenoune et al., 2001; Mohr et<br />

al., 1998).


DISEASE RESISTANCE IN PLANTS THROUGH<br />

ALLELOCHEMICALS<br />

DIMBOA (2,4-dihydroxy-7-methoxy-1 ,4-benzoxazin-3-one) play an important<br />

role in the chemical defense of cereals against insects, pathogenic fungi and bacteria<br />

(Niemeyer, 1988). Researches with maize indicate that mycorrizal colonization induced<br />

accumulation of DIMBOA and increase in transcript levels of Bx9. Concentrations of<br />

DIMBOA in maize roots inoculated both G. mosseae and Rhizoctonia solani were<br />

significantly higher than those roots inoculated separately with G. mosseae or<br />

Rhizoctonia solani.<br />

Phenolic compounds in plants play a role in disease resistance (Morandi, 1996).<br />

Mycorrhizal plants of maize accumulated more phenolic compounds p-hydrocinnamic<br />

acid and ferulic acid in roots than non-mycorrhizal plants (Huang, 2003). Mycorrhizal<br />

Ri T-DNA transformed carrot roots accumulated more phenolic compounds when<br />

challenged by Fusarium oxysporum (Benhamou et al., 1994). All cultivars of pea<br />

colonized with G. mosseae accumulated more phenolic acids, and total phenolic<br />

accumulation were closely correlated to disease intensity (Singh et al., 2004).<br />

Many studies show that AM fungi initiate a host defence response which is<br />

subsequently suppressed (Lambais and Mehdy, 1993; Volpin et al. 1994, 1995). The<br />

decreases were accompanied by differential reductions in the levels of mRNAs encoding<br />

for different endochitinase and endoglucanase isoforms. But the activation of specific<br />

plant defence reactions by AM fungi could predispose the plant to an early response<br />

to attack by a root pathogen (Gianinazzi-Pearson et al., 1994).<br />

Although studies on growth inhibition of fungi by isolated plant compounds<br />

suggest a role in plant defense, such in vitro tests may not always give an accurate<br />

indication of the significance of these compounds in restricting fungal growth in the<br />

plant. Despite increasing efforts in research on metabolic changes in mycorrhizal<br />

plants, the precise understanding of the mechanisms is poorly understood, and the<br />

role of secondary metabolites induced by AM fungi in disease resistance is still obscure.<br />

5. CONCLUSIONS<br />

Mycorrhizal fungi protect plant roots from diseases in several ways : (i) by providing<br />

a physical barrier to the invading pathogens. Physical protection is more likely to<br />

exclude soil insects and nematode than bacteria or fungi in ectomycorrhizal plants.<br />

However some nematodes can penetrate the fungal mantle, (ii) by competing with the<br />

pathogen; (iii) by producing allelopathic chemicals like secondary metabolites,<br />

antagonistic chemicals like antibiotics, toxins etc., and amount and type of the root<br />

exudates.<br />

For effective and persistent disease management and biocontrol the need is to<br />

evaluate the mycorrhizal symbionts in the natural system under field conditions. The<br />

use of mixed inoculum of mycorrhizal symbionts can be more effective and give<br />

better results than use of a single species. Selection of superior indigenous mycorrhizal<br />

symbionts may have an adaptive advantage to the soils and environment in<br />

which pathogen and host co-occur as compared to non-indigenous mycorrhizal<br />

symbionts.<br />

187


188<br />

REN-SEN ZENG<br />

6. REFERENCES<br />

Abdalla M.E., Abdel-Fattah G.M. Influence of the endomycorrhizal fungus glomus mosseae on the development<br />

of peanut pod rot disease in Egypt. Mycorrhiza 2000; 10: 29-35.<br />

Mzcon-Aguilar C., Barea J.M. Arbuscular mycorrhizas and biological control of soil-borne plant pathogens - an<br />

overview of the mechanisms involved. Mycorrhiza 1996; 6: 457-464.<br />

Azcon-Aguilar C., Barea J.M. Applying mycorrhizal biotechnology to horticulture: significance and potentials.<br />

Sci Hortic 1997; 68: 1-24.<br />

Bansal, M., Chamola, B.P., Sarwar, N., Mukerji, K.G. Mycorrhizosphere : Interaction between Rhizosphere<br />

microflora and VAM fungi. In, Myocrrhizal Biology, K.G. Mukerji, B.P. Chamola, Jagjit Singh eds.,<br />

Kluwer Academic/Plenum Publishers, New York, Dordrecht, London; 2000; pp. 143-152.<br />

Barker, S.J., Tagu, D. The role of auxins and cytokinins in mycorrhizal symbiosis. J. Plant Growth Regul 2000;<br />

19: 144-154.<br />

Benhamou N. Elicitro-induced pant defense pathways. Trends in Plant Science 1996; 1:233-240.<br />

Benhamou N., Fortin J.A., Hamel C., St.-Arnaud M, Shatilla A. Resistance responses of mycorrhizal Ri T-<br />

DNA transformed carrot roots to infection by Fusarium oxysporum f. sp. chrysanthemi. Phytopathology<br />

1994; 84: 958-968.<br />

Bennett R.N., Wallsgrove R.M. Secondary metabolites in plant defense mechanisms. New Phytol 1994; 127:<br />

617-633.<br />

Bodker L., Kjoller R. Rosendah S. Effect of phosophate and the arbuscular mycorrhizal fungus Glomus<br />

intraradices on disease severity of root rot of peas (Pisum sativum) caused by Aphanomyces euteiches.<br />

Mycorrhiza 1998; 8: 169-174.<br />

Buckingham J. Dictionary of natural products on CD-ROM. Chapman & Hall, London, 1998.<br />

Caron M. Potential use of mycorrhizae in control of soil-borne disease. Can J Pl Pathol 1989; 11: 177-179.<br />

Caron M., Richard C., Fortin, J.A. Effect of prteinfestation of the soil by a vesicular-arbuscular mycorrhizal<br />

fungus, Glomus intraradices, on Fusarium crown and root rot of tomatoes. Phytoprotection. 1986; 67:<br />

15-19.<br />

Charitha Devi M., Readdy M.N. Phenolic acid metabolish of groundnut (Arachis hypogaea L.) plants inoculated<br />

with VAM fungus and Rhizobium. Pl Gr Regul 2002; 37: 151-156.<br />

Conrath U., Pieterse C.M.J., Mauch Mani B. Priming in plant pathogen interactions. Trends Pl Sci 2002; 7:<br />

210-216.<br />

Cordier C., Gianianzzi S., Gianinazzi-Perason V. Colonization patterns of root tissues by Phytophthora nicotianae<br />

var. parasitica related to reduced disease in mycorrhizal tomato. Plant Soil 1996; 185: 223-232.<br />

Cordier C., Pozo M.J. Barea J.M., Gianinazzi S. Gianinazzi-Pearson V. Cell defense responses associated with<br />

localized and systemic resistance to Phytophthora parasitica induced in tomato by an arbuscular<br />

mycorrhizal fungus. Mol Plant-Microbe Interactions 1998; 11:1017-1028.<br />

Cordier C., Trouvelot A., Gianinazzi S., Gianinazzi-Pearson V. Arbuscular mycorrhizal technology applied to<br />

micropropagated Prunus avium and to protection against Phytophthora cinnamomi. Agronomie 1997;<br />

17: 256-265.<br />

Dar G.H. Zargar M.Y. Beigh G.M. Biocontrol of Fusarium root rot in the common bean (Phaseolus vulgaris<br />

L.) by using symbiotic Glomus mosseae and Rhizobium leguminosarum. Microb Ecol 1997; 34: 74-80.<br />

Davis R.M., Menge J.A. Influence of Glomus fasciculatum and soil phosphorus on phytopththora root rot of<br />

citrus. Phytopathology 1980; 70:447-452.<br />

Dehne H.W. Interaction between vesicular arbuscular mycorrhizal fungi and plant pathogens. Phytopathology<br />

1982; 72: 1115-1119.<br />

Dehne, H.W., Schonbeck, F. Unterschungen Zum Einflu-β-der endotrophen Mykorrhiza auf Pflanzenkrankheiten,<br />

II. Phenol stoffwechsel and Lignitiezierung. Phytopathol Z 1979b, 19: 210-216.<br />

Dehne, H.W., Schonbeck, F. The infleunce of endotrophic mycorrhizae on plant disease. I colonization to tomato<br />

plants by Fusarium oxysproum f. sp. lyeopersici. Phytopatol Z 1979a; 95: 105-109.<br />

Dixon R.A. The phytoalexin response: elicitation, signaling and control of host gene expression. Biol Rev 1986;<br />

61: 239-291.<br />

Dixon R.A. Natural productions and plant disease resistance. Nature 2001; 411: 843-847.<br />

Dixon R.A. and Gerrish C. Lamb C.J., Robbins M.P. Elicitor-mediated induction of chalone isomerase in<br />

Phaseolus vulgaris cell suspension cultures. Planta 1984; 159: 561-569.


DISEASE RESISTANCE IN PLANTS THROUGH<br />

ALLELOCHEMICALS<br />

Dixon R.A. and Harrison M.J. Activation, structure and organization of genes involved in microbial defense in<br />

plants. Adv Gentics 1990; 28: 165-233.<br />

Dixon, R.K. The Global Carbon Cycle and Global change : Resposns and fedcdback from the Mycorrhizaosphre.<br />

In, Mycorrhizal Biology, K.G. Mukerji, B.P. Chamola, Jaggit Singh eds. Kluwer Academic/Plenum<br />

Publishers, New York, Dordrecht, London. 2000; pp. 85-100.<br />

Dumas-Gaudot E., Furlan V., Grenier J., Asselin A. New acidic chitinase isoforms induced in tobacco roots by<br />

vesicular-arbuscular mycorrhizal fungi. Mycorrhiza 1992a; 1: 133-136.<br />

Dumas-Gaudot E., Renier J. Furlan V., Asselin A. Chitinase, chitosanase and b-1, 3-glucanase activites in<br />

Allium and Pisum roots colonized by Glomus species. Plant Sci 1992b; 84: 17-24.<br />

Engström K., Widmark A.K., Brishammar S., Helmersoon S., Antifungal activity to Phytophthora infestans of<br />

sesquiterpendoids from infectd potato tyubers. Potato Res 1999; 42: 43-50.<br />

Fester T., Maier W., Strack D. Accumulation of secondary compounds in barley and wheat roots in response to<br />

inoculation with an arbuscular mycorrhizal fungus and co-inoculation with rhizsophsere bacteria.<br />

Mycorrhiza 1999; 8: 241-246.<br />

Fester T., Schmidt D., Lohse S., Walter M.H., Giuliano G., Bramley P.M., Fraser P.d., Hause B., Strack D.<br />

Stimualtion of carotenoid metabolism in arbuscular mycorrhizal roots. Planta 2002; 216: 148-154.<br />

Frankenberger Jr. W.T., Arshad M. Microbial synthesis of auxins. In Phytohormones in soils, W.T. Frankberger<br />

Jr., M. Arshad eds., Marcel Deeker Inc., N.Y. 1995, pp. 35-71.<br />

Fyson A., Oaks A. Rapid methods for quantifying VAM fungal infections in maize roots. Plant Soil 1992; 147:<br />

317-319.<br />

Garcia-Garrido J.M., Ocampo J.A. Regulation of the plant defence response in arbuscular mycorrhizal symbiosis.<br />

J Exp Bot 2002; 53: 1377-1386.<br />

Gianinazzi-Pearson V. Plant Cell responses to arbuscular mycorrhizal fungi getting to the roots of the symbiosis.<br />

Plant cell 1996; 8: 1871-1883.<br />

Giaminazzi-Pearson V., Dumas-Gaudot E., Gollote A., Tahiri-Alaoui A., Gianinazzi S. Cellular and molecular<br />

defence related root responses to invasion by arbuscular mycorrhizal fungi. New Phytol 1996; 133: 45-<br />

57.<br />

Gianinazzi-Pearson V., Gollottee A. Dumas-Gaudo E. Franken P., Gianinazzi S. Gene expression and molecular<br />

modifications associated with plant responses to infection by arbuscular mycorrhizal fungi. In: Advnaces<br />

in molecular genetics of plantmicrobe interactions. M.J. Daniels, J.A. Downie, A.E. Obsourn. eds.<br />

Kluwer, Boston, 1994; pp. 179-186.<br />

Go’mez Garibay F., Reyes Chilpa R., Quijano L., Caldero’n Pardo J.S. Ray’os’Castillo T., Methoxifurans<br />

auranols with fungostatic activity from Lonchocarpus castilloi. Phytochemistry 1990; 29: 459-463.<br />

Grayer R.J. Harborne J.B., A survey of antifungal compounds from higher plants, 1982-1993. Phytochemistry<br />

1994; 37: 19-42.<br />

Guenoune D., Galili S., Phillips D.A., Volpin H. Chet I., Okon Y., Kapulnik Y. The defense response elicted by<br />

the pathogen Rhizoctonia solani is suppressed by colonization of the AM-fungus Glomus intraradices.<br />

Pl Sci 2001; 160: 925-932.<br />

Hammerschmidt R. Induced disease resistance: How do induced plants stop pathogens? Physiol Molecular Pl<br />

Pathol 1999; 55: 77-84.<br />

Harborne J.B., Ingham J.L. Biochemical aspects of the co-evaluation of higher plants with their fungal parasites.<br />

In Biochemical Aspects of Plant and Animal Co-evaluation. J.B. Harborne. ed. Academic Press, London,<br />

1978; 343-405.<br />

Harrison M.J., Dixon A. Isoflavonid accumulation and expression of defense gene transcripts during the<br />

establishment of vesicular arbuscular mycorrhizal association in roots of Medicago truncatula. Mol Plant-<br />

Microbe Interactions 1993; 6: 643-659.<br />

Harrison M.H., Dixon R.A. Spatial patterns of expression of flavonoid/isoflavonoid pathway genes during<br />

interactions between roots of Medicago truncatula and the mycorrhizal fungus Glomus versiforme. Plant<br />

J 1994; 6: 9-20.<br />

Hooker J.E., Jaizme-Vega M., Atkinson D. Biocontrol of plant pathogens using arbuscular mycorrhizal fungi.<br />

In : Impact of Arbuscular Mycorrhizas on Sustainable Agriculture and Natural Ecosystems. ALS,<br />

Birkhauser, Basel 1994; pp. 191-200.<br />

Huang J.H., Chemical mechanisms of plant disease resistance induced by arbuscular mycorrhizal fungi toward<br />

fungal pathogens. Doctoral dissertation submitted to South China Agricultural University, Guangzhou,<br />

2003.<br />

189


190<br />

REN-SEN ZENG<br />

Ingham E.R. Interactions among mycorrhizal fungi, rhizosphere orgnaims, and plants. In, Microbial mediation<br />

of Plant-herbivore Interactions. John Wiley and Sons, Inc. 1991; p. 530.<br />

Kapulink Y., Volpin H., Itzhaki H., Ganon D., Galili S., David R., Shaul O., Elad Y., Chet I., Okon Y. Supprossion<br />

of defence responses in mycorrhizal alfalfa and tobacco roots. New Phytol 1996; 133: 59-64.<br />

Klinger A., Bothe H., Wray V., Marner F.J. Identification of a yellow pigment formed in maize roots upon<br />

mycorrhizal colonization. Photochemistry 1995; 38: 53-55.<br />

Krishna K.R., Bagyaraj D.J. Interaction between Glomus fasciculatum and Sclerotium rolfsii in peanut. Can J<br />

Bot 1984; 61: 2349.<br />

Krupa S., Anderson J., Marx D.H. Studies on ectomycorrhizae of pine. IV. Volatile organic compounds in<br />

mycorrhizal and non mycorrhizal root systems of Pinus echinata Eur J For Pathol 1973; 3: 194-200.<br />

Lambais M.R., Mehdy M.C. Suppression of endochitinase, β-1, 3-endoglucanase and chalcone isomerase<br />

expression in bean vesicular-arbuscular mycorrhizal roots under different soil phosphate conditions. Mol<br />

Plant-Microbe Interactions 1993; 6: 75-83.<br />

Lambais M.R. Regulation of plant defense-related genes in arbuscular mycorrhizae. In: Current advances in<br />

mycorrhizae research. G.K. Podila, D.D. Douds eds. APS Press, Minnesota U.S.A. 2000.<br />

Levin D. Alkaloid-bearing plants: An ecogeographi perspective. Am Natur 1976; 110: 261-284.<br />

Linderman R.G. Role in VAM fungi in biocontrol. In: Mycorrhiza and Plant Health, F.L. Pfleger, R.G. Linderman<br />

eds. APS Press, St. Paul, 1994; pp. 1-26.<br />

Lingua G. D’Agostino G. Massa N., Antosiano M., Berta G. Mycorrhiza induced differential response to a<br />

yellows disease in tomato. Mycorrhiza 2002; 12: 191-198.<br />

Ludwig-Müller, J. Indole-3-butyric acid in plant growth and development. Plant Growth Regul 2000a; 32:<br />

219-230.<br />

Ludwig-Müller J. Hormonal balance in plants during colonization by mycorrhial fungi. In, Arbuscular<br />

Mycorrhizas : Physiology and function, D.D. Douds, Y. Kapulnik eds. Kluwer Academic Publishers,<br />

Dordrecht, the Netherlands. 2000b; pp. 263-285.<br />

Ludwig-Müller J. From auxin homeostasis to understanding plant pathogen and plant symbiont interaction :<br />

editor’s research interests JPGR 2004; 23: 1-8.<br />

Maier W., Hammer K. Dammann U., Schulz B., Strack D., Accumulation of sesquiterpenoid cyclohexenone<br />

dervatives induced by an arbuscular fungus in members of the Poaceae. Planta 1997; 202: 36-42.<br />

Maier W., Peipp H., Schmidt J., Wray V., Strack D. Levels of a terpenoid glycoside (blumenin) and cell wallbound<br />

phenolics in some cereal mycorrhizas. Pl Physiol 1995; 109: 465-470.<br />

Maier W., Schmidt J., Nimtz M., Wray V., Strcak D. Secondary products in mycorrizal roots of tobacco and<br />

tomato. Phytochemistry 2000; 54: 473-479.<br />

Maier W., Schmidt J. Wray V., Walter M.H. Strack D. The arbuscular mycorrhizal fungus Glomus intraradices<br />

induces the accumulation of cyclohexenone derivatives in tobacco roots. Planta 1999; 207: 620-623.<br />

Maier W., Schneier B., Strack D. Biosynthesis of sesquiterpenoid cyclohezenone derivatives in mycorrhizal<br />

barley roots proceeds via the glyceraldehydes 3-phosphate/pyruvate pathway. Tetrahedron Letter 1998;<br />

39: 521-524.<br />

Maronek, D.M. Mycorrhizal fungi and their importance in horticulture crop production. Hort Rev 1981; 3:<br />

172-213.<br />

Marx D.H. Ectomyocrrhizae as biological deterents to pathogenic root infections. Ann Rev Phytopath 1972; 10:<br />

429-441.<br />

Matsubara Y., Tamura H., Harada T. Growth enhancement and Verticillium wilt control vesicular-arbuscular<br />

mycorrhiza fungus inoculation in eggplant. J Japan Soc Hort Sci. 1995; 555-561.<br />

Mauch-Mani B., Metraux J.P. Salicylic acid and systemic acquired resistance to pathogen attack. Ann Bot<br />

1998; 82: 535-561.<br />

Mendelsohn R., Balick M.J. The value of undiscovered pharmaceuticals in tropical forests. Eco Bot 1995; 49:<br />

223-228.<br />

Mohr U., Lange J., Boller T., Wiemken A., Vogeli-Lange R. Plant defence genes are induced in teh pathogenic<br />

interaction between bean roots and Fusarium solani, but not in the symbiotic interaction with the arbuscular<br />

mycorrhizal fungus Glomus mosseae. New Phytol 1998; 138: 589-598.<br />

Morandi D. Isoflavonoid accumulation in soybean roots infected with vesicular-arbuscular mycorrhizal fungus.<br />

Physiol Pl Pathol 1984; 24: 357-364.


DISEASE RESISTANCE IN PLANTS THROUGH ALLELOCHEMICALS<br />

Morandi D. Occurrence of phytoalexins and phenolic compounds in endomyocrrhizal interactions, and their<br />

potential role in biological control. Plant Soil. 1996; 185: 241-251.<br />

Morandi D., LeQuerre J.L. Influence of nitrogen on accumulation of isosojagol (a new deterred coumestan in<br />

soybean) and associated isoflavonoids in roots and nodules of mycorrhizal and non-mycorrhizal soybean.<br />

New Phytol 1991; 117: 75-79.<br />

Mukerji K.G. Mycorrhiza in control of plant pathogens : Molecular approaches. In, Biotechnological Approaches<br />

in Biocontrol of Plant Pathogens, K.G. Mukerji, B.P. Chamola, R.K. Upadhyay eds. Kluwer Academic/<br />

Plenum Publishers, New York, Dordrecht, London, 1999; pp. 135-155.<br />

Mukerji K.G. Rhizosphere Biology. In, Techniques in Mycorrhizal Studies, K.G. Mukerji, C. Manoharachary,<br />

B.P. Chamola eds. Kluwer Academic Publishers, Dordrecht, The Netherlands, 2002; pp. 87-101.<br />

Mukerji K.G., Upadhyay R.K., Kaushik A. Mycorrhiza and Integrated Disease Management. In, IPM Systems<br />

in Agriculture. Vol 2. Biocontrol in Emerging Biotechnology. R.K. Upadhyay, K.G. Mukerji, R.L. Rajak<br />

eds. Aditya Books Pvt. Ltd., New Delhi, India, 1997; pp. 423-452.<br />

Neswham K.K., Fitter A.H. Watkinson A.R. Arbuscular mycorhriza protect an annual gras from root pathogenic<br />

fungi in the field. J Ecol 1995; 83: 991-1000.<br />

Niemeyer H.M. 1988. Hydroxamic acids (4-hydroxy-1, 4-benzoxazin-3-ones), defense chemicals in the<br />

GRamineae. Phytochemistry 1988; 27: 3349-3358.<br />

Niemira B.A., Hammerschmidt R., Sar G.R. Postharvest suppression of potato dry rot (Fusarium sambucinum)<br />

in prenuclear minitubers by arbuscular mycorrhizal fungal inoculum. An Potato J 1996; 73: 509-515.<br />

Paxton J.D. Phytoalexins-a working redefinition. Phytopathol 1981; 101: 106-109.<br />

Peipp H., Maier W., Schmidt J. Wray V. Strack D. Arbuscular mycorrhizal fungus-induced changes in the<br />

accumulation of secondary compounds in barely roots. Phytochemistry 1997; 44: 581-587.<br />

Pozo M.J., Cordier C., Dumas-Gaudot E., Gianinazzi S., Barea J.M., Azcon-Aguilar C. Localized versus systemic<br />

effect of arbuscular mycorrhizal fungi on defence responses to Phytophthora infection in tomato plants. J<br />

Exp Bot 2002; 525-534.<br />

Pozo M.T., Dumas-Gaudot e., Slezack S., Cordier C., Asselin A., Gianinazzi S., Gianinazzi-Pearson V., Azcon-<br />

Aguilar C., Barea J.M. Infuction of new chitinase isoforms in tomato roots during interactions with Glomus<br />

mosseae and/or Phytophthora nicotianae var parasitica. Agronomie 1996; 16: 689-697.<br />

Raina S., Chamola B.P., Mukerji K.G. Evolution of Mycorrhiza. In, Mycorhizal Biology, K.G. Mukerji, B.P.<br />

Chamola, J. Singh eds. Kluwer Academic / Plenum Publishers. New York, pp. 1-26.<br />

Ramaraj B., Shanmugam N., Reddy D.A. Biocontrol of Macrophomina root rot of cowpea and Fusarium wilt<br />

of tomato by using VAM fungi. In: Mycorrhizae for green Asia. A Mahadevan ed. Proc. 1st Asian Conf<br />

Mycorrhizae, University of Madras, Madras, India 1988; pp. 250-251.<br />

Rambelli A. The rhizosphere of mycorrhizae. In, Ectomycorrhizae, G.L. Marks, T.t. Koslowski eds. Academic<br />

Press, New York, 1973; pp. 299-343.<br />

Redecker D., Kodner R., Graham L.E. Glomalean fungi from the Ordovician. Science 2000; 289: 1920-1921.<br />

Reid C.P.P. Mycorrhizas. In, The Rhizosphere, J.M. Lynch ed. John Wiley and Sons, Chickester, England;<br />

1990; pp. 281-315.<br />

Safir G. The influence of vesicular mycorrhiza on the resistance of onion of Pyrenochaeta terrestris. M.S.<br />

Thesis, Univeristy of Illinois, Urbana. 1968.<br />

Sampangi R.K. Some recent advances in the study of fungal root diseases. Ind Phytopth 1989; 22: 1-17.<br />

Secilia J., Bagyaraj D.J. Bacteria and actinomycetes associated with pot cultures of vesicular arbuscular<br />

mycorrhiza. Can J Microbol 1987; 33: 1069-1073.<br />

Shaul O., Galili S., Volpin H., Ginzberg I.I., Elad Y., Chet I. I., Kapulnik Y. Mycorrhiza - induced changes in<br />

disease severity and PR protein expression in tobacco laves. Mol Plant-Microbe Interact 1999; 12: 1000-<br />

1007.<br />

Singh D.P., Srivastava J.S., Bahadur A., Singh U.P., Singh S.K. Arbuscular mycorrhizal fungi induced biochemical<br />

changes in pea (Pisum sativaum) and their effect on powdery mildew (Erysiphe pisi). Phytopathol Z fur<br />

Pelanzekrankheiten und Pelanzenschutz 2004; 111: 266-272.<br />

Singh G., Mukerji K.G. Root exudates as determinant of rhizospheric microbial diversity. In, Microbial activity<br />

in the rhizosphere. K.G. Mukerji, C. Manoharachary, J. Singh eds. Springer Verlag, Berlin, Heidelberg.<br />

2006; 39-55.<br />

Singh R., Adholeya A., Mukerji K.G. Mycorrhiza in control of soiolborne pathogens. In: Mycorrhizal biology.<br />

Mukerji K.G., Chamola B.P., Singh J. eds. Kluwer Academic / Plenum Publishers, New York, Dordrecht,<br />

London, 2000; pp. 173-196.<br />

191


192<br />

REN-SEN ZENG<br />

Smith S.E., Reid D.J. Mycorrhizal Symbiosis. Academic Press, London, New York. 1997; p. 605.<br />

Srivastava D., Kapoor R., Srivastava S.K., Mukerji K.G. Vesicular-arbuscular mycorrhiza - an overview. In,<br />

Concept in mycorrhizal research, K.G. Mukerji ed. Kluwer Academic Publishers, Dordrecht, The<br />

Netherlands, 1996; pp. 1-39.<br />

St-Arnaud M., Hamel C., Carom M, Fortin J.A. Inhibition of Pythium ultimum in roots and growth substrate of<br />

mycorrhizal Tagetes patula colonized with Glomus intraradices. Can J Path 1995; 16: 187-194.<br />

St-Arnaud M., Hamel C., Vimard B., Caron M., Fortin J.A. Inhibition of Fusarium oxysporum f. sp. dianthi in<br />

the non-VAM species Dianthus caryophyllus by co-culture with Tagetes patula companion plants colonized<br />

by Glomus intraradices. Can J Bot 1997; 75: 998-1005.<br />

Strobel N.E., Sinclair W.A. Role of flavonilic wall inclusions in the resistance induced by Laccaria bicolor to<br />

Fusarium oxysporum in primary roots of Donglas fir. Phytopathology 1991; 81: 420-425.<br />

Tagu D., Martin F. Molecular anlaysis of cell wall proteins expressed during early steps of ectomycorrhiza<br />

development. New Phytol 1996; 133: 73-85.<br />

Torres-Barragan A., Zavaleta-Mejia E., Gonzalez-Chavez C., Ferrera-Cerrato R. The use of arbuscular<br />

mycorrhizae to control onion white rot (Sclerotium cepivorum Berk.) under field conditions. Mycorrhiza<br />

1996; 6: 253-257.<br />

Trotta A., Varese G.C., UNAVI E., Fusconi A., Sampo S., Berta G. Interactions between the soil-borne root<br />

pathogen Phytophthora nicotianae var parasitica and arbuscular mycorrhizal fungus Glomus mosseae<br />

in tomato plants. Plant Soil 1996; 185: 199-209.<br />

Vierhilig H., Bago B., Albrecht C., Poulin M.J., Piche Y. Flavonoids and arbuscular mycorrhizal fungi. In :<br />

Falvonoids in the living system. J. Manthey, B. Buslig eds. Plenum, New York, 1998; pp. 9-33.<br />

Vierheilig H., Gagnon H., Strack D., Maier W. Accumulation of cyclohexenone derivatives in barley, wheat and<br />

maize roots in response to inoculation with different arbuscular mycorrhizal fungi. Mycorrhiza 2000; 9:<br />

291-293.<br />

Volpin H., Elkind Y. Okon Y., Kapulnik Y. A vesicular arbuscular mycorrhizal fungus (Glomus intraradix)<br />

induce a defense response in alfalfa roots. Pl Physiol 1994; 104: 683-689.<br />

Volpin H., Kapulnik Y. Interaction of Azospirillum with beneficial soil microorganisms. In: Azospirillum /<br />

Plant Associations. Okon Y. ed. CRC Press Boca Baton FL 1994; pp. 111-118.<br />

Volpin H., Phillips D.A., Okon Y., Kapulnik Y. Suppression of an isoflavonoid phytoalexin defense response in<br />

myocrrhizal alfalfa roots. Plant Physiol 1995; 108: 1449-1545.<br />

Walker C. Systamatics and Taxonomy of the arbuscular endomycorrhizal fungi (Glomales) - a possible way<br />

forward. Agronomie 1992; 12: 887-892.<br />

Walter M.H., Festr T. Strack D. Arbuscular mycorrhizal fungi induce the non-mevalonate methylerythritol<br />

phosphate pathway of isoprenoid biosynthesis correlated with accumulation of the ‘yellow pigment’ and<br />

other apocarotenoids. Plant J 2000; 21: 571-578.<br />

Walter M.H., Hans J., Strack D. Two distantly related genes encoding 1-deoxy-D-xylulose 5-phosphate synthases:<br />

differential regulation in shoots and apocarotenoid-accumulating mycorrhizal roots. Plant J 2002; 31:<br />

243-254.<br />

Yao M.K. Desilets H., Charles M.T., Boulanger R. Tweddell R.J. Effect of mycorrhization on the accumulation<br />

of rishitin and solavetivone in potato plantlets challenged with Rhizoctonia solani. Mycorrhiza 2003;<br />

13:333-336.<br />

Yao M., Tweddell R., Desilets H. Effect of two vesicular-arbuscular mycorrhizal fungi on the growth of<br />

micropropagated potato plantlets and on the extent of disease caused by Rhizotonia solani. Mycorrhiza<br />

2002; 12: 235-242.


CHUIHUA KONG<br />

ALLELOCHEMICALS FROM Ageratum conyzoides L.<br />

AND Oryza sativa L. AND THEIR EFFECTS<br />

ON RELATED PATHOGENS<br />

Institute of Applied Ecology, Chinese Academy of Sciences, Shenyang 110016,<br />

China, and South China Agricultural University, Guangzhou 510642, China.<br />

E-mail: kongch@mail.edu.cn<br />

Abstract. <strong>Allelochemicals</strong> play an important role in biological control of plant pathogens and diseases. Weed<br />

Ageratum conyzoides L. and food crop Oryza sativa L. can produce and release many kinds of allelochemicals<br />

participating in their defense against pathogens. The essential oil from A. conyzoides has been found to have<br />

significant negative effects on several plant pathogens. In the A. conyzoides intercropped citrus orchard, A.<br />

conyzoides released allelopathic flavones and agreatochromene into the soil to reduce the populations of soil<br />

pathogenic fungi Phytophthora citrophthora, Pythium aphanidermatum and Fusarium solani. Further research<br />

revealed that ageratochromene underwent a reversible transformation in the soils, that is, ageratochromene<br />

released from A. conyzoides plants was transformed into its dimers, and the dimers can be remonomerized in the<br />

soils. The reversible transformation between ageratochromene and its dimers in the A. conyzoides intercropped<br />

citrus orchard soil can be an important mechanism maintaining bioactive allelochemicals at an effective<br />

concentration, thus, sustaining the inhibition of pathogenic fungi in soil. Many kinds of allelochemicals in rice<br />

were identified. Among them, alkylresorcinols, flavone and cyclohexenone had high antifungal activities on<br />

Pyricularia oryzae and Rhizoctonia solani. Furthermore, these antifungal allelochemicals formed by rice can<br />

be triggered by a large number of abiotic and biotic factors. Antifungal allelochemicals from rice mainly involved<br />

two types of diterpenes and flavones, including momilactones A and B, oryzalexins A-F and S, phytocassanes<br />

A-E and sakuranetin. These compounds help rice establishing its own pathogen defense mechanism. However,<br />

it remains obscure which allelochemicals in rice are predominantly involved in defense mechanisms against the<br />

pathogens. Therefore, further clarification of the resistance mechanism and multiple functions of these compounds<br />

on rice are warranted.<br />

1. INTRODUCTION<br />

Plants produce many kinds of low-molecular-mass secondary metabolites that are<br />

generally non-essential for the basic metabolic processes of the plant. Among these<br />

secondary plant metabolites, some are known as allelochemicals that improved defense<br />

against other plant competition, microbial attack or insect/animal predation. Plants<br />

cannot move to escape pathogens. However, plants have evolved to successfully<br />

withstand infection by a vast majority of pathogens that attack them (Stuiver and<br />

Custers, 2001). It was found that plants biosynthesize phytoalexins as soon after<br />

pathogenic attack (Dangl and Jones, 2001). The concept of phytoalexins as induced<br />

anti-microbial allelochemicals in plant was first developed in 1940 (Muller and Borger,<br />

193<br />

Inderjit and K.G. Mukerji (eds.),<br />

<strong>Allelochemicals</strong>: <strong>Biologica</strong>l Control of Plant Pathogens and Diseases, 193– 206.<br />

© 2006 Springer. Printed in the Netherlands.


194<br />

CHUIHUA KONG<br />

1940; Bailey and Mansfield, 1982). Anti-microbial allelochemicals or phytoalexins<br />

have been extensively investigated and play important roles in plant defense (Dixon,<br />

2001). Most of anti-microbial allelochemicals have relatively broad-spectrum activity<br />

on pathogens and specificity is often occurred in cropping systems. Accordingly, an<br />

understanding of the interactions between allelochemicals and pathogenic organisms<br />

is essential in disease control in agro-ecosystems.<br />

The structures and sources of many allelochemicals with anti-microbial activity<br />

have been documented (Dixon, 2001; Grayer and Harborne, 1994; Harborne, 1999).<br />

In this chapter, evidence for anti-microbial functions of allelochemicals from weed<br />

Ageratum conyzoides L. and food crop Oryza sativa L. has been reviewed. Effects of<br />

these allelochemicals on related pathogen management in the A. conyzoides<br />

intercropped citrus orchard and the paddy ecosystem were discussed.<br />

2. ALLELOCHEMICALS OF A. conyzoides AND THEIR EFFECT ON<br />

RELATED PATHOGENS<br />

2.1. Ageratum conyzoides L.<br />

Ageratum conyzoides of the family Compositae (Asteraceae) is native to Central<br />

America (Kossmann and Groth, 1993) Caribbean and Florida (USA). It has spread to<br />

West Africa, Southeast Asia, South China, India, Australia and South America<br />

(Okunade, 2002; Stadler et al., 1998). A. conyzoides is an annual erect, branched<br />

herb growing 15 to 100 cm tall. Its stem is covered with fine white hairs, leaves are<br />

opposite, pubescent with long petioles and include glandular trichomes. It has a shallow<br />

tap root system. The inflorescence contains 30 to 50 pink or purple flowers arranged<br />

in a corymb and are self-incompatible (Jhansi and Ramanujam, 1987). The fruit is an<br />

achene with an aristate pappus and is easily dispersed by wind and animals fir. Seeds<br />

are positively photoblastic and remain viable upto 12 months (Okunade, 2002). The<br />

seeds germinate between 20-25°C. It prefers a moist, well drained soil but may tolerate<br />

dry conditions (Ladeira et al., 1987). This species has great morphological variations<br />

and appears highly adaptable to varying ecological conditions (Hu and Kong, 2002a).<br />

It is a pioneer plant growing in ruined sites and cultivated fields and often becomes<br />

dominant and forms a stand in natural community and is resistant to common insects<br />

or diseases (Liang and Hunag, 1994). Although it is harmful to crops and invades<br />

cultivated fields and interferes with the natural community compositions, it has been<br />

used as folk medicine in several countries and it has anti-microbial, insecticidal and<br />

nematicidal properties (Ming, 1999; Okunade, 2002). In Central America A.<br />

conyzoides has been bred for many colours of flowers (Stadler et al., 1998). In South<br />

China A. conyzoides is traditionally used as green manure in fields to increase the<br />

crop yields, and usually is intercropped as understory in citrus orchards to suppress<br />

weeds and control other pests (Liang and Hunag, 1994; Kong et al., 2004b). This<br />

species appears to be a valuable agricultural resource (Ming, 1999).


ALLELOCHEMICALS FROM AGERATUM CONYZOIDES AND ORYZA SATIVA<br />

2.2 The essential oil of A. conyzoides and their biological activities on related<br />

pathogens<br />

A. conyzoides has a wide range of secondary metabolites including flavonoids,<br />

chromenes, benzofurans and terpenoids. Among these secondary metabolites, some<br />

are allelochemicals inhibiting the growth of other organisms (Okunade, 2002; Pari et<br />

al., 1998). Usually, A. conyzoides can produce and release volatile chemicals into the<br />

environment. The concentration of its released volatiles is so high that the unpleasant<br />

odor can be smelled in the fields. Therefore, most investigations have focused on<br />

chemical components of its essential oil (Albersberg and Singh, 1991; Ekundayo et<br />

al., 1988; Menut et al., 1993; Wandji et al., 1996). It was found that ageratochromenes<br />

and their derivatives, monoterpenes and sesquiterpenes were the major components<br />

of the essential oil from A. conyzoides (Kong et al., 1999; 2002a; Pari et al., 1998).<br />

The allelopathic potential of volatile allelochemicals from A. conyzoides has been<br />

reported in our previous papers (Kong et al., 1999; 2002a; 2004a). Anti-microbial<br />

effects of the essential oil from A. conyzoides have been confirmed for a long time<br />

(Biond et al., 1993; Dixit et al., 1995; Rao et al., 1996). Table 1 showed that several<br />

fungal pathogens, such as Rhizoctonia solani, Botrytis cinerea, Sclerotinia sclerotiorum,<br />

were significantly inhibited by the essential oils of A. conyzoides (Kong et al., 2001;<br />

2002a). The quantity and variety of allelochemical produced by A. conyzoides varies<br />

depending on its growth stages and habitats and so do their growth inhibitory effects<br />

on the pathogens (Kong et al., 2002a, 2004a). A. conyzoides produces different volatiles<br />

in larger quantities when infected with Erysiphe cichoracearum (Kong et al., 2002a).<br />

Table 1. Inhibitory effects of essential oil from the A. congzoides collected from<br />

different growth stages and habitats on fungal pathogens.<br />

The essential oil collected Pathogens<br />

from R. solani B. cinerea S. sclerotiorum<br />

Growth stages<br />

4-leaf 23.8±3.8a 19.6±5.3a 39.8±7.6a Pre-flowering 56.3±10.1b 49.7±9.1b 60.2±7.9b Peak-flowering 100c 82.3±11.5c 100c Mature 60.6±8.8b 58.3±7.4b 40.8±9.5a Habitats<br />

Cultivated field 79.6±10.2d 83.5±6.3c 93.8±7.7c Under citrus canopy 58.4±4.3b 45.6±6.1b 66.2±5.5b Roadside 100 c 100d 100c Control<br />

50% Carbendazin 100c 59.8±5.2c 92.3±7.9c Test concentrations of the essential oil were 100 µg/ml. All data are inhibitory percentage of spore germination<br />

of fungal pathogens tested and mean of 3 replicates with standard error. Data in a column not followed by the<br />

same letter are significantly different, p=0.05, ANVOA with Ducan’s multiple range test.<br />

195


196<br />

CHUIHUA KONG<br />

2.3. <strong>Allelochemicals</strong> and pathogen management in the A. conyzoides intercropped<br />

citrus orchard<br />

Besides the volatiles, A. conyzoides can biosynthesize and release non-volatile<br />

allelochemicals into the soil, thus, inhibiting the growth of other plants and<br />

microorganisms in soils. Polymethoxyflavones, ageratochromene and its analogues<br />

are rare in natural products but they have been found in A. conyzoides (Adesogan and<br />

Okunade, 1979; Gonzalez et al., 1991; Horie, et al 1993; Okunade, 2002). These<br />

compounds have obvious anti-microbial activity and have been used in managed<br />

ecosystem.<br />

In South China A. conyzoides is often intercropped in citrus orchards as an<br />

understory plant that quickly becomes dominant in citrus orchards. In addition,<br />

intercropping A. conyzoides makes the citrus orchard ecosystem more favorable for<br />

predatory mites (Amblyseius spp.). These mites are effective natural enemies of the<br />

citrus red mite (Panonychus citri). Further investigations showed that the pathogenic<br />

fungi Phytophthora citrophthora, Pythium aphanidermatum and Fusarium solani were<br />

isolated from both the A. conyzoides intercropped and non-intercropped citrus orchards<br />

soils. However, populations of these fungi were lower in the A. conyzoides intercropped<br />

citrus orchard than in the non-intercropped citrus orchard (Figure 1), indicating that<br />

intercropping with A. conyzoides in citrus orchards markedly decreased the population<br />

of soil pathogenic fungi. It may have resulted from the phytotoxins in the A. conyzoides<br />

intercropped citrus orchard soil (Kong et al., 2004c).<br />

P. citrophthora P. aphanidermatum F. solani<br />

Figure 1. Population of major pathogenic fungi in the A. conyzoides intercropped and nonintercropped<br />

citrus orchards soils. Population density was mean individual in per gram soil.


ALLELOCHEMICALS FROM AGERATUM CONYZOIDES AND ORYZA SATIVA 197<br />

Several flavones, ageratochromene and its two dimers were isolated and identified<br />

from the A. conyzoides intercropped citrus orchard soil (Figures 2 and 3), their amounts<br />

ranged from 11 to 93µg g -1 in the air dried soil. However, these chemicals could not be<br />

found in the non-intercropped citrus orchard soil. The results showed that these<br />

chemicals were primarily released from ground cover plants of A. conyzoides and<br />

accumulated in the soil year by year.<br />

A: R 1 =R 2 =OCH 3 ; B: R 1 =H, R 2 =OCH 3 ; C: R 1 =OCH 3 , R 2 =H<br />

D: R 1 =R 3 =OCH 3 , R 2 =H; E: R 1 =R 2 =R 3 =OCH 3 ;<br />

F: R 1 =R 3 =H, R 2 =OCH 3 ; G: R 1 =H, R 2 =R 3 =OCH 3 ;<br />

H: R 1 =R 3 =OCH 3 , R 2 =OH; I: R 1 =R 3 =OCH 3 , R 2 =α- rhamnosyl.<br />

Figure 2. Flavones produced and released from the A. conyzoides<br />

intercropped citrus orchard soil.<br />

Bioassays showed that ageratochoromene and flavones could significantly inhibit<br />

spore germination of the pathogenic fungi P. citrophthora, P. aphanidermatum and F.<br />

solani, but two dimers of ageratochromene had no inhibitory effects on them (Table<br />

2). Thus, the flavones and ageratochromene could be one of the key factors that A.<br />

conyzoides plants are able to reduce the populations of soil pathogenic fungi in the<br />

citrus orchard. Two dimers, though not biologically active, may be the products of<br />

ageratochromene transformation in soil.<br />

Further studies revealed that ageratochromene underwent a reversible<br />

transformation in the soils, that is, ageratochromene released from ground A.<br />

conyzoides plants was transformed into its dimers, and the dimers can be<br />

remonomerized in the soils (Kong et al., 2004c). However, this dynamic transformation<br />

did not occur in the soil with low organic matter and fertility (Figure 3). The reversible<br />

transformation between ageratochromene and its dimers in the A. conyzoides<br />

intercropped citrus orchard soil can be an important mechanism maintaining bioactive


198<br />

CHUIHUA KONG<br />

Table 2. Inhibitiory effects of allelochemicals isolated from the A. conyzoides<br />

intercropped citrus orchard soil on major pathogenic fungi.<br />

Chemicals P. citrophthora P. aphanidermatum F. solani<br />

Flavone A 100 a 90.6±8.6 a 72.1±6.5 a<br />

Flavone B 90.5±7.6 b 75.9±8.1 b 39.9±5.8 b<br />

Flavone C 91.2±6.3 b 76.3±6.5 b 35.6±7.1 b<br />

Flavone D 100 a 90.5±8.7 a 59.6±7.1 c<br />

Flavone E 100 a 93.2±9.2 a 60.2±7.3 c<br />

Flavone f 87.5±7.9 b 80.5±7.2 b 38.9±4.5 b<br />

Flavone G 77.3±6.1 c 74.2±6.2 b 40.2±5.3 b<br />

Flavone H 98.6±8.0 b 80.3±6.6 b 57.7±4.1 c<br />

Flavone I 39.8±2.1 d 39.7±7.5 c 35.3±5.0 b<br />

Ageratochromene 43.5±6.4 d 37.6±5.4 c 60.8±9.9 c<br />

Dimer A 6.1±1.1 e 0.9±0.3 d 1.8±0.6 d<br />

Dimer B 5.0±0.9 e 1.6±0.3 d 7.2±1.6 d<br />

50% Carbendazin 100 a 74.8±6.1 b 58.3±3.2 c<br />

Test concentrations of the essential oil were 100 µg/ml. All data are inhibitory percentage and<br />

mean of 3 replicates with standard error. Data in a column not followed by the same letter are<br />

significantly different, p=0.05, ANVOA with Ducan’s multiple range test.<br />

Figure 3. Transformation and degrading of ageratochormene in the soil.<br />

Ageratochormene transformation was in soil with high organic matter content and high fertility.<br />

Transformation occurred from step 1 to step 3, only in soil with low organic content and fertility.<br />

allelochemicals at an effective concentration, thus, sustaining the inhibition of<br />

pathogenic fungi in soil.<br />

Therefore, A. conyzoides has been advocated for intercropping in citrus orchards<br />

and utilized on more than 150,000 ha of citrus orchards in South China and gave<br />

substantial ecological and economic benefits (Liang and Huang, 1994). It is an excellent<br />

example of applied aspects of allelopathy in agro-ecosystem.


ALLELOCHEMICALS FROM AGERATUM CONYZOIDES AND ORYZA SATIVA 199<br />

3.1. Rice antifungal allelochemicals<br />

3. RICE ALLELOCHEMICALS AND THEIR<br />

EFFECTS ON RELATED PATHOGENS<br />

Rice (Oryza sativa L.) is one of the principal food crops in the world. Its production is<br />

characterized by heavy use of herbicides and fungicides that may cause environmental<br />

problems in the paddy ecosystem (Kim and Shin, 2000). Accordingly, rice<br />

allelochemicals can potentially be used to improve weed and pathogen management<br />

in rice production. Therefore, search for allelochemicals from rice has been extensively<br />

studied (Chung et al., 2001; Kong et al., 2002b; Mattice et al., 1998; Rimando et al.,<br />

2001). A range of phenolic acids was identified as potent allelochemicals from rice<br />

tissues and root exudates (Chung et al., 2001; Mattice et al., 1998; Rimando et al.,<br />

2001). However, these phenolic acids are unlikely to explain the allelopathy of rice<br />

since their soil concentrations never reach phytotoxic levels (Olofsdotter et al., 2002).<br />

More recently, an increasing number of studies have shown that a few flavones,<br />

diterpenoids and other types of compounds are also the potent allelochemicals from<br />

rice (Kato-Noguchi et al., 2002; 2003; Kong et al., 2002b; 2004d,e; Lee et al., 1999).<br />

These allelochemicals can be biosynthesized in rice seedlings and then released into<br />

their surroundings at ecologically relevant concentrations to inhibit the germination<br />

and growth of associated weeds. Similarly, rice allelochemicals may participate in the<br />

defense mechanisms of rice against pathogens.<br />

In our laboratory, a flavone (5,7,4.-trihydroxy-3,5.-dimethoxyflavone), a<br />

cyclohexenone (3-isopropyl-5-acetoxycyclohexene-2-one-1) and a liquid mixture of<br />

low polarity, containing long-chain and cyclic hydrocarbons (Table 3), were isolated<br />

and identified from leaves of allelopathic rice accession PI 312777 (Kong et al., 2004e).<br />

Both the flavone and cyclohexenone significantly inhibited spore germination of<br />

Pyricularia oryzae and Rhizoctonia solani, but the mixture containing low polarity<br />

constituents did not show any inhibitory effect on them, even at high concentrations<br />

(Figure 4a,b). The IC 50 values of the flavone on spore germination of P. oryzae and R.<br />

solani were ca 50 and 70.g. g -1 , while the cyclohexenone were ca 75 (P. oryzae) and<br />

95 (R. solani), respectively. At all concentrations tested, the inhibitory activity of the<br />

flavone on pathogens was slightly higher than that of the cyclohexenone. The complete<br />

inhibition of both compounds on spore germination of the pathogens was observed at<br />

250.g. g -1 . (Figure 5a, b).<br />

a. b.<br />

5,7,4¡Ç-trihydroxy-3¡Ç,5¡Ç-dimethoxyflavone 3-isopropyl-5-acetoxycyclohexene-2-one-1<br />

Figure 4. a. Flavone; b. Cyclohexenone.


200<br />

CHUIHUA KONG<br />

(a) (b)<br />

Figure 5. Inhibitory activities of flavone, cyclohexnone and mixture with low polarity against<br />

spore germination of (a) P. oryzae and (b) R. solani at different concentrations.<br />

Table 3. Chemical constituents of the mixture with low<br />

polarity isolated from rice leaves<br />

Retention time (min) Chemical constituents Relative amount (%)<br />

1.89 2-methylhexane 8.77<br />

5.40 4,6-dimethylundecane 2.48<br />

6.28 2,2,6-trimethyldecane 2.35<br />

7.26 2-hexyl-1-decanol 2.11<br />

7.58 1-butyl-2-propylcyclopentane 8.45<br />

8.19 2,6-dimethyl-decahydronaphthalene 22.45<br />

8.40 trans, trans-1,10-dimethylspiro[4,5]decane 8.03<br />

8.48 2,3-dimethyl-decahydronaphthalene 3.37<br />

8.63 trans, cis-1,8-dimethylspiro[4,5]decane 8.19<br />

8.78 1,2-dimethyldecahydronaphthalene 11.49<br />

9.10 cis, cis-1,1-dimethylspiro[4,5]decane 11.60<br />

10.18 1,4,6-trimethyl-1,2,3,4-tetrahydronapthalene 1.77<br />

14.35 hexadecanoic acid, methyl ester 1.85<br />

- Other unknown components* 7.09<br />

*A total of 11 components and their relative amounts was less than 1% of the mixture with low<br />

polarity<br />

Alkylresorcinols (Figure 6) are another kind of antifungal agent that were isolated<br />

and identified from etiolated rice seedlings (Suzuki et al., 1996; 1998). Antifungal<br />

activities of alkylresorcinol against spore germination of P. oryzae were correlated<br />

with their concentration and structure, but complete inhibition of their spore<br />

germination occurred only at a concentration of over 100.g.ml -1 .


ALLELOCHEMICALS FROM AGERATUM CONYZOIDES AND ORYZA SATIVA 201<br />

R=-(CH 2 ) 12 CH 3 , -(CH 2 ) 14 CH 3 , -(CH 2 ) 16 CH 3 , -(CH 2 ) 7 CH=CH(CH 2 ) 5 CH 3 , -<br />

(CH 2 ) 7 CH=CH(CH 2 ) 7 CH 3 , -(CH 2 ) 7 CH=CHCH 2 CH=CH(CH 2 ) 4 CH 3<br />

Figure 6. Alkylresorcinols.<br />

3.2. Diterpene and flavone phytoalexins from rice<br />

<strong>Allelochemicals</strong> play an important role in rice disease resistance (Bailey and Mansfield,<br />

1982). Antifungal allelochemicala, phytoalexins produced by rice in response to injury,<br />

physiological stimuli or in the presence of infectious agents (Hammerschmidt, 1999).<br />

In particular, phytoalexins can be induced and accumulated by rice after fungal<br />

infection. Phytoalexins from rice mainly involve two types of diterpenes and flavones,<br />

including momilactones A and B, oryzalexins A-F and S, phytocassanes A-E and<br />

sakuranetin (Figures 7, 8).<br />

Momilactone A Momilactone B Sakuranetin<br />

Oryzalexin A: R 1 =OH, R 2 =CH 3 Oryzalexin E: R=CH 3 Oryzalexin S<br />

Oryzalexin B: R 1 =O, R 2 =OH Oryzalexin F: R=CH 2 OH<br />

Oryzalexin C: R 1 =O, R 2 =O<br />

Oryzalexin D: R 1 =OH, R 2 =O<br />

Figure 7. Typical diterpene and flavone phytoalexins from rice.


202<br />

CHUIHUA KONG<br />

Momilactones A and B are the first phytoalexins characterized from any member of<br />

the Gramineae. They were isolated initially as plant inhibitors from rice husk (Kato et al.,<br />

1977), and were subsequently found to be produced by rice in response to either<br />

infection by the blast disease that caused by the pathogen Pyricularia oryzae or<br />

irradiation with UV light (Cartwright et al., 1977; 1981). They had significant<br />

phytoalexin-like activity in P. oryzae and Helminthosporium oryzae. More recently,<br />

momilactones A and B have been found in rice root exudates, which have been found<br />

to participate in defense against weeds (Kato-Noguchi et al., 2002; 2003; Kong et al.,<br />

2004d; Lee et al., 1999).<br />

Phytocassane B Phytocassane C<br />

(ED 50 =20ì g/mg) (ED 50 =4ì g/mg) (ED 50 =7ì g/mg)<br />

Phytocassane D Phytocassane E<br />

(ED 50 =25ì g/mg) (ED 50 =6ì g/mg)<br />

Figure 8. Structures of phytocassanes A-E and their antifungal activities (the ED 50<br />

values) on spore germination of Magnaporthe grisea.<br />

Oryzalexins A-F and S, phytocassanes A-E and sakuranetin were isolated from<br />

rice leaves infected with blast fungus, Magnaporthe grisea (Akatsuka et al., 1985;<br />

Kato et al., 1993; 1994; Kodama et al., 1992; Koga et al., 1995; 1997; Nakazato Y et<br />

al., 2000). Oryzalexins A-C strongly inhibited the spore germination of P. oryzae.<br />

Their ED 50 values were 130, 68 and 136 ppm, respectively. Complete inhibition of P.<br />

oryzae spore germination was observed at 200 ppm. Oryzalexins A-C also strongly<br />

suppressed the germ tube elongation of P. oryzae. Their ED 50 values on P. oryzae<br />

germ tube elongation were 35, 18 and 35 ppm, respectively (Akatsuka et al., 1985).<br />

Similarly, oryzalexins D-F and S significantly inhibited spore germination of the rice<br />

blast fungus (Kato et al., 1993; 1994; Kodama et al., 1992). Noteworthy, oryzalexins<br />

E had slightly lower antifungal activity than oryzalexins D, but higher than Oryzalexins<br />

A-C (Kato et al., 1993). Phytocassanes A-E are the diterpenes with a cassane skeleton


ALLELOCHEMICALS FROM AGERATUM CONYZOIDES AND ORYZA SATIVA 203<br />

(Figure 8). They were produced in rice leaves and stems with M. grisea and R. solani.<br />

Phytocassanes A-E had high antifungal activity against the pathogenic fungi, M.<br />

grisea and R. solani (Koga et al., 1995; 1997). Their ED 50 values in prevention of<br />

spore germination and germ tube growth of M. grisea were very low (Figure 8).<br />

Sakuranetin is a flavonoid-type phytoalexin that was identified from rice plants. It<br />

had high antifungal activity and a large amount of accumulation in rice leaves<br />

(Nakazato et al., 2000).<br />

It has been shown in host-pathogen interactions that resistance reactions can be<br />

triggered by a large number of abiotic and biotic factors. Among the chemical factors,<br />

macromolecules of microbial origin are very important. Plant defense responses are<br />

stimulated by very low concentrations of these molecules (Darvill and Albersheim,<br />

1984). It was confirmed that a chemical for plant disease control might function by<br />

activating the natural resistance mechanisms of the host, for example, 2,2-dichloro-<br />

3,3-dimethyl cyclopropane carboxylic acid may exert its systemic fungicidal activity<br />

against the rice blast disease caused by P. oryzae (Cartwright et al., 1977). The<br />

application of methionine on wounded rice leaves induced the production of rice<br />

phytoalexins, sakuranetin and momilactone A. In the paddy field, methionine treatment<br />

has been demonstrated to reduce rice blast (Nakazato et al., 2000).<br />

An increasing number of studies have shown that rice phytoalexins are induced<br />

by elicitor that are produced by pathogenic microorganisms and make field disease<br />

control by inducing the pathogen defense mechanism in rice (Schaffrath et al., 1995;<br />

Tamogami et al., 1997a,b). Elicitors have been investigated extensively. It has been<br />

shown that jasmonic acid and its related compounds play important roles as the<br />

signaling molecules that elicit the production of phytoalexins in rice. Sakuranetin<br />

production may be elicited by exogenously applied jasmonic acid in rice leaves.<br />

Furthermore, sakuranetin production by exogenously applied jasmonic acid was<br />

significantly counteracted by amino acid, cytokinin, kinetin and zeatin (Tamogami et<br />

al., 1997a,b).<br />

4. CONCLUSION<br />

Many interactions between plant pathogens and their hosts are allelopathic.<br />

<strong>Allelochemicals</strong> can be applied in biological control of weeds and plant diseases (Rice,<br />

1995). Our research suggested that allelochemicals produced and released from A.<br />

conyzoides intercropping in citrus orchard did play important roles in integrated pest<br />

management. Many kinds of allelochemicals in rice not only inhibit the germination<br />

and growth of weeds, but also participate in the defense against pathogens. However,<br />

it remains unclear which allelochemicals in rice are predominantly involved in defense<br />

mechanisms against the pathogens. Therefore, further clarification of the resistance<br />

mechanism and multiple functions of rice allelochemicals are warranted.<br />

Acknowledgments : The author thanks Dr. Inderjit and anonymous reviewers for<br />

thoughtful criticisms and correcting English on earlier versions of the manuscript. The<br />

work was supported by National Natural Science Foundation of China (NSFC<br />

No.30170182; 30430460) and Hundreds-Talent Program of Chinese Academy of Sciences.


204<br />

CHUIHUA KONG<br />

5. REFERENCES<br />

Adesogan, E.K., Okunade, A.L. A new flavone from Ageratum conyzoides. Phytochemistry 1979;<br />

18:1863-1864.<br />

Akatsuka, T., Kodama, Q, Sekido, H., Kono, Y, Takeuchi, S. Novel phytoalexins ( Oryzalexins A, B and C)<br />

isolated from rice blast leaves infected with Pyricularia oryxae. Part I: Isolation, characterization and<br />

biological activities of oryzalexins. Agric Biol Chem 1985; 49:1689-1694.<br />

Albersberg, W. G. L. Singh, Y. Essential oil of Ageratum conyzoides. Flavor Fragrance J 1991; 6: 117-120.<br />

Bailey, J. A., Mansfield, J. W. eds. Phytoalexins. John Wiley and Sons Inc., New York, 1982.<br />

Biond, D., Cianci, P., Geraci, C. Antimicrobial activity and chemical composition of essential oil from Sicilian<br />

aromatic plants. Flavor Fragrance 1993; 8:331-337.<br />

Cartwright, D., Langcake, P., Pryce, R. J., Leworthy, D. P. Chemical activation of host defence mechanisms as<br />

a basis for crop protection. Nature 1977; 267:511-523.<br />

Cartwright, D., Langcake, P., Pryce, R.J., Leworthy, D.P. Ride, J.P. Isolation and characterization of two<br />

phytoalexins from rice as momilactones A and B. Phytochemistry 1981; 20:535-537<br />

Chung, I.M., Ahn, J.K., Yun, S.J. Identification of allelopathic compounds from rice (Oryza sative L.) straw<br />

and their biological activity. Can J Plant Sci 2001; 81:815-819.<br />

Dangl, J.L., Jones, J.D.G. Plant pathogens and integrated defence responses to infection. Nature 2001; 411:826-<br />

8335.<br />

Darvill, A.G., Albersheim, P. Phytoalexins and their elicitors-A defense against microbial infection in plants.<br />

Ann Rev Plant Pathol 1984; 35:243-275.<br />

Dixit, S.N., Chandra, H., Tiwari, R., Dixit, V. Development of a botanical fungicide against blue mould of<br />

mandarins. J Stored Prod Res 1995; 31:165-172.<br />

Dixon, R.A. Natural products and plant disease resistance. Nature 2001; 411:843-847.<br />

Ekundayo, O., Laakso, I., Hiltunen, R. Essential oil of A. conyzoides. Planta Med 1988; 54:55-57.<br />

Gonzalez, A.G., Aguiar, Z.E., Grillo, T.A., Luis, J.G., Rivera, A., Calle, J. Methoxyflavones from Ageratum<br />

conyzoides. Phytochemistry 1991; 30:1269-1271.<br />

Grayer, R.J. Harborne, J.B. A survey of antifungal compounds from higher plants. Phytochemistry 1994; 37:19-<br />

42.<br />

Hammerschmidt, R. Phytoalexins: what have we learned after 60 years? Annu Rev Phytopathol 1999; 37:285-<br />

306.<br />

Harborne, J.B. The comparative biochemistry of phytoalexin induction in plants. Biochem System Ecol 1999;<br />

27:335-367.<br />

Horie, T., Tominaga, H. Kawamura, Y. Revised structure of a natural flavone from Ageratum conyzoides.<br />

Phytochemistry 1993; 32:1078-1077<br />

Hu, F., Kong, C.H. Allelopathy of Ageratum conyzoides. VI. Effects of meteorological conditions on allelopathy<br />

of Ageratum conyzoides. Chin J Appl Ecol 2002a ; 13:76-80. (In Chinese)<br />

Hu, F., Kong, C.H. Inhibitory effect of flavones from Ageratum conyzoides on the major pathogens in citrus<br />

orchard. Chin J Appl Ecol 2002b ; 13:1166-1168 (In Chinese)<br />

Jhansi, P., Ramanujam, C.G.K. Pollen analysis of extracted and squeezed honey of Hyderabad. India Geophytol<br />

1987; 17:237–240.<br />

Kato, H., Kodama, O., Akatsuka, T. Oryzalexin E, a diterpene phytoalexin from UV-irradiat rice leaves.<br />

Phytochemistry 1993; 33:79-81.<br />

Kato, H., Kodama, O., Akatsuka, T. Oryzalexin F, a diterpene phytoalexin from UV-irradiat rice leaves.<br />

Phytochemistry 1994; 36:299-301.<br />

Kato-Noguchi, H., Ino, T., Sata, N., Yamamura, S. Isolation and identification of a potent allelopathic substance<br />

in rice root exudates. Physiol Planta 2002; 115:401-405.<br />

Kato-Noguchi, H., Ino, K. Rice seedlings release momilactone B into the environment. Phytochemistry 2003;<br />

63:551-554.<br />

Kim, K.U., Shin, D.H. eds. Rice Allelopathy Kyungpook National University. Taegu(Korea) 2000.<br />

Kato, T., Tsunakawa, M., Sasaki, N. Growth and germination inhibitors in rice husks. Phytochemistry 1977;<br />

16:45-48.<br />

Kodama, O., Li, W. X., Tamogami, S., Akatsuka, T. Oryzalexin S, a novel stemarane-type diterpene rice<br />

phytoalexin. Biosci Biotech Biochem 1992; 56:1002-1003.


ALLELOCHEMICALS FROM AGERATUM CONYZOIDES AND ORYZA SATIVA 205<br />

Koga, J., Ogawa, N., Yamauchi, T., Kikuchi, M., Ogasawara, N., Shimura, M. Functional moiety for<br />

the antifungal activity of phytocassane E, a diteroene phytoalexin from rice. Phytochemistry<br />

1997; 44:249-253.<br />

Koga, J., Shimura, M., Oshima, K., Ogawa, N., Yamauchi, T., Ogasawara, N. Phytocassanes A, B, C and D,<br />

novel diterpene phytoalexins from rice, Oryza sativa L. Tetrahedron 1995; 51:7907-7918.<br />

Kong, C.H., Hu, F., Xu, T., Lu, Y.H. Allelopathic potential and chemical constituents of volatile oil from Ageratum<br />

conyzoides. J Chem Ecol 1999; 25:2347-2356.<br />

Kong, C.H., Huang, S.S., Hu, F. Allelopathy of Ageratum conyzoides. V. <strong>Biologica</strong>l activities of the volatile oil<br />

from Ageratum on fungi, insects and plants and its chemical constituents. Acta Ecol Sin 2001; 21:584-<br />

587. (In Chinese).<br />

Kong, C.H., Hu, F., Xu, X. H. Allelopathic potential and chemical constituents of volatile from Ageratum<br />

conyzoides under stress. J Chem Ecol 2002a; 28:1173-1182.<br />

Kong, C.H., Xu, X.H., Hu, F., Cheng, X.H., Lin, B., Tan, Z.W. Using specific secondary metabolite as marker to<br />

evaluate allelopathic potential of rice variety and individual plant. Chin Sci Bull 2002b; 47:839-843.<br />

Kong, C.H., Hu, F., Liang, W.J., Wang, P., Jiang, Y. Allelopathic potential of Ageratum conyzoides with different<br />

growth stages and habitats. Allelopathy J 2004a ; 13:233-240.<br />

Kong, C.H., Hu, F., Xu, X.H., Liang, W.J., Zhang, C.X. Allelopathic plant: Ageratum conyzoides L. Allelopathy<br />

J 2004b; 14:5-13.<br />

Kong, C.H., Liang, W. J., Hu, F., Xu, X. H. Wang, P., Jiang, Y. <strong>Allelochemicals</strong> and their transformations in the<br />

Ageratum conyzoides intercropped the citrus orchard soil. Plant Soil 2004c; 264:149-157.<br />

Kong, C.H., Liang, W.J., Xu, X.H., Hu, F., Wang, P., Jiang, Y. Release and activity of allelochemicals from<br />

allelopathic rice seedlings. J Agric Food Chem 2004d ; 52:2861-2865.<br />

Kong, C.H., Xu, X. H., Zhou, B., Hu, F., Zhang, C.X., Zhang, M.X. Two compounds from allelopathic rice<br />

accession and their inhibitory effects on weeds and fungal pathogens. Phytochemistry 2004e ; 65:1123-<br />

1128.<br />

Kossmann, G., Groth, D. Plantea Infestantes e Nocivas. BASF Brasileira, São Paulo, 1993.<br />

Ladeira, A.M., Zaidan, L.B.P., Figueiredo, Ribeiro, R.D.C.L. Ageratum conyzoides L.(Compositae): germination,<br />

flowering and occurrence of phenolic derivatives at different stages of development. Hoehnea 1987; 14:53-<br />

62.<br />

Lee, C.W., Yoneyama, K., Takeuchi, Y., Konnai, M., Tamogami, S., Kodama, O. Momilactones A and B in rice<br />

straw harvested at different growth stages. Biosci Biotech Biochem 1999; 63:1318-1320.<br />

Liang, W.G., Hunag, M.D. Influence of citrus orchard ground cover plants on anthropod communities in China:<br />

a review. Agric Ecosyst Environ 1994; 50:29-37.<br />

Mattice, J., Lavy, T., Skulman, B., Dilday, R. Searching for allelochemicals in rice that control ducksalad. In<br />

Allelopathy In Rice; Olofsdotter, M. Ed. International Rice Research Institute: Manila, Philippines, 1998;<br />

pp. 81-97.<br />

Menut, C., Lamaty, G., Zollo, P.H.A. Aromatic plants of tropical central Africa: Part X. Chemical composition<br />

of essential oil of Ageratum houstonianum and Ageratum conyzoides from Cameroon. Flavor Fragrance<br />

J 1993; 8:1-4.<br />

Ming, L.C. Ageratum conyzoides: A tropical source of medicinal and agricultural products. In: Perspectives<br />

on New Crops and New Uses, Janick, J. eds. Alexandria, VA USA: ASHS Press 1999; pp 469-473.<br />

Muller, K.O., Borger, H. Experimentelle Untersuchungen uber die Phytophthora-Resistenz der Kartoffel-zugleich<br />

ein Bewitrag zum Problem der “erworbenen Resistenz” im Pflanzenreich. Arb Biol Anst Reichsanst (Berl)<br />

1940; 23:189-231.<br />

Nakazato, Y., Tamogami, S., Kawai, H., Hasegawa, M., Kodama, O. Methionine-induced phytoalexin production<br />

in rice leaves. Biosci Biotech Biochem 2000; 64:577-583.<br />

Okunade, A. L. Ageratum conyzoides L.(Asteraceae). Fitoterapia 2002; 73:1-16.<br />

Olofsdotter, M., Rebulanan, M., Madrid, A., Wang, D.L., Navarez, D., Olk, D.C. Why phenolic acids are<br />

unlikely primary allelochemicals in rice. J Chem Ecol 2002; 28:229-241.<br />

Pari, K., Rao, P. J., Subrahmanyam, B., Rasthogi, J. N., Devakumar, G. Benzofuran and other constituents of<br />

the essential oil of Ageratum conyzoides. Phytochemistry 1998; 49:1385-1388.<br />

Rao, G. P., Pandey, A. K., Shukla, K. Essential oils of some higher plants vis-à-vis some legume viruses. Indian<br />

Perfumer 1986; 30:483-486.<br />

Rice, E.L. <strong>Biologica</strong>l Control of Weeds and Plant Diseases: Advances in Applied Allelopathy. University of


206<br />

CHUIHUA KONG<br />

Oklahoma Press, Norman, OK, 1995.<br />

Rimando, A.M., Olofsdotter, M., Dayan, F.E., Duke, S.O. Searching for rice allelochemicals: An<br />

example of bioassay-guided isolation. Agron J 2001; 93:16-20.<br />

Schaffrath, U., Scheinpflug, H., Reisener, H.J. An elicitor from Pyricularia oryzae induced resistance<br />

responses in rice: isolation, Characterization and physiological properties. Physiol Mol Plant<br />

Pathol 1995; 46:293-307.<br />

Stadler, J., Mungai, G., Brandl, R. Weed invasion in East Africa: insights from herbarium records. African J<br />

Ecol 1998; 36:15-22.<br />

Stuiver, M.H., Custers, J.H.H.V. Engineering disease resistance in plants. Nature 2001; 411:865-868.<br />

Suzuki, Y., Esumi, Y., Hyakutake, H., Kono, Y., Sakurai, A. Isolation of 5-(8.Z-heptadecenyl)-resorcinol from<br />

etiolated rice seedlings as an antifungal agent. Phytochemistry 1996; 41:1485-1489.<br />

Suzuki, Y., Esumi, Y., Saito T, Kishimoto Y, Morita T. Identification of 5-n-(2.-oxo)alkylresorcinols from etiolated<br />

rice seedlings. Phytochemistry 1998; 47:1247-1252.<br />

Tamogami, S., Rakwal, R., Kodama, O. Phytoalexin production by amino acid conjugates of jasmonic acis<br />

through induction of naringenin-7-o-methyltransferase, a key enzyme on phytoalexin biosynthesis in rice<br />

(Oryza sativa L.). FEBS Lett 1997a; 401:239-242.<br />

Tamogami, S., Rakwal, R., Kodama, O. Phytoalexin production elicited by exogenously applied jasmonic acid<br />

in rice leaves (Oryza sativa L.) is under the control of cytokinins and ascorbic acid. FEBS Lett 1997b;<br />

412:61-64.<br />

Wandji, J., Bissangou, M. F., Ouambra, J. M., Silou, T., Abena, A. A., Keita, A. The essential oil from Ageratum<br />

conyzoides. Fitoterapia 1996; 67, 427-431.


Index<br />

A<br />

Abutilon theophrasti, 38, 83, 148<br />

Acacia sedillense, 47<br />

Acalypha indica, 114<br />

Acidovorax delafieldii, 148<br />

Adenostoma fasciculatum, 55<br />

Aeschynomene sp., 17<br />

Ageratum conyzoides, 42, 92, 194, 195, 196, 197,<br />

198, 203<br />

Aggressiveness, 93 , 94<br />

Agrobacterium radiobacter, 63<br />

Agrobacterium sp. 130<br />

Alcaligenes piechaudii, 63<br />

Allium cepa, 57<br />

Alnus firmifolia, 51<br />

Alternanthera philoxeroides, 111<br />

Alternaria alternata, 111, 164<br />

Alternaria alternata f sp. lycopersici, 112<br />

Alternaria eichhorniae, 111<br />

Alternaria kikuchiana, 178<br />

Alternaria panax, 164, 165<br />

Alternaria solani, 37, 53<br />

Amaranthus hypochondriacus, 37, 46, 53<br />

Amaranthus retroflexus, 43, 92<br />

Amaranthus spp., 57<br />

Amblyseius spp, 196<br />

Ambrosia cumanensis, 38<br />

Ammania coccinea, 104<br />

Anabaena sp, 110<br />

antifungal, 165, 178, 200, 203<br />

Aphanomyces euteiches, 183<br />

Arabidopsis thaliana, 87, 160<br />

Arachis hypogaea, 56, 92<br />

Arthrobacter spp., 66<br />

Ascochyta caulina, 83<br />

Aspergillus flavus, 9<br />

Aspergillus fumigatus, 9<br />

Aspergillus niger, 9<br />

Aspergillus parasiticus, 9<br />

Aspergillus sp., 68<br />

Avena fatua, 92<br />

Avena sativa, 64, 86<br />

Azadirachta indica, 21<br />

Azolla nilotica, 108<br />

B<br />

Bacillus cereus, 66, 127<br />

Bacillus mycoides, 61<br />

Bacillus subtilis, 3, 66<br />

Bacillus sp., 130<br />

Belonolaimus longicaudatus, 65<br />

207<br />

Berula erecta, 51<br />

Beta vulgaris, 83<br />

Bievundimonas vesicularis, 63<br />

biocontrol , 66, 110, 111, 118, 124, 128, 129, 130,<br />

131, 132, 133, 134, 145, 146, 148, 149, 150,<br />

169 , 170, 183, 187<br />

bioherbicides, 148, 149, 150, 151<br />

biological control, 25, 56, 70, 71, 111, 125, 131, 134,<br />

143, 145, 146, 148, 151, 152, 182<br />

biopesticide, 16, 15, 17<br />

biotrophic, 80, 93, 94<br />

biotrophs, 79<br />

Botrytis cinerea, 3, 5, 7, 8, 9, 92, 130, 159, 164,<br />

195<br />

Botrytis sp., 3, 12<br />

Brassica juncea, 42, 108<br />

Brassica kaber, 50<br />

Brassica napus, 87<br />

Brassica rapa var. pervidis, 38<br />

Brassica spp., 21<br />

Bromus japonicus, 148<br />

Bromus mollis, 92<br />

Bromus tectorum, 147<br />

Burkholderia cepacia, 66<br />

Burkholderia pickettii, 63<br />

Burkholderia sp., 20<br />

C<br />

Caenorhabditis elegans, 23<br />

Callicarpa acuminata, 37<br />

Calotropis gigantea, 106<br />

Canavalia ensiformis, 17, 71<br />

Capsella bursa-pastoris, 84, 152<br />

Capsicum annuum, 42, 57<br />

Capsicum frutescens, 9, 45<br />

Cassia fasciculata, 17<br />

Cassia obtusifolia, 57<br />

Cassytha sp., 108<br />

Centaurea solstitialis, 107<br />

Ceratophyllum demersum, 108<br />

Cercospora alternantherae, 111<br />

Cercospora amboinicus,114<br />

Cercospora kaki, 183<br />

Cercospora sp., 111<br />

Chaetomella raphigera, 111<br />

chemotaxis, 18<br />

Chenopodium album, 43, 83, 149<br />

Chondrilla juncea, 83<br />

Cicer arietinum, 86<br />

Cichorium intybus, 177<br />

Cirsium maculosa, 45


208<br />

Cirsium sp., 43<br />

Cladosporium herbarum, 178<br />

Cleistopholis acutatum, 12<br />

Cleistopholis fragariae, 11<br />

Cleistopholis patens, 9<br />

Cleome viscosa, 114<br />

Coleus amboinicus, 109, 114 , 115, 116, 117<br />

Colletotrichum acutatum, 3, 5, 7, 8, 9, 12<br />

Colletotrichum coccodes, 83<br />

Colletotrichum fragariae , 3, 5, 8 , 9, 11<br />

Colletotrichum gloeosporioides, 3, 5, 8, 9, 10<br />

Colletotrichum sp., 3, 4, 8, 12<br />

commensal, 124, 125, 127, 128<br />

compatible, 57<br />

contamination, 143, 149<br />

Criconemella sp., 65<br />

Crotalaria juncea, 17, 20, 57, 69, 70, 71<br />

Crotalaria retusa, 71<br />

Crotalaria spectabilis, 17, 70<br />

Crotalaria sp.,16, 65, 67, 69, 71<br />

Croton sparsiflorus, 114<br />

Cucumis sativus cv. Wisconsin, 47<br />

Cucumis sativus, 92<br />

Cucurbita pepo, 47<br />

Cuminum cyminum, 86<br />

Curcuma longa, 114<br />

Cyamopsis tetragonoloba, 52<br />

Cylindrocarpon destructans, 164, 165, 169, 170<br />

Cylindrocarpon sp., 170<br />

Cynoches sp., 3<br />

Cynodon dactylon, 107<br />

Cyperus difformis,104, 106, 117<br />

Cyperus littoralis, 106<br />

Cyperus rotundus, 106, 111<br />

D<br />

Dactylaria higginsi, 111<br />

Dactylis glomerata, 88, 92<br />

Daucus carota, 86<br />

Delonix regia, 35<br />

Digitaria decumbens, 35<br />

Diospyros lotus, 183<br />

Drechmeria coniospora, 69<br />

Drecshlera sp., 111<br />

E<br />

Echinochloa crus-galli, 37, 46, 104, 111<br />

Echinochloa spp, 106<br />

Eclipta alba, 114<br />

Egeria densa, 111<br />

Eichhornia crassipes, 108, 109, 114<br />

elicitors, 203<br />

Empetrum hermaphroditum, 41<br />

Enterobacter asburiae, 63<br />

epidemics, 88<br />

INDEX<br />

Erwinia carotovora, 87, 126, 127<br />

Erwinia sp, 130<br />

Erysiphe cichoracearum, 92, 195<br />

Eucalyptus globulus, 106<br />

Eucalyptus sp., 106<br />

Euphorbia esula, 147<br />

Euphorbia hirta, 114<br />

Exostema caribaeum, 47<br />

Exserohilum monocerus, 111<br />

exudate, 19, 35, 40, 69, 87, 107, 124, 148, 160, 168,<br />

170, 184, 199<br />

F<br />

Fagus sylvatica, 47<br />

Festuca arundinacea, 91<br />

fungistasis, 84<br />

fungitoxic, 157, 158, 185<br />

Fusarium graminearum, 111<br />

Fusarium oxysporum f. sp. cucumerinum, 130<br />

Fusarium oxysporum f.sp. cumini, 86<br />

Fusarium oxysporum f. sp. lycopersici, 159<br />

Fusarium oxysporum, 3, 5, 8, 9, 64, 66, 86, 164,<br />

165, 187<br />

Fusarium sambucinum, 183<br />

Fusarium solani, 23, 164, 165 , 183, 196, 197<br />

Fusarium sp., 60, 61, 165, 170, 183<br />

G<br />

Gaeumannomyces graminis var. tritici, 60<br />

Gaeumannomyces graminis, 158, 159<br />

Gigaspora rosea, 186<br />

Globodera pallida, 67<br />

Globodera rostochiensis, 18<br />

Glomus etunicatum, 185<br />

Glomus fasciculatum, 183<br />

Glomus intraradices, 68, 183, 186<br />

Glomus mosseae, 68, 183, 186, 187<br />

Glomus versiforme, 183<br />

Glomus sp., 183<br />

Glycine max, 56, 65, 66, 69, 83<br />

Gnomonia comari, 7, 12<br />

Gossypium spp., 56<br />

H<br />

Hebeloma crustuliniforme, 44<br />

Helianthus annuus, 39, 57, 86<br />

Helicobacter pylori, 49<br />

Helicotylenchus dihystera, 64, 65<br />

Helicotylenchus sp.,24, 65<br />

Helminthosporium longirostratum, 37<br />

Helminthosporium oryzae, 202<br />

Heteranthera limosa, 104<br />

Heterodera glycines, 24, 64, 65<br />

Heterodera schachtii, 22, 62<br />

Hintonia latiflora, 47


Hippeastrum hybridum, 48<br />

Hordeum leporinum, 92<br />

Hymenoscyphus ericae, 44<br />

Hyrdilla verticilata, 107, 108<br />

I<br />

Imperata cylindrica, 107<br />

Indigofera hirsuta, 17, 70, 71<br />

Indigofera nummularifolia, 71<br />

Indigofera spicata, 71<br />

Indigofera suffruticosa, 71<br />

Indigofera tinctoria, 71<br />

inoculum, 12, 63, 83, 111, 125, 145, 183, 187<br />

Ipomoea batatas, 54<br />

Ipomoea tricolor, 46<br />

Ipomoea sp., 57<br />

Iva axillaris, 38<br />

J<br />

Juglans nigra, 82<br />

Juncus sp., 51<br />

K<br />

Kocuria kristinae, 63<br />

Kocuria varians, 63<br />

L<br />

Laccaria bicolor, 184<br />

Lactuca sativa, 37, 41, 83<br />

Lactuca serriola, 89<br />

Lantana camara, 47<br />

Larrea tridentata, 41<br />

Lemna leucocephala, 114<br />

Lemna paucicostaha, 108<br />

Lemna perpusilla, 108<br />

Leptinotarsa decemlineata, 37<br />

Leptosphaeria maculans, 87<br />

Leptosphaerulina trifolii, 85<br />

Leucaena leucocephala, 35, 52, 106, 114<br />

Leucaena sp., 53, 106<br />

Leucas aspera, 114<br />

Linum usitatissimum, 57<br />

Lolium multiflorum, 92, 107<br />

Lolium perenne, 84, 87<br />

Lolium sp., 64<br />

Lycopersicon esculentum, 47, 53, 70, 82, 108<br />

Lythium salicaria, 107<br />

M<br />

Macaranga monandra, 9<br />

Macrophomina sp., 183<br />

Magnaporthe grisea, 202, 203<br />

Manihot esculenta, 107<br />

Medicago truncatula, 185<br />

Meloidogyne arenaria, 18, 20, 24, 65<br />

Meloidogyne hapla, 18<br />

INDEX 209<br />

Meloidogyne incognita, 18, 20, 21, 23, 24, 25, 61,<br />

63, 65, 66, 68, 69, 70<br />

Meloidogyne javanica, 18, 20, 21, 22, 23, 68, 70<br />

Meloidogyne spp, 19, 18, 24, 64, 65, 71<br />

Microbacterium esteraromaticum, 63<br />

Mucuna deeringiana, 17, 20, 62, 64, 66, 70<br />

mycoherbicide, 111<br />

Mycosphaerella sp, 111<br />

Myriophyllum aquaticum, 111<br />

Myriophyllum spicatum, 110<br />

Myrothecium roridum, 111<br />

Myrothecium sp., 68<br />

N<br />

Najas graminea, 108<br />

necrotrophic, 182<br />

necrotrophs, 79<br />

nematophagous, 67, 69<br />

Neochetina bruchii, 118<br />

Neotyphodium coenophialum, 91<br />

Neotyphodium lolii, 90, 94<br />

Nicotiana tabaccum, 57, 86<br />

Nimbya alternantherae, 111<br />

O<br />

Orobanche aegyptica, 57<br />

Orobanche cernua, 57, 40<br />

Orobanche crenata, 40, 57<br />

Orobanche cumana, 40<br />

Orobanche ramosa, 40, 57<br />

Orobanche sp., 40, 57<br />

Oryza sativa, 57, 194, 199<br />

P<br />

Paecilomyces sp., 20<br />

Paenibacillus sp, 130<br />

Panax cactorum, 164<br />

Panax ginseng, 162<br />

Panax quinquefolius, 157, 161<br />

Panax spp, 161<br />

Panicum miliaceum, 56<br />

Panonychus citri, 196<br />

Paratrichodorus minor, 19, 65<br />

Paratichodorus sp., 24<br />

Parthenium hysterophorous, 38, 53, 108, 114<br />

Paspalum sp., 17<br />

Pellicularia sasaki, 178<br />

Penicillium sp., 68<br />

Pennisetum americanum, 52<br />

Phaseolus aureus, 57<br />

Phaseolus mungo, 57<br />

Phaseolus vulgaris, 47, 86, 183<br />

Phaseolus sp., 41<br />

Phoenix dactylifera, 86<br />

Phoma lingam, 159


210<br />

Phoma typhae-domingensis, 111<br />

Phomopsis obscurans, 3, 5, 9<br />

Phomopsis viticola, 3, 5, 9<br />

Phomopsis sp., 8<br />

phyllosphere, 62<br />

phytoalexins, 3, 56, 85, 86, 159, 182, 184, 185, 186,<br />

193, 194, 201, 202, 203<br />

phytophagus, 34<br />

Phytophthora cactorum, 7, 66, 164,165,166, 169,<br />

170<br />

Phytophthora cinnamomi, 61<br />

Phytophthora citrophthora, 196, 197<br />

Phytophthora infestans, 86, 166<br />

Phytophthora nicotianae, 6, 9<br />

Phytophthora parasitica, 183<br />

Phytophthora sp., 5, 9, 166, 170, 183<br />

phytoprotection, 132<br />

phytotoxic, 25, 46, 52, 57, 64, 87, 106, 124, 160,<br />

199<br />

phytotoxin, 35, 42, 44, 45, 56, 57, 58, 59, 105, 111,<br />

147, 150<br />

Picea abies, 44<br />

Pinus laricio, 47<br />

Pistia stratiotes, 108, 109<br />

Pisum sativa, 86<br />

Plantago lanceolata, 43<br />

Pluchea lanceolata, 42<br />

Plutella xylostella, 54<br />

Portulaca oleracea, 57, 83<br />

Pratylenchus brachyurus, 68<br />

Pratylenchus neglectus, 21<br />

Pratylenchus penetrans, 63<br />

Pratylenchus vulnus, 22<br />

Pratylenchus sp., 19, 24, 65<br />

predators, 69, 181<br />

propagule, 147, 183<br />

Prunus dulcis, 23<br />

Psacalium decompositum, 37<br />

Pseudomonas aeruginosa, 130<br />

Pseudomonas aureofaciens, 127<br />

Pseudomonas chlororaphis, 63<br />

Pseudomonas fluorescens, 56, 63, 147<br />

Pseudomonas putida, 56, 148<br />

Pseudomonas syringae, 147<br />

Pseudomonas sp., 60, 61, 63, 66, 147<br />

Pseudopeziza coronata, 94<br />

Pseudopeziza medicagnis, 94<br />

Puccinia chondrillina, 83<br />

Puccinia coronata f.sp. avenae, 86<br />

Puccinia coronata f.sp. lolii, 88<br />

Puccinia coronata, 89, 92<br />

Puccinia graminis, 86, 92<br />

Puccinia hordei, 92<br />

Puccinia lagenophorae, 83<br />

INDEX<br />

Puccinia pulsatillae, 84<br />

Puccinia recondita, 92<br />

Pusatilla pratensis, 84<br />

Pyrenochaeta terrestris, 183<br />

Pyricularia oryzae, 178, 199, 200, 202<br />

Pythium aphanidermatum, 196, 197<br />

Pythium irregulare, 164, 165, 166,168,169, 170<br />

Pythium ultimum, 66, 164, 169<br />

Pythium sp., 50, 159, 166, 168, 170<br />

R<br />

Radopholus similis, 65<br />

Ralstonia solanacearum, 126<br />

Ralastonia sp., 127<br />

Ratibida mexicana, 46<br />

Renibacterium salmoninarum, 61<br />

Reynoutria sachalinensis, 3<br />

Rhizobium etli, 67<br />

Rhizobium japonicum, 111<br />

Rhizobium leguminosarum biovar phaseoli, 41<br />

Rhizobium leguminosarum, 183<br />

Rhizobium sp., 41, 51, 67, 170<br />

Rhizoctonia solani, 60, 64, 66, 92, 159, 164, 183,<br />

185, 187, 203,195,199<br />

rhizoplane, 108, 124, 131, 146, 147, 160, 169, 180,<br />

184<br />

rhizosphere, 17, 20, 23, 35, 36, 42, 44, 57, 58, 62,<br />

67, 68, 108, 131, 133, 134, 144, 146, 147, 148,<br />

151, 157, 169, 170, 183, 184, 185<br />

Ricinus communis, 17, 70<br />

rotation, 16, 17, 18, 19, 20, 22, 35, 41, 64, 66, 71,<br />

117, 134, 150, 152<br />

Rotylenchulus reniformis, 22, 24, 65, 70, 71<br />

Rumex crispus, 43<br />

Ruta graveolens, 8<br />

S<br />

Saccharum officinarum, 86<br />

Salvinia auriculata, 111<br />

Salvinia molesta, 108,109, 111<br />

Schistocerca gregaria, 177<br />

Sclerotinia sclerotiorum, 87, 92, 195<br />

Sclerotinia sp., 3<br />

Sclerotium cepivorum, 183<br />

Scutellonema sp., 65<br />

Sebastiania adenophora, 47<br />

Secale cereale, 57, 64<br />

Senecio vulgaris, 83<br />

Septoria lycopersici, 159<br />

Serratia plymuthica, 130<br />

Sesamum bicolor, 17<br />

Sesamum indicum, 17, 19, 57<br />

Sesbania aculeata, 117<br />

Sesbania exaltata, 107<br />

Sesbania grandiflora, 114


Setaria faberi, 151<br />

Setaria virdis, 148<br />

Sicyos deppei, 47<br />

Sida spinosa, 57<br />

Simsia amplexicaulis, 41<br />

Sinapis arvensis, 54<br />

Solanum nigrum, 43<br />

Solanum tuberosum, 62, 86<br />

Solanum, 86<br />

Sorghum bicolor,17, 19, 56, 57, 86<br />

Sorghum sudanense, 56<br />

Sorghum vulgare var, sudanense,17<br />

Sorghum vulgare, 17 , 52<br />

spermosphere, 144<br />

Sphenoclea zeylanica, 106<br />

Spirodella polyrhiza, 108<br />

Stellaria media, 149<br />

Stenotrophomonas sp., 130<br />

Streptococcus pneumoniae, 61<br />

Streptomyces hygroscopicus, 56, 112<br />

Streptomyces viridichromogens, 56, 111<br />

Streptomyces sp.,112, 130<br />

Striga asiatica, 40, 56<br />

Striga gesnerioides, 56<br />

Striga hermonthica, 32, 56, 152<br />

Striga sp., 40, 132<br />

suppression, 17, 18, 20, 21, 34, 55, 60, 61, 70, 72,<br />

84, 88, 89, 90, 91, 104, 105, 106, 107, 118,<br />

128, 132, 133, 149, 150, 151, 152<br />

susceptible, 4, 8, 13, 55, 68, 86, 88, 90, 93, 108,<br />

112, 160<br />

synergistic, 103, 118<br />

systemic acquired resistance, 13<br />

T<br />

Tagetes erecta, 18, 53, 62<br />

Tagetes patula, 18, 62<br />

Tagetes signata, 18<br />

Tagetes spp., 17, 63<br />

Tephrosia adunca, 71<br />

Thlaspi arvense, 149<br />

Thymus vulgaris, 24<br />

INDEX 211<br />

Tradescantia crassifolia, 41<br />

Trianthema portulacastum, 114<br />

Trichoderma hamatum, 165, 166, 169<br />

Trichoderma harzianum, 165<br />

Trichoderma viride, 165<br />

Trichoderma sp., 68, 165, 169<br />

Trifolium incarnatum, 57, 66<br />

Trifolium pratense, 50, 54, 85<br />

Trifolium repens, 84, 87<br />

Trifolium subterraneum, 57, 83, 92<br />

Trifolium sp., 64<br />

Triticum aestivum, 42, 54, 64, 66, 70, 86<br />

Tsukamurella paurometabolum, 63<br />

Tylenchorhynchus claytoni, 64, 65<br />

Tylenchulus semipenetrans, 21<br />

Tylenchorhynchus sp., 24<br />

Typha domingensis, 111<br />

U<br />

Uredo eichhorniae, 111<br />

Uromyces troflii-repentis, 92<br />

Urtica urens, 132<br />

V<br />

Vaccinium myrtillus, 44<br />

Variovorax paradoxus, 127<br />

Verticillium albo-atrum, 159<br />

Verticillium biguttatum, 64<br />

Vicia villosa, 64<br />

Vicia sp., 17<br />

Vigna catjang, 56<br />

Vigna unguiculata, 64, 66<br />

virulence, 93, 170<br />

Vitex negundo, 35<br />

Vitis vinifera, 3<br />

X<br />

Xanthium strumarium, 57<br />

Xanthomonas campestris, 126<br />

Xiphinema sp., 24<br />

Z<br />

Zea mays, 47, 50, 52, 56, 57, 83, 107


List of Contributors<br />

Ana Luisa Anaya<br />

Laboratorio de Alelopatía, Instituto de Ecología, Universidad Nacional<br />

Autónoma de México. Circuito Exterior, Ciudad Universitaria. 04510 México,<br />

D.F. Email: alanaya@miranda.ecologia.unam.mx<br />

Mark A. Bernards<br />

Department of Biology, University of Western Ontario, London, ON, Canada,<br />

N6A 5B7, Email: bernards@uwo.ca<br />

Ramanathan Kathiresan<br />

Professor, Department of Agronomy, Annamalai University, Annamalainagar,<br />

Tamilnadu 608 002, India. E-mail : rm.kathiresan@sify.com<br />

Cilfford H. Koger<br />

Weed Ecologist, Southern Weed Science Research Unit, USDA-ARS,<br />

P. O. Box 350, Stoneville, Mississippi 38776, USA.<br />

Nancy Kokalis-Burelle<br />

Research Ecologist, USDA, Agricultural Research Service, U.S. Horticultural<br />

Research Lab, Fort Pierce, FL, 34945, USA<br />

Email:NBurelle@ushrl.ars.usda.gov<br />

Chuihua Kong<br />

Institute of Applied Ecology, Chinese Academy of Sciences, Shenyang 110016,<br />

China, and South China Agricultural University, Guangzhou 510642, China.<br />

E-mail: kongch@mail.edu.cnw<br />

Robert J. Kremer<br />

U.S.D.A., Agricultural Research Service, Cropping Systems &<br />

Water Quality Unit, University of Missouri, Columbia, Missouri, U.S.A.<br />

Email: KremerR@missouri.edu<br />

Scott W. Mattner<br />

Department of Primary Industries, Knoxfield Centre, Victoria,<br />

Private Bag 15, Ferntree Gully Delivery Centre, 3156, VIC, Australia,<br />

scott.mattner@dpi.vic.gov.au<br />

Robert W. Nicol<br />

NovoBiotic Pharmaceuticals, Cambridge MA, 02138, USA<br />

Email: bernards@uwo.ca<br />

Hiroyuki Nishimura<br />

Department of Biosciences and Technology, School of Engineering, Hokkaido<br />

Tokai University, Sapporo 005-8601, Japan<br />

Email:nishimura@db.htokai.ac.jp<br />

213


214<br />

LIST OF CONTRIBUTORS<br />

Krishna N. Reddy<br />

Plant Physiologist, Southern Weed Science Research Unit, USDA-ARS,<br />

P. O. Box 350, Stoneville, Mississippi 38776, USA;<br />

Email: kreddy@ars.usda.gov<br />

Rodrigo Rodríguez-Kábana<br />

Distinguished University Professor, Auburn University, Auburn,<br />

AL, 36849, USA. Email : NBurelle@ushrl.ars.usda.gov<br />

Atsushi Satoh<br />

Department of Biosciences and Technology, School of Engineering, Hokkaido<br />

Tokai University, Sapporo 005-8601, Japan Email:nishimura@db.htokai.ac.jp<br />

Barbara J. Smith<br />

Small Fruit Research Station, 306 S. High St., Poplarville, MS 39470.<br />

Email : dwedge@olemiss.edu<br />

Antony V. Sturz<br />

Prince Edward Island Department of Agriculture, Fisheries, Aquaculture and<br />

Forestry, P.O. Box 1600, Charlottetown, P.E.I., C1A 7N3, Canada.<br />

Tel: 902-368-5664, FAX: 902-368-5661, E-mail:avsturz@gov.pe.ca<br />

David E. Wedge<br />

United States Department of Agriculture, Agricultural Research Service,<br />

Natural Products Utilization Research Unit, The Thad Cochran National<br />

Center for Natural Products Research, University of Mississippi, University,<br />

MS 38677, USA Email : dwedge@olemiss.edu<br />

Lina F. Yousef<br />

Department of Biology, University of Western Ontario, London, ON, Canada,<br />

N6A 5B7<br />

Ren-Sen Zeng<br />

Chemical Ecology Lab., Institute of Tropical & Subtropical Ecology,<br />

Agricultural College, South China Agricultural University Wushan,<br />

Tianhe District, Guangzhou, 510642 P.R. China<br />

Email: rszeng@scau.edu.cn, zengrs8@hotmail.com

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!