04.07.2015 Views

Mycosphaerella leaf spot diseases of bananas - CBS

Mycosphaerella leaf spot diseases of bananas - CBS

Mycosphaerella leaf spot diseases of bananas - CBS

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>:<br />

present status<br />

and outlook<br />

Proceedings <strong>of</strong> the 2 nd International workshop<br />

on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> held in San José,<br />

Costa Rica, 20-23 May 2002<br />

L. Jacome, P. Lepoivre, D. Marin, R. Ortiz, R. Romero<br />

and J.V. Escalant, editors


The mission <strong>of</strong> the International Network for the Improvement <strong>of</strong> Banana and Plantain (INIBAP) is<br />

to sustainably increase the productivity <strong>of</strong> banana and plantain grown on smallholdings for domestic<br />

consumption and for local and export markets.<br />

The programme has four specific objectives:<br />

•To organize and coordinate a global research effort on banana and plantain, aimed at the development,<br />

evaluation and dissemination <strong>of</strong> improved cultivars and at the conservation and use <strong>of</strong><br />

Musa diversity<br />

•To promote and strengthen collaboration and partnerships in banana-related research activities<br />

at the national, regional and global levels<br />

•To strengthen the ability <strong>of</strong> NARS to conduct research and development activities on <strong>bananas</strong> and<br />

plantains<br />

•To coordinate, facilitate and support the production, collection and exchange <strong>of</strong> information and<br />

documentation related to banana and plantain.<br />

INIBAP is a programme <strong>of</strong> the International Plant Genetic Resources Institute (IPGRI), a Future<br />

Harvest centre.<br />

The International Plant Genetic Resources Institute is an autonomous international scientific<br />

organization, supported by the Consultative Group on International Agricultural Research (CGIAR).<br />

IPGRI’s mandate is to advance the conservation and use <strong>of</strong> genetic diversity for the well being <strong>of</strong><br />

present and future generations. IPGRI’s headquarters is based in Rome, Italy, with <strong>of</strong>fices in another<br />

19 countries worldwide. It operates through three programmes: (1) the Plant Genetic Resources<br />

Programme, (2) the CGIAR Genetic Resources Support Programme, and (3) the International Network<br />

for the Improvement <strong>of</strong> Banana and Plantain (INIBAP).<br />

The international status <strong>of</strong> IPGRI is conferred under an Establishment Agreement which, by January<br />

2000, had been signed and ratified by the Governments <strong>of</strong> Algeria, Australia, Belgium, Benin, Bolivia,<br />

Brazil, Burkina Faso, Cameroon, Chile, China, Congo, Costa Rica, Côte d’Ivoire, Cyprus, Czech Republic,<br />

Denmark, Ecuador, Egypt, Greece, Guinea, Hungary, India, Indonesia, Iran, Israel, Italy, Jordan, Kenya,<br />

Malaysia, Mauritania, Morocco, Norway, Pakistan, Panama, Peru, Poland, Portugal, Romania, Russia,<br />

Senegal, Slovakia, Sudan, Switzerland, Syria, Tunisia, Turkey, Uganda and Ukraine.<br />

Cover illustration: Microscope photo <strong>of</strong> the causal agent <strong>of</strong> black <strong>leaf</strong> streak disease, Paracercospora fijiensis<br />

(photo: J. Carlier, CIRAD).<br />

Citation: Jacome L., P. Lepoivre, D. Marin, R. Ortiz, R. Romero and J.V. Escalant (eds). 2003. <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook. Proceedings <strong>of</strong> the Workshop on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

held in San Jose, Costa Rica on 20-23 May 2002. The International Network for the Improvement <strong>of</strong><br />

Banana and Plantain, Montpellier, France.<br />

INIBAP-ISBN: 2-910810-57-7<br />

© International Plant Genetic Resources Institute 2003<br />

IPGRI Headquarters<br />

INIBAP<br />

Via dei Tre Denari 472/a<br />

Parc Scientifique Agropolis II<br />

000 57 Maccarese (Fiumicino) 34397 Montpellier Cedex 5<br />

Rome, Italy<br />

France


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>:<br />

present status<br />

and outlook<br />

Proceedings <strong>of</strong> the 2 nd International workshop<br />

on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> held in San José,<br />

Costa Rica, 20-23 May 2002<br />

L. Jacome, P. Lepoivre, D. Marin, R. Ortiz, R. Romero<br />

and J.V. Escalant, editors


Acknowledgments<br />

INIBAP would like to thank all those who helped in the organization <strong>of</strong> the 2 nd<br />

International workshop on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> and contributed to<br />

the publication <strong>of</strong> these proceedings.<br />

CATIE, Chiquita, CORBANA, Dole, EARTH, Lapanday, Syngenta,TADECO and Total Fina<br />

Elf for their financial support to the organization <strong>of</strong> the meeting and the publication<br />

<strong>of</strong> these proceedings.<br />

Jorge A. Sandoval (CORBANA), Galileo Rivas (CATIE), Franklin Rosales and Luis<br />

Pocasangre (INIBAP-LAC) and Ronald Madrigal and Arllen Carpio (EARTH) for<br />

helping organize the workshop.<br />

Luis Jacome, Philippe Lepoivre, Douglas Marin, Rodomiro Ortiz and Ronald Romero<br />

for efficiently chairing the sessions and for their work as scientific editors.<br />

Jean-Vincent Escalant and Claudine Picq for overseeing the organization <strong>of</strong> the<br />

workshop and the production <strong>of</strong> these proceedings.<br />

Andrew Entwistle and Anne Vézina for the technical editing <strong>of</strong> these proceedings.<br />

3


Contents<br />

Foreword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

Introduction to workshop<br />

Overview <strong>of</strong> progress and results since the first international workshop<br />

on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong> in 1989<br />

X. MOURICHON . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

Session 1 – Impact <strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong><br />

Introduction - The spread, detection and impact <strong>of</strong> black <strong>leaf</strong> streak disease<br />

and other <strong>Mycosphaerella</strong> species in the 1990s<br />

R. A. ROMERO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

The distribution and importance <strong>of</strong> the <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> banana<br />

D. R. JONES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

Integrating morphological and molecular data sets on <strong>Mycosphaerella</strong>,<br />

with specific reference to species occurring on Musa<br />

P. W. CROUS, J. Z. GROENEWALD, A. APTROOT, U. BRAUN, X. MOURICHON and J. CARLIER . . . . . . . . . . . . . . . 43<br />

Improved PCR-based detection <strong>of</strong> Sigatoka disease and black <strong>leaf</strong> streak disease<br />

in Australian banana crops<br />

J. HENDERSON, K. GRICE, J. PATTEMORE, R. PETERSON and E. AITKEN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

Impact <strong>of</strong> minor <strong>Mycosphaerella</strong> pathogens on <strong>bananas</strong> (Musa) in South Africa<br />

A. VILJOEN, A. K. J. SURRIDGE and P. W. CROUS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

Economic impact and management <strong>of</strong> black <strong>leaf</strong> streak disease in Cuba<br />

L. PÉREZ VICENTE, J. M. ALVAREZ and M. PÉREZ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

Management <strong>of</strong> black <strong>leaf</strong> streak disease in tropical Asia<br />

A. B. MOLINA and E. FABREGAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

Impact <strong>of</strong> <strong>Mycosphaerella</strong> spp. in Brazil<br />

Z. J. MACIEL CORDEIRO and A. PIRES DE MATOS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

Poster - Fungi associated with banana foliage in South Africa<br />

A. K. J. SURRIDGE, A. VILJOEN and F. C. WEHNER . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

Session 2 – Population biology and epidemiology<br />

Introduction - Population biology and epidemiology<br />

L. H. JÁCOME . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

Airborne dispersal <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

P. J. A. BURT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

Genetic differentiation in <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens<br />

J. CARLIER, H. HAYDEN, G. RIVAS, M.-F. ZAPATER, C. ABADIE and E. AITKEN . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

Development and application <strong>of</strong> molecular markers<br />

in <strong>Mycosphaerella</strong> populations in Colombia<br />

C. MOLINA, S. APONTE, A. GUTIÉRREZ, V. NÚÑEZ and G. KAHL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

Poster - An electrophoretic karyotype for <strong>Mycosphaerella</strong> fijiensis<br />

L. CONDE-FERRÁEZ, CECILIA M. RODRÍGUEZ, L. PERAZA-ECHEVERRÍA and A. JAMES . . . . . . . . . . . . . . . . . . . . 141<br />

Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147<br />

5


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Session 3 – Host-pathogen interactions<br />

Introduction - Banana–<strong>Mycosphaerella</strong> fijiensis interactions<br />

P. LEPOIVRE, J. P. BUSOGORO, J. J. ETAME, A. EL HADRAMI, J. CARLIER, G. HARELIMANA,<br />

X. MOURICHON, B. PANIS, A. STELLA RIVEROS, G. SALLÉ, H. STROSSE and R. SWENNEN . . . . . . . . . . . . . . . . . 151<br />

Efficiency and durability <strong>of</strong> partial resistance against black <strong>leaf</strong> streak disease<br />

C. ABADIE, A. ELHADRAMI, E. FOURÉ and J. CARLIER . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161<br />

Poster - Early evaluation <strong>of</strong> black <strong>leaf</strong> streak resistance by using mycelial suspensions<br />

<strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

Y. ALVARADO CAPÓ, M. LEIVA MORA, M. A. DITA RODRÍGUEZ, M. ACOSTA, M. CRUZ,<br />

N. PORTAL, R. GÓMEZ KOSKY, L. GARCÍA, I. BERMÚDEZ and J. PADRÓN . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169<br />

Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176<br />

Session 4 – Genetic improvement for the management <strong>of</strong> resistance<br />

Introduction - Genetic improvement for a sustainable management <strong>of</strong> resistance<br />

K. CRAENEN and R. ORTIZ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181<br />

Conventional breeding <strong>of</strong> <strong>bananas</strong><br />

C. JENNY, K. TOMEKPÉ, F. BAKRY and J.V. ESCALANT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199<br />

Transgenic approaches for resistance to <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in Musa spp.<br />

R. SWENNEN, G. ARINAITWE, B.P.A. CAMMUE, I. FRANCOIS, B. PANIS, S. REMY, L. SÁGI , E. SANTOS,<br />

H. STROSSE and I. VAN DEN HOUWE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209<br />

Mutagenesis and somaclonal variation to develop new resistance<br />

to <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

N. ROUX, A. TOLOZA, J. P. BUSOGORO, B. PANIS, H. STROSSE, P. LEPOIVRE,<br />

R. SWENNEN and F. J. ZAPATA-ARIAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239<br />

Reaction <strong>of</strong> banana genotypes to black <strong>leaf</strong> streak disease in the State <strong>of</strong> Acre in Brazil<br />

M. DE J. B. CAVALCANTE, A. DA S. LEDO, F. H. S . COSTA, Z. J. M. CORDEIRO and A. P. MATOS . . . . . . . . . . . . 251<br />

The International Musa testing programme (IMTP): a worldwide programme<br />

to evaluate elite Musa cultivars<br />

J.-V. ESCALANT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257<br />

Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266<br />

Session 5 – Integrated disease management<br />

Management <strong>of</strong> Mycospharella <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in Australia<br />

R. PETERSON, K. GRICE and S. DE LA RUE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271<br />

Spread and management <strong>of</strong> black <strong>leaf</strong> streak disease in the Dominican Republic<br />

P. E. JORGE and T. POLANCO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277<br />

Microbiological control <strong>of</strong> black <strong>leaf</strong> streak disease<br />

A. STELLA RIVEROS, C. INÉS GIRALDO and A. GAMBOA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287<br />

Precision agriculture to improve management decisions and field research<br />

E. SPAANS and L. QUIROS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297<br />

Poster - The role <strong>of</strong> managing resistance to fungicides in maintaining<br />

the effectiveness <strong>of</strong> integrated strategies to control black <strong>leaf</strong> streak disease<br />

S. KNIGHT, M. WIRZ, A. AMIL, A. HALL and M. SHAW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303<br />

Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308<br />

List <strong>of</strong> participants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311<br />

6


Foreword<br />

The rapid expansion in the 1980s <strong>of</strong> black <strong>leaf</strong> streak disease, which is caused by<br />

<strong>Mycosphaerella</strong> fijiensis, resulted in such damage to small producers that it<br />

encouraged INIBAP to organize the 1 st International workshop on Sigatoka <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> <strong>of</strong> Banana held in San José, Costa Rica in March 1989. Coming 13 years<br />

after this meeting, the 2 nd International workshop on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

<strong>of</strong> <strong>bananas</strong> provided a timely opportunity to analyse the current situation regarding<br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> at the global level.<br />

Black <strong>leaf</strong> streak disease has been spreading for more than 20 years and is now<br />

reported in most parts <strong>of</strong> the world. During that time, considerable research<br />

efforts have been expended to develop alternatives to allow small producers to<br />

continue producing banana and plantain. Efforts to create new resistant varieties were<br />

part <strong>of</strong> a broad spectrum <strong>of</strong> activities including classical and modern tools for genetic<br />

improvement. Studies are ongoing to develop a better understanding <strong>of</strong> host-pathogen<br />

interactions. The epidemiology, distribution and population structure <strong>of</strong> the<br />

<strong>Mycosphaerella</strong> pathogens are being investigated at national, regional and international<br />

levels. Research is also conducted to develop new methods to control the<br />

disease based on a rational use <strong>of</strong> fungicides.<br />

All those who are involved and interested in the sustainability <strong>of</strong> the small banana<br />

and plantain producers know that the state <strong>of</strong> the research and the impact <strong>of</strong><br />

Mycospharella <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> have radically changed over the last decade. Sigatoka<br />

disease (caused by <strong>Mycosphaerella</strong> musicola) is still important in some parts <strong>of</strong> the<br />

world and a previously undescribed <strong>leaf</strong> <strong>spot</strong> disease, eumusae <strong>leaf</strong> <strong>spot</strong> disease<br />

(caused by <strong>Mycosphaerella</strong> eumusae), has recently been discovered in southern and<br />

southeastern Asia.<br />

By organizing this workshop, in collaboration with EARTH, CORBANA and CATIE<br />

and in the framework <strong>of</strong> PROMUSA, INIBAP hopes to strengthen collaborations to<br />

ensure that the benefits <strong>of</strong> the research efforts reach the smallholders and to accelerate<br />

the creation <strong>of</strong> new varieties resistant to Mycospharella <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>.<br />

7


Editorial note<br />

Throughout the proceedings, black <strong>leaf</strong> streak disease, also known as black Sigatoka, is used to<br />

refer to the disease caused by <strong>Mycosphaerella</strong> fijiensis, Sigatoka disease, also known as yellow<br />

Sigatoka, refers to the disease caused by <strong>Mycosphaerella</strong> musicola, and eumusae <strong>leaf</strong> <strong>spot</strong> disease,<br />

which was called Septoria <strong>leaf</strong> <strong>spot</strong> disease when first identified, refers to the disease caused by<br />

<strong>Mycosphaerella</strong> eumusae.


Introduction to workshop<br />

Overview <strong>of</strong> progress and results<br />

since the first international workshop<br />

on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

<strong>of</strong> <strong>bananas</strong> in 1989<br />

X. Mourichon<br />

At the San José workshop, 1989, the main topics about banana pathogens were:<br />

1) improvements in knowledge,<br />

2) geographical distribution,<br />

3) epidemiology,<br />

4) mechanisms <strong>of</strong> host-parasite interactions,<br />

5) sources <strong>of</strong> resistance and genetic improvement in Musa and<br />

6) efficacy <strong>of</strong> new fungicides and their use.<br />

This paper reviews and evaluates progress made in recent years, especially the<br />

major developments, and highlights topics in which efforts have perhaps not been<br />

sufficiently sustained.<br />

In general, the main results obtained on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> over<br />

the past 10 years have been widely published in refereed journals together with several<br />

reviews. The last book edited by CABI (Jones 2000) is a very good synthesis <strong>of</strong> present<br />

knowledge <strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>.<br />

CIRAD-AMIS, Montpellier, France<br />

11


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Identification, taxonomy and diagnosis<br />

In 1989, the discussions at the San José meeting concentrated on the availability <strong>of</strong><br />

diagnostic methods for the pathogens causing Sigatoka disease, black <strong>leaf</strong> streak disease<br />

and black Sigatoka. The main observation was the risk <strong>of</strong> confusing the species when<br />

diagnosis was based on the observation <strong>of</strong> symptoms alone.<br />

In situ or in vitro observation <strong>of</strong> the anamorph was considered to be the most reliable<br />

method to identify the <strong>diseases</strong>:<br />

1) the anamorph Cercospora musae, now Pseudocercospora musae, for Sigatoka<br />

disease caused by <strong>Mycosphaerella</strong> musicola,<br />

2) the anamorph Cercospora fijiensis, now Paracercospora fijiensis, for black <strong>leaf</strong><br />

streak disease caused by <strong>Mycosphaerella</strong> fijiensis,<br />

3) the two anamorphs for black Sigatoka which at the time were attributed<br />

to <strong>Mycosphaerella</strong> fijiensis var. difformis were in reality the anamorphs <strong>of</strong><br />

M. fijiensis and M. musae. The species was described only in Latin America.<br />

There has been much work on these aspects in recent years, using the methods<br />

developed for molecular taxonomy on different populations <strong>of</strong> the pathogens<br />

(Carlier et al., 1994, 2000; Johanson and Jeger, 1993; Johanson et al., 1994).<br />

Molecular markers are highly sensitive at discriminating between fungal<br />

species and have clarified the taxonomy <strong>of</strong> the banana pathogens.<br />

Molecular markers have made it possible to distinguish clearly between<br />

M. fijiensis and M. musicola and to confirm that M. fijiensis and M. fijiensis var.<br />

difformis are synonymous (Carlier et al., 1994).<br />

More recently, markers have been used to identify a new species pathogenic to<br />

banana, <strong>Mycosphaerella</strong> eumusae. Initially, M. eumusae was thought to have a<br />

Septoria anamorph (Carlier et al., 2000) therefore the new disease was called<br />

Septoria <strong>leaf</strong> <strong>spot</strong> disease. Additional detailed work on the morphotaxonomy<br />

attributed M. eumusae with a Pseudocercospora anamorph and the name was<br />

changed to eumusae <strong>leaf</strong> <strong>spot</strong> disease, ELSD (Crous and Mourichon, 2002).<br />

Other methods based on serology are still being developed. The methods are<br />

intended above all to be sufficiently simple to diagnose the early stages <strong>of</strong> disease,<br />

for example, within the framework <strong>of</strong> preventive control measures with rational<br />

use <strong>of</strong> fungicides (Etienne et al., 1995). More specific methods to identify<br />

pathogens as well as opportunist or non-pathogenic species <strong>of</strong> <strong>Mycosphaerella</strong> are<br />

also being developed.<br />

The use <strong>of</strong> molecular markers has provided a great deal <strong>of</strong> information in recent<br />

years. In particular, molecular markers have made it possible to perform many<br />

analyses <strong>of</strong> genetic diversity, mainly in M. fijiensis.<br />

A high level <strong>of</strong> genetic diversity has been demonstrated in M. fijiensis<br />

particularly in the genetic structure <strong>of</strong> populations at a macrogeographical<br />

scale (Carlier et al., 1996). The diversity and geographical distribution <strong>of</strong> popu-<br />

12


Introduction to workshop<br />

X. Mourichon<br />

lations are mainly explained by genetic recombination in the teleomorph <strong>of</strong><br />

<strong>Mycosphaerella</strong>. Genetic differences have also been observed at smaller scales e.g.<br />

at a field scale (Müller et al., 1997).<br />

M. musicola also shows considerable intraspecific diversity, again involving sexual<br />

reproduction (Hayden et al., 2000, 2002).<br />

There is universal agreement about the extent <strong>of</strong> genetic diversity in M. fijiensis<br />

and M. musicola, and the implications for the capacity <strong>of</strong> the two fungi to evolve.<br />

Thus, genetic diversity must be taken into account when devising strategies to<br />

improve disease resistance in banana. A knowledge <strong>of</strong> the genetic variation in<br />

different geographical populations <strong>of</strong> the pathogens is important for the management<br />

<strong>of</strong> resistance genes.<br />

Geographical distribution <strong>of</strong> <strong>Mycosphaerella</strong><br />

Black <strong>leaf</strong> streak disease (BLSD) and Sigatoka disease (SD) are widespread in the main<br />

banana production zones, particularly Southeast Asia, which is the zone <strong>of</strong> origin<br />

<strong>of</strong> the pathogens, and the Pacific, and also in Latin America and Africa (Jones, 2000).<br />

Since the San José workshop, the main change has been the rapid spread <strong>of</strong> BLSD<br />

in Latin America from Central America northwards to Mexico and Florida, USA, and<br />

southwards to Colombia, Peru, Venezuela, Bolivia and Brazil. BLSD has spread to<br />

the Caribbean e.g. Cuba, Jamaica, the Dominican Republic and Haiti, and the rest<br />

<strong>of</strong> the Caribbean arc is threatened.<br />

BLSD has also spread to the western central and eastern parts <strong>of</strong> the African<br />

continent and, recently, to Madagascar. M. fijiensis is no longer confined to northern<br />

Australia and is a new constraint that has to be managed in other Australian<br />

commercial plantations.<br />

Very little information is available on the geographical distribution <strong>of</strong> the new<br />

species, M. eumusae, cause <strong>of</strong> ELSD. The species appears to be centred in India but<br />

probably has a wider distribution and requires detailed study.<br />

Finally, a fourth species, M. musae, causing speckle disease, is widespread<br />

throughout the world but generally causes little damage except in the sub-tropical<br />

areas <strong>of</strong> Australia and South Africa.<br />

Epidemiology <strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

At the San José meeting, it was recognized that knowledge <strong>of</strong> the epidemiology <strong>of</strong><br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> was weak. Just one recommendation regarding<br />

“the urgent need to concentrate efforts on a better understanding <strong>of</strong> the different<br />

epidemiological components” was made.<br />

Nevertheless, important research (unpublished) had been done on various epidemiological<br />

aspects <strong>of</strong> BLSD in Latin America and Africa. The work was mainly on the effects<br />

<strong>of</strong> abiotic factors on the different phases <strong>of</strong> monocyclic infection by M. fijiensis i.e.<br />

infection processes, incubation period, rates <strong>of</strong> symptom development, reproduction and<br />

dispersal (Gauhl, 1994; Fouré, 1992; Rutter et al., 1998; Smith et al., 1997).<br />

At the time, this work aimed to make the control strategies more scientific but<br />

the question is whether the results were sufficiently exploited and used.<br />

13


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Other biotic factors e.g. genetic host resistance also affect parts <strong>of</strong> the infection<br />

cycle (see: management <strong>of</strong> genetic resistance).<br />

Host-pathogen interactions<br />

In the past, host resistance was evaluated in the field under natural, and hence variable,<br />

pathogen pressure, and in environmental conditions that differed considerably between<br />

locations. Environmental factors can have major effects on the expression <strong>of</strong> resistance,<br />

especially partial resistance.<br />

Variability was reduced by using tissue-cultured plants inoculated with different<br />

isolates <strong>of</strong> <strong>Mycosphaerella</strong>. The plants could be maintained in controlled environmental<br />

chambers but the system was criticised for such things as using host plants that were<br />

too young, which could give rise to resistant phenotypes that were not true-to-type<br />

and results that were not reproducible. A miniaturized technique was thus developed.<br />

Using <strong>leaf</strong> fragments kept under artificial survival conditions, made it possible to work<br />

with older leaves, to work under closely controlled environmental conditions, and to<br />

allow the different banana genotypes to express the type and level <strong>of</strong> resistance to<br />

M. fijiensis typical <strong>of</strong> those expressed in the field. Crucially, all stages <strong>of</strong> development<br />

in a monocyclic infection were represented in the miniaturised system. The method can<br />

be used to analyse the nature <strong>of</strong> host-parasite interactions and to study the variability<br />

in virulence and aggressiveness in different populations <strong>of</strong> the pathogens.<br />

Research over the last 10 years has made use <strong>of</strong> various models: banana plants in<br />

natural conditions, young plants or leaves under artificial survival conditions, providing<br />

complementary data about the nature <strong>of</strong> compatible and incompatible interactions.<br />

Host-pathogen interactions occur in nature as the following phenotypes: very or<br />

highly resistant banana cultivars and partially resistant <strong>bananas</strong> with resistance<br />

varying from very marked partial resistance to very susceptible. All banana germplasm<br />

can be classified by these types <strong>of</strong> behaviour (Fouré et al., 2000). Only a few wild, and<br />

cultivated diploid and triploid acuminata are resistant to BLSD. High resistance in the<br />

triploid acuminata is present only in cultivars <strong>of</strong> the Ibota subgroup.<br />

Phenotypes with different levels <strong>of</strong> partial resistance appear to be widely distributed<br />

among all diploid and triploid acuminata and balbisiana genotypes. Genome B seems<br />

to give higher levels <strong>of</strong> partial resistance.<br />

The idea that the relationships between <strong>Mycosphaerella</strong> and banana were based on<br />

compatible and incompatible interactions was proposed at the San José workshop.<br />

Compatible interactions refer to susceptible, or partially resistant <strong>bananas</strong> that display<br />

different degrees <strong>of</strong> partial resistance. M. fijiensis can complete its entire infection cycle<br />

under this type <strong>of</strong> interaction. It was also suggested that partial resistance might result<br />

from a constitutive polyphenolic compound.<br />

The various cytological, ultrastructural and biochemical studies, and genetic<br />

analyses <strong>of</strong> different varieties <strong>of</strong> banana with different levels <strong>of</strong> partial resistance, clearly<br />

show that proanthocyanidins play a role in partial resistance. It is also<br />

clear that the compounds are not essential for the expression <strong>of</strong> partial resistance.<br />

Proanthocyanidin is probably involved in the rate <strong>of</strong> lesion elongation, but other factors<br />

may be involved at other stages <strong>of</strong> the infection monocycle and should be identified<br />

(Beveraggi et al., 1995; Mourichon et al., 2000).<br />

14


Introduction to workshop<br />

X. Mourichon<br />

Using a model under controlled conditions, it is fairly easy to dissect partial resistance<br />

and evaluate the importance <strong>of</strong> certain sequences <strong>of</strong> monocyclic infection, e.g.<br />

incubation period, rate <strong>of</strong> lesion development, effectiveness <strong>of</strong> infection, latent periods<br />

and different parameters <strong>of</strong> sexual and asexual sporulation. Recent studies suggest that<br />

the presence <strong>of</strong> several components <strong>of</strong> resistance act on different stages <strong>of</strong> the infection<br />

cycle. Depending on the banana cultivar, partial resistance may be the result <strong>of</strong> different<br />

mechanisms. Thus, similar expressions <strong>of</strong> partial resistance could depend on different<br />

genetic interactions, e.g. efficacy <strong>of</strong> infection or level <strong>of</strong> sexual reproduction. These<br />

possibilities should be taken into account by breeders, for whom partial resistance is a<br />

major objective.<br />

Partial resistance in banana is important because it is considered to be more durable.<br />

However, the great diversity <strong>of</strong> pathogen populations and their capacity to evolve should<br />

be taken into account. Specific interactions between the pathogen and the host plant<br />

must be established for certain sequences <strong>of</strong> the infectious monocycle (Abadie et al.,<br />

2001a, b). It is possible that some specific interactions may select for more aggressive<br />

pathotypes <strong>of</strong> <strong>Mycosphaerella</strong>. The result would be gradual erosion, rather than a sudden<br />

decrease, <strong>of</strong> partial resistance.<br />

Incompatible interactions. Banana cultivars that are very resistant rapidly block<br />

progress <strong>of</strong> the fungus in the early stages <strong>of</strong> disease. At the San José workshop, it was<br />

suggested that such behaviour could be governed by an active defence mechanism.<br />

Studies <strong>of</strong> host-parasite interactions using cytology, particularly at the ultrastructural<br />

scale, provided accurate images <strong>of</strong> the interactions that occur after inoculation. There<br />

is clear evidence <strong>of</strong> active mechanisms such as cell collapse occurring after penetration<br />

<strong>of</strong> stomata (Beveraggi et al., 1995; Mourichon et al., 2000). Similarly, hypersensitive<br />

reactions have been elicited, and necrosis induced experimentally by fungal compounds<br />

<strong>of</strong> high molecular weight.<br />

Other research reported at the San José workshop demonstrated that phytotoxic<br />

compounds or toxins were released by M. musicola and M. fijiensis. This raised the<br />

question <strong>of</strong> the role <strong>of</strong> these compounds in the infection process. Breeders were interested<br />

in the use <strong>of</strong> toxic compounds in schemes for the early selection <strong>of</strong> <strong>bananas</strong> resistant<br />

to M. fijiensis, in particular.<br />

During the last decade, a large number <strong>of</strong> phytotoxic compounds produced by<br />

M. fijiensis have been described in the literature, e.g. juglone which displays a high<br />

level <strong>of</strong> biological activity (Stierle et al., 1991).<br />

Several research projects have shown that such compounds are not primary<br />

determinants <strong>of</strong> the disease but are secondary determinants <strong>of</strong> pathogenicity (Harelimana<br />

et al., 1997). The role <strong>of</strong> these compounds as agents in the selection <strong>of</strong> resistance, as<br />

originally considered, deserves discussion, in particular for studies on partial resistance.<br />

Breeding for resistance to <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong><br />

At the San José meeting, it was stated that “little is known <strong>of</strong> the genetics and<br />

inheritance <strong>of</strong> resistance to Sigatoka <strong>diseases</strong>”. Several strategies and programmes<br />

for genetic improvement were presented but these had been developed mainly<br />

for the control <strong>of</strong> Fusarium wilt.<br />

15


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

A great deal <strong>of</strong> effort has been made over the past decade by several<br />

institutions that started programmes to breed for resistance to BLSD in dessert<br />

and cooking <strong>bananas</strong>. The priority <strong>of</strong> the breeding programmes has been to search<br />

for high levels <strong>of</strong> partial resistance, which is considered to be more durable in<br />

the presence <strong>of</strong> diverse and evolving populations <strong>of</strong> a pathogen. In some breeding<br />

programmes, molecular marker–assisted selection was developed to introduce<br />

resistance. Several partially resistant hybrids were produced and tested in multisite<br />

setups such as INIBAP’s International Musa testing programme. Some hybrids<br />

survived the validation stage and were distributed more widely.<br />

Other breeding approaches were developed using biotechnology, and in<br />

particular the production <strong>of</strong> transgenic plants using genes coding for antifungal<br />

proteins (AFPs) (see session 4: R. Swennen). However, this approach has the<br />

disadvantage <strong>of</strong> conferring monogenic resistance and is considered unstable in the<br />

presence <strong>of</strong> diverse populations <strong>of</strong> <strong>Mycosphaerella</strong> species. Nevertheless, the strategy<br />

deserves further study. For example, introducing specific resistance genes in<br />

<strong>bananas</strong> that possess a high level <strong>of</strong> partial resistance might be an attractive<br />

approach.<br />

Control strategies<br />

In San José, there was much discussion about the potential <strong>of</strong> new fungicides, rational<br />

ways <strong>of</strong> using them, the advantages <strong>of</strong> forecasting systems, and the management<br />

<strong>of</strong> resistance to fungicides. It was emphasized that effective and rational control<br />

required a greater knowledge <strong>of</strong> different aspects <strong>of</strong> the epidemiology <strong>of</strong> the<br />

pathogens.<br />

It is generally agreed that this theme is probably the one which received the least<br />

attention and hence could still produce important results. The use <strong>of</strong> fungicides<br />

remains the strategy by which other strategies are compared. In the past, the selection<br />

pressure by different active ingredients has given rise to the disastrous situation where<br />

fungicide-resistant pathotypes are continuously selected. This strategy is no longer<br />

acceptable in a society increasingly concerned about the environment (Romero, 2000).<br />

In future, the aim will be to propose alternatives to chemical control but without<br />

excluding them completely. Integrated control strategies are needed that combine<br />

several different methods <strong>of</strong> control, methods which individually are only partially<br />

effective. Integrated control strategies would have to be adapted to the different<br />

farming systems <strong>of</strong> banana production on a large scale and production <strong>of</strong> plantains<br />

and cooking <strong>bananas</strong> on smallholdings where rational chemical control is difficult.<br />

1) Chemical control can still be considered, provided that its use is strictly limited.<br />

Forecasting methods should be improved or adapted to different environmental<br />

conditions.<br />

2) Control measures based on cultural practices are known to affect inoculum<br />

pressure in the field e.g. <strong>leaf</strong> removal, methods <strong>of</strong> irrigation, management <strong>of</strong><br />

planting density. There is potential for the information to be used more<br />

effectively.<br />

16


Introduction to workshop<br />

X. Mourichon<br />

3) The use <strong>of</strong> genetic resistance is important for the future. With the production<br />

<strong>of</strong> resistant varieties as an objective, it is necessary to consider strategies to<br />

manage resistance, in order to maximize the durability <strong>of</strong> resistance.<br />

The key link between these three aspects is the need to obtain more information<br />

about the epidemiology <strong>of</strong> these pathogens, to make better use <strong>of</strong> what has been<br />

achieved in the past and to propose new lines <strong>of</strong> research into qualitative and<br />

quantitative aspects <strong>of</strong> epidemiology. A particular effort in modelling is expected<br />

to take into account the dynamics <strong>of</strong> pathogen populations.<br />

References<br />

Abadie C., A. El Hadrami and J. Carlier. 2001a. Banana partial resistance against<br />

<strong>Mycosphaerella</strong> fijiensis: studies <strong>of</strong> efficiency and durability. P. 43 in Symposium on<br />

Durable Disease Resistance, Wageningen, The Netherlands, November 2001.<br />

Abadie C., A. El Hadrami, G. Rivas, M.F. Zapater and J. Carlier. 2001b. Studies <strong>of</strong><br />

<strong>Mycosphaerella</strong> fijiensis population structure and partial resistance <strong>of</strong> <strong>bananas</strong>. INFOMUSA<br />

10(1): XIV-XV.<br />

Beveraggi A., X. Mourichon and G. Salle. 1995. Etude des interactions hôte-parasite chez<br />

des bananiers sensibles et résistants inoculés par Cercospora fijiensis (<strong>Mycosphaerella</strong><br />

fijiensis) responsable de la maladie des raies noires. Canadian Journal <strong>of</strong> Botany 73:1328-<br />

1337.<br />

Carlier J., X. Mourichon, D. Gonzales de León, M.F. Zapater and M.H. Lebrun. 1994. DNA<br />

restriction fragment length polymorphisms in <strong>Mycosphaerella</strong> species causing banana <strong>leaf</strong><br />

<strong>spot</strong> <strong>diseases</strong>. Phytopathology 84:751-756.<br />

Carlier J., M.H. Lebrun, M.F. Zapater, C. Dubois and X. Mourichon. 1996. Genetic structure<br />

<strong>of</strong> the global population <strong>of</strong> banana black <strong>leaf</strong> streak fungus <strong>Mycosphaerella</strong> fijiensis.<br />

Molecular Ecology 5:499-510.<br />

Carlier J., X. Mourichon and D.R. Jones. 2000. Black <strong>leaf</strong> streak. The causal agent.<br />

Pp. 46-56 in Diseases <strong>of</strong> Banana, Abacá and Enset. (D.R. Jones, ed.). CABI Publishing,<br />

Wallingford, UK.<br />

Carlier J., M.F. Zapater, F. Lapeyre, D.R, Jones and X. Mourichon. 2000. Septoria <strong>leaf</strong> <strong>spot</strong><br />

<strong>of</strong> banana: a newly discovered disease caused by <strong>Mycosphaerella</strong> eumusae (anamorph<br />

Septoria eumusae). Phytopathology 90:884-890.<br />

Crous P.W and X. Mourichon. 2002. <strong>Mycosphaerella</strong> eumusae and its anamorph<br />

Pseudocercospora eumusae spp. nov., causal agent <strong>of</strong> Eumusae Leaf Spot Disease <strong>of</strong> Banana.<br />

Sydowia 54:35-43.<br />

Etienne J.L., A. Binder, H. Steiner and J.F. Rodriguez. 1995. Detection <strong>of</strong> black Sigatoka disease<br />

in banana leaves using Elisa immuno diagnostics. Pp. 213-218 in Proceedings <strong>of</strong> the<br />

XI Acorbat meeting. (V. Morales Soto, ed.), CORBANA, San Jose, Costa Rica.<br />

Fouré E. and A. Moreau. 1992. Contribution à l’étude épidémiologique de la cercosporiose<br />

noire dans la zone du Mungo au Cameroun. Fruits 47:3-16.<br />

Fouré E., X. Mourichon and D.R. Jones. 2000. Black <strong>leaf</strong> streak. Host reaction. Evaluating<br />

germplasm for reaction to black <strong>leaf</strong> streak. Pp. 62-67 in Diseases <strong>of</strong> Banana, Abacá and<br />

Enset. (D.R. Jones, ed.). CABI Publishing, Wallingford, UK.<br />

Gauhl F. 1994. Epidemiology and Ecology <strong>of</strong> Black Sigatoka (<strong>Mycosphaerella</strong> fijiensis Morelet)<br />

on Plantain and Banana in Costa Rica, Central America. Translation <strong>of</strong> a PhD thesis<br />

originally in German. INIBAP, Montpellier, France, 120pp.<br />

17


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Harelimana G., P. Lepoivre, H. Jijakli and X. Mourichon. 1997. Use <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

toxins for the selection <strong>of</strong> banana cultivars resistant to black <strong>leaf</strong> streak. Euphytica<br />

96(1):125-128.<br />

Hayden H.L., J. Carlier and E.A.B. Aitken. 2000. The population genetics <strong>of</strong> <strong>Mycosphaerella</strong><br />

musicola in Australia. 2 nd International Symposium on Molecular and Cellular Biology<br />

<strong>of</strong> Banana. Brisbane, Australia, 20 October-3 November 2000.<br />

Hayden H.L., J. Carlier and E.A.B. Aitken. In press. The genetic structure <strong>of</strong> <strong>Mycosphaerella</strong><br />

fijiensis from Australia, Papua New Guinea and the Pacific Islands. Plant Pathology.<br />

Johanson A. and M.J. Jeger. 1993. Use <strong>of</strong> PCR for detection <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis and<br />

M. musicola, the causal agents <strong>of</strong> Sigatoka <strong>leaf</strong> <strong>spot</strong>s in banana and Plantains. Mycological<br />

Research 97:670-674.<br />

Johanson A., R.N. Crowhurst, E.H.A. Rikkerink, R.A. Fullerton and M.D. Templeton. 1994.<br />

The use <strong>of</strong> species-specific DNA probes for the identification <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

and M. musicola, the agents <strong>of</strong> Sigatoka <strong>diseases</strong> <strong>of</strong> banana. Plant Pathology 44:701-707.<br />

Jones D.R. 2000. Diseases <strong>of</strong> Banana, Abacá and Enset. (D.R. Jones ed.). CABI Publishing,<br />

Wallingford, UK, 544pp.<br />

Mourichon X., P. Lepoivre and J. Carlier. 2000. Black <strong>leaf</strong> streak. Host-pathogen interactions.<br />

Pp. 67-72 in Diseases <strong>of</strong> Banana, Abacá and Enset. (D.R. Jones, ed.). CABI Publishing,<br />

Wallingford, UK.<br />

Müller R., C. Pasberg-Gauhl, F. Gauhl, J. Ramser and G. Kahl. 1997. Oligonucleotide<br />

fingerprinting detects genetic variability at different levels in Nigerian <strong>Mycosphaerella</strong><br />

fijiensis. Journal <strong>of</strong> Phytopathology 145:25-30.<br />

Romero R.A. 2000. Black <strong>leaf</strong> streak. Control. Pp. 72-79 in Diseases <strong>of</strong> banana, Abacá and<br />

Enset. (D.R. Jones, ed.). CABI Publishing, Wallingford, UK.<br />

Rutter J., P.J.A. Burt and F. Ramirez. 1998. Movement <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis spores and<br />

Sigatoka disease development on plantain close to an inoculum source. Aerobiologia<br />

14:201-208.<br />

Smith M.C., J. Rutter, P.J.A. Burt, F. Ramirez and O.E.H. Gonzales. 1997. Black Sigatoka disease<br />

<strong>of</strong> banana: spatial and temporal variability in disease development. Annals <strong>of</strong> applied<br />

Biology 131:63-77.<br />

Stierle A.A., R. Upadhyay, J. Hershenhorn, G.A. Strobel and G. Molina. 1991. The phytotoxins<br />

<strong>of</strong> <strong>Mycosphaerella</strong> fijiensis, the causative agent <strong>of</strong> black Sigatoka disease <strong>of</strong> <strong>bananas</strong> and<br />

plantains. Experientia 47:853-859.<br />

18


Session 1<br />

Impact <strong>of</strong> <strong>Mycosphaerella</strong><br />

<strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>


Session 1<br />

R.A. Romero<br />

Introduction<br />

The spread, detection and impact <strong>of</strong><br />

black <strong>leaf</strong> streak disease and other<br />

<strong>Mycosphaerella</strong> species in the 1990s<br />

R. A. Romero<br />

Abstract<br />

By the end <strong>of</strong> the 1980s, black <strong>leaf</strong> streak disease caused by <strong>Mycosphaerella</strong> fijiensis was present<br />

in all continents where <strong>bananas</strong> or plantains were grown, although distribution in some regions<br />

was limited to a few countries. In this presentation, the spread <strong>of</strong> the disease during the 1990s<br />

in several countries and regions, and the important socio-economic consequences are discussed.<br />

From 1990 to 1999, new records <strong>of</strong> black <strong>leaf</strong> streak disease were reported from six countries in<br />

Africa, eight in Asia, eight in Latin America and the Caribbean, and one from Australasia/Oceania.<br />

M. fijiensis has also spread within countries to ecological niches that were previously occupied<br />

by M. musicola, the causal agent <strong>of</strong> Sigatoka disease, thus threatening the survival <strong>of</strong> this<br />

pathogen. This presentation also discusses the methods, developed in the last decade, to<br />

identify the species <strong>of</strong> <strong>Mycosphaerella</strong> that cause <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in banana.The methods were<br />

used to confirm the synonymy <strong>of</strong> M. fijiensis var. difformis and M. fijiensis, and have provided the<br />

opportunity to study the genetic diversity <strong>of</strong> pathogen populations among isolates from<br />

different geographical regions.<br />

Resumen - Propagación, detección e impacto de la Sigatoka negra y otras enfermedades<br />

foliares causadas por <strong>Mycosphaerella</strong> de bananos en la década de los 90<br />

Al finales de la década de los 80, la Sigatoka negra, causada por <strong>Mycosphaerella</strong> fijiensis, ya se<br />

encontraba presente en todos los continentes donde se cultivan bananos y plátanos, pero su<br />

distribución en algunas regiones estaba limitada a unos pocos países. En esta presentación, se<br />

discutirá brevemente la propagación de la enfermedad durante los 90 abarcando varios países<br />

y regiones, y su impacto socioeconómico. De 1990 a 1999, se reportaron nuevos registros de la<br />

enfermedad desde seis países en Africa, ocho en Asia, ocho en América Latina y el Caribe, y solo<br />

un informe de un país en Australasia/Oceania. M. fijiensis también ha estado progresando en<br />

los países para llegar a los nichos ecológicos que anteriormente estaban ocupados solo por<br />

Chiquita Brands, San José, Costa Rica<br />

21


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

M. musicola, agente causal de la Sigatoka amarilla, amenazando la supervivencia de este<br />

patógeno. La presentación también discutirá el desarrollo de los métodos más precisos para<br />

identificar las diferentes especies de <strong>Mycosphaerella</strong> que causan las enfermedades de las<br />

manchas foliares en banano en la última década, que han mejorado nuestra habilidad de detectar<br />

el patógeno. Los mismos métodos permitieron confirmar la sinonimia entre M. fijiensis var. difformis<br />

y M. fijiensis, así como el estudio de la diversidad genética de las poblaciones del patógeno entre<br />

los aislados de diferentes regiones geográficas.<br />

Résumé - La propagation, la détection et l’impact de la maladie des raies noires et<br />

d’autres espèces de <strong>Mycosphaerella</strong> dans les années 1990<br />

A la fin des années 1980, la maladie des raies noires, causée par <strong>Mycosphaerella</strong> fijiensis, était<br />

présente sur tous les continents où les bananes et les plantains étaient cultivés, bien que dans<br />

certaines régions sa distribution était limitée à quelques pays seulement. Dans cet exposé, nous<br />

présenterons la propagation de la maladie dans plusieurs pays et régions dans les années 1990<br />

ainsi que les prinicipales répercutions socio-économiques. Entre 1990 et 1999, de nouveaux cas<br />

de maladie des raies noires ont été enregistrés dans six pays d’Afrique, huit d’Asie, huit d’Amérique<br />

latine et des Caraïbes et un d’Australasie/Océanie. Dans certains pays, M. fijiensis, s’est également<br />

répandu dans les niches écologiques occupées au préalable par M. musicola,l’agent provoquant<br />

la maladie de Sigatoka, mettant ainsi en danger la survie de ce pathogène. Nous présenterons<br />

également les méthodes développées dans la dernière décennie qui permettent d’identifier les<br />

espèces responsables des maladies foliaires causées par les <strong>Mycosphaerella</strong> chez la banane. Ces<br />

méthodes ont été utilisées afin de confirmer la synonymie entre M. fijiensis var. difformis et M.<br />

fijiensis, et ont permis l’étude de la diversité génétique des populations de pathogènes parmi des<br />

isolats de différentes régions géographiques.<br />

Introduction<br />

This paper describes the most important events that have characterized the situation<br />

<strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in the decade 1990-2000. Emphasis is given to<br />

black <strong>leaf</strong> streak disease because <strong>of</strong> its greater importance in comparison with other<br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>. Black <strong>leaf</strong> streak disease continued to spread to<br />

new areas and remains a threat to other countries and regions. From the end<br />

<strong>of</strong> the 1980s to 1999, black <strong>leaf</strong> streak disease was newly reported in six countries<br />

in Africa, eight in Asia, eight in Latin America and in one location in<br />

Australasia/Oceania. In America, the French West Indies and the Windward Islands,<br />

two regions that produce <strong>bananas</strong> for export are threatened by black <strong>leaf</strong> streak<br />

disease.<br />

In the near future, it is possible that black <strong>leaf</strong> streak disease may attain a<br />

distribution similar to that <strong>of</strong> Sigatoka disease caused by M. musicola. Details <strong>of</strong><br />

the spread and the distribution <strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> are described<br />

by Pasberg-Gauhl et al. (2000) and as updated by D. Jones elsewhere in this volume.<br />

The continued spread <strong>of</strong> black <strong>leaf</strong> streak disease poses two questions 1) are<br />

there preventative measures that could delay or prevent further spread <strong>of</strong> the disease?<br />

and 2) are those countries that are threatened by the disease taking measures to<br />

prevent the entry <strong>of</strong> the pathogen?<br />

During the last decade there have been considerable advances in the development<br />

<strong>of</strong> precise techniques to detect <strong>Mycosphaerella</strong> spp. pathogens (Carlier et al., 1994;<br />

Johanson and Jeger, 1993a, 1993b). However, there are few examples where these<br />

22


Session 1<br />

R.A. Romero<br />

molecular methods are being used to prevent or delay the spread <strong>of</strong> these pathogens<br />

to new areas. Little is known about the integration <strong>of</strong> these methods with strategies<br />

to prevent the dissemination <strong>of</strong> the <strong>diseases</strong> caused by <strong>Mycosphaerella</strong> spp. A possible<br />

limitation to the use <strong>of</strong> molecular methods is a lack <strong>of</strong> adequate infrastructure and<br />

trained personnel in developing countries.<br />

The arrival <strong>of</strong> black <strong>leaf</strong> streak disease to a new area has resulted in the replacement<br />

<strong>of</strong> Sigatoka disease as the predominant disease <strong>of</strong> <strong>bananas</strong> and plantains. The<br />

ability <strong>of</strong> M. fijiensis to displace M. musicola has not been adequately studied.<br />

However, it is known that M. fijiensis has several biological characteristics that make<br />

it more competitive than M. musicola, e.g. greater ascospore production, more sexual<br />

cycles a year, and a higher rate <strong>of</strong> colonization <strong>of</strong> host tissue (Stover, 1980; Mouliom<br />

Pefoura et al., 1996). Little information is available about differences in pathogenicity<br />

between the two species.<br />

The replacement <strong>of</strong> M. musicola by M. fijiensis, i.e. the elimination <strong>of</strong> one species<br />

by another in a short time, is interesting from an ecological and evolutionary point<br />

<strong>of</strong> view. Unfortunately, there are no studies that show whether M. musicola has in<br />

fact disappeared. It is possible that M. musicola coexists with M. fijiensis at low<br />

frequencies that are difficult to detect. If both pathogens survive in mixed<br />

populations, with M. musicola at a low frequency and M. fijiensis as the predominant<br />

species, it would be preferable to refer to a complex <strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> rather than to black <strong>leaf</strong> streak disease. Breeding strategies for resistance<br />

would need to take this possibility into consideration because resistance to<br />

M. fijiensis does not necessarily imply resistance to M. musicola. Similarly,<br />

resistance genes incorporated by genetic engineering would need to be evaluated<br />

against both pathogens. Current molecular techniques may be able to clarify whether<br />

M. musicola survives at low frequencies with M. fijiensis.<br />

Reports describe how M. musicola is better adapted than M. fijiensis to low<br />

temperatures, and how M. musicola is prevalent at higher altitudes where<br />

temperatures are cooler (Mouliom Pefoura et al., 1996; Romero and Gauhl, 1988;<br />

Tapia, 1993). However, over the years, M. fijiensis has been replacing M. musicola<br />

at higher altitudes in Costa Rica (Gauhl et al., 2000) suggesting that M. fijiensis may<br />

be adapting to the environment. Plantains and other cooking <strong>bananas</strong> commonly<br />

grown at higher altitudes, sometimes in combination with c<strong>of</strong>fee, can be severely<br />

affected due to the greater aggressiveness <strong>of</strong> black <strong>leaf</strong> streak disease in comparison<br />

with Sigatoka disease.<br />

The socio-economic impact <strong>of</strong> black <strong>leaf</strong> streak disease has continued to increase<br />

as the pathogen reaches new areas. The impact has also increased as the disease<br />

becomes more difficult to control because <strong>of</strong> increased resistance <strong>of</strong> the pathogen<br />

to new fungicides. Thus, fungicide resistance is increasingly an important constraint<br />

to control black <strong>leaf</strong> streak disease in plantations dedicated to the export market.<br />

Recently M. eumusae, a new species pathogenic to <strong>bananas</strong>, has been described<br />

(Anon., 1995; Carlier et al., 2000). Little is known about the biology and epidemiology<br />

<strong>of</strong> this species, and hence it is not possible to evaluate whether or not it represents<br />

a threat to small-scale and large–scale banana production. Research is urgently<br />

needed to characterize the pathogenicity and aggressiveness <strong>of</strong> M. eumusa. In<br />

particular, studies are needed on its competitive ability with respect to M. musicola<br />

23


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

and M. fijiensis to determine the potential <strong>of</strong> M. eumusae as a pathogen <strong>of</strong> <strong>bananas</strong><br />

and other Musa genotypes.<br />

During the course <strong>of</strong> this workshop, we will have the opportunity to examine<br />

how these events have taken place or are currently underway. We may be able to<br />

learn from the difficulties and progress on technical assistance and research and hence<br />

increase our ability to improve the control <strong>of</strong> these <strong>diseases</strong> in the future.<br />

References<br />

Anonymous. 1995. Musanews. INFOMUSA 4(2):26-30.<br />

Carlier J., X. Mourichon, D. Gonzales de León, M.F. Zapater and M.H. Lebrun. 1994. DNA<br />

restriction fragment length polymorphisms in <strong>Mycosphaerella</strong> species causing banana <strong>leaf</strong><br />

<strong>spot</strong> <strong>diseases</strong>. Phytopathology 84:751-756.<br />

Carlier J., X. Mourichon and D.R. Jones. 2000. Fungal <strong>diseases</strong> <strong>of</strong> the foliage. Sigatoka like<br />

<strong>leaf</strong> <strong>spot</strong>s, Septoria <strong>leaf</strong> <strong>spot</strong>. Pp. 93-96 in Diseases <strong>of</strong> Bananas, Abacá and Ensete (D.R.<br />

Jones, ed.). CAB International, Wallingford, UK.<br />

Gauhl F., C. Pasberg-Gauhl and D.R. Jones. 2000. Fungal <strong>diseases</strong> <strong>of</strong> the foliage. Sigatoka<br />

<strong>leaf</strong> <strong>spot</strong>s. Black <strong>leaf</strong> streak, disease cycle and epidemiology. Pp. 56-62 in Diseases <strong>of</strong><br />

Bananas, Abacá and Ensete (D.R. Jones, ed.). CAB International, Wallingford, UK.<br />

Johanson A. and M.J. Jeger. 1993a. Use <strong>of</strong> PCR for detection <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis and<br />

M. musicola, the causal agents <strong>of</strong> Sigatoka <strong>leaf</strong> <strong>spot</strong>s in banana and plantain. Mycological<br />

Research 97:670-674.<br />

Johanson A. and M.J. Jeger. 1993b. Detection <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis and M. musicola<br />

in banana <strong>leaf</strong> tissue using the polymerase chain reaction. Pp. 227-236 in Breeding Banana<br />

and Plantain for Resistance to Diseases and Pests. Proceedings <strong>of</strong> an International<br />

symposium on Genetic improvement <strong>of</strong> <strong>bananas</strong> for resistance to <strong>diseases</strong> and pests. CIRAD<br />

and INIBAP, France.<br />

Mouliom Pefoura A., A. Lassoudière, J. Foko and D.A Fontem. 1996. Comparison <strong>of</strong><br />

development <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis and <strong>Mycosphaerella</strong> musicola on banana and<br />

plantain in the various ecological zones in Cameroon. Plant Disease 80:950-953.<br />

Pasberg-Gauhl C., F. Gauhl and D.R. Jones. 2000. Fungal <strong>diseases</strong> <strong>of</strong> the foliage. Sigatoka<br />

<strong>leaf</strong> <strong>spot</strong>s. Black <strong>leaf</strong> streak, distribution and economic importance. Pp. 37-44 in Diseases<br />

<strong>of</strong> Bananas, Abacá and Ensete (D.R. Jones, ed.). CAB International, Wallingford, UK.<br />

Romero R.A. and F. Gauhl. 1988. Determinación de la severidad de la Sigatoka negra<br />

(<strong>Mycosphaerella</strong> fijiensis var. difformis) en bananos a diferentes altitudes sobre el nivel<br />

del mar. Revista de la Asociación Bananera Nacional (ASBANA), San José, Costa Rica<br />

12(29):7-10.<br />

Stover R. H. 1980. Sigatoka <strong>leaf</strong> <strong>spot</strong>s <strong>of</strong> <strong>bananas</strong> and plantains. Plant Disease 64:750-755.<br />

Tapia A. 1993. Distribución altitudinal de la Sigatoka amarilla (<strong>Mycosphaerella</strong> musicola) y<br />

la Sigatoka negra (<strong>Mycosphaerella</strong> fijiensis) en Costa Rica. Tesis. Universidad de Costa<br />

Rica, 76pp.<br />

24


Session 1<br />

D.R. Jones<br />

The distribution and importance <strong>of</strong><br />

the <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

<strong>of</strong> banana<br />

D. R. Jones<br />

Abstract<br />

The three main fungal <strong>leaf</strong> <strong>spot</strong> pathogens <strong>of</strong> banana are <strong>Mycosphaerella</strong> musicola, M. fijiensis and<br />

M. eumusae.All result in serious economic damage to cultivars in the Cavendish subgroup grown<br />

for local consumption and export. M. musicola was the first foliar pathogen to be identified as a<br />

major problem. Its spread from the Southeast Asian/Pacific region, where it was established by the<br />

1920s, to North and South America in the 1930s, resulted in widespread disruption to the export<br />

trade. Since the 1960s, M. musicola has been steadily and largely replaced worldwide by M. fijiensis.<br />

This pathogen is more damaging to Cavendish cultivars than M. musicola and attacks a wider range<br />

<strong>of</strong> banana clones. It was first recognised in the Pacific region and has since become a major problem<br />

in most banana-growing areas. The susceptibility <strong>of</strong> plantain and other subsistence banana types<br />

is <strong>of</strong> special concern in Africa. By the late 1980s, reports in the scientific literature <strong>of</strong> records <strong>of</strong><br />

M. fijiensis and M. musicola in Asia led to some confusion as to the true distribution <strong>of</strong> the two<br />

pathogens in this region. Many specimens collected in the region were identified as M. eumusae,<br />

a new <strong>leaf</strong> <strong>spot</strong> pathogen <strong>of</strong> banana.This fungus was observed damaging Cavendish cultivars and<br />

plantain. Most records for M. eumusae have been from Asia but the pathogen has also been found<br />

on islands in the Indian Ocean and in West Africa. Because <strong>of</strong> the similarity <strong>of</strong> symptoms caused<br />

by M. musicola, M. fijiensis and M. eumusae,itis proposed that the <strong>diseases</strong> caused by these three<br />

fungi be known collectively as <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>.<br />

Resumen - Distribución e importancia de las enfermedades de las manchas foliares<br />

causadas por <strong>Mycosphaerella</strong> en banano<br />

Los tres principales patógenos fungosos de las manchas foliares del banano son <strong>Mycosphaerella</strong><br />

musicola, M. fijiensis y M. eumusae.Todos ellos producen serios daños económicos a los cultivares<br />

en el subgrupo Cavendish cultivados para el consumo local y para la exportación. M. musicola fue<br />

el primer patógeno foliar identificado como el principal problema en la región productora de bananos<br />

de América Latina y el Caribe. Se propagó de la región de Sudeste asiático y el Pacífico, donde se<br />

había establecido allá por la década de los 20, al Nuevo Mundo en la década de los 30, resultando<br />

en una quiebra del comercio de exportación. Desde los años 60, M. musicola fue reemplazada firme<br />

y extensamente en todo el mundo por M. fijiensis. Este patógeno es más dañino para los cultivares<br />

Consultant in International Agriculture,Worcestershire, United Kingdom<br />

25


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Cavendish que M. musicola y ataca un rango más amplio de cultivares. Por primera vez esta<br />

enfermedad fue reconocida en el Pacífico. Actualmente, M. fijiensis es uno de los principales<br />

problemas en la mayoría de las áreas productoras de banano y ha reemplazado a M. musicola como<br />

patógeno dominante de la mancha foliar. La susceptibilidad del plátano y otros tipos de bananos<br />

de subsistencia representa la principal preocupación en Africa. A finales de la década de los 80, los<br />

informes encontrados en la literatura científica de los registros de M. fijiensis y M. musicola en el<br />

Sudeste de Asia llevaron a algunas confusiones en cuanto a la verdadera distribución de los dos<br />

patógenos en esta región. Muchos especímenes recolectados en la región fueron identificados como<br />

M. eumusae,un nuevo patógeno de la mancha foliar del banano. Este hongo se observó como dañino<br />

para los cultivares Cavendish y plátano. Aunque hasta la fecha la mayoría de los registros sobre la<br />

M. eumusae provienen de las regiones de Sudeste y Sur de Asia, el patógeno también fue encontrado<br />

en las islas del Océano Indico y en Africa Occidental. Debido a la similitud de los síntomas causados<br />

por M. musicola, M. fijiensis y M. eumusae, se sugiere que las enfermedades causadas por estos tres<br />

hongos se denominen manchas foliares causadas por <strong>Mycosphaerella</strong>.<br />

Résumé - La distribution et l’importance des maladies foliaires causées par les<br />

<strong>Mycosphaerella</strong><br />

Les trois principaux pathogènes responsables des maladies foliaires de la banane sont<br />

<strong>Mycosphaerella</strong> musicola, M. fijiensis et M. eumusae. Tous trois provoquent de sérieux dégâts<br />

économiques aux cultivars du sous-groupe Cavendish, cultivés pour la consommation locale et<br />

l’exportation. M. musicola a été le premier pathogène de <strong>Mycosphaerella</strong> a être identifié<br />

provoquant des problèmes majeurs. Sa propagation de l’Asie du Sud-Est/région du Pacifique, où<br />

il était établi dans les années 1920, à l’Amérique du Nord et du Sud dans les années 1930, a<br />

sérieusement affecté le commerce international. Depuis les années 1960, M. musicola a été<br />

graduellement et largement remplacé à travers le monde par M. fijiensis. Ce pathogène cause<br />

plus de dommages aux Cavendish que M. musicola et attaque un spectre plus large de cultivars<br />

de bananiers. Il a d’abord été identifié dans la région du Pacifique avant de devenir un problème<br />

majeur dans la plupart des zones cultivées de bananes. La sensibilité du plantain et d’autres types<br />

de bananes servant d’aliment de base est particulièrement préoccupante en ce qui a trait à<br />

l’Afrique. A la fin des années 1980, les études rapportant la présence de M. fijiensis et de M. musicola<br />

en Asie, ont prêté à confusion quant à la véritable distribution des deux pathogènes dans cette<br />

région. De nombreux échantillons récoltés dans la région ont été identifiés comme étant<br />

M. eumusae, un nouveau pathogène responsable d’une maladie foliaire des bananiers. Il a été<br />

observé que ce champignon attaquait les Cavendish et le plantain. C’est en Asie que la majorité<br />

des observations de M. eumusae ont été faites mais ce pathogène a également été trouvé dans<br />

des îles de l’Océan Indien et en Afrique de l’Ouest. Du fait des symptômes similaires causés par<br />

M. musicola, M. fijiensis et M. eumusae, nous proposons que les pathologies causées par ces trois<br />

champignons soient connues collectivement en tant que maladies foliaires causées par les<br />

<strong>Mycosphaerella</strong>.<br />

Introduction<br />

The most serious <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> banana are caused by three species <strong>of</strong><br />

<strong>Mycosphaerella</strong>. All three were discovered and became important constraints to<br />

commercial production in the 20 th century. One has long since reached the limits<br />

<strong>of</strong> its distribution; a second is now approaching the limits <strong>of</strong> its distribution and<br />

a third is possibly only now beginning its global spread. The evidence suggests<br />

that all three may have arisen in the Southeast Asian/Australasian region, which<br />

is the centre <strong>of</strong> origin <strong>of</strong> Musa species and also the centre <strong>of</strong> evolution <strong>of</strong> cultivated<br />

banana (Jones, 2000). Some accessions <strong>of</strong> M. acuminata ssp. banksii, which is a<br />

wild diploid banana that has contributed genetic components to most edible banana<br />

26


Session 1<br />

D.R. Jones<br />

clones (Carreel, 1995), are known to be susceptible to at least two <strong>of</strong> the<br />

<strong>Mycosphaerella</strong> species causing <strong>leaf</strong> <strong>spot</strong> (Carlier et al., 2000a) and so coevolution<br />

is a strong possibility.<br />

It is not clear why the <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens were not disseminated<br />

during the first movements <strong>of</strong> banana planting material out <strong>of</strong> the Southeast<br />

Asian/Australasian region, as they could have been carried as infections <strong>of</strong> scale<br />

<strong>leaf</strong> tissue associated with sword suckers (Stover, 1978). Possibly, the fungi were<br />

not widespread pathogens on subspecies <strong>of</strong> M. acuminata or cultivated banana.<br />

Alternatively, the fungi, which initially were saprophytes surviving by colonizing<br />

dead <strong>leaf</strong> tissue, may have evolved to become pathogenic quite recently. Stover<br />

(1978) speculated that M. fijiensis might have evolved on a susceptible wild diploid<br />

and then spread to edible cultivars. The histories <strong>of</strong> the three <strong>Mycosphaerella</strong> fungi<br />

that continue to cause so much damage to cultivated banana are contrasted and<br />

compared in chronological order <strong>of</strong> discovery in the following text.<br />

<strong>Mycosphaerella</strong> musicola, the cause <strong>of</strong> Sigatoka disease<br />

First records<br />

<strong>Mycosphaerella</strong> musicola was the first <strong>leaf</strong> pathogen to become a serious problem<br />

on commercial banana plantations. The fungus, originally named Cercospora musae<br />

from its imperfect stage, was first described as a pathogen <strong>of</strong> banana on the island<br />

<strong>of</strong> Java in Indonesia at the beginning <strong>of</strong> the 20 th century (Zimmermann, 1902).<br />

However, it wasn’t until 10 years later in Fiji that it became prominent as the cause<br />

<strong>of</strong> an important disease. The pathogen was first found in Fiji in the Sigatoka<br />

(pronounced ‘Singatoka’) Valley on the main island Viti Levu and quickly became<br />

an important constraint to production (Philpott and Knowles, 1913; Massee, 1914).<br />

The disease became known as Sigatoka disease, and more recently as yellow Sigatoka,<br />

to distinguish it from the disease caused by M. fijiensis, which is widely known as<br />

black <strong>leaf</strong> streak disease or black Sigatoka.<br />

Global spread<br />

The subsequent spread <strong>of</strong> M. musicola, as it was called after the perfect stage was<br />

discovered (Leach, 1941), around the world to practically all banana-growing regions<br />

is well documented (Meredith, 1970) (Table 1). The chronological sequence <strong>of</strong> disease<br />

occurrences led to speculation that ascopores borne on high altitude wind currents<br />

may have been responsible for the intercontinental dissemination <strong>of</strong> the pathogen<br />

from Australia to Africa and Central America (Stover, 1962). However, the latest<br />

information on survivability and movement <strong>of</strong> windblown ascospores, which are the<br />

spores implicated in long distance spread, indicate that this hypothesis is unlikely<br />

(Parnell et al., 1998). Stover (1980) changed his opinion on the distances ascospores<br />

<strong>of</strong> <strong>Mycosphaerella</strong> pathogens could spread by stating that wind dispersal <strong>of</strong><br />

M. fijiensis from small areas <strong>of</strong> infection to new areas was probably slight when<br />

distances exceeded 50 km. Natural ultraviolet radiation, which would be a factor<br />

limiting high-altitude dispersal on clear days, kills ascospores <strong>of</strong> M. fijiensis within<br />

27


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Table 1. Year Sigatoka disease was first reported in a given country until 1970 (modified from information published<br />

in Table 2 <strong>of</strong> Meredith,1970).<br />

Region/Country Year Region/Country Year<br />

Australasia/Oceania<br />

Fiji 1912<br />

Australia (Queensland) 1924<br />

Australia (New South Wales) 1927<br />

Solomon Islands 1946<br />

Papua New Guinea 1951<br />

New Caledonia 1951<br />

Indonesia (Irian Jaya) 1953<br />

Norfolk Island 1954<br />

Wallis Island 1954<br />

USA (Hawaii) 1958<br />

Samoa 1961<br />

American Samoa 1961<br />

French Polynesia 1962<br />

Tonga 1965<br />

Asia<br />

Indonesia (Java) 1902<br />

Sri Lanka 1919<br />

Philippines 1921<br />

Malaysia (Peninsula) 1933<br />

China 1936<br />

India (Assam) 1946<br />

Taiwan 1953<br />

Malaysia (Sabah and Sarawak) 1959<br />

Cambodia 1960<br />

Thailand 1962<br />

Hong Kong 1966<br />

Vietnam 1966<br />

Laos 1967<br />

Brunei 1968<br />

Latin America/Caribbean<br />

Guadeloupe 1932 ?<br />

Surinam 1933<br />

Trinidad 1933<br />

Guyana 1935<br />

Honduras 1935<br />

Jamaica


Session 1<br />

D.R. Jones<br />

Movement between continents and between isolated countries, such as those found<br />

in the Pacific, is probably the result <strong>of</strong> the transfer <strong>of</strong> diseased material by humans.<br />

The appearance <strong>of</strong> the disease in Australia in the mid-1920s may have been due<br />

to the same movements <strong>of</strong> planting material from the South Pacific that are believed<br />

to have introduced bunchy top disease (Magee, 1953). Another possibility is that<br />

diseased banana leaves may have been used as packing material for goods shipped<br />

into Australia from the South Pacific.<br />

Sigatoka disease may have been introduced to East Africa in the late 1930s by<br />

the movement <strong>of</strong> planting material/diseased leaves from Asia. During the colonial<br />

period, there was much migration to East Africa from India and settlers and/or British<br />

administrators undoubtedly transported infected suckers <strong>of</strong> their favourite cultivars<br />

from one location to another.<br />

The appearance <strong>of</strong> Sigatoka disease in European colonies in the eastern Caribbean<br />

region in the mid-1930s could also have been due to colonial-inspired movements<br />

<strong>of</strong> Musa germplasm from Asia. The almost simultaneous discovery <strong>of</strong> the disease at<br />

several locations may have been due to multiple introductions <strong>of</strong> the pathogen or<br />

to rapid spread from one point after an undetected build-up <strong>of</strong> inoculum. Prevailing<br />

landward winds <strong>of</strong>f the Caribbean Sea would have carried spores <strong>of</strong> M. musicola<br />

across the north <strong>of</strong> South America and into Central America.<br />

Today, Sigatoka disease is regarded as having a worldwide distribution, although<br />

it has not been recorded in the Canary Islands, Egypt or Israel (Meredith, 1970). The<br />

dry summer climates <strong>of</strong> these countries may make the local environment unsuitable<br />

for disease establishment.<br />

Impact<br />

The most efficient leaves for photosynthesis on a vegetatively growing banana are<br />

the second to fifth counting down from the top <strong>of</strong> the plant. Therefore, if optimal<br />

assimilation potential <strong>of</strong> the plant is to be maintained, it is important that leaves<br />

2-5 are free <strong>of</strong> excessive shade, severe <strong>leaf</strong> tearing and disease. In a vigorous plant<br />

growing in the tropics, this critical <strong>leaf</strong> area is renewed monthly (Robinson, 1996)<br />

and the pathogen does not cause enough damage to have an appreciable effect on<br />

growth (Leach, 1946). Damage comes after bunch emergence when <strong>leaf</strong> production<br />

ceases and <strong>leaf</strong> tissue cannot be renewed. The greater the damage on remaining leaves<br />

and the earlier it occurs after shooting, the greater the effects on yield. Sigatoka<br />

disease also affects the physiology <strong>of</strong> developing fruits causing premature ripening<br />

(Meredith, 1970). This occurs in the field if the plant is severely diseased or in transit<br />

to markets if moderately affected. For these reasons it was important to control<br />

Sigatoka disease in commercial plantations.<br />

The economic impact <strong>of</strong> Sigatoka disease has been tw<strong>of</strong>old. First, there was the<br />

direct effect on production when the disease first became established and control<br />

measures were being developed. Between 1937 and 1941, production in Mexico was<br />

halved as a direct result <strong>of</strong> Sigatoka disease. In Honduras, production declined to<br />

less than one third <strong>of</strong> the pre-disease level (Meredith, 1970). These were enormous<br />

losses. Secondly, there were the on-going costs associated with <strong>leaf</strong> <strong>spot</strong> control once<br />

effective measures were developed and adopted. The struggle against Sigatoka disease<br />

29


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

in this regard has been well documented (Meredith, 1970; Stover 1972, 1990; Carlier<br />

et al., 2000a). Bordeaux mixture, the first effective fungicide, had to be applied as<br />

a high volume spray and large pipeline systems were installed at great expense to<br />

deliver the chemical in plantations. Later, petroleum oil applied as a low-volume<br />

spray from aircraft proved effective and costs were reduced, especially when combined<br />

with forecasting systems. Protectant fungicides, such as dithiocarbomates, and<br />

systemic fungicides, such as benomyl, improved the standard <strong>of</strong> control. However,<br />

control measures, which included labour intensive cultural practices, such as<br />

pruning old diseased leaves, was an expense that was borne by growers. In 1990 in<br />

Queensland, where Sigatoka disease is still the dominant <strong>leaf</strong> <strong>spot</strong> disease, control<br />

measures were estimated as 14% <strong>of</strong> total production costs.<br />

There are no real figures on the impact Sigatoka disease had on small-scale<br />

growers in developing countries producing fruit for families or local markets.<br />

Plantains and other cooking <strong>bananas</strong> favoured by many subsistence farmers are<br />

resistant to Sigatoka disease in lowland areas, and therefore the impact in these cases<br />

would be minimal. Those growing the more susceptible dessert banana cultivars, such<br />

as in the Pome subgroup (AAB), would either have had to change to cooking types,<br />

accepted a loss in yield, or invested in backpack sprayers. However, the worst was<br />

yet to come.<br />

<strong>Mycosphaerella</strong> fijiensis, the cause <strong>of</strong> black <strong>leaf</strong> streak<br />

disease<br />

First records<br />

The first report <strong>of</strong> M. fijiensis causing damage was in the same Sigatoka valley on<br />

the island <strong>of</strong> Vitu Levu in Fiji where M. musicola was first recognised as a major<br />

pathogen <strong>of</strong> banana fifty years earlier. In February 1963, the disease caused by<br />

M. fijiensis was reported to be spreading rapidly in the Sigatoka Valley (Rhodes,<br />

1964) and was predicted to affect the whole island by the end <strong>of</strong> 1964 (Leach, 1964a).<br />

The causal organism was also described for the first time from material collected in<br />

Fiji (Leach, 1964b). The disease caused by M. fijiensis was called black <strong>leaf</strong> streak<br />

disease by Rhodes (1964). Leach (1964b) described the risk <strong>of</strong> spread <strong>of</strong> this new<br />

disease <strong>of</strong> banana as “a grave threat” and feared that the abundance <strong>of</strong> airborne<br />

ascospores produced by the pathogen may lead to dissemination around the world<br />

faster than for Sigatoka disease. He also warned that an outbreak <strong>of</strong> a new, perhaps<br />

worse, disease might happen again in the future in another location.<br />

The problem caused by M. fijiensis in Fiji became apparent when the mist-sprays<br />

<strong>of</strong> light mineral oil being used to control M. musicola lost their effectiveness. The<br />

recognition <strong>of</strong> yet another important banana pathogen in Fiji before anywhere else<br />

can probably be attributed to the fact that at this location there were sizeable<br />

plantations <strong>of</strong> susceptible dessert banana cultivars and an efficient plant protection<br />

service experienced in banana problems.<br />

Surveys undertaken after black <strong>leaf</strong> streak disease was discovered in Fiji led to<br />

the conclusion that the pathogen had most likely been present in the Pacific and<br />

parts <strong>of</strong> the Pacific rim for many years (Meredith, 1970; Stover, 1976; Stover, 1978;<br />

30


Session 1<br />

D.R. Jones<br />

Long, 1979). It was suggested that M. fijiensis may have been in the Hawaiian Islands<br />

in 1958 (Meredith and Lawrence, 1969). An analysis <strong>of</strong> herbarium specimens by<br />

Stover (1976), showed that M. fijiensis was present in Papua New Guinea by at<br />

least 1957 and in Taiwan as early as 1927. The similarity <strong>of</strong> symptoms with those<br />

<strong>of</strong> Sigatoka disease most likely masked the arrival <strong>of</strong> this new disease in many<br />

countries. Because <strong>of</strong> this, the year that black <strong>leaf</strong> streak disease was first<br />

discovered in many countries in this region (Table 2) does not reflect the order <strong>of</strong><br />

spread <strong>of</strong> the pathogen.<br />

Global spread<br />

When the black <strong>leaf</strong> streak pathogen was first found in Honduras in 1972, it was<br />

thought from its morphology to be a variant <strong>of</strong> M. fijiensis and was named<br />

M. fijiensis var. difformis (Mulder and Stover, 1976). However, it was later shown<br />

that M. fijiensis and M. fijiensis var. difformis were synonymous (Pons, 1987). Black<br />

<strong>leaf</strong> streak disease has precedence as the common name for the disease caused by<br />

M. fijiensis and was adopted by Carlier et al. (2000a) in the publication ‘Diseases<br />

<strong>of</strong> Banana, Abacá and Enset’, however it is widely known in as black Sigatoka or<br />

Sigatoka negra in Spanish. The choice <strong>of</strong> which name to use in publications is one<br />

<strong>of</strong> personal preference.<br />

A measure <strong>of</strong> the rate <strong>of</strong> spread <strong>of</strong> M. fijiensis between countries can be<br />

gained from an examination <strong>of</strong> the records from the Latin American/Caribbean<br />

region (Table 2). Within three years <strong>of</strong> its detection in Honduras in 1972, M. fijiensis<br />

was reported in Belize to the north and by 1977 had arrived in Guatemala to the<br />

west. Local spread was quicker in the direction <strong>of</strong> prevailing winds from the east<br />

and northeast (Stover, 1980). In 1979, it appeared in El Salvador, Nicaragua and<br />

Costa Rica and by 1981 had spread north to Mexico and south to Panama and<br />

northern Colombia (Carlier et al., 2000a). Spread was believed to have been<br />

accelerated in Central America by the movement <strong>of</strong> diseased banana leaves and<br />

<strong>leaf</strong> trash across international boundaries with road-transported banana and<br />

plantain fruit (Stover, 1980). By 1986, commercial plantations in northern Ecuador<br />

were affected and plantains in western Venezuela succumbed in 1991. Spread to<br />

northern Peru occurred in 1994 and to Bolivia in 1996. The first report from western<br />

Brazil came in 1996 and since then M. fijiensis has been advancing in a<br />

southwesterly direction towards the Brazilian coast. In 2001, movements <strong>of</strong> banana<br />

fruit and associated banana leaves from inland areas where M. fijiensis was found<br />

to coastal cities were being controlled in an effort to delay spread (R. S. Moreira,<br />

Brazil, personal communication).<br />

In the Caribbean, black <strong>leaf</strong> streak disease was first found in Cuba in 1990,<br />

Jamaica in 1995 and the Dominican Republic in 1996. The first authenticated report<br />

from Haiti, which has a dry climate, was in 1999. Natural spread to other<br />

Caribbean countries is inevitable, but may be slowed considerably by the prevailing<br />

winds, which blow from the east. Recent investigations involving the analysis <strong>of</strong><br />

isolates have revealed that the source <strong>of</strong> inoculum for the outbreak in Jamaica may<br />

have come form Central America and not windblown from Cuba as originally<br />

suspected (G. Rivas, Costa Rica, personal communication). The outbreak in the<br />

31


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Table 2. Countries where black <strong>leaf</strong> streak disease has been detected (modified from Carlier et al., 1999 with information<br />

supplied by X. Mourichon and D. R. Jones).<br />

Region/Country Year 1 Region/Country Year 1<br />

Australasia/Oceania<br />

Solomon Islands 1957 (1946)<br />

Papua New Guinea 1957 (1951)<br />

Fiji 1963<br />

French Polynesia 1964-67<br />

Micronesia 1964-67<br />

New Caledonia 1964-67<br />

Vanuatu 1964-67<br />

Tonga 1965<br />

Samoa 1965<br />

USA (Hawaii) 1969 (1958)<br />

American Samoa 1975<br />

Cook Islands 1976<br />

Niue 1976<br />

Norfolk Island 1980<br />

Australia (Torres Strait/Cape York) 1981<br />

Wallis and Fortuna Islands 1996<br />

Australia (North Queensland,<br />

currently under eradication) 2001<br />

Asia<br />

Taiwan 1927<br />

Philippines (Luzon) 1964<br />

Singapore 1964-67<br />

Philippines (Mindanao) 1965<br />

Malaysia (Peninsula) 1965<br />

Thailand 1969<br />

Indonesia (Java) 1969<br />

Indonesia (Halmahera) 1970<br />

China (Hainan) 1980<br />

Bhutan 1985<br />

China (Guangdong) 1990<br />

China (Yunnan) 1993<br />

Vietnam 1993<br />

Indonesia (Sumatra) 1993<br />

Indonesia (Kalimantan) 1996<br />

East Malaysia (Sarawak) 1996<br />

Latin America/Caribbean<br />

Honduras 1972 (1969)<br />

Belize 1975<br />

Guatemala 1977<br />

Nicaragua 1979<br />

Costa Rica 1979<br />

El Salvador 1979 2<br />

Mexico 1980<br />

Panama 1981<br />

Colombia 1981<br />

Ecuador 1986<br />

Cuba 1990<br />

Venezuela 1991<br />

Peru 1994<br />

Jamaica 1995<br />

Bolivia 3 1996<br />

Dominican Republic 1996<br />

Brazil 1998<br />

USA (Florida) 1998<br />

Haiti 1999<br />

Africa<br />

Zambia 1973 4<br />

Gabon 1978<br />

Cameroon (south-east) 1980<br />

Cameroon (south-west) 1983<br />

São Tomé 1983<br />

Côte d’Ivoire 1985<br />

Congo 1985<br />

Nigeria 1986<br />

Ghana 1986<br />

Rwanda 1986<br />

Burundi 1987<br />

Tanzania (inc. Pemba and Zanzibar) 1987<br />

Democratic Republic <strong>of</strong> Congo<br />

(highlands) 1987<br />

Democratic Republic <strong>of</strong> Congo<br />

(lowlands) 1988<br />

Togo 1988<br />

Kenya 1988<br />

Malawi 1990<br />

Uganda 1990<br />

Benin 1993<br />

Comoros 1993<br />

Mayotte 1993<br />

Central African Republic 1996<br />

Madagascar 2000<br />

1<br />

Most likely year <strong>of</strong> introduction, but no definite record published at<br />

this time.<br />

2<br />

Year disease was first reported or year present from herbarium<br />

specimens (earliest year disease was believed present in hindsight<br />

within brackets). 3 In 2001, black <strong>leaf</strong> streak disease was spreading<br />

south-eastwards across Brazil from the Mato Grosso towards the<br />

Atlantic coast. 4 Authenticity <strong>of</strong> this record has been challenged.<br />

Dominican Republic has also been linked to Central America, though the evidence<br />

is more circumstantial. In both examples, the disease appeared shortly after<br />

banana fruit was shipped to the islands. Could inoculum on the surface <strong>of</strong> fruit or<br />

in associated <strong>leaf</strong> trash have introduced black <strong>leaf</strong> streak to these countries?<br />

32


Session 1<br />

D.R. Jones<br />

The first record <strong>of</strong> black <strong>leaf</strong> streak disease in Africa was from Zambia in 1973<br />

(Raemaekers, 1975). The publication <strong>of</strong> this outbreak is convincing but the identity<br />

<strong>of</strong> the pathogen could not be confirmed from specimens sent to the UK, therefore<br />

doubt remains as to the authenticity <strong>of</strong> the report (Dabek and Waller, 1990). The<br />

next record was from Gabon in 1978. Frossard (1980) believed it might have been<br />

introduced on planting material from Asia. The disease then spread steadily through<br />

Central and West Africa reaching Côte d’Ivoire, Nigeria and Ghana in 1985-1986,<br />

and Uganda and Malawi in 1990 (Table 2). A second, separate introduction <strong>of</strong><br />

M. fijiensis into Africa is thought to have occurred in 1987 on the island <strong>of</strong> Pemba.<br />

This outbreak is believed to have led to the pathogen spreading to the island <strong>of</strong><br />

Zanzibar and coastal areas <strong>of</strong> Tanzania and Kenya (Carlier et al., 2000a). In 2000,<br />

M. fijiensis was recorded in Madagascar for the first time.<br />

The Australian experience<br />

Stover (1978) believed that M. fijiensis may have originated in the Papua New<br />

Guinea-Solomon Islands area and disseminated around the South Pacific with<br />

banana leaves or planting material. This possibility is suggested by the discovery<br />

that isolates <strong>of</strong> M. fijiensis are more diverse in the Papua New Guinea /Philippines<br />

region than elsewhere, an indication that the area may be the centre <strong>of</strong> origin <strong>of</strong><br />

the pathogen (Carlier et al., 2000a). Therefore, it is likely that M. fijiensis may have<br />

been present on banana on islands in the Torres Strait and on the tip <strong>of</strong> Cape York<br />

Peninsula, Australia long before its discovery on the first plant pathological survey<br />

<strong>of</strong> the area in 1981 (Jones and Alcorn, 1982). The pathogen may not have spread<br />

further south in Australia because <strong>of</strong> the barrier presented by the Cape York<br />

Peninsula, which is a large, remote area <strong>of</strong> native bush with comparatively few<br />

communities and banana plants. After 1981, better land and air communications,<br />

which encouraged more tourists and people seeking an alternative lifestyle, led to<br />

a higher risk <strong>of</strong> spread. During the 1990s, M. fijiensis was regularly eradicated from<br />

isolated outbreaks on small plantings within the Peninsula. In all cases, the origin<br />

<strong>of</strong> the inoculum could not be positively determined. In 2000, an outbreak occurred<br />

on a commercial banana planting at Daintree on the northern fringe <strong>of</strong> the more<br />

heavily populated coastal strip centred on Cairns. The grower was compensated for<br />

the destruction <strong>of</strong> his crop by the Australian banana industry. Although eradicated,<br />

the close proximity <strong>of</strong> this outbreak to the main banana growing area was worrying.<br />

Towards the end <strong>of</strong> the wet season in April 2001, M. fijiensis was detected on<br />

unmanaged (feral) banana plants and also on cultivated plants in an adjacent farm<br />

in the Tully Valley, which is in the heart <strong>of</strong> the commercial banana-growing area<br />

in North Queensland centred south <strong>of</strong> Cairns. Subsequently, the pathogen was<br />

reported from other locations in the same area. An eradication campaign was<br />

immediately mounted. This campaign gathered momentum when the governments<br />

<strong>of</strong> banana-growing states and the Commonwealth Government pledged funds.<br />

Measures included: (1) establishment <strong>of</strong> a special banana quarantine area, (2) a ban<br />

on the movement <strong>of</strong> fruit from this area to other banana-growing areas in Australia,<br />

(3) close monitoring <strong>of</strong> crops and the diagnosis <strong>of</strong> any <strong>leaf</strong> <strong>spot</strong>s detected,<br />

(4) destruction <strong>of</strong> fields where affected plants were found, (5) drastic pruning <strong>of</strong> all<br />

33


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

banana plants in the growing area, (6) regular application <strong>of</strong> systemic fungicides<br />

and (7) zero tolerance for <strong>leaf</strong> <strong>spot</strong> disease. The campaign was conducted during<br />

the 2001 dry season, which also markedly reduced the chances <strong>of</strong> spore release and<br />

infection, with the co-operation <strong>of</strong> most growers.<br />

A total <strong>of</strong> 25 plants have been found infected with M. fijiensis in the Tully area.<br />

The last seven plants were either growing in private gardens or were unmanaged.<br />

At the time <strong>of</strong> the International Sigatoka Workshop on 20-23 May 2002, M. fijiensis<br />

had not been detected for over five and a half months. There are hopes for the<br />

successful eradication <strong>of</strong> M. fijiensis, which will be the first time that this will have<br />

been achieved anywhere in the world.<br />

Impact<br />

Black <strong>leaf</strong> streak disease is the major constraint to cultivation in commercial<br />

plantations producing dessert banana fruit for export. It is also a limiting factor<br />

for small-scale and subsistence farmers growing plantain. The disease has had a<br />

much greater impact than Sigatoka disease because the life cycle and epidemiology<br />

<strong>of</strong> the causal pathogen makes it more difficult to control and it attacks a wider<br />

range <strong>of</strong> banana clones. M. fijiensis attacks younger leaves on more susceptible<br />

banana clones than those affected by M. musicola. On dessert clones in the<br />

Cavendish subgroup, which are extremely susceptible, the pathogen can kill much<br />

more <strong>leaf</strong> tissue in the critical 2-5-<strong>leaf</strong> range than M. musicola. After flowering,<br />

remaining leaves are rapidly killed resulting in premature ripening and large<br />

reductions in yield.<br />

The arrival <strong>of</strong> black <strong>leaf</strong> streak disease in Latin America coincided with the<br />

introduction <strong>of</strong> systemic fungicides to control Sigatoka disease on plantations.<br />

The fungicides were also effective against black <strong>leaf</strong> streak disease, but many more<br />

applications were needed to maintain control with corresponding increases in<br />

production costs. It also became much more important to prevent the disease<br />

building-up too much as loss <strong>of</strong> control could have serious consequences. Good<br />

disease control management became essential and those that couldn’t were in<br />

serious difficulties.<br />

Initial effects on commercial banana production soon after the arrival <strong>of</strong> the<br />

disease were devastating. In the South Pacific, only 49% <strong>of</strong> unsprayed fruit reached<br />

export quality (Firman, 1972). Fiji ceased exporting banana fruit in 1974 and<br />

Samoa in 1984. Exports also dropped in Tonga and the Cook Islands because<br />

export quality could not be achieved (Fullerton, 1987). In Central America, export<br />

banana cultivation has survived, but at a price. Costs <strong>of</strong> control in Costa Rica<br />

are now US$900-1500 hectare/year. The overuse <strong>of</strong> fungicides with the same mode<br />

<strong>of</strong> action, and the use <strong>of</strong> fungicides below recommended doses has led to increased<br />

resistance to fungicides in M. fijiensis populations. Different strategies such as<br />

alternation <strong>of</strong> fungicides with different modes <strong>of</strong> action and the use <strong>of</strong> fungicide<br />

mixtures had to be adopted. Monitoring the resistances <strong>of</strong> local populations <strong>of</strong><br />

M. fijiensis to a range <strong>of</strong> fungicides is now routine. In Costa Rica, the rising costs<br />

<strong>of</strong> labour and <strong>leaf</strong> <strong>spot</strong> control are making commercial banana cultivation increasingly<br />

uneconomical. Other dessert <strong>bananas</strong> grown for local consumption, such<br />

34


Session 1<br />

D.R. Jones<br />

as clones in the Pome subgroup (AAB) and ‘Silk’ (AAB), which are popular in<br />

Brazil and India, are also susceptible to black <strong>leaf</strong> streak disease.<br />

Plantain cultivation has also been seriously affected by M. fijiensis. The pathogen<br />

has caused a considerable decrease in the availability <strong>of</strong> fruit for local consumption<br />

with a corresponding substantial increase in market price. In many areas <strong>of</strong> Central<br />

America, growers have either gone out <strong>of</strong> business or have formed cooperatives to<br />

share resources to cover the costs <strong>of</strong> spray equipment and chemicals. In Panama,<br />

plantain production was estimated to have decreased by 69% and prices to have<br />

increased by 50% between 1979 and 1984 (Bureau, 1990). In Costa Rica, black <strong>leaf</strong><br />

streak disease was calculated to have reduced production by 40% by 1982 (Romero,<br />

1986). Similar effects occurred in the plantain industry in Colombia, South America.<br />

After the introduction <strong>of</strong> black <strong>leaf</strong> streak disease, plantains became scarce and<br />

expensive and consumers changed to cheaper foods (Belalcazar, 1991). Black <strong>leaf</strong> streak<br />

disease also affects plantain production in Africa and endangers food security for<br />

many poor people. On poor sandy soils in West Africa, it has been estimated that<br />

yield losses are 33% and 76% during the plant and ratoon cropping cycle respectively<br />

(Mobambo et al., 1996). This has led to small-scale farmers abandoning plantain<br />

cultivation.<br />

Plantain is not the only cooking banana type grown by subsistence farmers and<br />

other small-scale growers to be affected by black <strong>leaf</strong> streak. In the Great Lakes area<br />

<strong>of</strong> Africa, the East African highland cultivars in the Lujugira-Mutika subgroup (AAA)<br />

are susceptible. Losses <strong>of</strong> 37% due to the combined affects <strong>of</strong> black <strong>leaf</strong> streak disease<br />

and Cladosporium speckle have been reported (Tushemereirwe, 1996). In the Pacific,<br />

cultivars in the popular Maia Maoli-Popoulu subgroup (AAB) are susceptible (Carlier<br />

et al., 2000a).<br />

Interactions between <strong>Mycosphaerella</strong> fijiensis<br />

and <strong>Mycosphaerella</strong> musicola<br />

Soon after M. fijiensis was discovered on Fiji, it was reported to be displacing<br />

M. musicola as the dominant <strong>leaf</strong> <strong>spot</strong>. Displacement occurred in many coastal areas<br />

in the tropics particularly in Latin America. At elevation, M. musicola has an<br />

advantage being suited to cooler conditions and there are a number <strong>of</strong> reports (include<br />

references here) <strong>of</strong> the two pathogens co-existing at heights <strong>of</strong> around 1200m to<br />

1500m with M. musicola dominating at higher altitudes and M. fijiensis at lower<br />

ones. There is also evidence that M. fijiensis may be slowly adapting to higher<br />

elevations (Carlier et al., 2000a).<br />

M. musicola has probably not disappeared completely from banana in coastal<br />

areas dominated by M. fijiensis. Evidence <strong>of</strong> small amounts <strong>of</strong> survival within the<br />

<strong>leaf</strong> <strong>spot</strong> population comes from Nigeria where ‘SH-3362’ (AA), a hybrid resistant<br />

to M. fijiensis and susceptible to M. musicola, was found to be affected by<br />

M. musicola at planting (C. Pasberg-Gauhl and F. Gauhl, Nigeria, personal<br />

communication). In the Philippines, M. musicola has also still been reported to be<br />

present despite the dominance <strong>of</strong> M. fijiensis in commercial plantations. This may<br />

be because the great genetic diversity <strong>of</strong> cultivated banana in this country has meant<br />

some clones, like ‘Amas’ (AA, syn. ‘Sucrier’), which is more susceptible to M. musicola<br />

35


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

than M. fijiensis, have maintained the former pathogen. In other areas <strong>of</strong> Southeast<br />

Asia, the situation has been more difficult to understand. This is discussed more fully<br />

under “<strong>Mycosphaerella</strong> eumusae”.<br />

<strong>Mycosphaerella</strong> eumusae, the cause <strong>of</strong> eumusae<br />

<strong>leaf</strong> <strong>spot</strong><br />

Confusion between <strong>leaf</strong> <strong>spot</strong>s in Asia<br />

The latest <strong>of</strong> the three <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens to affect banana was only<br />

recognised in the mid-1990s (Carlier et al., 2000a). This was because <strong>of</strong> uncertainties<br />

concerning the distribution <strong>of</strong> Sigatoka disease and black <strong>leaf</strong> streak disease in<br />

Southeast and South Asia. The uncertainty arose because the usual rapid displacement<br />

<strong>of</strong> M. musicola by M. fijiensis in tropical lowland areas, as had occurred in the<br />

Americas and West Africa, did not seem to have happened in Java (Indonesia), West<br />

Malaysia and Thailand (Jones, 1990). Although M. fijiensis had first been recorded<br />

at these locations in the mid to late 1960s, M. musicola was still present near Bogor<br />

in Java and Kuala Lumpur in West Malaysia in 1976 (Stover, 1976). From<br />

observations ten years later in 1988, the author believed Sigatoka disease was still<br />

the dominant <strong>leaf</strong> <strong>spot</strong> occurring in Java, West Malaysia and Thailand. If M. fijiensis<br />

was present, what was stopping it from becoming the dominant <strong>leaf</strong> <strong>spot</strong>? There<br />

was also a problem concerning the <strong>leaf</strong> <strong>spot</strong> situation in South Asia. If M. fijiensis<br />

had been recorded in Bhutan in 1985 (Peregrine, 1989), then why hadn’t it since<br />

been found in neighbouring India?<br />

First records<br />

It became possible between 1992 and 1995 for the author, who was employed at<br />

the time by the International Network for the Improvement <strong>of</strong> Banana and Plantain<br />

(INIBAP), to collect specimens <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> in the Southeast Asian/South Asian<br />

region during visits for diagnosis. He hoped that this would help determine the<br />

distribution <strong>of</strong> the two main <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens, which may help<br />

explain the apparent lack <strong>of</strong> expansion <strong>of</strong> M. fijiensis in the region. Specimens,<br />

which were thought to be mainly <strong>of</strong> M. musicola, were collected at different<br />

locations and on different banana clones in southern India, Sri Lanka, West<br />

Malaysia and Thailand. The specimens were sent by courier to Drs Xavier<br />

Mourichon and Jean Carlier at the Centre de coopération internationale en<br />

recherche agronomique pour le développement (CIRAD) in Montpellier for<br />

identification. Unexpectantly, M. musicola was not identified from any <strong>of</strong> the<br />

specimens collected. Some specimens from Johore in West Malaysia were found<br />

to be M. fijiensis, but the majority <strong>of</strong> <strong>leaf</strong> <strong>spot</strong>s from all countries were caused<br />

by a fungus that was unknown. This fungus had <strong>Mycosphaerella</strong> as the perfect<br />

(teleomorph) stage and what first appeared to be Septoria as its imperfect<br />

(anamorph) stage (Anon., 1995; Carlier et al., 2000b). Initially, it was believed<br />

that the pathogen might have been Phaeoseptoria musae (Anon.,1995), which has<br />

been reported to have <strong>Mycosphaerella</strong> as a perfect stage and is fairly widespread<br />

36


Session 1<br />

D.R. Jones<br />

(Carlier et al., 2000a). However, this turned out not to be the case (Carlier et al.,<br />

2000b). The new fungus was named M. eumusae (Carlier et al., 2000b).<br />

Because <strong>of</strong> the similarity <strong>of</strong> symptoms with those <strong>of</strong> Sigatoka disease (Carlier et al.,<br />

2000a), it is likely that the <strong>leaf</strong> <strong>spot</strong> caused by M. eumusae was seen by the author in<br />

Malaysia and Thailand in 1988. Evidence suggested that it might have been the common<br />

<strong>leaf</strong> <strong>spot</strong> <strong>of</strong> banana in the South Asia and parts <strong>of</strong> Southeast Asia. If so, M. eumusae<br />

competed effectively with M. fijiensis and prevented it from becoming dominant.<br />

A specimen <strong>of</strong> a <strong>leaf</strong> <strong>spot</strong> collected in the Mekong Delta in Vietnam by Ivan<br />

Buddenhagen in 1995 was later identified as M. eumusae at CIRAD, as were specimens<br />

from Mauritius in 1997 (Carlier et al., 2000b) and Réunion in 2000 (X. Mourichon,<br />

France, personal communication). Re-examination <strong>of</strong> specimens collected at Onne,<br />

Nigeria in 1989 and 1990, when M. fijiensis was presumed to be present, also revealed<br />

the presence <strong>of</strong> M. eumusae (Carlier et al., 2000b). Details <strong>of</strong> findings <strong>of</strong> M. eumusae<br />

that are documented by CIRAD are summarised in Table 3.<br />

Table 3. Countries where <strong>Mycosphaerella</strong> eumusae, the causal agent <strong>of</strong> eumusae <strong>leaf</strong> <strong>spot</strong> disease, has been detected<br />

in chronological order <strong>of</strong> records (modified from Table 1 <strong>of</strong> Carlier et al., 2000b with additional information supplied<br />

by X. Mourichon). All identifications made by J. Carlier and M.F. Zapater, CIRAD.<br />

Country Banana host Year specimen was<br />

collected and collector<br />

Nigeria (Onne) AAB clone, most likely in plantain subgroup 1989 (IITA)<br />

Nigeria (Onne) AAB clone, most likely in plantain subgroup 1990 (IITA)<br />

India (Bangalore) ‘Grande naine’ (AAA, Cavendish subgroup) 1992 (D.R. Jones, INIBAP)<br />

Malaysia (Johor State) ‘Pisang kapas’ (AA or AAB) 1993 (D.R. Jones, INIBAP)<br />

Thailand (Sukothai) ‘Grande naine’ (AAA, Cavendish subgroup) 1994 (D.R. Jones, INIBAP)<br />

Thailand (Surat Thani) ‘Williams’ (AAA, Cavendish subgroup) 1994 (D.R. Jones, INIBAP)<br />

Thailand (Tha Yang) ‘Kluai hom thong’ (AAA) 1994 (D.R. Jones, INIBAP)<br />

India (Kannara) ‘Grande naine’ (AAA, Cavendish subgroup) 1995 (D.R. Jones, INIBAP)<br />

Sri Lanka (Gannoruwa) AAA clone in Cavendish subgroup 1995 (D.R. Jones, INIBAP)<br />

Sri Lanka (Nugahena) ‘Anamala’ (AAA, syn. ‘Gros Michel’) 1995 (D.R. Jones, INIBAP)<br />

Vietnam (Mekong Delta) ‘Sucrier’ (AA) 1995 (I. Buddenhagen)<br />

Mauritius ‘Grande naine’ (AAA, Cavendish subgroup) 1997 (S.P. Beni Madhu, AREU)<br />

Réunion ‘Grande naine’ (AAA, Cavendish subgroup) 2000 (C. Lavigne, CIRAD)<br />

Recently, a re-examination <strong>of</strong> M. eumusae specimens and cultures has revealed<br />

that the imperfect stage <strong>of</strong> the fungus is likely to be Pseudocercospora and not<br />

Septoria as originally thought. However, the fungus can be distinguished from<br />

M. musicola and M. fijiensis on morphological grounds (Crous and Mourichon, 2002)<br />

and by ITS sequence analysis (Carlier et al., 2000b). A change in the common name<br />

<strong>of</strong> the disease from Septoria <strong>leaf</strong> <strong>spot</strong> disease to eumusae <strong>leaf</strong> <strong>spot</strong> disease has been<br />

proposed (Crous and Mourichon, 2002).<br />

Impact<br />

The economic impact <strong>of</strong> M. eumusae as a <strong>leaf</strong> pathogen <strong>of</strong> banana has still to be<br />

evaluated. However, it is known to seriously affect cultivars in the important AAA<br />

Cavendish and AAB plantain subgroups in southern India. It has also been observed<br />

37


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

causing large areas <strong>of</strong> necrosis on leaves <strong>of</strong> ‘Anamala’ (AAA, syn. ‘Gros Michel’) in<br />

Sri Lanka. Other cultivars/clones seen with symptoms are ‘Kluai lep mu nang’ (AA),<br />

‘Pisang mas’ (AA), ‘Pisang kapas’ (AA) and ‘Mysore’ (AAB) (Carlier et al., 2000a).<br />

More research needs to be undertaken on the effect <strong>of</strong> M. eumusae on important<br />

clones.<br />

Summary and discussion<br />

The three <strong>Mycosphaerella</strong> species causing <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> banana reported<br />

above are serious pathogens. M. musicola and M. fijiensis were well documented<br />

in the scientific literature because <strong>of</strong> their steady global or near global spread and<br />

impact on banana cultivation. However, M. fijiensis, the more recent has had a<br />

far greater impact. It is more difficult to control on plantations <strong>of</strong> dessert <strong>bananas</strong><br />

for export and local consumption, and also affects the production <strong>of</strong> cooking<br />

<strong>bananas</strong> grown by resource-poor farmers. The arrival <strong>of</strong> M. fijiensis in West Africa<br />

and the perceived threat that it posed to the livelihood <strong>of</strong> the peoples there led to<br />

the formation <strong>of</strong> INIBAP and the plantain breeding programmes <strong>of</strong> the International<br />

Institute <strong>of</strong> Tropical Agriculture (IITA) and the Centre africain de recherches sur<br />

bananiers et plantains (CARBAP).<br />

The third <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogen is not so well known because it<br />

has been overlooked until fairly recently. Evidence on distribution suggests that<br />

it arose, like the others, in the Southeast Asian/Australasian region. It is not known<br />

how long M. eumusae has affected banana in South and Southeast Asia, nor the<br />

extent <strong>of</strong> its distribution. Extensive surveys <strong>of</strong> <strong>leaf</strong> <strong>spot</strong>s in the region would help<br />

clarify the situation. An estimate <strong>of</strong> the severity <strong>of</strong> the disease and the name <strong>of</strong><br />

the clone affected would help determine host reactions. Basic information on the<br />

biology and epidemiology <strong>of</strong> the pathogen is also needed.<br />

M. eumusae has been present in Onne, Nigeria for at least 13 years but was<br />

not found in a recent thorough survey <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> organisms in neighbouring<br />

Cameroon (C. Abadie, France, personal communication). Therefore, in West Africa,<br />

the pathogen may still be confined to southeast Nigeria. If so, one must speculate<br />

on how it got here in the first place. Was M. eumusae introduced with planting<br />

material from Asia?<br />

M. eumusae seems to be an important pathogen and may be able to compete<br />

with M. fijiensis. The apparent dominance <strong>of</strong> M. eumusae in parts <strong>of</strong> Asia suggests<br />

that it was established before the introduction <strong>of</strong> M. fijiensis and thus perhaps<br />

able to resist intrusion by the latter. The situation at Onne, Nigeria, may be different<br />

with M. eumusae because arrival was after M. fijiensis became dominant over<br />

M. musicola and subsequent competitive interactions favoured the established<br />

pathogen. Work is needed to determine if this speculation has any basis in fact.<br />

The identification <strong>of</strong> the <strong>leaf</strong> <strong>spot</strong> pathogens found on plantain at Onne may help<br />

determine relative competitiveness on the main susceptible host in the area. The<br />

identification <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> pathogens on cultivars in the IITA germplasm collection<br />

may help to determine the nature <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> interactions on other clones.<br />

When M. fijiensis first appeared, some believed that it might have arisen in<br />

Fiji by mutation from M. musicola. Stover (1969) initially considered M. fijiensis<br />

38


Session 1<br />

D.R. Jones<br />

to be a physiological strain <strong>of</strong> M. musicola. The latter hypothesis is now considered<br />

to be unlikely. How do we view the appearance <strong>of</strong> M. eumusae? How are the <strong>leaf</strong><br />

<strong>spot</strong> pathogens evolving? Recent phylogenetic studies indicate that M. musicola,<br />

M. fijiensis and M. eumusae may once have had a common origin (P. Crous, South<br />

Africa, personal communication). All may have arisen from similar saprophytic<br />

or weakly pathogenic fungi growing on damaged or weakened <strong>leaf</strong> tissues <strong>of</strong><br />

banana. Stover (1969) reported M. minima as a saprophytic co-inhabitant with<br />

M. musicola in <strong>leaf</strong> <strong>spot</strong>s. He has also recorded M. musae was an endophyte in<br />

Sigatoka <strong>leaf</strong> <strong>spot</strong>s (Stover, 1994). Other fungi could be evolving as parasites in<br />

senescing <strong>leaf</strong> tissue. Further investigations <strong>of</strong> speciation in <strong>Mycosphaerella</strong> and<br />

other related genera found on banana leaves in the centre <strong>of</strong> origin <strong>of</strong> banana<br />

and elsewhere may prove interesting.<br />

The similarity <strong>of</strong> the necrotic symptoms caused by M. musicola, M. fijiensis<br />

and M. eumusae and the fact that all three have <strong>Mycosphaerella</strong> as the perfect<br />

stage suggests that the <strong>diseases</strong> should be grouped together as Sigatoka <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong>. These <strong>diseases</strong> warrant fundamental research to clarify their evolution<br />

and adaptability, and to help to find ways <strong>of</strong> breeding resistant <strong>bananas</strong>.<br />

Acknowledgements<br />

The author thanks the Australian Banana Growers’ Council for financial support to attend<br />

the workshop on <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens.<br />

References<br />

Anonymous 1995. Musanews. INFOMUSA 2(2):26-30.<br />

Belalcazar S.L. 1991. El Cultivo del Plátano en el Trópico. Pp. 288-289 in Manual de Assistencia<br />

Técnica No. 50. Instituto Colombiano Agropecuario, Armenia, Colombia.<br />

Bureau E. 1990. Adoption <strong>of</strong> a forecasting system to control black Sigatoka (<strong>Mycosphaerella</strong><br />

fijiensis Morelet) in plantain plantations in Panama. Fruits 45:329-338.<br />

Carlier J. E. Fouré, F. Gauhl, D.R. Jones, P. Lepoivre, X. Mourichon and C. Pasberg-Gauhl.<br />

2000a. Fungal <strong>diseases</strong> <strong>of</strong> the foliage. Pp. 37-142 in Diseases <strong>of</strong> Banana, Abacá and Enset<br />

(D.R. Jones, ed.). CABI Publishing, Wallingford, UK.<br />

Carlier J., M.F. Zapater, F. Lapeyre, D.R. Jones and X. Mourichon. 2000b. Septoria <strong>leaf</strong> <strong>spot</strong><br />

<strong>of</strong> banana: a newly discovered disease caused by <strong>Mycosphaerella</strong> eumusae (anamorph<br />

Septoria eumusae). Phytopathology 90:884-890.<br />

Carreel F. 1995. Etude de la diversité génétique des bananiers (genre Musa) à l’aide des<br />

marqueurs RFLP. PhD thesis, Institut National Agronomique, Paris-Grignon, France.<br />

Crous P. and X. Mourichon. 2002. <strong>Mycosphaerella</strong> eumusae and its anamorph Pseudocercospora<br />

eumusae spp. nov.: causal agent <strong>of</strong> eumusae <strong>leaf</strong> <strong>spot</strong> disease <strong>of</strong> banana. Sydowia 54:<br />

35-43.<br />

Dabek A.J. and J.M. Waller. 1990. Black <strong>leaf</strong> streak and viral streak: new banana <strong>diseases</strong> in<br />

east Africa. Tropical Pest Management 36(2):157-158.<br />

Firman I.D. 1972. Susceptibility <strong>of</strong> banana cultivars to fungus <strong>diseases</strong> in Fiji. Tropical<br />

Agriculture (Trinidad) 49:189-196.<br />

Frossard P. 1980. Apparition d’une nouvelle et grave maladie foliaire des bananiers et<br />

plantains au Gabon: la maladie des raies noires, <strong>Mycosphaerella</strong> fijiensis Morelet. Fruits<br />

35:443-453.<br />

39


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Fullerton R.A. 1987. Banana production in selected Pacific islands. Pp. 57-62 in Banana<br />

and Plantain Breeding Strategies, Proceedings <strong>of</strong> an International Workshop held in<br />

Cairns, Australia, 13-17 October 1986 (G.J. Persley and E.A. De Langhe, eds). ACIAR<br />

Proceedings 21, Australian Centre for International Agricultural Research, Canberra,<br />

Australia.<br />

Jones D.R. 1990. Black Sigatoka in the Southeast Asian–Pacific region. Musarama 3 (1):2-5.<br />

Jones D.R. 2000. Introduction to banana, abacá and enset. Pp. 1-36 in Diseases <strong>of</strong> Banana,<br />

Abacá and Enset (D.R. Jones, ed.). CABI Publishing, Wallingford, UK.<br />

Jones D.R. and J.L. Alcorn. 1982. Freckle and black Sigatoka <strong>diseases</strong> <strong>of</strong> banana in far north<br />

Queensland. Australasian Plant Pathology 11:7-9.<br />

Leach R. 1941. Banana <strong>leaf</strong> <strong>spot</strong> <strong>Mycosphaerella</strong> musicola, the perfect stage <strong>of</strong> Cercospora<br />

musae Zimm. Tropical Agriculture (Trinidad) 18:91-95.<br />

Leach R. 1946. Banana Leaf Spot (<strong>Mycosphaerella</strong> musicola) on the Gros Michel Variety in<br />

Jamaica. Investigations into the Aetiology <strong>of</strong> the Disease and Principles <strong>of</strong> Control by<br />

Spraying. Bulletin, Government Printer, Kingston, Jamaica.<br />

Leach R. 1964a. Report on Investigations into the Cause and Control <strong>of</strong> the New Banana<br />

Diseases in Fiji, Black Leaf Streak. Council Papers Fiji 38, Suva.<br />

Leach R. 1964b. A new form <strong>of</strong> banana <strong>leaf</strong> <strong>spot</strong> in Fiji, black <strong>leaf</strong> streak. World Crops 16:D60-<br />

64.<br />

Long P.G. 1979. Banana black <strong>leaf</strong> streak disease (<strong>Mycosphaerella</strong> fijiensis) in Western Samoa.<br />

Transactions <strong>of</strong> the British Mycological Society 72:299-310.<br />

Magee C.J.P. 1953. Some aspects <strong>of</strong> the bunchy top disease <strong>of</strong> banana and other Musa spp.<br />

Journal and Proceedings <strong>of</strong> the Royal Society <strong>of</strong> New South Wales 87:3-18.<br />

Massee G. 1914. Fungi exotici: XVIII. Kew Bulletin 1914, 159.<br />

Meredith D.S. 1970. Banana Leaf Spot Disease (Sigatoka) caused by <strong>Mycosphaerella</strong> musicola<br />

Leach. Phytopathological Papers, No. 11, Commonwealth Mycological Institute, Kew,<br />

Surrey, UK, 147pp.<br />

Meredith D.S. and J.S. Lawrence. 1969. Black <strong>leaf</strong> streak disease <strong>of</strong> <strong>bananas</strong> (<strong>Mycosphaerella</strong><br />

fijiensis): Symptoms <strong>of</strong> disease in Hawaii, and notes on the conidial state <strong>of</strong> the causal<br />

fungus. Transactions <strong>of</strong> the British Mycological Society 52:459-476.<br />

Mobambo K.N., F. Gauhl, R. Swennen and C. Pasberg-Gauhl. 1996. Assessment <strong>of</strong> the cropping<br />

cycle effects <strong>of</strong> black <strong>leaf</strong> streak severity and yield decline <strong>of</strong> plantain and plantain hybrids.<br />

International Journal <strong>of</strong> Pest Management 42:1-8.<br />

Mulder J.L. and R.H. Stover. 1976. <strong>Mycosphaerella</strong> species causing banana <strong>leaf</strong> <strong>spot</strong>.<br />

Transactions <strong>of</strong> the British Mycological Society 67:77-82.<br />

Parnell M., P.J.A. Burt and K. Wilson. 1998. The influence <strong>of</strong> exposure to ultraviolet radiation<br />

in simulated sunlight on ascospores causing Black Sigatoka disease <strong>of</strong> banana and plantain.<br />

International Journal <strong>of</strong> Biometeorology 42:22-27.<br />

Peregrine W.T.H. 1989. Black <strong>leaf</strong> streak <strong>of</strong> banana <strong>Mycosphaerella</strong> fijiensis Morelet. FAO Plant<br />

Protection Bulletin 37(3):130.<br />

Philpott J. and C.H. Knowles. 1913. Report on a Visit to Sigatoka. Phamphlet <strong>of</strong> the Department<br />

<strong>of</strong> Agriculture, Fiji 3, Suva.<br />

Pons N. 1987. Notes on <strong>Mycosphaerella</strong> fijiensis var. difformis. Transactions <strong>of</strong> the British<br />

Mycological Society 89:120-124.<br />

Raemaekers R. 1975. Black <strong>leaf</strong> streak disease in Zambia. PANS 21:396-400.<br />

Rhodes P.L. 1964. A new banana disease in Fiji. Commonwealth Phytopathological News 10:<br />

38-41.<br />

Robinson J.C. 1996. Bananas and Plantains. Crop Production Science in Horticulture 5, CAB<br />

International, Wallingford, UK, 238pp.<br />

40


Session 1<br />

D.R. Jones<br />

Romero C. R. 1986. Impacto de Sigatoka negra y roya del cafeto en actividad platanera<br />

nacional. Revista de la Asociación Bananera Nacional (ASBANA), San José, Costa Rica<br />

12(9):7-10.<br />

Stover R.H. 1962. Intercontinental spread <strong>of</strong> banana <strong>leaf</strong> <strong>spot</strong> (<strong>Mycosphaerella</strong> musicola Leach).<br />

Tropical Agriculture (Trinidad) 29:327-338.<br />

Stover R.H. 1969. The <strong>Mycosphaerella</strong> species associated with banana <strong>leaf</strong> <strong>spot</strong>s. Tropical<br />

Agriculture (Trinidad) 46:325-332.<br />

Stover R.H. 1972. Banana, Plantain and Abaca Diseases. Commonwealth Mycological<br />

Institute, Kew, Surrey, UK, 316pp.<br />

Stover R.H. 1976. Distribution and cultural characteristics <strong>of</strong> the pathogens causing <strong>leaf</strong> <strong>spot</strong>.<br />

Tropical Agriculture (Trinidad) 53:111-114.<br />

Stover R.H. 1978. Distribution and probable origin <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis in Southeast<br />

Asia. Tropical Agriculture (Trinidad) 55:65-68.<br />

Stover R.H. 1980. Sigatoka <strong>leaf</strong> <strong>spot</strong>s <strong>of</strong> banana and plantain. Plant Disease 64:750-755.<br />

Stover R.H. 1990. Sigatoka <strong>leaf</strong> <strong>spot</strong>s: thirty years <strong>of</strong> changing control strategies: 1959-1989.<br />

Pp. 66-74 in Sigatoka Leaf Spot Diseases <strong>of</strong> Bananas, Proceedings <strong>of</strong> an International<br />

Workshop held in San José, Costa Rica, March 28-April 1, 1989 (R.A. Fullerton and<br />

R.H. Stover, eds). INIBAP, Montpellier, France.<br />

Stover R.H. 1994. <strong>Mycosphaerella</strong> musae and Cercospora “non-virulentum” from Sigatoka<br />

<strong>leaf</strong> <strong>spot</strong>s are identical. Fruits 49:187-190.<br />

Tushemereirwe W.K. 1996. Factors influencing the expression <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> highland<br />

<strong>bananas</strong> in Uganda. PhD thesis, University <strong>of</strong> Reading, UK, 197 pp.<br />

Zimmerman A. 1902. Uber einige tropischer Kulturpflanzen beobachtete Pilze. Zentralblatt<br />

für Bakteriologie, Parasitenkunde, Infektionskrankheiten und Hygiene 8:219 (abs).<br />

41


Session 1<br />

P.W. Crous et al.<br />

Integrating morphological and<br />

molecular data sets on <strong>Mycosphaerella</strong>,<br />

with specific reference to species<br />

occurring on Musa<br />

P. W. Crous 1 ,J.Z. Groenewald 1 ,A.Aptroot 2 ,U.Braun 3 ,<br />

X. Mourichon 4 and J. Carlier 4<br />

Abstract<br />

The genus <strong>Mycosphaerella</strong> (= Sphaerella) is one <strong>of</strong> the largest genera <strong>of</strong> ascomycetes, containing<br />

more than 3000 named taxa. Approximately 23 anamorph genera have been linked to<br />

<strong>Mycosphaerella</strong> via cultural studies. Several <strong>of</strong> these anamorph genera have recently been<br />

reduced to synonymy based on phylogenetic studies derived from ITS1, 5.8S and ITS2 DNA sequence<br />

data. In addition, several genera not previously associated with <strong>Mycosphaerella</strong>,have also been<br />

shown to cluster within <strong>Mycosphaerella</strong>, which has proved to be largely monophyletic. From these<br />

results, as well as a re-evaluation <strong>of</strong> the criteria upon which anamorph genera are distinguished<br />

in this complex, a reduced set <strong>of</strong> informative criteria and genera are proposed.The degree <strong>of</strong> scar<br />

thickening, darkening and refraction, as well as the presence or absence <strong>of</strong> pigmentation in<br />

conidiophores and conidia appear to be useful features delimiting anamorph genera <strong>of</strong><br />

<strong>Mycosphaerella</strong>. Species, however, are still separated on a combination <strong>of</strong> characters such as<br />

conidiomatal structure, the nature and arrangement <strong>of</strong> conidiophores, conidiogenesis, dehiscence<br />

scars and pigmentation. For the species that occur on Musa, anamorph morphology appears to<br />

be more informative than the more conserved teleomorph morphology, and can be used to<br />

separate the major pathogens, namely M. fijiensis (the causal agent <strong>of</strong> black <strong>leaf</strong> streak disease),<br />

M. musicola (the causal agent <strong>of</strong> Sigatoka disease), M. eumusae (the causal agent <strong>of</strong> eumusae<br />

<strong>leaf</strong> <strong>spot</strong> disease), as well as other reputedly less important pathogens.<br />

Resumen - Integración de los conjuntos de datos morfológicos y moleculares en<br />

<strong>Mycosphaerella</strong>, con referencia específica a las especies que ocurren en Musa<br />

El género <strong>Mycosphaerella</strong> (= Sphaerella) es uno de los géneros más grandes de ascomicetos, que<br />

contiene más de 3000 taxa nombrados. Aproximadamente 23 géneros anamorfos diferentes han<br />

estado vinculados con <strong>Mycosphaerella</strong> mediante estudios culturales. Varios de estos géneros<br />

anamorfos fueron reducidos recientemente debido a la sinonimia basada en estudios filogenéticos<br />

1<br />

University <strong>of</strong> Stellenbosch, Matieland, South Africa<br />

2<br />

Centraalbureau voor Schimmelcultures,Utrecht, The Netherlands<br />

3<br />

Martin-Luther-Universität, Halle (Saale), Germany<br />

4<br />

CIRAD, Montpellier, France<br />

43


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

derivados de los datos de las secuencias de ADN ITS1, 5.8S y ITS2. En adición, se ha demostrado que<br />

varios géneros que no estaban asociados previamente con <strong>Mycosphaerella</strong>, actualmente están<br />

agrupados dentro del género <strong>Mycosphaerella</strong>, el cual está comprobado, que es extensamente<br />

mon<strong>of</strong>ilético. De estos resultados, así como de una nueva evaluación de los criterios de acuerdo a<br />

los cuales se distinguen géneros anamorfos en este complejo, se propone un conjunto reducido de<br />

criterios y géneros informativos. El grado del espesor, oscurecimiento y refracción de las cicatrices,<br />

así como la presencia o ausencia de pigmentación en conidióforos y conidios parecen representar<br />

características útiles que delimitan los géneros anamorfos de <strong>Mycosphaerella</strong>. Sin embargo, las<br />

especies aún están separadas en una combinación de caracteres como la estructura conidiomatal,<br />

la naturaleza y el arreglo de los conidióforos, conidiogénesis, cicatrices de dehiscencia y pigmentación.<br />

Para las especies que ocurren en Musa, la morfología anamorfa parece ser más informativa que la<br />

morfología teleomorfa más conservada, y puede ser utilizada para separar los principales patógenos,<br />

a saber M. eumusae (agente causal del ELSD), M. musicola (agente causal de la enfermedad de<br />

Sigatoka), M. fijiensis (agente causal de la raya negra de la hoja).<br />

Résumé - L’intégration des données morphologiques et moléculaires sur <strong>Mycosphaerella</strong>,<br />

particulièrement celles des espèces présentes sur Musa<br />

Le genre <strong>Mycosphaerella</strong> (= Sphaerella) est un des genres les plus représentés des ascomycètes<br />

avec plus de 3000 taxa. Environ 23 genres anamorphes ont été liés à <strong>Mycosphaerella</strong> à l’aide<br />

d’études sur les cultures. Des études phylogénétiques à partir de séquences d’ADN ITS1, 5.8S<br />

et ITS2, ont permis d’identifier les synonymes parmi ces genres anamorphes. De plus,<br />

plusieurs genres qui n’étaient pas associés auparavant à <strong>Mycosphaerella</strong>, se regroupent dans<br />

ce genre qui s’est avéré être principalement monophylétique. A partir de ces résultats ainsi<br />

que par la réévaluation des critères à partir desquels on peut distinguer les genres<br />

anamorphes, nous proposons une réduction du nombre de critères pertinents et de genres.<br />

Le degré d’épaississement, de noircissement et de réfraction des cicatrices, ainsi que la<br />

présence ou l’absence de pigmentation dans les conidiophores et les conidies, semblent être<br />

des critères intéressants pour délimiter les genres anamorphes de <strong>Mycosphaerella</strong>. Les<br />

espèces se distinguent toutefois par une combinaison de caractères tels que la structure des<br />

conidioma, la nature et la disposition des conidiophores, la conidiogénèse, les cicatrices de<br />

déhiscence et la pigmentation. Pour les espèces qui se trouvent sur Musa, la morphologie<br />

de l’anamorphe est plus instructive que la morphologie moins variable du téléomorphe et<br />

peut être utilisée pour distinguer les principaux pathogènes, soit M. fijiensis (l’agent causal<br />

de la maladie des raies noires), M. musicola (l’agent causal de la maladie de Sigatoka),<br />

M. eumusae (l’agent causal de l’ELSD, eumusae <strong>leaf</strong> <strong>spot</strong> disease), ainsi que d’autres pathogènes<br />

considérés moins importants.<br />

What is <strong>Mycosphaerella</strong>?<br />

Corlett (1991, 1995) lists more than 2000 species belonging to the genus<br />

<strong>Mycosphaerella</strong> Johanson (Dothideales: <strong>Mycosphaerella</strong>ceae) (including the fungi<br />

described in Sphaerella Ces. et De Not.). This makes it one <strong>of</strong> the largest genera<br />

<strong>of</strong> ascomycetes known. Furthermore, this genus has been confirmed to have<br />

anamorphs in at least 23 different genera (Crous et al., 2000), including Cercospora<br />

Fres. and Septoria Sacc., which alone encompass several thousand species (Pollack,<br />

1987; Sutton, 1980). Some species are saprobes, but most are known from necrotic<br />

lesions that are associated with leaves, stems or fruit <strong>of</strong> various hosts (Park and<br />

Keane, 1984). Host specificity still plays a major role in the identification <strong>of</strong> species,<br />

where taxa are chiefly compared with those that occur on a specific host genus<br />

or family <strong>of</strong> phanerogamic plants (Chupp, 1954; Braun, 1995). Von Arx (1949)<br />

44


Session 1<br />

P.W. Crous et al.<br />

regarded some species as polyphagous but the concept <strong>of</strong> host specificity is still<br />

strongly adhered to, especially with species shown to be plant pathogens.<br />

Species <strong>of</strong> <strong>Mycosphaerella</strong> are usually defined as having small, black, immersed<br />

or erumpent to almost superficial pseudothecial ascomata, with various degrees<br />

<strong>of</strong> stromatic tissue surrounding the ascomata, and pale to brown superficial<br />

mycelium being present or absent, smooth or verruculose. The ascomatal wall<br />

consists <strong>of</strong> 3–6 layers <strong>of</strong> pseudoparenchyma cells but in some taxa this can also<br />

be more prominently developed, especially around the ostiole. Ostioles are<br />

singular, central, and usually lined with periphyses; in some taxa, however, this<br />

develops further, giving the impression <strong>of</strong> periphysoids. The hamathecium dissolves<br />

at maturity, and no stromatic tissue remains between the asci. Asci are bitunicate,<br />

8-spored, sessile, arranged in a cluster, hyaline, ovoid to obovoid, ellipsoidal or<br />

cylindrical, rarely clavate. Ascospores are bi to multi-seriate, thin to thick-walled,<br />

guttulate or not, 1-septate, constricted at the septum or not, usually hyaline<br />

and smooth, rarely with a mucous sheath, fusoid to obovoid, ellipsoid or elongate.<br />

Anamorphs are highly variable, including numerous hyphomycete and coelomycete<br />

genera, namely Cercospora Fres., Cercosporella Sacc., Cladosporium Link, Clypeispora<br />

Ramaley, Colletogloeopsis Crous et M.J. Wingf., Dissoconium De Hoog,<br />

Oorschot et Hijwegen, Fusicladiella Höhn., Miuraea Hara, Passalora Fr.,<br />

Phaeophleospora Rangel, Phloeospora Wallr., Pseudocercospora Speg., Pseudocercosporella<br />

Deighton, Ramularia Unger, Septoria Sacc., Sonderhenia H.J. Swart<br />

and J. Walker, Stenella Syd., Thedgonia B. Sutton, Uwebraunia Crous et M.J. Wingf.,<br />

Xenostigmina Crous (Crous et al., 2000; 2001).<br />

Given the wide range <strong>of</strong> variation among anamorphs, it was not clear whether<br />

<strong>Mycosphaerella</strong> was monophyletic, paraphyletic or polyphyletic. Crous (1998)<br />

hypothesized that <strong>Mycosphaerella</strong> could be split into groups as indicated by the<br />

various anamorph genera. Based on the revision <strong>of</strong> the monograph on<br />

<strong>Mycosphaerella</strong> by A. Aptroot (<strong>CBS</strong>, The Netherlands), several sections are recognized<br />

(modified from Barr 1972):<br />

• Section <strong>Mycosphaerella</strong> is characterized by cylindrical asci and mostly<br />

uniseriate, thin-walled, <strong>of</strong>ten small ascospores that are constricted at the septum<br />

and inequilateral, with rounded upper ends. Anamorphs: typically Ramularia with<br />

Asteromella spermatial states. Representative species: the common polyphagous<br />

M. punctiformis (Pers. : Fr.) Starb.<br />

• Section Tassiana M.E. Barr is characterized by pyriform asci and irregularly<br />

arranged, thick-walled ascospores that are <strong>of</strong>ten large and constricted at the<br />

septum and nearly equilateral, relatively broad with rounded ends. Anamorph:<br />

typically Cladosporium. Representative species: the common polyphagous species<br />

M. tassiana (de Not.) Joh. (with large ascospores) and M. longissima (Fuck.) Lindau<br />

(with small ascospores). Further research is still required to determine whether<br />

the teleomorphs <strong>of</strong> Cladosporium subgen. Heterosporium (David, 1997) can also<br />

be accommodated in this section. Preliminary data suggest, however, that<br />

Cladosporium may fall outside <strong>Mycosphaerella</strong> (Crous et al., 2001, unpublished<br />

data).<br />

45


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

• Section Caterva M.E. Barr is characterized by cylindrical asci and irregularly<br />

arranged, thin-walled, <strong>of</strong>ten medium-sized ascospores that are rarely constricted<br />

at the septum and inequilateral, with more or less pointed ends. Asteroma and<br />

Asteromella spermatial forms are typical. Representative species: the common<br />

polyphagous M. subradians (Fr. : Fr.) Schroeter.<br />

• Section Longispora M.E. Barr is characterized by cylindrical asci with<br />

aggregated, thin-walled, long and slender ascospores that are rarely constricted<br />

at the septum and mostly equilateral, long but slender ascospores, characteristically<br />

with rounded upper and pointed lower ends. Anamorphs: Phloeospora or Septoria<br />

sensu lato. Representative species: M. millegrana (Cooke) Schroeter (with short<br />

spores), M. latebrosa (Cooke) Schroeter (with longer spores) and M. populi<br />

(Auersw.) J. Schröt. (with the longest spores in the genus). The genus Sphaerulina<br />

Sacc., which differs only by additional septa, may be synonymous.<br />

• Section Fusispora M.E. Barr is characterized by pyriform asci and irregularly<br />

arranged, thin-walled ascospores that are rarely constricted at the septum and<br />

mostly equilateral, fusiform, pointed ascospores. Anamorphs have not been proved.<br />

• Section Plaga M.E. Barr (including Section Macula M.E. Barr) includes endophytic<br />

species sporulating on <strong>leaf</strong> <strong>spot</strong>s, many <strong>of</strong> which are described as plant pathogens.<br />

This section is characterized by obovoid to ellipsoidal or cylindrical asci, small to<br />

medium sized ascopores, fusiform to obovoid with rounded ends. Many species have<br />

been described in these groups, and the majority originate from warm-temperate<br />

and tropical areas. Anamorphs include Colletogloeopsis, Mycovellosiella, Phaeophleospora,<br />

Pseudocercospora, Pseudocercosporella, Sonderhenia, Stenella Syd., Dissoconium,<br />

Uwebraunia and possibly others. Representative species: listed by Crous (1998)<br />

on Eucalyptus.<br />

How do we separate species in this complex based<br />

on morphology?<br />

<strong>Mycosphaerella</strong> encompasses species with a saprobic, plant pathogenic as well<br />

as a hyperparasitic habit. In general, however, most species are found to be<br />

associated with <strong>leaf</strong> <strong>spot</strong>s. Some species have been isolated as endophytes (Crous,<br />

1998) but this is not the norm. Importantly, up to four taxa have been reported<br />

from the same lesion on leaves <strong>of</strong> Eucalyptus (Crous, 1998) suggesting that some<br />

act as primary and others as secondary or weak pathogens.<br />

In the past, the taxonomy <strong>of</strong> <strong>Mycosphaerella</strong> relied mostly on aspects such as<br />

host, symptom type, and teleomorph morphology (pseudothecium, ascus and<br />

ascospore morphology). Very few studies focused on cultures, therefore ascospore<br />

germination patterns and cultural characteristics are unknown for most species.<br />

Furthermore, no ex-type cultures are available for molecular studies.<br />

Most links to anamorphs have also been based on association. Given<br />

the concept <strong>of</strong> sympatric colonisation <strong>of</strong> host tissue discussed above, this has led<br />

to numerous erroneous reports <strong>of</strong> anamorph/teleomorph links. Anamorph<br />

46


Session 1<br />

P.W. Crous et al.<br />

morphology has focused strongly on aspects such as conidiomatal structure, and<br />

mode <strong>of</strong> conidiogenesis (von Arx, 1983; Sutton and Hennebert, 1994). An<br />

important overlap has been observed between different conidiomatal types (Nag<br />

Raj, 1993; Braun 1995), hence species have been transferred from one anamorph<br />

genus to another (Sutton et al., 1996; Braun, 1998). The separation <strong>of</strong> some<br />

coelomycete and hyphomycete anamorphs <strong>of</strong> <strong>Mycosphaerella</strong> is, therefore,<br />

debatable.<br />

Subsequent to the wide taxonomic concept employed for the cercosporoid fungi<br />

by Chupp (1954), Deighton (1973, 1976, 1987, 1990) recognised the value <strong>of</strong><br />

pigmentation, conidiogenesis and conidium release. Different types <strong>of</strong> dehiscence<br />

scars (David, 1993) have subsequently been recognised to separate genera<br />

such as Cladosporium, Cercospora and Stenella. Conidiogenesis is variable in this<br />

complex (Verkley, 1997), as are dehiscence scars (Crous, 1998), but it still remains<br />

a useful feature to separate species (Braun, 1995).<br />

Integrating morphological and molecular data sets<br />

The issue <strong>of</strong> integrating morphological and molecular data sets in <strong>Mycosphaerella</strong><br />

is beset with numerous problems. Several hypotheses have been proposed to divide<br />

the genus into separate parts, groups, sections or genera. The molecular work to<br />

date, however, mainly supported one major clade <strong>of</strong> <strong>Mycosphaerella</strong> (Table 1), as<br />

well as two minor clades, typified by Dissoconium and Cladosporium (Crous et<br />

al., 2000, 2001). That anamorph concepts have not always correlated with different<br />

groups in <strong>Mycosphaerella</strong>, but have been shown to have evolved more than once<br />

within the genus, has caused confusion. Sometimes, however, integration has<br />

simplified or reduced the genera. For example, species <strong>of</strong> Mycovellosiella<br />

(superficial mycelium; conidia in chains), Phaeoramularia (no superficial<br />

mycelium, conidia in chains), and Passalora (no superficial mycelium, conidia<br />

solitary) consistently cluster together, leading to synonymy <strong>of</strong> the genera under<br />

the older name Passalora (Crous et al., 2001). Pseudocercospora is another<br />

confusing example. Published work (Crous et al., 2001) indicates that several other<br />

genera are also clustered here, including Pseudophaeoramularia U. Braun, and<br />

Stigmina, while the position <strong>of</strong> Stenella was still not clearly resolved. As more<br />

taxa have been added to our analyses (unpublished data), it has become clear that<br />

Stenella and Stigmina are in fact good genera, whereas Pseudophaeoramularia<br />

clusters in Pseudocercospora. These findings are still in agreement with those <strong>of</strong><br />

Crous et al. (2001), namely that conidial catenulation is not a good feature at the<br />

generic level, and that the degree <strong>of</strong> thickening <strong>of</strong> spore scars is important, rather<br />

than just presence or absence.<br />

Molecular data will also influence the way we define the teleomorph.<br />

Preliminary data suggest that ascospore septation, and the presence <strong>of</strong><br />

pseudoparenchymatal stromatic tissue around the ostiole is <strong>of</strong> lesser taxonomic<br />

value at the generic level (unpublished data). This will result in many older names<br />

having to be re-evaluated, and may even eventually require the conservation <strong>of</strong><br />

<strong>Mycosphaerella</strong>, to safely preserve the name for future use over older, but lesser<br />

known names.<br />

47


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Integrating morphology and molecular data sets on a species level is also beset<br />

with problems. Firstly, as discussed above, most species are known from dried<br />

herbarium material only and there are no reliable reference strains. Most strains<br />

that can be ordered from culture collections are sterile (a common phenomenon<br />

associated with species <strong>of</strong> <strong>Mycosphaerella</strong>) therefore their identities cannot be<br />

confirmed. Many <strong>of</strong> the species already sequenced in GenBank originate from such<br />

strains, and thus these data that are routinely used by plant pathologists obscure<br />

the issue even further, even for the few species presently known on Musa.<br />

Table 1. Fungal isolates included for ITS sequence analysis<br />

Accession no. Teleomorph Anamorph Origin<br />

U04234 Leptosphaeria microscopica Unknown ATCC 42652 (Saccharum<br />

<strong>of</strong>ficinarum, Kenya)<br />

PCR18 <strong>Mycosphaerella</strong> cruenta Pseudocercospora cruenta ATCC 262271 (Vigna,Puerto Rico)<br />

PP15 M. berkeleyi Passalora personata MPPD L2121 (Arachis,<br />

Oklahoma, U.S.A.)<br />

458 M. eumusae Pseudocercospora eumusae Musa, Malaysia<br />

487 M. eumusae Pseudocercospora eumusae Musa, Thailand<br />

PF7 M. fijiensis Paracercospora fijiensis ATCC 22116 (Musa, Philippines)<br />

PF8 M. fijiensis Paracercospora fijiensis ATCC 22117 (Musa, Hawaii)<br />

01A M. fijiensis Paracercospora fijiensis Musa, Philippines<br />

009 M. fijiensis Paracercospora fijiensis Musa, Ntoum, Gabon<br />

PFD9 M. fijiensis var. difformis Paracercospora fijiensis ATCC 36054 (Musa, Honduras)<br />

var. difformis<br />

PM10 M. musicola Pseudocercospora musae ATCC 22115 (Musa,Philippines)<br />

PM11 M. musicola Pseudocercospora musae ATCC 36143 (Musa, Honduras)<br />

121 M musicola Pseudocercospora musae Musa, Indonesia, West Java<br />

090 M musicola Pseudocercospora musae Musa, Armenia, Colombia<br />

CA1 <strong>Mycosphaerella</strong> Cercospora apii ATCC 12246<br />

state unknown<br />

CH5, CH6 <strong>Mycosphaerella</strong> Cercospora hayi ATCC 12234<br />

state unknown<br />

(Musa, Cuba)<br />

MA12 Unknown Mycocentrospora acerina ATCC 34539 (Daucus carota,<br />

Norway)<br />

IMI 271341 Unknown Phaeoseptoria musae Musa, Honduras<br />

Re-evaluating species occurring on Musa<br />

<strong>Mycosphaerella</strong> eumusae, M. fijiensis and M. musicola<br />

The <strong>Mycosphaerella</strong> state <strong>of</strong> M. eumusae is morphologically very similar to that<br />

<strong>of</strong> M. musicola. As expected, reports from literature also suggest that these two<br />

pathogens have commonly been confused in the past, thereby also questioning<br />

the value <strong>of</strong> much <strong>of</strong> the published literature (and distribution records) on this<br />

disease complex. Leaf <strong>spot</strong> symptoms attributed to M. eumusae were first seen<br />

after a survey conducted in Southeast Asia during 1992–1995 (Carlier et al.,<br />

2000b). The anamorph <strong>of</strong> M. eumusae was initially regarded as a species <strong>of</strong><br />

48


Session 1<br />

P.W. Crous et al.<br />

Septoria (Carlier et al., 2000a, b). Pseudocercospora eumusae, the anamorph <strong>of</strong><br />

M. eumusae, is characterized by having predominantly epiphyllous sporodochia<br />

that form on dark brown substomatal stromata. The sporodochia are mingled with<br />

developing spermatogonia. Young sporodochia are subepidermal and substomatal,<br />

and initially produce conidia that appear to be exuding from a<br />

subepidermal, substomatal pycnidium. In cross section, however, the subepidermal<br />

and substomatal structure is seen to be a sporodochium, not a pycnidium. As more<br />

stromatal tissue is formed, conidiophores become erumpent, and sporodochia burst<br />

through the epidermis, almost appearing acervular, but in fact being subepidermal<br />

sporodochia. Conidiophores are subhyaline to pale olivaceous, becoming pale<br />

brown at the base, subcylindrical, 0 to 3-septate, 10 to 25 x 3–5 µm, with<br />

conidiogenous cells terminating in truncate ends. Sporodochia <strong>of</strong> M. eumusae<br />

develop in a similar fashion to those <strong>of</strong> M. musicola but the conidiophores are<br />

much longer and more septate in the former. Conidia <strong>of</strong> P. eumusae are<br />

subhyaline to pale olivaceous, subcylindrical, (18–)30–50(–65) x (2–)2.5–3 µm,<br />

3 to 8-septate, and have subtruncate ends without visible scars. Conidia can be<br />

distinguished from those <strong>of</strong> M. musicola by their more cylindrical shape,<br />

subtruncate ends, and shorter dimensions (Crous and Mourichon, 2002). Based<br />

on ITS sequence data (Figure 1), the two species are also very closely related.<br />

Isolates <strong>of</strong> M. fijiensis (anamorph: P. fijiensis) are easily distinguished from P.<br />

musae and P. eumusae by their minutely thickened spore scars (Deighton, 1979).<br />

These scars have been shown to be <strong>of</strong> lesser phylogenetic importance in the<br />

cercosporoids (Stewart et al., 1999), and Paracercospora should be merged back<br />

into Pseudocercospora (Crous et al., 2000, 2001), but they are still valuable at the<br />

species level and should be used to separate taxa. Nevertheless, the anamorph <strong>of</strong><br />

M. fijiensis should now correctly be referred to as Pseudocercospora fijiensis.<br />

Other species <strong>of</strong> <strong>Mycosphaerella</strong><br />

As can be seen below, numerous additional species <strong>of</strong> <strong>Mycosphaerella</strong> (or their<br />

anamorphs) have been described from Musa. Little is known about many <strong>of</strong> these<br />

taxa, but they are expected to occur on lesions typically associated with the major<br />

pathogens discussed above. Some species, e.g. Cercospora hayi, appear to have<br />

a wider host range than just Musa. From general morphology, C. hayi resembles<br />

what is currently referred to as the Cercospora apii sensu lato morphotype. It is<br />

suspected that Cercospora apii is a weak or secondary pathogen with a wide host<br />

range, and has been described from numerous hosts worldwide. Based on<br />

morphology, C. hayi is indistinguishable from C. apii. From the ITS sequence data<br />

(Figure 1), the two species are clearly similar, and should be regarded as<br />

conspecific, with preference being given to the older name, C. apii (Braun et al.,<br />

unpublished data). Cladosporium musae, and several other species <strong>of</strong> Cladosporium<br />

have also been recorded from Musa. The generic affinities, and circumscription<br />

<strong>of</strong> genera within the Cladosporium-complex, however, remain to be resolved<br />

pending a full morphological and molecular study (Braun et al. unpublished data).<br />

<strong>Mycosphaerella</strong> musae is another interesting species and is commonly associated<br />

with <strong>Mycosphaerella</strong> speckle. From an analysis <strong>of</strong> published literature, and<br />

49


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

purported anamorphs associated with this species (Stover, 1994; Carlier et al.,<br />

2000a), it is clear that several different biological species have in the past been<br />

treated as representative <strong>of</strong> M. musae.<br />

Mycocentrospora acerina MA12<br />

100<br />

Phaeoseptoria musae IMI271341<br />

Leptosphaeria microscopica U04234<br />

M. musae 122<br />

<strong>Mycosphaerella</strong> musae<br />

100<br />

M. berkeleyi PP15<br />

C. hayi CH5<br />

63<br />

100<br />

61<br />

C. apii CA1<br />

Cercospora apii sensu lato<br />

C. hayi CH6<br />

69<br />

M. cruenta PCR18<br />

100<br />

Ps. musae PM11<br />

79<br />

99<br />

89<br />

88<br />

Ps. musae PM10<br />

M. musicola 121<br />

M. musicola 090<br />

<strong>Mycosphaerella</strong> musicola<br />

100<br />

51<br />

M. eumusae 458<br />

M. eumusae 487<br />

<strong>Mycosphaerella</strong> eumusae<br />

92<br />

M. fijiensis 01A<br />

10 changes<br />

54<br />

74<br />

M. fijiensis 009<br />

Pa. fijiensis PF8<br />

Pa. fijiensis var.<br />

difformisPFD9<br />

Pa. fijiensis PF7<br />

<strong>Mycosphaerella</strong> fijiensis<br />

Figure 1. One <strong>of</strong> eight most parsimonious trees (length = 697 steps, CI = 0.803, RI = 0.795, RC = 0.639). Bootstrap<br />

support from 1000 replicates is shown above the lines. Mycocentrospora acerina, Phaeoseptoria lysae and<br />

Leptosphaeria microscopica were used as outgroups.<br />

M. = <strong>Mycosphaerella</strong>,Ps.= Pseudocerospora and Pa. = Paracercospora.<br />

50


Session 1<br />

P.W. Crous et al.<br />

Major <strong>Mycosphaerella</strong> <strong>diseases</strong><br />

Black <strong>leaf</strong> streak disease<br />

<strong>Mycosphaerella</strong> fijiensis M. Morelet, Ann. Soc. Sci. Nat. Archéol. Toulon Var. 21:105.<br />

1969.<br />

= <strong>Mycosphaerella</strong> fijiensis var. difformis J.L. Mulder & R.H. Stover, Trans. Brit. Mycol.<br />

Soc. 67:82. 1976.<br />

Anamorph: Pseudocercospora fijiensis (M. Morelet) Deighton, Mycol. Pap. 140:144.<br />

1976.<br />

≡ Cercospora fijiensis M. Morelet, Ann. Soc. Sci. Nat. Archéol. Toulon Var. 21:105.<br />

1969.<br />

≡ Paracercospora fijiensis (M. Morelet) Deighton, Mycol. Pap. 144:51. 1979.<br />

= Cercospora fijiensis var. difformis J.L. Mulder & R.H. Stover, Trans. Brit. Mycol.<br />

Soc. 67:82. 1976.<br />

≡ Paracercospora fijiensis var. difformis (J.L. Mulder & R.H. Stover) Deighton, Mycol.<br />

Pap. 144:52. 1979.<br />

Host(s) and Distribution: Musa acuminata, Musa spp.; American Samoa,<br />

Australia, Belize, Benin, Bhutan, Bolivia, Brazil, Burundi, Cameroon, Central African<br />

Republic, China, Colombia, Comoros, Congo, Cook Islands, Costa Rica, Côte d’Ivoire,<br />

Cuba, Dominican Republic, Ecuador, El Salvador, Fiji, French Polynesia, Gabon,<br />

Ghana, Guatemala, Guinea-Bissau, Guyana, Haiti, Honduras, Indonesia, Jamaica,<br />

Kenya, Malawi, Malaysia, Mayotte, Mexico, Micronesia, Netherlands Antilles, New<br />

Caledonia, Nicaragua, Niger, Nigeria, Niue, Norfolk Island, Northern Mariana<br />

Islands, Panama, Papua New Guinea, Peru, Philippines, Rwanda, Samoa, São Tomé<br />

and Principe, Singapore, Solomon Islands, Somalia, Tahiti, Taiwan, Tanzania<br />

(Pemba, Zanzibar), Thailand, Togo, Tonga, Uganda, USA (FL, HI), Vanuatu,<br />

Venezuela, Vietnam, Wallis and Futuna Islands, Zambia.<br />

Eumusae <strong>leaf</strong> <strong>spot</strong> disease<br />

<strong>Mycosphaerella</strong> eumusae Crous et X. Mourichon, Sydowia 54:36. 2002.<br />

Anamorph: Pseudocercospora eumusae Crous et X. Mourichon, Sydowia, 54:36. 2002.<br />

Leaf <strong>spot</strong>s amphigenous, initially visible as faint brown streaks, developing<br />

into oval or elliptical light brown lesions with grey centres and dark brown<br />

borders, coalescing to form large, brown, necrotic areas under favourable<br />

conditions. Grey <strong>spot</strong>s and patches are visible in necrotic areas, and lesions are<br />

surrounded by a chlorotic yellow zone. Pseudothecia amphigenous, predominantly<br />

hypophyllous, black, subepidermal, becoming slightly erumpent, globose, up to<br />

80 µm diam., apical ostiole 10–15 µm wide; wall consisting <strong>of</strong> 2–3 layers <strong>of</strong><br />

medium brown textura angularis. Asci aparaphysate, fasciculate, bitunicate,<br />

subsessile, obovoid, straight or slightly incurved, 8-spored, 30–50 x 9–15 µm.<br />

Ascospores tri- to multiseriate, overlapping, hyaline, guttulate, thick-walled,<br />

51


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

straight, obovoid with obtuse ends, widest in the middle <strong>of</strong> apical cell, medianly<br />

1-septate or basal cell slightly longer than apical cell, tapering towards both ends,<br />

but with a more prominent taper towards lower end, (11–)12–13(–16.5) x<br />

(3–)3.5–4(–4.5) µm. Spermogonia predominantly hypophyllous, subepidermal,<br />

substomatal, globose, dark brown, up to 75 µm diam. Spermatia hyaline, rodshaped,<br />

3–6 x 1–2 µm. Mycelium internal, pale brown, consisting <strong>of</strong> septate,<br />

branched, smooth hyphae, 1.5–2.5 µm wide. Caespituli sporodochial, subepidermal,<br />

substomatal, predominantly epiphyllous, grey, up to 100 µm wide. Conidiophores<br />

aggregated in dense fascicles arising from the upper cells <strong>of</strong> a brown stroma up<br />

to 70 µm wide; conidiophores subcylindrical, smooth, hyaline or pale brown below,<br />

0–3-septate, straight to geniculate-sinuous, unbranched or branched below,<br />

10–25 x 3–5 µm. Conidiogenous cells terminal, unbranched, hyaline, smooth,<br />

tapering to flat-tipped apical loci, proliferating sympodially, or 1–4 times<br />

percurrently near the apex, 10–20 x 3–4 µm; scars inconspicuous. Conidia solitary,<br />

subhyaline to pale olivaceous, thick-walled, smooth, subcylindrical, apex obtuse,<br />

base subtruncate, straight to variously curved, 3–8-septate, (18–)30–50(–65) x<br />

(2–)2.5–3 µm; hila inconspicuous. Cultural characteristics: colonies pale olivaceous<br />

grey (23””’d according to Rayner, 1970) to rosy vinaceous (7”d) (surface), and<br />

brown vinaceous (5”’m) (bottom), with even margins and moderate aerial<br />

mycelium, obtaining 10 mm diam. after 2 months at 25°C in the dark (Crous and<br />

Mourichon, 2002).<br />

Host(s) and Distribution: Musa spp.; Southern India, Sri Lanka, Thailand,<br />

Malaysia, Vietnam, Mauritius, Nigeria.<br />

Sigatoka disease<br />

<strong>Mycosphaerella</strong> musicola R. Leach ex J.L. Mulder, Trans. Brit. Mycol. Soc. 67:77.<br />

1976.<br />

≡ <strong>Mycosphaerella</strong> musicola R. Leach, Trop. Agric. 18:92. 1941. (nom. nud.).<br />

Anamorph: Pseudocercospora musae (Zimm.) Deighton, Mycol. Pap. 140:148. 1976.<br />

≡ Cercospora musae Zimm., Centralbl. Bakteriol. Parasitenk. 2. Abth. 8:219. 1902.<br />

= Cercospora musae Massee, Bull. Misc. Inform. 28:159. 1914.<br />

Host(s) and Distribution: Musa acuminata, M. banksii, M. basjoo, M. liukiuensis,<br />

M. paradisiaca, M. textiles; widely distributed, including American Samoa, Angola,<br />

Antigua and Barbuda, Argentina, Australia, Bahamas, Barbados, Belau, Belize, Bolivia,<br />

Brazil, Brunei Darussalam, Bhutan, Cambodia, Cameroon, Cape Verde, Cayman<br />

Islands, China, Colombia, Congo, Cook Islands, Costa Rica, Côte d’Ivoire, Cuba,<br />

Dominica, Dominican Republic, Ecuador, El Salvador, Ethiopia, Fiji, French Guiana,<br />

French Polynesia, Gabon, Ghana, Grenada, Guadeloupe, Guam, Guatemala, Guinea,<br />

Guinea-Bissau, Guyana, Haiti, Honduras, Hong Kong, India, Indonesia, Jamaica,<br />

Kenya, Kiribati, Laos, Madagascar, Malawi, Malaysia, Martinique, Mauritius, Mexico,<br />

Micronesia, Montserrat, Mozambique, Nepal, New Caledonia, Nicaragua, Nigeria,<br />

Niue, Norfolk Island, Panama, Papua New Guinea, Peru, Philippines, Puerto Rico,<br />

Saint Kitts and Nevis, Saint Lucia, Saint Vincent and the Grenadines, Samoa,<br />

52


Session 1<br />

P.W. Crous et al.<br />

Sao Tome and Principe, Sierra Leone, Solomon Islands, Somalia, South Africa,<br />

Sri Lanka, Surinam, Taiwan, Tanzania, Thailand, Togo, Tonga, Trinidad and Tobago,<br />

Tuvalu, Uganda, USA (FL, HI), Vanuatu, Venezuela, Vietnam, Wallis and Futuna<br />

Islands, Yemen, Zambia, Zimbabwe.<br />

Other <strong>diseases</strong> caused by species <strong>of</strong> <strong>Mycosphaerella</strong><br />

and its anamorphs<br />

Cercospora hayi Calp., Studies on the Sigatoka Disease <strong>of</strong> Bananas and its Fungus<br />

Pathogen, Atkins Garden and Research Laboratory, Cuba, p. 63. 1955.<br />

Host(s) and Distribution: Musa paradisiaca, Musa sp.; Brazil, Cuba, Philippines.<br />

Notes: Part <strong>of</strong> the C. apii sensu lato species complex.<br />

Cercospora musae var. paradisiaca Bat. et R. Garnier, Revista Agric. (Recife), 3:54.<br />

1963.<br />

Host(s) and Distribution: Musa paradisiaca; Brazil.<br />

Notes: Status unknown, has not been treated.<br />

Cercospora pingtungensis T.Y. Lin et J.M. Yen, Bull. Soc. Mycol. France 87:427. (1971)<br />

1972.<br />

Host(s) and Distribution: Musa acuminata, M. cavendishii; China, Taiwan.<br />

Notes: Conidia pigmented with thickened hila, not a Cercospora.<br />

Cladosporium bisporum, Matsush., Icones micr<strong>of</strong>ungorum a Matsushima lectorum<br />

(Kobe):33. 1975.<br />

Host(s) and Distribution: Musa paradisiaca; Japan.<br />

Cladosporium <strong>leaf</strong> speckle<br />

Cladosporium musae E.W. Mason, Mycol. Pap. 13:2. 1945.<br />

Host(s) and Distribution: Musa paradisiaca, M. textiles, Musa spp.; Bangladesh,<br />

Brunei, Burundi, Cameroon, Congo, Côte d’Ivoire, Cuba, Ecuador, Egypt, Ethiopia,<br />

France, Ghana, Ghana, Guinea, Guinea, Honduras, Honduras, Hong Kong, Indonesia,<br />

Jamaica, Malaysia, Nepal, Papua New Guinea, Rwanda, Sierra Leone, Solomon Islands,<br />

South Africa, Sri Lanka, Sudan, Uganda, Thailand, Togo, Uganda, Vietnam, Western<br />

Samoa, Zimbabwe.<br />

Cladosporium oxysporum Berk. et M.A. Curtis, J. Linn. Soc. Bot. 10:362. 1868.<br />

Host(s) and Distribution: Musa paradisiaca; Brazil, Venezuela.<br />

<strong>Mycosphaerella</strong> formosana T.Y. Lin et J.M. Yen, Rev. Mycol. 35:323. 1971.<br />

Host(s) and Distribution: Musa sp.; Taiwan.<br />

<strong>Mycosphaerella</strong> henriquesiana G. Winter, Bol. Soc. Brot. 4:13. 1886.<br />

Host(s) and Distribution: Musa sp.; Africa.<br />

53


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

<strong>Mycosphaerella</strong> liukiuensis Sawada, Special Publ. Coll. Agric. Natl. Taiwan Univ.<br />

63. 1959.<br />

Host(s) and Distribution: Musa formosana (?), M. liukiuensis; Taiwan.<br />

<strong>Mycosphaerella</strong> speckle<br />

<strong>Mycosphaerella</strong> musae (Speg.) Syd. et P. Syd., Phillipp. J. Sci. 8:482. 1913.<br />

≡ Sphaerella musae Speg., Anal. Mus. Nac. Hist. Nat. Buenos Aires 19:354. 1909.<br />

= Sphaerella musae Sacc., Atti Accad. Sci. Veneto-Trentino-Instriana, Ser. 3, 10:67.<br />

1917, homonym.<br />

Host(s) and Distribution: Musa acuminata, M. banksii, M. cavendishii, M. coccinea,<br />

M. paradisiaca, M. paradisiaca, M. textilis, M. uranoscopos, Musa sp.; Argentina,<br />

Australia, Malaysia, Philippines, Puerto Rico, USA (HI), Virgin Islands.<br />

Pseudocercospora fengshanensis (T.Y. Lin & J.M. Yen) J.M. Yen & S.K. Sun,<br />

Cryptogam. Mycol. 4:197. 1983.<br />

≡ Cercospora fengshanensis T.Y. Lin & J.M. Yen, Rev. Mycol. 35:320. (1970) 1971.<br />

Host(s) and Distribution: Musa acuminata; China, Taiwan.<br />

Pseudocercospora musae-sapienti (A.K. Kar & M. Mandal) U. Braun & Mouch.,<br />

New Zealand J. Bot. 37:317. 1999.<br />

≡ Cercospora musae-sapienti A.K. Kar & M. Mandal, Norweg. J. Bot. 22:105. 1975.<br />

Host(s) and Distribution: Musa paradisiaca; India, Wallis.<br />

Pseudocercospora musicola U. Braun, New Zealand J. Bot. 37:317. 1999.<br />

≡ Cercospora musicola Sawada (musaecola), Rep. Gov. Agric. Res. Inst. Taiwan 85:116.<br />

1943 (nom. inval.).<br />

≡ Cercospora musicola Goh & W.H. Hsieh, Cercospora and similar fungi from Taiwan<br />

(1990, p. 242). (nom. inval.).<br />

= Cercospora musae-liukiuensis Sawada, Special Publ. Coll. Agric. Natl. Taiwan Univ.<br />

8:221. 1959. (nom. nud.).<br />

Host(s) and Distribution: Musa acuminata, M. basjoo, M. liukiuensis,<br />

M. paradisiaca; China, Japan, Taiwan.<br />

Excluded from <strong>Mycosphaerella</strong><br />

Deightoniella <strong>leaf</strong> <strong>spot</strong><br />

Teleomorph state unknown (presumed not <strong>Mycosphaerella</strong>)<br />

Anamorph: Deightoniella torulosa (Syd.) M.B. Ellis, Mycol. Pap. 66:7 (1957).<br />

≡ Brachysporium torulosum Syd., Hedwigia 49: 83 (1909).<br />

= Cercospora musarum S.F. Ashby, Bull. Dept. Agric. Jamaica N.S. 2:109. 1913.<br />

Host(s) and Distribution: Musa paradisiaca, Musa spp.; Australia, Egypt, Jamaica,<br />

Malaysia, Sierra Leone, Somalia, Taiwan, Thailand, Vietnam.<br />

54


Session 1<br />

P.W. Crous et al.<br />

Current and future research prospects<br />

1. Given the obvious problems surrounding <strong>Mycosphaerella</strong> research discussed<br />

above, it is imperative that all researchers agree that they are working with the<br />

same disease. This needs to be firmly established by means <strong>of</strong> molecular techniques.<br />

Primers have been developed to identify M. musicola and M. fijiensis (Johanson<br />

et al., 1994) but little is known about similar, closely related species, and whether<br />

they could be separated using these primers which are based on ITS sequence data.<br />

Although ITS has thus far proved to be very valuable in <strong>Mycosphaerella</strong> systematics,<br />

additional genes also need to be sequenced, as ITS alone indicates similarity, not<br />

necessarily conspecificity.<br />

2. The data presently available in GenBank for species occurring on Musa indicate<br />

some variation within well-known taxa. This could either be due to sequencing<br />

errors, intraspecific variation, or to different researchers working with different<br />

species. To standardize the taxonomy and pathology research being conducted<br />

on these organisms, we need to agree on what they are. This can only be achieved<br />

by designating epitypes <strong>of</strong> the various species following a thorough taxonomic<br />

study. These cultures should then be lodged in culture collections and be readily<br />

available to those wishing to study the organisms occurring on Musa. In all these<br />

instances, it is best for mycologists to derive the cultures from fresh specimens,<br />

so that the identity can be confirmed on the host material, and once again in<br />

culture.<br />

3. To fully understand species within <strong>Mycosphaerella</strong>, we need to collect both<br />

species and populations. We need to address questions relating to host specificity,<br />

speciation and, in respect to plant pathology, epidemiology, fungicide sensitivity,<br />

the importance <strong>of</strong> the different morphs and mechanisms <strong>of</strong> variation and<br />

dispersal. We also need to learn how the species are migrating around the world.<br />

To address these questions, we need to develop the correct molecular markers that<br />

can be used to for investigations at a species and a population level. Once again,<br />

these populations need to be deposited in reputable culture collections (i.e. <strong>CBS</strong>,<br />

IMI or ATCC) so that they can be studied and re-studied in years to come.<br />

References<br />

Arx J.A. von 1949. Beitrage zur Kenntnis der Gattung <strong>Mycosphaerella</strong>. Sydowia 3:28–100.<br />

Arx J.A. von 1983. <strong>Mycosphaerella</strong> and its anamorphs. Proc. K. Nederl. Akad. Wet. Ser. C<br />

86:15–54.<br />

Barr M.E. 1972. Preliminary studies on the Dothideales in temperate North America. Contrib.<br />

Univ. Michigan Herb. 9:523–638.<br />

Braun U. 1995. A monograph <strong>of</strong> Cercosporella, Ramularia and allied genera (Phytopathogenic<br />

Hyphomycetes). Vol. 1. IHW-Verlag, Eching, Germany.<br />

Braun U. 1998. A monograph <strong>of</strong> Cercosporella, Ramularia and allied genera (Phytopathogenic<br />

Hyphomycetes). Vol. 2. IHW-Verlag, Eching, Germany.<br />

55


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Carlier J., E. Fouré, F. Gauhl, D.R. Jones, P. Lepoivre, X. Mourichon, Pasberg-Gauhl C. and<br />

R.A. Romero. 2000a. Fungal <strong>diseases</strong> <strong>of</strong> the foliage. Pp. 37–141 in Diseases <strong>of</strong> Banana,<br />

Abacá and enset. (D.R. Jones, ed.). CAB International, Wallingford, UK.<br />

Carlier J., M.F. Zapater, F. Lapeyre, D.R. Jones and X. Mourichon. 2000b. Septoria <strong>leaf</strong> <strong>spot</strong><br />

<strong>of</strong> banana: a newly discovered disease caused by <strong>Mycosphaerella</strong> eumusae (anamorph<br />

Septoria eumusae). Phytopathology 90:884-890.<br />

Chupp C. 1954. A monograph <strong>of</strong> the fungus genus Cercospora. Ithaca, New York. Published<br />

by the author.<br />

Corlett M. 1991. An annotated list <strong>of</strong> the published names in <strong>Mycosphaerella</strong> and Sphaerella.<br />

Mycol. Mem. 18:1–328.<br />

Corlett M. 1995. An annotated list <strong>of</strong> the published names in <strong>Mycosphaerella</strong> and Sphaerella:<br />

corrections and additions. Mycotaxon 53:37–56.<br />

Crous P.W. 1998. <strong>Mycosphaerella</strong> spp. and their anamorphs associated with <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

<strong>of</strong> Eucalyptus. Mycol. Mem. 21:1–170.<br />

Crous P.W., A. Aptroot, J.C. Kang, U. Braun and M.J. Wingfield. 2000. The genus<br />

<strong>Mycosphaerella</strong> and its anamorphs. in Molecules, morphology and classification: towards<br />

monophyletic genera in the Ascomycetes. (K.A. Seifert, W. Gams, P.W. Crous and G.J.<br />

Samuels, eds). Stud. Mycol. 45:107-121.<br />

Crous P.W., J.C. Kang, and U. Braun. 2001. A phylogenetic redefinition <strong>of</strong> anamorph genera<br />

in <strong>Mycosphaerella</strong> based on ITS rDNA sequence and morphology. Mycologia 93:1081–1101.<br />

Crous P.W. and X. Mourichon. 2002. <strong>Mycosphaerella</strong> eumusae and its anamorph<br />

Pseudocercospora eumusae spp. nov.: causal agent <strong>of</strong> eumusae <strong>leaf</strong> <strong>spot</strong> disease <strong>of</strong> banana.<br />

Sydowia 54:35–43.<br />

David J.C. 1993. A revision <strong>of</strong> taxa referred to Heterosporium Klotzsch ex Cooke (Mitosporic<br />

Fungi). PhD Dissertation, University <strong>of</strong> Reading, UK.<br />

David J.C. 1997. A contribution to the systematics <strong>of</strong> Cladosporium: revision <strong>of</strong> fungi<br />

previously referred to Heterosporium. Mycol. Pap. 172:1–157.<br />

Deighton F.C. 1973. Studies on Cercospora and allied genera. IV. Cercosporella Sacc.,<br />

Pseudocercosporella gen. nov. and Pseudocercosporidium gen. nov. Mycol. Pap. 133:1–62.<br />

Deighton F.C. 1976. Studies on Cercospora and allied genera. VI. Pseudocercospora Speg.,<br />

Pantospora Cif. and Cercoseptoria Petr. Mycol. Pap. 140:1–168.<br />

Deighton F.C. 1979. Studies on Cercospora and allied genera. VII. New species and<br />

redispositions. Mycol. Pap. 144:1–56.<br />

Deighton F.C. 1987. New species <strong>of</strong> Pseudocercospora and Mycovellosiella, and new<br />

combinations into Pseudocercospora and Phaeoramularia. Trans. Brit. Mycol. Soc.<br />

88:365–391.<br />

Deighton F.C. 1990. Observations on Phaeoisariopsis. Mycol. Res. 94:1096-1102.<br />

Johanson A., R.N. Crowhurst, E.H.A. Rikkerink, R.A. Fullerton and M.D. Templeton. 1994.<br />

The use <strong>of</strong> species-specific DNA probes for the identification <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

and M. musicola, the causal agents <strong>of</strong> Sigatoka disease <strong>of</strong> banana. Plant Pathol.<br />

44:701–707.<br />

Nag Raj T.R. 1993. Coelomycetous anamorphs with appendage-bearing conidia. Mycologue<br />

Publications, Waterloo, ON, Canada.<br />

Park R.F. and P.J. Keane. 1984. Further <strong>Mycosphaerella</strong> species causing <strong>leaf</strong> <strong>diseases</strong> <strong>of</strong><br />

Eucalyptus. Trans. Brit. Mycol. Soc. 83:93–105.<br />

Pollack F.G. 1987. An annotated compilation <strong>of</strong> Cercospora names. Mycol. Mem. 12:1–212.<br />

Stewart E.L., Z. Liu, P.W. Crous and L. Szabo. 1999. Phylogenetic relationships among some<br />

cercosporoid anamorphs <strong>of</strong> <strong>Mycosphaerella</strong> based on rDNA sequence analysis. Mycol. Res.<br />

103:1491–1499.<br />

56


Session 1<br />

P.W. Crous et al.<br />

Stover R.H. 1994. <strong>Mycosphaerella</strong> musae and Cercospora ‘non-virulentum’ from Sigatoka <strong>leaf</strong><br />

<strong>spot</strong>s are identical. Fruits 49:187–190.<br />

Sutton B.C. 1980. The Coelomycetes. Commonwealth Mycological Institute, Kew, England.<br />

Sutton B.C. and G.L. Hennebert. 1994. Interconnections amongst anamorphs and their possible<br />

contribution to Ascomycete systematics. Pp. 77–98. in Ascomycete Systematics. Problems<br />

and perspectives in the nineties. (D.L. Hawksworth, ed.). Plenum Press, New York.<br />

Sutton B.C., S.F. Shamoun and P.W. Crous. 1996. Two <strong>leaf</strong> pathogens <strong>of</strong> Ribes spp. in North<br />

America, Quasiphloeospora saximontanensis (Deighton) comb. nov. and Phloeosporella<br />

ribis (J.J. Davis) comb. nov. Mycol. Res. 100:979–983.<br />

Verkley G.J.M. 1997. Ultrastructural evidence for two types <strong>of</strong> proliferation in a single<br />

conidiogenous cell <strong>of</strong> Septoria chrysanthemella. Mycol. Res. 102:368-372.<br />

57


Session 1<br />

J. Henderson et al.<br />

Improved PCR-based detection <strong>of</strong><br />

Sigatoka disease and black <strong>leaf</strong> streak<br />

disease in Australian banana crops<br />

J. Henderson 1 ,K.Grice 2 ,J.Pattemore 1 ,R.Peterson 2 and E. Aitken 1<br />

Abstract<br />

Accurate diagnosis <strong>of</strong> black <strong>leaf</strong> streak disease is <strong>of</strong>ten complicated by the presence <strong>of</strong> other fungal<br />

pathogens and in particular by the morphological similarity <strong>of</strong> the related species <strong>Mycosphaerella</strong><br />

musicola, the causal agent <strong>of</strong> Sigatoka disease. In addition, high rainfall <strong>of</strong>ten washes away fungal<br />

structures making microscopic identification difficult. Starting in 1998, the Queensland<br />

Department <strong>of</strong> Primary Industries has been using molecular methods to help diagnose black <strong>leaf</strong><br />

streak disease. A polymerase chain reaction (PCR) assay was used, but the protocol was found<br />

to lack specificity when applied to Australian isolates <strong>of</strong> the fungi. In July 2000, a project aimed<br />

at improving the sensitivity and specificity <strong>of</strong> the PCR as well as streamlining the assay was<br />

initiated. Various components <strong>of</strong> the PCR test were targeted for improvement. Homogenization<br />

<strong>of</strong> banana <strong>leaf</strong> tissue has eliminated possible cross-contamination while tripling batch<br />

throughput. An improved DNA extraction method produces cleaner DNA in less than half the<br />

time <strong>of</strong> the prior extraction method. Flexibility and sensitivity <strong>of</strong> the PCR has been improved by<br />

introducing a new enzyme while the new format PCR thermal cyclers have increased sample<br />

throughput. Importantly, specificity has been enhanced with the design <strong>of</strong> new diagnostic primers.<br />

These changes produce a definitive result during the first PCR in more than 98% <strong>of</strong> samples while<br />

increasing daily throughput more than eight-fold.<br />

Resumen - Mejoramiento de la detección de la Sigatoka negra y Sigatoka amarilla<br />

basada en PCR en los cultivos bananeros de Australia<br />

A menudo el diagnóstico preciso de la Sigatoka negra es complicado debido a la presencia de<br />

otros patógenos fungosos y en particular por la similitud morfológica de la especie relacionada<br />

<strong>Mycosphaerella</strong> musicola, agente causal de la Sigatoka amarilla. En adición, fuertes precipitaciones<br />

a menudo se llevan las estructuras fungosas dificultando la identificación microscópica. El QDPI<br />

ha estado utilizando métodos moleculares para confirmar el diagnóstico de la Sigatoka negra<br />

desde 1998. Se utilizó un ensayo de la reacción en cadena de polimerasa (PCR), sin embargo se<br />

descubrió que al protocolo le faltaba la especificidad al aplicarlo a los aislados australianos de<br />

los hongos. En julio de 2000, se inició un proyecto dirigido a mejorar la sensibilidad y especificidad<br />

del PCR así como la modernización del ensayo. Se seleccionaron varios componentes del examen<br />

1<br />

Cooperative Research Centre for Tropical Plant Protection, Indooroopilly, Australia<br />

2<br />

Queensland Department <strong>of</strong> Primary Industries, Mareeba, Australia<br />

59


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

PCR para ser mejorados. La homogenización del tejido foliar del banano con la ayuda de una<br />

micromano de mortero plástica dentro del tubo ha eliminado una posible contaminación<br />

cruzada al triplicar el rendimiento de los lotes. Un método mejorado de extracción de ADN produce<br />

un ADN más limpio en menos de la mitad del tiempo, que el método de extracción anterior. La<br />

flexibilidad y sensibilidad de PCR fueron mejoradas introduciendo una nueva enzima, mientras<br />

que nuevos variadores térmicos para el formateo de PCR han aumentado el rendimiento de las<br />

muestras. La especificidad de PCR ha sido mejorada a través del diseño de nuevos iniciadores de<br />

diagnóstico. Combinadas, estas mejoras producen un resultado definitivo durante el primer ensayo<br />

de PCR en más del 98% de las muestras, mientras que el rendimiento diario de la muestra es<br />

8 veces mayor.<br />

Résumé - Détection améliorée basée sur la PCR de la maladie de Sigatoka et de la<br />

maladie des raies noires dans les plantations de bananes en Australie<br />

Le diagnostic exact de la maladie des raies noires est souvent rendu plus difficile par la présence<br />

d’autres pathogènes fongiques, et en particulier par la similarité morphologique d’une espèce<br />

voisine <strong>Mycosphaerella</strong> musicola, l’agent causal de la maladie de Sigatoka. De plus, de fortes<br />

précipitations éliminent souvent des structures fongiques ce qui rend l’identification au<br />

microscope difficile. En 1998, le Queensland Department <strong>of</strong> Primary Industries (QDPI) a commencé<br />

à utiliser des méthodes moléculaires afin de mieux identifier la maladie des raies noires. Un essai<br />

basé sur la réaction en chaîne par polymérase (PCR, polymerase chain reaction) a été utilisé mais<br />

ce protocole était peu spécifique quant aux isolats australiens du champignon. En juillet 2000,<br />

une étude a été initiée afin d’augmenter la sensibilité et la spécificité de la PCR et pour<br />

rationaliser le protocole. Divers composants du test PCR ont été ciblés afin d’être améliorés.<br />

L’homogénéisation des tissus de la feuille de banane a permis d’éliminer les contaminations<br />

extérieures tout en triplant le débit. Une méthode améliorée d’extraction de l’ADN permet<br />

d’obtenir un ADN plus pur en deux fois moins de temps. La flexibilité et la sensibilité de la PCR<br />

ont été améliorées grâce à l’utilisation d’une nouvelle enzyme. De plus, le nouveau format des<br />

thermocycleurs PCR a permis d’accroître le débit. Il est important de noter que la spécificité a<br />

été mise en valeur par la conception de nouvelles amorces diagnostiques. Ces changements<br />

produisent un résultat définitif dans la première PCR dans plus de 98% des cas et multiplient<br />

par huit le nombre d’échantillons traités par jour.<br />

The Tully 2001 black <strong>leaf</strong> streak disease outbreak<br />

The value <strong>of</strong> the Australian banana industry is estimated to be A$357 million (US$193<br />

million) per year. In 2000, nearly 250 000 tonnes <strong>of</strong> <strong>bananas</strong> were produced by 100<br />

growers in Australia. All Australian <strong>bananas</strong> are produced for consumption locally<br />

and 85% are <strong>of</strong> the ‘Cavendish’ variety. The majority <strong>of</strong> <strong>bananas</strong> are grown in<br />

northern Queensland, with 67% <strong>of</strong> the crop concentrated in Tully, Cairns and Innisfail<br />

(Figure 1).<br />

In April 2001, the Australian banana industry suffered a potentially devastating<br />

outbreak <strong>of</strong> black <strong>leaf</strong> streak disease (caused by <strong>Mycosphaerella</strong> fijiensis) in Tully<br />

(Figure 1). This is the pathogen’s first incursion in a major commercial region in<br />

Australia; failure to control the pathogen would have far reaching effects on the<br />

industry.<br />

The Queensland Department <strong>of</strong> Primary Industries (QDPI) has had considerable<br />

success eradicating previous outbreaks <strong>of</strong> black <strong>leaf</strong> streak disease by plant<br />

destruction and replacement with resistant varieties. Since the initial discovery <strong>of</strong><br />

black <strong>leaf</strong> streak disease in 1981 at Bamaga, an Aboriginal community located 40km<br />

60


Session 1<br />

J. Henderson et al.<br />

from the tip <strong>of</strong> the Cape York Peninsula, black <strong>leaf</strong> streak disease has been detected<br />

and eradicated eight times in far north Queensland. This ninth outbreak was in Tully<br />

where crops are estimated to be worth A$119 million per year (US$64 million).<br />

Figure 1. Map <strong>of</strong> Queensland (Australia)<br />

and inset showing location <strong>of</strong> major<br />

commercial banana growing region<br />

and Tully, the site <strong>of</strong> the 2001 black <strong>leaf</strong><br />

streak disease outbreak.<br />

Diagnosis <strong>of</strong> black <strong>leaf</strong> streak disease in Australia<br />

Banana crops are routinely surveyed for black <strong>leaf</strong> streak disease by QDPI scientists<br />

at the Centre for Tropical Agriculture, Mareeba. Accurate diagnosis <strong>of</strong> black <strong>leaf</strong> streak<br />

disease is complicated by the morphological similarity <strong>of</strong> M. fijiensis to a related species<br />

M. musicola, which causes Sigatoka disease. Usually, experienced plant pathologists<br />

distinguish the two <strong>diseases</strong> by the development <strong>of</strong> symptoms and microscopical<br />

features <strong>of</strong> the fungi. In Tully, conidia were absent because <strong>of</strong> prolonged rainfall, and<br />

identification <strong>of</strong> morphological characters was not possible. Therefore, molecular<br />

methods were used for diagnosis.<br />

The QDPI has used the polymerase chain reaction (PCR) to confirm diagnoses <strong>of</strong><br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> since 1998 (Johanson, 1997). Approximately 10%<br />

<strong>of</strong> laboratory samples required confirmation by PCR. However, the method was slow<br />

and lacked specificity to some Australian and Torres Strait Island isolates <strong>of</strong> the fungi.<br />

The lack <strong>of</strong> specificity was possibly due to the high variability among the Southeast<br />

Asian populations <strong>of</strong> the pathogen. Populations <strong>of</strong> M. fijiensis from the Torres Strait<br />

were found to differ from those <strong>of</strong> the Pacific (Hayden, 2001). However, the Torres<br />

Strait populations and Pacific populations were found to be related to those from Papua<br />

New Guinea where there is a considerable diversity (Carlier, pers. comm.). In the original<br />

61


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

study by Johanson et al. (1994) isolates from the Torres Strait were not included and<br />

it is possible that these isolates could be the source <strong>of</strong> the variability not detected by<br />

the original PCR primers.<br />

In July 2000, a project between QDPI and the Cooperative Research Centre for<br />

Tropical Plant Protection (CRCTPP) was initiated with the aim <strong>of</strong> improving the<br />

specificity <strong>of</strong> the diagnostic procedure and increasing throughput in readiness for<br />

outbreaks <strong>of</strong> the disease 1 . Aspects <strong>of</strong> the PCR diagnostic procedure that were targeted<br />

for improvement included sample excision, homogenisation <strong>of</strong> banana <strong>leaf</strong> tissue, DNA<br />

extraction, PCR protocol, PCR primer design and equipment.<br />

Flame sterilised cork-borers 4 mm in diameter have replaced scalpels for the removal<br />

<strong>of</strong> suspect lesions from banana leaves. The method is quick and simple and there is<br />

no cross-contamination between samples. Plastic micropestles have replaced ceramic<br />

mortar and pestles for homogenising <strong>leaf</strong> tissue. Micropestles have reduced the potential<br />

for cross-contamination between samples, eliminated transfer from mortar to tube, and<br />

have tripled the throughput. The rapid cetyltrimethyl ammonium bromide (CTAB) DNA<br />

extraction method (Stewart and Via, 1993) was adopted. It produces cleaner DNA in<br />

half the time.<br />

New PCR primer sequences specific to M. musicola and M. fijiensis, and a<br />

modification <strong>of</strong> published ribosomal gene primer sequences (White et al., 1990) to<br />

improve specificity to increase duplex stability <strong>of</strong> the primers with the target DNA<br />

(Rychlik, 1993), has improved specificity <strong>of</strong> the PCR assay. A size difference was also<br />

included in the PCR assay with the specific primer for M. fijiensis designed to the ITS1<br />

and the specific primer for M. musicola designed to the ITS2 (Figure 2). Flexibility<br />

and sensitivity was improved by introducing the hot-start enzyme, TaqGold DNA<br />

polymerase (PE Biosystems). TaqGold DNA polymerase requires heat-activation before<br />

amplification can proceed. Therefore, non-specific amplification products are reduced<br />

and reactions can be left at 4°C until the addition <strong>of</strong> template. New equipment at the<br />

Centre for Tropical Agriculture has also improved throughput <strong>of</strong> assays. New format<br />

PCR thermal cyclers have increased tube capacity from 30 to 192 per run, and new<br />

electrophoresis equipment can analyse 52 samples for Sigatoka disease and black <strong>leaf</strong><br />

streak disease at the same time.<br />

The new methods produce a high quality DNA preparation and provide a definitive<br />

result during the first PCR in more than 98% samples. In addition, extraction time is<br />

more than halved and daily throughput increased by more than eight-fold.<br />

Application <strong>of</strong> new molecular test<br />

Use <strong>of</strong> the new methods in April 2001 coincided with Australia’s most severe<br />

outbreak <strong>of</strong> black <strong>leaf</strong> streak disease. This was the first outbreak in a commercial<br />

growing area; previous outbreaks had been further north and in places where<br />

containment was easy. Fungal structures were absent on banana samples because<br />

<strong>of</strong> high rainfall at Tully. Therefore, diagnosis <strong>of</strong> up to 50% <strong>of</strong> samples was confirmed<br />

using the PCR assay. The PCR assay provided the Australian government and<br />

1<br />

The project is co-managed by Ron Peterson (Principal Plant Pathologist, Mareeba QDPI) and Juliane Henderson (Research<br />

Officer, CRCTPP).<br />

62


Session 1<br />

J. Henderson et al.<br />

Sigatoka disease<br />

MMfor<br />

R635<br />

ITS1<br />

ITS2<br />

1 8S (partial) 5.8S 25S (partial)<br />

MFfor<br />

R635mod<br />

Black <strong>leaf</strong> streak disease<br />

Figure 2. Location <strong>of</strong> primers for the diagnostic PCR assay.The black <strong>leaf</strong> streak disease specific forward primer<br />

(MFfor) is located on the ITS1 region while the Sigatoka disease specific forward primer (Mmfor) is located<br />

on the ITS2 region.<br />

banana industry the confidence to start a A$20 million (US$10.8 million)<br />

surveillance and eradication plan in Tully, and more than 2500 PCR tests have<br />

been done.<br />

The CRCTPP and QDPI continue to monitor and improve the PCR diagnostic.<br />

Thus, scientists from the CRCTPP improved homogenization <strong>of</strong> banana tissue by<br />

the use <strong>of</strong> glass beads shaken at high speeds. The commercially available<br />

“FastPrep” Instrument (Q-Biogene) processes 12 samples in 45 seconds and<br />

eliminates cross-contamination between samples by single-use O-ring tubes. Use<br />

<strong>of</strong> the method at the Centre for Tropical Agriculture is dependant on funds.<br />

Opportunities to automate and improve specificity and sensitivity <strong>of</strong> the assay<br />

are being studied as part <strong>of</strong> the CRCTPP’s plan to use new technology. Development<br />

<strong>of</strong> a real-time PCR assay to detect and differentiate M. fijiensis, M. musicola and<br />

M. eumusae is in progress. A fluorescent PCR format increases sensitivity and<br />

specificity, reduces cross-contamination, and increases throughput because post-<br />

PCR processing is not required.<br />

To ensure the robustness <strong>of</strong> the PCR diagnostic and to facilitate development<br />

<strong>of</strong> new diagnostic assays for <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in Australia, the<br />

sequence variability in Australasian isolates <strong>of</strong> M. musicola and M. fijiensis will<br />

be studied 2 . First, the region incorporating the ITS1, 5.8S ribosomal gene and ITS2<br />

will be cloned and sequenced from <strong>Mycosphaerella</strong> isolates from Australia, the<br />

Torres Strait Islands and Fiji. In collaboration with other groups studying sequences<br />

pertaining to the disease, we will compare our database with overseas isolates.<br />

If further sequence information is required, other conserved fungal genes, e.g.<br />

ß-tubulin, histone-4, glyceraldehyde-3-phosphate, will be investigated. The<br />

information from this study will help us to understand how the 2001 Australian<br />

outbreak arose, e.g. whether from one or several sources.<br />

2<br />

Drs Elizabeth Aitken (Department <strong>of</strong> Botany, University <strong>of</strong> Queensland) and Juliane Henderson will use joint funding from<br />

the Australian Banana Industry Protection Board (BIPB) and Horticulture Australia Limited (HAL) to investigate sequence<br />

variability <strong>of</strong> <strong>Mycosphaerella</strong> causing <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> on banana.<br />

63


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

The status <strong>of</strong> black <strong>leaf</strong> streak disease in Australia is yet to be confirmed and<br />

the future application <strong>of</strong> diagnostic tests is uncertain. The method could be used<br />

to maintain Australia’s disease-free status, as far as black <strong>leaf</strong> streak disease is<br />

concerned, or to monitor pathogen populations for control measures should the<br />

disease become endemic. Either would ensure that the best diagnostic assay is<br />

available to the Australian banana industry.<br />

References<br />

Hayden H. 2001. Genetic variability in populations <strong>of</strong> pathogens causing black and yellow<br />

Sigatoka <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>. PhD Thesis, University <strong>of</strong> Queensland, Australia.<br />

Johanson A. 1997. Detection <strong>of</strong> Sigatoka Leaf Spot Pathogens <strong>of</strong> Banana by the Polymerase<br />

Chain Reaction. Natural Resources Institute, Chatham, UK.<br />

Rychlik W. 1993. Selection <strong>of</strong> primers for polymerase chain reaction. Pp. 31-40 in Methods<br />

in Molecular Biology, Vol. 15: PCR Protocols: Current Methods and Applications (B.A.<br />

White, ed.). Humana Press Inc, Totowa, New Jersey.<br />

Stewart C.N. and L.E. Via. 1993. A rapid CTAB DNA isolation technique useful for RAPD<br />

fingerprinting and other PCR applications. BioTechniques 14(5):748-750.<br />

White T.J., T. Bruns, S. Lee and J.W. Taylor. 1990. Amplification and direct sequencing <strong>of</strong><br />

fungal ribosomal RNA genes for phylogenetics. Pp. 315-322 in PCR Protocols: A Guide<br />

to Methods and Applications (M.A. Innis, D.H. Gelfand, J.J. Sninsky and T.J. White, eds.).<br />

Academic Press Inc., San Diego, USA.<br />

64


Session 1<br />

A. Viljoen et al.<br />

Impact <strong>of</strong> minor <strong>Mycosphaerella</strong><br />

pathogens on <strong>bananas</strong> (Musa) in<br />

South Africa<br />

A. Viljoen 1 ,A.K.J. Surridge 1 and P.W. Crous 2<br />

Abstract<br />

Of the species <strong>of</strong> <strong>Mycosphaerella</strong> known to occur on <strong>bananas</strong>, only M. musicola and M. musae<br />

occur in South Africa. Since both species are less damaging than M. fijiensis and M. eumusae,they<br />

are considered minor <strong>Mycosphaerella</strong> pathogens <strong>of</strong> this host. However, both M. musicola and<br />

M. musae can cause significant damage to <strong>bananas</strong> in the subtropics. For several years, M. musicola<br />

seemed to be the dominant pathogen <strong>of</strong> banana leaves in South Africa. It was very severe in<br />

banana plantations in Southern KwaZulu-Natal in the early 1990s, and caused losses <strong>of</strong> up to<br />

50% in Cavendish <strong>bananas</strong> due to early ripening and lower yields in the Komatipoort area in 1999<br />

and 2000. A highly coordinated disease management programme involving severe de<strong>leaf</strong>ing and<br />

fungicidal sprays has reduced the impact <strong>of</strong> Sigatoka disease in the country since 2001. However,<br />

<strong>Mycosphaerella</strong> speckle now appears to have replaced Sigatoka disease as the dominant <strong>leaf</strong><br />

pathogen in all banana growing areas <strong>of</strong> South Africa. Management strategies for Sigatoka disease<br />

seem to be less effective against <strong>Mycosphaerella</strong> speckle. Although this fungus primarily affects<br />

older leaves, the disease has become very severe in southern KwaZulu-Natal during 2002. Its<br />

economic impact and epidemiology, however, still have to be determined.<br />

Resumen - Impacto de los patógenos de <strong>Mycosphaerella</strong> de menor importancia sobre<br />

los bananos (Musa) en Africa del Sur<br />

De las varias especies de <strong>Mycosphaerella</strong> que ocurren en los bananos, solo M. musicola y M. musae,<br />

ocurren en Africa del Sur. Ya que ambas especies causan menores daños que M. fijiensis y<br />

M. eumusae, ellas se consideran patógenos de <strong>Mycosphaerella</strong> de menor importancia en este<br />

hospedante. Sin embargo, tanto M. musicola como M. musae pueden causar daños significativos<br />

a los bananos en los subtrópicos. Durante varios años, M. musicola se consideró el patógeno<br />

dominante en las hojas de los bananos en Africa del Sur. A principios de la década de los 90, este<br />

patógeno afectó severamente las plantaciones bananeras en el sur de KwaZulu-Natal, y causó<br />

pérdidas de hasta 50% en los bananos Cavendish, debido a una maduración precoz y bajos<br />

rendimientos en el área de Komatipoort en 1999 y 2000. Un programa coordinado de manejo<br />

de la enfermedad, que incluyó deshoje y rociados de funguicidas redujo el impacto de la<br />

1<br />

University <strong>of</strong> Pretoria, Pretoria, South Africa<br />

2<br />

University <strong>of</strong> Stellenbosch, Matieland, South Africa<br />

65


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Sigatoka amarilla en el país desde 2001. Sin embargo, la mancha causada por <strong>Mycosphaerella</strong><br />

parece haber reemplazado actualmente la Sigatoka amarilla como patógeno foliar dominante<br />

en todas las zonas productoras de banano en Africa del Sur. Parece que las estrategias de manejo<br />

de la Sigatoka amarilla son menos eficaces contra la mancha causada por <strong>Mycosphaerella</strong>.Aunque<br />

este hongo afecta principalmente las hojas viejas, la enfermedad se hizo muy severa en el sur<br />

de KwaZulu-Natal durante el año 2002. No obstante, aún falta determinar su impacto económico<br />

y la epidemiología.<br />

Résumé - Impact des pathogènes mineurs de <strong>Mycosphaerella</strong> sur les bananiers (Musa)<br />

en Afrique du Sud<br />

De toutes les espèces connues de <strong>Mycosphaerella</strong> affectant les bananiers, seules M. musicola<br />

et M. musae se trouvent en Afrique du Sud.Vu que ces deux espèces provoquent moins de dégâts<br />

que M. fijiensis et M. eumusae,elles sont considérées comme étant des pathogènes mineurs de<br />

cet hôte. Toutefois, M. musicola et M. musae peuvent toutes deux provoquer des dégâts<br />

significatifs aux cultures de bananes dans la zone subtropicale. Pendant plusieurs années, il<br />

semblait que M. musicola était le pathogène dominant des feuilles de bananiers en Afrique du<br />

Sud. L’infection était même très grave dans les plantations du sud du KwaZulu-Natal au début<br />

des années 1990 et en 1999 et 2000 a provoqué chez les bananiers Cavendish de la région de<br />

Komatipoort des pertes pouvant aller jusqu’à 50% dues à un mûrissement prématuré des fruits<br />

et à des rendements réduits. Un programme de gestion de la maladie parfaitement coordonné<br />

impliquant un défeuillage massif ainsi que des traitements fongicides a réduit l’impact de la<br />

maladie de Sigatoka dans le pays depuis 2001. Toutefois, <strong>Mycosphaerella</strong> speckle semble<br />

maintenant avoir remplacé la maladie de Sigatoka et se trouve être le pathogène dominant dans<br />

les régions de culture de la banane en Afrique du Sud. Les stratégies de gestion de la maladie<br />

de Sigatoka semblent moins efficaces envers le <strong>Mycosphaerella</strong> speckle. Bien que ce pathogène<br />

affecte en premier les feuilles les plus âgées, la maladie est devenue très grave dans le sud du<br />

KwaZulu-Natal en 2002. Son impact économique ainsi que son épidémiologie restent encore à<br />

être déterminés.<br />

Introduction<br />

Fungi that cause disease on leaves <strong>of</strong> banana and plantain include <strong>Mycosphaerella</strong><br />

fijiensis M. Morelet (the causal agent <strong>of</strong> black <strong>leaf</strong> streak disease), M. musicola<br />

Leach ex J.L. Mulder (the causal agent <strong>of</strong> Sigatoka disease), M. eumusae Crous et<br />

X. Mourichon (the causal agent <strong>of</strong> eumusae <strong>leaf</strong> <strong>spot</strong> disease) and M. musae (Speg.)<br />

Syd. et P. Syd. (the causal agent <strong>of</strong> <strong>Mycosphaerella</strong> speckle). M. fijiensis is the most<br />

virulent and economically important <strong>of</strong> the four <strong>Mycosphaerella</strong> spp. Since its<br />

discovery in Fiji in 1963, M. fijiensis has replaced M. musicola as the main <strong>leaf</strong><br />

pathogen in all tropical countries that produce banana (Jones, 2000). However,<br />

Sigatoka disease is still the main <strong>leaf</strong> disease in the subtropics and at higher altitudes<br />

in the tropics. In 1995, a new disease, eumusae <strong>leaf</strong> <strong>spot</strong>, was reported on Musa<br />

(Carlier et al., 2000). Eumusae <strong>leaf</strong> <strong>spot</strong> has only been found in Southeast Asia and<br />

in Nigeria, West Africa, (Jones, these proceedings) but is very damaging there. Black<br />

<strong>leaf</strong> streak disease, Sigatoka disease and eumusae <strong>leaf</strong> <strong>spot</strong> disease comprise the<br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> disease complex on banana. <strong>Mycosphaerella</strong> speckle is not<br />

considered to be important on banana. It is limited to the subtropics and is severe<br />

only on Cavendish <strong>bananas</strong> in Australia (Jones, 2000).<br />

The dominant <strong>Mycosphaerella</strong> spp. pathogens that occur in the subtropics are<br />

M. musicola and M. musae. Since neither consistently damage banana leaves, they<br />

66


Session 1<br />

A. Viljoen et al.<br />

are considered minor pathogens. The objective <strong>of</strong> this manuscript is to report the<br />

impact <strong>of</strong> such minor <strong>Mycosphaerella</strong> pathogens on <strong>bananas</strong> in South Africa.<br />

Banana production in South Africa<br />

Bananas are produced in six areas in South Africa (Figure 1). The crop was<br />

introduced from India at the beginning <strong>of</strong> the 19 th century. Production began along<br />

the southern and northern sections <strong>of</strong> the KwaZulu-Natal (KZN) coast, then<br />

introduced in the former Transvaal province and planted in Kiepersol, Tzaneen<br />

and Levubu. The largest production area, the Onderberg (4600 ha), became<br />

important in the 1990s. The total area <strong>of</strong> commercial banana production is<br />

12 500 ha, and is with Cavendish cultivars only. Almost 90% <strong>of</strong> new plantings<br />

are from tissue culture, and transport <strong>of</strong> banana plants between production areas<br />

is strictly controlled. All <strong>bananas</strong> are consumed locally, but there is a possibility<br />

<strong>of</strong> export to the Middle East.<br />

Leaf <strong>diseases</strong> <strong>of</strong> banana in South Africa<br />

Since 1999, regular surveys <strong>of</strong> areas where banana is cultivated have shown that<br />

Sigatoka disease, <strong>Mycosphaerella</strong> speckle and Cordana <strong>leaf</strong> <strong>spot</strong> (caused by<br />

Cordana musae [Zimm.] Höhn.) are present in all production areas. Cladosporium<br />

speckle (Cladosporium musae E.W. Mason) was found only in Levubu. Sigatoka<br />

disease and <strong>Mycosphaerella</strong> speckle were the most important.<br />

The banana <strong>leaf</strong> <strong>diseases</strong> in the southern part <strong>of</strong> Africa have not been studied<br />

very much. Black <strong>leaf</strong> streak disease is present in most tropical African countries<br />

(Jones, 2000), and has been reported as far south as northern Malawi (Ploetz,<br />

1992). However, black <strong>leaf</strong> streak disease is not known in Zimbabwe (which borders<br />

the Levubu area) and Mozambique (which borders the Onderberg area) (Figure 1).<br />

Figure 1.<br />

The six banana production areas<br />

<strong>of</strong> South Africa: Levubu, Letaba<br />

(Tzaneen), Hazyview (Kiepersol),<br />

Onderberg, and northern and<br />

southern KwaZulu-Natal.<br />

67


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

A severe outbreak <strong>of</strong> Sigatoka disease in 1999-2000, prompted an investigation<br />

into the identity <strong>of</strong> the causal agent. Samples were collected in all production areas,<br />

and the fungi identified using morphological and molecular techniques (Surridge et<br />

al., these proceedings). PCR primers developed by Johansen and Jeger (1994) were used<br />

to distinguish between Sigatoka disease and black <strong>leaf</strong> streak disease. Isolates from<br />

single conidia were further sequenced (ITS region) and compared with sequence data<br />

<strong>of</strong> M. musicola, M. fijiensis and M. eumusae from GenBank. All local isolates proved<br />

to be M. musicola, which causes Sigatoka disease. There was no evidence <strong>of</strong><br />

M. fijiensis and M. eumusae, suggesting they have not been introduced in South Africa.<br />

The severity <strong>of</strong> the outbreaks was attributed to favourable weather conditions and<br />

increases in the amount <strong>of</strong> inoculum.<br />

The life cycles <strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> have a sexual (teleomorph) stage,<br />

which produces ascospores, and an asexual (anamorph) stage, which produces conidia<br />

(Jones, 2000). Conidia are the main spore produced by M. musicola (Meredith, 1970).<br />

Conidia are dispersed within the <strong>leaf</strong> canopy and to neighbouring plants by rain, which<br />

dislodges and washes them onto adjacent leaves. Ascospores are forcefully discharged<br />

and spread by wind currents over bigger distances than conidia. Both types <strong>of</strong> spore<br />

require moisture for production, release, infection, growth and sporulation. Most stages<br />

in the life cycle take place over a wide range <strong>of</strong> temperatures; however, minimum night<br />

temperatures <strong>of</strong> 18 o C and 21 o C are needed for the production <strong>of</strong> conidia and<br />

ascospores <strong>of</strong> M. musicola, respectively (Meredith, 1970). Conidia are produced on both<br />

<strong>leaf</strong> surfaces, while ascospore production is almost three times greater on the upper<br />

(adaxial) than lower (abaxial) <strong>leaf</strong> surface (Meredith, 1970).<br />

Climatic conditions in South Africa<br />

The banana production areas <strong>of</strong> South Africa are located in the east between<br />

25 o and 30 o latitude and 30 o and 32 o longitude. The areas have a subtropical<br />

climate with cool, dry winters and warm, wet summers. Rainfall and temperature<br />

data for the Onderberg over a period <strong>of</strong> 10 years showed that November and<br />

March were the most favourable months for infection and disease development<br />

(Figure 2). During this time minimum night temperatures exceed 18 o C, which is<br />

necessary for the production <strong>of</strong> conidia. Minimum night temperatures exceed 21 o C<br />

only in January and February, therefore the period for ascospore production is<br />

short. Disease development is most rapid between November and March but slows<br />

substantially during the cooler months from May through September. Climatic<br />

conditions in South Africa provide ample opportunities for the management <strong>of</strong><br />

Sigatoka disease.<br />

The impact <strong>of</strong> <strong>Mycosphaerella</strong> <strong>diseases</strong><br />

Van den Boom and Kuhne (1969) first reported Sigatoka disease in South Africa, although<br />

the disease was also mentioned in 1954 (Meredith, 1970). The first report <strong>of</strong> <strong>Mycosphaerella</strong><br />

speckle in South Africa was in 1973 (Brodrick, 1973). Despite these late reports,<br />

both <strong>diseases</strong> have been associated with banana leaves for as long as producers can<br />

remember.<br />

68


Session 1<br />

A. Viljoen et al.<br />

35<br />

140<br />

30<br />

120<br />

25<br />

100<br />

Temperature<br />

20<br />

15<br />

10<br />

5<br />

0<br />

January<br />

February<br />

March<br />

April<br />

May<br />

June<br />

July<br />

August<br />

September<br />

October<br />

November<br />

December<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Rainfall<br />

Ave. min. temperature (°C) Ave. max. temperature (°C) Ave. rainfall (mm)<br />

Figure 2. Average annual temperature range and rainfall in the banana growing areas <strong>of</strong> South Africa.<br />

Sigatoka disease first became severe in South Africa during the 1960’s (Van den<br />

Boom and Kuhne, 1969). The disease devastated production in southern KZN in the<br />

1990s and in the Onderberg in 2000 (Viljoen, unpublished data). <strong>Mycosphaerella</strong> musicola<br />

infects the first three leaves <strong>of</strong> the banana plant. The symptoms first become visible on<br />

the third or fourth <strong>leaf</strong> (Jones, 2000). Under favourable weather conditions and with<br />

large amounts <strong>of</strong> inoculum, M. musicola can destroy all leaves after the stage when<br />

bunches are produced. This is what happened during the 1999-2000 outbreaks <strong>of</strong><br />

Sigatoka disease in the Onderberg. Damage included smaller fruits, smaller bunches,<br />

and premature fruit ripening in the field and in boxes. Farmers reported losses <strong>of</strong> up<br />

to 50% <strong>of</strong> the crop. An extensive disease management programme was implemented<br />

in October 2000 to halt the devastation. All leaves with Sigatoka lesions were cut and<br />

turned over on the plantation floor to limit the release <strong>of</strong> air-borne ascospores. Many<br />

bunches were sacrificed, in one instance amounting to nearly 18 000 bunches on a<br />

farm <strong>of</strong> about 40 ha. A fungicidal spray programme with protectant and systemic<br />

fungicides when the rainy season started and night temperatures exceeded 18 o C was<br />

recommended to growers. A total <strong>of</strong> six sprays <strong>of</strong> systemic fungicide were recommended,<br />

interrupted with a protectant fungicide after every second application <strong>of</strong> systemic<br />

fungicide (Peterson et al. these proceedings).<br />

None <strong>of</strong> the farmers applied the recommended number <strong>of</strong> sprays, and only 2-4 sprays<br />

were applied in total. The cost <strong>of</strong> fungicides, therefore, was small compared with the<br />

costs <strong>of</strong> fungicide sprays used to control black <strong>leaf</strong> streak disease in the tropics. Sigatoka<br />

disease was almost absent from banana fields in 2001, and current control strategies<br />

are now limited to cutting leaves and the application <strong>of</strong> one or two sprays, dependant<br />

on pre-seasonal <strong>leaf</strong> <strong>spot</strong> incidence, per year.<br />

<strong>Mycosphaerella</strong> musae infects older leaves <strong>of</strong> banana plants (Jones, 2000).<br />

<strong>Mycosphaerella</strong> speckle is rarely visible above the fifth fully open <strong>leaf</strong>, and seldom affects<br />

fruit quality and quantity after bunching. Since 2000, <strong>Mycosphaerella</strong> speckle has become<br />

69


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

more severe, and now is the main <strong>leaf</strong> disease <strong>of</strong> banana in South Africa. The symptoms<br />

are <strong>leaf</strong> yellowing (chlorosis) and death (necrosis). Necrosis is most visible on the older<br />

leaves, but, in 2002, chlorosis affected leaves as young as the third <strong>leaf</strong> after bunching<br />

in southern KZN. The effects on yield have not yet been determined. Control strategies<br />

are similar to those for Sigatoka disease, and include removing leaves and applying<br />

fungicides.<br />

Conclusion<br />

<strong>Mycosphaerella</strong> musicola and M. musae are the only <strong>Mycosphaerella</strong> <strong>leaf</strong> pathogens<br />

<strong>of</strong> banana in South Africa. They are considered to be minor pathogens, but become<br />

damaging under favourable weather conditions and in the presence <strong>of</strong> large<br />

amounts <strong>of</strong> inoculum. Subtropical climatic conditions and a clear understanding <strong>of</strong><br />

the biology and epidemiology <strong>of</strong> M. musicola make the management <strong>of</strong> Sigatoka<br />

disease relatively easy. The increased severity <strong>of</strong> <strong>Mycosphaerella</strong> speckle may result<br />

from the management <strong>of</strong> Sigatoka disease. The quantity <strong>of</strong> M. musicola ascospores<br />

released into the air is reduced by placing leaves upside down on the ground, but<br />

this probably increases the quantity <strong>of</strong> M. musae ascospores, which are mainly<br />

released from the lower <strong>leaf</strong> surface (Jones, 2000). A better understanding <strong>of</strong> the<br />

biology and epidemiology <strong>of</strong> M. musae is needed to develop the necessary<br />

management practices for <strong>Mycosphaerella</strong> speckle in the subtropics.<br />

References<br />

Brodrick H.T. 1973. Spikkelblaar. Banana Series Journal J4:1-2.<br />

Carlier J., M.F. Zapater, F. Lapeyre, D.R. Jones, and X. Mourichon. 2000. Septoria <strong>leaf</strong> <strong>spot</strong><br />

<strong>of</strong> banana: A newly discovered disease caused by <strong>Mycosphaerella</strong> eumusae (anamorph<br />

Septoria eumusae). Phytopathology 90(8):884-890.<br />

Johanson A. and M.J. Jeger. 1993. Use <strong>of</strong> PCR for detection <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis and<br />

M. musicola, the causal agents <strong>of</strong> Sigatoka <strong>leaf</strong> <strong>spot</strong>s in banana and plantain. Mycological<br />

Research 96(6):670-674.<br />

Jones D.R. 2000. Fungal <strong>diseases</strong> <strong>of</strong> the foliage. Pp. 37-141 in Disease <strong>of</strong> banana, abacá and<br />

enset. (D.R. Jones ed.) CAB International, Wallingford, UK, 544pp.<br />

Meredith D.S. 1970. Banana <strong>leaf</strong> <strong>spot</strong> disease (Sigatoka) caused by <strong>Mycosphaerella</strong> musicola<br />

Leach. Phytopathological Papers no 11, Commonwealth Mycological Institute, Kew, Surrey,<br />

UK. 147pp.<br />

Ploetz R.C. 1992. A current appraisal <strong>of</strong> banana and plantain <strong>diseases</strong> in Malawi. Tropical<br />

Pest Management 38:36-42.<br />

Van den Boom T. and F.A. Kuhne. 1969. First report <strong>of</strong> Sigatoka disease <strong>of</strong> banana in South<br />

Africa. Citrus Journal 428:17-18.<br />

70


Session 1<br />

L. Pérez Vicente et al.<br />

Economic impact and management<br />

<strong>of</strong> black <strong>leaf</strong> streak disease in Cuba<br />

L. Pérez Vicente,J.M. Alvarez and M. Pérez<br />

Abstract<br />

Black <strong>leaf</strong> streak disease (caused by <strong>Mycosphaerella</strong> fijiensis) is the most damaging disease<br />

found in Musa plantations in Cuba. Four years after the appearance <strong>of</strong> the disease in 1990,<br />

it had replaced Sigatoka disease (caused by M. musicola) in all areas <strong>of</strong> the country. During<br />

1991 and 1992, plantations <strong>of</strong> ‘Cavendish’ <strong>bananas</strong> received 23 fungicide applications per year,<br />

which accounted for 15% <strong>of</strong> total production costs. A system <strong>of</strong> integrated control based on<br />

cultural practices, a bioclimatic warning system for timing fungicide application and the use<br />

<strong>of</strong> systemic fungicides and mineral oil, reduced the number <strong>of</strong> applications to 13-15 per year<br />

and the cost <strong>of</strong> control to less than 40%. Black <strong>leaf</strong> streak disease has seriously affected the<br />

production <strong>of</strong> susceptible varieties. In 1989, more than 40 000 ha <strong>of</strong> plantain (Musa cv. AAB)<br />

and 14 000 ha <strong>of</strong> ‘Cavendish’ (Musa cv. AAA) were treated with fungicide. However, by<br />

the end <strong>of</strong> 1995 the areas had decreased by 69% and 51% respectively. Since 1994, ‘FHIA-18’,<br />

‘FHIA-03’, ‘FHIA-01-1’, ‘FHIA-02’ and ‘FHIA-21’ with partial resistance to black <strong>leaf</strong> streak<br />

disease were introduced into Cuba. Currently, there are 10 000 ha planted with these clones<br />

resulting in an 80% reduction in the use <strong>of</strong> fungicide. The severity <strong>of</strong> black <strong>leaf</strong> streak disease<br />

on ‘FHIA-18’ has been inversely correlated with the availability <strong>of</strong> total K and with the ratio<br />

K/(K+Ca+Mg) in soil and foliage. Variability in the pathogenicity <strong>of</strong> M. fijiensis populations<br />

has been studied in order to identify possible changes that result from large scale cultivation<br />

<strong>of</strong> FHIA hybrids with partial resistance.<br />

Resumen - Impacto económico y manejo de la enfermedad de la raya negra en Cuba<br />

La enfermedad de la raya negra o Sigatoka negra, causada por <strong>Mycosphaerella</strong> fijiensis, es la<br />

enfermedad más nociva presente en las plantaciones de musáceas en Cuba. Cuatro años<br />

después de su aparición en 1990, reemplazó a la Sigatoka amarilla (M. musicola) en todas las<br />

áreas del país. Durante 1991 y 1992 se realizaron hasta 23 aplicaciones de fungicida por año<br />

en plantaciones de banano ‘Cavendish’ con un costo de protección alcanzando el 15% del costo<br />

total de la producción. Un sistema de manejo integrado basado en prácticas culturales y<br />

pronóstico bioclimático de los momentos de tratamiento permitió reducir a 13-15 los<br />

tratamientos con funguicidas sistémicos y aceite mineral al año en las principales plantaciones<br />

de producción con un costo menor de un 40%. La Sigatoka negra ha tenido un serio impacto<br />

en la producción de plátanos susceptibles. En 1989, existían más de 40 000 ha de plátanos<br />

INISAV, Cuba<br />

71


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

(Musa cv. AAB) y 14 000 ha de banana ‘Cavendish’ (Musa cv. AAA) bajo protección fúngica. A<br />

finales de 1995 se habían reducido en un 69 y 51% respectivamente. A partir de 1994, se<br />

introdujeron los clones ‘FHIA-18’,‘FHIA-03’,‘FHIA-01-1’,‘FHIA-02’ y ‘FHIA-21’ con resistencia parcial<br />

a Sigatoka negra. En la actualidad existen 10 000 ha plantadas de estos clones y se ha reducido<br />

el consumo de fungicidas en un 80%. Se ha observado una correlación inversa entre la<br />

severidad de los ataques de Sigatoka negra en el clon FHIA-18 y la disponibilidad total de K<br />

y con la relación K/(K+Ca+Mg) en suelo y hojas. Se desarrollan estudios de la variabilidad<br />

genética de las poblaciones de M. fijiensis con el objetivo de determinar cambios de<br />

patogenicidad a causa del cultivo a gran escala de clones con resistencia parcial.<br />

Résumé - Impact économique et gestion de la maladie des raies noires à Cuba<br />

La maladie des raies noires, causée par <strong>Mycosphaerella</strong> fijiensis, est la maladie la plus<br />

destructrice des plantations de bananes à Cuba. Quatre années après l’apparition de la<br />

maladie en 1990, elle avait remplacé dans tout le pays la maladie de Sigatoka, causée par<br />

M. musicola.Pendant les années 1991 et 1992, les plantations de bananes ‘Cavendish’ ont reçu<br />

23 traitements de fongicides par an , représentant 15% des coûts totaux de production. Un<br />

système de lutte intégrée basé sur les pratiques culturales, un système de prévision<br />

bioclimatique pour déterminer le moment d’application des fongicides et l’utilisation de<br />

fongicides systémiques et d’huile minérale, ont permis de réduire le nombre de traitements<br />

à 13-15 par an et les coûts de plus de 60%. La maladie des raies noires a sérieusement affecté<br />

la production de variétés sensibles. En 1989, plus de 40 000 ha de plantain (Musa cv. AAB) et<br />

14 000 ha de bananes ‘Cavendish’ (Musa cv. AAA) ont été traités avec des fongicides. Toutefois,<br />

fin 1995, les surfaces traitées ont été réduites respectivement de 69% et 51%. Depuis 1994,‘FHIA-<br />

18’,‘FHIA-03’,‘FHIA-01-1’,‘FHIA-02’ et ‘FHIA-21’, des hybrides possédant une résistance partielle<br />

à la maladie des raies noires, ont été introduits à Cuba. Actuellement, 10 000 ha sont plantés<br />

avec ces clones ce qui entraîné une baisse de 80% de l’utilisation de fongicides. La gravité<br />

de la maladie des raies noires sur ‘FHIA-18’ a été inversement corrélée avec la disponibilité du<br />

K total et avec le rapport K/(K+Ca+Mg) dans le sol et le feuillage. La variabilité du pouvoir<br />

pathogène des populations de M. fijiensis a été étudiée afin d’identifier des modifications<br />

potentielles qui pourraient découler de la culture à grande échelle d’hybrides FHIA partiellement<br />

résistants.<br />

Introduction<br />

Many fungal, bacterial and viral <strong>diseases</strong> affect Musa but undoubtedly the one<br />

that has had the most socio-economic impact at a world level has been black <strong>leaf</strong><br />

streak disease caused by the fungus <strong>Mycosphaerella</strong> fijiensis Morelet.<br />

Black <strong>leaf</strong> streak disease was first reported in Cuba in 1991 (Vidal, 1992).<br />

Previously, the main banana and plantain disease in Cuba was Sigatoka disease<br />

caused by M. musicola Leach ex Mulder, for which a warning system (Ganry and<br />

Meyer, 1972b) for timing the application <strong>of</strong> fungicides in oil emulsions had been<br />

established (Perez, 1983, 1989). <strong>Mycosphaerella</strong> musicola causes considerable<br />

economic losses in ‘Cavendish’ cultivars (AAA) and occasionally in plantains<br />

(AAB), which generally show an acceptable level <strong>of</strong> resistance to the pathogen<br />

(Vakili, 1968; Perez et al., 1981).<br />

This article reviews the economic impact <strong>of</strong> black <strong>leaf</strong> streak disease on banana<br />

and plantain production in Cuba, and summarises studies on the epidemiology,<br />

disease management and resistance <strong>of</strong> cultivars carried out over several years<br />

in Cuba.<br />

72


Session 1<br />

L. Pérez Vicente et al.<br />

Impact <strong>of</strong> disease<br />

Protection cost<br />

In the 1980s, 15 to 16 applications <strong>of</strong> mineral oil were carried out each year to control<br />

M. musicola. A quarter <strong>of</strong> applications contained mixtures <strong>of</strong> dithiocarbamate<br />

fungicides and two or three contained benomyl or propiconazole. By the end <strong>of</strong> the<br />

80s, the use <strong>of</strong> a bioclimatic model to forecast the treatments, based on the method<br />

<strong>of</strong> scoring the rate <strong>of</strong> development <strong>of</strong> the disease (Ganry and Meyer, 1972a, b), led<br />

to a 30% reduction in the cost <strong>of</strong> controlling M. musicola (Perez, 1989). The cost per<br />

hectare in ‘Cavendish’ plantations varied between US $134 and US $241 (Figure 1).<br />

In 1991 and 1992, the first and second years after the first outbreak <strong>of</strong> black <strong>leaf</strong><br />

streak disease, the cost per hectare rose to US$640 and US $801. The adoption <strong>of</strong> a<br />

warning system to time the applications <strong>of</strong> oil-based systemic fungicides (Perez et<br />

al., 1993a, 2002) reduced the cost in later years (Perez et al., 1993b, 1997, 2000a, b;<br />

Porras and Perez, 1997; Perez, 1998).<br />

Changes in areas occupied by the cultivars<br />

Banana and plantain production in Cuba is for local consumption due to a lack<br />

<strong>of</strong> export markets and to insufficient production levels to satisfy the high demand.<br />

The areas planted with different cultivars are shown in Figure 2. In 1990, more<br />

than 14 000 ha were planted with ‘Cavendish’ cultivars and treated with fungicide<br />

to control Sigatoka disease. In 2001, ten years after the arrival <strong>of</strong> black <strong>leaf</strong> streak<br />

disease, the area was reduced to 15% <strong>of</strong> that figure. In 1990, more than<br />

43 000 ha were planted with plantains but black <strong>leaf</strong> streak disease reduced this<br />

to 18% in later years.<br />

The ‘Cavendish’ plantations have been replanted with resistant FHIA hybrids,<br />

particularly ‘FHIA-23’ and ‘FHIA-18’ (AAAB). As a result, the amount <strong>of</strong> fungicide<br />

imported for controlling black <strong>leaf</strong> streak disease in ‘Cavendish’ has declined<br />

(Figure 3). The plantains have been replaced by ‘Burro CEMSA 3/4’ (ABB) and<br />

‘FHIA-03’ (AABB).<br />

Epidemiology <strong>of</strong> black <strong>leaf</strong> streak disease:<br />

relation betwen weather and disease development<br />

Tables 1 and 2 show the correlation matrix between weather variables – the cumulated<br />

quantity and duration <strong>of</strong> rain over a 10 or 14-day period, the weekly Piche<br />

evaporation, the weekly accumulated hours with relative humidity over 95%<br />

(RH>95%) - and the state <strong>of</strong> development <strong>of</strong> the disease (Fouré, 1988) recorded weekly<br />

for eight weeks in ‘Grande naine’ and ‘CEMSA 3/4’. High significative correlations<br />

were found between the accumulated quantity and duration <strong>of</strong> rain and RH>95%,<br />

and the state <strong>of</strong> development <strong>of</strong> the disease in banana and plantain.<br />

In ‘Cavendish’ <strong>bananas</strong>, Perez et al. (1993b, 2000a) found the highest correlations<br />

between the cumulated quantity and duration <strong>of</strong> rain over a 10, 14 or 28-day period,<br />

and the state <strong>of</strong> development <strong>of</strong> the disease four to six weeks later. Regression<br />

73


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

50<br />

0<br />

Banao, 1984<br />

Artemisa, 1987<br />

Sagua, 1989<br />

Sagua, 1990<br />

Contramaestre, 1990<br />

Protection cost<br />

in US dollars per hectare<br />

Horquita, 1990<br />

La Cuba, 1991<br />

La Cuba, 1992<br />

Limoncito, 1994<br />

Menendez, 1994<br />

Menendez, 1995<br />

Nueva Paz, 1994<br />

Figure 1. The cost <strong>of</strong> protection per hectare in various banana producing enterprises before and after the<br />

first report <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis in Cuba. Bars for black <strong>leaf</strong> streak disease represent the costs in the<br />

first and second years after the disease appeared in each plantation.<br />

60<br />

50<br />

Planted area (x 1000 ha)<br />

40<br />

30<br />

20<br />

0<br />

Cavendish (AAA) Plantains (AAB) ABB cultivars FHIA cultivars<br />

1990 1995 1997 1999 2000<br />

Figure 2. Surface area planted with various types <strong>of</strong> banana and plantain in Cuba, 1990-2000.<br />

74


Session 1<br />

L. Pérez Vicente et al.<br />

2500<br />

12<br />

Cost <strong>of</strong> fungicides (x 1000 US dollars)<br />

2000<br />

1500<br />

1000<br />

10<br />

8<br />

6<br />

4<br />

2<br />

Planted area (X 1000 ha)<br />

0<br />

1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001<br />

0<br />

Cost <strong>of</strong> fungicides<br />

Planted area<br />

Figure 3. Cost <strong>of</strong> fungicides to protect ‘Cavendish’ plantations and area planted with FHIA hybrids that are<br />

partially resistant to black <strong>leaf</strong> streak disease.<br />

equations were used for develop models to predict the development <strong>of</strong> black <strong>leaf</strong><br />

streak disease as a function <strong>of</strong> the duration and intensity <strong>of</strong> rainfall. The observed<br />

and estimated state <strong>of</strong> development <strong>of</strong> the disease in ‘Grande naine’ as a function<br />

<strong>of</strong> rainfall four weeks before is shown in Figure 4.<br />

In ‘CEMSA 3/4’, Perez et al. (2000b), found high correlations between the state<br />

<strong>of</strong> development <strong>of</strong> the disease and the cumulated quantity and duration <strong>of</strong> rain over<br />

a 10 or 14-day period four to six weeks before, and the weekly Piche evaporation<br />

(PwEv) three to six weeks before (Table 2). The highest correlation coefficient was<br />

obtained between the cumulated quantity <strong>of</strong> rain over a 14-day period and the state<br />

<strong>of</strong> development <strong>of</strong> the disease five weeks later. In ‘CEMSA 3/4’, a highly significant<br />

negative correlation was found between PwEv and the state <strong>of</strong> development <strong>of</strong> black<br />

<strong>leaf</strong> streak disease three to five weeks later.<br />

In general, the state <strong>of</strong> development <strong>of</strong> the disease in any week <strong>of</strong> the year is highly<br />

dependent on <strong>leaf</strong> wetness four and five weeks before. Many <strong>of</strong> the biological process<br />

<strong>of</strong> the fungus, such as mating between compatible isolates and pseudothecia<br />

development (Mourichon and Zapater, 1990), conidia development, ascospore and<br />

conidia germ tube growth (Porras y Perez, 1997) and ascospore release from<br />

pseudothecia (Stover 1980), are dependent on the presence <strong>of</strong> a water layer on the<br />

<strong>leaf</strong> surface or on high relative humidity.<br />

In ‘CEMSA 3/4’, the cumulated duration <strong>of</strong> rain over a 14-day period and the state<br />

<strong>of</strong> development <strong>of</strong> the disease four weeks later are shown in Figure 5. The progressions<br />

and regressions <strong>of</strong> the disease were highly correlated with rainfall and <strong>leaf</strong> wetness.<br />

75


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Table 1. Correlation matrix between climatic factors and the state <strong>of</strong> development <strong>of</strong> black <strong>leaf</strong> streak disease in ‘Grande<br />

naine’, La Cuba 1995–1996. (Adapted from Perez et al., 2000a).<br />

Number <strong>of</strong> weeks after the recording <strong>of</strong> the independent variable<br />

Dependent Independent<br />

variable variable 0 1 2 3 4 5 6 7<br />

SD4L Rf7mm n.s. 0.34* n.s. 0.56 *** 0.61 *** 0.74 *** 0.41 * n.s.<br />

RfD7 min n.s. n.s. n.s. 0.45 ** 0.48 ** 0.77 *** 0.52 ** n.s.<br />

Rf10 mm n.s. n.s. 0.38 * 0.54 *** 0.80 *** 0.71 *** n.s. n.s.<br />

RfD10 min n.s. n.s. 0.37 * 0.39 * 0.74 *** 0.73 *** n.s. n.s.<br />

Rf14 mm n.s. n.s. 0.45 ** 0.64 *** 0.77 *** 0.69 *** n.s. n.s.<br />

RfD14 min n.s. n.s. 0.38 * 0.51 ** 0.75 *** 0.73 *** n.s. n.s.<br />

H7 n.s. n.s. n.s. n.s. 0.72 ** 0.71 ** 0.53 * n.s<br />

H10 n.s. n.s. 0.55* n.s. 0.79** 0.70 ** n.s. n.s.<br />

H14 n.s. n.s. n.s. n.s. n.s. n.s. n.s. n.s.<br />

RH>95% n.s. n.s. n.s. 0.316* 0.276* 0.308* 0.36** 0.36**<br />

PwEv -0.27 * -0.29 * n.s. n.s. n.s. n.s. n.s. n.s.<br />

TMed 0.32 * n.s. 0.29 * 0.29 * n.s. n.s. n.s. n.s.<br />

* Statistically significant at probability 0.05. ** Statistically significant at probability 0.01. *** Statistically significant at probability 0.001.<br />

n.s.: Not statistically significant.<br />

SD4L: State <strong>of</strong> development <strong>of</strong> black <strong>leaf</strong> streak disease in the four youngest leaves. Rf(n)mm: Cumulated quantity, in mm, <strong>of</strong> rain over a<br />

period <strong>of</strong> n days. RfD(n)min: Cumulated duration, in min, <strong>of</strong> rain over a period <strong>of</strong> n days. H(n): Cumulated quantity, in mm, <strong>of</strong> water on the<br />

leaves over a period <strong>of</strong> n days. PwEv: Weekly Piche evaporation.<br />

Table 2. Correlation matrix between climatic factors and the state <strong>of</strong> development <strong>of</strong> black <strong>leaf</strong> streak disease in ‘CEMSA<br />

3/4’ (Musa spp., AAB), La Cuba 1996. (Adapted from Perez et al., 2000b).<br />

Number <strong>of</strong> weeks after the recording <strong>of</strong> the independent variable<br />

Dependent Independent<br />

variable variable 0 1 2 3 4 5 6 7<br />

SD4L Rf7 mm n.s. n.s. n.s. 0.46 ** 0.48 ** 0.60 *** 0.58 *** 0.55 **<br />

RfD7 min n.s. n.s. n.s. 0.34 * 0.39 * 0.64 *** 0.55 ** 0.54 **<br />

Rf10 mm n.s. n.s. n.s. 0.51 ** 0.59 *** 0.74 *** 0.62 *** 0.40 *<br />

RfD10 min n.s. n.s. n.s. 0.40 * 0.50 ** 0.74 *** 0.57 *** 0.38 *<br />

Rf14 mm n.s. n.s. n.s. 0.52 ** 0.65 *** 0.74 *** 0.64 *** 0.39 *<br />

RfD14 min n.s. n.s. n.s. 0.41 * 0.58 *** 0.72 *** 62 *** 0.38 *<br />

H 7<br />

(mm) n.s. n.s. n.s. n.s. n.s. n.s. n.s. n.s.<br />

H 10<br />

(mm) 0.49* 0.50* n.s. n.s. 0.73** n.s. n.s. n.s.<br />

H 14<br />

(mm) 0.57** 0.57** 0.58* n.s. 0.81** 0.55* n.s. n.s.<br />

PwEv -0.44 ** -0.58 *** -0.60*** -0.70*** -0.74*** -0.75*** -0.52** n.s.<br />

TMed n.s. n.s. n.s. 0.40 * 0.45 ** 0.48 ** 0.51 ** 0.43 **<br />

* Statistically significant at probability 0.05. ** Statistically significant at probability 0.01. *** Statistically significant at probability 0.001.<br />

n.s.: Not statistically significant.<br />

SD4L: State <strong>of</strong> development <strong>of</strong> black <strong>leaf</strong> streak disease in the four youngest leaves. Rf(n)mm: Cumulated quantity, in mm, <strong>of</strong> rain over a<br />

period <strong>of</strong> n days. RfD(n)min: Cumulated duration, in min, <strong>of</strong> rain over a period <strong>of</strong> n days; H(n): Cumulated quantity, in mm, <strong>of</strong> water on the<br />

leaves over a period <strong>of</strong> n days. PwEv: Weekly Piche evaporation.<br />

76


Session 1<br />

L. Pérez Vicente et al.<br />

2500<br />

2000<br />

Cumulated duration <strong>of</strong> rain over a 14-day period<br />

2000<br />

1500<br />

1000<br />

500<br />

Duration <strong>of</strong> rainfall<br />

State <strong>of</strong> development<br />

<strong>of</strong> the disease 4 weeks later<br />

R 2 = 0.56*<br />

1750<br />

1500<br />

1250<br />

1000<br />

750<br />

500<br />

250<br />

State <strong>of</strong> development <strong>of</strong> the disease<br />

47 49 51 1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35<br />

Weeks<br />

0<br />

Figure 4. Correlation between the cumulated duration <strong>of</strong> rain, in minutes, over a 14-day period and the state<br />

<strong>of</strong> development <strong>of</strong> black <strong>leaf</strong> streak disease in the four youngest leaves (SD4L) <strong>of</strong> ‘Grande naine’ 4 weeks later.<br />

(From Perez et al., 2000b).<br />

Cumulated duration <strong>of</strong> rain over a 14-day pediod (min)<br />

2500<br />

2250<br />

2000<br />

1750<br />

1500<br />

1250<br />

1000<br />

750<br />

500<br />

250<br />

r = 0.72***<br />

Duration <strong>of</strong> rainfall<br />

State <strong>of</strong> development<br />

<strong>of</strong> the disease 4 weeks later<br />

48 50 52 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30<br />

Weeks<br />

2750<br />

2500<br />

2250<br />

2000<br />

1750<br />

1500<br />

1250<br />

1000<br />

750<br />

500<br />

250<br />

0<br />

State <strong>of</strong> development <strong>of</strong> disease<br />

Figure 5. Curves <strong>of</strong> the cumulated duration <strong>of</strong> rain, in minutes, over a 14-day period and the state <strong>of</strong><br />

development <strong>of</strong> black <strong>leaf</strong> streak disease in the four youngest leaves (SD4L) <strong>of</strong> ‘CEMSA 3/4’5 weeks later. (From<br />

Perez et al., 2000b).<br />

77


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

The curves <strong>of</strong> the observed and the estimated state <strong>of</strong> development <strong>of</strong> the<br />

disease using the model SD4L = 6.24 (S 14d<br />

Rfmm) – 198.2 PwEv + 1812 are shown<br />

in Figure 6. The model predicts the progress <strong>of</strong> the disease three weeks later.<br />

State <strong>of</strong> development <strong>of</strong> the disease (SD4L)<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

R2 = 0.85*<br />

Observed state <strong>of</strong> development<br />

Estimated state <strong>of</strong> development<br />

35 37 39 41 43 45 47 49 51 1 3 5 7 9 11 13 15 17 19 21<br />

Weeks<br />

Figure 6. Observed and estimated state <strong>of</strong> development <strong>of</strong> the disease in the four youngest leaves (SD4L).<br />

The estimated values were derived using the model SD4L = 6.24 (S 14d<br />

Rfmm) – 198.2 EvPp + 1812. (From Perez<br />

et al., 2000b).<br />

Between 1993 and 1996, 13 to 15 treatments <strong>of</strong> systemic fungicides per year were<br />

applied in ‘Cavendish’ plantations using the pattern <strong>of</strong> rainfall as the variable on<br />

which to base the decision <strong>of</strong> using fungicides. For example, the state <strong>of</strong> development<br />

<strong>of</strong> black <strong>leaf</strong> streak disease and the timing <strong>of</strong> fungicide treatments, using the<br />

bioclimatic model, are shown in Figure 7 for La Cuba in 1994.<br />

Between 8 and 10 treatments <strong>of</strong> fungicides are required each year in plantain<br />

plantations in Cuba, to achieve an adequate control <strong>of</strong> black <strong>leaf</strong> streak disease. The<br />

treatments carried out using a bio-climatic model for the cultivar ‘CEMSA 3/4’ in<br />

La Cuba plantations during 1996 are shown in Figure 8. However, the low yields <strong>of</strong><br />

plantain cultivars and the low prices paid for the product do not cover the costs <strong>of</strong><br />

controlling black <strong>leaf</strong> streak disease.<br />

A comparison <strong>of</strong> the cost <strong>of</strong> controlling black <strong>leaf</strong> streak disease using the<br />

bioclimatic model for timing applications with the cost <strong>of</strong> using a predetermined schedule<br />

<strong>of</strong> fungicide applications are shown in Table 3. The use <strong>of</strong> the bioclimatic model led to<br />

a 40% reduction in total costs, which resulted in an important reduction <strong>of</strong> the quantity<br />

<strong>of</strong> fungicides imported in Cuba to control black <strong>leaf</strong> streak disease in ‘Cavendish’<br />

plantations (Perez et al., 1993a, b, 1997, 2000a, b; Porras and Perez, 1997; Perez, 1998).<br />

The model for timing applications <strong>of</strong> fungicides has shown to be effective in<br />

regions where the total annual rainfall is under 2000 mm. The system depends on<br />

the use <strong>of</strong> systemic fungicides able to inhibit the evolution <strong>of</strong> infections already<br />

established in the host at the time <strong>of</strong> the application.<br />

78


Session 1<br />

L. Pérez Vicente et al.<br />

State <strong>of</strong> development <strong>of</strong> the disease (SD4L)<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

1 5 9 13 17 21 25 33 37 41 45 49 53<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS)<br />

WEEKS<br />

SD4L SD4L YLS YLS<br />

Figure 7. Black <strong>leaf</strong> streak disease control in a ‘Grande naine’ (AAA) plantation using the bioclimatic model<br />

for timing applications in La Cuba in 1994 (fields 1 and 2). Arrows indicate the moments <strong>of</strong> the application.<br />

(Adapted from Perez, 1998).<br />

State <strong>of</strong> development <strong>of</strong> the disease (SD4L)<br />

3000<br />

2500<br />

2000<br />

Pr<br />

1500<br />

1000<br />

500<br />

0<br />

33 36<br />

Te<br />

Te<br />

Tr<br />

Be<br />

Oil<br />

39 42 45 48 51 2 5 8<br />

Weeks<br />

Bi<br />

Oil<br />

11 14 17 20 23 26 29 32<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS)<br />

SD4L<br />

YLS<br />

Figure 8. Black <strong>leaf</strong> streak disease control in a ‘CEMSA 3/4’ plantation using the bioclimatic model for timing<br />

applications in La Cuba, 1995–1996. Arrows indicate the moments <strong>of</strong> the application with tebuconazole (Te),<br />

propiconazole (Pr), mineral oil (Oil), benomyl (Be) and tridemorph (Tr). (Adapted from Perez et al. 2000b).<br />

Resistance <strong>of</strong> cultivars<br />

Different studies have been reported on the resistance <strong>of</strong> banana and plantains<br />

to M. fijiensis (Meredith and Lawrence 1970, Firman 1972, Fouré et al.1984, Fouré<br />

et al. 1990). Fouré et al. (1990) reported two different types <strong>of</strong> resistant reaction<br />

79


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

against black <strong>leaf</strong> streak disease in Musa: the hypersensitive reaction observed<br />

in ‘Yangambi km 5’, and the partial resistance, that is expressed by the duration<br />

<strong>of</strong> the cycle <strong>of</strong> evolution <strong>of</strong> the disease and a reduction in reproduction <strong>of</strong> the<br />

pathogen. Studies done in 1995 (Hernandez and Perez, 2001), showed the<br />

reaction and components <strong>of</strong> the resistance to black <strong>leaf</strong> streak disease <strong>of</strong> various<br />

FHIA hybrids and Musa genotypes from the Cuba collection. Table 4 shows a<br />

significant increase in the incubation period (from <strong>leaf</strong> emergence and the<br />

appearance <strong>of</strong> the first symptoms), and in the transition period (from streak<br />

stage to necrotic <strong>spot</strong>s), as well as a significant reduction in the production <strong>of</strong><br />

pseudothecia. As a result, a reduction <strong>of</strong> the logistic rates <strong>of</strong> increment <strong>of</strong> infection<br />

(typical <strong>of</strong> partial resistance) takes place and the plants reach the flowering stage<br />

with a greater number <strong>of</strong> functional leaves.<br />

Table 3. Comparison <strong>of</strong> the number and the cost <strong>of</strong> fungicide applications in ‘Cavendish’ plantations using a bioclimatic<br />

model and following a pre-determined schedule. (Adapted from Perez, 1998).<br />

Timing <strong>of</strong> application<br />

Plantations<br />

Using a predetermined schedule Using a bioclimatic model<br />

Year Number Total cost Year Number Total cost<br />

<strong>of</strong> treatments (US$) <strong>of</strong> treatments (US$)<br />

La Cuba 1991 21 801.24<br />

1992 23 619.66<br />

1993 15 568.74<br />

1994 13 303.29<br />

1995 12 299.33<br />

1996 12 269.52<br />

Limoncito 1994 22 476.19<br />

1995 13 246.48<br />

1996 13 288.72<br />

Quemado 1995 8 172.39<br />

De Guines 1996 4 71.21<br />

up to June<br />

Menendez 1994 18 412.09<br />

1995 23 518.49<br />

Sola 1994 11 221.56<br />

1995 12 282.78<br />

1996 11 237.31<br />

Nueva Paz 1994 23 599.66 1996 13 326.56<br />

Guines 1995 13 308.96<br />

1996 10 219.78<br />

(Hurricane)<br />

A comparison <strong>of</strong> the infection indices and the youngest <strong>leaf</strong> <strong>spot</strong>ted (Stover<br />

and Dickson, 1970) are shown for FHIA hybrids plantations from four locations<br />

in Cuba, during the most favourable periods for disease development in each<br />

locality (Table 5).<br />

80


Session 1<br />

L. Pérez Vicente et al.<br />

Table 4. Reaction to black <strong>leaf</strong> streak disease <strong>of</strong> a group <strong>of</strong> FHIA hybrids in Cuba without fungicide protection. (Adapted<br />

from Hernandez and Perez, 2001).<br />

Clones Incubation period Transition period Number <strong>of</strong> functional<br />

(days) (days) leaves at harvest<br />

February June February June Mother First<br />

plant follower<br />

Spot size<br />

(mm)<br />

Number <strong>of</strong><br />

pseudothecia<br />

FHIA-1-1 43.5 b 28.2 a* 76.4** 75.5 9 8 17.3 15.9<br />

FHIA-02 46.9 b 31.0 a >150** 86.6 8 10 13.3 34.8<br />

FHIA-03 60.4 a 24.1 b >150** 107.7 10 8 15.5 31.6<br />

FHIA-18 52.8 ab 28.7 a >150** 119.0 12 9 14.3 35.0<br />

SH 3436 35.8 c 28.0 a 84.3* 80.2 10 7 12.7 9.5<br />

Grande naine 27.9 c 16.7 c 36.6 43.0 1 0 17.5 73.6<br />

* Different letters indicate significant differences at probability 0.05.<br />

** Most <strong>spot</strong>s stopped developing at Fouré’s stage 3 (Fouré, 1984).<br />

Table 5. Infection index (II) and youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS) in FHIA hybrids and ‘Cavendish’ at the moment <strong>of</strong> maximum<br />

disease severity.<br />

Cultivar Locality and year II (%) YLS<br />

FHIA-23 La Cuba, 1996 20.3 6.8<br />

La Cuba, 2001 16.3 7.8<br />

Alquízar, 2001 21.7 6.5<br />

FHIA-18 La Cuba, 1996 2.1 11.0<br />

La Cuba, 2000 (Fca. Berlier) 28.2 5.6<br />

La Cuba, 2000 (Fca. Cozola) 23.2 6.4<br />

La Cuba, 2001 10.3 7.6<br />

Alquizar, 2001 14.2 10.6<br />

Baracoa, 2002 15.8 9.6<br />

Cavendish La Cuba, 1996 35.6 4.3<br />

La Cuba, 2001 - -<br />

Alquizar, 2001 36.3 5.3<br />

Baracoa, 2002 59.8 4.3<br />

In the last two years, there has been an increase in the severity <strong>of</strong> black <strong>leaf</strong> streak<br />

disease in experimental plots in Ciego de Avila compared with the levels observed<br />

in 1996, despite the fact that in 1996 the plots were largely surrounded by fields <strong>of</strong><br />

‘Cavendish’ <strong>bananas</strong>. Increased disease severity in Ciego de Avila has been specially<br />

marked on the ‘FHIA-18’ and ‘FHIA-23’ hybrids. The causes <strong>of</strong> the changes in the<br />

susceptibility are under study. A negative correlation has been observed between<br />

the level <strong>of</strong> K in soil and the severity <strong>of</strong> black <strong>leaf</strong> streak disease. Table 6 shows<br />

values for the concentration <strong>of</strong> K and the K/K+Ca+Mg ratio in the soil <strong>of</strong> farms with<br />

red latosolic soils planted with the cultivar ‘FHIA-18’ and not sprayed against black<br />

<strong>leaf</strong> streak disease, and disease severity expressed as the youngest <strong>leaf</strong> <strong>spot</strong>ted.<br />

Farms with the lowest content <strong>of</strong> K in the soil had the highest severity <strong>of</strong> black<br />

<strong>leaf</strong> streak disease (Table 6). Potassium seems to have an important role in the defence<br />

mechanisms <strong>of</strong> banana plants <strong>of</strong> the FHIA cultivars to pathogens. Peng et al. (1999)<br />

81


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

reported that soils deficient in K are conducive to Fusarium oxysporum f.sp. cubense<br />

in Australia. Different stress factors have been shown to have an influence on<br />

peroxidase and phenylalanine ammonia lyase enzyme systems which are at the same<br />

time related to defence mechanism in plants to pathogens (Aguiar et al., 2000).<br />

A nationwide survey <strong>of</strong> pathogenic variability <strong>of</strong> populations <strong>of</strong> M. fijiensis has<br />

been undertaken based on the widespread use <strong>of</strong> FHIA hybrids in Cuba. A collection<br />

<strong>of</strong> single ascospore isolates <strong>of</strong> M. fijiensis from different cultivars collected at<br />

locations where FHIA hybrids have been extensively planted and from regions where<br />

they have not been introduced has been carried out. Artificial inoculations on selected<br />

hybrids under controlled conditions are in progress.<br />

Table 6. Disease severity expressed as the youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS) in farms planted with ‘FHIA-18’ and availability<br />

<strong>of</strong> potassium in the soil, La Cuba, Ciego de Avila, in Nov 2000.<br />

Farms YLS K K/(K+Ca+Mg)<br />

(meq/100g soil)<br />

El Berlier 5.6 0.44 1/59<br />

Cozola 6.4 0.52 1/56<br />

El Transformador (Cuba 3) 7.0 0.65 1/38<br />

El Colorado (Farm 10) 8.0 0.85 1/39<br />

La pista (Tin) 10.5 1.5 1/14<br />

Cooperative 11.2 1.2 1/15<br />

Conclusion<br />

1. Black <strong>leaf</strong> streak disease has had a strong impact on the economy <strong>of</strong> growers<br />

and on banana and plantain production since the first appearance in Cuba. The<br />

costs/ha <strong>of</strong> control <strong>of</strong> black <strong>leaf</strong> streak disease increased fourfold due to the increase<br />

<strong>of</strong> the number <strong>of</strong> fungicide treatments. The annual cost <strong>of</strong> fungicides in Cuba for<br />

black <strong>leaf</strong> streak disease control in the first year after disease outbreak in Cuba reached<br />

US$2 million.<br />

2. The area planted with ‘Cavendish’ cultivars and with plantains had been reduced<br />

to 15% and 18% respectively <strong>of</strong> the existing area previous to the introduction <strong>of</strong><br />

black <strong>leaf</strong> streak disease. At the same time, the area planted with FHIA hybrids that<br />

are resistant to black <strong>leaf</strong> streak disease is steadily increasing, leading to a 19%<br />

reduction in the amounts <strong>of</strong> fungicides imported during the two first years following<br />

the arrival <strong>of</strong> black <strong>leaf</strong> streak disease.<br />

3. The development <strong>of</strong> the disease in Cuba is highly correlated with the cumulated<br />

quantity and duration <strong>of</strong> the rain over a 14-day period, four weeks before in the<br />

case <strong>of</strong> ‘Cavendish’, and five weeks before in the case <strong>of</strong> plantains (AAB). These<br />

relationships can be useful for timing fungicide treatments in banana and plantains.<br />

4. From 1993 to 1997, bioclimatic warnings were used in the main ‘Cavendish’<br />

production plantations in Cuba, which allowed an optimization <strong>of</strong> control measures<br />

and a 40% reduction in the costs <strong>of</strong> protecting against black <strong>leaf</strong> streak disease.<br />

82


Session 1<br />

L. Pérez Vicente et al.<br />

5. Resistance <strong>of</strong> the different FHIA hybrids is expressed as a longer period <strong>of</strong><br />

incubation, a longer period <strong>of</strong> transition from streaks to <strong>spot</strong>s, and a reduction <strong>of</strong><br />

the production <strong>of</strong> pseudothecia, all <strong>of</strong> which are typical <strong>of</strong> partial resistance. The<br />

severity <strong>of</strong> black <strong>leaf</strong> streak disease on the cultivar ‘FHIA-18’ is currently higher in<br />

soils with low K contents.<br />

6. Studies are in progress to determine the potential for M. fijiensis populations to<br />

change pathogenicity as a result <strong>of</strong> the extensive planting <strong>of</strong> resistant hybrids.<br />

References<br />

Aguilar E.A., D.W. Turner and K. Sivasithamparam. 2000. Fusarium oxysporum f.sp. cubense<br />

inoculation and hypoxia alter peroxidase and phenylalanine ammonia lyase enzyme<br />

activities in nodal roots <strong>of</strong> banana cultivars (Musa sp.) differing in their susceptibility to<br />

Fusarium wilt. Australian Journal <strong>of</strong> Botany 48:589–596.<br />

Firman I.D. 1972. Susceptibility <strong>of</strong> banana cultivars to fungus <strong>diseases</strong> in Fiji. Tropical<br />

Agriculture Trinidad 49:189-196.<br />

Fouré E. 1988. Stratégies de lutte contre la Cercosporiose noire des bananiers et des plantains<br />

provoquée par <strong>Mycosphaerella</strong> fijiensis Morelet. L’avertissement biologique au Cameroun.<br />

Evaluation des possibilités d’amélioration. Fruits 43(5):269-274.<br />

Fouré E., M. Grisoni and R. Zurfluh. 1984. Les Cercosporioses du bananier et leurs traitements.<br />

Comportement des variétés. II. Etude de la sensibilité des bananiers et plantains<br />

á <strong>Mycosphaerella</strong> fijiensis Morelet et des quelques caractéristiques biologiques de la maladie<br />

des raies noires au Gabon. Fruits 39:365-378.<br />

Fouré E. A. Moulioum Pefoura and X. Mourichon. 1990. Etude de la sensibilité variétale des<br />

bananiers et des plantains à <strong>Mycosphaerella</strong> fijiensis Morelet au Cameroun. Caractérisation<br />

de la résistance au champ des bananiers appartenant à divers groupes génétiques. Fruits<br />

45:339-345.<br />

Ganry J. and J.P. Meyer. 1972a. La lutte contrôlée contre la Cercosporiose aux Antilles. Bases<br />

climatiques de l’avertissement. Fruits 27:665-676.<br />

Ganry J. and J.P. Meyer. 1972b. La lutte contrôlée contre le Cercospora aux Antilles.<br />

Application de techniques d’observation et numération de la maladie. Fruits 27:767-774.<br />

Hernandez A. and L. Perez. 2001. Reaction <strong>of</strong> banana and plantain cultivars to black Sigatoka<br />

disease caused by <strong>Mycosphaerella</strong> fijiensis, Morelet. Epidemiological components <strong>of</strong><br />

the resistance. Fitosanidad 5(3):9-15.<br />

Meredith D.S. and J.S. Lawrence. 1970. Black <strong>leaf</strong> streak <strong>of</strong> <strong>bananas</strong> (<strong>Mycosphaerella</strong> fijiensis).<br />

Susceptibility <strong>of</strong> cultivars. Tropical Agriculture Trinidad 47:275-287.<br />

Peng H.X., K. Sivasithamparam and D.W. Turner. 1999. Chlamydospore germination and<br />

Fusarium wilt <strong>of</strong> banana plantlets in suppressive and conducive soils are affected by<br />

physical and chemical factors. Soil Biology and Biochemistry 31(10):1363-1374.<br />

Pérez L. 1983. Epifitiología de la mancha de la hoja del plátano (Sigatoka) causada por<br />

<strong>Mycosphaerella</strong> musicola. Factores que influyen en el período de incubación y el desarrollo<br />

de la enfermedad en Cuba. Agrotecnia de Cuba 15(1):55–64.<br />

Pérez, L. 1989. Sistema de pronóstico climático fenológico de los tratamientos contra la<br />

mancha de la hoja (Sigatoka) causada por <strong>Mycosphaerella</strong> musicola en plátano fruta (Musa<br />

acuminata AAA). Agrotecnia de Cuba 21(2):35–46.<br />

Pérez L., 1998. Black Sigatoka disease control in banana and plantains plantations in Cuba.<br />

Management <strong>of</strong> the disease based on an integrated approach. INFOMUSA. Vol. 7(1):27-30.<br />

83


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Pérez L., C. Torres, M. Delgado and F. Mauri. 1981. Resistencia de diferentes clones de plátano<br />

a la Sigatoka causada por <strong>Mycosphaerella</strong> musicola Leach. Agrotecnia de Cuba<br />

13(2):51–66.<br />

Pérez L., F. Mauri, B. Barranco and G. García. 1993a. Efficacy <strong>of</strong> EBI’s fungicides in the<br />

control <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis Morelet on banana and plantains with treatments based<br />

on stage <strong>of</strong> evolution <strong>of</strong> the disease (biological warnings) in Cuba. P.55 in Proceedings<br />

<strong>of</strong> the 6 th International Congress <strong>of</strong> Plant Pathology, Montreal.<br />

Pérez L., F. Mauri, A. Hernández and A. Porras. 1993b. Efficacy <strong>of</strong> a biological warning system<br />

for timing fungicide treatments for the control <strong>of</strong> black Sigatoka disease (<strong>Mycosphaerella</strong><br />

fijiensis Morelet) in banana plantations in Cuba. Proceedings <strong>of</strong> the 6 th International<br />

Congress <strong>of</strong> Plant Pathology, Montreal.<br />

Pérez L., A. Hernández, A. Porras, E. Abreu, A. Guzmán, J. Montero, A. Méndez, H. Martínez,<br />

A. Aguirre y R. Pupo. 1997. Generalización del manejo integrado de Sigatoka negra en<br />

bananos y plátanos. Balance de cuatro años de su aplicación en áreas de producción.<br />

XII Fórum de Ciencia y Técnica.<br />

Pérez L., F. Mauri, A. Hernández, E. Abreu y A. Porras. 2000a. Epidemiología de la Sigatoka<br />

negra (<strong>Mycosphaerella</strong> fijiensis Morelet) en Cuba. Pronóstico bio-climático de los<br />

tratamientos en bananos (Musa acuminata AAA). Revista Mexicana de Fitopatología<br />

18(1):15–26.<br />

Pérez L., A. Hernández y A. Porras. 2000b. Epidemiología de la Sigatoka negra (<strong>Mycosphaerella</strong><br />

fijiensis Morelet) en Cuba. Pronóstico bio-climático de los tratamientos en plátanos (Musa<br />

spp. AAB). Revista Mexicana de Fitopatología 18:27–35.<br />

Porras A. y L. Pérez. 1997. Efecto de la temperatura en el crecimiento de los tubos germinativos<br />

de las ascósporas de <strong>Mycosphaerella</strong> spp. causantes de Sigatoka en plátanos. Cálculo de<br />

las sumas de velocidades de desarrollo para el pronóstico de los tratamientos a partir de<br />

la temperatura máxima y mínima diarias en Cuba. INFOMUSA 6(2):27-31.<br />

Stover R.H. and J.D. Dickson. 1970. Leaf <strong>spot</strong> <strong>of</strong> <strong>bananas</strong> caused by <strong>Mycosphaerella</strong> musicola<br />

Leach. Methods <strong>of</strong> measuring <strong>spot</strong>ting prevalence and severity. Tropical Agriculture<br />

Trinidad 47: 289-302.<br />

Vakili N.G. 1968. Response <strong>of</strong> Musa acuminata species and edible cultivars to infection by<br />

<strong>Mycosphaerella</strong> musicola. Tropical Agriculture Trinidad 45:13-22.<br />

Vidal A. 1992. Sigatoka negra en Cuba. En nuevos focos de plagas y enfermedades. Boletín<br />

Fitosanitario de la FAO 40:1-2.<br />

84


Session 1<br />

A.B. Molina and E. Fabregar<br />

Management <strong>of</strong> black <strong>leaf</strong> streak<br />

disease in tropical Asia<br />

A. B. Molina 1 and E. Fabregar 2<br />

Abstract<br />

In the Philippines, <strong>diseases</strong> are the major production constraint in the region. Leaf <strong>spot</strong> <strong>diseases</strong><br />

cause significant fruit yield and quality reduction in both commercial banana plantations and<br />

in small-scale farms. Black <strong>leaf</strong> streak disease is the most important banana <strong>leaf</strong> <strong>spot</strong> disease<br />

in the region. Although eumusae <strong>leaf</strong> <strong>spot</strong> disease has been reported in India, Sri Lanka,Thailand<br />

and Malaysia, the extent <strong>of</strong> damage has not yet been established. Sigatoka disease is no longer<br />

a major concern.The damage due to black <strong>leaf</strong> streak disease affects mostly smallholders as they<br />

generally do not implement any systematized disease management programme. Disease<br />

management in commercial plantations is based on the use <strong>of</strong> fungicides together with cultural<br />

practices that reduce the inoculum <strong>of</strong> the disease and therefore its development. Changes in<br />

the cropping/production system in the commercial plantations <strong>of</strong> Southern Philippines have had<br />

an impact on the management <strong>of</strong> <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>. The adoption <strong>of</strong> annual<br />

cropping reduced disease pressure and the need for year-round fungicide applications.<br />

Resumen - Estado del manejo de la Sigatoka negra en Asia Tropical<br />

El las Filipinas, las enfermedades representan la principal limitación en la región. Las enfermedades<br />

de las manchas foliares provocadas por el género <strong>Mycosphaerella</strong> causan una reducción<br />

significativa del rendimiento y calidad de la fruta tanto en las plantaciones comerciales como<br />

en pequeñas fincas bananeras. La Sigatoka negra es el problema de las manchas foliares más<br />

importante en la región. Aunque se ha reportado la presencia de <strong>Mycosphaerella</strong> eumusae (mancha<br />

foliar eumusae) en India, Sri Lanka, Tailandia y Malasia, la dimensión de los daños que ella causa<br />

aún no se ha establecido. La Sigatoka amarilla ya no representa la principal preocupación. El daño<br />

debido a la Sigatoka negra se observa principalmente al nivel de los pequeños agricultores, ya<br />

que, básicamente, ellos no implementan ningún programa de manejo sistematizado en<br />

comparación con las plantaciones comerciales. El manejo de las enfermedades empleado en las<br />

plantaciones comerciales es un programa basado en fungicidas reforzado por prácticas culturales<br />

que reduce el inóculo de la enfermedad y reduce así su desarrollo. Los cambios en los sistemas<br />

de cultivo y producción tuvieron cierto impacto sobre el manejo de la mancha foliar de Sigatoka<br />

en las plantaciones comerciales en el sur de Filipinas. La adopción del cultivo anual redujo la presión<br />

de la enfermedad y así se evita la necesidad de la aplicación de fungicidas durante todo el año.<br />

1<br />

INIBAP-Asia and the Pacific, Los Baños, Philippines<br />

2<br />

Lapanday Fruits Development Corporation, Philippines<br />

85


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Résumé - Gestion de la maladie des raies noires en Asie tropicale<br />

Aux Philippines, les maladies constituent la principale contrainte de la production. Les maladies<br />

foliaires causées par les <strong>Mycosphaerella</strong> réduisent le rendement et la qualité des fruits dans les<br />

plantations commerciales, ainsi que dans les petites fermes. Dans la région, la maladie des raies<br />

noires est la plus importante maladie foliaire attaquant la banana. Bien que l’ELSD (eumusae <strong>leaf</strong><br />

<strong>spot</strong> disease) ait été observée en Inde, au Sri Lanka, en Thaïlande et en Malaisie, l’étendue des<br />

dégâts n’a pas encore été établie. La maladie de Sigatoka n’est plus considérée comme importante.<br />

Les dégâts dus à la maladie des raies noires affectent principalement des petits fermiers car ils<br />

ne peuvent généralement pas mettre en place de programme systématique de gestion des<br />

maladies. La gestion des maladies dans les grandes plantations est basée sur l’utilisation de<br />

fongicides ainsi que sur des pratiques culturales visant à réduire la quantité d’inoculum et donc<br />

le développement de la maladie. Des changements dans le système de production dans les<br />

plantations commerciales du sud des Philippines ont eu un impact sur la gestion des maladies<br />

foliaires causées par les <strong>Mycosphaerella</strong>. Le fait d’avoir adopté une culture annuelle a réduit la<br />

pression de la maladie et le recours aux fongicides tout au long de l’année.<br />

Introduction<br />

Banana (Musa spp.), a group <strong>of</strong> plants that comprises many different types <strong>of</strong> sweet<br />

dessert <strong>bananas</strong>, cooking <strong>bananas</strong> and plantains, is an important fruit crop in Asia.<br />

Bananas are grown largely by smallholder farmers, traded by local entrepreneurs<br />

and consumed locally. Thus, banana plays a major role in food security and is<br />

a source <strong>of</strong> income for the rural poor. The Philippines is the only major banana-exporting<br />

country in Asia and <strong>bananas</strong> generate important foreign exchange. In summary, banana<br />

is an important food and source <strong>of</strong> income for local farmers and for the region.<br />

The main constraints to banana production and threats to the industry in the region<br />

are from pests and <strong>diseases</strong>. The region is the centre <strong>of</strong> origin <strong>of</strong> Musa, and hence many<br />

serious pests and <strong>diseases</strong> affect the crop, e.g. <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> caused by <strong>Mycosphaerella</strong><br />

spp. are responsible for a reduction in fruit yield and quality in commercial plantations<br />

and on small-scale farms.<br />

Three species <strong>of</strong> <strong>Mycosphaerella</strong> are present in Asia. M. fijiensis, responsible for black<br />

<strong>leaf</strong> streak disease, is the most important pathogen because <strong>of</strong> its wide distribution and<br />

its virulence which gives rise to epidemics. M. eumusae, a newly reported pathogen, is<br />

potentially devastating and has been reported in India, Sri Lanka, Thailand and Malaysia<br />

but not, so far, in the Philippines. M. musicola, the causal agent <strong>of</strong> Sigatoka disease,<br />

is present in Southeast Asia but is no longer <strong>of</strong> major importance.<br />

In commercial plantations, even if the disease does not affect yield, it can still reduce<br />

fruit quality and render the fruit unfit for export. Leaf <strong>spot</strong> <strong>diseases</strong> also cause fruits<br />

to ripen prematurely during transport to the market.<br />

Disease management practices<br />

Banana production systems in Southeast Asia are classified as follows (1) backyard,<br />

(2) mixed crop, (3) commercial smallholder plantation and (4) large commercial<br />

export-oriented plantation. The first three are intended for local markets and<br />

production is on a small scale. Management <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> varies according<br />

to the system <strong>of</strong> production.<br />

86


Session 1<br />

A.B. Molina and E. Fabregar<br />

Small-scale farming<br />

Leaf <strong>spot</strong> disease management by small-scale banana growers ranges from minimal<br />

to none, that is there is no systematic management <strong>of</strong> the disease. The disease reduces<br />

fruit size on susceptible cultivars but the fruits are still acceptable to the local market.<br />

Local consumers are not exacting in terms <strong>of</strong> fruit quality, size and ripening<br />

characteristics, unlike the commercial export market. Thus, small-scale banana<br />

growers make a certain amount <strong>of</strong> pr<strong>of</strong>it even when <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> are not controlled.<br />

In addition, small-scale banana growers plant many varieties <strong>of</strong> banana. Figure 1<br />

shows the various local cultivars grown in the Philippines and their proportion relative<br />

to total production. This variety <strong>of</strong> cultivars caters to the local demand for different<br />

uses or consumption <strong>of</strong> <strong>bananas</strong>. Some varieties are used cooked, processed or as fresh<br />

banana. The planting <strong>of</strong> different varieties provides genetic diversity against black <strong>leaf</strong><br />

streak disease. Several important local varieties, e.g. ‘Saba’ and ‘Pelipita’ in the<br />

Philippines and ‘Pisang kepok’ in Indonesia, are resistant to <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong>. Hence, for banana growers who specialize in cooking <strong>bananas</strong>, <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> are not important. ‘Lakatan’, ‘Pisang berangan’, and some ‘Cavendish’<br />

varieties are susceptible to black <strong>leaf</strong> streak disease, but still produce yields that are<br />

acceptable to the local market.<br />

Lakatan<br />

13%<br />

Latundan<br />

8%<br />

Bungulan<br />

5%<br />

Others<br />

3%<br />

Saba<br />

39%<br />

Cavendish-type<br />

32%<br />

Figure 1. Most popular banana cultivars grown in the Philippines.<br />

Chemical control is not practised by small-scale growers because <strong>of</strong> the expense.<br />

Better-<strong>of</strong>f local growers may remove severely diseased leaves as a means <strong>of</strong> reducing<br />

inoculum and infection but none use fungicides. The variety <strong>of</strong> cultivars planted and<br />

the diversity <strong>of</strong> growing conditions reduce the risk <strong>of</strong> disease in comparison with largescale<br />

monocrop commercial plantations.<br />

Commercial plantations<br />

A high level <strong>of</strong> disease control is required in commercial plantations for the export<br />

market. In Southeast Asia, only the Philippines export large quantities <strong>of</strong> <strong>bananas</strong>.<br />

87


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

The high standards <strong>of</strong> fruit quality and the characteristically large-scale monocultures<br />

used in the production <strong>of</strong> banana for export require intensive disease management.<br />

Thus, the cultivation <strong>of</strong> a single banana variety ‘Cavendish’ and monocropping make<br />

disease management in commercial plantations very challenging. However, the value<br />

<strong>of</strong> the export market supports the use <strong>of</strong> expensive disease management practices.<br />

The high levels <strong>of</strong> disease control achieved by a mixture <strong>of</strong> fungicides and cultural<br />

practices reduce inoculum and conditions that favour disease development. In the<br />

Philippines, fungicides may be applied up to 40 times per year (Table 1). The control<br />

programme includes contact fungicides, which are alternated or mixed with systemic<br />

fungicides applied in water or as an oil-water emulsion. The fungicides and doses used<br />

by plantation owners are approved by the local regulatory agencies. In addition, the<br />

fungicide doses are within the tolerances permitted by the importing countries. Table<br />

2 lists the fungicides used in a typical commercial plantation in the Philippines.<br />

Table 1. Spray programme <strong>of</strong> LADC, Davao, Philippines.<br />

Number <strong>of</strong> cycles<br />

Fungicide 1997 1998 1999 2000 2001<br />

Strobilurin 0 0 4 6 6<br />

Triazole 9 11 8 8 8<br />

Dithane/Vondoze/Maneb 4 4 11 12 12<br />

Daconil 11 8 7 6 5<br />

Calixin 7 9 8 8 7<br />

Total cycles 31 32 38 40 38<br />

Oil used (L//ha/yr) 76 82 147 119 130<br />

Table 2. Fungicides used by commercial growers in the Philippines.<br />

Trade name Chemical name Volume/hectare<br />

Systemic<br />

Bankit 25 SC Azoxystrobin 0.4 L<br />

Baycor 300 EC Bitertanol 0.5 L<br />

Bumper 25 EC Propiconazole 0.4 L<br />

Calixin 75 EC Tridemorph 0.6 L<br />

Folicur 250 EC Tebuconazole 0.23 L<br />

Indar 2F Fenbuconazole 0.4 L<br />

Siganex Pyrazophos 0.5 L<br />

Sico 250 EC Propiconazole 0.4 L<br />

Tega 075 EC Trifloxystrobin 1.0 L<br />

Tilt 250 EC Propiconazole 0.4 L<br />

Protectant<br />

Daconil 720 SC Chlorothalonil 1.38 L<br />

Dithane F448 Mancozeb 4.0 L<br />

Dithane M-45 Mancozeb 2.5 kg<br />

Dithane OS Mancozeb 1.75 L<br />

Vondozeb 42 SC Mancozeb 4.0 L<br />

The high dependence on the very specific fungicide propiconazole in the late 80s<br />

to early 90s resulted in a considerable increase in insensitivity <strong>of</strong> <strong>Mycosphaerella</strong>.<br />

88


Session 1<br />

A.B. Molina and E. Fabregar<br />

However, the loss <strong>of</strong> effectiveness was less and occurred later than in Central America.<br />

The prolonged effectiveness in the Philippines is possibly because spray programmes<br />

have always been based on the principle <strong>of</strong> rotation and/or combinations <strong>of</strong> fungicides,<br />

avoiding block or consecutive applications <strong>of</strong> the same products. The fungicide Benomyl<br />

(Benlate) was used unwisely in the late 70s and resulted in a high degree <strong>of</strong> fungicide<br />

resistance. Since then, Benlate has been withdrawn from use in the Philippines. The<br />

introduction <strong>of</strong> newer chemicals such as Azoxystrobin has provided a much-needed<br />

opportunity for fungicide control in combination with triazoles.<br />

Commercial plantations also integrate cultural and other practices to their disease<br />

management in order to reduce conditions that favour disease development. Good<br />

drainage and irrigation practices receive attention. The removal and destruction <strong>of</strong><br />

severely affected leaves also reduce inoculum.<br />

Monitoring <strong>of</strong> disease severity on plants with or without a flower is regularly<br />

practiced in commercial plantation. In some plantations, early detection and<br />

quantification <strong>of</strong> symptoms are also done. The data are used to schedule fungicide<br />

treatments as well as a guide for harvest to reduce the risk <strong>of</strong> the effects <strong>of</strong> <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> on fruit quality. Frequently, when disease is severe and reduces the numbers<br />

<strong>of</strong> functional leaves, fruits are harvested earlier to avoid the risk <strong>of</strong> premature fruit<br />

ripening during transit to the market. This also removes a source <strong>of</strong> inoculum from<br />

the plantation.<br />

Changes in cropping/production systems have also had important effects on the<br />

management <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in commercial plantations in the southern Philippines.<br />

Adoption <strong>of</strong> annual cropping reduced disease pressure and the need for year round<br />

fungicide application. In Taiwan, where annual cropping was introduced more than<br />

a decade ago, black <strong>leaf</strong> streak disease is no longer a problem. Expanded banana<br />

commercial production in the Philippines included a few thousand plantations <strong>of</strong><br />

‘Lakatan’ to supply the lucrative local market, but not controlling <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

as intensely is providing an abundant source <strong>of</strong> inoculum for the export ‘Cavendish’<br />

plantations, making control programmes more difficult. The export market has also<br />

increased considerably during the last decade (Figure 2). The increased area devoted<br />

to monoculture crops has increased the risk <strong>of</strong> epidemics.<br />

Use <strong>of</strong> resistant varieties<br />

As mentioned before, some popular local cultivars are resistant to black <strong>leaf</strong> streak<br />

disease. However, many <strong>of</strong> the acuminata dessert type <strong>bananas</strong>, e.g. ‘Cavendish’, ‘Lakatan’<br />

and‘Pisang berangan’, are susceptible to black <strong>leaf</strong> streak disease.<br />

Several <strong>of</strong> the varieties provided by INIBAP through its International Musa Testing<br />

Programme (IMTP) proved to be resistant to black <strong>leaf</strong> streak disease in Southeast Asia.<br />

In particular, the FHIA series are very resistant to this disease in field trials. The high<br />

yield potential and disease resistance <strong>of</strong> these varieties make them attractive to tropical<br />

Asian farmers and consumers. It is worth noting that, being the centre <strong>of</strong> origin <strong>of</strong><br />

banana, Asian consumers in the tropics are used to eating different kinds <strong>of</strong> banana<br />

<strong>of</strong> different size, colour and taste. Moreover, Asians prepare <strong>bananas</strong> for a variety <strong>of</strong><br />

uses. Hence, disease-resistant, high-yielding hybrids have the potential to increase<br />

productivity <strong>of</strong> <strong>bananas</strong> in tropical Asia.<br />

89


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

120<br />

100<br />

80<br />

Million boxes<br />

50<br />

40<br />

20<br />

0<br />

1985 1990 1995 1998 1999 2000 2001<br />

Year<br />

Figure 2. Number <strong>of</strong> 13-kg boxes exported by the Philippines.<br />

Literature consulted<br />

Carlier J., M.F. Zapater, F. Lapeyre, D.R. Jones and X. Mourichon. 2000. Septoria <strong>leaf</strong> <strong>spot</strong> <strong>of</strong><br />

banana: A newly discovered disease caused by <strong>Mycosphaerella</strong> eumusae (anamorph<br />

Septoria eumusae). Phytopathology 90:884:890.<br />

Jones D.R., S.H. Jamaluddin and N.H. Nik Masdek. 1994. Banana disease survey <strong>of</strong> west<br />

Malaysia. Pp. 49-62 in Proceedings <strong>of</strong> the Fourth Meeting <strong>of</strong> the Regional Advisory<br />

Committee <strong>of</strong> INIBAP-Asia and the Pacific Network held in Taiwan Banana Research<br />

Institute, Chiuju, Pingtung, Taiwan, November 21-25, 1994.<br />

Magnaye, L.V. and L.E. Herradura. 1995. Rescue and Conservation <strong>of</strong> the Southeast Asian<br />

Regional Banana Germplasm Collection (A Terminal Report <strong>of</strong> the INIBAP/IPGRI-funded<br />

project <strong>of</strong> the same title). 47pp.<br />

Molina G.C. and V.N. Villegas. 2001. Etiology and survey <strong>of</strong> banana <strong>leaf</strong> <strong>spot</strong> in the Philippines.<br />

A final report <strong>of</strong> a collaborative project submitted to INIBAP Asia and the Pacific.<br />

Qi Pei-Kun, Jiang Zi-De and Xi Ping-Gen. 2001. Etiology and preliminary survey <strong>of</strong> banana<br />

<strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in Guangdong Province in China. A final report <strong>of</strong> a collaborative project<br />

submitted to INIBAP-Asia and the Pacific. College <strong>of</strong> resources and environmental sciences,<br />

South China Agricultural University, Guangzhou, China.<br />

Selvarajan R., S. Uma and S. Sathiamoorthy. 2000. Etiology and survey <strong>of</strong> banana <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> in India. Pp. 94-102 in Advancing banana and plantain R&D in Asia and the<br />

Pacific, Vol. 10 (A.B. Molina, V.N. Roa and M.A.G. Maghuyop, eds) INIBAP-ASPNET, Los<br />

Baños, Laguna, Philippines.<br />

Sirisena J.A. 1997. Status <strong>of</strong> banana production in Sri Lanka. Pp. 160-180 in Proceedings <strong>of</strong><br />

the seventh meeting <strong>of</strong> the Regional Advisory Committee <strong>of</strong> INIBAP-ASPNET held in VASI,<br />

Hanoi, Vietnam, October 21-23, 1997.<br />

90


Session 1<br />

Z.J. Maciel Cordeiro and A. Pires de Matos<br />

Impact <strong>of</strong> <strong>Mycosphaerella</strong> spp.<br />

in Brazil<br />

Z.J. Maciel Cordeiro and A. Pires de Matos<br />

Abstract<br />

Brazil ranks second in the world for banana production. Bananas are grown throughout the country,<br />

mainly by smallholders. Of the <strong>Mycosphaerella</strong> species present in Brazil, M. musicola (anamorph<br />

Pseudocercospora musae), the causal agent <strong>of</strong> Sigatoka disease, and M. fijiensis (anamorph<br />

Pseudocercospora fijiensis), the causal agent <strong>of</strong> black <strong>leaf</strong> streak disease, are the most serious. They<br />

can cause a 100% yield loss on susceptible varieties from the Cavendish (Musa cv. AAA) and Prata<br />

(Musa cv. AAB) groups. Sigatoka disease is present in all banana growing areas <strong>of</strong> the country. Yield<br />

loss is dependent on environmental conditions but is estimated at 50% on average. Black <strong>leaf</strong> streak<br />

disease is still confined to the north region <strong>of</strong> the country plus the State <strong>of</strong> Mato Grosso (centrewest<br />

region). It causes a 100% yield loss in dessert <strong>bananas</strong> and around a 70% yield loss in plantain,<br />

a crop which is very important as a staple food in northern Brazil. Since 1999, susceptible cultivars<br />

have been gradually replaced by resistant cultivars, e.g.‘Caipira’ (AAA),‘Thap maeo’ (AAB),‘FHIA-18’<br />

(AAAB) and ‘Pacovan Ken’(AAAB), especially in the State <strong>of</strong> Amazonas, following the recommendations<br />

<strong>of</strong> the Brazilian agriculture research corporation (Embrapa).Traditional varieties have been replaced<br />

by cultivars resistant to black <strong>leaf</strong> streak disease because <strong>of</strong> the high yield losses and the lack <strong>of</strong><br />

opportunity for growers to use fungicides. In addition there is the socio-economic impact <strong>of</strong> the<br />

ban by the Federal and State authorities on the transport <strong>of</strong> banana fruits from infected areas,<br />

designed to prevent the spread <strong>of</strong> black <strong>leaf</strong> streak disease to other banana growing regions <strong>of</strong> the<br />

country.<br />

Resumen - Impacto de <strong>Mycosphaerella</strong> spp. en el banano en Brasil<br />

La producción bananera es una actividad muy importante en Brasil, el segundo productor del<br />

mundo de este cultivo. El banano se cultiva en todo el país, principalmente por pequeños<br />

agricultores. Entre las especies de <strong>Mycosphaerella</strong> que se dan en Brasil, M. musicola (anamorpho<br />

Pseudocercospora musae), agente causal de la Sigatoka amarilla, y M. fijiensis (anamorpho<br />

Paracercospora fijiensis), agente causal de la Sigatoka negra, son las más serias. Ellas han sido<br />

responsables por pérdidas de rendimiento de hasta 100% en las variedades susceptibles como<br />

las que pertenecen a los grupos ‘Cavendish’ (Musa cv. AAA) y Prata (Musa cv. AAB). La Sigatoka<br />

amarilla se encuentra en todas las áreas productoras de banano del país causando pérdidas<br />

de rendimiento que varía de acuerdo a las condiciones ambientales prevalecientes. En<br />

Embrapa Mandioca e Fruticultura, Cruz das Almas, Brazil<br />

91


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

promedio, las pérdidas debido a la Sigatoka amarilla en Brasil se estiman en un 50%. La Sigatoka<br />

negra está aún confinada a la región norte, más, el estado de Mato Grosso (región central<br />

occidental). Además de causar pérdidas de rendimiento del 100% en los bananos de postre,<br />

la Sigatoka negra también causa pérdidas de rendimiento de un 70% en plátano, muy<br />

importante para la región norte de Brasil. En esta región, especialmente en el estado de<br />

Amazonas, a partir de 1999, los cultivares susceptibles han sido reemplazados gradualmente<br />

por los cultivares resistentes, recomendados por la Corporación Brasileña de Investigación<br />

Agrícola (Embrapa) como: ‘Caipira’ (AAA), ‘Thap maeo’ (AAB), ‘FHIA-18’ (AAAB) y ‘Pacovan Ken’<br />

(AAAB). El reemplazo de las variedades tradicionales fue forzado tanto por las altas pérdidas<br />

de rendimiento, como por la inhabilidad de los productores de insertar los agroquímicos como<br />

parte del sistema de producción. Además de la sostenibilidad ecológica de variedades<br />

resistentes, también se observan los impactos socioeconómicos en las acciones de las<br />

organizaciones federales y estatales que previenen el transporte de las frutas de banano de<br />

las áreas infectadas, dirigidas a proteger las regiones del país libres de las enfermedades<br />

producto de la diseminación de la Sigatoka negra.<br />

Résumé - Impact des <strong>Mycosphaerella</strong> spp. au Brésil<br />

Le Brésil est le deuxième producteur mondial de banane. Les bananiers sont cultivés dans tout<br />

le pays surtout par des petits producteurs. De toutes les espèces de <strong>Mycosphaerella</strong> présentes<br />

au Brésil, M. musicola (anamorphe Pseudocercospora musae), l’agent causal de la maladie de<br />

Sigatoka, et M. fijiensis (anamorphe Pseudocercospora fijiensis),l’agent causal de la maladie des<br />

raies noires, sont les plus sérieuses. Elles peuvent provoquer jusqu’à 100% de perte de production<br />

sur des variétés sensibles comme celles des groupes Cavendish (Musa cv. AAA) et Prata (Musa<br />

cv. AAB). La maladie de Sigatoka est présente dans toutes les régions productrices de bananes<br />

du pays. La réduction du rendement dépend des conditions environnementales mais en<br />

moyenne elle est estimée à 50%. La maladie des raies noires est encore confinée dans la région<br />

Nord du pays ainsi que dans l’état du Mato Grosso (région centre-ouest). Elle provoque des pertes<br />

de 100% chez la banane dessert et d’environ 70% chez la banane plantain, un aliment de base<br />

important dans le nord du Brésil. Depuis 1999, les cultivars sensibles ont été progressivement<br />

remplacés par des cultivars résistants, par exemple ‘Caipira’ (AAA),‘Thap maeo’ (AAB),‘FHIA-18’<br />

(AAAB) et ‘Pacovan Ken’ (AAAB), particulièrement dans l’état de l’Amazone suivant les<br />

recommandations de la société brésilienne de recherche en agriculture (Embrapa). Étant donné<br />

les réductions importantes du rendement et l’impossibilité qu’ont les cultivateurs à utiliser des<br />

fongicides, les variétés traditionnelles ont été remplacées par des cultivars résistants à la maladie<br />

des raies noires. De plus, l’interdiction, par les autorités fédérales, de transporter les bananes<br />

provenant de zones infectées, afin de limiter la propagation de la maladie des raies noires vers<br />

d’autres régions du pays, a un impact socioéconomique certain.<br />

Introduction<br />

The cultivation <strong>of</strong> banana in Brazil plays an important economic and social role in large<br />

and small plantations. Smallholders grow <strong>bananas</strong> mainly as a subsistence crop. Plantains<br />

are grown as a staple food, mainly in the north and northeastern regions <strong>of</strong> the country.<br />

Brazil is the world’s second largest producer <strong>of</strong> <strong>bananas</strong> and production is estimated<br />

at six million metric tons per year. The total area cultivated with banana is about<br />

533 000 hectares, distributed throughout the country: 8% in the south, 28% in the<br />

southeast, 8% in the center west, 32% in the northeast and 24% in the north.<br />

Of the various phytosanitary problems that affect banana production in Brazil, the<br />

most important are those related to <strong>leaf</strong> <strong>diseases</strong> caused by <strong>Mycosphaerella</strong> musicola<br />

and M. fijiensis, wilt <strong>diseases</strong> caused by Fusarium oxysporum f. sp. cubense and Ralstonia<br />

92


Session 1<br />

Z.J. Maciel Cordeiro and A. Pires de Matos<br />

solanacearum race 2, and s<strong>of</strong>t rot caused by Erwinia carotovora. Nematodes, mainly<br />

Radopholus similis, also affect the banana crop, and the weevil borer Cosmopolites<br />

sordidus is perhaps the most important insect pest. Due to the increasing demands by<br />

the consumer for fruits <strong>of</strong> better quality and appearance, pre and postharvest fruit <strong>diseases</strong><br />

and damage to fruit by arthropods, e.g. thrips, have become increasingly important.<br />

Because they reduce the quantity and quality <strong>of</strong> <strong>bananas</strong>, and are difficult and expensive<br />

to control, <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> caused by M. musicola and M. fijiensis are considered to<br />

be the most serious <strong>diseases</strong> affecting the Brazilian banana industry.<br />

Impact <strong>of</strong> <strong>Mycosphaerella</strong> musicola<br />

Sigatoka disease, caused by M. musicola Leach (anamorph Pseudocercospora musae<br />

(Zumm) Deighton) was first reported in Java, in 1902. Sigatoka disease is found in all<br />

banana growing regions except for Israel, Egypt and the Canary Islands. The disease<br />

was first reported in 1944 in the Amazon region <strong>of</strong> Brazil. Eight years later, the disease<br />

was found in the southeast region, the largest banana production region <strong>of</strong> Brazil. Today,<br />

Sigatoka disease is widespread throughout Brazil causing severe yield losses in those<br />

regions where environmental conditions favour disease development.<br />

In Brazil, <strong>Mycosphaerella</strong> musicola results in an average yield loss <strong>of</strong> about 50%<br />

which may be higher in particular regions <strong>of</strong> the country. In banana orchards affected<br />

by Sigatoka disease, the number <strong>of</strong> bunches and <strong>of</strong> hands per bunch is lower and the<br />

fruits, in addition to being smaller and lighter, ripen prematurely. A high incidence <strong>of</strong><br />

Sigatoka disease causes early decline <strong>of</strong> the banana orchard. Plant vigour declines and<br />

yield is reduced in later crop cycles (Cordeiro and Matos, 2001) and orchards have to<br />

be replanted at shorter intervals than usual.<br />

According to Cordeiro (1990), by the end <strong>of</strong> the 80s, the cost <strong>of</strong> control using five<br />

applications per year <strong>of</strong> systemic fungicides plus mineral oil accounted for 9% <strong>of</strong> the<br />

total production cost estimated at US$1350/ha/year. In recent years, the cost has<br />

increased because <strong>of</strong> the need to increase the numbers <strong>of</strong> applications to seven per<br />

year. Today, the cost <strong>of</strong> controlling Sigatoka disease amounts to about 10% <strong>of</strong> the<br />

total production cost. In areas where fungicide control measures are not used, yield<br />

losses can exceed 80%<br />

and fruit quality is very<br />

poor (Figure 1).<br />

Figure 1.<br />

Damage from Sigatoka<br />

disease on ‘Prata aña’ in the<br />

absence <strong>of</strong> control measures.<br />

93


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Impact <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

Black <strong>leaf</strong> streak disease was first reported in 1963, in the Fiji Islands, district <strong>of</strong><br />

Sigatoka. In America, the disease was first reported in Honduras in 1972, where it<br />

was called black Sigatoka. The disease spread from Honduras throughout the banana<br />

growing areas <strong>of</strong> Central and South America. In Brazil, black <strong>leaf</strong> streak disease was<br />

first discovered in 1998 (Pereira et al., 1998; Cordeiro et al., 1998), in banana orchards<br />

located in the municipalities <strong>of</strong> Tabatinga and Benjamim Constant, State <strong>of</strong><br />

Amazonas, on the borders <strong>of</strong> Brazil, Colombia and Peru.<br />

Distribution <strong>of</strong> black <strong>leaf</strong> streak disease in Brazil<br />

Following the discovery <strong>of</strong> black <strong>leaf</strong> streak disease in Brazil, surveys were carried out<br />

to follow its spread. By the end <strong>of</strong> 1998, black <strong>leaf</strong> streak disease was found in the<br />

State <strong>of</strong> Acre, probably introduced accidentally from Bolivia. A survey in early 1999,<br />

revealed high levels <strong>of</strong> black <strong>leaf</strong> streak disease in Rodônia, in the northern region <strong>of</strong><br />

the state, and Mato Grosso, in the central western region (Cordeiro et al., 2000). Recent<br />

surveys in the northern region <strong>of</strong> Brazil showed that the disease had spread throughout<br />

the region, e.g. the States <strong>of</strong> Pará, Amapá, Roraima, Amazonas, Acre and Rondônia<br />

(Gasparotto et al. 2001a). Figure 2 shows the distribution <strong>of</strong> black <strong>leaf</strong> streak disease<br />

in northern Brazil, inclu0ding the State <strong>of</strong> Mato Grosso, and the central western region.<br />

Phytosanitary defence strategy<br />

The yellow band in Figure 2 corresponds to a buffer zone, and includes the States<br />

<strong>of</strong> Mato Grosso do Sul, Goiás, Tocantins and Maranhão, where a phytosanitary<br />

defence strategy has been put in place to delay as much as possible the spread<br />

<strong>of</strong> M. fijiensis to other banana growing areas currently free <strong>of</strong> the disease<br />

(Figure 2). Planned actions include training for technicians and banana growers<br />

with emphasis on the recognition <strong>of</strong> the symptoms <strong>of</strong> black <strong>leaf</strong> streak disease,<br />

field evaluations <strong>of</strong> resistant varieties, e.g. ‘Caipira’ (AAA), ‘Thap Maeo’ (AAB),<br />

‘FHIA-18’ (AAAB), ‘Pacovan Ken’ (AAAB) which are recommended by Embrapa.<br />

Dependent on the agreement <strong>of</strong> growers, orchards in the buffer zone will be<br />

replanted with varieties resistant to black <strong>leaf</strong> streak disease in order to establish<br />

a barrier to prevent the spread <strong>of</strong> the pathogen. The movement <strong>of</strong> <strong>bananas</strong> from<br />

areas affected by black <strong>leaf</strong> streak disease to disease-free areas is restricted; a<br />

phytosanitary certificate <strong>of</strong> origin is required before permitting transport <strong>of</strong> fruits<br />

to other areas.<br />

Impact <strong>of</strong> black <strong>leaf</strong> streak disease<br />

It is not yet possible to quantify exactly the impact <strong>of</strong> black <strong>leaf</strong> streak disease<br />

on banana production in Brazil. In general, banana cultivation is characterized<br />

by a low level <strong>of</strong> technology, mainly subsistence cultivation without the use <strong>of</strong><br />

fungicidal control measures. Nevertheless, field evaluations <strong>of</strong> fungicides to control<br />

black <strong>leaf</strong> streak disease in the Amazon region have shown that approximately<br />

94


Session 1<br />

Z.J. Maciel Cordeiro and A. Pires de Matos<br />

26 applications with systemic fungicide and up to 52 applications <strong>of</strong> protectant<br />

fungicides would be necessary to give adequate control <strong>of</strong> black <strong>leaf</strong> streak disease<br />

in areas where weather favours the disease (Gasparotto et al., 2001b; Pereira and<br />

Gasparotto, 2001). The large numbers <strong>of</strong> fungicide applications are ten times higher<br />

than the total fungicide applied for the control <strong>of</strong> Sigatoka disease in some areas<br />

<strong>of</strong> Brazil. In areas with clear dry periods, 15 fungicide applications give sufficient<br />

control. Even in these conditions, the numbers <strong>of</strong> applications is twice that needed<br />

to control Sigatoka disease.<br />

Figure 2. Distribution <strong>of</strong> black <strong>leaf</strong> streak disease in Brazil (April 2002).<br />

The biggest impact <strong>of</strong> black <strong>leaf</strong> streak disease in the Amazon region is the<br />

need to change the varieties that are cultivated. According to growers <strong>of</strong> ‘Prata<br />

comum’ in the Amazon region, banana production fell almost to zero after the<br />

arrival <strong>of</strong> black <strong>leaf</strong> streak disease in the municipalities <strong>of</strong> Benjamim Constant<br />

and Tabatinga in the State <strong>of</strong> Amazonas. The only way to continue banana<br />

production in the area would be the use <strong>of</strong> varieties resistant to black <strong>leaf</strong> streak<br />

disease. To support the use <strong>of</strong> resistant varieties, the government should acquire<br />

micropropagated plantlets for distribution to growers. At present, however, a clear<br />

decision has yet to be made about the preferred resistant cultivar, based on<br />

information about yield performance and consumer acceptance. Since 1999, the<br />

Embrapa tissue culture laboratory has sold 1 384 003 plantlets <strong>of</strong> varieties resistant<br />

to black <strong>leaf</strong> streak disease to the states affected by the disease (Table 1).<br />

95


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Table 1. Varieties resistant to black <strong>leaf</strong> streak disease and number <strong>of</strong> plantlets sold to the area <strong>of</strong> incidence <strong>of</strong> the<br />

disease in Brazil.<br />

Variety Number <strong>of</strong> plantlets Number contracted Total number<br />

sold by 2001 from 2002 <strong>of</strong> plantlets<br />

Thap maeo (AAB) 301 968 350 000 651 968<br />

Caipira (AAA) 648 100 200 000 848 100<br />

FHIA-01 (AAAB) 42 000 - 42 000<br />

FHIA-18 (AAAB) 163 735 300 000 463 735<br />

SH36-40 (AAAB) 128 150 - 128 150<br />

FHIA-21 (AAAB) 9 000 - 9 000<br />

FHIA-03 (AABB) 5 900 - 5 900<br />

FHIA-10 3 000 - 3 000<br />

Ouro (AA) 4 650 - 4 650<br />

Prata zulu (ABB) 3 000 150 000 150 000<br />

PV03-44 (AAAB) 74 500 - 74 500<br />

PV42-85 (AAAB) 50 - 50<br />

Total 1 384 053 1 000 000 2 384 053<br />

Research activities<br />

Before the discovery <strong>of</strong> black <strong>leaf</strong> streak disease in Brazil, all banana genotypes<br />

generated by Embrapa’s breeding programme were sent to Costa Rica to evaluate<br />

their resistance in collaboration with CATIE, INIBAP and CORBANA. At present,<br />

genotypes have been evaluated in the states <strong>of</strong> Acre and Amazonas, in<br />

collaboration with Embrapa Acre and Embrapa Western Amazon. In addition to<br />

supporting the banana breeding programme in Brazil, the knowledge generated<br />

by researchers has also helped the phytosanitary defence strategy. One <strong>of</strong> the most<br />

important contributions to fight black <strong>leaf</strong> streak disease in Brazil has been the<br />

delivery <strong>of</strong> resistant varieties. In addition, experiments have showed that spores<br />

<strong>of</strong> M. fijiensis can survive for 60 days on several types <strong>of</strong> surface including clothes,<br />

fruits, wood and iron (Hanada et al., 2000). That observation gave strong support<br />

to the legal actions that ban the transport and sale <strong>of</strong> fruits from affected to disease<br />

free areas.<br />

References<br />

Cordeiro Z.J.M. 1990. Economic impact <strong>of</strong> Sigatoka disease in Brazil. Pp. 56-60 in Sigatoka<br />

<strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong> (R.H. Stover and R. Fullerton, eds). Proceedings <strong>of</strong> an<br />

international workshop, San José, Costa Rica, March 28–April 1,1989.<br />

Cordeiro Z.J.M and A.P. de Matos. 2001. Sigatoka-amarela no Norte de Minas Gerais. Simpósio<br />

Norte Mineiro sobre a cultura da banana, Nova Porteirinha, 6-9 novembro. Anais I,<br />

pp. 238-247.<br />

Cordeiro Z.J.M, A.P. de Matos and S. De O. Silva. 1998. Black Sigatoka confirmed in Brazil.<br />

INFOMUSA 7(1):31.<br />

Cordeiro Z.J.M., A.P. de Matos, L. Gasparotto and M. de J.B. Cavalcante. 2000. Disseminação<br />

da Sigatoka-negra no Brasil. Summa Phytopathologica 26:110. (Abstract 062).<br />

96


Session 1<br />

Z.J. Maciel Cordeiro and A. Pires de Matos<br />

Gasparotto L., J.C.R. Pereira and D.R. Trindade. 2001a. Situação atual da Sigatoka negra da<br />

bananeira. Fitopatologia brasileira (suplemento) 26:449. (Abstract 692).<br />

Gasparotto L., J.C.R. Pereira, M.M. Costa and M.C.N. Pereira. 2001b. Fungicidas para o controle<br />

da Sigatoka negra da bananeira. Fitopatologia brasileira (suplemento) 26:434. (Abstract<br />

636).<br />

Hanada R.E., L.Gasparotto and J.C.R. Pereira. 2000. Sobrevivência de conídios de<br />

<strong>Mycosphaerella</strong> fijiensis em diferentes materiais. Fitopatologia brasileira (suplemento)<br />

25:380. (Abstract 303).<br />

Pereira, J. C. R., L. Gasparotto, A. F. Da S. Coelho, and A.F. Urben. 1998. Ocorrência da Sigatoka<br />

negra no Brasil. Fitopatologia brasileira (suplemento) 23:295.<br />

Pereira, J. C. R. and L. Gasparotto. 2001. Sigatoka negra da bananeira. Simpósio Norte Mineiro<br />

sobre a cultura da banana. Pp. 102-104 in Nova Porteirinha, 6-9 de novembro, Anais I.<br />

97


Session 1<br />

A.K.J. Surridge et al.<br />

Poster<br />

Fungi associated with banana foliage<br />

in South Africa<br />

A.K.J.Surridge,A.Viljoen and F.C. Wehner<br />

Abstract<br />

A comprehensive investigation was conducted to determine the identity, distribution and<br />

importance <strong>of</strong> fungi associated with banana leaves in South Africa. Banana leaves were randomly<br />

collected from the five banana growing areas in the country. Spores were isolated from <strong>leaf</strong> lesions<br />

following surface sterilization and incubation in moisture chambers or taken directly collected<br />

from lesions. Single spores were then cultured on half-strength potato dextrose agar. Both<br />

molecular and morphological techniques were applied to identify the isolates. Four main<br />

<strong>diseases</strong> were found in the different banana growing areas. Yellow Sigatoka (caused by<br />

M. musicola), <strong>Mycosphaerella</strong> speckle (caused by M. musae) and Cordana <strong>leaf</strong> <strong>spot</strong> (caused by<br />

Cordana musae) were present in all five areas, whereas, Cladosporium speckle (caused by<br />

Cladosporium musae) only occurred in Levubu. Many other fungi, predominantly saprophytes and<br />

endophytes, were also isolated. The most common species include (in order <strong>of</strong> predominance)<br />

Nigrospora sacchari, N. sphaerica and N. oryzae.<br />

Resumen - Hongos asociados con el follaje del banano en Africa del Sur<br />

En Africa del Sur, se llevó a cabo una amplia investigación con el fin de determinar la identidad,<br />

distribución e importancia de los hongos asociados con las hojas de banano. Las hojas de banano<br />

se recolectaron al azar en cinco zonas bananeras del país. Los aislados se hicieron de las lesiones<br />

foliares después de la esterilización de su superficie, incubación en cámaras húmedas, o las esporas<br />

se recolectaron directamente de las lesiones. Luego, las esporas individuales fueron cultivadas<br />

en el agar de dextrosa de patata de fuerza media. Para identificar los aislados se emplearon<br />

técnicas tanto moleculares como morfológicas. En diferentes zonas productoras de banano se<br />

descubrieron cuatro enfermedades principales. La Sigatoka amarilla (causada por M. musicola),<br />

la mancha <strong>Mycosphaerella</strong> (causada por M. musae) y la mancha foliar Cordana (causada por<br />

Cordana musae) se encontraron presentes en todas las cinco áreas, mientras que la mancha por<br />

Cladosporium (causada por Cladosporium musae) solo se encontró en Levubu. También se<br />

aislaron muchos otros hongos, predominantemente sapr<strong>of</strong>íticos y end<strong>of</strong>íticos. Las especies más<br />

comunes incluyen (en orden de predominancia) Nigrospora sacchari, N. sphaerica y N. oryzae.<br />

University <strong>of</strong> Pretoria, Pretoria, South Africa<br />

99


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Résumé - Champignons associés au feuillage du bananier en Afrique du Sud<br />

Une recherche détaillée a été faite afin de déterminer l’identité, la répartition et l’importance des<br />

champignons associés aux feuilles de bananier en Afrique du Sud. Des feuilles de bananiers ont<br />

été récoltées au hasard dans les cinq régions productrices du pays. Des spores ont été isolées des<br />

lésions foliaires après stérilisation de la surface foliaire et mise en incubation dans des chambres<br />

humides ou prélevées directement sur les lésions. Des spores isolées ont ensuite été cultivées sur<br />

un milieu gélifié à l’agar à demi-concentration de dextrose de pomme de terre. Des techniques<br />

moléculaires et morphologiques ont été appliquées afin d’identifier les isolats. Quatre maladies<br />

principales ont été trouvées dans les différentes régions de culture de la banane. La maladie de<br />

Sigaoka (causée par M. musicola), le <strong>Mycosphaerella</strong> speckle (causé par M. musae) et le Cordana (causé<br />

par Cordana musae) étaient présents dans les cinq zones, alors que le Cladosporium (causé par<br />

Cladosporium musae) n’a été trouvé qu’à Levubu. De nombreux autres champignons, surtout des<br />

saprophytes et des endophytes, ont également été isolés. Les espèces les plus communes sont, dans<br />

l’ordre d’importance, Nigrospora sacchari, N. sphaerica et N. oryzae.<br />

Introduction<br />

Among the various fungi associated with the foliage <strong>of</strong> banana plants, pathogens<br />

such as <strong>Mycosphaerella</strong> musicola, M. fijiensis and M. eumusae cause significant<br />

losses. Others, e.g. M. musae, Cladosporium musae and Cordana musae, can become<br />

damaging under certain climatic conditions. In addition to these pathogens, various<br />

species <strong>of</strong> endophytic fungi have also been reported on Musa species. The most<br />

commonly isolated are Colletotrichum gloeosporioides, Nigrospora oryzae,<br />

Pestalotiopsis palmarum and Phoma spp. (Brown et al., 1998 ; Photita et al., 2001).<br />

Some <strong>of</strong> these, e.g. N. oryzae, are also known to cause minor disease on their host<br />

plant (Ellis, 1971).<br />

In South Africa, the fungal pathogens previously reported on banana include<br />

M. musicola (Van den Boom and Kuhne, 1969), M. musae (Brodrick, 1973) and<br />

Cordana musae (Roth, 1965). However, these reports were based solely on symptom<br />

sightings and not on isolation <strong>of</strong> the causal organisms. This study was conducted<br />

to determine, and update existing knowledge <strong>of</strong>, the identity <strong>of</strong> fungi associated with<br />

banana leaves in South Africa.<br />

Materials and methods<br />

Banana leaves were randomly collected from the five banana growing areas in<br />

South Africa during 2000-2001, namely Levubu, Tzaneen, Kiepersol, Komatipoort<br />

and southern Kwa-Zulu Natal. Samples were taken from diseased leaves incubated<br />

in moisture chambers and cultured on half-strength potato-dextrose agar. Isolates<br />

were identified morphologically. The identity <strong>of</strong> the causal organism was confirmed<br />

using diagnostic PCR and the species-specific primers <strong>of</strong> Johansen and Jeger (1993).<br />

Results<br />

Four <strong>leaf</strong> <strong>diseases</strong> were identified. Sigatoka disease, <strong>Mycosphaerella</strong> speckle and<br />

Cordana <strong>leaf</strong> <strong>spot</strong> were present in all five areas. Cladosporium speckle occurred<br />

only in Levubu, the most northern <strong>of</strong> the areas. M. musicola was the most<br />

100


Session 1<br />

A.K.J. Surridge et al.<br />

commonly isolated pathogen, followed by M. musae (Table 1). The presence <strong>of</strong><br />

M. musicola and the absence <strong>of</strong> other <strong>Mycosphaerella</strong> pathogens was confirmed<br />

by PCR. Various saprobes and endophytes representing 23 species were also<br />

isolated from banana <strong>leaf</strong> material (Table 1). Nigrospora oryzae was the species<br />

isolated most frequently.<br />

Table 1. Fungi associated with banana foliage in South Africa.<br />

Species<br />

Number <strong>of</strong> isolates<br />

Alternaria alternata 21<br />

Alternaria cf. citri 3<br />

Alternaria tenuissima 4<br />

Bipolaris cynodontis 3<br />

Cladosporium musae 5<br />

Colletotrichum gloeosporioides 5<br />

Colletotrichum musae 1<br />

Cordana musae 30<br />

Curvularia lunata 1<br />

Curvularia pallescens 1<br />

Diaporthe sp. 3<br />

Drechslera dematoidea 1<br />

Drechslera phlei 1<br />

Epicoccum nigrum 3<br />

Exserohilum rostratum 1<br />

Harpographium sp. 1<br />

<strong>Mycosphaerella</strong> musae 30<br />

<strong>Mycosphaerella</strong> musicola 66<br />

Myrothecium verrucaria 1<br />

Nigrospora oryzae 81<br />

Nigrospora sacchari 10<br />

Nigrospora sphaerica 25<br />

Pestalotiopsis guepinii 9<br />

Phoma glomerata 7<br />

Phyllosticta sp. 1<br />

Pithomyces sacchari 1<br />

Selenophoma asterina 10<br />

Selenophoma juncea 5<br />

Discussion<br />

M. musicola is the most common and severe pathogen <strong>of</strong> banana foliage in South<br />

Africa. It was identified morphologically and its presence was confirmed using<br />

molecular markers (Johanson and Jeger, 1993). The second most prevalent pathogen<br />

was M. musae which in some cases caused severe symptoms resulting in <strong>leaf</strong> death.<br />

The subtropical conditions in the banana plantations <strong>of</strong> South Africa appear to be<br />

conducive to Sigatoka disease. Cordana musae is considered to be a minor/secondary<br />

<strong>leaf</strong> pathogen and was most <strong>of</strong>ten observed infecting in conjunction with Sigatoka<br />

disease or a speckle. Its presence and that <strong>of</strong> Cladosporium musae was<br />

morphologically confirmed, as well as according to lesion appearance.<br />

101


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Most fungi isolated were strictly saprobes. The genus Nigrospora, particularly<br />

N. oryzae, was the most commonly isolated. This conforms to literature <strong>of</strong> endophytes<br />

isolated from banana leaves in Hong Kong (Brown et al., 1998). All species isolated<br />

are the first recordings on banana leaves in South Africa. Colletotrichum<br />

gloeosporioides, C. musae, N. oryzae, and some Curvularia, Pestalotiopsis, Phoma,<br />

Phyllosticta species have previously been reported from Musa in Thailand (Photita<br />

et al., 2001), Hong Kong and Northern Queensland, Australia (Brown et al., 1998).<br />

However, no report could be traced referring to the presence <strong>of</strong> Alternaria cf. citri,<br />

A. tenuissima, Bipolaris cynodontis, Diaporthe sp., Drechslera dematoidea, D. phlei,<br />

Exserohilum rostratum, Harpographium sp., Myrothecium verrucaria, N. sacchari,<br />

N. sphaerica, Pithomyces sacchari, Selenophoma asterina and S. juncea on Musa<br />

species.<br />

References<br />

Brodrick H. T. 1973. Spikkelblaar. Banana Series Journal J4:1-2.<br />

Brown K. B., K.D. Hyde and D.I. Guest. 1998. Preliminary studies on endophytic fungal<br />

communities <strong>of</strong> Musa acuminata species complex in Hong Kong and Australia. Fungal<br />

Diversity 1:27-51.<br />

Ellis M. B. 1971. Dematiaceous Hyphomycetes. Commonwealth Mycological Institute, Kew,<br />

Surrey, United Kingdom.<br />

Johanson A. and M.J. Jeger. 1993. Use <strong>of</strong> PCR for detection <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis and<br />

M. musicola, the causal agents <strong>of</strong> Sigatoka <strong>leaf</strong> <strong>spot</strong>s in banana and plantain. Mycological<br />

Research 96:670-674.<br />

Photita W., S. Lumyong, P. Lumyong and K.D. Hyde. 2001. Endophytic fungi <strong>of</strong> wild<br />

banana (Musa acuminata) at Doi Suthep Pui National Park, Thailand. Mycological Research<br />

105: 1508-1513.<br />

Roth G. 1965. A new <strong>leaf</strong> <strong>spot</strong> disease <strong>of</strong> dwarf Cavendish banana in South Africa. South<br />

African Journal <strong>of</strong> Agricultural Science 8:87-92.<br />

Van den Boom T. and F.A. Kuhne. 1969. First report <strong>of</strong> Sigatoka disease <strong>of</strong> banana in South<br />

Africa. Citrus Journal 428:17-18.<br />

102


Session 1<br />

Recommendations<br />

Recommendations <strong>of</strong> session 1<br />

Dispersal <strong>of</strong> <strong>Mycosphaerella</strong> spp.<br />

<strong>Mycosphaerella</strong> fijiensis continues to spread to new areas. It is the dominant <strong>leaf</strong> <strong>spot</strong> pathogen<br />

in West Africa. In 2000-2002, the pathogen was identified for the first time in Madagascar,<br />

the Bahamas, the Galapagos Islands <strong>of</strong> Ecuador and in the north Queensland banana growing<br />

area where eradication is being attempted. The encroaching threat <strong>of</strong> M. fijiensis in the eastern<br />

Caribbean is <strong>of</strong> concern. It has been estimated that 40% <strong>of</strong> banana growers in the French<br />

Antilles would stop cultivating banana if the pathogen became established. The effects in the<br />

Windward Islands would also be significant.<br />

Quarantine may need to be either strengthened or reinforced to prevent entry. The<br />

introduction <strong>of</strong> an appropriate monitoring system to detect any incursion <strong>of</strong> M. fijiensis<br />

should be encouraged.<br />

M. eumusae is currently limited in extent throughout most <strong>of</strong> Asia, although there is some<br />

evidence that the pathogen may have reached Africa. Eumusae <strong>leaf</strong> <strong>spot</strong> disease has been<br />

observed on ‘Mysore’ (AAB) in Sri Lanka. As this clone is highly resistant to M. musicola<br />

and M. fijiensis, there is some cause <strong>of</strong> concern. Information suggests that Cavendish and<br />

plantain cultivars are also very susceptible. The dynamics <strong>of</strong> the disease are not fully<br />

understood. In order to prepare adequate disease control strategies, a detailed knowledge <strong>of</strong><br />

the epidemiology <strong>of</strong> this pathogen is urgently required.<br />

The exact distribution <strong>of</strong> M. eumusae needs to be known. Further surveys in South and<br />

Southeast Asia, to determine where M. musicola, M. fijiensis and M. eumusae occur,<br />

are necessary.<br />

More information on the effect <strong>of</strong> M. eumusae on the growth and yield <strong>of</strong> banana clones<br />

is needed.<br />

More laboratory research needs to be undertaken with M. eumusae to determine its<br />

optimum growth temperature and other relevant biological information. This data would<br />

underpin epidemiological studies.<br />

The name <strong>of</strong> the banana clone affected, an indicator <strong>of</strong> the severity <strong>of</strong> the <strong>leaf</strong> <strong>spot</strong> and<br />

local environmental data would be useful as this may help explain distribution. IMTP<br />

trials are seen as ideal locations for assessing the reaction <strong>of</strong> different clones to the<br />

different <strong>leaf</strong> <strong>spot</strong> pathogens. The collection and diagnosis <strong>of</strong> specimens <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> from<br />

IMTP trials sites needs to be continued. The cooperation and collaboration <strong>of</strong> scientists<br />

in South and Southeast Asia is viewed as essential. Identification tools should be provided<br />

to enable diagnoses to be undertaken locally.<br />

103


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

M. eumusae was also identified from <strong>leaf</strong> <strong>spot</strong> specimens collected in 1989 and 1990 at Onne<br />

in Nigeria. Eumusae <strong>leaf</strong> <strong>spot</strong> disease has therefore been at this location for at least 13 years.<br />

The IITA germplasm collection and clones in local banana farms in and around Onne<br />

should be surveyed for <strong>leaf</strong> <strong>spot</strong> pathogens. The results would indicate the relative<br />

competitiveness <strong>of</strong> M. fijiensis and M. eumusae on different cultivars in West Africa.<br />

Although M. eumusae has not been identified from many isolates <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> pathogens<br />

collected in neighbouring Cameroon, surveys <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> in Nigeria and other countries<br />

in West Africa may indicate if spread has occurred<br />

Taxonomy<br />

Anamorph morphology is more important than teleomorph morphology in distinguishing<br />

the <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens. It has been proposed that the anamorph stage <strong>of</strong><br />

M. fijiensis be renamed Pseudocercospora fijiensis as the phylogenetic studies do not support<br />

keeping the name Paracercospora fijiensis.<br />

Diagnostics<br />

Several fungi <strong>diseases</strong> that attack leaves have been reported on Musa and other related species.<br />

A greater knowledge <strong>of</strong> <strong>Mycosphaerella</strong> pathogens/saprophytes, and <strong>of</strong> those in related genera,<br />

is a prerequisite to the development <strong>of</strong> rapid diagnostic tests to distinguish <strong>leaf</strong> <strong>spot</strong> pathogens.<br />

Diagnostic tools depend on the development <strong>of</strong> species-specific primers, such as microsatellites<br />

and ITS-sequences, tested on <strong>Mycosphaerella</strong> isolates from all over the world.<br />

More taxonomic information about species <strong>of</strong> <strong>Mycosphaerella</strong> and other related genera<br />

that either form or occur in banana <strong>leaf</strong> lesions would be beneficial.<br />

Diagnostic tools specific to the three main species <strong>of</strong> <strong>Mycosphaerella</strong> pathogen on Musa:<br />

M. fijiensis, M. musicola and M. eumusae should be developed.<br />

The currently available molecular methods should be assessed for their specificity.<br />

A manual with descriptions <strong>of</strong> symptoms and morphological characters should be<br />

produced.<br />

Protocols for collection and analysis <strong>of</strong> samples should be developed.<br />

PROMUSA participants should be trained on the different technologies required:<br />

collecting and sampling, single-ascosopore cultures, and molecular markers.<br />

Resistant cultivars<br />

Reports <strong>of</strong> black <strong>leaf</strong> streak disease on two resistant FHIA hybrids are <strong>of</strong> concern. Stress caused<br />

by adverse growing conditions may be responsible.<br />

Factors responsible for breakdowns need to be investigated, including the possibility that<br />

resistance is being eroded.<br />

104


Session 2<br />

Population biology<br />

and epidemiology


Session 2<br />

L.H. Jácome<br />

Introduction<br />

Population biology and epidemiology<br />

L. H. Jácome<br />

Introduction<br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> comprise one <strong>of</strong> the most important disease complex<br />

limiting banana and plantain production. Sigatoka disease is still important in some<br />

areas. In the 1990s, black <strong>leaf</strong> streak disease was spreading and is now reported from<br />

most parts <strong>of</strong> the world. Recently, a new disease, eumusae <strong>leaf</strong> <strong>spot</strong> was reported in<br />

South and Southeast Asia.<br />

Black <strong>leaf</strong> streak disease reached most South American countries by the early<br />

1990s and was reported in Bolivia in 1996 and Brazil in 1998. The disease was<br />

reported in the Caribbean in the Dominican Republic in 1996 and in Haiti in 2000.<br />

It threatens banana production in Puerto Rico and the Lesser Antilles. The disease<br />

was reported in the United States in 1998. In 2000, two new outbreaks <strong>of</strong> black <strong>leaf</strong><br />

streak disease were reported in the Caribbean and Indian Ocean region. In Latin<br />

America the latest report was in the Galapagos Islands in early 2001. In April 2001,<br />

it was reported for the first time in a commercial production area near Tully in North<br />

Queensland, Australia. This illustrates the spreading capacity <strong>of</strong> this disease.<br />

The continued spread <strong>of</strong> black <strong>leaf</strong> streak disease in the tropics within the last<br />

decade has made the disease the most economically important disease <strong>of</strong> <strong>bananas</strong><br />

and plantains. Except in the Philippines, Sigatoka disease has been replaced by black<br />

<strong>leaf</strong> streak disease. Sigatoka disease is better adapted to cooler areas and dominates<br />

at altitudes above 1200 metres above sea level. Therefore, knowledge <strong>of</strong> the<br />

distribution and variability <strong>of</strong> the pathogens is needed to ensure the efficient<br />

introduction and sustainability <strong>of</strong> resistant hosts.<br />

Restriction enzyme analysis has revealed unique differences in the banding pattern<br />

<strong>of</strong> DNA fragments. Genetic variability among populations <strong>of</strong> <strong>Mycosphaerella</strong> from<br />

different geographical regions has been identified using DNA restriction fragment<br />

Pathotec, La Lima, Honduras<br />

107


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

length polymorphism (RFLP) markers. RFLPs are used as markers on physical and<br />

genetic linkage maps. To complement RFLP tests, methods based on polymerase<br />

chain reaction (PCR) could be used to obtain DNA pr<strong>of</strong>iles. Differentiation <strong>of</strong> the<br />

pathogens by PCR could be useful in determining their distribution, spread and<br />

prevalence. Additional work is now needed to obtain accurate information on the<br />

geographical distribution <strong>of</strong> genetic variants <strong>of</strong> M. fijiensis, and to test hypotheses<br />

about the origin and spread <strong>of</strong> the pathogen, and its population structure.<br />

Given that the disease caused by M. fijiensis is more serious than the one caused<br />

by M. musicola, research is needed on the pathogenic variability <strong>of</strong> M. fijiensis,<br />

the distribution <strong>of</strong> its pathogenic variants and the identification <strong>of</strong> sources <strong>of</strong><br />

resistance. Studies on genetic diversity in M. fijiensis using RFLP have shown high<br />

levels <strong>of</strong> polymorphism. Genetic variability was greatest in the Philippines and Papua<br />

New Guinea. Isolates from Africa, the Pacific Islands and Latin America formed<br />

genetically homologous groups that were specific to each region. Groups from the<br />

Pacific and Latin America also appeared to be related. Are these independent<br />

introductions <strong>of</strong> the pathogen? Are we dealing with the effects <strong>of</strong> genetic drift on<br />

the population structure?<br />

Gametic disequilibrium analysis among RFLP loci has shown that genetic<br />

recombination plays an important role in the population structure <strong>of</strong> M. fijiensis.<br />

Gene pyramiding may not be durable. Mixing varieties or partial resistance could<br />

be more appropriate. Observations indicate that resistance to black <strong>leaf</strong> streak disease<br />

may be breaking down in some hybrids. It is not known whether this is due to more<br />

pathogenic variants or to favourable environmental conditions resulting in severe<br />

disease pressure. In addition to studying the effects <strong>of</strong> climatic conditions on the<br />

infection process, managing the inoculum using cultural practices should be<br />

considered. The fitness <strong>of</strong> fungicide-resistant <strong>Mycosphaerella</strong> deserves attention, as<br />

does the potential contribution <strong>of</strong> Paracercospora fijiensis to epidemics <strong>of</strong> black<br />

<strong>leaf</strong> streak disease in some areas.<br />

Disease dynamics has been studied by calculating rates <strong>of</strong> disease development<br />

in relation to climatic factors. Epidemiological studies demonstrated that ascospores<br />

were the predominant inoculum <strong>of</strong> black <strong>leaf</strong> streak disease and their release was<br />

correlated with rainfall. Therefore, it should be possible to predict future rates <strong>of</strong><br />

disease development from previous patterns <strong>of</strong> spore release in relation to<br />

temperature and rainfall. However, no consistency has been observed between<br />

ascospore release and disease development, probably because the conditions for spore<br />

release are not always conducive to infection.<br />

There is considerable variation in disease progress curves and the relative duration<br />

<strong>of</strong> epidemics in pathosystems <strong>of</strong> Musa-<strong>Mycosphaerella</strong> spp. and plantation<br />

management (treated with fungicide versus untreated). For each epidemic, it is<br />

possible to determine the time <strong>of</strong> disease onset, the initial amount <strong>of</strong> disease, the<br />

rate <strong>of</strong> disease increase, the area under the disease progress curve, the shape <strong>of</strong> the<br />

curve, the maximum disease, final amount <strong>of</strong> disease and the duration <strong>of</strong> the<br />

epidemic. In tropical and semitropical climatic zones, for epidemics that are not<br />

curtailed by the harvest <strong>of</strong> an annual crop, e.g. black <strong>leaf</strong> streak disease, the<br />

progressive and regressive phases <strong>of</strong> an epidemic correspond mainly to seasonal<br />

changes in weather conditions.<br />

108


Session 2<br />

L.H. Jácome<br />

At one level, a temporal analysis seeks gross comparisons between experimental<br />

treatments, e.g. fungicide spray schedules in order to evaluate strategies for disease<br />

management. At a second, more complex level, changes in specific environmental<br />

factors, pathogens or host resistance lead to changes in the epidemic that are reflected<br />

by changes in the disease progress curves. A third level <strong>of</strong> analysis corresponds to<br />

comparative epidemiology, where the purpose is to identify similarities and<br />

differences between epidemics based on the shape <strong>of</strong> the disease progress curves and<br />

to look for the elements that serve as primary determinants. Geographical populations<br />

<strong>of</strong> <strong>Mycosphaerella</strong>, with marked genetic differentiation, could be considered as<br />

separate epidemiological units requiring independent disease management. This<br />

suggests that studies and modelling <strong>of</strong> the epidemiology, distribution and population<br />

structure <strong>of</strong> the three <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens should be undertaken at<br />

the national, regional and global levels.<br />

In general, there is little information on the biology, population structure and<br />

epidemiology <strong>of</strong> the <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens. As a result, areas requiring<br />

further investigation are pathogen variability, distribution <strong>of</strong> variants, sources <strong>of</strong><br />

resistance, epidemiology and population structure. The papers and posters in this<br />

session have been selected to improve our understanding <strong>of</strong> the population biology<br />

and epidemiology <strong>of</strong> the <strong>Mycosphaerella</strong> pathogens <strong>of</strong> Musa. Aspects <strong>of</strong> the<br />

aerobiological pathway <strong>of</strong> M. fijiensis ascospores and conidia at small-scale and<br />

mesoscale levels are discussed. Such studies are needed to clarify the temporal and<br />

spatial patterns <strong>of</strong> the spread <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> within or across banana cropping<br />

sequences, and in relation to environmental and host factors. Several forecasting<br />

schemes for black <strong>leaf</strong> streak disease have been introduced and are based on those<br />

developed for Sigatoka disease in the French Antilles. The forecasting schemes<br />

combine local weather and ascospore trapping data. The value <strong>of</strong> ascospore trapping<br />

for forecasting black <strong>leaf</strong> streak disease is discussed.<br />

Aspects <strong>of</strong> the genetic structure and evolution <strong>of</strong> <strong>Mycosphaerella</strong> pathogens at<br />

the global, regional and local levels are discussed. Genetic differentiation and<br />

independency <strong>of</strong> introductions <strong>of</strong> the pathogens according to the region are also<br />

described. Knowledge <strong>of</strong> the variability within <strong>Mycosphaerella</strong> is necessary for<br />

breeding and management <strong>of</strong> disease resistance. The usefulness <strong>of</strong> microsatellite<br />

markers for the study <strong>of</strong> fungal populations having high evolutionary rate is presented<br />

by an analysis <strong>of</strong> the introduction and spread <strong>of</strong> black <strong>leaf</strong> streak disease in Mexico.<br />

The presence <strong>of</strong> polymorphisms in chromosome length between molecular karyotypes<br />

<strong>of</strong> M. fijiensis is discussed. Understanding the organization <strong>of</strong> the genome could<br />

lead to the development <strong>of</strong> new strategies for disease control management.<br />

Given that pathogens can evolve to break down host resistance, further research<br />

on the evolution <strong>of</strong> the three <strong>Mycosphaerella</strong> pathogens on resistant hosts is needed.<br />

As stated in the Musa disease fact sheet No. 8, published by INIBAP, population<br />

studies <strong>of</strong> <strong>Mycosphaerella</strong> pathogens <strong>of</strong> banana are required to determine whether<br />

the banana-producing regions correspond to one or several epidemiological units.<br />

The studies should use molecular markers and determine pathogenicity. Techniques<br />

are also needed to determine quickly and reliably host-pathogen interactions in<br />

controlled conditions. The variability in pathogenicity in genetically differentiated<br />

populations <strong>of</strong> the three pathogens could then be evaluated by means <strong>of</strong> a standard<br />

109


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

set <strong>of</strong> Musa cultivars. Population studies should help define, for the different regions,<br />

a set <strong>of</strong> <strong>Mycosphaerella</strong> isolates that are representative <strong>of</strong> the variability in virulence<br />

and aggressiveness for use in resistance screening. Knowing the components <strong>of</strong> partial<br />

resistance which greatly reduce the rate <strong>of</strong> disease development in the field is also<br />

important. Finally, the evolution <strong>of</strong> the pathogen population in response to the<br />

selection pressure exerted by resistant cultivars should be evaluated if durable<br />

resistance is to be achieved.<br />

110


Session 2<br />

P.J.A. Burt<br />

Airborne dispersal<br />

<strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

P. J. A. Burt<br />

Abstract<br />

Dispersal <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis ascospores and conidia within, above and outside unsprayed<br />

banana plantations was studied in a series <strong>of</strong> field experiments at CATIE in Costa Rica. Laboratory<br />

experiments were also conducted in Costa Rica and the UK, to estimate the potential viability<br />

<strong>of</strong> spores dispersing through the atmosphere.Three field regimes were used to assess windborne<br />

spore dispersal on a local and mesoscale, in relation to wet and dry seasons. Spore catches were<br />

analysed in relation to the time <strong>of</strong> day <strong>of</strong> capture, temperature, rainfall and behaviour <strong>of</strong> the wind.<br />

Results showed that ascospores and conidia are windborne within infected plantations and up<br />

to several tens <strong>of</strong> kilometres away from disease sources. Laboratory studies <strong>of</strong> simulated spore<br />

release under field conditions showed that significantly fewer spores entered the air than might<br />

be expected on the basis <strong>of</strong> field surveys. The reason for this is unclear. There is also evidence<br />

that a major constraint on the airborne dispersal <strong>of</strong> viable spores is their duration <strong>of</strong> exposure<br />

to ultraviolet radiation in sunlight. A greater understanding <strong>of</strong> the microscale processes occurring<br />

on the surface <strong>of</strong> an infected banana <strong>leaf</strong> is required in order to resolve the role <strong>of</strong> the wind in<br />

the epidemiology <strong>of</strong> black <strong>leaf</strong> streak disease. A more accurate quantification <strong>of</strong> the numbers <strong>of</strong><br />

spores undertaking long-distance dispersal (with assessments <strong>of</strong> their viability in the field) is<br />

also essential. Future research needs are discussed.<br />

Resumen - Dispersión aérea de <strong>Mycosphaerella</strong> fijiensis<br />

Se han investigado los aspectos de la vía aerobiológica de las ascosporas y conidias de<br />

<strong>Mycosphaerella</strong> fijiensis en experimentos en el campo en CATIE, Costa Rica.También se realizaron<br />

investigaciones en el laboratorio en Costa Rica y Reino Unido con el fin de evaluar la viabilidad<br />

potencial de la dispersión de esporas en la atmósfera. Se utilizaron tres regímenes de campo para<br />

evaluar la dispersión de esporas por el viento a escalas local y media en relación con las<br />

estaciones húmeda y seca. La captura de esporas fue analizada basándose en la hora del día de<br />

la captura, temperatura, precipitación y comportamiento del viento. Los resultados mostraron<br />

que las ascosporas y conidias se propagan por el viento dentro de las plantaciones infectadas y<br />

hasta decenas de kilómetros fuera de las fuentes de la enfermedad. Sin embargo, de los estudios<br />

de laboratorio que simularon la liberación de las esporas bajo condiciones de campo, está claro<br />

que se podría esperar que una cantidad significativamente menor de esporas penetre en el aire<br />

en base a las encuestas en el campo. La razón de este hecho no está clara. También existen<br />

Natural Resources Institute, Kent, United Kingdom<br />

111


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

evidencias de que la duración de su exposición a la radiación ultravioleta del sol es una de las<br />

principales limitaciones para la dispersión aérea. Se requiere un mayor entendimiento de los<br />

procesos a pequeñas escalas que ocurren en la superficie de una hoja de banano infectado con<br />

el fin de resolver con mayor pr<strong>of</strong>undidad el papel del viento en la epidemiología de la Sigatoka<br />

negra. También es esencial realizar una cuantificación más precisa de las cantidades de esporas<br />

que se dispersan a largas distancias (con la evaluación de su viabilidad en el campo). Se debe<br />

discutir estas futuras investigaciones.<br />

Résumé - Dispersion aérienne de <strong>Mycosphaerella</strong> fijiensis<br />

La dispersion des ascospores et des conidies de <strong>Mycosphaerella</strong> fijiensis a été étudiée au cours<br />

d’une série d’essais en champ au CATIE, au Costa Rica. Des essais en laboratoire ont également<br />

été effectués au CATIE et en Grande-Bretagne, afin d’estimer la viabilité potentielle des spores<br />

au cours de leur dispersion dans l’atmosphère.Trois régimes en champ ont été utilisés pour évaluer<br />

la dispersion par le vent des spores à l’échelle locale et moyenne, en relation avec la saison sèche<br />

et humide. Les spores récoltées ont été analysées en relation avec l’heure de capture, la<br />

température, la pluviométrie, et la force et la direction du vent. Les résultats ont montré que les<br />

ascospores et les conidies sont transportées par le vent au sein des plantations infectées et jusqu’à<br />

plusieurs dizaines de kilomètres des sources d’infection. Les simulations en laboratoire de la<br />

dissémination des spores en conditions naturelles ont montré qu’un nombre de spores<br />

significativement plus faible que celui attendu sur la base des mesures en champ entrait dans<br />

l’air. Les raisons de ce phénomène ne sont pas claires. Il existe également des preuves qu’une<br />

contrainte majeure à la dispersion aérienne de spores viables est la durée de leur exposition aux<br />

rayons ultraviolets du soleil. Une meilleure compréhension des processus à l’échelle microscopique<br />

qui ont lieu à la surface d’une feuille de bananier infectée est nécessaire, afin de comprendre le<br />

rôle du vent dans l’épidémiologie de la maladie des raies noires. Une quantification plus précise<br />

du nombre de spores qui sont dispersées à longue distance (avec une évaluation de leur viabilité<br />

en champ) est également essentielle. Les besoins futurs de recherches sont discutés.<br />

Introduction<br />

Banana and plantain are major subsistence crops for small-scale farmers in the<br />

developing world, and production is increasing worldwide. Crops are affected by a<br />

range <strong>of</strong> pests, including <strong>Mycosphaerella</strong> musicola, the causal agent <strong>of</strong> Sigatoka<br />

disease and M. fijiensis, the causal agent <strong>of</strong> black <strong>leaf</strong> streak disease. The infective<br />

agents are ascospores and conidia. Black <strong>leaf</strong> streak disease has gradually replaced<br />

Sigatoka disease in most banana growing areas and can reduce yields by up to 50%<br />

(Stover and Simmonds, 1987). Fungicides can control black <strong>leaf</strong> streak disease but<br />

they are too expensive for small-scale farmers and affect the environment. Black<br />

<strong>leaf</strong> streak disease is currently absent from most <strong>of</strong> the Caribbean. Its arrival would<br />

be disastrous for the smallholders. Quarantine regulations being strict, the risk <strong>of</strong><br />

black <strong>leaf</strong> streak disease arriving in the Caribbean by windborne dispersion requires<br />

quantification.<br />

Previous studies have implicated wind and water in the release <strong>of</strong> ascospores and<br />

conidia <strong>of</strong> M. fijiensis, but there is disagreement in the literature about their relative<br />

importance. Conidial dispersal appears to occur primarily in water, either as run<strong>of</strong>f<br />

in dew or by rain splash (Leach, 1946; Stover, 1968, 1972; Meredith et al., 1973;<br />

Stover and Simmonds, 1987; Gauhl, 1989) whereas ascospores are primarily removed<br />

from diseased leaves by wind (Leach, 1946). There is also evidence that the conidia<br />

may be blown <strong>of</strong>f infected leaves (Stover and Simmonds, 1987), and that ascospores<br />

112


Session 2<br />

P.J.A. Burt<br />

are released by water (Leach, 1941; Meredith, 1962; Meredith and Lawrence, 1970;<br />

Stover, 1970; Meredith et al., 1973). One thing is sure: within infected plantations,<br />

ascospores and conidia are present in the air in varying amounts (Gauhl, 1989).<br />

Long-distance airborne dispersal is known for some fungal pathogens e.g.<br />

Peronospora tabacina, the causal agent <strong>of</strong> tobacco blue mould (Davies et al., 1985);<br />

Cochliobolus heterostrophus, which causes corn <strong>leaf</strong> blight (Pedgley, 1982);<br />

Melampsora spp., the causal agent <strong>of</strong> poplar rust; and Hemileia vastatrix, which causes<br />

c<strong>of</strong>fee <strong>leaf</strong> rust (Burdekin, 1960; Bowden et al., 1971; Pedgley, 1982). There are no<br />

records <strong>of</strong> M. fijiensis ascospores or conidia high in the atmosphere. In the absence<br />

<strong>of</strong> such records, it is necessary to look at disease incidence, but this can be unreliable,<br />

as the infective agents may enter an area long before the disease is first observed<br />

or reported. The global pattern <strong>of</strong> Sigatoka disease suggests windborne dispersal<br />

from east to west. The pattern is less obvious for black <strong>leaf</strong> streak disease but disease<br />

records suggest that some intra-continental spread may have been windborne (Burt,<br />

1994).<br />

Even assuming that airborne spores travel in large enough numbers to overcome<br />

dilution in the atmosphere and that they enter an area where there are suitable hosts<br />

present, they may be affected by environmental conditions during transport. Of these,<br />

temperature and ultraviolet radiation in sunlight are probably the most significant.<br />

High temperatures destroy spore walls and denature DNA (Parnell et al., 1998).<br />

Ultraviolet (UV) radiation, depending on wavelength, also denatures DNA.<br />

Wavelengths below 320 nm, especially 250-270 nm (which do not reach the ground),<br />

are the most lethal (Setlow, 1974; Rotem et al., 1985; Chuang and Su, 1988; Rotem<br />

and Aust, 1991). UV radiation at wavelengths above 290 nm does not reach the<br />

ground, and kills spores within a few hours (Maddison and Manners, 1972; Bashi<br />

and Aylor 1983; Rotem et al. 1985). The ascospores <strong>of</strong> M. fijiensis have thin walls<br />

and are hyaline in contrast to the thick-walled spores <strong>of</strong> many windborne fungi.<br />

It is clear from the literature that ascospores and conidia <strong>of</strong> M. fijiensis have<br />

the capacity to be windborne over long and short distances, and that spores are<br />

present in the air within diseased plantations. Water (dew and rain) has been<br />

implicated in short-distance dispersal, although the precise role <strong>of</strong> wind in spore<br />

dispersal is unclear.<br />

The aerobiology <strong>of</strong> M. fijiensis was studied in 1992-1995 by investigating spore<br />

dispersal within a plantation; spore movement across a small experimental site;<br />

mesoscale dispersal; potential inoculum loads and spore viability. The studies were<br />

conducted at the Centro Agronómico Tropical de Investigación y Enseñanza (CATIE),<br />

Costa Rica, and in the UK. This paper summarizes the results.<br />

Methods<br />

Dispersal <strong>of</strong> spores within a plantation<br />

Airborne spores <strong>of</strong> M. fijiensis were monitored at four sites using rotorod spore traps<br />

within a naturally infected plantation at CATIE; samples were taken below the canopy,<br />

at mid-canopy and above the canopy (Burt et al., 1997). Wind speed, temperature,<br />

relative humidity and rainfall were recorded. Spore concentrations were meas-<br />

113


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

ured with a Burkard volumetric spore trap (Burkard Manufacturing Company,<br />

Rickmansworth, UK) at a height <strong>of</strong> 1.6 m above ground level. Disease incidence and<br />

disease development were also recorded (Burt et al., 1997). No attempt was made to<br />

control black <strong>leaf</strong> streak disease by treatment with fungicide.<br />

Small-scale dispersal <strong>of</strong> ascospores and conidia<br />

A small-scale field investigation was undertaken between May and August 1995 in<br />

order to clarify some <strong>of</strong> the results obtained by spore trapping and to investigate<br />

the role <strong>of</strong> wind and water in spore dispersal (Rutter et al., 1998). Disease<br />

development and spore movement was monitored in a small experimental plot in<br />

which 100 healthy plantain plants were arranged 2 m apart (Figure 1). Data were<br />

collected from quadrats and analysed. The numbers <strong>of</strong> ascospores and conidia in<br />

and around the plot were measured using four sets <strong>of</strong> rotorods at 0.5 m and 1.5 m<br />

above ground level for 20 minutes at 07.00 local time, the time when spore<br />

concentration is highest (Rutter and Burt, 1997). Wind, temperature, relative<br />

humidity and rainfall were also recorded. The ambient spore concentration was<br />

recorded continuously using a Burkard spore trap set at 6 m above ground level.<br />

An inoculum made up <strong>of</strong> a bunch <strong>of</strong> diseased leaves was placed in the centre <strong>of</strong> the<br />

plot (Figure 1) and the progress <strong>of</strong> the disease measured across the plot.<br />

B<br />

N<br />

Q2<br />

Q3<br />

R3/4<br />

X<br />

R7/8<br />

R5/6<br />

Q1<br />

Q4<br />

Figure 1. Small plot experimental design. Plants, shown by circles, were 2 m apart. R refers to the position<br />

and number <strong>of</strong> rotorod traps, X the location <strong>of</strong> the source <strong>of</strong> inoculum, the two rotorod traps and the weather<br />

instruments.The Burkard trap (B) was located outside the plot. Q refers to quadrats. (From Rutter et al., 1998).<br />

114


Session 2<br />

P.J.A. Burt<br />

Mesoscale dispersal<br />

Mesoscale dispersal <strong>of</strong> spores was investigated by sampling along a 5 km transect<br />

across the floor and up one side <strong>of</strong> a valley outside CATIE, between April and August<br />

1995 with 3 continuously recording Burkard spore traps. Trap 1 was located in the<br />

middle <strong>of</strong> the plantation, trap 2 in the small experimental plot and trap 3 was 5 km<br />

north-north-west <strong>of</strong> the plantation, on the side <strong>of</strong> the valley at an elevation <strong>of</strong> 1000<br />

m (Burt et al., 1998). There were no obstacles between trap 3 and the other two traps<br />

and there were no sources <strong>of</strong> inoculum close to trap 3 (the nearest large source was<br />

the plantation at CATIE). Daily spore counts were analysed in relation to the speed<br />

and direction <strong>of</strong> the wind at the three sites (Burt et al., 1998).<br />

Assessment <strong>of</strong> inoculum<br />

Disease surveys suggest that the infected plants in a plantation would act as a vast<br />

reservoir <strong>of</strong> inoculum, but trapping data suggest that, whilst both ascospores and<br />

conidia are windborne, they are not abundant in the air (particularly ascospores).<br />

Ten banana leaves showing necrosis on at least 16% <strong>of</strong> the surface <strong>of</strong> the <strong>leaf</strong><br />

(using the modified Stover scale to assess the levels <strong>of</strong> infection) were selected at<br />

random from the CATIE plantation during two periods: October 1993–February 1994<br />

and April–September 1995. Each <strong>leaf</strong> was divided into three parts, each with different<br />

amounts <strong>of</strong> necrosis (Figure 2). Necrotic tissue from each part was excised, the area<br />

measured and the perithecia counted. A regression equation relating the number <strong>of</strong><br />

perithecia to the necrotic area was calculated (Burt et al., 1999).<br />

A random selection <strong>of</strong> the <strong>leaf</strong> sections was exposed to a regime <strong>of</strong> wetting and<br />

drying under simulated field conditions (21°C and 100% relative humidity and 28°C<br />

at 60% relative humidity) in order to assess spore release (Burt et al., 1999). The<br />

released spores were deposited on agar in Petri dishes and counted using a<br />

compound microscope. A second regression equation was derived, relating the number<br />

<strong>of</strong> ascospores to perithecia in infected <strong>leaf</strong> tissue (Burt et al., 1999).<br />

Plant<br />

Area A<br />

Area B<br />

Area C<br />

1/3 <strong>leaf</strong> length 1/3 <strong>leaf</strong> length<br />

Typical <strong>leaf</strong> dimensions: 50 - 120 cm long, 25 - 50 cm wide<br />

Figure 2. Leaf parts used for counts <strong>of</strong> perithecia. (From: Rutter et al., 1998).<br />

115


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Viability <strong>of</strong> dispersed spores<br />

Spore viability under conditions <strong>of</strong> temperatures and solar radiation likely to be<br />

encountered during atmospheric transport was investigated (Parnell et al., 1998).<br />

Ascospores released from wetted <strong>leaf</strong> sections were exposed to simulated sunlight<br />

with UV wavelengths between 270 and 800 nm typical <strong>of</strong> the tropics, and<br />

incubated under a light/dark cycle at 24-26°C (Parnell et al. 1998). Control plants<br />

were subjected to the same regime minus the UV radiation.<br />

Results and discussion<br />

Dispersal <strong>of</strong> spores within a plantation<br />

Spore numbers at all sampling heights were generally low and the majority were<br />

found at sunrise, the time <strong>of</strong> dew evaporation (Table 1). More conidia than<br />

ascospores were caught at all heights within the plantation. Spores were also present<br />

after rain in the afternoon, again in low numbers (Burt et al., 1997). More conidia<br />

than ascospores were found at all heights within the plantation (Table1).<br />

The number <strong>of</strong> spores that got into the air was much smaller than the one expected<br />

from looking at disease incidence alone. These low numbers made it impossible to<br />

investigate the relationship between spore capture and weather, especially rainfall<br />

which was low during sampling. It was also unclear whether the larger number <strong>of</strong><br />

conidia in the air is atypical or indicates that dispersal is mainly by conidia. At a<br />

lowland site, Gauhl (1989) had found more ascospores than conidia. The different<br />

results obtained at CATIE suggest that ascospore trapping may not be a reliable<br />

method to forecast black <strong>leaf</strong> streak disease, at least at highland sites.<br />

Table 1. Summary <strong>of</strong> spore trapping <strong>of</strong> M. fijiensis within an infected plantation, December 1993–February 1994.<br />

Position relative Ascospores Conidia<br />

to the canopy<br />

Mean spore count Number Mean spore count Number <strong>of</strong><br />

(min-max) <strong>of</strong> samples (min-max) samples<br />

Above 1.35 (0-6) 40 5.75 (0-54) 40<br />

Middle 1.15 (0-63) 39 6.59 (0-63) 39<br />

1.18 (0-5) 40 7.65 (0-44) 40<br />

Bottom 1.20 (0-8) 40 8.52 (0-64) 40<br />

From Burt et al., 1997<br />

Small-scale dispersal <strong>of</strong> ascospores and conidia<br />

More conidia than ascospores were caught in the rotorod traps within and just<br />

above the plants in the small plot, but catches were again low (Figure 3). Initially,<br />

it appeared that many more ascospores than conidia were recorded in the Burkard<br />

spore trap but when the data were for sampling volume , the Burkard trap caught<br />

fewer spores overall (Rutter et al., 1998). There was no evidence that spores entered<br />

the plantation from outside and caused disease.<br />

116


Session 2<br />

P.J.A. Burt<br />

Disease progress across the small plot was uneven; plants in separate quadrats<br />

showed different rates <strong>of</strong> disease progression (Figure 4). Wind direction was not<br />

consistent but northerly winds were the most frequent. At these times, plants in<br />

quadrat 1 were directly downwind <strong>of</strong> the inoculum source, which might account<br />

for the higher spore catches there (Figure 3). Otherwise, wind direction and the<br />

appearance <strong>of</strong> symptoms were not related (Rutter et al., 1998). Southerly winds<br />

were the second most common and would have blown inoculum over quadrat 3.<br />

However neither the pattern <strong>of</strong> disease spread nor the rotorod catches reflected<br />

this. Southerly winds blowing across quadrat 1 might explain the high spore<br />

catches by rotorods in quadrat 4. Patterns <strong>of</strong> disease spread in the small plot<br />

showed no evidence <strong>of</strong> splash dispersal. Importantly, symptoms were not seen in<br />

the plants immediately surrounding the source <strong>of</strong> innoculum until the middle <strong>of</strong><br />

June, when the upper leaves <strong>of</strong> the plants were affected. Rotorod traps 50 m<br />

downwind <strong>of</strong> the plot had no ascospores, and only a few conidia (up to 12 per<br />

sampling run) (Rutter et al., 1998).<br />

Catch<br />

180<br />

150<br />

120<br />

90<br />

60<br />

30<br />

0<br />

R1&2 (centre) R3&4 (Q2/Q3) R5&6 (Q4) R7&8 (Q1)<br />

Trap number and quadrat<br />

Ascospores<br />

Conidia<br />

Figure 3. Total spore catch in the small plot during the sampling period 1 May-3 August 1995. (From: Rutter<br />

et al., 1998).<br />

% <strong>of</strong> infected plants<br />

150<br />

100<br />

50<br />

Q1<br />

Q2<br />

Q3<br />

Q4<br />

0<br />

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20<br />

Sample number (every 4 days starting 2/5/95)<br />

Figure 4. Appearance <strong>of</strong> disease symptoms in the quadrats <strong>of</strong> the small plot. (From: Rutter et al., 1998).<br />

Mesoscale dispersal<br />

Ascospores and conidia <strong>of</strong> M. musicola and M. fijiensis were present in the air at<br />

all sites, usually between 0530-0830 in the morning with an abrupt cut-<strong>of</strong>f at 08.30<br />

117


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

(Burt et al., 1998). Many more spores <strong>of</strong> M. fijiensis than <strong>of</strong> M. musicola were caught.<br />

The analysis was therefore based on M. fijiensis only (Table 2). Despite the apparent<br />

relationship between spore numbers in traps 1 and 3 (Table 2) the numbers <strong>of</strong> spores<br />

caught in each trap were not statistically related to each other (Burt et al., 1998).<br />

No differences in wind direction were recorded at the three trapping sites. The<br />

winds blew most frequently from north to south, or vice versa, along the valley<br />

axis, as might be expected (Burt et al., 1998). Winds blowing from between the<br />

north-east and south-east would have crossed over trap 3 and the large plantation<br />

at CATIE, the nearest large source. But there was no evidence that the spores were<br />

being blown directly from this source, either directly or after allowing for a lag<br />

between times <strong>of</strong> release and capture (Burt et al., 1998).<br />

There were no other large sources <strong>of</strong> inoculum near trap 3. However, there are<br />

many large plantations on the Atlantic coast, 40 km east <strong>of</strong> the trapping sites.<br />

NE and SE winds could have blown spores from the plantations to the study area<br />

but the cut-<strong>of</strong>f in spore captures around 0830 in the morning does not support<br />

this. It is possible that spores were descending at night at the time when the<br />

nocturnal inversion would have been breaking up.<br />

Table 2. Mean daily spore catches <strong>of</strong> M. fijiensis during the sampling period <strong>of</strong> 29 April–30 August 1995.<br />

Mean daily catches <strong>of</strong> ascospores Mean daily catched <strong>of</strong> conidia n<br />

(standard error <strong>of</strong> the mean)<br />

(standard error <strong>of</strong> the mean)<br />

Trap 1 153.5 72 116<br />

(14.6) (12.1)<br />

Trap 2 124.2 8.5 127<br />

(17.4) (1..2)<br />

Trap 3 144.6 8.2 118<br />

(16.4) (1..2)<br />

From: Burt et al., 1998<br />

Assessment <strong>of</strong> potential innoculum<br />

The numbers <strong>of</strong> perithecia was related to the area <strong>of</strong> necrotic <strong>leaf</strong> tissue at different<br />

stages <strong>of</strong> disease development (Burt et al., 1999):<br />

ln(perithecial number) = [1.173 x ln(necrotic area)] + 4.624<br />

A second regression equation was calculated from data from the release and rewetting<br />

experiments (Burt et al., 1999):<br />

ln(no. ascospores) = [1.173 x ln(necrotic area)] + 6.128<br />

where necrotic area is measured in cm 2 .<br />

This gave a mean <strong>of</strong> 4.5 ascospores per perithecium, a result which does not resolve<br />

the ambiguities surrounding the number <strong>of</strong> ascospores present in each perithecium<br />

reported in the literature, with values ranging from between 1 and 27 (Stahel, 1937;<br />

Stover, 1972) and up to 160 (Meredith and Lawrence, 1970; Stover, 1970) being<br />

reported. Even if there were some errors in the counting <strong>of</strong> the released ascospores,<br />

it is clear that the number <strong>of</strong> ascospores available for release is not being represented<br />

in spore trap catches.<br />

118


Session 2<br />

P.J.A. Burt<br />

Ascospore viability<br />

Exposure to UV radiation at tropical sunlight levels killed ascospores after six hours<br />

continuous exposure (Parnell et al., 1998). Although spores in the air are unlikely<br />

to be exposed to continuous sunlight or temperature for these lengths <strong>of</strong> time, this<br />

gives some indication <strong>of</strong> the potential for long-distance transport <strong>of</strong> viable spores.<br />

A 12-hour transport time in winds <strong>of</strong>, for example, 10 m/s, would permit transport<br />

over distances up to 400 km. This might not preclude the transport <strong>of</strong> viable spores<br />

over that distance. In Costa Rica, however, the prevailing winds are <strong>of</strong>f the sea on<br />

both coasts, so windborne transport towards the Caribbean is unlikely. This<br />

combination <strong>of</strong> unfavourable winds and viability may explain why black <strong>leaf</strong> streak<br />

disease has not reached the Caribbean in significant amounts, and why localised<br />

disease outbreaks there have not spread. Looking at spore viability in relation to<br />

distance (Burt, 1994), suggests that wind dispersal <strong>of</strong> M. fijiensis ascospores over<br />

longer distances is unlikely. For example, a trans-Atlantic transport time <strong>of</strong> even 5<br />

days–which is probably the shortest period possible (Rosenberg and Burt, 1999)–is<br />

significantly longer than the time spores are likely to remain viable, even assuming<br />

that they were carried under cloudy skies all the time.<br />

Suggestions for future research<br />

It is clear that more research is required if the full contribution <strong>of</strong> airborne dispersal to<br />

the epidemiology <strong>of</strong> M. fijiensis is to be fully understood. It is clear that much more<br />

inoculum is being produced on the leaves than is entering the air. Consequently, there<br />

is a need to understand more fully the relationship between wind patterns and spore<br />

movement on all scales, but particularly updraughts in relation to spore movement inside<br />

and out <strong>of</strong> a canopy.<br />

Detailed studies <strong>of</strong> the <strong>leaf</strong>-surface are also required, to measure run<strong>of</strong>f (the number<br />

<strong>of</strong> spores transported and their destination) and plant architecture and micrometeorology.<br />

Rather than relying on rewetting experiments, which may be prone to error, detailed<br />

histological investigations <strong>of</strong> necrotic tissue at various stages <strong>of</strong> the disease may also<br />

reveal more information about the supply <strong>of</strong> inoculum (especially the number <strong>of</strong><br />

ascospores).<br />

Finally, a more complex field investigation is needed to resolve whether or not spores<br />

actually travel over long distances. This should be undertaken in association with a<br />

study to determine the effect <strong>of</strong> natural environmental conditions on the viability <strong>of</strong><br />

spores found at various distances from known sources.<br />

Acknowledgements<br />

The material summarised in this paper was prepared over a period <strong>of</strong> seven years and involved<br />

many people. Specifically, the support and assistance <strong>of</strong> Dr Elkin Bustamente, Principal<br />

Pathologist, CATIE, Costa Rica, is most gratefully acknowledged. The following also made<br />

valuable contributions to the research: John Rutter, Herbert Gonzales, Francisco Ramirez,<br />

Kate Wilson, Mark Parnell, Margaret Smith, Jane Rosenberg, Sheila Green, John Sherington<br />

and Philip Shannon. Funding was provided by the Crop Protection Programme <strong>of</strong> the<br />

Department <strong>of</strong> International Development <strong>of</strong> the United Kingdom, who can accept no<br />

responsibility for any information provided, nor views expressed.<br />

119


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

References<br />

Bashi E. and D.E. Aylor 1983. Survival <strong>of</strong> detached sporangia <strong>of</strong> Peronospora destructor and<br />

Peronospora tabacina. Phytopathology 73:1135-1139.<br />

Bowden J., P.H. Gregory and C.G. Johnson. 1971. Possible wind transport <strong>of</strong> c<strong>of</strong>fee rust across<br />

the Atlantic Ocean. Nature 229:500-501.<br />

Burdekin D.A. 1960. Wind and water dispersal <strong>of</strong> c<strong>of</strong>fee <strong>leaf</strong> rust in Tanganyika. Kenya C<strong>of</strong>fee,<br />

212-213.<br />

Burt P.J.A. 1994. Windborne dispersal <strong>of</strong> Sigatoka <strong>leaf</strong> <strong>spot</strong> pathogens. Grana 33:108-111.<br />

Burt P.J.A., J. Rutter and H. Gonzales. 1997. Short-distance windborne dispersal <strong>of</strong> the fungal<br />

pathogens causing Sigatoka <strong>diseases</strong> in banana and plantain. Plant Pathology 46:451-458.<br />

Burt P.J.A., J. Rutter and F. Ramirez. 1998. Airborne spore loads and mesoscale dispersal <strong>of</strong><br />

the fungal pathogens causing Sigatoka <strong>diseases</strong> in banana and plantain. Aerobiologia<br />

14:209-214.<br />

Burt P.J.A., L.J. Rosenberg, J. Rutter, F. Ramirez and H. Gonzales. 1999. Forecasting the airborne<br />

spread <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis, a cause <strong>of</strong> black Sigatoka disease on <strong>bananas</strong>:<br />

estimations <strong>of</strong> numbers <strong>of</strong> perithecia and ascospores to aid forecasting. Annals <strong>of</strong> Applied<br />

Biology 135:367-377.<br />

Chuang T.Y. and H.J. Su. 1988. Physiological study <strong>of</strong> Fusarium oxysporum f.sp. cubense.<br />

Memoirs <strong>of</strong> the College <strong>of</strong> Agriculture, National Taiwan University 28:19-26.<br />

Davies J.M., C.E. Main and W.C. Nesmith. 1985. The biometeorology <strong>of</strong> blue mold <strong>of</strong> tobacco:<br />

Part 2, the evidence for long-range sporangiospore transport. Pp. 473-498 in Movement<br />

and dispersal <strong>of</strong> agriculturally-important biotic agents (Mackenzie, Barfield, Kennedy,<br />

Berger and Taranto, eds).<br />

Gauhl F. 1989. Untersuchingen zur Epidemiologie und Okologie der Schwarzen Sigatoka-<br />

Krankheit (<strong>Mycosphaerella</strong> fijiensis Morelet) an Kochbananen (Musa sp.) in Costa Rica.<br />

PhD thesis, University <strong>of</strong> Gottingen, Germany.<br />

Leach R. 1941. Banana <strong>leaf</strong><strong>spot</strong> (<strong>Mycosphaerella</strong> musicola) the perfect stage <strong>of</strong> Cercospora<br />

musae Zimm. Tropical Agriculture Trinidad 18:91-95.<br />

Leach R. 1946. Banana <strong>leaf</strong><strong>spot</strong> (<strong>Mycosphaerella</strong> musicola) on the Gros Michel variety in<br />

Jamaica. Kingston, Jamaica: The Government Printer.<br />

Maddison A.C. and J.G. Manners. 1972. Sunlight and viability <strong>of</strong> cereal rust uredospores.<br />

Transactions <strong>of</strong> the British Mycological Society 59:429-443.<br />

Meredith D.S. 1962. Some components <strong>of</strong> the air-spora in Jamaican banana plantations. Annals<br />

<strong>of</strong> Applied Biology 50:577-594.<br />

Meredith D.S. and J.S. Lawrence. 1970. Black <strong>leaf</strong> streak disease <strong>of</strong> <strong>bananas</strong> (<strong>Mycosphaerella</strong><br />

fijiensis): symptoms <strong>of</strong> disease in Hawaii, and notes on the conidial state <strong>of</strong> the causal<br />

fungus. Transactions <strong>of</strong> the British Mycological Society 52:459-476.<br />

Meredith D.S., J.S. Lawrence and I.D. Firman. 1973. Ascospore release and dispersal in black<br />

<strong>leaf</strong> streak <strong>of</strong> <strong>bananas</strong> (<strong>Mycosphaerella</strong> fijiensis). Transactions <strong>of</strong> the British Mycological<br />

Society 60:547-554.<br />

Parnell M., P.J.A. Burt and K. Wilson. 1998. The influence <strong>of</strong> exposure to ultraviolet radiation<br />

in simulated sunlight on ascospores causing black Sigatoka disease <strong>of</strong> banana and plantain.<br />

Int. J. Biometeorol. 42(1):22-27.<br />

Pedgley D.E. 1982. Windborne pests and <strong>diseases</strong>. Chichester: Ellis Horwood.<br />

Rosenberg L.J. and P.J.A. Burt. 1999. Windborne displacements <strong>of</strong> Desert Locusts from Africa<br />

to the Caribbean and South America. Aerobiologia 15:167-175.<br />

Rotem J. and H.J. Aust. 1991. The effect <strong>of</strong> ultraviolet and solar radiation and temperature<br />

on survival <strong>of</strong> fungal propagates. Journal <strong>of</strong> Phytopathology 133:76-84.<br />

120


Session 2<br />

P.J.A. Burt<br />

Rotem J., B. Wooding and D.E. Aylor 1985. The role <strong>of</strong> solar radiation, especially ultraviolet,<br />

in the mortality <strong>of</strong> fungal; spores. Phytopathology 75:510-514.<br />

Rutter J. and P.J.A. Burt. 1997. An assessment <strong>of</strong> the levels <strong>of</strong> inoculum <strong>of</strong> <strong>Mycosphaerella</strong><br />

fijiensis in a banana plantation. Pp. 229-242 in Aerobiology (S.N. Agashe ed.), Oxford<br />

and IBH Publishing Co. Pvt Ltd, New Delhi.<br />

Rutter J., P.J.A. Burt and F. Ramirez. 1998. Movement <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis spores and<br />

Sigatoka disease development on plantain close to an inoculum source. Aerobiologia<br />

14:201-208.<br />

Setlow R.B. 1974. The wavelength in sunlight effective in producing skin cancer: a theoretical<br />

analysis. Proceedings <strong>of</strong> the National Academy <strong>of</strong> Science 71:3363-3366.<br />

Stahel G. 1937. Notes on Cercospora <strong>leaf</strong> <strong>spot</strong> <strong>of</strong> <strong>bananas</strong> (Cercospora musae). Tropical<br />

Agriculture Trinidad 14:2587-264.<br />

Stover R.H. 1968. Leaf <strong>spot</strong> <strong>of</strong> <strong>bananas</strong> caused by <strong>Mycosphaerella</strong> musicola: perithecia and<br />

sporodochia production in different climates. Tropical Agriculture Trinidad 45:1-12.<br />

Stover R.H. 1970. Leaf <strong>spot</strong> <strong>of</strong> <strong>bananas</strong> caused by <strong>Mycosphaerella</strong> musicola: role <strong>of</strong> conidia<br />

in epidemiology. Phytopathology 60:856-860.<br />

Stover R.H. 1972. Banana, plantain and abaca <strong>diseases</strong>. Kew: Commonwealth Mycological<br />

Institute.<br />

Stover R.H. and N.W. Simmonds. 1987. Bananas. Harlow, England: Longman<br />

121


Session 2<br />

J. Carlier et al.<br />

Genetic differentiation in<br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> pathogens<br />

J. Carlier 1 ,H.Hayden 2 ,G.Rivas 3 ,M.-F.Zapater 1 ,C.Abadie 1 and E. Aitken 4<br />

Abstract<br />

Black <strong>leaf</strong> streak disease and Sigatoka disease, caused respectively by two related fungi,<br />

<strong>Mycosphaerella</strong> fijiensis and M. musicola, are the most important <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>.<br />

Understanding the genetic structure <strong>of</strong> the populations and the evolution <strong>of</strong> these pathogens is<br />

an important aid in breeding and managing disease resistance. The population structure <strong>of</strong> each<br />

pathogen was analysed using molecular markers mainly at the global and continental scales.<br />

Features common to both were observed: 1) the centre <strong>of</strong> diversity is located in Southeast Asia<br />

and founder events accompanying the introduction <strong>of</strong> the pathogens in other regions led to a<br />

reduction <strong>of</strong> genetic diversity; 2) genetic diversity is maintained in all populations and is also present<br />

at the scale <strong>of</strong> the plant; 3) genetic recombination played an important role in the genetic structure<br />

<strong>of</strong> both pathogens; 4) genetic differentiation exists between populations from the global to the<br />

local level. The main difference between the two species had to do with the measures <strong>of</strong> genetic<br />

differentiation. Whereas the African populations <strong>of</strong> M. fijiensis were significantly different from<br />

the Latin American/Caribbean ones, no significant difference was observed between the African<br />

and Latin American/Caribbean populations <strong>of</strong> M. musicola.This suggests independent introductions<br />

<strong>of</strong> M. fijiensis but not <strong>of</strong> M. musicola.Except for this situation, the genetic differentiation observed<br />

between populations at the global and continental scales indicate an important effect <strong>of</strong> genetic<br />

drift and limited gene flow.<br />

Resumen - Diferenciación genética en los patógenos de la mancha foliar Mycosphaeralla<br />

La enfermedades de la mancha foliar de los bananos más importantes se deben a dos hongos<br />

relacionados: <strong>Mycosphaerella</strong> fijiensis y M. musicola, los agentes causales de la enfermedad de la<br />

raya negra de la hoja y de la enfermedad de Sigatoka, respectivamente. El entendimiento, tanto<br />

de la estructura genética de la población, como de la evolución de estos patógenos proporciona<br />

información importante para brindar asistencia al mejoramiento y manejo de la resistencia a la<br />

enfermedad. La estructura de la población de ambos patógenos fue analizada utilizando<br />

marcadores moleculares a escalas global y continental. Se observaron las siguientes características<br />

comunes: 1) el centro de la diversidad está localizado en el Sudeste de Asia y los eventos de<br />

1<br />

CIRAD, Montpellier, France<br />

2<br />

Institute <strong>of</strong> Land and Food Resources, Victoria, Australia<br />

3<br />

CATIE, Costa Rica<br />

4<br />

CRCTPP, Brisbane, Australia<br />

123


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

colonización que acompañaron la introducción de los patógenos en otras regiones han llevado a<br />

una reducción de la diversidad genética en comparación con el Sudeste de Asia; 2) la diversidad<br />

genética se mantiene en todas las poblaciones y se distribuye a escala de la planta, 3) la<br />

recombinación genética desempeña un papel importante en la estructura genética de ambos<br />

patógenos; 4) existe una diferenciación genética entre las poblaciones a escalas de global a local.<br />

La principal diferencia observada es la existencia de una diferenciación genética entre las<br />

poblaciones africanas y las poblaciones latinoamericanas y caribeñas de M. fijiensis pero no de<br />

M. musicola. Este resultado sugiere introducciones independientes de M. fijiensis pero no de<br />

M. musicola. Con excepción de la situación descrita arriba, la diferenciación genética observada<br />

en ambos patógenos entre las poblaciones a escalas global y continental indica un efecto<br />

importante de la genética y un flujo de genes bajo.<br />

Résumé - Différenciation génétique chez les <strong>Mycosphaerella</strong> pathogènes<br />

La maladie des raies noires et la maladie de Sigatoka, respectivement causées par deux<br />

champignons apparentés, <strong>Mycosphaerella</strong> fijiensis et M. musicola, sont les maladies foliaires les<br />

plus importantes chez le bananier. La compréhension de la structure génétique des populations<br />

et de l’évolution de ces pathogènes représente une aide importante pour l’amélioration et la gestion<br />

de la résistance à ces maladies. La structure de la population de chaque pathogène a été analysée<br />

en utilisant des marqueurs moléculaires, principalement à l’échelle globale et continentale. Des<br />

caractéristiques communes aux deux pathogènes ont été observées : 1) leur centre de diversité<br />

est localisé en Asie du Sud-est et des événements fondateurs accompagnant l’introduction des<br />

pathogènes dans d’autres régions ont conduit à une réduction de la diversité génétique ; 2) la<br />

diversité génétique est maintenue dans toutes les populations et est également présente à l’échelle<br />

de la plante ; 3) la recombinaison génétique a joué un rôle important dans la structure génétique<br />

des deux pathogènes ; 4) une différentiation génétique existe entre populations, du niveau global<br />

au niveau local. La principale différence entre les deux espèces concerne les niveaux de la<br />

différentiation génétique. Alors que les populations africaines de M. fijiensis sont significativement<br />

différentes de celles d’Amérique latine/Caraïbes, aucune différence significative n’a été observée<br />

entre les populations de M. musicola originaires d’Afrique et d’Amérique latine/Caraïbes. Ceci suggère<br />

des introductions indépendantes de M. fijiensis,mais pas de M. musicola.A part cette situation,<br />

la différentiation génétique observée entre les populations à l’échelle globale et continentale indique<br />

un effet important de la dérive génétique et des flux géniques limités.<br />

Introduction<br />

Black <strong>leaf</strong> streak disease, caused by <strong>Mycosphaerella</strong> fijiensis, and Sigatoka disease,<br />

caused by M. musicola, are the most important <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong> (Jones,<br />

2000). The fungi are haploid and heterothallic. The anamorph and teleomorph stages<br />

are both present on infected leaves, and the ascospores produced during the sexual<br />

stage play an important role in epidemics. The first species to be described was<br />

M. musicola, in Java in 1902. The rapid dissemination <strong>of</strong> Sigatoka disease round the<br />

world in the 1930s, led to speculations that the spores were carried by air currents<br />

between continents: from Asia to the Pacific, from the Pacific to Australia, from<br />

Australia to Africa and from Africa to Latin America (Stover, 1962). In 1962,<br />

M. musicola was present in all banana-producing regions, making Sigatoka disease<br />

one <strong>of</strong> the most important plant <strong>diseases</strong>. Black <strong>leaf</strong> streak disease was reported in<br />

Fiji in 1964 and since then has been reported throughout the Pacific and Asia. The<br />

chronology <strong>of</strong> records suggests that M. fijiensis originated, as M. musicola, in<br />

Southeast Asia (Mourichon and Fullerton, 1990), which is also the centre <strong>of</strong> origin<br />

124


Session 2<br />

J. Carlier et al.<br />

<strong>of</strong> the host genus Musa. Starting in the 1970s, M. fijiensis spread to Africa, Latin<br />

America and the Caribbean. Being more aggressive, M. fijiensis replaced M. musicola<br />

as the dominant <strong>leaf</strong> <strong>spot</strong> pathogen in many areas. Although the distribution <strong>of</strong> both<br />

pathogens is well documented in Australia, the Pacific region, Africa, Latin America<br />

and the Caribbean, it is still not well understood in Southeast Asia.<br />

Why and how to analyse populations <strong>of</strong> pathogens?<br />

Knowledge <strong>of</strong> the genetic population structure and evolution <strong>of</strong> the pathogens is an<br />

important aid in breeding and managing disease resistance. The main objective <strong>of</strong><br />

such study is to provide information on the level and distribution <strong>of</strong> variability.<br />

Molecular markers are <strong>of</strong>ten used to analyse population structure at different<br />

geographical scales. It should make it possible to identify potential sources <strong>of</strong><br />

resistance, which are expected to be in areas where the diversity <strong>of</strong> pathogens and<br />

host are high. It should also ensure that the diversity <strong>of</strong> pathogens used when screening<br />

for resistance is representative <strong>of</strong> the one in the regions where resistant hosts are<br />

intended to be used. Pathogens can evolve to break down total resistance or erode<br />

partial resistance. The evolution <strong>of</strong> pathogen populations depends on mutation,<br />

recombination, genetic drift, gene flow and the selection pressure exerted by the host.<br />

It should be possible to limit and restrict the evolution <strong>of</strong> pathogenicity by varying<br />

host resistance in space and time. Such strategies should improve the durability <strong>of</strong><br />

the types <strong>of</strong> resistance used and ensure the durability <strong>of</strong> the culture.<br />

Another objective <strong>of</strong> pathogen population studies is to evaluate the relative<br />

importance <strong>of</strong> the evolutionary factors on the evolution <strong>of</strong> pathogens. Such<br />

information would make it possible to model and test the effect <strong>of</strong> different<br />

management strategies on the evolution <strong>of</strong> the pathogen. Analysing population<br />

structure through space allows us to evaluate the effects <strong>of</strong> genetic recombination,<br />

genetic drift and gene flow on the evolution <strong>of</strong> the pathogen. This paper reviews the<br />

results obtained at global and local scales for M. musicola and M. fijiensis. A second<br />

approach consists in studying the evolution <strong>of</strong> the pathogen in fields <strong>of</strong> resistant hosts<br />

by using molecular markers and by characterizing pathogenicity. This allows us to<br />

evaluate the effect <strong>of</strong> the selection pressure exerted by the host on the pathogen. This<br />

second approach is described in another paper (see Abadie et al. in this volume).<br />

Global population structure<br />

RFLP markers were used to study the genetic structure <strong>of</strong> the global population <strong>of</strong><br />

M. musicola and <strong>of</strong> M. fijiensis (Carlier et al., 1996; Hayden et al., in prep.) Random<br />

single-locus probes were used on samples from Southeast Asia, Australia, the Pacific<br />

Islands, Africa, Latin America and the Caribbean. Features common to both pathogens<br />

were observed.<br />

Southeast Asia has the highest level <strong>of</strong> gene diversity (Table 1) and the majority<br />

<strong>of</strong> alleles found in this region were also present in the other regions. This supports<br />

the hypothesis that the pathogens originated in Southeast Asia. Founder events<br />

accompanying the introduction <strong>of</strong> the pathogens to other regions have led to a reduction<br />

in genetic diversity in comparison with Southeast Asia. Nevertheless, genetic diversity<br />

125


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

is maintained in all populations. Ecological conditions being favourable for disease<br />

development and banana cultivation almost year round in most growing areas, low<br />

genetic drift in large pathogen populations can maintain the high levels <strong>of</strong> genetic<br />

diversity observed. Therefore, a certain level <strong>of</strong> variability in pathogenicity might also<br />

be maintained in pathogen populations, allowing the pathogen to attack newly<br />

introduced resistant genotypes, as has been observed with M. fijiensis on ‘Paka’ and<br />

‘T8’ in Rarotonga, Cook Islands (Fullerton and Olsen, 1995). The existence <strong>of</strong> specific<br />

interactions between the host and M. fijiensis isolates was suggested for highly resistant<br />

genotypes (Fullerton and Olsen, 1995). Variability in aggressiveness was evaluated for<br />

two M. fijiensis samples from Cameroon and the Philippines by inoculating five partially<br />

resistant cultivars using a <strong>leaf</strong> piece assay (El Hadrami, 2000). Variability was similar<br />

but low for both countries, however genetic diversity in the Philippines was much higher<br />

(Carlier et al., 1996). Specific interactions between the isolates and the cultivar were<br />

not detected. Only susceptible hosts are cultivated in these countries, and the lack <strong>of</strong><br />

a selection pressure being exerted by the host on the pathogens could explain the results.<br />

The potential <strong>of</strong> pathogen populations to overcome partial resistance should be analysed<br />

by following their evolution in fields <strong>of</strong> resistant hosts (see Abadie et al., this<br />

proceedings).<br />

Genetic recombination plays an important role in the genetic structure <strong>of</strong><br />

M. musicola and M. fijiensis. Genetic markers were statistically independent therefore<br />

characteristics <strong>of</strong> pathogenicity could not be related to RFLP genotypes. With regards<br />

to breeding programmes, introducing specific resistance genes in individual cultivars<br />

(pyramiding) may not be a strategy for durable resistance in banana. Mixing varieties<br />

or using partially resistant hosts might be more appropriate.<br />

Table 1. Nei’s gene diversity estimates for populations <strong>of</strong> M. fijiensis (Carlier et al., 1996) and M. musicola (Hayden<br />

et al., in prep.) from different geographical regions.<br />

Species Asia Africa Latin America and Caribbean Australia and Pacific<br />

M. fijiensis 0.57 0.25 0.40 0.28<br />

M. musicola 0.41 0.27 0.21 0.33<br />

A high level <strong>of</strong> genetic differentiation was observed between populations <strong>of</strong><br />

M. musicola and M. fijiensis from different regions (Figure 1). The Fst parameter (Wright,<br />

1951) estimated for all loci between pairs <strong>of</strong> populations varied between 0.14 and 0.58<br />

for M. fijiensis and between 0.025 and 0.55 for M. musicola. But whereas the African<br />

populations <strong>of</strong> M. fijiensis were significantly different from the Latin<br />

American/Caribbean ones (Fst = 0.49), no significant difference was observed between<br />

the African and Latin American populations <strong>of</strong> M. musicola (Fst = 0.025, not significant).<br />

This suggests a separate introduction <strong>of</strong> M. fijiensis in each region but a common one<br />

for M. musicola.<br />

On the other hand, the high levels <strong>of</strong> genetic differentiation observed between<br />

Australian and African populations <strong>of</strong> M. musicola (Fst = 0.47) does not support the<br />

hypothesis <strong>of</strong> Stover (1962) whereby spores <strong>of</strong> M. musicola were carried by air currents<br />

from Australia to Africa. In general, the high level <strong>of</strong> genetic differentiation <strong>of</strong> both<br />

pathogens at a global scale suggests occasional migration events between continents.<br />

126


Session 2<br />

J. Carlier et al.<br />

Long distance dissemination <strong>of</strong> the disease around the world was more likely to have<br />

occurred by movement <strong>of</strong> infected plant material, as proposed by Mourichon and<br />

Fullerton (1990).<br />

Africa<br />

Latin America<br />

Latin America/<br />

Caribbean<br />

Indonesia<br />

Pacific<br />

Islands<br />

Africa<br />

A<br />

Philippines<br />

Papua New Guinea<br />

B<br />

Australia<br />

Figure 1. Global population structure <strong>of</strong> M. fijiensis (Carlier et al., 1996) and M. musicola (Hayden etal., in prep.).<br />

Additive trees constructed from estimates <strong>of</strong> Wright’s Fst among pairs <strong>of</strong> geographical populations.<br />

Continental and local population structures<br />

Population structures at a continental scale were studied in Australia for<br />

M. musicola (Hayden, 2000). Collections <strong>of</strong> isolates from twelve sites along the<br />

east coast were analysed using 15 RFLP markers. The level <strong>of</strong> gene diversity (Nei,<br />

1978), varied between 0.14 and 0.37. On a plant, the pathogen isolated from a<br />

given lesion would <strong>of</strong>ten be a haplotype not found in the other lesions, meaning<br />

that diversity is also present at a fine scale. Low to high levels <strong>of</strong> genetic<br />

differentiation were observed between populations (Fst = 0.04-0.45). There was<br />

no apparent correlation between genetic and geographical distances as high levels<br />

<strong>of</strong> genetic differentiation were observed between neighbouring populations and<br />

low levels were observed in populations separated by long distances.<br />

The population structure <strong>of</strong> M. fijiensis was analysed in Africa, Latin America<br />

and the Caribbean (Rivas et al., subm.). Samples from different countries were<br />

characterized using CAPS (Cleaved Amplified Polymorphic Sequence) markers<br />

(Zapater et al., subm.). The results obtained for both continents were similar. The<br />

value <strong>of</strong> gene diversity varied between 0.19 and 0.38 for Africa, and 0.16 and<br />

0.36 for the Latin America/Caribbean region. The low levels detected in some<br />

populations suggest that founder effects occurred during the spread <strong>of</strong> the disease<br />

on both continents. In the Latin America and Caribbean region, the highest levels<br />

are observed in populations from Honduras and Costa Rica, supporting the<br />

suggestion that the pathogen entered the continent in this area. In one locality<br />

in Cameroon, the values <strong>of</strong> gene diversity estimated at the scale <strong>of</strong> the field and<br />

<strong>of</strong> the plant are similar suggesting, as with M. musicola, that diversity is distributed<br />

at a fine scale.<br />

Important levels <strong>of</strong> genetic differentiation were detected between most <strong>of</strong> the<br />

populations (Fst = 0.04-0.45 for Africa and 0.01-0.56 for Latin America/Caribbean).<br />

127


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

There is sufficient differentiation between populations in the Caribbean islands<br />

to support the hypothesis that there was more than one introduction from Latin<br />

America. In Cameroon, genetic differentiation was detected between localities 300<br />

km apart but not between localities 200 km apart.<br />

Finally, the genetic structure <strong>of</strong> M. fijiensis was studied in the Australasia/<br />

Pacific region using RFLP markers (Hayden et al., subm.). Genetic differentiation<br />

was detected between the Torres Straits Islands, Pacific Islands and Papua New<br />

Guinea. At a local scale, there was no differentiation between two sites in the<br />

small Mer island.<br />

The levels <strong>of</strong> genetic differentiation observed at the continental scale for both<br />

pathogens suggest an important effect <strong>of</strong> genetic drift on population structure<br />

and limited gene flow. Thus, spread <strong>of</strong> the <strong>diseases</strong> within continents could be<br />

due to the movement <strong>of</strong> infected plants and/or very restricted ascospore dispersal.<br />

As populations are probably not in genetic equilibrium, gene flow resulting from<br />

ascospore dispersal may be underestimated. However, preliminary results from<br />

an epidemiological study <strong>of</strong> black <strong>leaf</strong> streak disease suggests that dispersal <strong>of</strong><br />

the pathogen is more restricted than previously thought. The results suggest a<br />

dispersal gradient <strong>of</strong> about 25 m from an inoculum source (Abadie et al., this<br />

proceedings).<br />

Conclusion and perspectives<br />

The population structures <strong>of</strong> M. musicola and M. fijiensis are now better known<br />

at different geographical scales. However, at a regional scale few samples from<br />

Southeast Asia have been analysed. A new pathogen, <strong>Mycosphaerella</strong> eumusae,<br />

was recently discovered and detected mainly in Southeast Asia (Carlier et al., 2000).<br />

Southeast Asia is not only the centre <strong>of</strong> origin <strong>of</strong> all three pathogens but also <strong>of</strong><br />

the host genus Musa. The distribution <strong>of</strong> the pathogens and their population<br />

structure should now be determined in detail for this region. Host-pathogen<br />

interactions could differ for each pathogen. One hypothesis to explain the<br />

continued presence <strong>of</strong> the three pathogens in Southeast Asia is the high diversity<br />

<strong>of</strong> host species. The hypothesis could be tested by surveying the fungal species<br />

in relation to host diversity. If host-pathogen interactions differ, the resistance<br />

genes introduced to produce new varieties could be more or less efficient<br />

depending on the pathogen they are exposed to. Their utilization should take into<br />

account the distribution <strong>of</strong> pathogen species. Zones <strong>of</strong> co-evolution for the three<br />

pathogens could be localized in Southeast Asia. This area is a potential source<br />

<strong>of</strong> resistance, therefore a study <strong>of</strong> pathogen populations in natural systems should<br />

provide us with information to complement evaluations <strong>of</strong> the relative importance<br />

<strong>of</strong> the different evolutionary forces.<br />

The results to date suggest that genetic drift has an important effect on the<br />

structure <strong>of</strong> pathogen populations and that gene flow is limited. The limit <strong>of</strong><br />

ascospore dispersal should be estimated indirectly at a country scale using genetic<br />

models such as the isolation by distance model (Rousset, 1997). However, we can<br />

already predict that the improvement <strong>of</strong> quarantine measures at the continental<br />

scale might limit the risk <strong>of</strong> introducing the disease to new areas, and limit the<br />

128


Session 2<br />

J. Carlier et al.<br />

exchange between existing pathogen populations from different countries. At a<br />

country and local level, geographical obstacles could also limit exchange between<br />

populations from different fields by playing on gene flow. Such a measure could<br />

have an impact on the durability <strong>of</strong> the resistances and <strong>of</strong> the management<br />

strategies.<br />

References<br />

Carlier J., M.F Zapater, F. Lapeyre, D.R. Jones and X Mourichon. 2000. Septoria <strong>leaf</strong> <strong>spot</strong> <strong>of</strong><br />

banana: A newly discovered disease caused by <strong>Mycosphaerella</strong> eumusae (anamorph<br />

Septoria eumusae). Phytopathology 90(8):884-890.<br />

Carlier J., M.H. Lebrun, M.F. Zapater, C. Dubois and X. Mourichon. 1996. Genetic structure<br />

<strong>of</strong> the global population <strong>of</strong> <strong>bananas</strong> black <strong>leaf</strong> streak fungus <strong>Mycosphaerella</strong> fijiensis.<br />

Molecular Ecology 5:499-510<br />

El Hadrami A. 2000. Caractérisation de la résistance partielle des bananiers à la maladie des<br />

raies noires et évaluation de la variabilité de l’agressivité de l’agent causal, <strong>Mycosphaerella</strong><br />

fijiensis. Thèse d’Université. Faculté Universitaire des Sciences Agronomiques de Gembloux,<br />

Belgium. 153pp.<br />

Fullerton R.A. and T.L. Olsen. 1995. Pathogenic variability in <strong>Mycosphaerella</strong> fijiensis Morelet<br />

cause <strong>of</strong> black Sigatoka in banana and plantain. New Zealand Journal <strong>of</strong> Crop and<br />

Horticultural Science 23:39-48.<br />

Hayden H.L., J. Carlier and E.A.B Aitken. (In preparation). Population differentiation in the<br />

banana <strong>leaf</strong> <strong>spot</strong> pathogen <strong>Mycosphaerella</strong> musicola, examined at a global scale.<br />

Hayden H.L., J. Carlier and E.A.B. Aitken. (Submitted). The genetic structure <strong>of</strong> <strong>Mycosphaerella</strong><br />

fijiensis from Australia, Papua New Guinea and the Pacific Islands.<br />

Hayden H.L. 2000. Population genetic studies <strong>of</strong> <strong>Mycosphaerella</strong> species infecting banana.<br />

Thesis, University <strong>of</strong> Queensland, Australia.<br />

Jones D. 2000. Diseases <strong>of</strong> Banana, Abacá and Enset. CAB International, Wallingford, UK.<br />

Mourichon X. and R.A. Fullerton. 1990. Geographical distribution <strong>of</strong> the two species<br />

<strong>Mycosphaerella</strong> musicola Leach (Cercospora musae) and M. fijiensis Morelet (C. fijiensis),<br />

respectively agents <strong>of</strong> Sigatoka disease and black <strong>leaf</strong> streak disease in Bananas and<br />

Plantains. Fruits 45:213-218.<br />

Nei M. 1978. Estimation <strong>of</strong> average heterozygosity and genetic distances from a small number<br />

<strong>of</strong> individuals. Genetics 89:583-590.<br />

Rivas G. G., M.-F. Zapater, C. Abadie and J. Carlier. (Submitted). Founder effect and stochastic<br />

dispersal at the continental scale <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis, a tropical fungal pathogen<br />

<strong>of</strong> <strong>bananas</strong> that has recently spread in Latin America, the Caribbean and Africa.<br />

Rousset F. 1997. Genetic differentiation and estimation <strong>of</strong> gene flow from F-statistic under<br />

isolations by distance. Genetics 145:1219-1228.<br />

Stover R.H. 1962. Intercontinental spread <strong>of</strong> banana <strong>leaf</strong> <strong>spot</strong> (<strong>Mycosphaerella</strong> musicola Leach).<br />

Tropical Agriculture Trinidad 39(4):327-338.<br />

Wright S. 1951. The genetical structure <strong>of</strong> populations. Annals <strong>of</strong> Eugenics 15:323-354.<br />

Zapater M.F., A. Rakotonantoandro, F. Cohen, J. Carlier. (Submitted). CAPS (Cleaved<br />

Amplified Polymorphic Sequence) markers for the fungal banana pathogen <strong>Mycosphaerella</strong><br />

fijiensis.<br />

129


Session 2<br />

C. Molina et al.<br />

Development and application <strong>of</strong><br />

molecular markers in <strong>Mycosphaerella</strong><br />

populations in Colombia<br />

C. Molina 1,2 , S. Aponte 1 ,A.Gutiérrez 1 ,V.Núñez 1 and G. Kahl 2<br />

Abstract<br />

In order to design effective strategies against the <strong>Mycosphaerella</strong> banana pathogens M. fijiensis<br />

and M. musicola, it is essential to have information on genetic diversity and population<br />

composition. Information to understand the population dynamics <strong>of</strong> these banana pathogens<br />

was based on microsatellite markers. The present study reports tests on 48 primer pairs<br />

designed for M. musicola, <strong>of</strong> which 26 proved to be polymorphic and four were transferable<br />

to M. fijiensis. Based on microsatellites, a comparison was made <strong>of</strong> the genetic variability in<br />

M. fijiensis and M. musicola populations from Colombia, Costa Rica and Venezuela. Dendograms<br />

for each species were based on the Dice similarity algorithm and grouped with the UPGMA<br />

clustering method. With the exception <strong>of</strong> a few isolates, most clusters coincided with the<br />

geographical locations (sampling sites).<br />

Resumen - Desarrollo y aplicación de los marcadores moleculares en las poblaciones<br />

de <strong>Mycosphaerella</strong> en Colombia<br />

Con el fin de diseñar estrategias eficaces contra los patógenos de <strong>Mycosphaerella</strong>, M. fijiensis<br />

y M. musicola, en el banano, es esencial disponer de la información sobre la diversidad<br />

genética y composición de sus poblaciones. La información, con ayuda de la cual se entienden<br />

las dinámicas de las poblaciones de estos patógenos de banano, se basa en los marcadores<br />

de microsatélites. El estudio actual presenta los informes de las pruebas de 48 pares de<br />

iniciadores diseñados para M. musicola, de los cuales 26 resultaron ser polimórficos y cuatro<br />

transferibles a M. fijiensis. Basándose en los microsatélites, se hizo la comparación de la<br />

variabilidad genética en las poblaciones de M. fijiensis y M. musicola procedentes de Colombia,<br />

Costa Rica y Venezuela. Los dendogramas para cada especie se basaron en el algoritmo de<br />

similitud de Dice y se agruparon con el método de análisis de conglomerados UPGMA. Con<br />

excepción de unos pocos aislados, la mayoría de los conglomerados coincidieron con las<br />

localizaciones geográficas (sitios de muestreo).<br />

1<br />

CORPOICA, Bogotá, Colombia<br />

2<br />

University <strong>of</strong> Frankfurt, Frankfurt/Main, Germany<br />

131


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Résumé - Développement et application des marqueurs moléculaires chez des populations<br />

de <strong>Mycosphaerella</strong> en Colombie<br />

Afin de développer des stratégies efficaces contre les <strong>Mycosphaerella</strong> pathogènes du bananier,<br />

M. fijiensis et M. musicola, il est essentiel d’avoir des informations sur la diversité génétique<br />

et la composition des populations. Les informations visant à comprendre la dynamique des<br />

populations de ces pathogènes du bananier ont été basées sur les marqueurs microsatellites.<br />

Cette étude présente les tests effectués avec 48 paires d’amorces conçues pour M. musicola,<br />

dont 26 se sont avérées polymorphes et quatre ont été transférables à M. fijiensis.En se basant<br />

sur les microsatellites, une comparaison de la variabilité génétique a été faite chez des<br />

populations de M. fijiensis et M. musicola originaires de Colombie, du Costa Rica et du<br />

Venezuela. Les dendrogrammes pour chaque espèce ont été basés sur l’algorithme de<br />

similarité de Dice et groupés avec la méthode d’agrégation UPGMA. A l’exception de quelques<br />

isolats, la plupart des agrégats coïncidaient avec les localisations géographiques (lieux de<br />

collecte).<br />

Introduction<br />

<strong>Mycosphaerella</strong> fijiensis and M. musicola, the two most severe fungal pathogens<br />

<strong>of</strong> plantain and banana, are the major cause <strong>of</strong> economic losses in commercial<br />

plantations and in numerous smallholdings. Both pathogens spread around the<br />

world through demographic events such as founder effects. Population bottlenecks<br />

have been caused by increased doses <strong>of</strong> fungicides, the introduction <strong>of</strong> partly<br />

resistant host varieties, and isolation by distance and geographical barriers between<br />

populations.<br />

M. musicola was first reported from Java in 1902 (Mourichon et al., 1997),<br />

and from the Sigatoka Valley, Fiji in 1912 (Leach, 1941), where it caused an<br />

epidemic. About 50 years later a more aggressive pathogen, M. fijiensis, was<br />

detected in the same region (Leach, 1964). Both pathogens rapidly colonized the<br />

South Pacific Islands, Asia, Africa and America (Stover, 1976). M. fijiensis has<br />

been reported from sites where M. musicola was formerly present, suggesting a<br />

gradual displacement <strong>of</strong> M. musicola to higher altitudes (inter-Andean valley<br />

populations). Therefore, the dynamics <strong>of</strong> population structure <strong>of</strong> both pathogens<br />

is in some way interdependent and most likely to be influenced by parameters<br />

common to both.<br />

Direct comparison <strong>of</strong> the populations <strong>of</strong> both pathogens would help understand<br />

local and regional genetic diversity and differentiation, and the influence <strong>of</strong><br />

environmental pressures on the spread <strong>of</strong> the disease to new sites. It could also<br />

help predict the behaviour <strong>of</strong> new epidemics.<br />

Molecular markers have become important tools for the investigation <strong>of</strong> the<br />

genetic composition <strong>of</strong> fungal populations (Groppe and Boller, 1997; Bucheli et<br />

al., 2001). Restriction Fragment Length polymorphism (RFLP) markers were<br />

developed for the M. fijiensis genome, and used to characterize populations <strong>of</strong><br />

M. fijiensis at a regional and global scale (Carlier et al., 1994, 1996; Müller et<br />

al., 1997). More recently, simple sequence repeat (SSR) markers have been<br />

established for M. fijiensis (Neu et al., 1999) and M. musicola (Molina et al., 2001)<br />

which, together with other PCR-based DNA pr<strong>of</strong>iling methods, provide a new<br />

method to compare the genetic diversity <strong>of</strong> both pathogens.<br />

132


Session 2<br />

C. Molina et al.<br />

In the present study, dendograms based on Dice similarity algorithms were obtained<br />

for each species. With the exception <strong>of</strong> a few isolates that grouped outside the main<br />

clusters, the majority <strong>of</strong> individuals grouped according to their geographical location.<br />

Materials and methods<br />

Sampling sites and fungal material<br />

Eighty isolates <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis and 64 <strong>of</strong> M. musicola were collected from<br />

12 locations. Isolates from Colombia were from two adjacent locations on the<br />

Caribbean coast (Santa Marta), and from three sites in the inner valleys <strong>of</strong> the Andes,<br />

Caldas, La Mesa and Cachipay. Isolates from Costa Rica and Venezuela were included<br />

for comparison. Costa Rican isolates were collected in “Valle Central” and those from<br />

Venezuela were collected in Mérida, a mid-altitude region located in the eastern<br />

branch <strong>of</strong> the Andes. There are two types <strong>of</strong> banana and plantain crop management<br />

in these areas (i) extensive banana plantations that are typical <strong>of</strong> the Colombian<br />

Atlantic coast and (ii) smallholdings that are characteristic <strong>of</strong> higher altitudes such<br />

as “Valle Central” in Costa Rica and the inner Andean valleys <strong>of</strong> Venezuela and<br />

Colombia (Price 1999).<br />

DNA isolation<br />

Infected plantains and banana leaves with advanced stages <strong>of</strong> lesion development<br />

were transferred to a humid chamber to allow ascospores to discharge onto 1.5%<br />

water agar. Single ascospores were identified microscopically, transferred to V8<br />

medium and incubated for 12 days at 25ºC. DNA was isolated from mycelium using<br />

a FastDNA Kit, Bio 101. DNA preparations were further purified by phenol:chlor<strong>of</strong>orm<br />

extraction (24:1, v/v) and ethanol precipitation. After washing with 70% ethanol,<br />

the final pellets were dissolved in an appropriate volume <strong>of</strong> 10 mM Tris-HCl, 1 mM<br />

EDTA, pH 8.<br />

SSR design for M. musicola<br />

The single ascospore culture MmCol-LM9.5.1 (collected from plantations severely<br />

affected by Sigatoka disease in a mid-altitude region <strong>of</strong> Colombia) was used as<br />

source material for the construction <strong>of</strong> a genomic library. Fungal DNA was isolated<br />

according to Weising et al. (1991) and purified by cesium chloride gradient<br />

centrifugation. Microsatellite enriched libraries were constructed and screened<br />

according to Fischer and Bachmann (1998). Detailed information about methodology,<br />

primer sequences and genebank accession numbers for the 26 polymorphic<br />

M. musicola primers can be found in Molina et al. (2001).<br />

Microsatellite analysis<br />

PCR amplifications were performed in a Perkin Elmer 9700 thermocycler in 10 µl<br />

reactions containing 5 ng <strong>of</strong> genomic DNA template, 0.5 µM <strong>of</strong> each forward and<br />

133


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

reverse primer, following PCR conditions according to Neu et al. (1999) and Molina<br />

et al. (2001). Products were separated on 6% sequencing gels and autoradiographed<br />

(Sambrook et al., 1989).<br />

For microsatellite allelic data, an initial matrix containing allele sizes was<br />

constructed, from which a 0/1 matrix was also derived. Dice similarity coefficient<br />

matrices and their corresponding dendrograms, grouped by the UPGMA agglomerative<br />

method, were calculated with the NTSYS s<strong>of</strong>tware package (Rohlf, 1993).<br />

Results<br />

SSR design for M. musícola<br />

Out <strong>of</strong> 1029 colonies screened, 205 yielded a positive hybridization signal (enrichment<br />

efficiency: 19.9%). Sixty-four clones were sequenced, and primers could be designed<br />

for 48 clones. Primer functionality was tested on a set <strong>of</strong><br />

24 template DNAs comprising 18 M. musicola isolates from a single field in Colombia,<br />

four isolates from Costa Rica and two from Mexico. The original clone served as a<br />

positive control. Primer transferability was tested with three M. fijiensis isolates:<br />

Mf-Col-LD9.1 (Colombia), Mf-Mex-015 (Mexico), Mf-PNG-294 (Papua New-Guinea).<br />

A total <strong>of</strong> 48 primer pairs were tested and 26 yielded single polymorphic bands <strong>of</strong><br />

the expected size. The characteristics are summarized in Table 1.<br />

Successful cross-priming with M. fijiensis DNA was observed at four loci. The<br />

availability <strong>of</strong> polymorphic microsatellite markers specific for M. musicola makes<br />

it possible to study the population structure <strong>of</strong> the pathogen in areas infested with<br />

Sigatoka disease, and to compare the pathogens using the same marker system.<br />

SSR typing<br />

SSR markers were used to type all isolates from both species, nine for M. fijiensis<br />

and eleven for M. musicola, showing high levels <strong>of</strong> polymorphism (Neu et al., 1999;<br />

Molina et al., 2001). Although the markers proved to be highly informative for the<br />

species for which they were developed, they were <strong>of</strong> limited value when transferred<br />

to other species. For example, M. fijiensis marker Mf-SSR-061 resulted in<br />

monomorphic patterns when used for typing M. musicola populations. The same was<br />

true for M. musicola marker Mm-SSR-024 in M. fijiensis isolates. An average <strong>of</strong> 3.4<br />

alleles were observed for M. fijiensis and 4.0 alleles for M. musicola. Polymorphic<br />

loci were 80.0% and 90.9% for M. fijiensis and M. musicola respectively.<br />

Cluster analysis based on Dice similarity index<br />

Dendograms based on Dice similarity indexes were produced for each species. The<br />

M. musicola tree was based on 42 SSR markers. Isolates grouped into six clusters<br />

(I to VI in Figure 1), which mostly paralleled the geographic locations. For example,<br />

in cluster I all isolates belonged to Cachipay and La Mesa, Colombia. These two<br />

locations are close to each other, have similar weather conditions and similar crop<br />

management practices. Cluster II mostly contained isolates from La Mesa, with some<br />

134


Session 2<br />

C. Molina et al.<br />

Table 1. Characteristics <strong>of</strong> microsatellites cloned from <strong>Mycosphaerella</strong> musicola.<br />

Locus Repeat <strong>of</strong> cloned allele Primer sequences (5’ – 3’)<br />

Mm SSR 01 (CA)9 G (CA)11 TAGTTGCAACCGAACAGG/ CTCCGTAGGTATGATGGTGT<br />

Mm SSR 03 (GA)4 (GC)2 (GA)32 CTCCGTAGGTATGATGGTGT/ GCTTCGTCAAGACCCTTAC<br />

Mm SSR 05 ((GT)4 (C))3 CCTCTTACGAAGTCTGTGGT/ TATCTCGGGAGACCAGACTA<br />

Mm SSR 06 (GA)8 CGAACAGGACGAAAGAATAG/ GTTTGTTCCAGTTCGTCAAG<br />

Mm SSR 07 (CA)50 ACGAGGTTTCAGAAGCAATA/ TCTTTCACCGAAGAAACCT<br />

Mm SSR 09 ((GT)6 AT (GT)3 GCn(GT)5 (GC) AGGGACGAACAAAACAGAG/ CCATGGTTTTCAAGCATATT<br />

Mm SSR 10 (CA)30 GAGAGCATGAAAAGTGGAAA/ CGTGACACTCGTCAGTTACA<br />

Mm SSR 14 (CA)7 CAA (CA)19 ATTTGGTGAATGGGGTAAG/ ACAGAGGGAAGCAAGTTTTT<br />

Mm SSR 15 (CA)27 CTACTGAGGCAGTCGCTAAC/ GGAGAGGTGGAAAAAGAAGT<br />

Mm SSR 16 (GA)6 AAA (GA)17 CCATCTGCCTTGAGATAGTC/ GAATTTATTCCAGCGAAGC<br />

Mm SSR 18 (GA)n ATCTGATTCGTATGGTGGAG/ TTGCTACTACTGGTGCTTCTC<br />

Mm SSR 21 (CTT)9 GTCGACCTCCATGACACTC/ TGCATGCAATCTGTAACCT<br />

Mm SSR 22 (GAA)9 CCAAAGCTTGAGTTGCTATT/ ACAACTTTTTGAGGAAAATGTAA<br />

Mm SSR 23 (CTT)27 CGACCTAGTCGAGGATGATA/ CGAAGACTTCTGAAAGGTCA<br />

Mm SSR 24 (GAA)2 GG (GAA)3GG (GAA)12 TCAAGAGGAGGAGAAGTTGA/ GGTTCTGATCAAGAGGAGGA<br />

Mm SSR 26 (CAA)8 ATATCTCTTCGTGTTTTGCG/ AAGTGTGGTCACAGCAAGTT<br />

Mm SSR 30 (CA)28 TGATGTTAAGTTGACGGACA/ CTAAGCCAAACCTCAATCAG<br />

Mm SSR 31 (AC)27 AACCACATCTTCGATCAGG/ CACATGGAATATCCTTGGTC<br />

Mm SSR 34 (CA)19 CTCGCTGCCTGATTATTCT/ AGATGGCATCGCTTCAC<br />

Mm SSR 35 (CA)4 AA (CA)26 TAACAATGTCCCTGAGAAGC/ GCCTTATCTGGAAAGTATCGT<br />

Mm SSR 36 (CA)13 ATTCCAGGTACGGCTACAC/ ATTCAGATCTGGTCTGGTTG<br />

Mm SSR 38 ((GT)n (CG))3 GAGAGTGAGATCGGTAGCAA/ CGGGATTAAGGTCTACCAA<br />

Mm SSR 39 (CA)19 TGCGAATTCCATTGATATG/ CGTGTGCTGACGAGAGAT<br />

Mm SSR 41 (GT)14 GGTGAGGTCGTTATTGTTGT/ GCTTTAGAGGTTTCGTTCTTC<br />

Mm SSR 44 (CA)9 (CT)14 CCTCACTCTCGCTCATACA/ AGAATGGACGAAAAACACTG<br />

Mm SSR 46 (CT)6 (GT)38 CGTGGACCTATTGTCAACTC/ TGGGTTACATTTACGAGAGAA<br />

isolates from Costa Rica and one from Venezuela. For clusters III and IV, all isolates<br />

were from Venezuela with each cluster having isolates from a single location. Cluster<br />

III represents isolates from Mérida and cluster IV represents isolates from Santa Rosa.<br />

Isolates from clusters V and VI were respectively from Costa Rica and Colombia.<br />

The M. fijiensis tree, based on 42 polymorphic SSR markers, shows four clusters<br />

(Figure 2). Cluster I comprises, almost exclusively, Santa Marta isolates, with a few<br />

isolates from Costa Rica and one from Caldas (Colombia). Clusters II and IV mostly<br />

comprise isolates from different locations in Costa Rica, with some isolates from<br />

135


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Colombia. Cluster III comprises 15 isolates from Caldas (Colombia), two isolates from<br />

Costa Rica and four from Santa Marta (Colombia). The 80 isolates used represent 51<br />

haplotypes. One particular haplotype was found in 10 isolates from Santa Marta<br />

(Colombia).<br />

Dice similarity index<br />

0.00 0.25 0.50 0.75 1.00<br />

M. musicola SSR<br />

Figure 1. Dendogram for <strong>Mycosphaerella</strong> musicola populations, based on Dice similarity indexes. Colombian<br />

populations are Mm-Col-CH: Clachipay and Mm-Col-LM: La Mesa;Venezuelan populations are Mm-Ven-Md:<br />

Merida and Mm-Ven-SR: Santa Rosa; Costa Rican populations are Mm-CR-DE: el Descanso, Mm-CR-ME: Ma.<br />

Eugenia and Mm-CR-QB: Quebrador.<br />

Discussion<br />

In the M. fijiensis and M. musicola dendograms, most clusters correspond to the<br />

original sampling sites and show a correlation between clusters and discrete populations.<br />

Studies <strong>of</strong> genetic diversity at a worldwide level (Carlier et al., 1996) and a<br />

136


Session 2<br />

C. Molina et al.<br />

Dice similarity index<br />

0.00 0.25 0.50 0.75 1.00<br />

M. fijiensis SSR<br />

Figure 2. Dendogram for <strong>Mycosphaerella</strong> fijiensis populations, based on Dice similarity indexes. Colombian<br />

populations are Mf-Col-SM: Santa Marta and Mf-Col-CD: Caldas; Costa Rican populations are Mf-CR-DE: el<br />

Descanso, Mf-CR-ME: Ma. Eugenia, Mf-CR-TR: Trsissia and Mf-CR-SR: San Rafael.<br />

regional level in Africa (Müller et al., 1997) suggest that in most situations<br />

colonization by M. fijiensis involves founder effects, a few individuals from an<br />

original population representing the haplotypic pool <strong>of</strong> a derived population.<br />

In the M. musicola tree, all but one <strong>of</strong> the Venezuelan isolates are grouped in two<br />

discrete clusters (III and IV). This is particularly true for the isolates from Mérida, which<br />

show very similar to identical DNA pr<strong>of</strong>iles. Only one Venezuelan isolate (Mm-Ven-<br />

Md-20) is in a separate group, with Colombian isolates from La Mesa (cluster II). M.<br />

137


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

musicola populations are mainly confined to mid-altitude regions <strong>of</strong> the Andes where<br />

<strong>bananas</strong> are grown in smallholdings. These smallholdings are separated by deep valleys<br />

and high mountains which are geographical barriers to gene flow. This may explain<br />

the separation <strong>of</strong> Colombian and Venezuelan populations in the M. musicola tree, despite<br />

being located in the same mountain range.<br />

In the M. fijiensis tree, it is important to highlight that the majority <strong>of</strong> isolates from<br />

the Atlantic coast (Santa Marta) are grouped in cluster I, whereas isolates from the mid-<br />

Andean region (Caldas) are in cluster III. Both regions are isolated from each other by<br />

physical distance and geographical barriers that prevent gene flow between populations.<br />

In addition, populations from the Atlantic coast are constantly under environmental<br />

pressure (high dosages <strong>of</strong> fungicides) whereas Andean populations are rarely treated<br />

with chemicals, keeping those areas as reservoirs <strong>of</strong> genetic diversity.<br />

The colonization dynamics <strong>of</strong> M. fijiensis for Central America has been documented<br />

(Unión de Países Exportadores de Banano, 1994). Colonization started from Honduras<br />

and continued southwards to Costa Rica and Colombia. Black <strong>leaf</strong> streak disease was<br />

first reported in 1981 in Urabá, Colombia, where it was confined for four years, until<br />

a first outbreak occurred along the shores <strong>of</strong> the Atrato River and then spread to the<br />

Atlantic coast and mid-Andean regions. Forty four isolates from Santa Marta (Atlantic<br />

Coast) corresponded to only 23 haplotypes, whereas 16 isolates from Caldas (mid-Andes)<br />

represented 10 haplotypes. This could be an indication <strong>of</strong> higher genetic diversity in<br />

the Caldas populations. In the Atlantic coast <strong>of</strong> Colombia, the genetic consequences <strong>of</strong><br />

the founder effect could have been enhanced by the pressure exerted by high doses <strong>of</strong><br />

fungicide and strict regulations on the transport <strong>of</strong> plant material between populations.<br />

M. fijiensis and M. musicola followed the same route, that is from Central America<br />

to Colombia. In cluster II <strong>of</strong> each phenogram (Figure 1 and 2), Costa Rican isolates are<br />

found in the same group with Colombian isolates. The genetic similarity <strong>of</strong> these two<br />

distant populations is consistent with the fact that Central American populations are<br />

ancestral and that the similarity cannot be explained by gene flow.<br />

Acknowledgements<br />

Plant Genetic Resources and Biotechnology (CORPOICA, Colombia) appreciates the<br />

support from COLCIENCIAS (223-95). C. Molina appreciates a fellowship from UNESCO,<br />

Paris (No. Sc-206.668.1) and Stiftung für Internationale Wissenschaftliche Zusammenarbeit<br />

(Frankfurt/Main, Germany). We would like to thank Luisa Pérez for commenting on the<br />

manuscript.<br />

References<br />

Bucheli E., B. Gautschi and J.A. Shyk<strong>of</strong>f. 2001. Differences in population structure <strong>of</strong> the anther<br />

smut fungus Microbotryum violaceum on two closely related host species, Silene latifolia<br />

and S. dioica. Molecular Ecology 10:285-294.<br />

Carlier J., X. Mourichon, D. Gonzalez-de-Leon and M.H. Lebrun. 1994. DNA restriction fragment<br />

length polymorphisms in <strong>Mycosphaerella</strong> species that cause banana <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>.<br />

Phytopathology 84:751–756.<br />

Carlier J., M.H. Lebrun, M.F. Zapater, C. Dubois and X. Mourichon. 1996. Genetic structure <strong>of</strong><br />

the global population <strong>of</strong> banana black <strong>leaf</strong> streak fungus, <strong>Mycosphaerella</strong> fijiensis.<br />

Molecular Ecology 5:499–410.<br />

138


Session 2<br />

C. Molina et al.<br />

Fischer D. and K. Bachmann. 1998. Microsatellite enrichment in organisms with large genomes<br />

(Allium cepa). Biotechniques 24:796–802.<br />

Groppe K. and T. Boller. 1997. PCR assay based on a microsatellite-containing locus for detection<br />

and quantification <strong>of</strong> Epichloe endophytes in grass tissue. Applied and Environmental<br />

Microbiology 63:1543-1550.<br />

Leach R. 1941. Banana <strong>leaf</strong> <strong>spot</strong>, <strong>Mycosphaerella</strong> musicola, the perfect stage <strong>of</strong> Cercospora musae<br />

Zimm. Tropical Agriculture 18:91-95<br />

Leach R. 1964. A new form <strong>of</strong> banana <strong>leaf</strong> <strong>spot</strong> in Fiji, black <strong>leaf</strong> streak. Wild Crops 16:60-64.<br />

Molina C., D. Kaemmer, S. Aponte, K. Weising and G. Kahl. 2001. Microsatellite markers for<br />

the fungal banana pathogen <strong>Mycosphaerella</strong> musicola. Molecular Ecology Notes 1:137-139.<br />

Mourichon X., J. Carlier and E. Fouré. 1997. Sigatoka Leaf Spot Diseases. INIBAP Musa disease<br />

fact sheet No.8. 4pp.<br />

Müller R., C. Pasberg-Gauhl, F. Gauhl, J. Ramser and G. Kahl. 1997. Oligonucleotide<br />

fingerprinting detects genetic variability at different levels in Nigerian <strong>Mycosphaerella</strong><br />

fijiensis. Journal <strong>of</strong> Phytopathology 145:25-30.<br />

Neu C., D. Kaemmer, G. Kahl, D. Fischer and K. Weising. 1999. Polymorphic microsatellite markers<br />

for the banana pathogen <strong>Mycosphaerella</strong> fijiensis. Molecular Ecology 8:523–525.<br />

Price N. 1999. Highland Bananas in Colombia. INFOMUSA 8(2):26-28.<br />

Rohlf J. 1993. Numerical Taxonomy and Multivariate Analysis System. Version 1.8.<br />

Rozen S. and H.J. Skaletsky. 1997. Primer3. Available at http://www-genome.wi.mit.edu/<br />

genome_s<strong>of</strong>tware/other/primer3.html.<br />

Sambrook J., E.F. Fritsch and T. Maniatis. 1989. Molecular Cloning: a Laboratory Manual.<br />

2 nd Edition. Cold Spring Harbor Laboratory Press, New York.<br />

Stover R.H. 1976. Distribution and cultural characteristics <strong>of</strong> the pathogens causing banana<br />

<strong>leaf</strong> <strong>spot</strong>. Tropical Agriculture 53:111-115.<br />

Unión de Países Exportadores de Banano (UPEB). 1994. The main lines <strong>of</strong> the banana industry<br />

in Latin America. INFOMUSA 5(1):14-19.<br />

Weising K.B. Beyermann, J. Ramser and G. Kahl. 1991. Plant DNA fingerprinting with<br />

radioactive and digoxigenated oligonucleotide probes complementary to simple repetitive<br />

DNA sequences. Electrophoresis 12:159–169.<br />

139


Session 2<br />

L. Conde-Ferráez et al.<br />

Poster<br />

An electrophoretic karyotype<br />

for <strong>Mycosphaerella</strong> fijiensis<br />

L. Conde-Ferráez,C.M.Rodríguez, L. Peraza-Echeverría and A. James<br />

Abstract<br />

In view <strong>of</strong> the current problems caused by black <strong>leaf</strong> streak disease in banana production, a<br />

knowledge and understanding <strong>of</strong> the genetics and organization <strong>of</strong> the genome <strong>of</strong> <strong>Mycosphaerella</strong><br />

fijiensis could lead to the development <strong>of</strong> new control strategies. Regarding the former, mycelium<br />

was obtained from isolates <strong>of</strong> M. fijiensis from three sites in Mexico (Veracruz, Colima and Chiapas)<br />

in order to estimate the size <strong>of</strong> the genome by using the CHEF (Contour clamped Homogeneous<br />

Electric Field) system. Different conditions <strong>of</strong> pulse field electrophoresis enabled the separation<br />

<strong>of</strong> M. fijiensis chromosomes and a preliminary estimate <strong>of</strong> the karyotype <strong>of</strong> each isolate was<br />

obtained. Isolates from Colima and Chiapas had bands corresponding to at least 10 chromosomes<br />

in the size range 0.71 to 2.2 Mb. The Veracruz isolate had at least 14 chromosomes in a size range<br />

<strong>of</strong> 0.67 to 5.6 Mb. Genome size calculated for the Veracruz isolate was at least 28 Mb, which is<br />

comparable to that <strong>of</strong> some ascomycete fungi. Attempts were made to estimate the genome<br />

size <strong>of</strong> the Colima and Veracruz isolates. Differences in the principal band suggested the<br />

presence <strong>of</strong> polymorphisms in chromosome length between the isolates studied, as reported<br />

for other species <strong>of</strong> fungi.<br />

Resumen - Cariotipo molecular de <strong>Mycosphaerella</strong> fijiensis<br />

Ante la problemática actual ocasionada por la Sigatoka negra en la producción de banano, el<br />

conocimiento y la comprensión de la genética y la organización del genoma de <strong>Mycosphaerella</strong><br />

fijiensis podrían conducir al desarrollo de nuevas estrategias para su control. Considerando lo<br />

anterior, se propuso obtener el cariotipo molecular de tres aislados de M. fijiensis por medio del<br />

sistema CHEF (Contour clamped Homogeneous Electric Field), así como estimar su tamaño<br />

genómico. Para ello, se utilizó el micelio de aislados procedentes de tres diferentes lugares de<br />

México (Veracruz, Colima y Chiapas). Se ensayaron diferentes condiciones de electr<strong>of</strong>oresis de<br />

campo pulsante que permitieron separar los cromosomas de M. fijiensis. Se obtuvo una estimación<br />

preliminar del cariotipo de cada aislado. En los aislados de Colima y Chiapas se observaron bandas<br />

correspondientes a por lo menos 10 cromosomas, en un rango de tamaño entre 0.71 y 2.2 Mb. En<br />

Centro de Investigación Científica de Yucatán, Mérida, México.<br />

141


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

el aislado de Veracruz se observaron por lo menos 14 cromosomas en un rango de tamaño entre<br />

0.67 y 5.6 Mb. El tamaño del genoma calculado para el aislado de Veracruz es de al menos 28 Mb<br />

lo cual es comparable con algunos hongos ascomicetos reportados en la literatura. Se pretende<br />

realizar experimentos para estimar el tamaño genómico de los aislados de Colima y Veracruz.<br />

Diferencias observadas en el patrón de bandeo sugieren la existencia de polimorfismos en la<br />

longitud de los cromosomas entre los aislados estudiados, lo cual también ha sido reportado en<br />

otras especies de hongos.<br />

Résumé - Un karyotype électrophorétique pour <strong>Mycosphaerella</strong> fijiensis<br />

Du fait des problèmes posés à l’heure actuelle par la maladie des raies noires pour la production<br />

de bananes, la connaissance et la compréhension de la génétique et de l’organisation du<br />

génome de M. fijiensis pourrait conduire au développement de nouvelles stratégies de contrôle.<br />

En ce qui concerne ce dernier point, du mycelium a été obtenu à partir d’isolats de M. fijiensis<br />

provenant de trois sites au Mexique (Veracruz, Colima et Chiapas) afin d’estimer la taille du<br />

génome en utilisant le système CHEF (Contour clamped Homogeneous Electric Field). Des<br />

conditions différentes d’électrophorèse en champ pulsé ont permis la séparation des chromosomes<br />

de M. fijiensis et une estimation préliminaire du karyotype de chaque isolat a été obtenue. Les<br />

isolats provenant du Colima et du Chiapas avaient des bandes correspondant à au moins 10<br />

chromosomes dont la taille variait entre 0,71 et 2,2 Mb. L’isolat de Veracruz avait au moins 14<br />

chromosomes dont la taille variait entre 0,67 et 5,6 Mb. La taille du génome calculée pour l’isolat<br />

provenant de Veracruz était d’au moins 28 Mb, ce qui est comparable à celle de certains<br />

champignons ascomycètes. Des essais ont été réalisés pour estimer la taille du génome des isolats<br />

provenant de Colima et de Veracruz. Les différences observées pour la bande principale suggèrent<br />

la présence de polymorphismes dans la longueur des chromosomes entre les isolats étudiés,<br />

comme cela a été rapporté chez d’autres espèces de champignons.<br />

Introduction<br />

The chromosomes <strong>of</strong> many fungi are too small to be identified by cytological<br />

methods therefore the detailed karyotype and genome size <strong>of</strong> most fungal species<br />

are unknown. But starting with the successful separation <strong>of</strong> Saccaromyces<br />

cerevisiae chromosomes (Schwartz and Cantor, 1984), Pulsed Field Gel<br />

Electrophoresis (PFGE) has been used to obtain the molecular karyotypes <strong>of</strong><br />

important fungi, including pathogenic fungi. New technologies and electrophoresis<br />

apparatus have since been developed, resulting in improved techniques for<br />

separating large DNA fragments and chromosomes.<br />

The CHEF (contour-clamped homogeneous electric field), electrophoresis system<br />

(Chu et al., 1986), has been used to separate fungal chromosomes and estimate<br />

genome size. Karyotyping procedures using PFGE generally involve the production<br />

<strong>of</strong> protoplasts or sphaeroplasts. However a different technique (McCluskey et al., 1990)<br />

permits the preparation <strong>of</strong> chromosome-sized DNA without the need for such<br />

laborious procedures.<br />

Even though it affects Musa worldwide, there are no data on the karyotype or<br />

genome size <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis. Knowing and understanding the genetics<br />

and genome organization <strong>of</strong> fungal pathogens could lead to the development <strong>of</strong> new<br />

strategies in controlling black <strong>leaf</strong> streak disease. This paper describes the use <strong>of</strong><br />

PFGE to obtain the molecular karyotype <strong>of</strong> M. fijiensis and estimates <strong>of</strong> genome<br />

size.<br />

142


Session 2<br />

L. Conde-Ferráez et al.<br />

Methods<br />

Leaves displaying symptoms <strong>of</strong> black <strong>leaf</strong> streak disease were collected from<br />

Veracruz, Colima and Chiapas in Mexico. Fungal isolates were obtained from single<br />

ascospores and grown on PDA medium. The identity <strong>of</strong> isolates was verified by the<br />

morphology <strong>of</strong> mycelium, Koch´s postulates and PCR (Johanson, 1997; Neu et al.,<br />

1999). For the preparation <strong>of</strong> agarose plugs, isolates were grown in liquid shake<br />

culture (PDB medium) for 6 days at 26 o C and 100 rpm. The mycelium was separated<br />

by centrifugation, ground, washed with buffer, mixed with melted agarose at 45°C<br />

and transferred to plug moulds. Plugs were incubated in proteinase K at 60°C for<br />

48 h, and washed and stored in 0.5 M EDTA at 4°C. Empirical methods were tried<br />

to optimize the resolution <strong>of</strong> chromosomes <strong>of</strong> different size ranges, using PFGE with<br />

CHEF DR II and CHEF DR III. Finally, genome size was estimated by adding the values<br />

assigned to each band resolved in pulsed field gels.<br />

Results and discussion<br />

We obtained a preliminary estimation <strong>of</strong> the karyotype <strong>of</strong> each isolate. For the<br />

Veracruz isolate, 12 bands were resolved, representing 14 chromosomes in a size<br />

range <strong>of</strong> 0.67 to 5.6 Mb (Figures 1 and 2). At least 10 chromosomes were resolved<br />

in the Colima and Chiapas isolates, in a size range <strong>of</strong> 0.71 to 2.2 Mb. The estimated<br />

genome size for the Veracruz isolate is at least 28 Mb (Figure 3), which is<br />

comparable to that reported for other ascomycetes (Cooley and Caten, 1991;<br />

McDonald and Martinez, 1991). The genome sizes <strong>of</strong> the other two isolates are to<br />

be determined in further experiments. Differences observed in banding patterns<br />

suggest the existence <strong>of</strong> chromosome length polymorphisms, which has been<br />

reported for other fungi. A broader study with different and well characterized<br />

haplotypes <strong>of</strong> M. fijiensis is now underway at the Centro de Investigación Científica<br />

de Yucatán.<br />

Conclusion<br />

This is the first estimate <strong>of</strong> the karyotype and genome size <strong>of</strong> M. fijiensis, a useful<br />

tool for constructing a physical and genetic map or for calculating the required size<br />

<strong>of</strong> a genomic library<br />

References<br />

Chu G., D. Vollrath and R.W. Davis. 1986. Separation <strong>of</strong> large DNA molecules by contourclamped<br />

homogeneous electric fields. Science, 234:1582-1585.<br />

Cooley R.N. and C. Caten. 1991. Variation in electrophoretic karyotype between strains <strong>of</strong><br />

Septoria nodorum. Molecular and General Genetics 228:17-23.<br />

Johanson A. 1997. Detection <strong>of</strong> Sigatoka <strong>leaf</strong> <strong>spot</strong> <strong>of</strong> banana by the Polymerase Chain<br />

Reaction. Natural Resources Institute. The University <strong>of</strong> Greenwich, U.K. 37pp.<br />

McCluskey K., B.W. Russell and D. Mills. 1990. Electrophoretic karyotyping without the need<br />

for generating protoplasts. Current Genetics 18:385-386.<br />

143


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

McDonald B.A. and J.P. Martinez. 1991. Chromosome length polymorphisms in Septoria tritici<br />

population. Current Genetics 19:265-271.<br />

Neu C., D. Kaemmer, G. Kahl, D. Fischer and K. Weising. 1999. Polymorphic microsatellite<br />

markers for the banana pathogen <strong>Mycosphaerella</strong> fijiensis. Molecular Ecology 8:523-525.<br />

Schwartz, D.C. and C.R. Cantor. 1984. Separation <strong>of</strong> yeast chromosome-sized DNAs by Pulsed<br />

Field Gradient Gel Electrophoresis. Cell 37:67-75.<br />

1 2 3 4 5<br />

2200<br />

+<br />

1600<br />

1900<br />

1500<br />

1300<br />

1125<br />

1200<br />

1070<br />

1050<br />

1020<br />

945<br />

825<br />

785<br />

750<br />

680<br />

610<br />

450<br />

365<br />

285<br />

225<br />

830<br />

740<br />

670<br />

Figure 1. CHEF gel stained with SYBR Green showing separation <strong>of</strong> medium size chromosomes. Lane 1:<br />

S. cerevisiae chromosome size standards. Lanes 2-5: M. fijiensis isolate from Veracruz. Numbers on the left give<br />

the siez <strong>of</strong> the standards. Numbers on the right correspond to the nine bands resolved. Numbers in bold<br />

represent comigrating bands.<br />

144


Session 2<br />

L. Conde-Ferráez et al.<br />

1 2 3 4 5<br />

5.7<br />

4.6<br />

5.7<br />

4.6<br />

4.0<br />

4.6<br />

Figure 2. CHEF gel stained with SYBR Green showing separation <strong>of</strong> largest chromosomes. Lane 1: S. pombe<br />

chromosome size standards. Lanes 2-5: M. fijiensis isolate form Veracruz. Numbers on the left give the size<br />

<strong>of</strong> the standards. Numbers on the right correspond to the three bands resolved. Numbers in bold represent<br />

comigrating bands.<br />

145


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

1 2 3 4 5<br />

2200<br />

1400<br />

2200<br />

+<br />

1600<br />

1055<br />

1125<br />

945<br />

805<br />

785<br />

750<br />

715<br />

1020<br />

945<br />

825<br />

785<br />

750<br />

680<br />

610<br />

450<br />

Figure 3. CHEF gel stained with SYBR Green showing separation <strong>of</strong> largest chromosomes. Lanes 1-4: M. fijiensis<br />

isolates from Veracruz, Colima, Chiapas #1 and Chiapas #2 respectively. Lane 5: S. cerevisiae chromosome size<br />

standards. Numbers on the right give the size <strong>of</strong> the standards. Numbers on the left correspond to the eight<br />

bands resolved. Numbers in bold represent comigrating bands.<br />

146


Session 1<br />

Recommendations<br />

Recommendations <strong>of</strong> session 2<br />

Information on the epidemiology and population structure <strong>of</strong> the main <strong>Mycosphaerella</strong> species<br />

(M. fijiensis, M. musicola and M. eumusae) at the national, regional and international levels<br />

are needed to better understand the distribution and the spread <strong>of</strong> the pathogens, to anticipate<br />

the evolution <strong>of</strong> pathogen populations and to define resistance management strategies. Such<br />

studies are particularly necessary in Asia, the centre <strong>of</strong> diversity <strong>of</strong> the three pathogens and<br />

where little research has so far been conducted.<br />

Distribution <strong>of</strong> <strong>Mycosphaerella</strong> spp.<br />

Knowing the name <strong>of</strong> the banana clones affected, the severity <strong>of</strong> the <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

and the local environmental conditions would help explain the distribution <strong>of</strong> M. fijiensis,<br />

M. musicola and M. eumusae. IMTP trials are seen as ideal locations for assessing the reaction<br />

<strong>of</strong> the banana clones to various <strong>leaf</strong> <strong>spot</strong> pathogens.<br />

The collection and diagnosis <strong>of</strong> specimens from IMTP trials sites should be continued.<br />

Identification tools should be provided to enable diagnoses to be undertaken locally.<br />

The exact distribution <strong>of</strong> M. eumusae needs to be known.<br />

Further surveys in South and Southeast Asia are necessary to determine where<br />

M. musicola, M. fijiensis and M. eumusae occur. The cooperation and collaboration <strong>of</strong><br />

scientists in South and Southeast Asia is viewed as essential. The INIBAP regional <strong>of</strong>fice<br />

for Asia and the Pacific should strengthen and facilitate any exchange between Asian<br />

partners and the rest <strong>of</strong> the PROMUSA community.<br />

National and international collections<br />

The creation <strong>of</strong> national collections <strong>of</strong> strains <strong>of</strong> <strong>Mycosphaerella</strong> pathogens is <strong>of</strong> special<br />

relevance to the understanding <strong>of</strong> population structure. The collections must be based on<br />

single-spore cultures with an in vitro characterization <strong>of</strong> the anamorph stage (in vitro<br />

sporulation <strong>of</strong> conidia). Diagnostic tools would help the development <strong>of</strong> collections <strong>of</strong><br />

<strong>Mycosphaerella</strong> isolates. It has been recommended to provide the participants with a<br />

protocol to sample, establish and maintain the collections (see the “Diagnostics” section in<br />

session 1 recommendations).<br />

A reliable, rapid test to distinguish M. musicola, M. fijiensis, M. eumusae and possible<br />

other <strong>Mycosphaerella</strong> pathogens/saprophytes needs to be developed. Information on how<br />

to distinguish the three pathogens on morphological characteristics also needs to be<br />

produced and circulated to banana scientists. INIBAP was asked to address this need.<br />

The establishment <strong>of</strong> a national collection should be promoted and facilitated through<br />

the organization <strong>of</strong> a training course; especially for those countries that develop breeding<br />

programmes, but also in places where banana resistant hybrids are used on an industrial<br />

scale, and where the high diversity <strong>of</strong> Musaceas has likely produced a similar diversity<br />

<strong>of</strong> pathogens.<br />

147


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

It was recommended to develop an international core collection <strong>of</strong> M. fijiensis,<br />

M. musicola and M. eumusae. The different strains should be conserved as fungal mycelia<br />

and DNA. CIRAD was suggested as host <strong>of</strong> the international collection, using a similar<br />

mechanism as INIBAP developed with KULeuven regarding Musa germplasm. INIBAP<br />

was asked to address this needs in collaboration with CIRAD.<br />

Genetic population structure<br />

The study <strong>of</strong> the genetic population structures <strong>of</strong> <strong>Mycosphaerella</strong> pathogens is already ongoing<br />

at the national, regional and international levels. However, the number <strong>of</strong> countries<br />

involved at the national level should be increased to refine regional and international studies.<br />

The sampling protocols should be standardized and widely distributed. INIBAP and CIRAD<br />

agreed to work together in the preparation <strong>of</strong> this information which should include<br />

several detailed illustrations <strong>of</strong> the different pathogens and their anamorph stages. This<br />

information should also be part <strong>of</strong> the IMTP guidelines.<br />

More molecular markers, such as SSR and CAPS, should be developed to improve the<br />

understanding <strong>of</strong> the different populations structures.<br />

Pathogenicity characterization<br />

In vitro and in vivo inoculation systems exist to evaluate the pathogenicity <strong>of</strong> the various<br />

<strong>Mycosphaerella</strong> strains. The different pathogens and their relation to their host need to be<br />

compared under controlled conditions using these methods.<br />

The methodologies that currently exist should be standardized. The in vitro inoculation<br />

on <strong>leaf</strong> fragment developed at CIRAD should be distributed together with the methodology<br />

to isolate, cultivate and produce the inoculum <strong>of</strong> the different pathogens. INIBAP and<br />

CIRAD have been requested to compile in a technical document, all the different<br />

information already published on these methods.<br />

Dispersal <strong>of</strong> <strong>Mycosphaerella</strong> spp.<br />

More research is necessary to understand spore dispersal.<br />

Disease incidence data should be collected from the field and the scientific literature.<br />

Laboratory methods to understand the mechanism <strong>of</strong> spore release, and spore survival<br />

in the atmosphere should be developed.<br />

The potential for windborne dispersal suspected from laboratory studies should be verified<br />

and assessed at the plantation level (as opposed to dispersal through other means, such<br />

as infected planting material).<br />

148


Session 3<br />

Host-pathogen interactions


Session 3<br />

P. Lepoivre et al.<br />

Introduction<br />

Banana–<strong>Mycosphaerella</strong> fijiensis<br />

interactions<br />

P. Lepoivre 1 ,J.P.Busogoro 1 ,J.J.Etame 1 ,A.El Hadrami 1,2 ,<br />

J. Carlier 2 ,G.Harelimana 1 ,X.Mourichon 2 ,B.Panis 3 ,<br />

A. Stella Riveros 4 ,G.Sallé 5 ,H.Strosse 3 and R. Swennen 3<br />

Abstract<br />

Using standard testing procedures, banana genotypes were classified as 1) highly resistant<br />

cultivars characterized by an early blockage <strong>of</strong> <strong>leaf</strong> infection (incompatible interactions), 2) partially<br />

resistant cultivars exhibiting a slow rate <strong>of</strong> symptom development (compatible reactions)<br />

and 3) susceptible cultivars, characterized by rapid development <strong>of</strong> necrotic lesions (compatible<br />

reaction).<br />

Most information on incompatible reactions comes from observations <strong>of</strong> early necrosis <strong>of</strong><br />

stomatal guard cells and the deposit <strong>of</strong> electron-dense compounds around the penetration sites<br />

<strong>of</strong> M. fijiensis on the cultivar ‘Yangambi km5’. Such rapid death <strong>of</strong> a few host cells, associated with<br />

the blockage <strong>of</strong> the progression <strong>of</strong> the infecting agent is usually defined as a hypersensitive<br />

reaction. Such a reaction <strong>of</strong>ten operates within a gene-for-gene relationship and as a consequence<br />

the resulting resistance may be unstable.<br />

As regards compatible interactions, cytological studies showed that M. fijiensis behaves first as<br />

a biotrophic parasite which colonizes exclusively the intercellular spaces without the formation<br />

<strong>of</strong> haustoria. Two main mechanisms have been investigated to explain the slow development<br />

<strong>of</strong> a lesion in partially resistant genotypes: preformed antifungal compounds and tolerance to<br />

putative toxin(s) produced by M. fijiensis.<br />

The mechanisms will be presented in relation to their possible use as early screening markers<br />

for selecting banana genotypes for durable resistance to M. fijiensis.<br />

1<br />

Faculté Universitaire des Sciences Agronomiques, Gembloux, Belgium<br />

2<br />

CIRAD, Montpellier, France<br />

3<br />

Katholieke Universiteit Leuven, Leuven, Belgium<br />

4<br />

CATIE, Costa Rica<br />

5<br />

Université Pierre et Marie Curie, Paris, France<br />

151


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Resumen - Interacciones banano – <strong>Mycosphaerella</strong> fijiensis<br />

Utilizando los procedimientos de evaluación estándar, los genotipos de banano fueron clasificados<br />

en tres categorías: 1) cultivares altamente resistentes caracterizados por un bloqueo temprano de<br />

la infección foliar (interacciones incompatibles), 2) cultivares parcialmente resistentes que exhiben<br />

una evolución lenta de los síntomas (reacciones compatibles), y 3) cultivares susceptibles,<br />

caracterizados por un desarrollo rápido de las lesiones necróticas (reacción compatible).<br />

La mayor parte de la información sobre las reacciones incompatibles proviene de los estudios del<br />

cultivar ‘Yangambi km5’. Se observaron la necrosis de las células de guarda estomatales y los depósitos<br />

de los compuestos con alta densidad de electrones alrededor de los sitios de penetración de M. fijiensis.<br />

La muerte tan rápida de unas cuantas células hospedantes asociada con el bloqueo de la progresión<br />

del agente infectante se define usualmente como una reacción hipersensible. Esta reacción a menudo<br />

opera dentro de una relación de gen por gen y podría convertir la resistencia en inestable.<br />

Con las reacciones compatibles, los estudios citológicos revelaron que M. fijiensis se comporta primero<br />

como un parásito biotrófico que coloniza exclusivamente los espacios intercelulares sin formar los<br />

haustorios. Dos mecanismos principales podrían estar involucrados en el desarrollo lento de las<br />

lesiones observado en los genotipos resistentes parcialmente:compuestos antifungosos sintetizados<br />

de manera constitutiva o tolerancia a la(s) toxina(s) putativa(s) producidas por M. fijiensis.<br />

Estos mecanismos se presentarán en relación con su posible utilidad como marcadores de cribado<br />

temprano en la selección de los genotipos de banano con respecto a la resistencia duradera a<br />

M. fijiensis.<br />

Résumé - Interactions bananier–<strong>Mycosphaerella</strong> fijiensis<br />

En utilisant des procédures de test standard, des génotypes de bananier ont été classés en :<br />

1) cultivars hautement résistants caractérisés par un blocage rapide de l’infection foliaire<br />

(interactions incompatibles), 2) cultivars partiellement résistants montrant un développement<br />

lent des symptômes (réactions compatibles) et 3) cultivars susceptibles caractérisés par un<br />

développement rapide de lésions nécrotiques (réaction compatible).<br />

L’essentiel des informations sur les réactions incompatibles provient d’observations de nécrose<br />

précoce des cellules de garde des stomates et du dépôt de composés denses en électrons autour<br />

des sites de pénétration de M. fijiensis chez le cultivar ‘Yangambi km5’. La mort aussi rapide d’un<br />

petit nombre de cellules hôtes, associée avec le blocage de la progression de l’agent infectieux,<br />

est habituellement définie comme une réaction hypersensible. Une telle réaction se produit souvent<br />

dans le cadre d’une relation gène pour gène et, en conséquence, la résistance qui en résulte peut<br />

être instable.<br />

Pour ce qui concerne les interactions compatibles, les études cytologiques ont montré que M. fijiensis<br />

se comporte d’abord comme un parasite biotrophique qui colonise exclusivement les espaces<br />

intercellulaires sans formation d’haustoria. Deux mécanismes principaux ont été étudiés pour<br />

expliquer le développement lent des lésions chez les génotypes partiellement résistants : des<br />

composés antifongiques préformés et la tolérance à une(des) toxine(s) putative(s) produite(s) par<br />

M. fijiensis.<br />

Les mécanismes sont présentés en relation avec leur utilisation possible comme marqueurs lors<br />

de criblage précoce pour sélectionner des génotypes de bananiers possédant une résistance durable<br />

à M. fijiensis.<br />

Introduction<br />

Black <strong>leaf</strong> streak disease is the most devastating disease <strong>of</strong> banana and plantain<br />

worldwide. The fungus induces foliar <strong>leaf</strong> streaks which, in highly susceptible<br />

cultivars, leads to the total collapse <strong>of</strong> the plant.<br />

Just as host plants evolved several defence mechanisms, pathogens have ways<br />

to evade or suppress these defence mechanisms. The response <strong>of</strong> the host and the<br />

152


Session 3<br />

P. Lepoivre et al.<br />

pathogen are crucial to the outcome <strong>of</strong> infection. A knowledge <strong>of</strong> the interactions<br />

is increasingly important for the rational selection <strong>of</strong> genotypes resistant to plant<br />

pathogens. The interactions between banana and <strong>Mycosphaerella</strong> fijiensis remained<br />

unknown for a long time.<br />

Although the field performance <strong>of</strong> an accession is the ultimate reference for<br />

evaluating its resistance, the method is not suitable for the study <strong>of</strong> host pathogen<br />

interactions. Nevertheless, the reaction to black <strong>leaf</strong> streak disease <strong>of</strong> about 50 Musa<br />

species belonging to various genetic groups was studied under natural infection<br />

conditions (Fouré et al., 1990). The study led to the grouping <strong>of</strong> the banana genotypes<br />

in three categories: 1) highly resistant (HR) cultivars characterized by an early<br />

blockage <strong>of</strong> <strong>leaf</strong> infection (incompatible interactions); 2) partially resistant (PR)<br />

cultivars exhibiting slow rates <strong>of</strong> symptom development (compatible interactions);<br />

and 3) susceptible (S) cultivars characterized by a rapid development <strong>of</strong> necrotic<br />

lesions (compatible interactions). Later, banana-M. fijiensis-interactions were<br />

studied under controlled conditions <strong>of</strong> inoculation (Mourichon et al., 1987) which<br />

reproduced the behaviour in the field <strong>of</strong> three reference cultivars: ‘Yangambi km5’<br />

(AAA; HR), ‘Fougamou’ (ABB; PR) and ‘Grande naine’ (AAA; S).<br />

These preliminary results, which were presented at the International workshop<br />

held in San José in 1989, were the start <strong>of</strong> investigations into banana-M. fijiensis<br />

interactions. They began with the microscopic events that take place in banana<br />

tissues and were followed by the analysis <strong>of</strong> the biochemical processes that culminate<br />

in the expression <strong>of</strong> resistance or susceptibility.<br />

Host-pathogen interactions<br />

Cytological studies <strong>of</strong> the interactions between M. fijiensis and the three reference<br />

cultivars ‘Yangambi km5’, ‘Fougamou’ and ‘Grande naine’ revealed that M. fijiensis<br />

enters banana leaves by the stomata.<br />

In compatible interactions (‘Grande naine’ and ‘Fougamou’ inoculated with<br />

M. fijiensis, strain 049 HND from Honduras), the pathogen colonized exclusively<br />

the intercellular spaces between mesophyll cells, without forming haustoria. There<br />

was a long period <strong>of</strong> biotrophy before the observation <strong>of</strong> the first cytological<br />

alterations to the mesophyll cells. Hyphae were observed between living cells ahead<br />

<strong>of</strong> the necrotic zone, a faster growth rate <strong>of</strong> hyphae being the main difference between<br />

susceptible (‘Grande naine’) and partially resistant (‘Fougamou’) cultivars (Beveraggi,<br />

1992; Beveraggi et al., 1995).<br />

In contrast, early necrosis <strong>of</strong> stomata guard cells and appositions around the<br />

penetration sites were observed with incompatible interactions (‘Yangambi km5’<br />

inoculated with M. fijiensis strain 049HND) (Beveraggi et al., 1995).<br />

The behaviour <strong>of</strong> partially and highly resistant genotypes <strong>of</strong> banana can be linked<br />

to major groups <strong>of</strong> plant-parasite interactions. The rapid death <strong>of</strong> only a few host<br />

cells, associated with the blockage <strong>of</strong> the progression <strong>of</strong> the infecting agent in the<br />

highly resistant cultivar ‘Yangambi km5’, is usually defined as an hypersensitive<br />

reaction. These <strong>of</strong>ten operate within a gene-for-gene relationship giving rise to<br />

resistance that is unstable. In comparison, the partial resistance <strong>of</strong> the reference<br />

cultivar ‘Fougamou’, for example, is usually considered polygenic and durable.<br />

153


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Incompatible interactions: Highly resistant cultivars<br />

The hypersensitive reaction operating within a gene-for-gene relationship is<br />

generally explained, either by the presence <strong>of</strong> a specific avirulence factor (or elicitor)<br />

or by the coordinated action <strong>of</strong> non-specific elicitor(s) and a specific suppressor (de<br />

Wit, 1992; Atkinson, 1993). In banana–M. fijiensis interactions, there is no<br />

experimental evidence <strong>of</strong> a gene-for-gene relationship because it is difficult to study<br />

the genetics <strong>of</strong> triploid genotypes such as ‘Yangambi km5’. But that such a<br />

relationship exists is supported by laboratory tests in which isolates <strong>of</strong> M. fijiensis<br />

overcame the resistance <strong>of</strong> ‘Yangambi km5’ (Fullerton and Olsen, 1995).<br />

Riveros and Lepoivre (1994) did preliminary experiments to identify the elicitors<br />

that induce resistance. Intercellular fluids (IF) from leaves <strong>of</strong> ‘Yangambi km5’<br />

(incompatible) and ‘Grande naine’ (compatible) inoculated with 049HND, and crude<br />

eliciting fractions (CEF) prepared from germinating spores <strong>of</strong> virulent and<br />

avirulent M. fijiensis isolates, elicited necrosis and appositions in banana<br />

cultivars (Riveros and Lepoivre, 1994).<br />

Regardless <strong>of</strong> the eliciting preparation (IFs or CEFs prepared with avirulent<br />

or virulent M. fijiensis isolates), the reaction was more intense and quicker<br />

in ‘Yangambi km5’ than in the susceptible ‘Grande naine’. The behaviour<br />

<strong>of</strong> ‘Yangambi km5’ cannot be explained by a race-specific eliciting activity<br />

in the IFs or the CEFs. However, the eliciting activity present in the IFs <strong>of</strong><br />

‘Yangambi km5’ inoculated with the avirulent strain 049 HND appeared to be<br />

higher than that in the compatible relationship between ‘Grande naine’ and the<br />

same isolate. Thus, we speculate that ‘Yangambi km5’ could have a higher<br />

sensitivity to the elicitor(s) but could also have a host-mediated effect on the<br />

release, production or stability <strong>of</strong> specific elicitor(s) produced by the fungal isolates.<br />

Such host-mediated effects have been reported in soybean tissues where plant<br />

enzymes are responsible for the release <strong>of</strong> elicitors from hyphal walls <strong>of</strong><br />

Phytophthora megasperma (Boller, 1987).<br />

A wide range <strong>of</strong> fungal compounds have been implicated as elicitors <strong>of</strong> HR:<br />

polysaccharides (Sharp et al., 1984), glycoproteins (Schaffrath et al., 1995),<br />

peptides (de Wit et al., 1985) and hydrolytic enzymes (Boller, 1987). In banana–<br />

M. fijiensis interactions, polysaccharide compounds may be involved in eliciting<br />

activity (Riveros, unpublished data).<br />

Evidence <strong>of</strong> a hypersensitive-like-reaction to M. fijiensis represents a first step<br />

towards a better characterization <strong>of</strong> that reaction. Independently <strong>of</strong> the<br />

mechanisms <strong>of</strong> resistance, there is the problem <strong>of</strong> the durability <strong>of</strong> resistance.<br />

Durability is <strong>of</strong> outmost importance because breakdown <strong>of</strong> resistance for a staple<br />

crop such as plantain would have dramatic effects. Because <strong>of</strong> the difficulties<br />

inherent in improving triploid <strong>bananas</strong>, breeding for resistance to black <strong>leaf</strong> streak<br />

disease <strong>of</strong>ten did not take into account host-pathogen interactions. The existence<br />

<strong>of</strong> M. fijiensis isolates able to overcome resistant ‘Yangambi km5’ in laboratory<br />

tests (Fullerton and Olsen, 1995) shows that highly resistant parents should not<br />

be used without appropriate management procedures, such as mixtures <strong>of</strong><br />

cultivars, choice <strong>of</strong> the “right gene combination” and co-ordinated regional<br />

deployment <strong>of</strong> genes.<br />

154


Session 3<br />

P. Lepoivre et al.<br />

Compatible reaction: partial resistant cultivars<br />

With partially resistant cultivars (compatible interactions), cytological studies showed<br />

that M. fijiensis behaves at first as a biotrophic parasite that colonizes exclusively the<br />

intercellular spaces (Beveraggi et al. 1995). Two possible mechanisms have been<br />

investigated to explain slow lesion development in partially resistant genotypes:<br />

preformed antifungal compounds and tolerance to putative toxin(s) produced by<br />

M. fijiensis.<br />

Constitutively synthesised antifungal compounds<br />

Many antimicrobial compounds produced by plants play an important role in the<br />

response to infection by cellular pathogens. Defence compounds may be classified<br />

into phytoanticipins, which are constitutive, and phytoalexins, which are synthesised<br />

in response to microorganisms. The two groups <strong>of</strong> secondary metabolites include<br />

a wide range <strong>of</strong> chemical families. However, phytoanticipins are primarily involved<br />

in non-host rather than varietal resistance.<br />

For ‘Fougamou’, histological analysis revealed the presence, in mesophyll<br />

layers, <strong>of</strong> many specialized cells containing vacuoles rich in polyphenol. The contents<br />

<strong>of</strong> the vacuoles were released into the intercellular spaces. The contents had a high<br />

affinity for fungal cell walls and their presence around hyphae seemed to be<br />

correlated with the slow growth <strong>of</strong> mycelium in parenchyma tissues (Beverragi et<br />

al., 1992, 1995). Gire (1994) identified soluble phenols in the <strong>leaf</strong> tissues <strong>of</strong> several<br />

banana cultivars with different levels <strong>of</strong> partial resistance. He also observed a close<br />

correlation between flavane (protoanthocyanidins) content and the level <strong>of</strong> partial<br />

resistance. However, a study conducted on a larger number <strong>of</strong> genotypes suggested<br />

that the role <strong>of</strong> these constitutive compounds in partial resistance is restricted to<br />

a limited number <strong>of</strong> cultivars (El Hadrami, 1997).<br />

The role <strong>of</strong> toxins in pathogenesis<br />

Pathogen toxins could constitute an alternative technique for rapidly screening<br />

resistant banana genotypes as in vitro plant tissues or young plants. The symptoms<br />

<strong>of</strong> black <strong>leaf</strong> streak disease suggest a possible involvement <strong>of</strong> phytotoxic compounds.<br />

Such compounds were found in culture filtrates <strong>of</strong> M. fijiensis (Molina and Krausz,<br />

1989; Lepoivre and Acuna, 1989; Upadhyay et al., 1990; Strobel et al., 1993). Stierle<br />

et al. (1991) reported that 2,4,8-tetrahydroxytetralone and juglone were the most<br />

abundant and most phytotoxic compounds.<br />

If toxins are involved in the development <strong>of</strong> black <strong>leaf</strong> streak disease, it may be<br />

possible to use them to identify resistant genotypes. During the previous workshop<br />

at San José, the use <strong>of</strong> M. fijiensis toxins for screening had four major limitations:<br />

1) a lack <strong>of</strong> quantitative and sensitive bioassays to measure the effects <strong>of</strong> M. fijiensis<br />

metabolites on banana genotypes; 2) insufficient characterization <strong>of</strong> the variability<br />

in toxin production <strong>of</strong> M. fijiensis; 3) a lack <strong>of</strong> experimental evidence for the role<br />

<strong>of</strong> the metabolites in the disease; and 4) the assurance that the susceptibility and/or<br />

resistance <strong>of</strong> cultured tissues reflected the reaction <strong>of</strong> the whole plant.<br />

155


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Bioassay to quantify the effect <strong>of</strong> M. fijiensis<br />

metabolites<br />

A set <strong>of</strong> bioassays was developed to quantify the toxic effects <strong>of</strong> the metabolites<br />

obtained from M. fijiensis culture filtrates. The induction <strong>of</strong> necrosis by a <strong>leaf</strong><br />

puncture bioassay on detached banana leaves, or the injection <strong>of</strong> ethyl acetate crude<br />

extract (EaCE) into the leaves are easy to perform but neither method is sensitive<br />

(injections <strong>of</strong> 250 ppm EaCE are required for ‘Grande naine’) or quantitative.<br />

The electrolyte leakage assay represented a quantitative but rather insensitive<br />

assessment <strong>of</strong> the toxicity <strong>of</strong> M. fijiensis metabolites. The test did not distinguish<br />

between cultivars. The most sensitive and accurate toxin assay was based on the<br />

measurement <strong>of</strong> chlorophyll fluorescence. The vitality index seemed to be the most<br />

sensitive method for early assessment <strong>of</strong> the effects <strong>of</strong> EaCE and a specific indicator<br />

<strong>of</strong> photosynthetic activity.<br />

Purification <strong>of</strong> the EaCE revealed the presence <strong>of</strong> different fractions with similar<br />

properties to the crude extracts. Juglone, a purified metabolite previously shown<br />

to be present in extracts <strong>of</strong> M. fijiensis culture filtrates, was identified in the extracts<br />

<strong>of</strong> all the strains analyzed. Injection <strong>of</strong> juglone into banana <strong>leaf</strong> tissues gave similar<br />

results to EaCE for ranking cultivars (Etame, unpublished data).<br />

Chloroplasts as target site <strong>of</strong> juglone<br />

The involvement <strong>of</strong> the photosynthetic apparatus in reaction to EaCE and juglone<br />

is in agreement with observation <strong>of</strong> light-dependent toxicity irrespective <strong>of</strong> the<br />

bioassay. The observation <strong>of</strong> swelling chloroplasts as the first abnormality observed<br />

by electron microscopy <strong>of</strong> EaCE-treated leaves also fits this pattern.<br />

Busogoro (unpublished data) developed a bioassay using isolated chloroplasts and<br />

measuring their capacity to reduce 2,6- dichlorophenoindolphenol (DCPIP) as a marker<br />

<strong>of</strong> the Hill reaction, which expresses electron transport from water to any electron<br />

acceptor by intact chloroplasts when exposed to light (Allen and Holmes, 1986).<br />

Juglone inhibited the Hill reaction in suspensions <strong>of</strong> banana chloroplasts. In<br />

addition, ‘Fougamou’ chloroplasts appeared to be less affected by juglone than<br />

‘Grande Naine’ chloroplasts. These results suggest that chloroplasts are one <strong>of</strong> the<br />

primary action sites <strong>of</strong> juglone.<br />

The role <strong>of</strong> toxins in banana-M. fijiensis interactions<br />

The electrolyte leakage assay and chlorophyll fluorescence were used to compare the<br />

sensitivity to EaCE <strong>of</strong> different banana cultivars with their behaviour in the field (highly<br />

resistant, sensitive or partially resistant) as scored using the rank <strong>of</strong> the youngest <strong>leaf</strong><br />

<strong>spot</strong>ted with necrotic lesions.<br />

These toxin assays confirmed that the incompatible interactions <strong>of</strong> ‘Yangambi km5’<br />

were not related to resistance to EaCE. The toxicity <strong>of</strong> EaCE preparations was<br />

independent <strong>of</strong> the virulence <strong>of</strong> the strain (unpublished data). Mechanisms <strong>of</strong> resistance<br />

in highly resistant cultivars were definitely not related to the action <strong>of</strong> these toxic<br />

metabolites.<br />

156


Session 3<br />

P. Lepoivre et al.<br />

Considering just the susceptible ‘Grande naine’ and the partially resistant<br />

‘Fougamou’ cultivars, susceptibility to EaCE was correlated with sensitivity to<br />

infection, suggesting that slow lesion development is associated with a lower<br />

sensitivity to M. fijiensis toxins.<br />

Such quantitative assessment is difficult to interpret because the concentrations<br />

<strong>of</strong> toxin(s) that were used in the bioassay could exceed the in vivo concentration<br />

and affect the mode <strong>of</strong> action <strong>of</strong> EaCE, hence affecting the rating <strong>of</strong> the cultivars.<br />

The hypothesis that M. fijiensis metabolites have a secondary role as determinants<br />

<strong>of</strong> pathogenicity agrees with cytological studies, which showed no evidence <strong>of</strong> an<br />

early effect <strong>of</strong> toxic compounds in the long period <strong>of</strong> biotrophy before observing<br />

the first cytological alterations in the mesophyll cells.<br />

Selection <strong>of</strong> banana tissues resistant to juglone<br />

The work was done with ‘Three hand planty’, a genotype susceptible to black <strong>leaf</strong><br />

streak disease, and juglone for which an embryo cell suspension was available.<br />

Juglone was toxic to embryogenic cell suspensions and somatic embryos <strong>of</strong> the<br />

cultivar. Necrosis <strong>of</strong> all cell suspensions and somatic embryos was quickly observed<br />

at 50 ppm or more <strong>of</strong> juglone, with the exception <strong>of</strong> some somatic embryos that<br />

continued development after treatment with 50 ppm <strong>of</strong> juglone.<br />

The plants regenerated from the surviving embryos showed a higher resistance<br />

to juglone: 250 ppm was required to induce necrosis in the <strong>leaf</strong> puncture bioassay<br />

with selected plantlets in comparison with 100 ppm for non selected plants.<br />

However, the selected plants did not show higher resistance to black <strong>leaf</strong> streak disease<br />

than the mother Three hand planty’ genotype following inoculation with M. fijiensis<br />

(El Hadrami, unpublished data).<br />

Daub (1986) advised caution when using metabolites from pathogens to screen<br />

tissue cultures for resistance. Nevertheless, fungal toxins have been proposed to screen<br />

banana in vitro (Strobel et al., 1993). There have been claims that resistant material<br />

has been produced by selecting callus <strong>of</strong> banana that survived increasing<br />

concentrations <strong>of</strong> M. fijiensis toxins (Okole and Shulz, 1993). Our results confirm<br />

the possibility <strong>of</strong> selecting banana plants resistant to M. fijiensis metabolites but<br />

this approach did not result in higher resistance to black <strong>leaf</strong> streak disease.<br />

References<br />

Allen J.F. and N.G. Holmes. 1986. Electron transport and Redox Titration. Pp. 103-141<br />

in Photosynthesis energy transduction. A practical approach. (M.F. Hipkins and N.R. Baker,<br />

eds). IRL PRESS, Washington.<br />

Atkinson M.M. 1993. Molecular mechanisms <strong>of</strong> pathogen recognition by plants. Advances<br />

in Plant Pathology 10:35-59.<br />

Beveraggi A. 1992. Etude des interactions hôte-parasite chez des bananiers sensibles et<br />

résistants inoculés par Cercospora fijiensis responsables de la maladie des raies noires.<br />

Thèse de 3 ème cycle, Université de Montpellier II, USTL.<br />

Beveraggi A., X. Mourichon and G. Salle. 1995. Etude comparée des premières étapes de<br />

l’infection chez les bananiers sensibles et résistants infectés par Cercospora fijiensis<br />

157


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

(<strong>Mycosphaerella</strong> fijiensis), agent responsable de la maladie des raies noires. Canadian<br />

Journal <strong>of</strong> Botany 73:1328-1337.<br />

Boller T. 1987. Hydrolytic enzymes in plant disease resistance. Pp. 385-413 in Plant-Microbe<br />

Interactions, vol.2 (Kosuge, T. and E.W. Nester, E. W. eds.). MacMillan, New York.<br />

Daub M.E. 1986. Tissue culture and the selection <strong>of</strong> resistance to pathogens. Annual Review<br />

<strong>of</strong> Phytopathology 24:159-186.<br />

de Wit P.G.J.M. 1992. Molecular characterization <strong>of</strong> gene-for-gene systems in plant fungus<br />

interactions and the application <strong>of</strong> avirulence genes in control <strong>of</strong> plant pathogens. Annual<br />

Review <strong>of</strong> Phytopathology 30:391-418.<br />

de Wit P.G.J.M., E.E. H<strong>of</strong>man, G.C.M. Velthius and J. Kuc. 1985. Isolation and characterization<br />

<strong>of</strong> an elicitor <strong>of</strong> necrosis isolated from intercellular fluids <strong>of</strong> compatible interactions <strong>of</strong><br />

Cladosporium fulvum (syn. Fulvia fulva) and tomato. Plant Physiology 77:642-647.<br />

El Hadrami A. 1997. Protoanthocyanidines constitutifs des feuilles de bananiers et résistance<br />

partielle vis-à-vis de <strong>Mycosphaerella</strong> fijiensis, l’agent causal de la maladie des raies noires.<br />

Thèse de DEA, Faculté universitaire des Sciences Agronomiques de Gembloux.<br />

Fullerton R.A. and T.L. Olsen. 1995. Pathogenic variability in <strong>Mycosphaerella</strong> fijiensis Morelet<br />

cause <strong>of</strong> black Sigatoka in banana and plantain. New Zealand Journal <strong>of</strong> Crop and<br />

Horticultural Science 23:39-48.<br />

Fouré E., A. Mouliom Pefoura and X. Mourichon. 1990. Etude de la sensibilité variétale des<br />

bananiers et plantains à <strong>Mycosphaerella</strong> fijiensis Morelet au Cameroun. Caractérisation<br />

de la résistance au champ de bananiers appartenant à divers groupes génétiques. Fruits<br />

45:339-345.<br />

Gire A. 1994. Relation entre la résistance partielle du bananier à Cercospora fijiensis et une<br />

composante cellulaire constitutive de nature polyphénolique. DEA, Université Montpellier II.<br />

Lepoivre P. and C.P. Acuna. 1989. Production <strong>of</strong> toxins by <strong>Mycosphaerella</strong> fijiensis var.<br />

difformis and induction <strong>of</strong> antimicrobial compounds in banana: their relevance in breeding<br />

for resistance to black Sigatoka. Pp. 201-207 in Sigatoka Leaf Spot Diseases <strong>of</strong> Bananas.<br />

Proceedings <strong>of</strong> an international workshop, San José, Costa Rica, March 28-April 1, 1989<br />

(Fullerton, R.A. and R.H. Stover eds). INIBAP Montpellier, France.<br />

Molina G., and J.P. Krausz. 1989. A phytotoxic activity in extracts <strong>of</strong> broth cultures <strong>of</strong><br />

<strong>Mycosphaerella</strong> fijiensis var. difformis and its use to evaluate host resistance to Black<br />

Sigatoka. Plant Disease 73:142-144.<br />

Mourichon X., D. Peter and M.F. Zapater. 1987. Inoculation expérimentale de M. fijiensis<br />

Morelet sur jeunes plantules de bananier issues de culture in vitro. Fruits 42:195-198.<br />

Okole B.N. and F.A. Shulz. 1993. Selection <strong>of</strong> banana and plantain (Musa spp.) resistant to<br />

toxins produced by <strong>Mycosphaerella</strong> species using in vitro culture techniques. Pp. 378 in<br />

Breeding banana and plantain for resistance to <strong>diseases</strong> and pests. Proceeding <strong>of</strong> the<br />

International Symposium on Genetic Improvement <strong>of</strong> Bananas to Diseases and Pests.<br />

(Ganry J. ed.). CIRAD-FHLOR, Montpellier, France, 7-9 septembre 1992.<br />

Riveros A.S. and P. Lepoivre. 1994. Early induction <strong>of</strong> non specific cultivar glucanase eliciting<br />

activity in the apoplastic fluids <strong>of</strong> banana infected by <strong>Mycosphaerella</strong> fijiensis. Arch. Int.<br />

Physiol. Bioch. Biophys. 102(5):9.<br />

Schaffrath U., H. Schelinpflug and H.J. Reisner. 1995. An elicitor from Pyricularia oryzae<br />

induces resistance response in rice: isolation, characterization and physiological properties.<br />

Physiol. and Mol. Plant Pathol. 46:293-307.<br />

Sharp J.K., M. McNeil and P. Albersheim. 1984. The primary structures <strong>of</strong> one elicitor-active<br />

and seven elicitor-inactive hexa (ß-D-glucopyranosyl)-D-glucitols isolated from mycelial<br />

walls <strong>of</strong> Phytophthora megasperma f. sp. glycinea. J. Biol. Chem. 259:11321-11336.<br />

158


Session 3<br />

P. Lepoivre et al.<br />

Stierle, A.A., R. Upadhyay, J. Hershenhorn, G.A. Strobel and G. Molina. 1991. The phytotoxins<br />

<strong>of</strong> <strong>Mycosphaerella</strong> fijiensis, the causative agent <strong>of</strong> black Sigatoka disease <strong>of</strong> <strong>bananas</strong> and<br />

plantains. Experientia 47:853-859.<br />

Strobel G.A., A.A. Stierle, R. Upadhyay, J. Hershenhorn and G. Molina. 1993. The phytotoxins<br />

<strong>of</strong> <strong>Mycosphaerella</strong> fijiensis, the causative agent <strong>of</strong> black sigatoka disease, and their potential<br />

use in screening for disease resistance. Pp. 93-103 in Biotechnology applications for banana<br />

and plantain improvement, 27-31 janvier 1992. INIBAP, Montpellier, France.<br />

Upadhyay, R., G.A. Strobel and S. Coval. 1990. Some toxins <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis.<br />

Pp. 231-236 in Sigatoka Leaf Spot Diseases <strong>of</strong> Bananas: Proceedings <strong>of</strong> an international<br />

workshop San José, Costa Rica, March 28-April 1, 1989 (Fullerton, R.A. and R.H. Stover<br />

eds). INIBAP Montpellier, France.<br />

159


Session 3<br />

C. Abadie et al.<br />

Efficiency and durability<br />

<strong>of</strong> partial resistance against<br />

black <strong>leaf</strong> streak disease<br />

C. Abadie 1,2 ,A.Elhadrami 2 ,E.Fouré 1 and J. Carlier 2<br />

Abstract<br />

Black <strong>leaf</strong> streak disease caused by <strong>Mycosphaerella</strong> fijiensis is the most destructive <strong>leaf</strong> disease<br />

<strong>of</strong> <strong>bananas</strong> and plantains. Genetic improvement for resistance appears as the most appropriate<br />

tool to control the disease. As a high level <strong>of</strong> diversity is maintained in pathogen populations,<br />

breeders prefer working with partial resistance, which is thought to be durable, instead <strong>of</strong> total<br />

resistance. Our aim is to evaluate the efficiency and durability <strong>of</strong> partial resistance. To achieve<br />

this objective, three complementary approaches were undertaken:<br />

1) Partial resistance was characterized by measuring various variables over the life cycle <strong>of</strong> the<br />

fungus under field and controlled conditions. The evaluation <strong>of</strong> 13 partially resistant varieties<br />

revealed the existence <strong>of</strong> several components acting at various stages <strong>of</strong> the infectious cycle.<br />

2) The efficiency <strong>of</strong> two resistant varieties which differ for two resistance components (infection<br />

efficacy, ascospores production) were studied. No difference in disease dispersal and incidence<br />

was observed between resistant varieties during the first year whereas small differences in disease<br />

severity, increasing over time, were measured. These results could be explained by differences<br />

in endogeneous inoculum production. Experiments are conduced to measure endogeneous<br />

inoculum in each field to confirm this hypothesis.<br />

3) The durability <strong>of</strong> resistance is being studied by analyzing the evolution <strong>of</strong> pathogen populations.<br />

Molecular characterization using CAPS markers was used on populations isolated after 6 and 25<br />

months <strong>of</strong> cultivation. No significant difference between the populations taken from susceptible<br />

and resistant <strong>bananas</strong> was observed after 6 months. Pathogenicity variability was undergone<br />

to assess an eventual selective effect <strong>of</strong> hosts.<br />

Resumen - Eficacia y durabilidad de la resistencia parcial contra la enfermedad de<br />

la raya negra de la hoja<br />

La enfermedad de la raya negra de la hoja causada por <strong>Mycosphaerella</strong> fijiensis es la enfermedad<br />

foliar más destructiva de bananos y plátanos. El mejoramiento genético con respecto a la<br />

resistencia parece convertirse en la medida de control más apropiada. Debido a un alto nivel de<br />

diversidad que se mantiene en las poblaciones del patógeno, los programas de mejoramiento<br />

se basan en la resistencia parcial que se considera hasta más durable que la resistencia total. El<br />

propósito de estos estudios consiste en evaluar la eficacia y durabilidad de la resistencia parcial<br />

1<br />

CARBAP, Douala, Cameroon<br />

2<br />

CIRAD, Montpellier, France<br />

161


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

para ayudar a los programas de mejoramiento. Para lograr este objetivo, se examinaron tres<br />

enfoques complementarios:<br />

1) La caracterización de la resistencia parcial se estudió a través de la medición de diferentes<br />

parámetros del ciclo de vida del hongo en condiciones de campo y controladas. La evaluación de<br />

13 variedades diferentes con resistencia parcial reveló la existencia de varios componentes de<br />

resistencia que actúan en diferentes etapas del ciclo de infección.<br />

2) La eficacia de dos variedades resistentes en las cuales los dos componentes de resistencia (la<br />

eficacia de la infección y la producción de ascosporas) se estudiaron. Durante el primer año no<br />

se observaron diferencias entre las dos variedades en cuanto a dispersión espacial o su incidencia.<br />

Sin embargo, las mediciones indicaron bajas diferencias en la severidad de la enfermedad, la cual<br />

aumentó en los años siguientes. Las diferencias en la producción del inóculo endógeno podrían<br />

explicar los resultados arriba mencionados. Varios experimentos que se están llevando a cabo<br />

actualmente para estudiar el inóculo endógeno podrían confirmar la hipótesis.<br />

3) Se está estudiando la durabilidad de la enfermedad a través de la evolución de las poblaciones<br />

del patógeno. La caracterización molecular con los marcadores CAPS se utilizó en las poblaciones<br />

aisladas después de 6 y 25 meses de cultivo. Después de seis meses no se observaron diferencias<br />

significativas entre las dos poblaciones provenientes de los bananos susceptibles y resistentes.<br />

Se estudió la variabilidad en el poder patógeno para evaluar un eventual efecto selectivo de los<br />

hospedantes.<br />

Résumé - Efficacité et durabilité de la résistance partielle contre la maladie des raies<br />

noires<br />

La maladie des raies noires, causée par <strong>Mycosphaerella</strong> fijiensis, est la maladie foliaire la plus<br />

destructrice chez les bananiers et les plantains. L’amélioration génétique pour la résistance apparaît<br />

comme le moyen le mieux approprié pour contrôler la maladie. Comme un niveau élevé de diversité<br />

est maintenu dans les populations de pathogènes, les sélectionneurs préfèrent travailler avec<br />

une résistance partielle, qui est considérée comme étant plus durable, plutôt qu’avec une<br />

résistance totale. Trois approches complémentaires ont été suivies pour atteindre cet objectif :<br />

1) La résistance partielle a été caractérisée en mesurant différentes variables pendant le cycle<br />

vital du champignon au champ et en conditions contrôlées. L’évaluation de 13 variétés partiellement<br />

résistantes a révélé l’existence de plusieurs composants agissant à des stades différents du cycle<br />

infectieux.<br />

2) L’efficience de deux variétés résistantes qui différaient pour leurs composants de résistance<br />

(efficacité de l’infection, production d’ascospores) a été étudiée. Aucune différence dans la<br />

dispersion et l’incidence de la maladie n’a été observée entre les variétés résistantes pendant la<br />

première année, alors que de petites différences ont été notées pour la sévérité de la maladie,<br />

qui augmentait avec le temps. Ces résultats pourraient être expliqués par des différences dans<br />

la production d’inoculum endogène. Des essais sont en cours pour mesurer l’inoculum endogène<br />

dans chaque champ pour confirmer cette hypothèse.<br />

3) La durabilité de la résistance a été étudiée en analysant l’évolution des populations de<br />

pathogènes. La caractérisation moléculaire avec des marqueurs CAPS a été utilisée sur des<br />

populations isolées après 6 et 25 mois de culture. Aucune différence significative n’a été trouvée<br />

entre des populations prélevées sur des bananiers résistants et sensibles. La variabilité de la<br />

pathogénicité a été étudiée pour évaluer la possibilité d’un effet sélectif des hôtes.<br />

Introduction<br />

Black <strong>leaf</strong> streak disease, caused by the ascomycete fungus <strong>Mycosphaerella</strong><br />

fijiensis, is the most important foliar disease <strong>of</strong> banana worldwide (Jones, 2000).<br />

Indeed, the main varieties cultivated industrially and by smallholders, and<br />

belonging to the Cavendish and plantains groups, are very susceptible to black<br />

162


Session 3<br />

C. Abadie et al.<br />

<strong>leaf</strong> streak disease. Chemical control strategies based on biological forecasting<br />

system were first developed in industrial plantations (Fouré, 1988). Although<br />

efficient, chemical control negatively affects the environment and human health,<br />

and is too expensive for poor smallholders. Managing black <strong>leaf</strong> streak disease<br />

must integrate the use <strong>of</strong> resistant varieties.<br />

Two types <strong>of</strong> resistance against black <strong>leaf</strong> streak disease have been described<br />

(Fouré, 1992). High resistance, characterized by a blockage <strong>of</strong> symptom expression<br />

and an absence <strong>of</strong> sporulation, and partial resistance, characterized by moderate<br />

disease expression and a normal but slow development <strong>of</strong> symptoms up to necrosis.<br />

High resistance is believed to be similar to hypersensitivity and under the control<br />

<strong>of</strong> a mono or oligogenic system which may be easily circumvented by pathogens<br />

(Fouré et al., 2000). For example, the high resistance <strong>of</strong> ‘Paka’ is no longer effective<br />

in the Cook Islands and studies <strong>of</strong> pathogenic variability showed that it was<br />

overcome by some virulent strains after 8 years <strong>of</strong> cultivation (Fullerton and Olsen,<br />

1995).<br />

Analysis <strong>of</strong> population genetic structure <strong>of</strong> M. fijiensis at various geographical<br />

scales revealed high levels <strong>of</strong> gene diversity and genetic differentiation (Carlier<br />

et al., 1996 and in this volume). Such structures suggest a high adaptation potential<br />

<strong>of</strong> the pathogen, hence the use in the banana breeding programmes <strong>of</strong> CARBAP<br />

and CIRAD <strong>of</strong> partial resistance instead <strong>of</strong> high resistance.<br />

Partial resistance is a complex character which could include several components<br />

corresponding to different stages <strong>of</strong> pathogen infectious cycle (Young, 1996). Until<br />

recently, only one variable, youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS), was used to evaluate partial<br />

resistance <strong>of</strong> <strong>bananas</strong> in screening trials. Such evaluation was useful to characterise<br />

germplasm but not sufficient to identify the components <strong>of</strong> partial resistance. For<br />

example, sporulation, which could have an important effect in the case <strong>of</strong> a<br />

polycyclic epidemic like black <strong>leaf</strong> streak disease, is not measured. Identifying the<br />

components coming into play at various stages <strong>of</strong> the infectious cycle can be<br />

conducted by inoculation under controlled conditions.<br />

The efficiency <strong>of</strong> the components <strong>of</strong> partial resistance can only be evaluated<br />

in the framework <strong>of</strong> an epidemiological study in a field <strong>of</strong> one variety. Pathogen<br />

populations could evolve and erode resistance depending on the evolutionary forces<br />

at work. Strategies based on the evolutionary potential <strong>of</strong> the pathogen could<br />

improve the durability <strong>of</strong> resistance. To define such strategies, the relative<br />

importance <strong>of</strong> the evolutionary forces acting on the pathogen. The effect <strong>of</strong> genetic<br />

recombination, genetic drift and gene flow on pathogen evolution can be evaluated<br />

by analyzing population structure (Carlier et al, 1996 and this volume).<br />

Resistant hosts exert a selection pressure on the pathogens. The evolution <strong>of</strong><br />

the pathogen can be studied in fields <strong>of</strong> resistant hosts by using molecular markers<br />

and characterizing its pathogenicity. The durability <strong>of</strong> partial resistance will be<br />

studied by following over time the evolution <strong>of</strong> pathogens in fields <strong>of</strong> resistant<br />

hosts.<br />

The objectives <strong>of</strong> this study are to identify components <strong>of</strong> partial resistance to<br />

black <strong>leaf</strong> streak disease in <strong>bananas</strong>, and to evaluate their efficiency and durability<br />

in controlling the disease. A three-step experimental approach was developed: 1)<br />

characterization <strong>of</strong> partial resistance <strong>of</strong> various cultivars under controlled<br />

163


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

conditions, 2) epidemiological study using one variety to evaluate the efficiency<br />

<strong>of</strong> the components <strong>of</strong> partial resistance, and 3) analysis <strong>of</strong> the pathogen population<br />

over time to evaluate the durability <strong>of</strong> partial resistance in relation to the selection<br />

pressure exerted by the host.<br />

Components <strong>of</strong> partial resistance<br />

To characterize partial resistance, variables corresponding to the different stages <strong>of</strong><br />

an infectious cycle were estimated and compared among various partially resistant<br />

varieties. Evaluations <strong>of</strong> resistance were conducted in Cameroon on 7 partially<br />

resistant and susceptible cultivars under field conditions and on 10 cultivars under<br />

controlled conditions. In the latter, pieces <strong>of</strong> banana <strong>leaf</strong> maintained on a culture<br />

media were inoculated (El Hadrami et al., 1998). Fifteen strains were used.<br />

The infectious cycle was dissected in eight stages (Figure 1) including incubation<br />

period; spore efficacy (number <strong>of</strong> lesions); size and growth rate <strong>of</strong> lesions; asexual<br />

and sexual latency period; asexual and sexual sporulation capacity.<br />

Many variables were significantly different between resistant varieties under field<br />

and controlled conditions. For example, the lesions were significantly smaller in<br />

‘Zebrina’ and ‘Pisang Ceylan’ than in other resistant varieties (Figures 2a and 2b).<br />

The number <strong>of</strong> lesions and the incubation period also varied significantly with the<br />

highest number <strong>of</strong> lesions found on ‘Pisang Berlin’ and ‘Pisang Ceylan’ and the<br />

Production <strong>of</strong> ascospores<br />

Production <strong>of</strong> conidia<br />

Number, size & growth rate <strong>of</strong> lesions<br />

Ascopores<br />

Conidia<br />

Penetration<br />

First lesions<br />

Sporulating lesions<br />

with conidia<br />

Necrosis<br />

Sporulating necrosis<br />

with ascospores<br />

Epiphyl phase<br />

Asexual phase<br />

Sexual phase<br />

Incubation period<br />

Asexual latency period<br />

Sexual latency period<br />

Infectious ascospores<br />

Figure 1. Infectious cycle <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis (from El Hadrami, 2000). The components <strong>of</strong> partial<br />

resistance being characterized are in italics.<br />

164


Session 3<br />

C. Abadie et al.<br />

Size <strong>of</strong> lesion<br />

2<br />

(% <strong>of</strong> a 25-cm piece <strong>of</strong> <strong>leaf</strong>)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

a<br />

de<br />

d<br />

Controlled conditions<br />

b<br />

e<br />

ef<br />

c<br />

fg<br />

ef<br />

g<br />

0<br />

CAV<br />

a<br />

Field conditions<br />

bc<br />

bc<br />

ab<br />

c<br />

ab<br />

c<br />

CAV<br />

FSO<br />

PMA<br />

PBE<br />

ZEB<br />

FOU<br />

PCE<br />

PMA<br />

ROS<br />

PBE<br />

TDE<br />

PKW<br />

TAN<br />

ZEB<br />

FOU<br />

PCE<br />

S<br />

PR<br />

30<br />

Size <strong>of</strong> lesion<br />

(% <strong>of</strong> <strong>spot</strong>ted area)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

S<br />

PR<br />

Figure 2. Size <strong>of</strong> lesions in partially resistant (PR) and susceptible (S) varieties under (a) controlled conditions<br />

and under (b) field conditions. (Cav: Cavendish (AAA), FSO: French sombre (AAB), PMA: Pisang madu (AAcv),<br />

ROS: Rose d’Ekona, PBE: Pisang Berlin (AAcv), TDE: Thong det (AAcv), PKW: Pisang klutuk wulung (BBw),<br />

TAN (BBw), ZEB: Zébrina (AAw), FOU: Fougamou (ABB), PCE: Pisang C eylan (ABB)).<br />

longest incubation period observed on ‘Fougamou’ and ‘Tani’ (data not shown). No<br />

difference was observed in the production <strong>of</strong> asexual spores whereas significant<br />

differences were observed in the production <strong>of</strong> sexual spores (Figure 3). ‘Pisang madu’<br />

and ‘Zebrina’ produced 3 to 8 times less than the other three partially resistant<br />

varieties, including ‘Pisang Berlin’. Thus, although ‘Pisang madu’ and ‘Pisang<br />

Berlin’ behave similarly in field trials as regards YLS, ‘Pisang madu‘ produced four<br />

times less perithecia/cm 2 <strong>of</strong> necrosis than ‘Pisang Berlin’.<br />

165


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

This first step allowed us to identify components <strong>of</strong> partial resistance which act<br />

at different stages <strong>of</strong> the infectious cycle : early stage (e.g. small lesions on ‘Zebrina’<br />

and ‘Pisang Ceylan’), intermediate stages (e.g. long incubation period on ‘Fougamou’<br />

and ‘Tani’) or end <strong>of</strong> the cycle (e.g. low production <strong>of</strong> sexual spores on ‘Pisang madu’<br />

and ‘Zebrina’).<br />

1600<br />

a<br />

Field conditions<br />

2<br />

Number <strong>of</strong> perithecia/cm <strong>of</strong> necrosis<br />

1200<br />

800<br />

400<br />

bc<br />

bc<br />

b<br />

bc<br />

c<br />

c<br />

0<br />

CAV<br />

FSO<br />

PMA<br />

PBE<br />

ZEB<br />

FOU<br />

PCE<br />

S<br />

PR<br />

Figure 3. Production <strong>of</strong> sexual spores in five partially resistant (PR) and two susceptible (S) varieties under<br />

field conditions.(Cav: Cavendish (AAA), FSO: French sombre (AAB), PMA: Pisang madu (AAcv), PBE: Pisang Berlin<br />

(AAcv), ZEB: Zébrina (AAw), FOU: Fougamou (ABB), PCE: Pisang Ceylan (ABB)).<br />

Efficiency <strong>of</strong> components<br />

An epidemiological study was conducted on two varieties to evaluate the efficiency<br />

<strong>of</strong> two components <strong>of</strong> partial resistance: spore efficacy and production <strong>of</strong> sexual<br />

spores. ‘Pisang madu’, ‘Pisang berlin’ and a susceptible control (‘Grande naine’) were<br />

cultivated in three rectangular plots containing 150 <strong>bananas</strong>/plot. Measures were<br />

taken over 3 cropping cycles.<br />

During the first year, no difference in disease dispersal and disease incidence was<br />

observed between the 3 plots. No spatial auto-correlation was measured at plot scale<br />

(Moran index not significant) and disease incidence was similar between plots (data<br />

not shown).<br />

Disease severity (visual quantification <strong>of</strong> percentage <strong>of</strong> <strong>spot</strong>ted surface per plant)<br />

was significantly different in resistant varieties (about 20 % for ‘Pisang Berlin’) and<br />

susceptible varieties (an average <strong>of</strong> 40%) (Table 1). This result shows the efficient<br />

role <strong>of</strong> partial resistance in controlling black <strong>leaf</strong> streak disease. Significant<br />

166


Session 3<br />

C. Abadie et al.<br />

differences were also observed between the two resistant varieties, differences<br />

which increased over time due to the declining disease severity on ‘Pisang madu’<br />

(Table 1). Production <strong>of</strong> sexual spores, on the other hand, was stable over the three<br />

cropping cycles. This result could be due to differences in components <strong>of</strong> resistance<br />

or an increase in resistance ‘Pisang madu’ over time. To test the first hypothesis,<br />

secondary inoculum was measured by using spore trap plants in each resistant plot.<br />

The precision <strong>of</strong> this method needs to be improved to be able to show eventual<br />

differences in auto-inoculum production.<br />

Table 1. Efficacy <strong>of</strong> partial resistance expressed as disease severity (% <strong>of</strong> <strong>spot</strong>ted area/banana) on susceptible (S)<br />

and partially resistant (PR) varieties over three cropping cycles.<br />

Varieties<br />

Disease severity (% <strong>of</strong> <strong>spot</strong>ted area)<br />

1 st cycle 2 nd cycle 3 rd cycle<br />

Grande naine (S) 37.5 a 39 a 36 a<br />

Pisang Berlin (PR) 22.7 b 18.6 b 17.2 b<br />

Pisang madu (PR) 17.5 c 7.1 c 0.8 c<br />

Durability <strong>of</strong> components partial resistance<br />

The population structures <strong>of</strong> about 50 isolates <strong>of</strong> M. fijiensis taken from resistant<br />

(‘Pisang madu’) and susceptible (‘Grande naine’) after 6 and 25 months <strong>of</strong> cultivation<br />

were analysed using molecular markers and a pathogenic test to evaluate the relative<br />

importance <strong>of</strong> genetic drift and selection by the host.<br />

Seven CAPS (cleaved amplified polymorphism sequences) neutrals markers were<br />

used and the pathogenicity <strong>of</strong> the isolates was measured by inoculating <strong>leaf</strong> pieces<br />

<strong>of</strong> ‘Pisang madu’ and ‘Grande naine’ maintained on a culture media (El Hadrami<br />

et al., 1998).<br />

No significant difference in genetic differentiation was detected between the<br />

samples isolated from resistant and susceptible <strong>bananas</strong>. The estimate <strong>of</strong> Wright’s<br />

Fst parameter over all loci between pairs <strong>of</strong> populations was low and ranged from<br />

0.007 to 0.0432. This absence <strong>of</strong> differentiation between populations suggests a low<br />

genetic drift effect during colonisation and/or important gene flow between fields.<br />

No clear evidence <strong>of</strong> differences in aggressiveness were observed between the<br />

two pathogens after 6 months <strong>of</strong> cultivation. Samples isolated from resistant and<br />

susceptible <strong>bananas</strong> seemed to have the same pathogenic behaviour. Further<br />

analyses will be done on populations after 25 months <strong>of</strong> cultivation.<br />

Conclusion and perspectives<br />

This study revealed the existence <strong>of</strong> various components <strong>of</strong> partial resistance under<br />

controlled and field conditions. Then, the efficiency <strong>of</strong> two components <strong>of</strong><br />

resistance (spore efficacy, production <strong>of</strong> sexual spores) on disease control were<br />

tentatively evaluated. A selective effect <strong>of</strong> partial resistance components on<br />

pathogen population was not detected but aggressiveness <strong>of</strong> isolates were evaluated<br />

only after six months <strong>of</strong> cultivation. We are analysing pathogens population after<br />

167


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

two years <strong>of</strong> cultivation. However, the absence <strong>of</strong> genetic differentiation between<br />

fields containing different varieties could be the result <strong>of</strong> high gene flow. High<br />

gene flow can counteract changes in gene frequency as a result <strong>of</strong> selection by<br />

the host. To clarify this point, spore dispersal needs to be specified directly in<br />

epidemiological studies or indirectly ing population genetic models.<br />

Computer models can now simulate epidemics using parameters corresponding<br />

to the different stages <strong>of</strong> the pathogen’s infectious cycle. The effect and the<br />

importance <strong>of</strong> the components <strong>of</strong> partial resistance can then be tested by<br />

comparing simulated and observed epidemics. Computer models can help in<br />

designing future experiments and in defining spatio-temporal management<br />

strategies <strong>of</strong> resistant varieties at different scales (from plot to regional scale) by<br />

testing different scenarios.<br />

References<br />

Carlier J., M.H. Lebrun, M. F. Zapater, C. Dubois and X. Mourichon. 1996. Genetic structure<br />

<strong>of</strong> the global population <strong>of</strong> <strong>bananas</strong> black <strong>leaf</strong> streak fungus <strong>Mycosphaerella</strong> fijiensis.<br />

Molecular Ecology 5:499-510.<br />

El Hadrami A., M.F. Zapater, F. Lapeyre, C. Abadie and J. Carlier. 1998. A <strong>leaf</strong> disk assay to<br />

assess partial resistance <strong>of</strong> banana germplasm and agressiveness <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis,<br />

the causal aget <strong>of</strong> black <strong>leaf</strong> streak disease. 7 th International Congress <strong>of</strong> Plant Pathology,<br />

Edinburgh, Scotland. BSPP vol. 2, p.1.1.24.<br />

El Hadrami, A. 2000. Caractérisation de la résistance partielle des bananiers à la maladie des<br />

raies noires et évaluation de la variabilité de l’agressivité de l’agent causal, <strong>Mycosphaerella</strong><br />

fijiensis. Thèse d’Université. Faculté Universitaire des Sciences Agronomiques de Gembloux,<br />

Belgique. 153pp.<br />

Fouré E. 1988. Stratégies de lutte contre la cercosporiose noire des bananiers et plantains<br />

provoquée par <strong>Mycosphaerella</strong> fijiensis Morelet. L’avertissement biologique au Cameroun.<br />

Evaluation des possibilités d’amélioration. Fruits, vol 43(5): 269-274.<br />

Fouré E. 1992. Characterization <strong>of</strong> the reactions <strong>of</strong> <strong>bananas</strong> cultivars <strong>Mycosphaerella</strong> fijiensis<br />

Morelet in Cameroon and genetics <strong>of</strong> resistance. Pp. 159-170 in Breeding banana<br />

and plantain for resistance to <strong>diseases</strong> and pests, Proceedings <strong>of</strong> the International<br />

Symposium on Genetic Improvement <strong>of</strong> Bananas for Resistance to Diseases and Pests<br />

7-9 september 1992, Montpellier, France.<br />

Fouré E., X. Mourichon and D. Jones. 2000. Evaluating germplasm for reaction to black <strong>leaf</strong><br />

streak. P 62-72 in Diseases <strong>of</strong> Banana, Abaca and Enset. (Jones D. ed.), CAB International,<br />

Wallingford, UK, 544pp.<br />

Fullerton R.A. and T. L. Olsen. 1995. Pathogenic variability in <strong>Mycosphaerella</strong> fijiensis Morelet<br />

cause <strong>of</strong> black Sigatoka in banana and plantain. New Zealand Journal <strong>of</strong> Crop and<br />

Horticultural Science 23:39-48.<br />

Jones D. 2000. Diseases <strong>of</strong> Banana, Abaca and Enset. CAB International, Wallingford, UK,<br />

544pp.<br />

Young N.D. 1996. QTL mapping and quantitative disease resistance in plants. Annual review<br />

<strong>of</strong> phytopathology, 34:479-501.<br />

168


Session 3<br />

Y. Alvarado Capó et al.<br />

Poster<br />

Early evaluation <strong>of</strong> black <strong>leaf</strong> streak<br />

resistance by using mycelial<br />

suspensions <strong>of</strong> <strong>Mycosphaerella</strong><br />

fijiensis<br />

Y. Alvarado Capó,M.Leiva Mora, M. A. Dita Rodríguez,<br />

M. Acosta, M. Cruz, N. Portal, R. Gómez Kosky,<br />

L. García, I. Bermúdez and J. Padrón<br />

Abstract<br />

A standardized method for the early evaluation <strong>of</strong> resistance to black <strong>leaf</strong> streak on in vitro Musa<br />

plants was developed using mycelial suspensions <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis. Seven cultivars:‘FHIA-<br />

18’, ‘FHIA-01’, ‘FHIA-21’, ‘Grande naine’, ‘Yangambi’, ‘Calcutta 4’ and ‘Niyarma yik’ were tested in a<br />

greenhouse. Inoculum was adjusted to 10 5 cfu/ml and applied to the lower surfaces <strong>of</strong> the first<br />

three open leaves. Plants were evaluated 15 days after inoculation and at 15-day intervals until<br />

60 days. A standardized scale <strong>of</strong> <strong>leaf</strong> symptoms ensured consistency between evaluators. All<br />

cultivars except ‘Yangambi’, showed a similar response to M. fijiensis in natural conditions. Partial<br />

resistance expressed in FHIA cultivars was characterized by a slow rate <strong>of</strong> symptom development<br />

with ‘Calcutta 4’ the slowest. ‘Grande naine’ and ‘Niyarma yik’ gave a susceptible reaction and<br />

their symptoms were more severe. Artificial inoculation <strong>of</strong> in vitro plants with mycelial suspensions<br />

was an easy, rapid and practicable method to determine resistance to M. fijiensis. An inoculum<br />

adjusted to an appropriate concentration gave uniform symptoms on the inoculated <strong>leaf</strong>. The<br />

method has promise for the evaluation <strong>of</strong> in vitro plants in breeding programmes.<br />

Resumen - Evaluación temprana de la resistencia a la raya negra de la hoja mediante<br />

el uso de la suspensión del micelio de <strong>Mycosphaerella</strong> fijiensis<br />

Se describen los métodos de normalización para la evaluación temprana de la resistencia a la<br />

raya negra de la hoja en las plantas in vitro de Musa mediante el uso de la suspensión del micelio<br />

Instituto de Biotecnología de las Plantas, Santa Clara, Cuba<br />

169


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

de <strong>Mycosphaerella</strong> fijiensis. Siete cultivares:‘FHIA-18’,‘FHIA-01’,‘FHIA-21’,‘Grande naine’,‘Yangambi’,<br />

‘Calcutta 4’ y ‘Niyarma Yik’ fueron utilizados para la prueba de inoculación en el invernadero. La<br />

concentración del inóculo fue ajustada a 10 5 ufc/ml y las primeras tres hojas en abrirse fueron<br />

inoculadas en la superficie inferior. El período de evaluación empezó a los 15 días y termino 60<br />

días después de la inoculación. En la evaluación de las hojas se utilizó una escala (unidad<br />

experimental) que permitió evitar la confusión del evaluador y describir el desarrollo de los<br />

síntomas de mejor manera. Con excepción del ‘Yangambí’, los cultivares mostraron un<br />

comportamiento similar contra los patógenos en condiciones naturales. La resistencia parcial<br />

expresada en los cultivares de la FHIA se caracterizó por una lenta evolución de los síntomas. El<br />

cultivar ‘Calcutta 4’mostró el tiempo más lento de desarrollo de los síntomas. Los cultivares ‘Grande<br />

naine’ y ‘Niyarma yik’ mantuvieron reacciones susceptibles y sus síntomas alcanzaron grados<br />

mayores de afectación. La inoculación artificial de las plantas in vitro utilizando la suspensión<br />

de micelio resultó ser un método fácil, rápido y factible para conocer la expresión de la resistencia<br />

de las plantas contra M. fijiensis. La utilización del inóculo con una concentración ajustable permitió<br />

obtener síntomas homogéneos y uniformes de la hoja inoculada. El mismo representa una<br />

herramienta útil para la evaluación de las plantas in vitro en los programas de mejoramiento.<br />

Résumé - Evaluation précoce de la résistance à la maladie des raies noires au moyen<br />

de suspensions mycéliennes de <strong>Mycosphaerella</strong> fijiensis<br />

Une méthode standardisée d’évaluation précoce de la résistance à la maladie des raies noires<br />

sur des vitroplants de Musa a été développée en utilisant des suspensions mycéliennes de<br />

<strong>Mycosphaerella</strong> fijiensis. Sept cultivars, ‘FHIA-18’, ‘FHIA-01’, ‘FHIA-21’, ‘Grande naine’, ‘Yangambi’,<br />

‘Calcutta 4’et ‘Niyarma yik’ont été étudiés en serre. L’inoculum a été ajusté à 10 5 cfu/ml et appliqué<br />

sur la surface inférieure des trois premières feuilles ouvertes. Les plantes ont été évaluées 15 jours<br />

après l’inoculation, puis à intervalles de 15 jours jusqu’à 60 jours. Une échelle standardisée de<br />

symptômes foliaires a permis d’assurer la consistance entre évaluateurs. Tous les cultivars sauf<br />

‘Yangambi’ ont montré une réponse similaire à M. fijiensis en conditions naturelles. La résistance<br />

partielle exprimée par les cultivars FHIA a été caractérisée par un faible taux de développement<br />

des symptômes, qui était le plus bas chez ‘Calcutta 4’. ‘Grande naine’ et ‘Niyarma yik’ ont donné<br />

une réaction susceptible et leurs symptômes étaient plus sévères. L’inoculation artificielle de<br />

vitroplants avec des suspensions mycéliennes s’est avérée une méthode simple, rapide et<br />

praticable pour déterminer la résistance à M. fijiensis. Un inoculum ajusté à une concentration<br />

appropriée a donné des symptômes uniformes sur la feuille inoculée. Cette méthode est<br />

prometteuse pour l’évaluation de vitroplants dans des programmes d’amélioration.<br />

Introduction<br />

Screening for resistance to black <strong>leaf</strong> streak in field condition is time-consuming<br />

and expensive. Early evaluation in controlled conditions is an important<br />

requirement to increase success and evaluate feasibility. Stover (1987)<br />

recommended the use <strong>of</strong> standard varieties and the development <strong>of</strong> standardized<br />

methods for preparing inoculum and measuring disease response in controlled<br />

environments.<br />

Mourichon et al. (1987) developed a greenhouse method for artificial<br />

inoculation with <strong>Mycosphaerella</strong> fijiensis conidia and mycelium onto three<br />

reference cultivars with different levels <strong>of</strong> resistance (‘Grande naine’, ‘Fougamou’<br />

and ‘Yangambi km 5 1/2 ’). Romero and Sutton (1997) used suspensions <strong>of</strong> conidia<br />

to evaluate the reaction <strong>of</strong> four Musa genotypes at three temperatures with isolates<br />

from different geographical regions. Jones (1995) used hyphal fragments <strong>of</strong><br />

M. musicola for the rapid assessment <strong>of</strong> different Musa spp., and Balint-Kurti et<br />

170


Session 3<br />

Y. Alvarado Capó et al.<br />

al. (2001) developed inoculation techniques using transgenic strains <strong>of</strong> M. fijiensis<br />

expressing GFP by using conidia and mycelia.<br />

The development <strong>of</strong> artificial inoculation techniques is also necessary to<br />

improve and simplify selection procedures in breeding programmes. Even though<br />

ascospores and conidia are commonly used for artificial inoculation, mycelial<br />

fragments may provide an alternative for evaluating disease development in<br />

controlled conditions.<br />

The objective <strong>of</strong> the work was to prepare a standardized method for the rapid<br />

evaluation <strong>of</strong> resistance to black <strong>leaf</strong> streak disease in in vitro Musa plants using<br />

mycelial suspensions <strong>of</strong> M. fijiensis, and its application in breeding programmes.<br />

Material and methods<br />

Plant material<br />

In vitro plants <strong>of</strong> ‘FHIA-18’ (AAAB), ‘FHIA-01’ (AAAB), ‘FHIA-21’ (AAAB), ‘Grande<br />

naine’ (AAA), ‘Yangambi’ (AAA), ‘Calcutta 4’ (AA) and ‘Niyarma yik’ (AA) were<br />

inoculated in a greenhouse. Plants were acclimatized for eight weeks on a substrate<br />

comprising 50% casting, 30% compost and 20% zeolite. Plants were grown in<br />

plastic pots 100 cm in diameter.<br />

Preparation <strong>of</strong> mycelial suspension<br />

The pathogenic monoascosporic strain CC-IBP-1 (isolated from ‘Grande naine’)<br />

was used. Colonies <strong>of</strong> M. fijiensis were grown on potato dextrose agar (PDA) at<br />

28 o C for 14 days. Pieces fragments were transferred to an Erlenmeyer flask<br />

containing 200 ml <strong>of</strong> V8 liquid medium (200 ml V8 juice, 0.3 g CaCO 3<br />

, 800 ml<br />

water, pH 6.0). The flask was incubated at 28 o C in a shaker (130 rpm) for 15 days.<br />

The mycelium was then blended and 1 ml was transferred to a Petri dish containing<br />

PDA to obtain homogeneous growth over the surface <strong>of</strong> the dish. The culture was<br />

incubated in the dark at 28 o C for 15 days. Two 1-cm 2 discs <strong>of</strong> mycelium were<br />

transferred to a 1-litre flask containing 400 ml <strong>of</strong> liquid V8, and shaken at 130<br />

rpm at 28 o C for 15 days. The mycelium (10 g in 900 ml <strong>of</strong> sterile distilled water)<br />

was blended and filtered through two layers <strong>of</strong> gauze to remove large fragments<br />

<strong>of</strong> hyphae. The concentration was determined with a haemacytometer and<br />

adjusted to 10 5 cfu/ml. Gelatin at 1% was added to the final inoculum.<br />

Inoculation<br />

Plants 20 cm in height and with four active leaves were selected for inoculation.<br />

There were 15 plants <strong>of</strong> each cultivar.<br />

The first three open leaves <strong>of</strong> each plant were inoculated on the under surface<br />

using a brush. Leaves were marked on the upper side. The plants were allowed<br />

to dry for 2 h, and the relative humidity maintained at 90-100% for the first three<br />

days by spraying continuously with water. Afterwards, the humidity was saturated<br />

during the night but spraying was suspended during the day. Sunlight was used.<br />

171


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Evaluation<br />

Inoculated leaves were examined every 15 days starting on the 15 th day after<br />

inoculation and ending on the 60 th . Symptoms corresponded approximately to<br />

the descriptions <strong>of</strong> Fullerton and Olsen (1995) for in vitro plants inoculated with<br />

conidia. Table 1 describes the scale used to evaluate symptom development and<br />

the classification <strong>of</strong> genotypes according to the stage <strong>of</strong> symptom development.<br />

Table 1. Stages <strong>of</strong> symptom development in in vitro Musa plants inoculated with mycelial suspensions <strong>of</strong><br />

<strong>Mycosphaerella</strong> fijiensis in the greenhouse.<br />

Stage Description<br />

0 Leaf symptoms mostly absent.<br />

1 Reddish flecks on lower <strong>leaf</strong> surface. No symptoms on the upper surface.<br />

2 Regular or irregular reddish circular <strong>spot</strong>s on the lower <strong>leaf</strong> surface. No symptoms on the upper surface.<br />

3 Regular or diffuse light brown circular <strong>spot</strong>s oin the upper <strong>leaf</strong> surface.<br />

4 Black or brown circular <strong>spot</strong>s, possibly with a yellow halo or chlorosis <strong>of</strong> adjacent tissues, on the upper <strong>leaf</strong><br />

surface. Areas <strong>of</strong> green tissue sometimes present.<br />

5 Black <strong>spot</strong>s with dry centre <strong>of</strong> grey colour. Leaf completely necrotic, sometimes hanging down.<br />

Classification <strong>of</strong> genotypes according to symptom development<br />

Resistant: stages 0-1<br />

Partially resistant: stages 2-3<br />

Susceptible: stages 4-5<br />

In vitro plants <strong>of</strong> ‘Grande naine’ and ‘FHIA-21’, obtained from IBP breeding<br />

programmes (by mutagenesis), were evaluated as described above. Figure 1<br />

describes the work schedule for the method.<br />

Results and discussion<br />

As described by Mourichon et al. (1987), symptoms developed on the leaves <strong>of</strong><br />

in vitro plants artificially inoculated with mycelial suspensions <strong>of</strong> M. fijiensis.<br />

The appearance <strong>of</strong> symptoms was similar to those observed on suckers in the field.<br />

The necrotic <strong>spot</strong>s were slightly circular, possibly because young plants derived<br />

from tissue culture have limited vein development and black <strong>leaf</strong> streak lesions<br />

tend to be spherical (Mourichon et al., 2000). The majority <strong>of</strong> lesions were <strong>of</strong>ten<br />

observed on the same <strong>leaf</strong> at the same stage <strong>of</strong> development. All cultivars, except<br />

‘Yangambi’, responded to M. fijiensis in a similar manner to that observed in<br />

natural conditions (Table 2).<br />

The reaction <strong>of</strong> ‘Yangambi’ was characterized by the presence <strong>of</strong> symptoms<br />

after the first 15 days, symptoms which reached stages 2 and 3 in 30-45 days<br />

similar to the behaviour <strong>of</strong> susceptible genotypes. Some leaves had necrotic <strong>spot</strong>s.<br />

Mourichon et al. (1987) had reported a hypersensitive reaction for in vitro plants<br />

<strong>of</strong> ‘Yangambi’ inoculated with conidia. In contrast, Fullerton and Olsen (1995)<br />

reported a typical susceptible response in ‘Yangambi’ when in vitro plants were<br />

inoculated with the conidia <strong>of</strong> a virulent strain <strong>of</strong> M. fijiensis.<br />

172


Session 3<br />

Y. Alvarado Capó et al.<br />

Schedule<br />

In vitro plants<br />

<strong>Mycosphaerella</strong> fijiensis<br />

Acclimatized 8 weeks<br />

20cm, 4 active leaves<br />

Mycelial suspension<br />

1x10 5 cfu/ml<br />

INOCULATION<br />

Under surface <strong>of</strong> first 3 open leaves<br />

Humidity 90-100% for 3 days-sunlight<br />

EVALUATION<br />

0 1 2<br />

15, 30, 45 and 60 days<br />

Stages 0-5<br />

3 4 5<br />

Figure 1. Schedule <strong>of</strong> the method for artificial inoculation <strong>of</strong> in vitro plants <strong>of</strong> Musa with mycelial suspensions<br />

<strong>of</strong> <strong>Mycosphaerella</strong> fijiensis in the greenhouse.<br />

Table 2. Reaction <strong>of</strong> seven Musa cultivars to artificial inoculation with mycelial suspensions <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

in the greenhouse.<br />

Cultivars Reaction in the field Symptom stage<br />

15 d* 30 d 45 d 60 d<br />

Grande naine Susceptible 1 1 2 4<br />

Niyarma yik Susceptible 1 1 1 4<br />

FHIA-01 Partially resistant 0 0 1 2<br />

FHIA-18 Partially resistant 0 0 1 2<br />

FHIA-21 Partially resistant 0 1 1 2<br />

Calcutta 4 Resistant 0 0 1 1<br />

Yangambi Resistant 1 2 3 3<br />

*d = days after inoculation.<br />

The partial resistance expressed by FHIA cultivars was characterized by a slow<br />

development <strong>of</strong> symptoms. Only after 60 days were reddish <strong>spot</strong>s seen on the upper<br />

surface <strong>of</strong> the leaves and the majority <strong>of</strong> leaves remained free <strong>of</strong> symptoms.<br />

Romero and Sutton (1997) reported similar results when they examined the<br />

173


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

response <strong>of</strong> ‘FHIA-01’ and ‘FHIA-02’ inoculated with conidia. They pointed out<br />

that although the mechanism <strong>of</strong> resistance to black <strong>leaf</strong> streak is not known, a<br />

low density <strong>of</strong> stomata, and increased production <strong>of</strong> cuticular wax, phytoalexin,<br />

suberin and lignin, or resistance to phytotoxins may be associated with partial<br />

resistance.<br />

‘Calcutta 4’ showed the slowest rate <strong>of</strong> symptom development. Most leaves had<br />

stage 1 symptoms although stages 2 and 3 were observed on a few leaves. ‘Grande<br />

naine’ and ‘Niyarma yik’ reacted, as expected. They were susceptible and the<br />

symptoms developed to the fullest extent. On ‘Grand naine’, symptoms were mostly<br />

at stages 4 and 5 whereas they were mostly at stages 3 and 4 on ‘Niyarma yik’. High<br />

humidity in the greenhouse caused the leaves <strong>of</strong> ‘Calcutta 4’, ‘Niyarma yik’,<br />

‘Yangambi’ and ‘FHIA-21’ to senesce rapidly.<br />

As for the in vitro plants from the IBP mutagenesis breeding programmes,<br />

differences in response were found between lines <strong>of</strong> the same genotype (‘FHIA-21’)<br />

and between ‘Grande naine’ and the control (Table 3).<br />

Table 3. Reaction <strong>of</strong> in vitro plants obtained from IBP mutagenesis breeding programmes to artificial inoculation with<br />

mycelial suspensions <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis in the greenhouse.<br />

Cultivar<br />

Symptom stage<br />

‘Grande naine’<br />

Line-GN-A1 4<br />

‘FHIA- 21’<br />

Line-F21-A 3<br />

Line-F21-B 3<br />

Line-F21-C 2<br />

Line-F21-D 1<br />

Line-F21-E 2<br />

Control<br />

‘Grande naine’ 5<br />

‘FHIA-21’ 2<br />

Conclusion<br />

Artificial inoculation <strong>of</strong> in vitro plants with mycelial suspensions was easy, rapid and<br />

practical for determining the expression <strong>of</strong> resistance <strong>of</strong> plants to M. fijiensis. Its use<br />

in breeding programs could reduce time and work.<br />

The standardized scale <strong>of</strong> symptoms was useful in order to ensure consistency<br />

between evaluators and to improve the description <strong>of</strong> symptom development. The<br />

use <strong>of</strong> inoculum with an adjusting concentration resulted in homogeneous and uniform<br />

symptoms on inoculated leaves. There was no interference from saprophytic fungi.<br />

Large quantities <strong>of</strong> mycelial fragments were produced in a short time and could solve<br />

the problem <strong>of</strong> loss <strong>of</strong> conidial production in vitro, which occurs with some isolates<br />

<strong>of</strong> M. fijiensis. An additional advantage is that mycelial fragments can be produced<br />

at different periods <strong>of</strong> the year. The method may have uses in other studies on hostpathogen<br />

interactions.<br />

174


Session 3<br />

Y. Alvarado Capó et al.<br />

References<br />

Balint-Kurti P.J., G.D. May and A. Churchill. 2001. Development <strong>of</strong> a Transformation System<br />

for <strong>Mycosphaerella</strong> Pathogens <strong>of</strong> Banana. FEMS Microbiology Letters 195:9-15.<br />

Fullerton R.A. and T.L. Olsen. 1995. Pathogenic variability in <strong>Mycosphaerella</strong> fijiensis Morelet,<br />

cause <strong>of</strong> black Sigatoka in banana and plantain. New Zealand Journal <strong>of</strong> Crop and<br />

Horticultural Science 23:39-48.<br />

Jones D.R. 1995. Rapid assessment <strong>of</strong> Musa for reaction to Sigatoka disease. Fruits 50(1):11-22.<br />

Mourichon X., D. Peter and M. Zapater. 1987. Inoculation expérimentale de <strong>Mycosphaerella</strong><br />

fijiensis Morelet sur de jeunes plantules de bananiers issues de culture in vitro. Fruits<br />

42:195-198.<br />

Mourichon X., P. Lepoivre and J. Carlier. 2000. Host-pathogens interactions. Chapter 2. Fungal<br />

disease <strong>of</strong> the foliage. Pp. 67-72 in Diseases <strong>of</strong> Banana, Abacá and Enset. (D.R. Jones,<br />

ed.). CABI publishing, Wallingford, Oxford, UK.<br />

Romero R.A. and T.B. Sutton. 1997. Reaction <strong>of</strong> four Musa genotypes at three temperatures<br />

to isolates <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis from different geographical regions. Plant Disease<br />

81:1139-1142.<br />

Stover R.H. 1987. Measuring response <strong>of</strong> Musa cultivars to Sigatoka pathogens and proposed<br />

screening procedures. Pp. 114-118 in Banana and Plantain breeding strategies. Proceedings<br />

<strong>of</strong> an international workshop, Cairns, Australia, 13-17 October 1986. (G.J. Persley and<br />

E.A. de Langhe, eds). ACIAR Proceeding No. 21.<br />

175


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Recommendations <strong>of</strong> session 3<br />

Several cases <strong>of</strong> an unexpected level <strong>of</strong> susceptibility to black <strong>leaf</strong> streak disease have been<br />

reported. Although different reasons have been <strong>of</strong>fered to explain the phenomenon (poor<br />

nutrition, environmental stress), the problem <strong>of</strong> the erosion <strong>of</strong> resistance cannot be ignored<br />

and requires a precise characterization <strong>of</strong> the pathogen population. A greater understanding<br />

<strong>of</strong> the mechanisms involved in plant-pathogen interactions continues to be needed to ensure<br />

the long term success <strong>of</strong> breeding programmes. The current programmes based on a priori<br />

hypothesis have shown their limits.<br />

Mechanisms <strong>of</strong> pathogenicity<br />

Other pathosystems (such as Magnaporthe grisea) have shown the powerful nature <strong>of</strong> the<br />

genetic approach to identify, without any a priori, the pathogenicity factors. These approaches<br />

include the study <strong>of</strong> gene expression during infection (differential display, DNA chip, SSH,<br />

etc.), production <strong>of</strong> pathogenicity mutants, comparative genomic and gene function validation<br />

techniques. A technique for the transformation <strong>of</strong> M. fijiensis is already available at the Boyce<br />

Thompson Institute (USA). The genetics and physical mapping <strong>of</strong> M. fijiensis genome in<br />

Mexico, in collaboration with PRI, Netherlands, should speed up the expected progress <strong>of</strong><br />

these experimental approaches.<br />

It was recommended to develop genetic and molecular biology tools for M. fijiensis in<br />

collaboration with groups working on M. graminicola.<br />

It was recommended launching a genomic initiative to access to genomic tools (EST<br />

collection, physical map, genome sequence) and set up a genomic-wide comparison <strong>of</strong><br />

M. fijiensis to M. graminocola<br />

Mechanisms involved in partial resistance<br />

We recommend studying differences in resistance and susceptibility in hosts which have<br />

a similar genetic background. Future work should concentrate on segregating populations<br />

to evaluate critically possible mechanisms <strong>of</strong> resistance.<br />

The genetic approach recommended to study pathogenicity factors should also be adopted<br />

to identify the mechanisms <strong>of</strong> partial resistance (patterns <strong>of</strong> genes expression during<br />

infection <strong>of</strong> resistant cultivars). These genetic studies should be accompanied by a<br />

dissection <strong>of</strong> the infection cycle to identify the components <strong>of</strong> resistance.<br />

176


Session 13<br />

Recommendations<br />

Vertical resistance<br />

It was considered that characterization <strong>of</strong> the different resistance genes represents a<br />

prerequisite to evaluate several strategies <strong>of</strong> gene deployment (pyramiding, mixture,<br />

multilines). Detection and identification <strong>of</strong> resistance genes in host plants relies on the<br />

availability <strong>of</strong> different isolates <strong>of</strong> the pathogen exhibiting different virulence phenotypes.<br />

It was recommended to collect pathogenic isolates on resistant cultivars for evaluation<br />

[genetic (to evaluate the genetic control <strong>of</strong> pathogenicity in resistant cultivars) and<br />

controlled inoculation on resistant cultivars (to evaluate the genetic control <strong>of</strong> resistance<br />

in banana)]. In this respect genetic crossings between isolates showing different behaviors<br />

on resistant cultivars are recommended to identify the genetic controls <strong>of</strong> virulence on<br />

resistant cultivars, thus leading to a better identification <strong>of</strong> genes for resistance.<br />

It was recommended to develop marker-assisted breeding (both for partial resistance<br />

and resistance genes). This will be facilitated by the activities <strong>of</strong> CIRAD on the genetic<br />

map <strong>of</strong> banana.<br />

177


Session 4<br />

Genetic improvement for the<br />

management <strong>of</strong> resistance


Session 4<br />

K. Craenen and R. Ortiz<br />

Introduction<br />

Genetic improvement for a sustainable<br />

management <strong>of</strong> resistance<br />

K. Craenen 1 and R. Ortiz 2<br />

Abstract<br />

In the 1990s, innovative cross-breeding and classic genetic analysis <strong>of</strong> segregation ratios allowed<br />

advances in the understanding <strong>of</strong> host plant response to black <strong>leaf</strong> streak disease. Partial resistance<br />

owing to a recessive major gene (bs 1<br />

) coupled with at least two additive minor genes (bsr i<br />

) appears<br />

to be durable because this genetic system slows disease development in the host plant. As a<br />

consequence, resistant hybrids show more healthy leaves, i.e. greater photosynthetic <strong>leaf</strong> area,<br />

than their susceptible full-sibs, which may partially account for their high yield. Although other<br />

breeding approaches such as genetic transformation, mutagenesis and somaclonal variation are<br />

advocated to develop new resistance to <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in Musa,farmers today<br />

are only adopting the research products from the so-called conventional breeding, i.e. tetraploid<br />

or triploid resistant hybrids from interspecific interploidy crosses. Recent findings on pathogenicity<br />

with molecular and cellular biology tools are providing new knowledge on host plant – pathogen<br />

interactions, which may result in science-led approaches for deploying resistance against<br />

sigatoka <strong>diseases</strong> within a holistic integrated disease management framework. For example,<br />

cultivar mixtures and gene pyramiding may be alternatives for potential durable resistance to<br />

sigatoka <strong>diseases</strong> <strong>of</strong> plantain and banana.<br />

Resumen - Mejoramiento genético para un manejo sostenible de la resistencia<br />

El cruzamiento innovador y el análisis genético clásico de las proporciones de segregación en las<br />

poblaciones resultantes de estos permitió avanzar en el entendimiento de la respuesta de la planta<br />

hospedante a la Sigatoka negra en Musa en la década de los 90. La resistencia parcial debido a<br />

un gen principal recesivo (bs 1<br />

) acoplado al menos a dos genes secundarios aditivos (bsr i<br />

) parece<br />

ser duradera, debido a que este sistema genético retarda el desarrollo de la enfermedad en la<br />

planta hospedante. Como consecuencia, los híbridos resistentes muestran mayor cantidad de<br />

hojas sanas, es decir, una mayor área foliar fotosintética que sus hermanos completos susceptibles,<br />

lo que puede explicar su alto rendimiento. Aunque se afirma que otros enfoques de mejoramiento<br />

como la transformación genética, mutagénesis y variación somaclonal desarrollan una nueva<br />

resistencia a las enfermedades de las manchas foliares causadas por <strong>Mycosphaerella</strong> en Musa,<br />

actualmente los agricultores están adoptando solo los productos de investigaciones del tal<br />

1<br />

Belgian Technical Cooperation, Kampala, Uganda<br />

2<br />

International Institute <strong>of</strong> Tropical Agriculture, Nigeria<br />

181


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

llamado mejoramiento “convencional”, es decir, híbridos tetraploides o triploides resistentes de<br />

los cruces interespecíficos, interploidia. Los recientes descubrimientos sobre la patogenecidad<br />

con la ayuda de las herramientas de la biología molecular y celular están proporcionando nuevos<br />

conocimientos sobre las interacciones planta hospedante-patógeno, que pueden resultar en<br />

enfoques guiados por la ciencia para desarrollar resistencia contra las enfermedades de la Sigatoka<br />

en el marco holístico de manejo integrado de la enfermedad. Por ejemplo, las mezclas de los<br />

cultivares y la construcción piramidal de genes podrían representar alternativas para una<br />

resistencia duradera potencial a las enfermedades de Sigatoka de bananos y plátanos.<br />

Résumé – L’amélioration génétique pour une gestion durable de la résistance<br />

Dans les années 1990, des croisements novateurs et l’analyse génétique classique des ratios de<br />

ségrégation ont permis de mieux comprendre la réaction des plantes-hôtes à la maladie des raies<br />

noires. Une résistance partielle due à un gène récessif principal (bs 1<br />

) couplée à au moins deux<br />

gènes mineurs additifs (bsr i<br />

) semble durable étant donné que le système génétique ralentit le<br />

développement de la maladie dans la plante-hôte. Par conséquent, les hybrides résistants ont<br />

plus de feuilles fonctionnelles, c.-à-d. une plus grande surface photosynthétique, que leurs parents<br />

susceptibles ce qui expliquerait en partie leur rendement élevé. Même si d’autres méthodes<br />

d’amélioration, comme la transformation génétique, la mutagénèse et la variation somaclonale<br />

sont prônées pour mettre au point de nouvelles résistances aux maladies foliaires causées par<br />

<strong>Mycosphaerella</strong> spp., de nos jours les fermiers adoptent seulement les produits issus des<br />

méthodes traditionnelles d’amélioration, c.-à-d. des hybrides tétraploïdes ou triploïdes résistants<br />

issus de croisements interploïdes interspécifiques. Des recherches récentes sur la pathogénicité<br />

en utilisant des outils de la biologie cellulaire et moléculaire nous renseignent sur les interactions<br />

plante-pathogène qui pourraient mener à des méthodes scientifiques pour déployer la résistance<br />

aux cercosporioses dans un cadre global de lutte intégrée aux maladies. Par exemple, l’assortiment<br />

de cultivars et le cumul des gènes (gene pyramiding) pourraient être des options pour créer une<br />

résistance durable aux cercosporioses qui affectent les bananiers.<br />

Dedication. To Dirk R. Vuylsteke (1958-2000) the ‘father’ <strong>of</strong> genetic-led Musa improvement,<br />

with whom we learned some <strong>of</strong> the issues discussed in this article. Dirk himself, his many<br />

articles and book chapters and research products ensuing from his breeding work will be<br />

always a source <strong>of</strong> inspiration to us and the new generation <strong>of</strong> plantain and banana breeders<br />

worldwide. We miss him greatly but Dirk will always live among us and those who share his<br />

humanitarian view <strong>of</strong> improving the livelihoods <strong>of</strong> the rural poor, particularly in Africa.<br />

Breeding for disease resistance<br />

Knowledge about the kind <strong>of</strong> disease, its effects and epidemics should be acquired<br />

before launching an efficient plant breeding program for disease resistance<br />

(Simmonds and Smartt, 1999). Throughout the breeding process information may<br />

be gathered on the kinds <strong>of</strong> resistance and the genetics <strong>of</strong> the host-plant reaction.<br />

At the same time research products, i.e., resistant germplasm from this approach,<br />

are made available for eco-friendly and sustainable disease control. The breeding<br />

strategy for durable host plant resistance needs to consider the populations and<br />

their genetic diversity (for both plant host and pathogen) and reliable screening<br />

methods.<br />

Interaction between the product <strong>of</strong> a resistance gene R in the plant host and the<br />

avirulence (avr) gene encoded by a given pathogen isolate, results in the specific<br />

recognition <strong>of</strong> the pathogen (Flor, 1971). If R or avr is absent, the pathogen continues<br />

182


Session 4<br />

K. Craenen and R. Ortiz<br />

colonising the plant host, reproduces and ultimately causes the disease (Holub, 2001).<br />

However, when R matches avr, the plant host recognises the pathogen, and a series<br />

<strong>of</strong> intracellular signal events occur in the plant host. The most common genetic<br />

interpretation for such an interaction claims that R products are receptors for avrencoded<br />

ligands, and that this recognition <strong>of</strong>ten leads to rapid, localized, cell death<br />

<strong>of</strong> those penetrated by the pathogen, i.e. hypersensitivity.<br />

Recent reports suggest that R genes are usually organized as clusters in plant<br />

genomes (Fluhr, 2001; Leister et al., 1998), which provides a comparative advantage<br />

for pyramiding specific resistance genes that may protect an individual plant host<br />

against many pathogen isolates (Dangl and Holub, 1997). Likewise, pyramiding will<br />

benefit the plant host because it may have a genetic reservoir from which new specific<br />

resistance may evolve. Advances in molecular breeding can assist in monitoring and<br />

accelerating the introgression <strong>of</strong> R genes into the host plant (Rommens and Kishore,<br />

2000).<br />

The crop<br />

Bananas and plantains are giant perennial herbs that thrive in the humid tropics<br />

and subtropics. The edible cultivars, in order <strong>of</strong> decreasing numerical importance,<br />

are triploid, diploid or tetraploid, and belong to the Eumusa series <strong>of</strong> the genus Musa.<br />

The warm, humid conditions required for banana and plantain also favour the<br />

development <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis, the causal agent <strong>of</strong> black <strong>leaf</strong> streak disease.<br />

All plantain cultivars and most triploid <strong>bananas</strong> are susceptible to black <strong>leaf</strong> streak<br />

disease. In the long-term, identification <strong>of</strong> resistant landraces or resistance breeding<br />

are generally considered as the most appropriate strategies to control the disease<br />

(Craenen et al., 2000).<br />

The disease and pathogen<br />

Black <strong>leaf</strong> streak disease has become a major constraint to expanding the cultivation<br />

<strong>of</strong> edible Musa. The causal pathogen <strong>of</strong> black <strong>leaf</strong> streak disease, <strong>Mycosphaerella</strong><br />

fijiensis Morelet, is a fungus that attacks the leaves. The fungal spores are<br />

disseminated by wind and infect the leaves as they unroll. The disease develops faster<br />

where humidity and rainfall are high. It has spread rapidly to all major banana and<br />

plantain growing areas and the spread is still continuing. Chemical control strategies<br />

exist, but are environmentally unsound and socio-economically inappropriate,<br />

particularly within the framework <strong>of</strong> the resource-poor smallholders that grow the<br />

crop in Africa (Craenen et al., 2000).<br />

Diversity and pathogenecity<br />

Isolates <strong>of</strong> M. fijiensis from different geographical origins were assayed with<br />

restriction fragment length polymorphisms (RFLP) (Carlier et al., 1994, 1996).<br />

Australasia and Southeast Asia isolates showed the greatest variation, suggesting the<br />

pathogen originated there. Genetically homogenous groups with low variation were<br />

observed in Africa, the Pacific Islands, and Latin America. These results indicate a<br />

183


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

few introductions <strong>of</strong> small size in each <strong>of</strong> these regions from Southeast Asia (Carlier<br />

et al., 1994). Genetic distances between regions were high but the isolates from the<br />

Pacific Islands and Latin America seem related, suggesting that the introductions to<br />

Latin America came from the Pacific Islands. Founder effects accompanied the<br />

introductions to Latin America, some Pacific Islands and Africa (Carlier et al., 1999),<br />

thereby reducing genetic diversity in these regions.<br />

Recent analysis with microsatellite polymorphisms revealed variation in the<br />

fingerprint patterns <strong>of</strong> M. fijiensis populations from Nigeria (Muller et al., 1995, 1997).<br />

Microsatellite (oligonucleotide) fingerprinting appears to be a reliable technique for<br />

assessing genetic variation among individuals as well as for defining clusters <strong>of</strong> related<br />

genotypes, i.e. detecting intraspecific variation even on a microgeographical scale.<br />

Fullerton and Olsen (1993) evaluated the pathogenic diversity within populations<br />

<strong>of</strong> M. fijiensis. They used as differential hosts most <strong>of</strong> the standard cultivars currently<br />

being used in the International Musa Testing Programme. A wide host response to the<br />

whole range <strong>of</strong> strains was reported. The most susceptible across all isolates were ‘Grande<br />

naine’ and ‘SF 215’, while ‘Calcutta 4’, the widest source <strong>of</strong> alleles resistant to black<br />

<strong>leaf</strong> streak disease, was susceptible to some strains, particularly those collected in the<br />

Pacific Islands and Papua New Guinea. However, the host reaction to the pathogen must<br />

be tested with adult plants and, as their results suggest, for probably more than one<br />

year in order to detect strains present at low frequency. Perhaps, resistant germplasm<br />

needs to be tested over several years in locations with very virulent strains in order to<br />

evaluate the durability <strong>of</strong> resistance.<br />

The data in Fullerton and Olsen (1993) were re-analysed using simple linear regression<br />

models to determine the stability <strong>of</strong> resistance to black <strong>leaf</strong> streak disease in Musa<br />

acuminata, and using principal component analysis to study the pattern <strong>of</strong> strain and<br />

genotype variation in the pathogen-host plant interaction (Ortiz et al., 2000). ‘Tuu Gia’<br />

was regarded as having non-specific resistance to the 33 strains <strong>of</strong> M. fijiensis used in<br />

the experiment. Furthermore, it seems that its resistance does not break down even under<br />

the pressure <strong>of</strong> high virulent strains. In contrast, the resistance <strong>of</strong> ‘Calcutta 4’ was very<br />

unstable and may break down in environments where highly virulent strains have<br />

evolved.<br />

Principal component analysis revealed that isolates <strong>of</strong> M. fijiensis from Papua New<br />

Guinea were the most virulent, and that ‘Calcutta 4’ accounted for most <strong>of</strong> the genotype<br />

x strain interaction. Strains collected from the same country could be clustered together<br />

(e.g. Nigeria or most <strong>of</strong> the Pacific Islands’ isolates), or had a continuous virulence<br />

distribution (e.g. Papua New Guinea), or were completely distinct (e.g. Central America).<br />

The latter suggests a different geographical origin for the introduced M. fijiensis or a<br />

change in virulence genes in the strains from Central America as a result <strong>of</strong> the intensive<br />

use <strong>of</strong> fungicides to control black <strong>leaf</strong> streak disease.<br />

Incidence and severity <strong>of</strong> black <strong>leaf</strong> streak disease<br />

Quantification <strong>of</strong> black <strong>leaf</strong> streak disease is necessary to evaluate resistance and to<br />

determine yield loss, the importance <strong>of</strong> black <strong>leaf</strong> streak disease in particular areas,<br />

and the efficacy <strong>of</strong> control measures (Craenen, 1998). Fouré (1985) described in detail<br />

the experimental methods to characterize the different host responses. A high level <strong>of</strong><br />

184


Session 4<br />

K. Craenen and R. Ortiz<br />

resistance to black <strong>leaf</strong> streak disease is characterized by hypersensitivity. Different<br />

levels <strong>of</strong> partial resistance, ranging from strong partial resistance to susceptibility have<br />

been observed in Musa germplasm (Fouré, 1994).<br />

Damage resulting from disease can be evaluated accurately only by measuring its<br />

incidence and severity (Gauhl et al., 1994; Jones, 1994). Severity <strong>of</strong> the disease relates<br />

to the intensity <strong>of</strong> damage to individual plants, while incidence deals with the<br />

percentage <strong>of</strong> plants affected in a population.<br />

The incidence and severity <strong>of</strong> black <strong>leaf</strong> streak disease on Musa can be assessed<br />

in the laboratory or in the field. The laboratory method involves determining<br />

ascospore and conidia production with the aid <strong>of</strong> a microscope, but this method is<br />

tedious, very time-consuming and not very accurate.<br />

Field evaluation under conditions <strong>of</strong> natural infection is the most common and<br />

preferred method to assess incidence and severity <strong>of</strong> black <strong>leaf</strong> streak disease (Craenen,<br />

1997). The nature and amount <strong>of</strong> lesions and the rate <strong>of</strong> their development on the<br />

leaves is observed in the field. This method does not require a full understanding <strong>of</strong><br />

host-pathogen interactions, nor plant population systems and is therefore suitable for<br />

field workers trained in symptom recognition.<br />

Although field-screening methods are relatively simple, they are also timeconsuming<br />

and influenced by environmental factors, such as weather and soil, which<br />

affect symptom expression. Therefore, it is recommended to compare the test plants<br />

with reference cultivars and to gather a large number <strong>of</strong> observations to validate the<br />

results.<br />

Laboratory evaluation<br />

The production <strong>of</strong> ascospores can be estimated by taking <strong>leaf</strong> samples with pseudothecia<br />

from the same plant at different times, as described in Stover (1976). For conidiophores,<br />

leaves at stage 2 <strong>of</strong> symptom development are collected in the field (Fouré, 1982).<br />

However, results can vary widely from <strong>leaf</strong> to <strong>leaf</strong>; <strong>spot</strong>s on each <strong>leaf</strong> can differ<br />

enormously in number <strong>of</strong> spore-producing organs, and subjective errors are the source<br />

<strong>of</strong> wrong results. Not being very valuable, this method is not described further.<br />

An inoculation technique using <strong>leaf</strong> pieces under controlled conditions has proven<br />

to be very simple and has many advantages (El Hadrami et al., 1998a, b). This technique<br />

enables the production <strong>of</strong> the sexual and asexual phases <strong>of</strong> M. fijiensis and is very<br />

useful to study host-parasite interactions (El Hadrami et al., 2000). The infection patterns<br />

and symptom evolution were the same as those observed in the field, and the <strong>leaf</strong><br />

piece assay allowed the expression <strong>of</strong> resistant phenotypes. Furthermore, this method<br />

may allow the epidemiological appraisal <strong>of</strong> partial resistance and the variability <strong>of</strong><br />

virulence in M. fijiensis.<br />

Field evaluation<br />

Field evaluation <strong>of</strong> disease severity requires knowledge not only <strong>of</strong> the stages <strong>of</strong> symptom<br />

development and the percentage <strong>of</strong> <strong>leaf</strong> area <strong>spot</strong>ted, but also <strong>of</strong> the different stages<br />

<strong>of</strong> unrolling <strong>of</strong> the <strong>leaf</strong> (Craenen, 2001). Disease severity is evaluated by recording the<br />

percentage <strong>of</strong> the <strong>leaf</strong> area that is <strong>spot</strong>ted using a 7-category scale (modified from Stover<br />

185


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

and Dickson, 1970), which ranges from 0 (no symptoms) to 6 (51-100% <strong>leaf</strong> area with<br />

symptoms) (Craenen 1997, 1998).<br />

Epidemiology<br />

In order to gain insight into the epidemiology <strong>of</strong> the disease, it is necessary to quantify<br />

the amount <strong>of</strong> inoculum in an area. At the IITA High Rainfall Station in Onne, Nigeria,<br />

spore-trapping and weather data were recorded for three years. This research indicated<br />

that spore concentrations in the air were lower during the dry season and higher in the<br />

rainy season. Ascospores were more frequent than conidiospores. During the dry season,<br />

ascospore concentration was only three times as high as conidiospore concentration,<br />

whereas during the rainy season ascospore concentration was found to be 40 times<br />

higher (IITA, 1995).<br />

Measuring the different characteristics <strong>of</strong> disease development on parents and their<br />

hybrids in epidemiological studies led to the identification <strong>of</strong> different types <strong>of</strong> host<br />

response and to the classification <strong>of</strong> Musa germplasm into different categories<br />

according to the response to the disease. The different characteristics to be recorded<br />

assume that M. fijiensis infects first the unfolded (> 10 cm) cigar <strong>leaf</strong> <strong>of</strong> the host plant<br />

(Fullerton, 1994; Jones,1994). These characteristics are:<br />

• Incubation time: calculated as the number <strong>of</strong> days between cigar <strong>leaf</strong> emergence<br />

(Brun’s stage 2) and the appearance <strong>of</strong> the initial chlorotic fleck symptoms (i.e. depigmentation<br />

<strong>spot</strong>) relating to symptom stage 1 <strong>of</strong> Fouré’s scale. Brun’s stage 2 refers<br />

to an upright cigar <strong>leaf</strong>, still strongly rolled and free from the petiole <strong>of</strong> the preceding<br />

<strong>leaf</strong>, but not reaching its full length. At the first stage <strong>of</strong> symptom development only<br />

minute yellowish specks (< 1 mm in length) are seen on the lower (abaxial) surface<br />

<strong>of</strong> the <strong>leaf</strong>. They are not visible in translucent light.<br />

• Evolution time: calculated as the number <strong>of</strong> days between first symptoms (Fouré’s<br />

stage 1) and the occurrence <strong>of</strong> mature lesions (Fouré’s stage 6). At this last stage <strong>of</strong><br />

symptom development, the centre <strong>of</strong> the <strong>spot</strong> dries out and fades into a clear gray.<br />

Often a black ring, surrounded by a yellow halo, encircles the gray centre. Necrotic<br />

<strong>spot</strong>s remain visible after the <strong>leaf</strong> has dried up completely.<br />

• Disease development time: defined as the number <strong>of</strong> days elapsing between Brun’s<br />

stage 2 <strong>of</strong> <strong>leaf</strong> emergence and Fouré’s stage 6 <strong>of</strong> symptoms, i.e. incubation time plus<br />

evolution time.<br />

• Lifetime <strong>of</strong> the <strong>leaf</strong>: recorded as the number <strong>of</strong> days between <strong>leaf</strong> emergence (Brun’s<br />

stage 2) and <strong>leaf</strong> death.<br />

•Youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS) at flowering: recorded as the first <strong>leaf</strong> (counting<br />

downwards from the first top unfurled <strong>leaf</strong>) that shows <strong>spot</strong>s (equal to or more than<br />

10) with a necrotic dry center (Vakili, 1968).<br />

The higher the YLS the more fully functional leaves on the plant, and hence, greater<br />

resistance to the fungus (Craenen and Ortiz, 1997). The YLS score correlates significantly<br />

with disease development time and other parameters to assess host plant response to<br />

black <strong>leaf</strong> streak disease (Craenen, 1994). The YLS is also very easy to score, heritable,<br />

and after scoring the level <strong>of</strong> host response can be defined, thereby allowing grouping<br />

186


Session 4<br />

K. Craenen and R. Ortiz<br />

<strong>of</strong> genotypes. Four distinct levels <strong>of</strong> host response to black <strong>leaf</strong> streak disease were defined<br />

by Ortiz and Vuylsteke (1994) as follows:<br />

susceptible (< 8 leaves without <strong>spot</strong>s before or at flowering),<br />

less susceptible (8–10),<br />

partially resistant (> 10)<br />

and highly resistant (none)<br />

These epidemiological characteristics do not only depend on the amount <strong>of</strong><br />

inoculum present, but also on climatological factors that may affect the development<br />

<strong>of</strong> the disease. Hence, it is essential to monitor environmental factors. Daily readings<br />

from a weather station near to the experimental site are required. If there is no weather<br />

station available, daily readings can be taken (if possible early in the morning) for the<br />

following factors:<br />

– rainfall (with a simple rain gauge);<br />

–minimum and maximum temperature (with a minimum/maximum thermometer);<br />

– relative humidity (with a hygrometer).<br />

Temperature and relative humidity can also be recorded continuously with a<br />

mechanical hygrothermograph.<br />

Variability may be introduced in data sets from different locations due to<br />

environmental or genetic causes (Ortiz et al., 1993). To minimise these effects the host<br />

response was re-defined as the ratio between the YLS and the total number <strong>of</strong> standing<br />

leaves (NSL) at the time <strong>of</strong> scoring. The index <strong>of</strong> non-<strong>spot</strong>ted leaves (INSL) to assess<br />

the host response to black <strong>leaf</strong> streak disease is calculated as follows:<br />

INSL = 100 – 100 x [(NSL – YLS + 1)/NSL]<br />

The INSL is the proportion <strong>of</strong> standing leaves without typical late stage symptoms<br />

<strong>of</strong> the disease (i.e. <strong>spot</strong>s with a necrotic centre). This index provides an estimation <strong>of</strong><br />

the available photosynthetic <strong>leaf</strong> area prior to fruit filling.<br />

Early screening<br />

Breeding resistance requires methods able to discriminate resistant and susceptible<br />

genotypes at different stages <strong>of</strong> plant development (Leproive et al., 1993). In vitro<br />

selection or field assessment using young plant materials are among some <strong>of</strong> the early<br />

screening methods.<br />

Inoculating Musa <strong>leaf</strong> tissue with a crude extract <strong>of</strong> M. fijiensis was suggested as<br />

an early screening method (Hernández 1995). The crude extract screening method allows<br />

a rapid (48 hours in greenhouse plants and 72 hours in callus tissue) identification <strong>of</strong><br />

host-plant resistance. For example, <strong>leaf</strong> tissue <strong>of</strong> susceptible cultivars such as ‘Grande<br />

naine’ or ‘Currare’ showed the highest levels <strong>of</strong> phenolic compounds while resistant<br />

germplasm (e.g. ‘Yangambi km5’, ‘Calcutta 4’ or ‘Saba’) showed low phenol content,<br />

after being inoculated with a crude extract <strong>of</strong> M. fijiensis.<br />

Pino (1997) indicated that 120 hours after inoculating in vitro plants with M. fijiensis<br />

toxins, lesions in susceptible banana cultivar ‘Grande naine’ were larger than in putative<br />

resistant mutants <strong>of</strong> ‘Grande naine’ (24.12 – 49.22 mm 2 vs. 1.07 mm – 4.07 mm 2 ,<br />

respectively). Similarly, Okole and Schulz (1997) reported as promising an in vitro<br />

selection technique using microsections (or callus cultures) <strong>of</strong> banana and plantain using<br />

187


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

a double selection system. The selection system uses in the first stage raw filtrate<br />

concentrations (10-100 mg/L) <strong>of</strong> the fungus, and then a specific purified fungal toxin<br />

(2,4,8-trihydroxyltetralone or 2,4,8-tht) isolated from M. fijiensis. This toxin plays an<br />

important role in the development <strong>of</strong> necrotic <strong>leaf</strong> symptoms that causes host-specific<br />

reactions depending on their concentration at different pathogenesis stages in tissue<br />

culture materials (Hoss, 1998). A Musa genotype resistant to black <strong>leaf</strong> streak disease<br />

with increased 2,4,8-tht content, produced the hypersensitive reaction and elicited<br />

postinfectional defense reactions in the host plant, which led to its incompatibility with<br />

the pathogen. A susceptible Musa genotype showed toxic doses <strong>of</strong> 2,4,8-tht but only<br />

after the establishment <strong>of</strong> a compatible interaction, which first helped the biotrophic<br />

nutrition <strong>of</strong> the pathogen, and acted as a virulence factor at the necrotrophic phase <strong>of</strong><br />

pathogenesis.<br />

Toxin-resistant plantlets <strong>of</strong> two Cavendish cultivars (‘Williams’ and ‘Petite naine’)<br />

and <strong>of</strong> ‘Horn plantain’ were regenerated using the above-mentioned method (Okole et<br />

al., 2000). Rooted plants were further transferred to soil infected with suspensions <strong>of</strong><br />

M. fijiensis spores (0.3 g/ml). About 11 to 19% <strong>of</strong> the plantlets resistant to the toxin<br />

were resistant to M. fijiensis in this culture chamber test, which reproduces the symptoms<br />

<strong>of</strong> black <strong>leaf</strong> streak disease. However, the plants that withstood the toxin injection to<br />

their tissues and the double selection procedure have not yet been field-tested.<br />

As pointed out by Harelimana et al. (1996), screening with toxins to select resistant<br />

germplasm has two major limitations. First, the lack <strong>of</strong> experimental evidence on the<br />

role <strong>of</strong> toxins in disease development, and second, the susceptibility or resistance <strong>of</strong><br />

the cultured tissues do not reflect those <strong>of</strong> the adult plant because <strong>of</strong> the mode <strong>of</strong> action<br />

<strong>of</strong> the toxin. It has not been demonstrated that toxins <strong>of</strong> M. fijiensis participate in the<br />

initiation <strong>of</strong> infection or in the hypersensitive reaction <strong>of</strong> highly resistant adult plants.<br />

Nonetheless, toxins could play a secondary role in pathogenicity, e.g. in disease<br />

development in partially resistant cultivars. Research showed that chloroplasts could<br />

be a precocious site <strong>of</strong> action <strong>of</strong> the toxins, suggesting that in vitro heterotrophic Musa<br />

tissues may not be suitable for early screening.<br />

Another early screening method for host response consists in using natural<br />

inoculum on young Musa plants (3-month-old micropropagated plants). This method<br />

confirmed the resistance <strong>of</strong> plantain hybrids and the susceptibility <strong>of</strong> their female plantain<br />

parent to this disease (Mobambo et al., 1994). Furthermore, the characteristics associated<br />

with disease development in these young plants were similar to those observed on adult<br />

plants <strong>of</strong> the same genotypes (Mobambo et al., 1997). This method was also faster,<br />

cheaper, less labour intensive and required less field space than screening <strong>of</strong> adult plants.<br />

Highly susceptible germplasm can be rejected at a precocious stage with this early<br />

screening method, thus reducing the sample size (and associated) costs for testing adult<br />

plants in the field.<br />

Host plant response to black <strong>leaf</strong> streak disease<br />

Genetics <strong>of</strong> resistance<br />

In plantains and <strong>bananas</strong>, resistance to black <strong>leaf</strong> streak disease is genetically<br />

controlled. Genetic analysis has been carried out on diploid and tetraploid progenies<br />

188


Session 4<br />

K. Craenen and R. Ortiz<br />

obtained from triploid (plantain) x diploid crosses (Vuylsteke et al., 1993). The diploid<br />

male parent was the resistant true-breeding line ‘Calcutta 4’ (Ortiz et al., 1998b).<br />

Thus, the populations produced can be considered to be genetically equivalent to<br />

test-crosses for the host response to this disease. Resistance to black <strong>leaf</strong> streak disease<br />

is mainly the result <strong>of</strong> the interaction <strong>of</strong> three independent alleles: a recessive allele<br />

at a major locus (bs1) and the alleles <strong>of</strong> at least two independent minor, modifying<br />

genes with additive effects (bsri) (Ortiz and Vuylsteke, 1994). These genes have a<br />

strong dosage effect at the tetraploid level that results in higher levels <strong>of</strong> resistance<br />

in tetraploid than in diploid hybrids. Resistance genes are present in the genome <strong>of</strong><br />

susceptible plantains, but their expression is masked by the dominant effect <strong>of</strong> the<br />

major gene for susceptibility (Ortiz and Vuylsteke, 1994).<br />

These results were further confirmed by investigating tetrasomic segregration in<br />

a cross between a resistant and a susceptible tetraploid hybrid (Ortiz, 2000). The<br />

resistant maternal genotype was a nulliplex for the major resistance locus and the<br />

paternal susceptible genotype was a duplex for the corresponding host response. Both<br />

parents were balanced diallelic for the two minor modifier loci with additive effects.<br />

The segregating tetraploid population from this cross showed a tri-modal frequency<br />

distribution in the population, which was not significantly different to the expected<br />

ratio (1.7 resistant : 1 susceptible) from the early genetic model defined by Ortiz<br />

and Vuylsteke (1994).<br />

Using the gene-for-gene hypothesis (Flor, 1971), a host-plant resistance system<br />

based on recessive alleles is difficult to overcome by the pathogen as this requires<br />

a mutation to the dominant allele <strong>of</strong> the virulence locus (Ortiz and Vuylsteke, 1994).<br />

Since such mutations are rare (Simmonds, 1979), resistance based on recessive alleles<br />

may prove to be durable.<br />

Resistance genotypes are expressed phenotypically in the plants. Highly resistant<br />

plants exhibit the longest incubation time and <strong>leaf</strong> life span as well as a hypersensitive<br />

reaction to black <strong>leaf</strong> streak disease (Craenen and Ortiz, 1998). This extremely resistant<br />

polygenic response (Ortiz and Vuylsteke, 1994) blocks disease development at an<br />

early stage, thereby impeding the occurrence <strong>of</strong> mature necrotic lesions in the leaves.<br />

In contrast, susceptible cultivars have a short incubation time, evolution time and<br />

disease development time. This indicates that after infection, disease symptoms evolve<br />

quickly into necrotic <strong>spot</strong>s, resulting in extensive <strong>leaf</strong> death and defoliation.<br />

Results from multilocational trials in Africa showed that all partially resistant<br />

hybrids had a homeostatic host response to black <strong>leaf</strong> streak disease (Ortiz et al.,<br />

1997). Also some <strong>of</strong> them achieved high and stable yields across environments due<br />

to their resistance (Ortiz and Vuylsteke, 1995), even under low organic matter inputs<br />

(Ortiz et al., 1995).<br />

Mechanisms <strong>of</strong> resistance<br />

A sample <strong>of</strong> 20 euploid Musa hybrids <strong>of</strong> various ploidy, exhibiting a range <strong>of</strong> resistant<br />

and susceptible responses, were used to investigate the role <strong>of</strong> stomatal density, stomatal<br />

length and the thickness <strong>of</strong> epicuticular wax in resistance to M. fijiensis (Craenen et<br />

al., 1997). The female parents <strong>of</strong> these hybrids were susceptible plantains, while the<br />

male parent was a wild, non-edible resistant banana (‘Calcutta 4’). Stomatal length was<br />

189


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

negatively correlated with the initial development (incubation time) <strong>of</strong> black <strong>leaf</strong> streak<br />

disease in the leaves <strong>of</strong> young diploids but not in those <strong>of</strong> polyploid hybrids. Stomatal<br />

density on the abaxial surface <strong>of</strong> young leaves was negatively correlated with<br />

incubation time only in polyploids. Incubation time was positively correlated with the<br />

accumulation <strong>of</strong> epicuticular wax in both diploid and polyploid hybrids. Although the<br />

resistant male parent lacked epicuticular wax, derived hybrids possessed epicuticular<br />

wax <strong>of</strong> various thickness which enhanced resistance. Hence, the two minor additive<br />

modifier genes (bsr i ) which enhanced resistance may control decreased stomatal density<br />

and increased <strong>leaf</strong> waxiness. Both characteristics may be two resistance mechanisms<br />

that lengthen the incubation time <strong>of</strong> the disease in the leaves.<br />

Craenen and Ortiz (1997) determined the role <strong>of</strong> the major gene for resistance (bs 1 )<br />

in a sample <strong>of</strong> euploid hybrids from triploid-diploid crosses <strong>of</strong> two French plantains<br />

and a diploid wild banana, and with a known genotype for the bs 1 locus. Their host<br />

response was assessed in the humid forest zone <strong>of</strong> Nigeria. Analysis <strong>of</strong> frequency<br />

distribution in each segregating population showed that almost all the traits displayed<br />

a normal distribution across ploidy level. This suggests that additive gene action plays<br />

an important role in the host plant response to the fungus. However, the environment<br />

and the genotype x environment interaction significantly affected the host response,<br />

which explains the low reproducibility <strong>of</strong> all traits. Intrafamily variation was larger than<br />

interfamily variation, and most <strong>of</strong> the genetic variation in each family depended on<br />

individual genotypes, regardless <strong>of</strong> their ploidy. The additive effect <strong>of</strong>, and the<br />

intralocus interaction at, the bs 1 locus were established by one-way analysis <strong>of</strong> variance<br />

and regression analysis. Intralocus interaction at the bs 1 locus apparently regulates the<br />

appearance <strong>of</strong> symptoms on the <strong>leaf</strong> surface, whereas the additive effect and the<br />

intralocus interaction <strong>of</strong> the bs 1 locus affect disease development in the host plant.<br />

Therefore, the gene action(s) at the bs 1 locus may provide durable resistance by slowing<br />

down disease development.<br />

Effect <strong>of</strong> black <strong>leaf</strong> streak disease on agronomic traits<br />

Yield loss due to black <strong>leaf</strong> streak disease is 33 to 50% in plantain (Stover, 1983;<br />

Mobambo et al., 1993) as a result <strong>of</strong> a reduced number <strong>of</strong> fruits per bunch and a lower<br />

fruit weight. Black <strong>leaf</strong> streak disease has no effect on plant height and suckering, but<br />

delays flowering and harvest by more than one month. The disease also causes premature<br />

fruit ripening (Stover, 1980; Mobambo et al., 1993). In plantain landraces, normal fruit<br />

filling or ripening time, i.e. time from flowering to harvest, is 91 days. Black <strong>leaf</strong> streak<br />

disease significantly reduces this time. This premature fruit ripening is expected to have<br />

adverse effects on postharvest characteristics, such as a reduced shelf life. In addition,<br />

fruits are shorter and thinner, which generally results in lower quality fruit and lower<br />

market value. Black <strong>leaf</strong> streak disease thus has a negative impact on fruit bulking,<br />

probably as the result <strong>of</strong> a reduction in healthy <strong>leaf</strong> area (Mobambo et al., 1993).<br />

Fruits <strong>of</strong> susceptible plantain hybrids are unable to bulk fully. Less susceptible<br />

and partially resistant plantains are bigger, longer and heavier. Resistant plants<br />

have increased bunch weight due to complete fruit filling as they have more<br />

functional leaves for photosynthesis during the period between flowering and<br />

harvest (Craenen and Ortiz, 1998).<br />

190


Session 4<br />

K. Craenen and R. Ortiz<br />

Genetic analysis <strong>of</strong> the plantain genome is difficult due to triploidy and high sterility.<br />

As shown earlier, ploidy manipulations (scaling up and down the number <strong>of</strong><br />

chromosomes) and interspecific plantain-banana hybridization opened the path for the<br />

genetic amelioration <strong>of</strong> the crop and for the investigation <strong>of</strong> its genome. There are several<br />

associated effects <strong>of</strong> ploidy, parthenocarpy and resistance to black <strong>leaf</strong> streak disease<br />

on growth and yield characteristics <strong>of</strong> euploid hybrids. The number <strong>of</strong> copies <strong>of</strong> the<br />

resistance allele (bs 1 ) and <strong>of</strong> the parthenocarpy gene (P 1 ), as well as their intralocus<br />

interaction and ploidy level, have all been found to significantly affect bunch and fruit<br />

characteristics <strong>of</strong> euploid hybrids (Ortiz et al., 1998a). Epistasis significantly affected<br />

fruit weight and size in one cross but not in another. Significant multiple regression<br />

models combining ploidy and genetic markers explained 15% to 85% <strong>of</strong> quantitative<br />

trait variation (QTV). The amount <strong>of</strong> QTV accounted by ploidy and genetic markers varied<br />

according to the characteristic and cross in which the markers were examined.<br />

Linear and multiple regression models, coefficients <strong>of</strong> determination, and Durbin-<br />

Watson statistics were used by Craenen and Ortiz (1996) to determine the single and<br />

combined effects <strong>of</strong> the major locus for resistance to black <strong>leaf</strong> streak disease (bs 1 )<br />

and <strong>of</strong> ploidy on bunch weight, fruit weight, fruit length and fruit girth in the<br />

progenies derived from crosses between a resistant diploid wild banana source and<br />

the susceptible French plantain landraces. Differences in yield were mainly due to<br />

changes in weight and circumference <strong>of</strong> the fruit, which are affected by the disease.<br />

The combined effect <strong>of</strong> ploidy and resistance to black <strong>leaf</strong> streak disease was partially<br />

responsible for QTV in yield. As a result <strong>of</strong> the gene interaction in the locus for<br />

resistance (bs 1 ), the partially resistant phenotypes showed higher yield than their<br />

more susceptible full sibs.<br />

The performance <strong>of</strong> 20 euploid hybrids was compared with that <strong>of</strong> their parents<br />

to determine the influence <strong>of</strong> the disease on growth parameters and components <strong>of</strong><br />

yield (Craenen and Ortiz, 1998). There were significant differences among the hybrids<br />

for all components <strong>of</strong> resistance, growth parameters, and yield components. For<br />

diploid hybrids, which <strong>of</strong>ten had a short growth cycle or early flowering, or both,<br />

the disease incubation time was significantly correlated with days to fruit filling<br />

(P < 0.05). However, for tetraploid hybrids that had a long growth cycle and delayed<br />

flowering, the correlation was not significant (P > 0.05). For diploid and tetraploid<br />

hybrids, disease evolution time and disease development time were both correlated<br />

(P < 0.05) with days to fruit filling. Bunch weight <strong>of</strong> tetraploid hybrids was correlated<br />

(P < 0.05) with disease development time as scored by the youngest <strong>leaf</strong> <strong>spot</strong>ted at<br />

flowering (r = 0.933; P < 0.001). This result confirms that resistant hybrids with<br />

potentially high yield could be selected efficiently by recording the youngest <strong>leaf</strong><br />

<strong>spot</strong>ted at flowering.<br />

Diploid or polyploid breeding<br />

Most banana and all plantain cultivars grown by farmers in the tropics are<br />

triploids. Euploid hybrids derived from triploid-diploid crosses are either mostly<br />

diploids but some are tetraploids. They are seldom triploids. The most important<br />

goal in genetic population improvement programs for managing disease resistance<br />

in large populations is to enhance the frequency <strong>of</strong> favourable alleles controlling<br />

191


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

the desired characteristics. Recurrent selection methods are the most common for<br />

increasing the frequency <strong>of</strong> favourable alleles in a cyclic fashion.<br />

Theoretically, sq 2 recessive individuals will be selected out from a large diploid<br />

population where q is the allele frequency <strong>of</strong> a recessive gene and s is the intensity<br />

<strong>of</strong> selection against the recessive genotype. Similarly, in a large tetraploid population,<br />

sq 4 represents the proportion <strong>of</strong> recessive genotypes that are selected out. Hence,<br />

the change <strong>of</strong> allele frequency appears to be faster at the diploid than at the tetraploid<br />

level. This occurs because the recessive allele is only hidden in the heterozygous<br />

genotype at the diploid level while it will be included in the triplex, duplex and<br />

simplex at the tetraploid level. For example, if q = 0.5 and selection efficiency s =1<br />

at both ploidy levels, then the change in allele frequency after selection will be four<br />

times larger in the diploids than that for the tetraploids (25% vs 6.25%, respectively).<br />

Population genetics theory also suggest that the smaller the s, the lower response<br />

to selection. When there are no escapes during screening, i.e. maximum selection<br />

efficiency, increases in the frequency <strong>of</strong> favourable allele will be maximum. With<br />

increased rates <strong>of</strong> escapes, breeding becomes inefficient and to the point that it may<br />

become worthless. As pointed out by Mendoza (1988) “the degree <strong>of</strong> success in altering<br />

the genotypic structure <strong>of</strong> the population, by modifying its gene frequency, is a function<br />

<strong>of</strong> the precision in identifying and isolating the individuals carrying the attributes<br />

under selection. Any errors or ‘escapes’ during the process, depending on their<br />

magnitude, could alter the response to selection… A breeding effort can only be as<br />

efficient as the screening procedure permits.”<br />

Selection appears to be more effective in the early cycles when the frequency <strong>of</strong><br />

the favourable allele is low, especially at the diploid level. However, if there are<br />

escapes owing to unreliable screening methods, particularly for a small population<br />

size, then a lowering <strong>of</strong> the response to selection will occur at a very low frequency<br />

<strong>of</strong> the favourable allele. With tetraploids, when the frequency <strong>of</strong> the favourable allele<br />

exceeds 0.4, the response to selection falls rapidly. When the frequency reaches 0.8<br />

in the tetraploid population, the response to selection becomes practically nil for<br />

breeding purposes.<br />

Outlook<br />

The results <strong>of</strong> this research support the early views <strong>of</strong> Musa breeders (Ortiz, 1997;<br />

Vuylsteke 2000), who claimed that a broad-based, improved Musa germplasm with<br />

pest and disease resistance was necessary to achieve the sustainable production <strong>of</strong><br />

this vegetatively propagated perennial crop. This germplasm was obtained by using<br />

conventional crossbreeding and may be further enhanced with the utilization <strong>of</strong><br />

innovative methods for the introduction <strong>of</strong> additional genetic variation, e.g. ideotype<br />

breeding, polycross mating design or marker-aided introgression (Ortiz, 2001). In<br />

short, “the prospects <strong>of</strong> banana and plantain breeding are unlimited and increased<br />

efforts will at once initiate a new phase <strong>of</strong> Musa evolution” (Vuylsteke, 2001).<br />

Manipulation <strong>of</strong> the Musa genome for its genetic betterment will be also facilitated<br />

by the available knowledge on the inheritance <strong>of</strong> most important characteristics in<br />

plantain and banana (Ortiz, 2000). Likewise, the information regarding fungal<br />

<strong>diseases</strong>, such as black <strong>leaf</strong> streak disease, and the interactions between the<br />

192


Session 4<br />

K. Craenen and R. Ortiz<br />

pathogen and its host plant, provide the basis for a rational integrated management<br />

strategy to control the disease. For example, the partial resistance provided by the<br />

bs 1<br />

gene (Craenen and Ortiz, 1997; Ortiz and Vuylsteke, 1994) can be easily<br />

incorporated into mixed cultivar systems common among the resource-poor farmers<br />

in the tropics (IITA, 1998). These farmers prefer cropping systems that provide<br />

intraspecific (cultivar mixtures) and interspecific (inter-cropping) diversity to<br />

maximise land, use labour efficiently and minimise the risk <strong>of</strong> crop failure.<br />

Deploying resistant hybrids in farmers’ cropping systems in association with their<br />

own landraces is regarded as non-disruptive (IITA, 1999). In this suggested cultivar<br />

mixture, the resistant hybrids serve as inoculum traps that reduce the spread <strong>of</strong> the<br />

disease to the susceptible plantain landraces and may increase the bunch weight <strong>of</strong><br />

the landraces that are preferred by farmers due to their culinary and rheological<br />

characteristics (IITA, 2000). On-farm participatory research undertaken by IITA will<br />

provide more insights into this proposal for deployment <strong>of</strong> resistance to black <strong>leaf</strong><br />

streak disease using a cultivar mixture system (IITA, 2001). Data are still being<br />

recorded in a farm in south-eastern Nigeria.<br />

A cultivar mixture system would preserve genetic diversity and provide new, highyielding<br />

hybrids that may be incorporated in the local diet through novel processing<br />

methods. Introducing new cultivars may lead to losses <strong>of</strong> diversity in farmer’s fields<br />

(Sharrock et al., 2000), particularly when single-cultivar plantations are preferred<br />

over mixed farming. Furthermore, <strong>diseases</strong>, e.g. black <strong>leaf</strong> streak disease, may spread<br />

quickly into single-cultivar plantations <strong>of</strong> susceptible germplasm or when resistance<br />

breaks down in improved germplasm. Hence, cultivar mixtures may provide an<br />

“insurance” for a sustainable farming system.<br />

Pyramiding genes from distinct germplasm sources may also enhance partial<br />

resistance in plantain and banana. For example, IITA hybrids, which show this kind<br />

<strong>of</strong> resistance, have alleles for resistance to black <strong>leaf</strong> streak disease from two sources:<br />

triploid plantains and diploid <strong>bananas</strong>. The resistance alleles are masked by intra<br />

and interlocus interactions in highly susceptible plantain parent landraces. The<br />

resistance alleles are mainly from the wild banana accession ‘Calcutta 4’ or the diploid<br />

banana cultivar ‘Pisang lilin’ (Hartman and Vuylsteke, 1999). The search for other<br />

sources <strong>of</strong> resistance against a wide range <strong>of</strong> strains (e.g. Fullerton and Olsen, 1991,<br />

1993) appears mandatory to develop a strategy for durable resistance to black <strong>leaf</strong><br />

streak disease. This requires urgent attention because resistance in at least two<br />

cultivars (‘Paka’ and ‘T8’) broke down after eight years <strong>of</strong> cultivation in the Cook<br />

Islands (Hartman and Vuylsteke, 1999).<br />

The incorporation <strong>of</strong> different resistance types in the same genotype could<br />

potentially confer durable resistance to black <strong>leaf</strong> streak disease. Indeed, resistance<br />

alleles may be more stable depending on their mode <strong>of</strong> action and the particular<br />

resistance they control. For example, ‘Calcutta 4’ has a polygenic hypersensitive<br />

response that stops all development <strong>of</strong> the pathogen, but it also possesses recessive<br />

alleles controlling partial resistance in its plantain hybrids. This partial resistance<br />

simply slows disease development and may be more difficult to circumvent than<br />

the hypersensitive response, which already failed when screening young plants with<br />

Papua New Guinea strains. Hypersensitive responses are <strong>of</strong>ten associated with a genefor-gene<br />

host-pathogen interaction (Flor, 1971), but this hypothesis has not been<br />

193


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

tested in Musa. Resistance to black <strong>leaf</strong> streak disease provided by the recessive bs 1<br />

gene may be stable in the host-plant because virulence requires a rare mutation to<br />

the dominant allele in the respective locus <strong>of</strong> the pathogen.<br />

In conclusion, improved propagules with partial resistance to black <strong>leaf</strong> streak<br />

disease coming from distinct genetic sources, along with crop husbandry techniques<br />

are part <strong>of</strong> a holistic approach for long-term, sustainable productivity in Musa<br />

farming systems.<br />

Acknowledgements<br />

To Sarah and Yannick Vuylsteke, who kindly allowed the senior author to take some<br />

<strong>of</strong> their time in order to write some <strong>of</strong> the reports included in this review article. The<br />

Directorate General <strong>of</strong> International Cooperation (DGIC, Belgium) supported this<br />

research at the International Institute <strong>of</strong> Tropical Agriculture (IITA, Nigeria).<br />

References<br />

Carlier J., M.H. Lebrun, M.F. Zapater, C. Dubois and X. Mourichon. 1996. Genetic structure<br />

<strong>of</strong> the global population <strong>of</strong> banana black <strong>leaf</strong> streak fungus <strong>Mycosphaerella</strong> fijiensis.<br />

Molecular Ecology 5:499-510.<br />

Carlier J., X. Mourichon, D. González de León, M.F. Zapater and M.H. Lebrun. 1994. DNA<br />

restriction fragment length polymorphism in <strong>Mycosphaerella</strong> fijiensis that cause banana<br />

<strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>. Phytopathology 84:751-756.<br />

Carlier J., A. El Hadrami, H. Hayden, M.F. Zapater and F. Lapeyre. 1999. Population study <strong>of</strong><br />

<strong>Mycosphaerella</strong> fijiensis and genetic improvement <strong>of</strong> <strong>bananas</strong> for resistance to black <strong>leaf</strong><br />

streak disease. Pp. 40 in Abstracts <strong>of</strong> the International Symposium on the Molecular and<br />

Cell Biology <strong>of</strong> Banana, Cornell University, New York.<br />

Craenen K. 1994. Assessment <strong>of</strong> black sigatoka resistance in segregating progenies. MusAfrica<br />

4:4-5.<br />

Craenen K. 1997. Technical Manual on Black Sigatoka Disease <strong>of</strong> Banana and Plantain. IITA,<br />

Ibadan, Nigeria. 23pp.<br />

Craenen K. 1998. Black Sigatoka Disease <strong>of</strong> Banana and Plantain: A Reference Manual. IITA,<br />

Ibadan, Nigeria. 60pp.<br />

Craenen K. 2001. Black Sigatoka Resistance in Plantain-Banana Hybrids: Assessment,<br />

Genetics, Resistance Mechanisms and their Effect on Yield. PhD Thesis, Katholieke<br />

Universiteit Leuven, Belgium. 155pp.<br />

Craenen K., J. Coosemans and R. Ortiz. 1997. The role <strong>of</strong> stomatal traits and epicuticular wax<br />

in resistance to <strong>Mycosphaerella</strong> fijiensis in banana and plantain Musa spp. Tropicultura<br />

15:136-140.<br />

Craenen K. and R. Ortiz. 1996. Effect <strong>of</strong> the black sigatoka resistance locus bs 1<br />

and ploidy<br />

level on fruit and bunch traits <strong>of</strong> plantain-banana hybrids. Euphytica 87:97-101.<br />

Craenen K. and R. Ortiz. 1997. Effect <strong>of</strong> the bs 1<br />

gene in plantain and banana hybrids in response<br />

to black sigatoka. Theoretical and Applied Genetics 95:497-505.<br />

Craenen K. and R. Ortiz. 1998. Influence <strong>of</strong> black sigatoka disease on the growth and yield<br />

<strong>of</strong> diploid and tetraploid plantains. Crop Protection 17:13-18.<br />

Craenen K., R. Ortiz, E. Karamura and D. Vuylsteke (eds). 2000. Proceedings <strong>of</strong> the<br />

First International Conference on Banana and Plantain for Africa. Acta Horticulturae 540.<br />

589pp.<br />

194


Session 4<br />

K. Craenen and R. Ortiz<br />

Dangl J. and E. Holub. 1997. La Dolce Vita: a molecular feast in plant-pathogen interactions.<br />

Cell 91:17-24.<br />

El-Hadrami A., M.F. Zapater, F. Lapeyre, X. Mourichon and J. Carlier. 1998a. A <strong>leaf</strong> disk assay<br />

to assess partial resistance <strong>of</strong> banana germplasm and aggressiveness <strong>of</strong> <strong>Mycosphaerella</strong><br />

fijiensis, the causal agent <strong>of</strong> black <strong>leaf</strong> streak disease in The 7 th International Congress<br />

<strong>of</strong> Plant Pathology, ICPP98, Vol. 2, Edinburg, Scotland, 9-16 August 1998.<br />

El-Hadrami A., M.F. Zapater, F. Lapeyre, C. Abadie, X. Mourichon and J. Carlier. 1998b.<br />

Evaluation sur fragments foliaires en survie de la résistance partielle du bananier et de<br />

l’agressivité de <strong>Mycosphaerella</strong> fijiensis, agent causal de la maladie des raies noires in<br />

Recontres de Mycologie-Phytopatologie, Aussois, 27 September – 1 October 1998. 1p.<br />

El-Hadrami A., C. Abadie and J. Carlier. 2000. Evaluation de la résistance partielle du bananier<br />

à <strong>Mycosphaerella</strong> fijiensis (maladie des raies noires) en conditions contrôlées et au champ<br />

in Rencontres de Mycologie-Phytopatologie, Aussois, 3-9 March 2000. 1p.<br />

Flor H.H. 1971. Current status <strong>of</strong> the gene-for-gene concept. Annual Review Phytopathology<br />

9:275-296.<br />

Fluhr R. 2001. Sentinels <strong>of</strong> disease. Plant resistance genes. Plant Physiology 127:1367-1374.<br />

Fouré E. 1982. Les cercosporioses du bananier et leurs traitements. Etude de la sensibilité<br />

variétale des bananiers et plantains à <strong>Mycosphaerella</strong> fijiensis Morelet au Gabon (maladie<br />

des raies noires). I. Incubation et évolution de la maladie. Fruits 37:749-771.<br />

Fouré E. 1985. Les cercosporioses du bananier et leurs traitements. Etude de la sensibilité<br />

variétale des bananiers et plantains à <strong>Mycosphaerella</strong> fijiensis Morelet au Gabon (maladie<br />

des raies noires). III. Comportement des variétés. Fruits 40:393-399.<br />

Fouré E. 1994. Leaf <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> banana and plantain caused by <strong>Mycosphaerella</strong> musicola<br />

and M. fijiensis. Pp. 37-46 in The Improvement and Testing <strong>of</strong> Musa: A Global<br />

Partnership (D.R. Jones, ed.). INIBAP, Montpellier, France.<br />

Fullerton R.A. 1994. Sigatoka <strong>leaf</strong> <strong>diseases</strong>. Pp. 12-14 in Compendium <strong>of</strong> Tropical Fruit<br />

Diseases (R.C. Ploetz, G.A. Zentmeyer, W.T. Nishijima, K.G. Rohrbach and H.D. Ohr, eds).<br />

APS Press, St. Paul, Minnesota.<br />

Fullerton R.A. and T.L. Olsen. 1991. Pathogen variability in <strong>Mycosphaerella</strong> fijiensis Morelet.<br />

Pp. 105-114 in Banana Diseases in Asia and the Pacific (R.V. Valmayor, B.E. Umaldi and<br />

C. Bejosano, eds). INIBAP, Los Baños, Philippines.<br />

Fullerton R.A. and T.L. Olesen. 1993. Pathogen diversity <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis Morelet.<br />

Pp. 201-211 in Breeding Banana and Plantain for Resistance to Diseases and Pests. (J.<br />

Ganry, ed.). CIRAD – INIBAP, Montpellier, France.<br />

Gauhl F., C. Pasberg-Gauhl, D. Vuylsteke and R. Ortiz.1994. Multilocational Evaluation <strong>of</strong><br />

Black Sigatoka Resistance in Banana and Plantain. IITA Research Guide 47. IITA, Ibadan,<br />

Nigeria. 59 pp.<br />

Harelimana G., P. Leproive, H. Jijakli and X. Mourichon.1996. Use <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

toxins for the selection <strong>of</strong> banana cultivars resistant to black <strong>leaf</strong> streak. Pp. 171-175 in<br />

Meeting on Tropical Plants, Montpellier, France, 11-15 March 1996. EUCARPIA,<br />

Montpellier, France.<br />

Hartman J. and D. Vuylsteke. 1999. Breeding for fungal resistance in Musa. Pp. 83-92 in<br />

Genetics and Breeding for Crop Quality and Resistance (G.T. Scarascia-Mugnozza,<br />

E. Porceddu and M.A. Pagnotta, eds). Kluwer Academic Press, Dordrecth.<br />

Hernández N.R. 1995. In vitro and greenhouse selection <strong>of</strong> Musa resistance in black sigatoka<br />

(<strong>Mycosphaerella</strong> fijiensis Morelet). INFOMUSA 4(1):15-16.<br />

Holub E.B. 2001. The arms race is ancient history in Arabidopsis, the wild flower. Nature<br />

Reviews 2:516-527.<br />

195


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Hoss R. 1998. Untersuchugen zur Funktion and Spezifitat pilzlicher Sekundamertaboliten im<br />

Pathosystem “schwarze Sigatokakrankheit” der Banane (Musa sp. – <strong>Mycosphaerella</strong><br />

fijiensis). PhD Thesis, 123 pp. (English abstract in Musarama 1999, 12(1):50)).<br />

IITA. 1995. Plant Health Management Division Annual Report 1994. IITA, Cotonou, Benin<br />

Republic.<br />

IITA. 1998. Project 7 – Improving plantain- and banana-based systems. Annual Report 1997.<br />

IITA, Ibadan, Nigeria.<br />

IITA. 1999. Project 7 – Improving plantain- and banana-based systems. Annual Report 1998.<br />

IITA, Ibadan, Nigeria.<br />

IITA. 2000. Project 7 – Improving plantain- and banana-based systems. Annual Report 1999.<br />

IITA, Ibadan, Nigeria.<br />

IITA. 2001. Project 2 – Improving plantain- and banana-based systems. Annual Report 2000.<br />

IITA, Ibadan, Nigeria.<br />

Jones, D.R. (ed.). 1994. The Improvement and Testing <strong>of</strong> Musa: A Global Partnership. INIBAP,<br />

Montpellier, France. 303pp.<br />

Leister D., J. Kurth, D.A. Laurie, M. Yano, T. Sasaki, K. Devos, A. Graner and P. Schulze-Lefert.<br />

1998. Rapid reorganization <strong>of</strong> resistance gene homologues in cereal genomes. Proceedings<br />

National Academy Sciences (USA) 95:370-375.<br />

Leproive P., C.P. Acuña and A.S. Riveros. 1993. Screening procedures for improving resistance<br />

to banana black <strong>leaf</strong> streak disease. Pp. 213-220 in Breeding Banana and Plantain for<br />

Resistance to Diseases and Pests. (J. Ganry, ed.). CIRAD – INIBAP, Montpellier, France.<br />

Mendoza. H.A. 1988. Progress in resistance breeding in potatoes as a function <strong>of</strong> efficiency<br />

<strong>of</strong> screenig procedures. Pp. 39-64 in Bacterial Diseases <strong>of</strong> the Potato. Centro Internacional<br />

de la Papa, Lima, Perú.<br />

Mobambo K.N., F. Gauhl, D. Vuylsteke, R. Ortiz, C. Pasberg-Gauhl and R. Swennen. 1993.<br />

Yield loss in plantain from black sigatoka <strong>leaf</strong> <strong>spot</strong> and field performance <strong>of</strong> resistant<br />

hybrids. Field Crops Research 35:35-42.<br />

Mobambo K.N., C. Pasberg-Gauhl, F. Gauhl and K. Zu<strong>of</strong>a. 1994. Early screening for black<br />

<strong>leaf</strong> streak/black sigatoka disease resistance under natural inoculation conditions.<br />

INFOMUSA 3(2):14-16.<br />

Mobambo K.N., C. Pasberg-Gauhl, F. Gauhl and K. Zu<strong>of</strong>a. 1997. Host response to black sigatoka<br />

in Musa germplasm <strong>of</strong> different ages under natural inoculation conditions. Crop<br />

Protection 16:359-363.<br />

Muller R., C. Pasberg-Gauhl, F. Gauhl, D. Kaemmer and G. Kahl. 1995. Tracing microsatellite<br />

polymorphisms within the Nigerian populations <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis. INFOMUSA<br />

4(1):9-11.<br />

Muller R., C. Pasberg-Gauhl, F. Gauhl, D. Kaemmer and G. Kahl. 1997. Oligonucleotide<br />

fingerprinting detects genetic variability at different levels in Nigerian <strong>Mycosphaerella</strong><br />

fijiensis. Journal <strong>of</strong> Phytopathology 145:25-30.<br />

Okole B., C. Memela, S. Rademan, K.J. Kunert and M. Brunette. 2000. Non-conventional<br />

breeding approaches for banana and plantain against fungal <strong>diseases</strong> at AECI. Acta<br />

Horticulturae 540:207-214.<br />

Okole B. and F.A. Schultz. 1997. Selection <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis – resistant cell lines<br />

from micro-cross sections <strong>of</strong> <strong>bananas</strong> and plantains. Plant Cell Reports 13:339-342.<br />

Ortiz R. 1997. Secondary polyploids, heterosis and evolutionary crop breeding for further<br />

improvement <strong>of</strong> the plantain and banana genome. Theoretical and Applied Genetics<br />

94:1113-1120.<br />

Ortiz R. 2000. Understanding the Musa genome: an update. Acta Horticulturae 540:157-168.<br />

196


Session 4<br />

K. Craenen and R. Ortiz<br />

Ortiz R. 2001. Dedication: Dirk. R. Vuylsteke: Musa scientist and humanitarian. Plant Breeding<br />

Reviews 21:1-25.<br />

Ortiz R., K. Craenen and D. Vuylsteke. 1998a. Ploidy manipulations and genetic markers as<br />

tools for analysis <strong>of</strong> quantitative trait variation in progeny derived from triploid plantain.<br />

Hereditas 126:255-259.<br />

Ortiz R., J.H. Crouch, D.R. Vuylsteke, R.S.B. Ferris and J. Okoro, 2000. Cultivar development,<br />

genotype x environment interaction and multi-site testing <strong>of</strong> improved plantain and<br />

banana germplam in sub-Saharan Africa. Pp. 84-106 in Genotype-by-environment<br />

interaction analysis <strong>of</strong> IITA mandate crops in sub-Saharan Africa (I.J. Ekanayake and<br />

R. Ortiz, eds). IITA, Ibadan, Nigeria..<br />

Ortiz R., J. Okoro, R. Apanisile and K. Craenen. 1995. Preliminary assessment <strong>of</strong> the yield<br />

potential <strong>of</strong> Musa hybrids under low external organic matter input. MusAfrica 7:15-17.<br />

Ortiz R. and D. Vuylsteke. 1994. Inheritance <strong>of</strong> black sigatoka resistance in plantain-banana<br />

(Musa spp.) hybrids. Theoretical and Applied Genetics 89:146-152.<br />

Ortiz R. and D. Vuylsteke. 1995. Genotype-by-environment interaction in Musa germplasm<br />

revealed by multi-site evaluation in sub-Saharan Africa. HortScience 30:795.<br />

Ortiz R., D. Vuylsteke and J.H. Crouch. 1998b. Musa genetics, ‘Calcutta 4’ and scientific ethics:<br />

reply to Shepherd’s letter. INFOMUSA 7(2):31-32.<br />

Ortiz R., D. Vuylsteke, R.S.B. Ferris, J.U. Okoro, A. N’Guessan, O.B. Hemeng, D.K. Yeboah, K.<br />

Afreh-Nuamah, E.K.S. Ahiekpor, E. Fouré, B.A. Adelaja, M. Ayodele, O.B. Arene, F.E.O.<br />

Ikiediugwu, A.N. Agbor, A.N. Nwongu, E. Okoro, G.O. Kayode, I.K. Ipinmoye, S.A. Akele<br />

and A. Lawrence. 1997. Developing new plantain varieties for Africa. Plant Varieties and<br />

Seeds 10:39.57.<br />

Ortiz R., D. Vuylsteke, J.U. Okoro, R.S.B. Ferris, O.B. Hemeng, D.K. Yeboah, C.C: Anojulu, B.A.<br />

Adelaja, O.B. Arene, A.N. Agbor, A.N. Nwongu, G. Kayode, I.K. Ipinmoye, S.A. Akele and<br />

A. Lawrence. 1993. Host response to black sigatoka across West and Central Africa.<br />

MusAfrica 3:8-10.<br />

Pino J.A. 1997. Selección temprana de mutantes de banana y plátano resistentes a<br />

<strong>Mycosphaerella</strong> fijiensis mediante fitotóxinas. Agrotecnia de Cuba 27:89-91.<br />

Rommens C.M. and G.M. Kishore. 2000. Exploiting the full potential <strong>of</strong> disease resistance<br />

genes for agricultural use. Current Opinion in Biotechnology 11:120-125.<br />

Sharrock S.L., J.-P. Horry and E. Frison. 2000. The state <strong>of</strong> use <strong>of</strong> Musa diversity. Pp. 223-<br />

244 in Broadening the Genetic Base <strong>of</strong> Crop Production (H.D. Cooper, C. Spillane and T.<br />

Hodgkin, eds). CABI Publishing – FAO – IPGRI, Wallingford.<br />

Simmonds N.W. 1979. Principles <strong>of</strong> Crop Improvement. Longman, London and New York.<br />

Simmonds N.W. and J. Smartt. 1999. Principles <strong>of</strong> Crop Improvement (2 nd edition). Blackwell<br />

Science, Oxford. Pp.227-261.<br />

Stover R.H. 1976. Distribution and cultural characteristics <strong>of</strong> the pathogen causing banana<br />

<strong>leaf</strong> <strong>spot</strong>. Tropical Agriculture (Trinidad) 53:111-114.<br />

Stover R.H. 1980. Sigatoka <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong> and plantains. Plant Disease 64:<br />

750-755.<br />

Stover R.H. 1983. Effet de la cercosporiose noire sur les plantains en Amérique Centrale. Fruits<br />

38:326-329.<br />

Stover R.H. and J.D. Dickson. 1970. Leaf <strong>spot</strong> <strong>of</strong> banana caused by <strong>Mycosphaerella</strong> musicola:<br />

methods <strong>of</strong> measuring <strong>spot</strong>ting prevalence and severity. Tropical Agriculture (Trinidad)<br />

47:289-302.<br />

Vakili N.G. 1968. Responses <strong>of</strong> Musa acuminata species and edible cultivars to infection by<br />

<strong>Mycosphaerella</strong> musicola. Tropical Agriculture (Trinidad) 45:13-22.<br />

197


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Vuylsteke D. 2000. Breeding <strong>bananas</strong> and plantains: from intractability to feasibility. Acta<br />

Horticulturae 540:149-156.<br />

Vuylsteke D. 2001. Strategies for utilization <strong>of</strong> genetic variation in plantain improvement.<br />

PhD Thesis, Katholieke Universiteit Leuven, Belgium.<br />

Vuylsteke D., R. Swennen and R. Ortiz. 1993. Development and performance <strong>of</strong> black sigatokaresistant<br />

tetraploid hybrids <strong>of</strong> plantains (Musa spp., AAB group). Euphytica 65:33-42.<br />

198


Session 4<br />

C. Jenny et al.<br />

Conventional breeding<br />

<strong>of</strong> <strong>bananas</strong><br />

C. Jenny 1 ,K.Tomekpé 2 ,F.Bakry 3 and J.V. Escalant 4<br />

Abstract<br />

Whereas ancestral <strong>bananas</strong> are fertile diploids, the main groups <strong>of</strong> <strong>bananas</strong> grown today are clones<br />

<strong>of</strong> plants, mostly triploids, which are reproduced entirely vegetatively and consequently difficult<br />

to breed. Conventional breeding techniques have yielded new varieties conventional breeding can<br />

accomplish only so much. Not all genetic combinations necessarily lead to useful hybrids. This<br />

communication presents the strategies which have been and are still being used in this<br />

field–namely the 3x/2x scheme and the creation <strong>of</strong> triploid hybrids from ancestral diploid<br />

varieties–and draws attention to their strengths and limitations.<br />

Resumen - Mejoramiento convencional de los bananos<br />

Mientras que los bananos ancestrales son diploides fértiles, los principales grupos de bananos<br />

que se cultivan actualmente son clones de las plantas, en su mayoría triploides, que se reproducen<br />

solo vegetativamente y, en consecuencia, son difíciles de mejorar. Las técnicas de mejoramiento<br />

convencional han producido nuevas variedades y el mejoramiento convencional tiene sur<br />

límites. Todas las combinaciones genéticas no llevan necesariamente a generar híbridos útiles.<br />

En este trabajo se presentan las estrategias que todavía están siendo utilizadas en este campo,<br />

a saber, el esquema 3x/2x y la creación de los híbridos triploides a partir de las variedades<br />

ancestrales diploides, y se destacan sus fortalezas y debilidades.<br />

Résumé - Amélioration conventionnelle des bananiers<br />

Alors que les bananiers ancestraux sont des diploïdes fertiles, les principaux groupes de<br />

bananiers cultivés aujourd’hui sont des clones de plantes, principalement triploïdes, qui se<br />

reproduisent selon un mode entièrement végétatif et sont donc difficiles à améliorer. Les<br />

techniques classiques d’amélioration ont produit de nouvelles variétés, et l’amélioration<br />

conventionnelle a ses limites. Toutes les combinaisons génétiques ne conduisent pas<br />

nécessairement à des hybrides utiles. Cette communication présente les stratégies qui ont été,<br />

et sont encore utilisées dans ce domaine, c’est-à-dire le schéma 3x/2x et la création d’hybrides<br />

triploïdes à partir de variétés ancestrales diploïdes, et insiste sur leurs avantages et leurs limites.<br />

1<br />

CIRAD-FHLOR, Guadeloupe<br />

2<br />

CARBAP, Douala, Cameroon<br />

3<br />

CIRAD, Montpellier, France<br />

4<br />

INIBAP, Montpellier, France<br />

199


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Introduction<br />

Bananas and plantains are difficult crops to breed because most <strong>of</strong> the important<br />

and popular varieties are highly sterile and therefore do not produce seeds.<br />

Furthermore, compared to many other important food crops, there is a relative lack<br />

<strong>of</strong> knowledge on Musa genetics and cytogenetics. Despite these constraints,<br />

important progress has been made in the genetic improvement <strong>of</strong> Musa in recent<br />

years and new varieties are now becoming available from breeding programmes.<br />

Major breeding programmes that use conventional breeding methodologies are<br />

located at the Fundación Hondureña de Investigación Agrícola (FHIA) in Honduras,<br />

the Centre de Coopération Internationale en Recherche Agronomique pour le<br />

Développement (CIRAD-FLHOR) in France and Guadeloupe, the International<br />

Institute <strong>of</strong> Tropical Agriculture (IITA) in Nigeria and Uganda, the Centre Africain<br />

de Recherches sur Bananiers et Plantains (CARBAP) in Cameroon and the Empresa<br />

Brasiliera de Pesquisa Agropecuaria (EMBRAPA) in Brazil.<br />

In the last decades new banana varieties were mainly created by using<br />

conventional breeding techniques (Bakry and Horry, 1992; Jenny et al., 1994;<br />

Menendez and Shepherd, 1975; Rowe and Rosales, 1992; Shepherd, 1968; Soares<br />

Filho et al., 1992; Swennen and Vuylsteke, 1990; Tomekpé et al., 1998). One <strong>of</strong> the<br />

peculiarities <strong>of</strong> <strong>bananas</strong> is the need to adapt these techniques to the genetics <strong>of</strong><br />

polyploid plants.<br />

This communication presents the strategies which have been and are still being<br />

used in this field, and draws attention to their strengths and limitations.<br />

Constraints to the improvement <strong>of</strong> <strong>bananas</strong><br />

Whilst ancestral <strong>bananas</strong> are fertile diploids, the main groups <strong>of</strong> <strong>bananas</strong> grown<br />

today are clones <strong>of</strong> plants, mostly triploids, which are reproduced entirely<br />

vegetatively. For producers and consumers, this feature presents two advantages:<br />

triploidy gives the plant vigour, making it easier to grow than diploids; clonal<br />

propagation assures uniformity which facilitates management, both in the field and<br />

throughout the distribution and sale chain. Finally, triploidy ensures sterility <strong>of</strong> the<br />

fruit, enabling it to be eaten.<br />

On the other hand, the nature <strong>of</strong> these plants also presents potential dangers<br />

for their cultivation, and obstacles to their improvement. First, the genetic<br />

uniformity <strong>of</strong> these plants facilitates the spread <strong>of</strong> <strong>diseases</strong> and increases the impact<br />

<strong>of</strong> the latter on banana plants. For example, the ‘Cavendish’ varieties throughout<br />

the world are all susceptible to <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>. The same is true for plantains<br />

with regard to black <strong>leaf</strong> streak disease, whether in Africa, South America or Asia.<br />

Furthermore, the sterility <strong>of</strong> the clones currently grown is a considerable hindrance<br />

to their genetic improvement.<br />

It is therefore clear that the primary needs in terms <strong>of</strong> breeding have to do with<br />

the various <strong>diseases</strong> which affect the crop. Among the most important is Fusarium<br />

wilt (caused by Fusarium oxysporum f.sp. cubense), and Sigatoka disease (caused<br />

by <strong>Mycosphaerella</strong> musicola) and black <strong>leaf</strong> streak (caused by <strong>Mycosphaerella</strong><br />

fijiensis). However there are also other improvement criteria, especially for dessert<br />

200


Session 4<br />

C. Jenny et al.<br />

<strong>bananas</strong> for export, which have to do with fruit quality and post-harvest<br />

characteristics.<br />

One <strong>of</strong> the first ideas was to draw on the natural existing genetic resources<br />

to find solutions for replacement varieties. Thus, during the 20 th century, all <strong>of</strong> the<br />

‘Gros Michel’, which was traditionally grown on commercial plantations, was<br />

gradually replaced by ‘Cavendish’ which is resistant to race 1 <strong>of</strong> Fusarium wilt. In<br />

certain areas, attempts were made to introduce ‘Pisang awak’ and ‘Bluggoe’ to replace<br />

or supplement the plantains susceptible to black <strong>leaf</strong> streak. These solutions are<br />

unfortunately <strong>of</strong> limited value because the natural varieties suitable for cultivation<br />

are very few, and are susceptible to other parasites, such as nematodes and weevils.<br />

Often the fruit produced is not the kind favoured locally and fails to gain acceptance.<br />

Finally, the problem <strong>of</strong> potential susceptibility ends up being passed on to the new<br />

clones through the creation <strong>of</strong> a new genetic uniformity which can easily be<br />

circumvented by the pathogens. The need for genuine genetic improvement is<br />

therefore real.<br />

An improvement strategy: the 3x/2x scheme<br />

The prerequisite to any genetic improvement strategy is to analyse the available tools.<br />

For conventional methods, it is necessary to identify the required qualities <strong>of</strong> fertility<br />

and useful characters <strong>of</strong> the germplasm which can be used. Since the 1920s, the<br />

questions tackled have been: which cultivars can be improved, and what<br />

characteristics can be introduced into them. Detailed analysis <strong>of</strong> the fertility <strong>of</strong> the<br />

main triploid cultivars resulted in the identification <strong>of</strong> a certain number <strong>of</strong> triploid<br />

clones which had retained residual female fertility, like ‘Gros Michel’ (Musa cv. AAA)<br />

and the French-type plantains (Musa cv. AAB). Conversely, certain clones turned<br />

out to be completely sterile, like ‘Cavendish’ 1 (Musa cv. AAA).<br />

The search for sources <strong>of</strong> resistance to the main <strong>diseases</strong> was pursued by<br />

looking for wild varieties present in the area <strong>of</strong> origin <strong>of</strong> <strong>bananas</strong>, Southeast Asia.<br />

In this way, the variety Musa acuminata ssp. burmannicoïdes ‘Calcutta’ 4, among<br />

others, was identified, notably for its resistance to M. musicola and to<br />

M. fijiensis.<br />

Armed with these tools, breeders crossed the various parents among themselves<br />

to improve the cultivated triploid clones, by combining the residual female fertility<br />

<strong>of</strong> the latter with the strong male fertility <strong>of</strong> the wild varieties (Figure 1). Detailed<br />

analysis <strong>of</strong> the mechanisms brought into play during these hybridizations has<br />

shown that the fertile triploid parent produced gametes during meiosis with a v<br />

ariable chromosome number, ranging from 11 to 66. Among these gametes, those<br />

with 33 chromosomes, i.e. with a true restitution nucleus, were particularly useful.<br />

Among the progeny the first priority was to look for tetraploid hybrids, the result<br />

<strong>of</strong> fusion between this restitution nucleus and a normal haploid gamete from the<br />

wild parent. The final result resembled more a fusion than a true hybridization,<br />

because recombination on the triploid side is low, and on the wild parent, being<br />

very homozygous, produces very homogeneous gametes. In this way hybrid<br />

1<br />

It is now known that under certain stress conditions, it is possible to obtain pollen from ‘Cavendish’ clones.<br />

201


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

tetraploids were produced with a genome similar to the complete genome <strong>of</strong> the<br />

triploid parent one is seeking to improve, supplemented by a wild haploid genome<br />

contributing sources <strong>of</strong> resistance. A certain degree <strong>of</strong> recombination would however<br />

explain the sometimes mediocre quality <strong>of</strong> the tetraploid hybrids obtained.<br />

Female parent<br />

2n = 3x = 33<br />

Male parent<br />

2n = 2x = 22<br />

MEIOSIS<br />

Female gametes<br />

n = variable<br />

Some non reduced gametes<br />

Male parent<br />

n = 11<br />

GAMETE FUSION<br />

Progeny: mix <strong>of</strong> euploid and eneuploid embryos<br />

2n = 22, 33 … 44 ...77<br />

Tetraploid hybrid<br />

Figure 1. Scheme for creating tetraploid hybrids from triploid and diploid parents.<br />

Two complementary methods exist to improve this strategy:<br />

• It is possible to produce diploid hybrids which appear to be useful, and which<br />

can themselves be used in an improvement programme.<br />

• The produced tetraploid hybrids, being more fertile than the triploid parent, can<br />

be reintroduced in the crossing schemes with a view to creating secondary triploid<br />

hybrids. One must not however forget that in this latter case the genetic gain obtained<br />

from nuclear restitution will be reduced by recombination which will occur during<br />

the meiosis <strong>of</strong> this tetraploid.<br />

Nevertheless, many research organizations throughout the world produced hybrids<br />

using this scheme. As for cooking banana hybrids, the cultivars ‘FHIA-21’ (FHIA,<br />

Honduras), ‘CRBP-39’ (CARBAP, Cameroon) and ‘Bita-3’ (IITA, Nigeria) (Ortiz and<br />

Vuylsteke, 1998) should be mentioned. In the majority <strong>of</strong> cases, these hybrids were<br />

found to be resistant to both Sigatoka disease and black <strong>leaf</strong> streak disease, which was<br />

202


Session 4<br />

C. Jenny et al.<br />

the primary aim <strong>of</strong> their breeding. This immediate success can be attributed to the<br />

type <strong>of</strong> resistance used. Very <strong>of</strong>ten the wild male parent chosen was ‘Calcutta 4’ which<br />

is highly resistant (HR) to <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>. It is generally accepted that this type <strong>of</strong><br />

resistance depends on a smaller number <strong>of</strong> genes than partial resistance (PR), and<br />

consequently can be more easily transmitted during hybridization. The result however<br />

poses two questions: the exact transmission mechanism <strong>of</strong> these resistances is not<br />

known; and as a corollary, one cannot predict the durability <strong>of</strong> these resistances over<br />

time. Logically one would prefer to try to introduce into these hybrids partial resistance,<br />

known to be more durable but more complex, which has, for example, been tried by<br />

using parents such as M-53.<br />

The 3x x 2x strategy described here also suffers from several limitations.<br />

• From a purely technical point <strong>of</strong> view, this strategy depends on the existence <strong>of</strong><br />

valuable triploid clones exhibiting exploitable female fertility. There are relatively<br />

few <strong>of</strong> these, limiting the possibilities <strong>of</strong> enlarging the genetic base <strong>of</strong> these crossings.<br />

• The hybrid populations obtained by crossing are usually small.<br />

• The percentage <strong>of</strong> tetraploid hybrids being small, the possibilities for selection<br />

are very limited. This technique cannot be used to work simultaneously on a large<br />

number <strong>of</strong> characters to be improved.<br />

More recently, the realization <strong>of</strong> the existence <strong>of</strong> potentially active sequences <strong>of</strong><br />

banana streak virus (BSV) contained within the balbisiana genome has further reduced<br />

the possibilities for using this strategy. Moreover, the most fertile triploid genomes<br />

are <strong>of</strong>ten <strong>of</strong> the AAB or ABB types. Such germplasm should not be used for breeding<br />

as long as the activation mechanisms <strong>of</strong> BSV are unknown. One might nevertheless<br />

use it to produce various diploid type AA hybrids, notably plantain diploids which<br />

constitute 50% <strong>of</strong> the 3x/2x descendants produced by certain plantain cultivars.<br />

Several dozen <strong>of</strong> these hybrids produced at CARBAP were found to be negative for<br />

IC-PCR and have never expressed any BSV symptoms in the field even under stress<br />

conditions in which nearly all tetraploid hybrids frequently present symptoms.<br />

Lastly, the tetraploid nature <strong>of</strong> the hybrids formed <strong>of</strong>ten leads to problems <strong>of</strong> fruit<br />

quality. In dessert-type hybrids the firmness <strong>of</strong> the pulp is less in the tetraploids.<br />

This problem also exists in cooking hybrids, albeit less pronounced. In numerous<br />

polyploid plants, it has been found that the water content <strong>of</strong> the cells increases with<br />

ploidy level. If this proves true for <strong>bananas</strong>, it might explain the phenomenon.<br />

Moreover, at the tetraploid level, female fertility might be restored, <strong>of</strong>ten leading to<br />

the presence <strong>of</strong> seed in the fruit, which is unacceptable to present-day consumers.<br />

Recent developments with biotechnological tools and a better understanding <strong>of</strong><br />

the evolutionary processes <strong>of</strong> <strong>bananas</strong> have led to the introduction <strong>of</strong> another<br />

breeding strategy aimed at the production <strong>of</strong> triploid hybrids.<br />

Creation <strong>of</strong> triploid hybrids from ancestral diploids<br />

The natural emergence <strong>of</strong> triploid cultivars derived from ancestral diploid varieties<br />

is due to the accidental production <strong>of</strong> unreduced gametes in one <strong>of</strong> the diploid<br />

parents during hybridization (Simmonds, 1962). The 4x x 2x strategy is a copy <strong>of</strong><br />

203


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

this natural evolutionary process. The meiotic error leading to these unreduced<br />

gametes is replaced by a chromosome doubling in one <strong>of</strong> the parents using<br />

colchicine (Vakili, 1967; Stover and Buddenhagen, 1986; Bakry et al., 1997).<br />

Unlike the scheme just described, this strategy does not attempt to improve<br />

existing varieties, but rather to create new improved varieties, close to the<br />

established targets, using ancestral varieties. These new hybrids should therefore<br />

combine all the classic characteristics <strong>of</strong> the banana requiring improvement, plus<br />

the improved characters for which the strategy was introduced.<br />

The triploid hybrids are obtained by simple hybridisation between a diploid<br />

parent and a tetraploid parent (Figure 2). This hybridisation has to be the end <strong>of</strong><br />

the procedure, since the product obtained is almost completely sterile, and can<br />

thus no longer be improved by conventional means. The tetraploid parent has<br />

previously been obtained by doubling with colchicine an ancestral diploid or an<br />

improved diploid. After treatment, the purely tetraploid nature <strong>of</strong> the parent is<br />

checked by flow cytometry. The success <strong>of</strong> the strategy rests on the judicious choice<br />

<strong>of</strong> the parents. These can either be natural diploid cultivars or improved diploid<br />

cultivars.<br />

Tetraploid development<br />

AAw<br />

AAcv<br />

AAcv<br />

x<br />

x<br />

x<br />

AAcv<br />

AAcv<br />

BBw<br />

AAcv<br />

ABcv<br />

colchicine<br />

colchicine<br />

AAAAcv<br />

AABBcv<br />

(1)<br />

(2)<br />

Triploid development<br />

AAcv<br />

AAw<br />

x<br />

x<br />

AAAAcv<br />

AAAAcv<br />

AAAAcv<br />

AAAAcv<br />

(3,4)<br />

BBw<br />

AAcv<br />

AAw<br />

x<br />

x<br />

x<br />

AAAAcv<br />

AABBcv<br />

AABBcv<br />

AABcv<br />

AABcv<br />

AABcv<br />

(3,5)<br />

1. Nearly 20 clones developed at CARBAP and CIRAD.<br />

2. Currently stopped because <strong>of</strong> BSV concern.<br />

3. Annual hybrid population size <strong>of</strong> nearly 400 plants in the field.<br />

4. 98% <strong>of</strong> the progeny is triploid.<br />

5. Currently stopped because <strong>of</strong> SV concern.<br />

Figure 2.Strategy for creating triploid hybrids from diploid material.<br />

204


Session 4<br />

C. Jenny et al.<br />

At CIRAD, where this method has been favoured for several years, progress has<br />

been made in characterizing available genetic resources and understanding the<br />

relationships between ancestral and cultivated varieties, thus facilitating the<br />

definition <strong>of</strong> pools <strong>of</strong> parent lines according to the desired results (Jenny et al., 1999).<br />

Moreover, the development <strong>of</strong> molecular tools has made, and continues to make,<br />

this strategy more efficient. Among the most important results, it has been possible<br />

to demonstrate the uniparental inheritance <strong>of</strong> cytoplasmic organelles, which helps<br />

in identifying the phylogeny <strong>of</strong> <strong>bananas</strong> (Fauré et al., 1994). Musa is one <strong>of</strong> the few<br />

species with biparental cytoplasmic inheritance: paternal inheritance <strong>of</strong> mitochondria<br />

and maternal inheritance <strong>of</strong> chloroplasts.<br />

The extent <strong>of</strong> genetic variability within the acuminata genome has been related<br />

to the variability in fruit quality in the main cultivated groups. The most striking<br />

example is probably the strict relationship between the sub-species M. acuminata<br />

banksii and cooking <strong>bananas</strong>. It has thus been possible to produce triploid cooking<br />

banana hybrids <strong>of</strong> purely acuminata origin. The variability <strong>of</strong> the acuminata genome<br />

also permits variation in the type <strong>of</strong> fruit type and plant obtained, whether dessert<br />

or cooking, but also with regards to the fruit’s sweetness or acidity, its length, the<br />

plant’s number <strong>of</strong> suckers, and its yield, etc.<br />

One <strong>of</strong> the main attractions <strong>of</strong> this strategy rests on making use <strong>of</strong> highly fertile<br />

parents, thus leading to a large progeny. In preliminary results, CARBAP identified<br />

in the 100 or so individuals <strong>of</strong> the progeny <strong>of</strong> a BB x AAA cross, about 20% <strong>of</strong><br />

hybrids having useful resistance to black <strong>leaf</strong> streak disease (BLSD).<br />

To sum up, the 4x x 2x strategy presents a certain number <strong>of</strong> undeniable<br />

advantages in genetics:<br />

• The number <strong>of</strong> available parents is only limited by our knowledge <strong>of</strong> the<br />

germplasm, knowing that subsequently pools <strong>of</strong> parents that produce the targeted<br />

results will be formed.<br />

• Using a tetraploid parent allows better control <strong>of</strong> the heritability <strong>of</strong> characters<br />

due to limited recombination within the polyploid parent.<br />

• Highly fertile parents lead to the production <strong>of</strong> large populations in which it is<br />

possible to set up a true selection programme, possibly based on several<br />

improvement criteria.<br />

In order to improve this strategy, it would be valuable to further enlarge the<br />

diploid crossing base either by crossing known diploids to obtain improved<br />

diploids or by more collection missions in the regions <strong>of</strong> interest. CARBAP is<br />

currently developing, from diploid plantain hybrids resistant to BLSD, second and<br />

third generation improved diploids (secondary and tertiary diploids obtained by<br />

crossing diploid plantain hybrids with different sources <strong>of</strong> resistance to BLSD) which<br />

are in the process <strong>of</strong> chromosome doubling so as to be integrated into this strategy.<br />

In the near future, the use <strong>of</strong> molecular markers should facilitate selection at the<br />

parent level (identification <strong>of</strong> the genes to be transferred) and hybrid level<br />

(identification <strong>of</strong> the genes effectively transferred). Nowadays for example, three<br />

QTL (Quantitative Trait Loci) have already been localized in relation to resistance<br />

to BLSD (Persley and George, 1999). Their use remains dependent on the completion<br />

<strong>of</strong> the genetic map being created by CIRAD. Among the future challenges,<br />

205


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

a molecular characterization <strong>of</strong> fruit quality will be particularly important for the<br />

creation <strong>of</strong> dessert-type varieties which are competitive in the world market.<br />

Once again however, BSV acts as a brake on the best use <strong>of</strong> this strategy. The<br />

B genome is excluded from breeding schemes because <strong>of</strong> possible activation <strong>of</strong><br />

integrated viral DNA sequences within the genome. It is known that activation is<br />

linked to certain stresses including hybridisation and in vitro multiplication. These<br />

two stresses do not necessarily activate the same integrated sequences, and could<br />

therefore have additive effects (Lheureux et al., 2003). It is therefore very important<br />

to study the B genome in more detail in order to be able, as soon as possible, to<br />

use it again for creating varieties.<br />

• The range <strong>of</strong> banana species is not so rich that we can manage for their<br />

improvement without the one or two species from which cultivated <strong>bananas</strong><br />

originated.<br />

• In particular, the use <strong>of</strong> M. balbisiana in crossing schemes confers resistance, in<br />

particular to <strong>Mycosphaerella</strong> musicola and M. fijiensis, on the hybrids produced.<br />

This increased vigour renders the plants less susceptible to growth stresses.<br />

• Finally, the high natural fertility <strong>of</strong> M. balbisana is an advantage for the<br />

production <strong>of</strong> large numbers <strong>of</strong> progeny.<br />

Among the avenues to explore, one should mention collecting in the regions <strong>of</strong><br />

origin <strong>of</strong> the species M. balbisiana, the analysis <strong>of</strong> the germplasm present in<br />

collections and genetic methods <strong>of</strong> improvement such as for example the extraction<br />

<strong>of</strong> the B genome from interspecific clones. In each case, it will be essential to gather<br />

international expertise and competence on the subject, and the PROMUSA workshop<br />

on the diversity <strong>of</strong> the M. balbisiana genome held in Bangkok in 2002 was a good<br />

starting point.<br />

Conclusion<br />

At a more technical level, it emerges from this presentation that conventional breeding<br />

can accomplish only so much. Not all genetic combinations necessarily lead to useful<br />

hybrids. Crossing cooking <strong>bananas</strong> with dessert ones, for example, generally leads<br />

to intermediate hybrids <strong>of</strong> no real value. Not all combinations are possible, and many<br />

simply do not work, when using conventional hybridization techniques.<br />

Conventional breeding methods should be viewed as just a part - admittedly a<br />

significant part - <strong>of</strong> a more general genetic improvement strategy. Unconventional<br />

techniques can usefully complete this arsenal (Novak, 1992; Sagi et al., 1995;<br />

Sharrock et al., 2000). For example, protoplast fusion is one <strong>of</strong> the ways which could<br />

increase the possibilities <strong>of</strong> overcoming certain fertility barriers in combining parental<br />

lines. This technique also allows transmission <strong>of</strong> intact genotypes by bypassing the<br />

recombination phenomena associated with meiosis, and also results in modification<br />

at the cytoplasmic level <strong>of</strong> the cells, thus potentially leading to novel results.<br />

Mutagenesis and genetic transformation - aside from the arguments as to their<br />

appropriateness - might improve either the parents or, at the other end <strong>of</strong> the chain,<br />

the hybrids created by conventional hybridization and which lack certain characters.<br />

A complementary approach, recombinant DNA, has led to the production <strong>of</strong> the first<br />

206


Session 4<br />

C. Jenny et al.<br />

transgenic banana and plantain, but also to the generation <strong>of</strong> a large number <strong>of</strong><br />

transgenic lines with agronomically useful genes. Transgenic plants transformed with<br />

genes encoding antifungal proteins are currently available for field-testing.<br />

It is essential that Musa breeders and biotechnologists work together to accelerate<br />

improvement. In view <strong>of</strong> the limited resources being devoted to research into Musa<br />

improvement, and knowing the scale <strong>of</strong> the problems to be overcome, it is important<br />

to strengthen collaboration between the various institutions working on the problem,<br />

and to take advantage <strong>of</strong> all the resources already available to develop research on<br />

Musa. The Global Programme for Musa Improvement (PROMUSA) aims to bring<br />

together all the scientists working on Musa genetic improvement, involving in the<br />

same programme geneticists, biotechnologists, but also pathologists and physiologists.<br />

References<br />

Bakry F. and J.P. Horry. 1992. Tetraploid hybrids from interploid 3x/2x crosses in cooking<br />

<strong>bananas</strong>. Fruits 47:641-655.<br />

Bakry F., F. Carreel, M.L. Caruana, F.X. Côte, C. Jenny and H. Tézenas du Montcel 1997.<br />

Les bananiers. Pp. 109-140 in L’amélioration des plantes tropicales (A. Charrier, M. Jacquot,<br />

S. Hamon and D. Nicolas, eds). CIRAD and ORSTOM, Paris and Montpellier, France.<br />

Fauré S., J.L. Noyer, F. Carreel, J.P. Horry, F. Bakry and C. Lanaud. 1994. Maternal inheritance<br />

<strong>of</strong> chloroplast genome ant paternal inheritance <strong>of</strong> mitochondrial genome in <strong>bananas</strong> (Musa<br />

acuminata). Current Genetics 25:265-269.<br />

Jenny C., E. Auboiron and A. Beveraggi. 1994. Breeding plantain-type hybrids at CRBP.<br />

Pp. 176-187 in The improvement and testing <strong>of</strong> Musa: a global partnership (D.R. Jones,<br />

ed.). Proceedings <strong>of</strong> the first global conference <strong>of</strong> the IMTP held at FHIA, Honduras, 27-<br />

30 April 1994. INIBAP, Montpellier, France.<br />

Jenny C., F. Carreel, K. Tomekpé, X. Perrier, C. Dubois, J.P. Horry and H. Tézenas du Montcel.<br />

1999. Les bananiers. Pp. 113-139 in Diversité génétique des plantes tropicales cultivées<br />

(P. Hamon, M. Seguin, X. Perrier and J.C. Glaszmann, eds). CIRAD, Montpellier, France.<br />

Lheureux F., F. Carreel, C. Jenny, B.E.L. Lockhart and M.L. Iskra-Caruana. 2003. Identification<br />

<strong>of</strong> genetic markers linked to banana streak disease expression in interspecific Musa hybrids.<br />

Theor. and Appl. Gen. 106(4):594-598.<br />

Menendez T. and K. Shepherd. 1975. Breeding new <strong>bananas</strong>. World crops (May/June):<br />

104-112.<br />

Novak F.J. 1992. Musa (<strong>bananas</strong> and plantains). Pp. 449-487 in Biotechnology <strong>of</strong> perennial<br />

fruit crops (F.A. Hammerschlag and R.E Litz, eds). CAB International, Wallington, UK.<br />

Ortiz R. and D. Vuylsteke. 1998. ‘Bita-3’: a starchy banana with partial resistance to black<br />

Sigatoka and tolerance to streak virus. HortScience 33:358-359.<br />

Persley G.J. and P. George (eds). 1999. Banana, Breeding and Biotechnology - Commodity<br />

advances through banana improvement project research, 1994 - 1998. The World Bank,<br />

Washington D.C. 62pp.<br />

Rowe P. and F. Rosales. 1992. Genetic improvement <strong>of</strong> <strong>bananas</strong>, plantains and cooking <strong>bananas</strong><br />

in FHIA, Honduras. Pp. 243-266 in Breeding <strong>bananas</strong> and plantains (J. Ganry, ed.).<br />

Proceedings <strong>of</strong> an International Symposium on Genetic Improvement <strong>of</strong> Bananas for their<br />

Resistance to Diseases and Pests. CIRAD-FLHOR, Montpellier, France.<br />

207


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Sagi L., B. Panis, S. Remy, H. Scho<strong>of</strong>s, K. de Smet, R. Swennen and B.P.A. Cammue. 1995.<br />

Genetic transformation <strong>of</strong> banana and plantain (Musa spp.) via particle bombardment.<br />

Biotechnology 13:481-485.<br />

Sharrock S., J.P. Horry and E.A. Frison. 2001. The state <strong>of</strong> the use <strong>of</strong> Musa diversity. Pp. 223-<br />

243 in Broadening the genetic base <strong>of</strong> crop production (H.D. Cooper, C. Spillane and<br />

T. Hodgkin, eds). IPGRI/FAO, Rome , Italy.<br />

Shepherd K. 1968. Banana breeding in the West Indies. Pest articles and news summaries<br />

14:370-379.<br />

Simmonds N.W. 1962. The evolution <strong>of</strong> the <strong>bananas</strong>. Longman, Green & Co, London, UK .<br />

Soares Filho W., S. Dos, Z.J.M. Cordeiro, K. Shepherd, J.L.L. Dantas, S. de Oliveira e Silva<br />

and M.A.P. da Cunha. 1992. The banana genetic improvement programme at<br />

CNPMF/EMBRAPA, Brazil. Pp. 339-346 in Breeding <strong>bananas</strong> and plantains (J. Ganry, ed.).<br />

Proceedings <strong>of</strong> an International Symposium on Genetic Improvement <strong>of</strong> Bananas for their<br />

Resistance to Diseases and Pests. CIRAD-FLHOR, Montpellier, France.<br />

Stover R.H. and I.W. Buddenhagen. 1986. Banana breeding: polyploidy, disease resistance<br />

and productivity. Fruits 41:175-191.<br />

Swennen R. and D. Vuylsteke. 1990. Aspects <strong>of</strong> plantain breeding at IITA. Pp. 252-266<br />

in Sigatoka <strong>leaf</strong> <strong>spot</strong> disease <strong>of</strong> <strong>bananas</strong> (R.A. Fullerton and R.H. Stover, eds). Proceedings<br />

<strong>of</strong> an international workshopheld at San José, Costa Rica, 28 March-1 April 1989. INIBAP,<br />

Montpellier, France.<br />

Tomekpé K., N. Noupadja, C. Abadie, E. Auboiron and J. Tchango Tchango. 1998. Genetic<br />

improvement <strong>of</strong> plantains at CRBP: performance <strong>of</strong> black Sigatoka resistant plantain<br />

hybrids. Pp. 45-50 in Actas: Seminario Internacional sobre Producción de Plátano.<br />

4-8 de Mayo 1998, Armenia, Quindío. CORPOICA, Colombia.<br />

Vakili N.G. 1967. The experimental formation <strong>of</strong> polyploidy and its effect in the genus Musa.<br />

Amer. J. Bot. 54:24-36.<br />

208


Session 4<br />

R. Swennen et al.<br />

Transgenic approaches for resistance<br />

to <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

in Musa spp.<br />

R. Swennen 1 ,G.Arinaitwe 1 ,B.P.A.Cammue 2 ,I.François 2 ,B.Panis 1 ,<br />

S. Remy 1 ,L.Sági 1 ,E.Santos 1 ,H.Strosse 1 and I. Van den Houwe 1<br />

Abstract<br />

In smallholdings, average banana and plantain yields per unit have not increased significantly<br />

in the last 30 years. Increases in production are due almost exclusively to an increase in the area<br />

under cultivation. Increasing pest and disease pressure, especially from <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>, and<br />

the deteriorating natural resource base have collectively been responsible for these low yields.<br />

Resistant high yielding <strong>bananas</strong> have been bred and supplied to smallholders in the 1990s after<br />

nearly 70 years <strong>of</strong> conventional breeding. This very slow progress was due to the high sterility,<br />

poor seed germination rate, need for interploidy crosses, the long generation cycle, which are<br />

inherent to <strong>bananas</strong> and plantains. A breeding program can only supply a few promising<br />

hybrids per year for further evaluation. The few selected hybrids are high yielding and resistant<br />

to some <strong>diseases</strong> but have usually lost other desired characteristics such as shelf life or pulp<br />

texture. Genetic transformation tools <strong>of</strong>fer an opportunity for plant breeders to overcome the<br />

constraints imposed by the high level <strong>of</strong> sterility <strong>of</strong> the most popular cultivars. Good progress<br />

has been made in the development <strong>of</strong> a molecular toolbox for <strong>bananas</strong> and plantains in the areas<br />

<strong>of</strong> 1) cell suspension, 2) genetic transformation (particle bombardment and Agrobacteriummediated<br />

transformation), 3) high expression <strong>of</strong> foreign genes, 4) insertion <strong>of</strong> multiple genes and<br />

5) identification <strong>of</strong> genes for resistance to fungal disease.<br />

Resumen - Enfoques transgénicos para la resistencia a las enfermedades de las<br />

manchas foliares en banano (Musa spp.)<br />

Durante los últimos 30 años, los rendimientos promedio de los bananos y plátanos no han<br />

aumentado significativamente en las pequeñas fincas y los aumentos de producción se deben<br />

casi exclusivamente a un aumento del área bajo cultivo. El aumento de la presión de plagas y<br />

enfermedades, especialmente de las enfermedades de las manchas foliares, y el deterioro de la<br />

base de recursos naturales han sido responsables de manera colectiva de estos rendimientos tan<br />

bajos. En la década de los 90, se seleccionaron bananos resistentes de alto rendimiento los cuales<br />

1<br />

Laboratory <strong>of</strong> Tropical Crop Improvement, KULeuven, Leuven, Belgium<br />

2<br />

Centre for Microbial and Plant Genetics, KULeuven, Leuven, Belgium<br />

209


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

fueron suministrados a los pequeños productores, después de casi 70 años de mejoramiento<br />

convencional. Este progreso tan lento se debió a una alta esterilidad, una tasa pobre de<br />

germinación de las semillas, una necesidad de cruzamientos interploídicos, un largo ciclo de<br />

regeneración, etc., inherentes a los bananos y plátanos. Básicamente, un programa de<br />

mejoramiento puede proporcionar solo unos pocos híbridos prometedores por año para realizar<br />

las evaluaciones consiguientes. Los pocos híbridos seleccionados son de alto rendimiento y<br />

resistentes a algunas enfermedades, pero usualmente pierden otras características deseadas como<br />

vida verde, textura de la pulpa, etc. Las herramientas de la transformación genética <strong>of</strong>recen una<br />

oportunidad a los fitomejoradores para vencer las limitaciones impuestas por el alto nivel de<br />

esterilidad en las variedades más populares. También se alcanzó un buen progreso en el<br />

desarrollo de una serie de herramientas moleculares para los bananos y plátanos en las áreas<br />

de (1) desarrollo de las suspensiones celulares; (2) tecnologías de transformación genética<br />

(bombardeo con partículas o transformación con Agrobacterium); (3) alta expresión de genes<br />

foráneos; (4) inserción de genes múltiples; (5) identificación de genes para la resistencia a<br />

enfermedades fungosas.<br />

Résumé - Approches transgéniques de la résistance aux maladies foliaires causées<br />

par les <strong>Mycosphaerella</strong> chez les Musa spp.<br />

Dans les exploitations de petite taille, les rendements en bananes et bananes plantain n’ont pas<br />

significativement augmenté au cours des 30 dernières années. L’augmentation de la production est<br />

due presque exclusivement à une augmentation de la surface cultivée. L’accroissement de la pression<br />

des maladies et ravageurs, et particulièrement des maladies foliaires, et la détérioration de la base<br />

de la ressource naturelle ont été collectivement responsables de ces faibles rendements. Des<br />

bananiers résistants et à rendement élevé ont été produits et distribués aux petits producteurs dans<br />

les années 90, après près de 70 ans d’amélioration conventionnelle. Ces progrès très lents sont dus<br />

à la stérilité élevée,au faible taux de germination des semences,au besoin de réaliser des croisements<br />

interploïdes et au long cycle de génération, qui sont propres aux bananiers et aux bananiers<br />

plantain. Un programme d’amélioration ne peut produire que quelques hybrides prometteurs par<br />

an pour leur évaluation ultérieure. Les quelques hybrides sélectionnés ont une production élevée et<br />

sont résistants à certaines maladies mais ont généralement perdu d’autres caractéristiques désirées,<br />

telles que la durée de conservation ou la texture de la pulpe. Les outils de transformation génétique<br />

<strong>of</strong>frent une opportunité aux sélectionneurs de surmonter les contraintes imposées par le niveau élevé<br />

de stérilité des cultivars les plus populaires.Des progrès importants ont été faits dans le développement<br />

d’une boîte à outils moléculaires pour les bananiers et les bananiers plantain dans les domaines :<br />

1) des suspensions cellulaires ; 2) de la transformation génétique (bombardement de particules et<br />

transformation avec Agrobacterium ; 3) du niveau d’expression élevé de gènes étrangers ;<br />

4) de l’insertion de gènes multiples et 5) de l’identification de gènes de résistance aux maladies<br />

fongiques.<br />

Introduction<br />

The predicted increase <strong>of</strong> the world’s population to 8 billion people by 2025 (Harris,<br />

1996) will require developing nations to dramatically increase crop yields.<br />

Technologies such as the application <strong>of</strong> fertilizers or pesticides will have to<br />

contribute, but the most environmentally safe and sustainable approach is the<br />

production and delivery <strong>of</strong> stress resistant high yielding cultivars. Until recently,<br />

new cultivars were produced by cross-breeding or the selection <strong>of</strong> induced or<br />

natural mutations. With the rapid advances in molecular biology, the genetic<br />

modification <strong>of</strong> tropical crops needs to be envisaged to accelerate and focus genetic<br />

improvement.<br />

210


Session 4<br />

R. Swennen et al.<br />

Bananas are one <strong>of</strong> the first domesticated crops (De Langhe and De Maret, 1999).<br />

Some 3000 years ago, between 3 to 8 plantain cultivars were introduced in Africa<br />

(De Langhe et al., 1995) (Mbida et al., 2001). Somatic mutations gave rise to about<br />

120 plantain cultivars (Swennen, 1990) which are all susceptible to black <strong>leaf</strong> streak<br />

disease. Clearly, plantains have a very narrow genetic basis but this is also true for<br />

the entire Musa genus despite the existence <strong>of</strong> about 1200 accessions (Van den houwe<br />

et al., 2000).<br />

Average yields <strong>of</strong> <strong>bananas</strong> and plantains, hereafter called <strong>bananas</strong>, have not<br />

increased significantly in the last 30 years and increases in production are due almost<br />

exclusively to an increase in the area under cultivation. Average yields on<br />

smallholdings remain below 8 t/ha/yr but yields up to 80 t/ha/yr are possible. The<br />

gradual decrease in yields in the major banana growing regions has been attributed<br />

to increased pest and disease pressure and a deteriorating natural resource base. As<br />

a result, many rural communities in Africa are now unable to meet their basic needs<br />

for food and income.<br />

Resistant high yielding <strong>bananas</strong> have been bred and supplied to smallholders in<br />

the 1990s after nearly 70 years <strong>of</strong> conventional breeding (Vuylsteke et al., 1993a,<br />

1993b, 1993c, 1994, 1995; Rowe, 1984; Rowe and Rosales, 1990). This extremely<br />

slow progress is due to high sterility, poor seed germination rate, the need for<br />

interploidy crosses (Swennen and Vuylsteke, 1993; Vuylsteke and Swennen, 1993;<br />

Ortiz and Vuylsteke, 1995) and the long generation cycle. Basically, a breeding<br />

program supplies only a few promising hybrids per year for further evaluation. Only<br />

0.1% <strong>of</strong> the selected hybrids are high yielding and resistant to some <strong>diseases</strong> but<br />

they have lost other desired characteristics such as shelf life or pulp texture. Genetic<br />

transformation tools <strong>of</strong>fer an opportunity for plant breeders to overcome the<br />

constraints imposed by the high level <strong>of</strong> sterility <strong>of</strong> the most popular cultivars<br />

(Swennen, 1994; Sági et al., 1995a, 1995c, 1998a, 1998b).<br />

In this article, we discuss the different molecular tools available for banana<br />

improvement, i.e. 1) embryogenic cell suspensions, 2) gene transfer technologies,<br />

3) expression <strong>of</strong> foreign genes, 4) insertion <strong>of</strong> multiple genes and 5) gene identification.<br />

The current technology has the potential <strong>of</strong> producing several hundreds<br />

transgenic plants per day, in contrast to conventional breeding methods. Possible<br />

future scenarios for the production <strong>of</strong> <strong>bananas</strong> resistant to <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong>, such<br />

as strategies relying on meristem transformation, R-genes from banana and pathogen<br />

inducible promoters, are presented.<br />

Transgenic <strong>bananas</strong> have been available at KULeuven since 1994. Yet, until now<br />

they could not be tested in the field because <strong>of</strong> a lack <strong>of</strong> national laws in African<br />

plantain-growing countries to regulate their release. This causes unnecessary delays<br />

in the further testing, fine-tuning and delivering <strong>of</strong> resistant <strong>bananas</strong> and plantains<br />

to smallholders.<br />

Cell and tissue culture<br />

In many plant species, genetic transformation is very simple and genes are<br />

transferred to callus obtained, for example, from wounded leaves as is the case with<br />

apples (De Bondt 1995). Even simpler is the flower dip method in Arabidopsis, which<br />

211


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

does not necessitate an in vitro process for transformation (Bent, 2000; Clough and<br />

Bent, 1998). In <strong>bananas</strong>, embryogenic cell suspensions are still needed and the<br />

procedure is far from routine (Scho<strong>of</strong>s et al., 1999). Unlike most dicots (De Vries et<br />

al., 1988; Meijer et al., 1999) and seedbearing monocots (Vasil and Vasil, 1986; Vasil,<br />

1987), <strong>bananas</strong> are highly recalcitrant to embryogenesis. Four main procedures have<br />

been developed, each relying on different explants: zygotic embryos (Cronauer and<br />

Krikorian, 1988; Escalant and Teisson, 1989), rhizome slices and <strong>leaf</strong> sheaths (Novak<br />

et al., 1989), immature (fe)male flowers (Escalant et al., 1994; Grapin et al., 1996;<br />

Grapin et al., 1998) and proliferating meristem cultures (Dhed’a et al., 1991; Scho<strong>of</strong>s,<br />

1997).<br />

Most embryogenic suspensions are produced from meristems or flowers, each<br />

method having advantages and disadvantages. For example, the former depends on<br />

extensive preparation <strong>of</strong> material before induction <strong>of</strong> embryogenesis, whereas the<br />

latter requires direct access to flowering banana plants.<br />

At KULeuven the ‘scalp’ method (Scho<strong>of</strong>s, 1997) relies on rapidly proliferating<br />

cultures initiated from a shoot-tip meristem cultured on a medium containing high<br />

levels <strong>of</strong> cytokinin. Embryogenesis-competent scalps contain a high number <strong>of</strong> tiny<br />

white meristems with only a small amount <strong>of</strong> corm or <strong>leaf</strong> tissue. The shoot-tip is<br />

first screened for endophytes and if found positive either cleaned-up or replaced by<br />

an endophyte-free shoot-tip (Van den houwe et al., 1998; Van den houwe and<br />

Swennen, 2000). The scalp method involves: 1) preparation <strong>of</strong> embryogenesis<br />

competent explants (scalps), which takes 5 to 14 months; 2) embryogenesis<br />

induction, which takes 4 to 7 months; and 3) suspension initiation and upgrading,<br />

which takes 3 to 6 months (Swennen et al., 1998). Hence 12 to 27 months, depending<br />

on the cultivar (Scho<strong>of</strong>s, 1997), are needed before a suspension is ready for<br />

transformation.<br />

The production <strong>of</strong> suspensions from East African Highland <strong>bananas</strong> is particularly<br />

cumbersome (Strosse et al. in press, Table 1). A broad range <strong>of</strong> cytokinins at varying<br />

concentrations was explored for scalp induction and it was found that TDZ (thidiazuron)<br />

was a good alternative to BAP (benzylaminopurine) (Table 2 and Figure 1). In fact,<br />

10 mM TDZ could reduce by threefold the embryogenesis induction time (Strosse<br />

et al., in press). The embryogenic response was found to depend on the genotype<br />

(Figures 2 and 3) and even on the selected line and the experiment, and varied between<br />

0 and 22.2% (Strosse et al., in press). Homogeneous complexes consisting <strong>of</strong> a high<br />

proportion <strong>of</strong> embryogenic callus and early-stage transparent embryos are preferred<br />

as inoculum but embryogenic cell suspensions remain more or less heterogeneous<br />

(Georget et al., 2000).<br />

Once cell suspensions are produced, they undergo quality control measurements<br />

at repeated intervals on regeneration potential, health status (Van den houwe et al.,<br />

1998), DNA content (Roux et al., in press a), true-to-typeness, etc. Between 10 4 to<br />

10 5 somatic embryos per ml settled cell volume can be obtained. Hence, a ‘Grande<br />

naine’ cell suspension can produce 14 580 to 100 980 plants while 27 000 to 117 000<br />

plants can be regenerated from an ‘Orishele’ suspension (Strosse et al., in press)<br />

(assuming an inoculum <strong>of</strong> 1.5% settled cell volume in a 60-ml cell suspension<br />

maintenance medium and a tw<strong>of</strong>old increase <strong>of</strong> cell volume after a two-week<br />

subculture). DNA content is assessed through flow cytometry and can show the loss<br />

212


Session 4<br />

R. Swennen et al.<br />

Table 1. Preliminary results using the scalp method (Jan 1998–Dec 2001).<br />

Cultivar Genome Type ITC Number <strong>of</strong> Responsive Frequency Highest EC ECS Ready for<br />

code inoculated scalps* (%)** frequency first applications<br />

scalps in a single<br />

experiment<br />

Calcutta 4 a AA wild diploid 1872<br />

Grande naine AAA Cavendish 1256 2040 112 5.5 11.7 yes yes yes (4) b<br />

GN FHIA AAA Cavendish 1296 11 0.8 2.9 yes yes<br />

GN JD AAA Cavendish 1872 16 0.9 4.2 Yes yes<br />

Williams BSJ AAA Cavendish 0570 456<br />

Williams JD AAA Cavendish 2808 99 3.5 22.2 yes yes yes (7) b<br />

Ingarama AAA-h highland 0160 1104<br />

Mbwazirume AAA-h highland 0084 1128 240<br />

Nyamwihogora AAA-h highland 0086 864<br />

Agbagba AAB-p plantain 0111 576 3 0.5 0.5 yes yes yes f (1) b<br />

Obino l’ewai AAB-p plantain 0109 336 3 0.9 2 yes yes yes (1) b<br />

Orishele AAB-p plantain 0517 1200 29 2.4 5.8 yes yes yes (4) b<br />

Burro cemsa ABB cooking 1259 504<br />

Total 16056 273<br />

Mean 1.1 3.8<br />

* Number <strong>of</strong> scalps forming embryogenic complexes<br />

** % <strong>of</strong> scalps forming embryogenic complexes (RS/NS3)x100, with NS3 being the total number <strong>of</strong> scalps longer than 3.5 months in culture (long enough for first embryogenic complexes to form)<br />

EC Embryogenic complex<br />

ECS Embryogenic cell suspension<br />

a Derived via zygotic embryo rescue. Seeds obtained from IITA, Nigeria<br />

b Number <strong>of</strong> independently established cell suspension lines ready for first applications at the end <strong>of</strong> 2001<br />

213


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Table 2. Preliminary results using TDZ scalps (Jan 1998–Dec 2001).<br />

Cultivar Genome Type ITC code Number <strong>of</strong> NS3* Responsive Frequency (%)*** Highest frequency EC ECS<br />

inoculated scalps scalps** in a single experiment<br />

Calcutta4 a AA wild diploid 96 24<br />

Williams AAA Cavendish 0365 960 720 13 1.8 2.8 yes yes<br />

Igisahira gizanswe AAA-h highland 0083 264 24<br />

Agbagba AAB-p plantain 0111 144 24 1 4.2 4.2 yes<br />

Bluggoe ABB cooking 0010 144 144 4 2.8 5.6 yes<br />

Cachaco b ABB cooking 0643 144 144 4 2.8 5.6 yes<br />

Cachaco ABB cooking 0643 120<br />

Total 1872 1080 22<br />

Mean 1.7 2.3<br />

* Total number <strong>of</strong> scalps longer than 3.5 months in culture (long enough for first embryogenic complexes to form)<br />

** Number <strong>of</strong> scalps forming embryogenic complexes<br />

*** % <strong>of</strong> scalps forming embryogenic complexes (RS/NS3)x100<br />

EC Embryogenic complex<br />

ECS Embryogenic cell suspension<br />

a Derived via zygotic embryo rescue. Seeds obtained from IITA, Nigeria<br />

b Scalps derived from 1 µM TDZ meristem cultures instead <strong>of</strong> 10µM TDZ meristem cultures as in all other cases<br />

214


Session 4<br />

R. Swennen et al.<br />

<strong>of</strong> even 1 chromosome (Dolozel et al., 1999). Rate <strong>of</strong> somaclonal variation can vary<br />

from very low (2%) (Côte et al., 2000) to very high (99%) (data not published). Since<br />

high quality suspensions are rare, they are cryopreserved for backup purposes (Panis<br />

et al., in press; Panis et al., 1990).<br />

Figure 1.<br />

Highly proliferating meristem cultures<br />

<strong>of</strong> ‘Williams’ (AAA) four months after<br />

inoculation <strong>of</strong> a 5-mm explant (apical<br />

dome fully covered by three to four <strong>leaf</strong><br />

primordia and a few millimeters <strong>of</strong> corm<br />

beneath the apical dome) on MS based<br />

medium supplemented with 10 µM TDZ,<br />

bar = 769 µm.<br />

Figure 2.<br />

Highly regenerable embryogenic<br />

cell suspensions <strong>of</strong> ‘Grande naine’<br />

(AAA). Embryogenic cell clusters<br />

observed with (left) light microscope,<br />

bar = 95 µm and (right)<br />

germinating embryos one month<br />

after culturing on regeneration<br />

medium, bar = 370 µm.<br />

Figure 3.<br />

Highly regenerable embryogenic<br />

cell suspensions <strong>of</strong> ‘Orishele’(AAB).<br />

Embryogenic cell clusters observed<br />

with (left) light microscope, bar =<br />

95 µm and (right) germinating<br />

embryos one month after culturing<br />

on regeneration medium, bar =<br />

370 µm.<br />

Gene transfer methods<br />

Regenerable embryogenic cell suspension (ECS) cultures are the material <strong>of</strong> choice<br />

(Dhed’a et al., 1991; Escalant et al., 1994; Côte et al., 1996, 1997) for the genetic<br />

engineering <strong>of</strong> <strong>bananas</strong> via particle bombardment-mediated transformation (PMT)<br />

and Agrobacterium-mediated transformation (AMT) (Sági et al., 2000). The former<br />

uses the biolistic gun device (Sági et al. 1995a), whereas the latter uses cocultivation<br />

215


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

with Agrobacterium (Pérez Hernández, 2000). Different methods and cultivars are<br />

being used (Sági et al., 1995a; May et al., 1995; Remy et al., 1998a, 1998b; Becker<br />

et al., 2000; Ganapathi et al., 2001). Our comparative study involved ECS <strong>of</strong> four<br />

cultivars in the exponential growth phase (4+1 days after subculture). DNA plasmid,<br />

PMT and plant regeneration were according to Sági et al. (1995a, 1995b). For AMT<br />

transformation, evaluation <strong>of</strong> transient gene expression and selection and<br />

regeneration <strong>of</strong> transformants was according to Pérez Hernández (2000), Pérez<br />

Hernández et al. (1999) and Arinaitwe et al. (in press).<br />

Results indicate that the efficiency <strong>of</strong> the two gene transfer methods is quite similar<br />

at the transient level with different promotors. In contrast, AMT was more efficient<br />

than PMT after selection (Arinaitwe et al., in press) (Table 3). The higher number <strong>of</strong><br />

plants regenerated using the AMT system in comparison with the PMT system<br />

confirmed AMT as the method <strong>of</strong> choice for transforming plant cells, as reported by<br />

Newell (2000).<br />

Table 3. Shoot regeneration <strong>of</strong> four cultivars using the AMT and PMT methods.<br />

Cultivar AMT PMT<br />

Grande naine 117 76<br />

Obino l’ewai 118 41<br />

Orishele 93 05<br />

Three hand planty 96 98<br />

ECSs <strong>of</strong> four cultivars were co-cultivated with an Agrobacterium tumefaciens<br />

strain: AGLO harbouring the binary plasmid pUbi-sgfpS65T; and EHA101 harbouring<br />

the binary plasmid pFAJ3000. Plasmid pFAJ3000 contains a gusA (ß-glucuronidase)<br />

gene driven by the CaMV 35S promoter and a neo gene under the control <strong>of</strong> the<br />

NOS promoter. Plasmid pUbi-sgfpS65T contains a gfp (green fluorescent protein)<br />

gene driven by the ubiquitin promoter (Arinaitwe et al., in press). There was a<br />

difference in the expression <strong>of</strong> the two reporter genes used (Figure 4). This was,<br />

probably, due to differences in efficiency <strong>of</strong> the two A. tumefaciens strains used<br />

and the variable embryogenesis.<br />

The effect <strong>of</strong> infection time on transformation frequency was investigated in<br />

‘Grande naine’ and ‘‘Three hand planty’ (AAB) by using transient gus expression<br />

(TGE) and transient green fluorescent protein expression (TGFPE). In both cultivars<br />

and both marker genes, transient expression increased with increasing infection time<br />

(Table 4). With ‘Grand naine’, maximum transient expression was reached after<br />

8 and 12 hours for TGE and TGFPE, respectively. More TGE was observed in Three<br />

hand planty’ than ‘Grande naine’, possibly due to differences in the quality <strong>of</strong> the<br />

cell line.<br />

Variable volumes (ml) <strong>of</strong> ECS were plated and uniformly spread over a 50-mm<br />

nylon mesh. Transient GFP expression indicates that T-DNA transfer was highest<br />

at 100±50 ml (Arinaitwe et al., in press) but dropped sharply when the volume <strong>of</strong><br />

ECS was increased to 300 and 600 ml. A decreased attachment and access to<br />

individual embryogenic cells or cell clusters by Agrobacterium is considered to be<br />

the cause. High TGFPE in small volumes is attributed to increased exposure <strong>of</strong> ECSs<br />

216


Session 4<br />

R. Swennen et al.<br />

% Regeneration<br />

60<br />

50<br />

40<br />

30<br />

20<br />

GUS<br />

GFP<br />

10<br />

0<br />

Grande naine<br />

Three hand planty<br />

Obino l'ewai<br />

Orishele<br />

Figure 4. Percentage <strong>of</strong> regenerated shoots <strong>of</strong> four cultivars transformed via Agrobacterium (AGLO, pUbisgfpS65T<br />

with gfp gene; EHA 101, pFAJ 3000 with gusA gene).<br />

to Agrobacterium since efficient spreading <strong>of</strong> thin layers <strong>of</strong> cells is achieved during<br />

the co-cultivation phase.<br />

Upon subculture an ECS starts to multiply but its growth (Scho<strong>of</strong>s et al. 1999)<br />

and cell cycle (Roux et al. in press b) changes with age. Hence an effect <strong>of</strong> ECS age<br />

on transformation frequency is expected and could be confirmed (Figure 5). Cell<br />

competence for transformation increased from day 1 until day 7, beyond which it<br />

dropped. This period is thought to coincide with the exponential growth phase <strong>of</strong><br />

the ECS (Sági et al. 1995a, b). Efficient transformation <strong>of</strong> 7-day-old ECSs has been<br />

reported in cultivar ‘Rasthali’ (AAB) (Ganapathi et al. 2001).<br />

Number <strong>of</strong> blue foci<br />

2500<br />

2000<br />

1500<br />

1000<br />

1 st trial<br />

2 nd trial<br />

500<br />

0<br />

1 3 5 7 9<br />

Age (days)<br />

Figure 5. Effect <strong>of</strong> age <strong>of</strong> embryogenic cell suspension on transformation frequency: transient gus expression<br />

in cultivar ‘Obino l’ewai’ transformed via Agrobacterium (EHA 101; pFAJ 3000) (n=3).<br />

217


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Table 4. The effect <strong>of</strong> infection time on transient gene expression frequency (Mean±SE <strong>of</strong> the number <strong>of</strong> <strong>spot</strong>s <strong>of</strong><br />

gene expression per cell sample <strong>of</strong> 50 mg (fresh weight); at least 4-5 replications).<br />

Marker Cultivar Infection time (hrs)<br />

gene 4 6 8 10 12 14<br />

gus Grande naine 794±87.9 881±91.8 >1500 >1500 >1500 >1500<br />

gfp Grande naine 209.7±24 272±41 365.7± 28 922.7±21 >1500 >1500<br />

gfp Three hand planty 1169.3±150 1311.7±95 >1500 >1500 >1500 >1500<br />

The combined use <strong>of</strong> these and other factors resulted in a five-fold increase in<br />

transient expression compared to the original procedure. Representative results <strong>of</strong><br />

these experiments are shown in Figure 6. One hundred mg (fresh weight) <strong>of</strong> control<br />

banana cells showed no background transient expression <strong>of</strong> the gus reporter gene<br />

(Figure 6A). In contrast, an average <strong>of</strong> 1500 blue foci was observed in the same<br />

amount <strong>of</strong> a cell line <strong>of</strong> the dessert banana ‘Grande naine’ (Figure 6B) using the<br />

improved method, in comparison with about 250 blue foci using the standard<br />

protocol. The uniform distribution <strong>of</strong> the transiently transformed cells also indicates<br />

the high efficiency <strong>of</strong> the improved procedure. Similarly increased transient gus<br />

expression rates have been observed in several <strong>bananas</strong> cultivars. Experiments are<br />

now in progress to determine if increased transient gene expression improves the<br />

yield <strong>of</strong> transgenic plants.<br />

Figure 6. Transient GUS expression in a cell suspension culture <strong>of</strong> ‘Grande naine’ after co-cultivation<br />

with Agrobacterium tumefaciens EHA105 harbouring the GUS-intron containing binary vector pFAJ3000.<br />

A) 100 mg control cells showing no GUS expression, B) the same amount <strong>of</strong> transformed cells after six days<br />

<strong>of</strong> co-cultivation using the improved method.<br />

Following confirmation that banana could be transformed (Sági et al., 1995a;<br />

May et al., 1995), several groups looked for suitable promoters to improve the<br />

expression <strong>of</strong> heterologous genes. More than 25 heterologous and 1 homologous<br />

promoters have been tested (Table 5). In general strong transcription is obtained when<br />

genes are driven by the constitutive promoters such as the promoter from the maize<br />

ubiquitin, or the rice actin gene and promoters from the pregenomic RNA <strong>of</strong> banana<br />

streak badnavirus. Few tissue-specific promotors have been identified in banana but<br />

some promoter regions from the banana bunchy top nanavirus (BBTV) seem to have<br />

a potential for expression in vascular tissue.<br />

218


Session 4<br />

R. Swennen et al.<br />

Table 5. Promoters used in banana genetic transformation.<br />

Promoter Source <strong>of</strong> the promoter Reference<br />

CaMV35S Cauliflower Mosaic Virus Sági et al., 1992; Sági et al., 1994; Sági et al., 1995b;<br />

Sági et al., 1995c; CIRAD, 2001; Dugdale et al., 1998;<br />

Pérez Hernández et al., 1998; Sági, 1998; Más et al.,<br />

1999; Schenk et al., 1999; Becker et al., 2000;<br />

Dugdale et al., 2001; Schenk et al., 2001<br />

CaMV35S a Cauliflower Mosaic Virus Más et al., 2000<br />

35S-AMV Cauliflower Mosaic Virus Sági et al., 1995c<br />

Alfalfa Mosaic Virus<br />

35S-35S b Cauliflower Mosaic Virus Moy et al., 1998; Moy et al., 1999; Schenk et al., 2001;<br />

Remy et al., 1998b; Sági et al., 1995a;<br />

Sági et al., 1995b; Sági et al., 1995c<br />

35S-35S-AMV Cauliflower Mosaic Virus Sági et al., 1994; Sági et al., 1995a; Sági et al., 1995b;<br />

Alfalfa Mosaic Virus<br />

Sági et al., 1995c; Sági et al., 1998a<br />

Emu c Recombinant ARE-ocs-adh1 Sági et al., 1995a; Sági et al., 1995b;<br />

Sági et al., 1998a; Remy, 2000<br />

Act-1 Rice actin gene act1D May et al. 1995; Sági et al., 1998a<br />

Act-1 Banana actin gene Hermann et al., 2001b<br />

Ubi Maize ubiquitin Grapin, 1995; Sági et al., 1995a; Grapin et al., 1996;<br />

CIRAD, 2001; Sági et al., 1998a; Dugdale et al., 1998;<br />

Moy et al., 1998; Remy et al., 1998b; Moy et al., 1999;<br />

Schenk et al., 1999; Becker et al., 2000;<br />

Pérez Hernández, 2000; Ganapathi et al., 2001;<br />

Schenk et al., 2001<br />

(ocs) 3<br />

mas d Recombinant ocs-mas Remy et al., 1998b; Moy et al., 1998; Moy et al., 1999;<br />

Remy, 2000; Ganapathi et al., 2001<br />

Sc Sugarcane bacilliform badnavirus Schenk et al., 1999<br />

(ScBV)<br />

Badnavirus Banana streak badnavirus Sági, 1998a; Schenk et al., 2001; Remans et al. 2000<br />

My<br />

My from Mysore cv. BSV<br />

Cv<br />

Cv from Cavendich cv. BSV<br />

Go<br />

Go from Goldfinger cv. BSV<br />

BBTV DNA 1 to 6 Banana Bunchy Top Virus Dugdale et al., 1998; Becker et al., 2000;<br />

S1<br />

Hermann et al., 2001a<br />

S2<br />

BT1, BT2, BT3, BT4, BT5 e Dugdale et al., 2000<br />

BT6.1 f Dugdale et al., 2001<br />

a<br />

OCS enhancer plus Cauliflower Mosaic Virus promoter plus rice act1 untranslated sequence<br />

b<br />

Enhanced Cauliflower Mosaic Virus promoter<br />

c<br />

Six copies <strong>of</strong> the 41-bp ARE (anaerobic responsive element) plus four copies <strong>of</strong> the 40-bp ocs (octopine synthase) enhancer plus the 5’<br />

end <strong>of</strong> a truncated adh1 promoter linked to its first intron<br />

d<br />

Atandem <strong>of</strong> three upstream activating sequences (UAS) <strong>of</strong> the octopine synthase gene (ocs) and a promoter/activator region <strong>of</strong> the mannopine<br />

synthase gene (mas)<br />

e<br />

Banana Bunchy Top Virus DNA 1-5 intergenic regions plus maize ubiquitin (ubi1) intron<br />

f<br />

Banana Bunchy Top Virus DNA-6 intergenic region plus intron mediated enhancement <strong>of</strong> maize ubi1, maize adh1, rice act1 and sugarcane<br />

rbcs genes<br />

Transformation for resistance to <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

Plants have developed a range <strong>of</strong> defense mechanisms against pathogens such<br />

as, the rapid death <strong>of</strong> the first infected cells (Colligne and Slusarenko, 1987) which<br />

prevents further pathogen spread. Other defense mechanisms include increased<br />

219


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

lignification <strong>of</strong> the cell wall (Vance et al., 1980), the synthesis <strong>of</strong> phytoalexines<br />

(Hahlbrock and Scheel, 1989) and the production <strong>of</strong> reactive oxygen species (Mehdy,<br />

1994). The biochemical complexity <strong>of</strong> these mechanisms makes it difficult to develop<br />

molecular methods to breed for fungus resistance.<br />

Of the large number <strong>of</strong> genes activated upon pathogen recognition by the plant<br />

is a group that encodes for pathogenesis-related (PR) proteins (Van Loon et al., 1987;<br />

Linthorst, 1991). PR proteins have been defined as proteins encoded by the host plant<br />

but induced only under pathological or related stress conditions (Antoniw et al.,<br />

1980). To date, more than 10 families <strong>of</strong> PR proteins have been classified, which<br />

include chitinases. In addition, plants express numerous other genes that encode for<br />

proteins with antimicrobial properties (Broekaert et al., 1997; García-Olmedo et al.,<br />

1998). These are small, stable and cysteine-rich peptides isolated from seeds <strong>of</strong> diverse<br />

plant species (Broekaert et al., 1992; Cammue et al., 1992; Terras et al., 1992a, b;<br />

Cammue et al., 1995; Osborn et al., 1995). These plant defensins, or antifungal<br />

peptides, are highly active against a broad spectrum <strong>of</strong> phytopathogenic fungi. For<br />

example, two <strong>of</strong> these proteins (Rs-AFP1 and Rs-AFP2) have been isolated from radish<br />

(Raphanus sativus) seeds. The latter appears to be the most potent with IC 50<br />

values<br />

ranging from 0.4 to 25 µg/mL (Terras et al., 1992a, b). Several <strong>of</strong> these antifungal<br />

peptide genes were shown to inhibit the growth <strong>of</strong> <strong>Mycosphaerella</strong> spp. under in<br />

vitro conditions (Cammue et al., 1993) but were not toxic to human fibroblasts and<br />

erythrocytes (Terras et al., 1992a, b; Cammue et al., 1995) or to banana cells (Cammue<br />

et al., 1993). We therefore focused our work on the insertion <strong>of</strong> these antifungal<br />

or antimicrobial proteins in banana with a different mode <strong>of</strong> action. Morphogenic<br />

defensins, like Rs-AFP2 from Raphanus sativus, cause a reduction in hyphal<br />

elongation and an increase in hyphal branching, whereas non-morphogenic ones,<br />

such as Dm-AMP1 from Dahlia merckii (Osborn et al. 1995), slow down hyphal<br />

elongation without a visible morphological effect.<br />

Since it was demonstrated in tobacco that disease resistance can be increased by<br />

simultaneously integrating different antifungal proteins (Jach 1995), the frequency<br />

<strong>of</strong> co-transformations with particle bombardment was evaluated in three independent<br />

experiments (Remy et al. 1998a). In experiment 1, the selectable marker gene (gene<br />

A) and an antifungal peptide gene (gene B) were introduced into embryogenic cells<br />

<strong>of</strong> the plantain cultivar ‘Three hand planty’ in a linked position, i.e. the two genes<br />

were present on the same plasmid. In experiments 2 and 3, the plasmid with genes<br />

A and B were co-transformed with another plasmid that carried a different<br />

antifungal protein gene (gene C) which was thus not linked to genes A or B.<br />

Transgenic shoots were then regenerated from all three experiments and analysed<br />

by PCR for the presence <strong>of</strong> each foreign gene. Integration <strong>of</strong> these genes was also<br />

confirmed by Southern gel blot hybridisation in a number <strong>of</strong> selected plants from<br />

each experiment.<br />

The number <strong>of</strong> plants carrying both gene A and gene B or C was used to calculate<br />

the co-transformation frequencies <strong>of</strong> linked genes and according to the following<br />

equation:<br />

{No. <strong>of</strong> (A+B or C) + / No. <strong>of</strong> A + } x 100<br />

As one can expect, the linked genes co-existed in the transgenic plants at a high<br />

frequency that ranged between 90 to 100% in the different experiments. Similarly,<br />

220


Session 4<br />

R. Swennen et al.<br />

as expected, the unlinked gene showed a lower co-transformation frequency than<br />

the linked genes. However, this frequency was still remarkably high, in the range<br />

<strong>of</strong> 70 to 80%, probably due to efficient co-precipitation <strong>of</strong> the two plasmids on<br />

microparticles. Similarly to the results obtained by Hadi et al. (1996), this observation<br />

indicates that simultaneous bombardment <strong>of</strong> different plasmid molecules may be a<br />

convenient way for the introduction <strong>of</strong> multiple genes into crop plants.<br />

Co-transformation frequency with Agrobacterium was also estimated with one<br />

<strong>of</strong> the several possible methods. Two Agrobacterium strains were used to introduce<br />

two reporter genes (gfp and luc) into <strong>bananas</strong>. The elegance <strong>of</strong> this experimental<br />

setup is the simultaneous and live observation <strong>of</strong> gene expression by imaging.<br />

This way, co-transformation frequencies were directly measured by co-expression<br />

frequency. As was expected, co-transformation frequency was relatively low because<br />

<strong>of</strong> the low probability that two bacterial cells will deliver their T-DNA molecules to<br />

the same plant cell. In our case, the average frequency <strong>of</strong> gfp and luc cotransformation<br />

in four cultivars was around 3% after three weeks and 4% two months<br />

after transformation (Ahmed, unpublished data).<br />

The expression <strong>of</strong> antimicrobial peptides in transgenic <strong>bananas</strong> was analysed<br />

by ELISA using specific antibodies. Out <strong>of</strong> more than 150 single transformants,<br />

i.e. transgenic plantains expressing only one antimicrobial peptide, more than 10%<br />

had a relatively high concentration (between 0.05-0.12% <strong>of</strong> total soluble protein)<br />

in the <strong>leaf</strong>. In contrast, out <strong>of</strong> 16 double transformants, i.e. plants expressing a<br />

Dm-AMP1 and another antimicrobial peptide from onions (Ac-AMP1), 6 (37%)<br />

accumulated one or both peptides to at least four times the background level<br />

(Remy, 2000).<br />

In order to assess tolerance to fungus in transgenic lines, a simple, sensitive and<br />

reproducible <strong>leaf</strong> disc bioassay has been developed (Remy, 2000). A 5-cm <strong>leaf</strong> disc<br />

was excised from transgenic plants grown in a greenhouse and inoculated in situ<br />

with fungi. Four days after infection, a differential disease response was observed<br />

between independent transformants, whereas a non-pathogenic fungus (e.g. Fusarium<br />

sp.) was unable to induce disease symptoms, indicating that the assay is specific for<br />

host-pathogen interactions. Transformations with different promoter-gene constructs<br />

resulted in a wide range <strong>of</strong> tolerance to fungus among the independent transgenic<br />

plants. Computer image capturing and s<strong>of</strong>tware-based area calculation has been used<br />

to precisely measure the area <strong>of</strong> infected <strong>leaf</strong> and to classify independent<br />

transformants according to their tolerance. This procedure was used to screen 42<br />

independent transgenic plantain lines expressing Ac-AMP1 (Pérez Hernández,<br />

2000), among which 6 lines had 2 to 3 times less necrosis upon infection with<br />

Colletotrichum musae than the untransformed controls.<br />

Towards an improved transformation technology<br />

Efficient transformation techniques exist in banana but they rely on labourintensive<br />

cell suspension technology. Moreover, controlled transgene expression needs<br />

to be developed for banana relying, among others, on native or heterologous<br />

developmental and tissue specific promoters and especially pathogen-inducible ones.<br />

This is needed because constitutive resistance leads to a decrease in fitness (Heil and<br />

221


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Baldwin, 2002). However, relatively few promoters are available for the regulated<br />

control <strong>of</strong> target gene expression.<br />

Meristem transformation<br />

Existing transformation technologies rely on cell suspensions, an elaborate, expensive<br />

and time-consuming process. A transformation procedure based instead on tissue<br />

culture <strong>of</strong> cell cultures would be more efficient and less dependent on the genotype.<br />

To this end, more than 12 000 explants <strong>of</strong> ‘Grande naine’ and ‘Williams’ were infected<br />

with Agrobacterium with or without sonication. Selection was performed either on<br />

a liquid or solid medium using different growth regulators. Independently <strong>of</strong><br />

treatment, the frequency <strong>of</strong> putative transgenic cultures was about 0.5% in both<br />

cultivars after 2 to 3 months. This consistently low transformation frequency indicates<br />

that proliferating meristematic cultures may not be the ideal target for genetic<br />

transformation. This could be attributed to the heterogenous structure <strong>of</strong> the<br />

meristem containing few embryogenic cells. Meristem transformation may produce<br />

chimeras that are not easily identified or dissociated (Roux et al., 2001). In addition,<br />

based on DNA behaviour in mouse embryonic stem cells (Lei et al., 1996),<br />

embryogenic plant tissues in their early development may be more suitable for gene<br />

targeting than tissues that underwent several divisions and differentiation (Kumar<br />

and Fladung, 2001). However, given the little effort directed to the optimization <strong>of</strong><br />

meristem transformation in banana, this avenue should be further explored.<br />

Positive selectable marker genes<br />

Once foreign genes are delivered to banana suspensions, transformed cells need to<br />

be harvested. With PMT, about 1 to 7 cells per 100 µl <strong>of</strong> a 33% settled cell volume<br />

suspension (± 25 mg fresh weight cells) are transformed, and up to 100 transgenic<br />

cells from twice the same volume with AMT. Because <strong>of</strong> this very low transformation<br />

frequency, selectable marker genes, e.g. the neomycin phosphotransferase gene (Fraley<br />

et al., 1983) which confers resistance to aminoglycoside-type antibiotics such as<br />

kanamycin, neomycin and G-418 (geneticin), are used. Occasionally herbicide<br />

resistance genes are also used. These are negative selection systems. In banana,<br />

antibiotic resistance genes should pose no concern to the environment because there<br />

is no pollen in edible <strong>bananas</strong>, yet there is concern that such genes in genetically<br />

modified food organisms pose a hazard to human health (Fuchs et al., 1992). Research<br />

is conducted to completely remove selectable marker genes (Puchta 2000).<br />

In positive selection systems, transformed cells can convert a physiologically<br />

inert substance into a compound that stimulates growth. Hence, transgenic cells<br />

overgrow non-transformed cells that are starved rather than killed, e.g. the gusA<br />

(ß-glucuronidase) gene from Escherichia coli that hydrolyzes benzyladenineglucuronide<br />

(Okkels et al. 1997) into active cytokinin and thus stimulates growth<br />

<strong>of</strong> transgenic cells. Other examples include the use <strong>of</strong> phosphomannose isomerase<br />

(PMI, or mannose-6-phosphate isomerase), an enzyme catalyzing the reversible<br />

isomerization <strong>of</strong> mannose-6-phosphate to fructose-6-phosphate, which serves as a<br />

precursor for the glycolytic pathway, and xylose isomerase (or D-xylose ketol-<br />

222


Session 4<br />

R. Swennen et al.<br />

isomerase) that interconverts D-xylose to D-xylulose, which after phosphorylation<br />

by xylulokinase enters the pentose phosphate pathway. The former enzyme is used<br />

to give transgenic cells a growth advantage over non-transgenic ones because most<br />

plant species cannot metabolize mannose (or mannose-6-phosphate) and in plant<br />

tissue cultures mannose has been known to be unable to support growth (Malca et<br />

al., 1967). The manA gene <strong>of</strong> E. coli (Miles and Guest, 1984), which codes for PMI,<br />

has been successfully used in transformation experiments with cassava, maize, rice,<br />

sugar beet and wheat (see Suprasanna et al., 2002 for references). No adverse<br />

nutritional effects were found (Reed et al., 2001). Xylose, like mannose, cannot be<br />

metabolized by plant cells, whereas they can utilize D-xylulose because they express<br />

xylulokinase. So far, transformation with the xylose isomerase gene (xylA) (Wong<br />

et al., 1991; Lee et al., 1990) as a selectable marker was demonstrated in potato,<br />

tobacco and tomato (Haldrup et al., 1998a, b). This enzyme is widely recognized as<br />

safe, since it is commercially used in the starch industry and for food processing.<br />

The use <strong>of</strong> positive selectable markers genes has the additional advantage that<br />

it can increase the transformation frequency dramatically because no toxic substances<br />

are released from dying cells. Positive selection systems have proven their value in<br />

several crops (sugar beet, cassava, maize, wheat and rice). For an overview the reader<br />

is referred to Suprasanna et al. (2002). These novel selectable marker systems are<br />

being tested in banana in order to increase the transformation frequency but also<br />

to allow repeated transformation operations, and minimize the use <strong>of</strong> antibiotic and<br />

herbicide selectable marker genes.<br />

Tagging, isolation and characterization <strong>of</strong> novel promoters<br />

and genes<br />

The isolation <strong>of</strong> promoters <strong>of</strong> differentially expressed or inducible genes can be<br />

accomplished indirectly or directly. In the first approach, the promoter is isolated<br />

in parallel or subsequently to the characterization <strong>of</strong> a gene <strong>of</strong> interest by molecular<br />

techniques. However, when dealing with multigene families and pseudogenes or<br />

with genes that are developmentally regulated or exert a cell-specific pattern <strong>of</strong><br />

expression this approach is likely to be extremely difficult.<br />

Promoters can also be identified directly within the genome via tagging by<br />

transformation with a promoterless reporter gene and screening for individual<br />

transformants, in which reporter gene expression is activated. After plasmid rescue<br />

<strong>of</strong> the respective region from the genome or via direct genome walking (e.g. by<br />

inverse PCR or various anchored PCR techniques), the promoter region is isolated<br />

and sequenced. Via a combined screening <strong>of</strong> a population <strong>of</strong> transgenic plants for<br />

different parameters (e.g. abiotic and biotic stress factors, development and tissuespecific<br />

expression), several promoters for various genes can simultaneously be<br />

identified and thoroughly characterized without the a priori isolation and analysis<br />

<strong>of</strong> the corresponding coding sequence(s).<br />

At KULeuven, several thousands <strong>of</strong> transgenic cultures can be produced in a<br />

relatively short time by using Agrobacterium-mediated transformation <strong>of</strong><br />

embryogenic cell suspension cultures. Such a large number <strong>of</strong> transgenic cultures<br />

provides a reasonable chance <strong>of</strong> tagging interesting promoters. At present, the target<br />

223


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

cultivars are ‘Grande naine’ and ‘Three hand planty’. The tagging construct<br />

contains a promoterless luciferase gene linked to a T-DNA border and, so far,<br />

approximately 2500 and 4000 transgenic cultures respectively have been selected<br />

after transformation. The expression <strong>of</strong> the luciferase gene (luc) (De Wet et al., 1985)<br />

is not destructive and therefore provides the opportunity for continuous detection<br />

<strong>of</strong> reporter gene activity in plants using low-light imaging techniques (Ow et al.,<br />

1986; Millar et al., 1992; Chia et al., 1994). Screening for activated luciferase<br />

expression has been carried out under a liquid nitrogen-cooled CCD camera coupled<br />

to a sophisticated image capture and analysis system (Remy et al., in press). A liquid<br />

nitrogen cooling system is used since it reduces the dark current to less than<br />

1 electron per pixel per hour allowing long exposures <strong>of</strong> up to tens <strong>of</strong> minutes.<br />

LUC activity could be detected 80 minutes after bombardment and was clearly<br />

visible 40 minutes later with the codon-modified luc+gene (Sherf and Wood, 1994).<br />

The luc+gene showed a much higher and faster LUC activity than the wild type<br />

luc gene (Remy et al., in press).<br />

As a wide range <strong>of</strong> LUC activities was detected (from less than 5 to more than<br />

300 relative grey levels/pixel), it is clear that this simple, fast and sensitive in vivo<br />

reporter gene assay can become a valuable tool in gene expression studies <strong>of</strong><br />

<strong>bananas</strong>. In parallel with the multiplication <strong>of</strong> the tagged population, preliminary<br />

screenings have been performed for constitutive activation and for promoters<br />

inducible by temperature shock, salt stress and herbicide treatment. The frequency<br />

<strong>of</strong> cultures with detectable (constitutive) activation has been around 10% in different<br />

tagging experiments with ‘Three hand planty’ (Figures 7A and B). On the other<br />

hand, as expected, the frequency <strong>of</strong> inducible activation by specific conditions has<br />

so far been well below 1% (Figure 7C and D for salt-induced activation).<br />

Figure 7.<br />

Luciferase imaging <strong>of</strong> transgenic<br />

cultures <strong>of</strong> the plantain cultivar<br />

‘Three hand planty’.<br />

A) Light image <strong>of</strong> a 24-well plate with<br />

cultures transformed with a 35Sluciferase<br />

construct for constitutive<br />

expression.<br />

B) The same cultures screened for<br />

luciferase expression.<br />

C) Light image <strong>of</strong> a 24-well plate<br />

with cultures transformed with a<br />

promoterless luciferase construct for<br />

tagging.<br />

D)The same cultures screened for<br />

luciferase activation after salt stress<br />

(arrow indicates a positive culture).<br />

In future, selected plants will be analysed by southern hybridization, as well as<br />

via a novel anchored PCR technique, to screen for individuals containing single<br />

insertions. Then, TAIL-PCR will be used to recover and clone the plant DNA flanking<br />

the luciferase gene. The cloned fragments will be sequenced, compared to sequence<br />

databases and analysed with standard bioinformatic tools for conserved regions and<br />

224


Session 4<br />

R. Swennen et al.<br />

putative transcription factor binding sites. Based on the sequence obtained, internal<br />

fragments <strong>of</strong> each promoter candidate will be hybridized to DNA purified from the<br />

original transformant as well as an untransformed control to confirm their genuine<br />

nature. Finally, for functional testing promoter-gfp fusions will be re-introduced<br />

to banana to confirm the constitutive or wound and pathogen pattern <strong>of</strong> expression<br />

as well the tissue specificity in an ectopic situation.<br />

Confirmed promoters will then be utilized for disease control, for example with<br />

antifungal genes, and associated genes isolated and characterized to understand<br />

their role in plant-pathogen interactions. Except for promoter tagging <strong>of</strong> genes<br />

related to nematode feeding structures (Bartels et al., 1997; Puzio et al., 1999), this<br />

method has not been applied to studies on plant-pathogen interactions. Since<br />

salicylic acid (SA) and its analogs play an important role in the defense response<br />

<strong>of</strong> many plant species to pathogen attack, promotorless luc tagging in banana would<br />

be useful to unravel upregulated genes that are involved in widely different<br />

metabolic pathways including pathogen defense. SA treatment mimics osmotic and<br />

oxidative stress, mediates the oxidative burst that leads to cell death in the<br />

hypersensitive response, and acts as a signal for the development <strong>of</strong> systemic acquired<br />

resistance (Shirasu et al., 1997).<br />

Resistance genes are <strong>of</strong>ten proposed to control pathogen attack. Based on a<br />

“guard model” it is proposed that R genes, which are generated randomly, most<br />

likely through a birth-and-death process (Michelmore and Meyers, 1998), stand a<br />

better chance to induce resistance if identified from the plant family <strong>of</strong> which a<br />

certain cultivar needs to acquire additional resistance (Van der Hoorn et al., 2002).<br />

Resistance genes from banana are currently unavailable for banana transformation<br />

although techniques are available and some sources do exist. A series <strong>of</strong> resistance<br />

gene analogs (RGAs) were isolated, using degenerate PCR primers targeting highly<br />

conserved regions in proven plant resistance genes (e.g. leucine-rich repeat<br />

sequences) (Wiame et al., 2000). For an overview <strong>of</strong> the current situation and a<br />

proposed strategy to correct this situation, the reader is referred to Kahl (in press).<br />

In any case, there is a need not to focus only on a few resistance genes but on the<br />

simultaneous detection, identification and quantification <strong>of</strong> all transcripts at a given<br />

time and monitoring <strong>of</strong> gene expression patterns at various developmental stages<br />

or after specific treatments (Matsumara et al., 1999) correlated to physiological or<br />

developmental processes. Much is expected from The Global Musa Genomics<br />

Consortium.<br />

Increased expression<br />

The search in badnaviruses for useful promoters to drive transgene expression in<br />

banana (Schenk et al., 1999) resulted in two novel DNA fragments <strong>of</strong> 2105 bp (My)<br />

and 1322 bp (Cv) amplified from the upstream region <strong>of</strong> the coding sequence <strong>of</strong><br />

two Australian banana streak badnavirus (BSV) isolates. Evaluation <strong>of</strong> the My and<br />

Cv promoters in transgenic banana demonstrated that these promoters could drive<br />

high-level expression <strong>of</strong> either the gusA or the gfp reporter gene in different tissues<br />

during vegetative development. For instance, gus activity in transgenic in vitro plants<br />

<strong>of</strong> the plantain ‘Three hand planty’ containing the My promoter were up to seven<br />

225


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

times higher in <strong>leaf</strong> tissue and up to four times stronger in root and corm tissue<br />

compared to plants harbouring the maize ubiquitin promoter (Schenk et al., 2001).<br />

The Cv promoter showed activities that were similar to the maize ubiquitin<br />

promoter in transgenic in vitro plants, but was significantly reduced in larger<br />

glasshouse-grown plants.<br />

Enhancement <strong>of</strong> transgene expression levels by translational fusion <strong>of</strong> transgenes<br />

has been observed by different research groups. In an experiment, the sequence<br />

coding for a naturally occurring plant linker peptide was used to connect the<br />

sequence coding for two antimicrobial proteins (AMPs) in a polyprotein construct<br />

that was transformed to Arabidopsis thaliana (François et al., 2002a, 2002b). The<br />

linker peptides used were based on the fourth linker peptide <strong>of</strong> the IbAMP<br />

polyprotein precursor isolated from the seed <strong>of</strong> Impatiens balsamina (Tailor et al.,<br />

1997). The heterologous polyprotein precursors were demonstrated to be cleaved<br />

post-translationally in A. thaliana thereby releasing the two AMPs (François et al.,<br />

2002a, 2002b). Cleavage appeared to be complete as no immunoreactive polyprotein<br />

precursor could be detected in the transformed A. thaliana plants. A striking<br />

observation from the experiments was that the expression levels <strong>of</strong> the first protein<br />

were several times higher in plants transformed with the polyprotein constructs<br />

compared to plants transformed with the single protein construct. Expression levels<br />

as high as 3.1% <strong>of</strong> total protein content, as seen in some lines transformed with<br />

polyprotein constructs, have so far never been reported in literature for the nuclear<br />

expression <strong>of</strong> a transgene in leaves <strong>of</strong> transgenic plants.<br />

In another experiment, enhanced expression <strong>of</strong> the gene coding for the<br />

antimicrobial peptide sarcotoxin IA was studied by fusing translationally the coding<br />

sequence <strong>of</strong> this gene to that <strong>of</strong> E. coli b-glucuronidase (GUS) (Okamoto et al., 1998).<br />

Western blot analysis <strong>of</strong> transgenic tobacco plants demonstrated that the amounts<br />

<strong>of</strong> sarcotoxin IA present in the form <strong>of</strong> sarcotoxin IA-GUS fusion proteins were<br />

considerably higher than in tobacco plants transformed with the single sarcotoxin<br />

IA peptide construct.<br />

It is assumed that a high transcription <strong>of</strong> genes coding for proteins that control<br />

fungi under in vitro conditions will increase resistance in plants. One <strong>of</strong> these<br />

strategies relies on strong promoters. However, much higher transgene expression<br />

levels can be achieved with chloroplast genetic engineering (Daniell et al., 2002)<br />

because chloroplasts are polyploid. Thousands <strong>of</strong> copies <strong>of</strong> foreign genes per plant<br />

cell will generate extraordinarily high levels <strong>of</strong> foreign protein. Consequently,<br />

chloroplast transgenic plants can show a 25-fold increase in the accumulation <strong>of</strong><br />

foreign gene products than nuclear transgenic plants (Lee et al., in press; Daniell<br />

et al. 2002). In tobacco this resulted in the accumulation <strong>of</strong> 45.3% foreign protein<br />

<strong>of</strong> total soluble protein (De Cosa et al., 2001).<br />

In related experiments, 21.5% <strong>of</strong> total soluble protein was demonstrated to be<br />

foreign and resulted in the protection against a fungal pathogen (DeGray et al.,<br />

2001). Sidorov et al. (1999) and Ruf et al. (2001) achieved expression levels up to<br />

5% and 50%, respectively. Moreover, in contrast to nuclear transformation where<br />

the integration <strong>of</strong> a transgene is random and in unpredictable numbers, chloroplast<br />

transformation facilitates the controlled integration in a pre-determined site,<br />

thereby influencing the expression <strong>of</strong> the transgene (Kumar and Fladung, 2001).<br />

226


Session 4<br />

R. Swennen et al.<br />

Field testing<br />

The genetic modification <strong>of</strong> many tropical and subtropical crops <strong>of</strong>fers the prospect<br />

<strong>of</strong> faster plant improvement (Ortiz, 1998; Sharma et al., 2000). Banana may be next.<br />

Since 1994, putative fungal resistant transgenic plantains have been obtained at<br />

KULeuven and analysed in partnership with scientists from Cuba, Ecuador, India and<br />

Uganda. The non-toxicity <strong>of</strong> the expressed proteins in fruits, and in feeding tests<br />

with rats, suggests that these plants should be field-tested in the tropics for<br />

confirmation <strong>of</strong> resistance and biosafety evaluation. Contained fields and nurseries<br />

have been put in place, but the LMOs (living modified organism) have not been<br />

exported due to the absence <strong>of</strong> competent national authorities to approve the request<br />

for import and risk assessment studies (Sági et al., 1998b).<br />

The absence <strong>of</strong> a regulatory framework in most tropical developing countries is<br />

delaying the evaluation <strong>of</strong> LMO <strong>bananas</strong> and plantains and the cultivation <strong>of</strong> resistant<br />

plantains by smallholders. This is occurring despite the ratification <strong>of</strong> the Cartagena<br />

Protocol (Cartagena Protocol, 2000) and article 19(3) <strong>of</strong> the Convention on Biological<br />

Diversity (CBD) (Convention on Biological Diversity, 1994). The objective <strong>of</strong> the<br />

Cartagena Protocol (adopted in January 2000) is to ensure an adequate level <strong>of</strong><br />

protection in the field for the safe transfer, handling and use <strong>of</strong> LMOs resulting from<br />

modern biotechnology that may have adverse effects on the conservation and<br />

sustainable use <strong>of</strong> biological diversity, taking also into account risks to human health,<br />

and specifically focusing on transboundary movements.<br />

Currently, public opinion is influenced by feelings about “Frankenstein food” and<br />

by the “precautionary principle”. The former calls for more scientific data whereas<br />

the latter should allow for approvals for field-testing. Indeed, the “precautionary<br />

principle” should not be used to stop field-testing but to guide scientists under what<br />

circumstances field-testing should be conducted. Besides, scientific data from the<br />

field can be used to further improve current requirements. Edible <strong>bananas</strong> are<br />

particularly suitable since they are both seed and pollen sterile. Thus, the introduced<br />

gene(s) remain confined to the transformed plant. Banana LMO plants should have<br />

been among the first plants to be tested in developing countries.<br />

The development and field release <strong>of</strong> transgenic plants have been much debated.<br />

It is clear that the deployment <strong>of</strong> transgenic plants should be safe (Custers, 2001).<br />

Therefore, ecological risk assessment studies need to be conducted on matters dealing<br />

with the invasiveness <strong>of</strong> the transgenic crop (can it become a weed in the natural<br />

habitat?), on the invasiveness <strong>of</strong> the transgene itself (gene flow into wild relatives)<br />

and the environmental side effects <strong>of</strong> the transgenic products (on non-target<br />

organisms, for example) (Amman, 2001). To avoid any type <strong>of</strong> invasiveness, research<br />

is conducted to introduce “reproductive isolation barriers” into crop plants, the<br />

biosafety <strong>of</strong> transgenic crops being one <strong>of</strong> the driving forces. Examples are male<br />

sterility (there are no viable pollen, hence no outcrossing) and the terminator<br />

technology (seeds cannot germinate without chemical application). Complementary<br />

strategies rely on the cultivation <strong>of</strong> sexually incompatible crops and respecting<br />

isolation barriers, i.e. crops that can intercross are separated by a crop that cannot<br />

intercross (Obrycki et al., 2001). The industry and environmentalists favour<br />

reproductive isolation barriers, but the strategy would seriously handicap future<br />

227


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

breeding as it would require that an interesting gene be inserted in each cultivar<br />

separately, which is very cumbersome and costly. However, a transgenic plant <strong>of</strong><br />

interest should become part <strong>of</strong> a conventional breeding programme and used for<br />

further crossing (Dodds et al., 2001). In the case <strong>of</strong> banana, this means that diploids<br />

should also be transformed for use in the current breeding programmes.<br />

Conclusion<br />

The predicted increase <strong>of</strong> the world’s population in the coming years poses many<br />

challenges to developing countries to feed their population as more than 90% <strong>of</strong><br />

the population increase will occur there. But since agricultural productivity is<br />

currently low, there are many opportunities to improve that situation. Technical<br />

solutions, such as biotechnology, are not the only solution but form part <strong>of</strong> a package<br />

to be used in synergy with an agro-ecological approach. The technology being in<br />

the plant material, biotechnology ensures benefits to smallholders without changing<br />

local cultural practices, as long as the appropriate features are considered. The 2001<br />

United Nations Human Development Report unequivocally states that biotechnology<br />

<strong>of</strong>fers “the hope <strong>of</strong> crops with higher yields, pest- and drought-resistant properties<br />

and superior nutritional characteristics - especially for farmers in ecological zones<br />

left behind by the green revolution” (UNDP, 2001).<br />

Many opponents raise ethical questions but blocking the development and<br />

application <strong>of</strong> biotechnology can also be construed as unethical. Zero risk does not<br />

exist. The important point is that biotechnology poses risks that are equal to the<br />

risks encountered in conventional breeding (NRC, 2000). “A process that is safer<br />

shouldn’t be given up because it cannot be elevated to an impossible standard <strong>of</strong><br />

absolute safety” (Trewavas, 2000). In the end, what counts is zero harm.<br />

Breeding has long been used to suppress plant <strong>diseases</strong> and pathogen-resistant<br />

cultivars quickly became popular and grown as homogeneous crops. Pathogens can<br />

eventually overcome resistance and become epidemic, forcing breeders to introduce<br />

a cultivar with a new resistance trait. The battle never ends as pathogens always try<br />

to circumvent recognition by resistant plants.<br />

To protect crops better, plant cultivars that differ in their resistance mechanisms<br />

should be mixed, as in natural plant populations (Dangl and Jones, 2001). For<br />

example, the deployment <strong>of</strong> different rice cultivars resulted in a 94% reduction in<br />

the occurrence <strong>of</strong> rice blast (Zhu et al., 2000). Smallholders would be best served<br />

by interplanting into their existing banana plot banana plants resistant to <strong>leaf</strong> <strong>spot</strong><br />

<strong>diseases</strong> consisting <strong>of</strong> cultivar(s) in which a single or a combination <strong>of</strong> foreign genes<br />

have been integrated. Given that it should be possible target proteins both intraand<br />

intercellularly, a broad-spectrum resistance to banana <strong>leaf</strong> <strong>spot</strong> should be<br />

achievable.<br />

References<br />

Antoniw J.F., C.E. Ritter, W.S. Piepont and L.C. Van Loon. 1980. Comparison <strong>of</strong> three<br />

pathogenesis-related proteins from plants <strong>of</strong> two cultivars <strong>of</strong> tobacco infected with TMV.<br />

J. Gen. Virol. 47:79-87.<br />

228


Session 4<br />

R. Swennen et al.<br />

Amman K. 2001. Safety <strong>of</strong> genetically engineered plants: an ecological risk assessment <strong>of</strong><br />

vertical gene flow. Custers R. (ed.). Safety <strong>of</strong> Genetically Engineered Crops. VIB, Belgium<br />

pp. 61-87. http://www.vib.be/frame.cfm<br />

Arinaitwe G., S. Remy, H. Strosse, R. Swennen and L. Sági. in press. Agrobacterium-and particle<br />

bombardment-mediated transformation <strong>of</strong> a wide range <strong>of</strong> banana cultivars (S.M. Jain<br />

and R. Swennen, eds) Banana Improvement: Cellular, Molecular Biology and Induced<br />

Mutations, Kluwer.<br />

Bartels N., F.M. van der Lee, J. Klap, O.J. Goddijn, M. Karimi, P. Puzio, F.M. Grindler,<br />

S.A. Ohl, K. Lindsey, L. Robertson, W.M. Robertson, M. Van Montagu, G. Gheysen and<br />

P.C. Sijmons. 1997. Regulatory sequences <strong>of</strong> Arabidopsis drive reporter gene expression<br />

in nematode feeding structures. Plant Cell 9:2119-2134.<br />

Becker D.K., B. Dugdale, M.K. Smith, R.M. Harding and J.L. Dale. 2000. Genetic transformation<br />

<strong>of</strong> Cavendish banana (Musa spp. AAA group) cv ‘Grand Nain’ via microprojectile<br />

bombardment. Plant Cell Rep. 19:229-234.<br />

Bent A.F. 2000. Arabidopsis in planta transformation. Uses, mechanisms, and prospects for<br />

transformation <strong>of</strong> other species. Plant Physiology 124:1540-1547.<br />

Broekaert W.F., X. Mariën, F.R.G. Terras, M.F.C. De Bolle, P. Proost, J. Van Damme, L. Dillen,<br />

M. Claeys, S.B. Rees, J. Vanderleyden and B.P.A. Cammue. 1992. Antimicrobial peptides<br />

from Amaranthus caudatus seeds with sequence homology to the cysteine/glycine-rich<br />

domain <strong>of</strong> chitin-binding proteins. Biochemistry 31:4308-4314.<br />

Broekaert W.F., B.P.A. Cammue, M.F.C. De Bolle, K. Thevissen, G.W. De Samblaux and<br />

R.W. Osborn. 1997. Antimicrobial peptides <strong>of</strong> plants. Crit. Rev. Plant Sci. 16:297-323.<br />

Cammue B.P.A., M.F.C. De Bolle, F.R.G. Terras, P. Proost, J. Van Damme, S.B. Rees,<br />

J. Vanderleyden and W.F. Broekaert. 1992. Isolation and characterisation <strong>of</strong> a novel class<br />

<strong>of</strong> plant antimicrobial peptides from Mirabilis jalapa L. seeds. J. Biol. Chem. 267:2228-<br />

2233.<br />

Cammue B.P.A., M.F.C. De Bolle, F.R.G. Terras and W.F. Broekaert. 1993. Fungal disease control<br />

in Musa: application <strong>of</strong> new antifungal proteins. Pp. 221-225 in Breeding Banana and<br />

Plantain for Resistance to Diseases and Pests. (Ganry J., ed.) Montpellier, 7-9 September<br />

1992, CIRAD, Montpellier, France.<br />

Cammue B.P., K. Thevissen, M. Hendrikx, K. Eggermont, I.J. Goderis, P. Proost, J. Van Damme,<br />

R.W. Osborn, F. Guerbette, J.C. Kader and W.F. Broekaert. 1995. A potent antimicrobial<br />

protein from onion (Allium cepa L.) seeds showing strong sequence homology to plant<br />

lipid transfer proteins. Plant Physiol. 109:445-455.<br />

Cartagena Protocol on Biosafety to the Convention on Biological Diversity. 2000. Secretariat<br />

<strong>of</strong> the Convention on Biological Diversity, Text and Annexes, Montreal.<br />

http://www.biodiv.org/biosafety/ and http://www.biodiv.org/doc/legal/cartagena-protocolen.pdf<br />

Chia T.-F., Y.S. Chan and N.H. Chua. 1994. The firefly luciferase gene as a non-invasive reporter<br />

for Dendrobium transformation. Plant J. 6:441-449.<br />

CIRAD. 2001. Banana genetic transformation. Plant genomics at CIRAD. At http://www.cirad.fr/<br />

presentation/programmes/biotrop/resultats/biositecirad/transfo/bananatg.htm<br />

Clough S.J. and A.F. Bent. 1998. Floral dip: a simplified method for Agrobacterium-mediated<br />

transformation <strong>of</strong> Arabidopsis thaliana. The Plant Journal 16(6):735-743.<br />

Colligne D.B. and A.J. Slusarenko. 1987. Plant gene expression in response to pathogens.<br />

Plant Mol. Biol 9:389-410.<br />

Côte F.X., R. Domergue, S. Mommarson, J. Schwendiman, C. Teisson and J.V. Escalant. 1996.<br />

Embryogenic cell suspension from the male flower <strong>of</strong> Musa AAA cv. Grand nain. Physiol.<br />

Plant. 97:285-290.<br />

229


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Côte F.X., T. Legavre, A. Grapin, B. Valentin, O. Frigout, J. Babeau, D. Meynard, F. Bakry and<br />

C. Teisson. 1997. Genetic transformation <strong>of</strong> embryogenic cell suspension in plantain (Musa<br />

AAB) using particle bombardmen. Proceedings <strong>of</strong> the International symposium on<br />

biotechnology <strong>of</strong> tropical and subtropical species: Part 1. (Drew R.A., A. Sasson, eds). Acta<br />

Hort. 460.<br />

Côte F.X., O. Goue, R. Domergue, B. Panis and C. Jenny. 2000. In-field behaviour <strong>of</strong> banana<br />

plants (Musa AA spp.) obtained after regeneration <strong>of</strong> cryopreserved embryogenic cell<br />

suspensions. Cryoletters 21:19-24.<br />

Convention on Biological Diversity. 1994. Convention on Biological Diversity Text and<br />

Annexes. Geneva: Interim Secreteriat for the Convention on Biological Diversity.<br />

http://www.biodiv.org/convention/articles.asp?lg=0<br />

Cronauer S.S. and A.D. Krikorian. 1988. Plant regeneration via somatic embryogenesis in the<br />

seeded diploid banana Musa ornata Roxb. Plant Cell Reports 7:23-25.<br />

Custers R. 2001. Safety <strong>of</strong> Genetically Engineered Crops. VIB publication: 160pp.<br />

http://www.vib.be/frame.cfm<br />

De Bondt A. 1995. Molecular Breeding <strong>of</strong> Apple (Malus domestica Borkh.) via Agrobacterium<br />

tumefaciens for increased fungal resistance. PhD thesis No 297, Katholieke Universiteit<br />

Leuven, Belgium. 103pp.<br />

De Langhe E., R. Swennen and D. Vuylsteke. 1995. Plantain in the Early Bantu World. Sutton<br />

J. (ed.). Azania XXIX-XXX 1994-1995. Special volume on ‘The Growth <strong>of</strong> Farming<br />

Communities in Africa from the Equator Southwards. Proceedings <strong>of</strong> a conference <strong>of</strong> the<br />

British Institute in Eastern Africa. Cambridge, 4-8 July 1994:147-160.<br />

De Langhe E. and P. De Maret. 1999. Tracking the banana: its significance in early agriculture.<br />

Pp. 377-396 in The prehistory <strong>of</strong> food. Appetites for change (C. Gosden and J. Hather,<br />

eds). Routledge, London and New York.<br />

Dangl J.L. and J.D.G. Jones. 2001. Plant pathogens and integrated defence responses to<br />

infection. Nature 411(6939):826-833.<br />

Daniell H., M.S. Khan and L. Allison. 2002. Milestones in chloroplast genetic engineering:<br />

an environmentally friendly era in biotechnology. Trends in Plant Science 7:84-91.<br />

De Cosa B., W. Moar, Seung-Bum Lee, M. Miller and H. Daniell. 2001. Overexpression <strong>of</strong><br />

the Bt Cry2Aa2 operon in chloroplasts leads to formation <strong>of</strong> insecticidal crystals. Nat.<br />

Biotechnol. 19: 71-74.<br />

DeGray G., K. Rajasekaran, F. Smith, J. Sanford and H. Daniell. 2001. Expression <strong>of</strong> an<br />

antimicrobial peptide via the chloroplast genome to control phytopathogenic bacteria and<br />

fungi. Plant Physiol. 127:852-862.<br />

De Vries S.C., H. Booij, P. Meyerink, G. Huisman, H.D. Wilde, T.L. Thomas and A. van Kammen.<br />

1988. Acquisition <strong>of</strong> embryogenic potential in carrot cell-suspension cultures. Planta<br />

176:196-204.<br />

De Wet J.R, K.V. Wood, D.R. Helinski and M. DeLuca. 1985. Cloning <strong>of</strong> firefly luciferase cDNA<br />

and the expression <strong>of</strong> active luciferase in Escherichia coli. Proc Natl Acad Sci USA 82:7870-<br />

7873.<br />

Dhed’a D., F. Dumortier, B. Panis, D. Vuylsteke and E. De Langhe. 1991. Plant regeneration<br />

in cell suspension cultures <strong>of</strong> cooking banana cv. Bluggoe (Musa spp. ABB group). Fruits<br />

46:125-135.<br />

Dodds J.H., R. Ortiz, J.H. Crouch, V. Mahalasksmi and K.K. Sharma. 2001. Biotechnology,<br />

the Gene Revolution, and Proprietary Technology in Agriculture: A Strategic Note for the<br />

World Bank. Strategy Today 2:29pp. http://www.biodevelopments.org/ip/ipst2.pdf or<br />

http://www.swiftt.cornell.edu<br />

230


Session 4<br />

R. Swennen et al.<br />

Dolezel J., M.A. Lysak, M. Dolezelova, M. Valarik, H. Simkova and J. Vrana. 1999. Analysis<br />

<strong>of</strong> Musa genome using flow cytometry and molecular cytogenetics. Pp.85-93 in Report<br />

<strong>of</strong> the 3 rd research co-ordination meeting <strong>of</strong> FAO/IAEA/BADC co-ordinated research project<br />

‘Cellular biology and biotechnology including mutation techniques for creation <strong>of</strong> new<br />

useful banana genotypes, Colombo, Sri Lanka, 4-8 October 1999.<br />

Dugdale B., P.R. Beetham, D.K. Becker, R.M. Harding and J.L. Dale. 1998. Promoter activity<br />

associated wit intergenic regions <strong>of</strong> banana bunchy top virus DNA-1 to –6 in transgenic<br />

tobacco and banana cells. Journal <strong>of</strong> General Virology 79:2301-2311.<br />

Dugdale B., D.K. Becker, P.R. Beetham, R.M. Harding and J.L. Dale. 2000. Promoters derived<br />

from banana bunchy top virus DNA-1 to DNA-5 direct vascular-associated expression in<br />

transgenic banana (Musa spp.) Plant Cell Rep. 19:810-814.<br />

Dugdale B., D.K. Becker, R.M. Harding and J.L. Dale. 2001. Intron-mediated enhancement <strong>of</strong><br />

the banana bunchy top virus promoter in banana (Musa spp.) embryogenic cells and plants.<br />

Plant Cell Rep. 20:220-226.<br />

Escalant J.V. C. Teisson and F. Côte. 1994. Amplified somatic embryogenesis from male flowers<br />

<strong>of</strong> triploid banana and plantain culivars (Musa spp.), In vitro Plant Cellular and<br />

Developmental Biology 30:181-186.<br />

Escalant J.V. and C. Teisson. 1989. Somatic embryogenesis from immature zygotic embryos<br />

<strong>of</strong> the species Musa acuminata and Musa balbisiana. Plant Cell Rep. 7:665-668.<br />

Fraley R.T., S.G. Rogers, R.B. Horsch, P.R. Sanders, J.S. Fick, S.P. Adams, M.L. Bittners,<br />

L.A. Brand, C.L. Fink, Y.S. Fry, G.R. Galluppi, S.B. Goldberg, N.L. H<strong>of</strong>fmann and S.C. Woo.<br />

1983. Expression <strong>of</strong> bacterial genes in plant cells. Proc. Natl. Acad. Sci USA 80:4803-4807.<br />

François I.E.J.A., M.F.C. De Bolle, I.J.W.M. Goderis, P. Verhaert, P. Proost, B.P.A. Cammue and<br />

W.F. Broekaert. 2002a. Transgenic expression in Arabidopsis thaliana <strong>of</strong> a polyprotein<br />

construct leading to production <strong>of</strong> two different antimicrobial proteins. Plant Physiol.<br />

128:1346-1358.<br />

François I.E.J.A., G.I. Dwyer, M.F.C. De Bolle, I.J.W.M. Goderis, P. Proost, P. Wouters, W.F.<br />

Broekaert and B.P.A. Cammue. 2002b. Processing in transgenic Arabidopsis thaliana plants<br />

<strong>of</strong> polyproteins with linker peptide variants derived from the Impatiens balsamina<br />

antimicrobial polyprotein precursor. Plant Physiol. Biochem. 40(10):871-879.<br />

Fuchs R.L., J.E. Ream, B.G. Gammond, M.W. Naylor, R.M. Leimgruber and S.A. Berberich.<br />

1992. Safety assessment <strong>of</strong> the neomycin phosphotransferase II (NPTII) protein.<br />

Bio/Technology 11:1543-1547.<br />

Ganapathi T.R., N.S. Higgs, P.J. Balint-Kurti and J. Van Eck. 2001. Agrobacterium-mediated<br />

transformation <strong>of</strong> embryogenic cell suspensions <strong>of</strong> the banana cultivar Rasthali (AAB).<br />

Plant Cell Rep. 20:157-162.<br />

Garcia-Olmedo F., A. Molina, J.M. Alamillo and P. Rodriguez-Palenzuela. 1998. Plant defense<br />

peptides. Biopolymers 47:479-491.<br />

Georget F., R. Domergue, N. Ferriere and F.X. Côte. 2000. Morphohistological study <strong>of</strong> the<br />

different constituents <strong>of</strong> a banana (Musa AAA, cv. Grande naine) embryogenic cell<br />

suspension. Plant Cell Reports 19:748-754.<br />

Grapin A. 1995. Régénération par embryogénèse somatique en milieu liquide et transformation<br />

génétique par biolistique de bananiers di- et triploïdes. (Regeneration by somatic<br />

embryogenesis in a liquid medium and genetic transformation <strong>of</strong> diploid and triploid<br />

<strong>bananas</strong> by biolistics.) ENSAM, Montpellier (FRA) 93pp.<br />

Grapin A., J. Schwendiman and C. Teisson. 1996. Somatic embryogenesis in plantain banana.<br />

In vitro Plant Cellular and Developmental Biology 32:66-71.<br />

Grapin A., O. Frigout, S. Monmarson, T. Legavre and F. Côte. 1996. GUS reporter gene<br />

expression in transformed embryogenic cell suspensions <strong>of</strong> banana (Musa sp.). Meeting<br />

231


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

on tropical plants. Communications and posters. - EUCARPIA, European Association for<br />

Research on Plant Breeding, Wageningen (NLD) 294pp.<br />

Grapin A., J.L. Ortiz, R. Domergue, J. Babeau, S. Monmarson, J.V. Escalant, C. Teisson and<br />

F. Côte. 1998. Establishment <strong>of</strong> embryogenic callus initiation and regeneration <strong>of</strong><br />

embryogenic cell suspensions from female and male immature flowers <strong>of</strong> Musa.<br />

INFOMUSA 7(1):13-15.<br />

Hadi M.Z., M.D. McMullen and J.J. Finer, 1996. Transformation <strong>of</strong> 12 different plasmids into<br />

soybean via particle bombardment. Plant Cell Rep. 15:500-505.<br />

Hahlbrock K. and D. Scheel. 1989. Physiology and molecular biology <strong>of</strong> phenylpropanoid<br />

metabolism. Annu. Rev. Plant Physiol. Plant Mol. Biol. 40:347-369.<br />

Haldrup A., S.G. Petersen and F.T. Okkels. 1998a. Positive selection: a plant selection principle<br />

based on xylose isomerase, an enzyme used in food industry. Plant Cell Rep. 18:76-81.<br />

Haldrup A., S.G. Petersen and F.T. Okkels. 1998b. The xylose isomerase gene from<br />

Thermoanaerobacterium thermosulfurogenes allows effective selection <strong>of</strong> transgenic plant<br />

cells using D-xylose as the selection agent. Plant Mol. Biol. 37:287-296.<br />

Harris J.M. 1996. World agricultural futures: regional sustainability and ecological limits.<br />

Ecological Economics 17:95-115.<br />

Heil M. and I.T. Baldwin. 2002. Fitness costs <strong>of</strong> induced resistance: emerging experimental<br />

support for a slippery concept. Trends in Plant Science 7:61-67.<br />

Hermann S.R., D.K. Becker, R.M. Harding and J.L. Dale. 2001a. Promoters derived from banana<br />

bunchy top virus-associated components S1 and S2. Plant Cell Rep. 20:642-646.<br />

Hermann S.R., R.M. Harding and J.L. Dale. 2001b. The banana actin 1 promoter drives nearconstitutive<br />

transgene expression in vegetative tissues <strong>of</strong> banana (Musa spp.). Plant Cell<br />

Rep. 20:525-530.<br />

Jach G., B. Gornhardt, J. Moundy, J. Logemann, E. Pinsdorf, R. Leah, J. Schell and C. Maas.<br />

1995. Enhanced quantitative resistance against fungal disease by combinatorial expression<br />

<strong>of</strong> different barley antifungal proteins in transgenic tobacco. Plant J. 8:97-109.<br />

Kahl G. in press. The banana genome in focus: a technical perspective. In Banana<br />

Improvement:. Cellular, Molecular Biology and Induced Mutations (Jain S.M. and<br />

Swennen R., eds). Kluwer.<br />

Kumar S. and M. Fladung. 2001. Controlling transgene integration in plants. Trends in Plant<br />

Science 6:155-159.<br />

Lee C., M. Bagdasarian, M. Meng and J.G. Zeikus. 1990. Catalytic mechanism <strong>of</strong> xylose<br />

(glucose) isomerase from Clostridium thermosulfurogenes: Characterization <strong>of</strong> the structural<br />

gene and function <strong>of</strong> active site histidine. J. Biol. Chem. 265:19082-19090.<br />

Lee S.B., H.B. Kwon, S.J. Kwon, S.C. Park, M.J. Jeong, S.E. Han, H. Daniell and M.O. Byun.<br />

In press. Drought tolerance conferred by the yeast trehalose-6 phosphate synthase gene<br />

engineered via the chloroplast genome. Transgenic Research.<br />

Lei H., S.P. Oh, M. Okano, R. Jüttermann, K.A. Goss and R. Jaenisch. 1996. De novo DNA<br />

cytosine methyltransferase activities in mouse embryonic stem cells. Development<br />

122:3195-3205.<br />

Linthorst H.J.M. 1991. Pathogenesis-related proteins <strong>of</strong> plants. Crit. Rev. Plant Sci. 10:123-<br />

150.<br />

Lysak M.A., M. Dolezelova, J.P. Horry, R. Swennen and J. Dolozel. 1999. Flow cytometric<br />

analysis <strong>of</strong> nuclear DNA content in Musa. Theoretical and Applied Genetics 98:1344-1350.<br />

Matsumara H., S. Nirasawa and R. Terauchi. 1999. Technical advance: Transcript pr<strong>of</strong>iling<br />

in rice (Oryza sativa L.) seedlings using serial analysis <strong>of</strong> gene expression. The Plant J.<br />

20:719–726.<br />

232


Session 4<br />

R. Swennen et al.<br />

Malca I., R.M. Endo and M.R. Long. 1967. Mechanism <strong>of</strong> glucose counteraction <strong>of</strong> inhibition<br />

<strong>of</strong> root elongation by galactose, mannose and glucosamine. Phytopathology 57:272-278.<br />

Más L., S. Martinez, G. Aguero, M. Reyes, R. Gomez, G. De la Riva, A. Coego, R. Vasquez,<br />

B. Ocaña and V. Gill. 1999. Expression <strong>of</strong> foreign genes in somatic embryos <strong>of</strong> banana<br />

cv ‘Gran Enano’, via gene gun. P. 51 in The International Symposium on the Molecular<br />

and Cellular Biology <strong>of</strong> Banana.<br />

Más L., G. Agüero, V. Gil, M. Reyes, R. Gómez, B. Ocaña and S. Martínez. 2000. Optimización<br />

de parámetros en la transformación de embriones somáticos de banano utilizando pistola<br />

de genes. Biotecnología Vegetal 1:51-54.<br />

May G., R. Afza, H. Mason, A. Wiecko, F. Novak and C. Arntzen. 1995. Generation <strong>of</strong> transgenic<br />

banana (Musa acuminata) plants via Agrobacterium-mediated transformation. Bio/<br />

Technology 13:486-492.<br />

Mbida M. C., H. Doutrelepont, L. Vrydaghs, R. Swennen, R.J. Swennen, H. Beeckman, E. De<br />

Langhe and P. De Maret. 2001. First archaeological evidence <strong>of</strong> banana cultivation in<br />

central Africa during the third millenium before present. Vegetation History and<br />

Archaeobotany 10:1-6.<br />

Mehdy M.C. 1994. Active oxygen species in plant defense against pathogens. Plant Physiol.<br />

105:467-472.<br />

Meijer E.A., S.C. de Vries and A.P. Mordhorst. 1999. Co-culture with Daucus carota somatic<br />

embryos reveals high 2,4-D uptake and release rates <strong>of</strong> Arabidopsis thaliana cultured cells.<br />

Plant Cell Reports 18(7-8):656-663.<br />

Michelmore R.W. and B.C. Meyers. 1998. Clusters <strong>of</strong> resistance genes in plants evolve by<br />

divergent selection and a birth-and-death process. Genome Res. 8:1113-1130.<br />

Miles J S. and J.R. Guest. 1984. Nucleotide sequence and transcriptional start point <strong>of</strong> the<br />

phosphomannose isomerase gene (manA) <strong>of</strong> Escherichia coli. Gene 32:41-48.<br />

Millar A.J., S.R. Short, K. Hiratsuka, N.-H. Chua and S.A. Kay. 1992. Firefly luciferase as a<br />

reporter <strong>of</strong> regulated gene expression in plants. Plant Mol Biol Rep 10:324-337.<br />

Moy Y., B. Hanson, N. Gutterson, and D. Engler. 1998. Development <strong>of</strong> an efficient banana<br />

transformation method and analysis <strong>of</strong> the expression <strong>of</strong> some constitutive promoters in<br />

young transgenic plants. Plant biotechnology and in vitro biology in the 21 st century:<br />

abstracts. IAPTC, International Association for Plant Tissue Culture, Jérusalem, Israel:140.<br />

Moy Y., B. Hanson, D. Engler and N. Gutterson. 1999. Analysis <strong>of</strong> uidA gene expression from<br />

transgenes in field-grown ‘Grand Nain’ (AAA) banana plants. The 1 st International<br />

symposium on the molecular and cellular biology <strong>of</strong> banana.<br />

Newell C.A. 2000. Plant transformation technology: Developments and applications. Molecular<br />

Biotechnology 16:53-65.<br />

Novak F.J., R. Afza, M. Van Duren, M. Perea-Dallos, B.V. Conger and T. Xiaolang. 1989. Somatic<br />

embryogenesis and plant regeneration in suspension cultures <strong>of</strong> dessert (AA and AAA)<br />

and cooking (ABB) <strong>bananas</strong> (Musa spp.). Bio/Technology 46:154-159.<br />

NRC (National Research Council). 2000. Genetically modified pest-protected plants: science<br />

and regulation. Washington, DC: National Academy Press. HTTP://WWW.NAP.EDU/BOOKS/<br />

0309069300.HTML<br />

Obrycki J.J., J.E. Losey, O.R. Taylor and L.C.H. Jesse. 2001. Analysis <strong>of</strong> transgenic insecticidal<br />

corn developed for lepidopteran pests reveals that the potential benefits <strong>of</strong> crop genetic<br />

engineering for insect pest management may not outweigh the potential ecological and<br />

economic risks. BioScience 51(5):353.<br />

233


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Okkels F. T., J.L. Ward and M. Joersbo. 1997. Synthesis <strong>of</strong> cytokinin glucuronides for the<br />

selection <strong>of</strong> transgenic plant cells. Phytochemistry 46:801-804.<br />

Okamoto M., I. Mitsuhara, M. Ohshima, S. Natori and Y. Ohashi. 1998. Enhanced expression<br />

<strong>of</strong> an antimicrobial peptide sarcotoxin IA by GUS fusion in transgenic tobacco plants.<br />

Plant Cell Physiol. 39:57-63.<br />

Ortiz R. 1998. Critical role <strong>of</strong> plant biotechnology for the genetic improvement <strong>of</strong> food crops:<br />

perspectives for the next millennium. Electronic Journal <strong>of</strong> Biotechnology 1(3).<br />

http://www.ejb.org/content/vol1/issue3/full/7<br />

Ortiz R. and D. Vuylsteke. 1995. Factors influencing seed set in triploid Musa spp. L. and<br />

production <strong>of</strong> euploid hybrids. Annals <strong>of</strong> Botany 75:151-155.<br />

Osborn R.W., G.W. De Samblanx, K. Thevissen, I. Goderis, S. Torrekens, F. Van Leuven, S.<br />

Attenborough, S.B. Rees and W.F. Broekaert. 1995. Isolation and characterisation <strong>of</strong> plant<br />

defensins from seeds <strong>of</strong> Asteraceae, Fabaceae, Hippocastanaceae and Saxifragaceae. FEBS<br />

Lett. 368:257-262.<br />

Ow D.W., K.V. Wood, M. DeLucia, J.R. De Wet, D.R. Helinski and S.H. Howell. 1986. Transient<br />

and stable expression <strong>of</strong> the firefly luciferase gene in plant cells and transgenic plants.<br />

Science 234:856-859.<br />

Panis B., L.A. Withers and E. De Langhe. 1990. Cryopreservation <strong>of</strong> Musa suspension cultures<br />

and subsequent regeneration <strong>of</strong> plants. Cryo-Letters 11:337-350.<br />

Panis B., H. Strosse, S. Remy, L. Sági and R. Swennen. in press. Cryopreservation <strong>of</strong> banana<br />

tissues: support for germplasm conservation and banana improvement. In Banana<br />

Improvement: Cellular, Molecular Biology and Induced Mutations (S.M. Jain and<br />

R. Swennen, eds). Kluwer.<br />

Pérez Hernández J.B. 2000. Development and application <strong>of</strong> Agrobacterium-mediated genetic<br />

transformation to increase fungus-resistance in banana (Musa spp). PhD thesis No. 442,<br />

Katholieke Universiteit Leuven, Belgium. 187pp.<br />

Pérez Hernández J.B., S. Remy, V. Galán Saúco, R. Swennen, and L. Sági. 1998. Chemotaxis,<br />

attachment and transgene expression in the Agrobacterium-mediated banana<br />

transformation system. Med. Fac. Landbouww. Univ. Gent 63 (4b):1603-1606.<br />

Pérez Hernández J.B., S. Remy, V. Galan Sauco, R. Swennen and L. Sági. 1999. Chemotactic<br />

movement and attachment <strong>of</strong> Agrobacterium tumefaciens to single cells and tissues <strong>of</strong><br />

banana. Journal <strong>of</strong> Plant Physiology 155:245-250.<br />

Puchta H. 2000. Removing selectable marker genes: taking the short cut. Trends Plant Sci.<br />

5:273-274.<br />

Puzio P.S., J. Lausen, J. Almeida-Engler, D.G. Cai, G. Gheysen and F.M.W. Grundler. 1999.<br />

Isolation <strong>of</strong> a gene from Arabidopsis thaliana related to nematode feeding structures. Gene<br />

239:163-172.<br />

Reed J., L. Privalle, L. Powell, M. Meghji, J. Dawson, E. Dunder, J. Suttie, A. Wenck, K. Launis,<br />

C. Kramer, Y.-F. Chang, G. Hansen and M. Wright. 2001. Phosphomannose isomerase: an<br />

efficient selectable marker for plant transformation. In Vitro Cell. Dev. Biol. 37:127-132.<br />

Remans T., L. Sági, A.R. Elliotts, R.G. Dietzgen, R. Swennen, P. Ebert, C.P.L. Gr<strong>of</strong>, J.M. Manners<br />

and P.M. Schenk. 2000. Banana streak virus promoters are highly active in transgenic<br />

banana and other monocot and dicot plants. 2 nd International Symposium on the Molecular<br />

and Cellular Biology <strong>of</strong> Banana. Byron Bay, Australia, October-November 2000.<br />

Remy S. 2000. Genetic transformation <strong>of</strong> banana (Musa spp.) for disease resistance by<br />

expression <strong>of</strong> antimicrobial proteins. PhD thesis No. 420, Katholieke Universiteit Leuven,<br />

Belgium. 341pp.<br />

234


Session 4<br />

R. Swennen et al.<br />

Remy S., I. François, B.P.A. Cammue, R. Swennen and L. Sági. 1998a. Co-transformation as<br />

a potential tool to create multiple and durable resistance in banana (Musa spp.). Acta<br />

Hort. 461:361-365.<br />

Remy S., A. Buyens, B.P.A Cammue, R. Swennen and L. Sági. 1998b. Production <strong>of</strong><br />

transgenic banana plants expressing antifungal proteins. Acta Hort. 490:433-436.<br />

Remy S., G. De Weerdt, I. Deconinck, R. Swennen and L. Sági. in press. An ultrasensitive<br />

luminescent detection system in banana biotechnology: from promotor tagging to<br />

southern hybridization. Banana Improvement: Cellular, Molecular Biology and Induced<br />

Mutations (S.M. Jain and R. Swennen, eds) Kluwer<br />

Roux N., H. Strosse, A. Toloza, B. Panis and J. Dolozel. In press a. Detecting ploidy level<br />

instability <strong>of</strong> banana embryogenic cell suspension cultures by flow cytometry in<br />

Banana Improvement: Cellular, Molecular Biology and Induced Mutations (S.M. Jain and<br />

R. Swennen, eds). Kluwer.<br />

Roux N., A. Toloza, J. Dolozel and B. Panis. In press b. Usefulness <strong>of</strong> embryogenic cell<br />

suspension cultures for the induction and selection <strong>of</strong> mutants in Musa spp. in Banana<br />

Improvement: Cellular, Molecular Biology and Induced Mutations (S.M. Jain and<br />

R. Swennen, eds). Kluwer.<br />

Roux N., Dolezel J., Swennen R. and F.J. Zapata-Arias. 2001. Effectiveness <strong>of</strong> three<br />

micropropagation techniques to dissociate cytochimeras in Musa spp. Plant Cell, Tissue<br />

and Organ Culture 66:189-197.<br />

Rowe P. 1984. Breeding <strong>bananas</strong> and plantains. Plant Breeding Reviews 2:135-155.<br />

Rowe P. and F. Rosales. 1990. Breeding <strong>bananas</strong> and plantains with resistance to black sigatoka.<br />

Pp. 243-251 in Sigatoka Leaf Spot Disease <strong>of</strong> Bananas. (R. Fullerton and R. Stover, eds).<br />

INIBAP, Montpellier, France.<br />

Ruf S., M. Hermann, I.J. Berger, H. Carrer and R. Bockl. 2001. Stable genetic transformation<br />

<strong>of</strong> tomato plastids and expression <strong>of</strong> a foreign protein in fruit. Nat. Biotechnol. 19:870-<br />

875.<br />

Sági L. 1998. Control <strong>of</strong> flowering time in banana (Musa spp.), a tropical monocot, by genetic<br />

transformation. Acta Agronomica Hungarica 46:439-448.<br />

Sági L., S. Remy, B. Panis, G. Volckaert and R. Swennen. 1992. Transient gene expression in<br />

banana protoplasts. Banana Newsletter 15:42.<br />

Sági L., S. Remy, B. Panis, R. Swennen and G. Volckaert. 1994. Transient gene expression in<br />

electroporated banana (Musa spp., cv. ‘Bluggoe’, ABB group) protoplasts isolated from<br />

regenerable embryogenic cell suspensions. Plant Cell Rep. 13:262-266.<br />

Sági L., B. Panis, S. Remy, H. Scho<strong>of</strong>s, K. De Smet, R. Swennen and B.P.A. Cammue. 1995a.<br />

Genetic transformation <strong>of</strong> banana and plantain (Musa spp.) via particle bombardment.<br />

Bio/Technology 13:481-485.<br />

Sági L., S. Remy, B. Verelst, B. Panis, B.P.A. Cammue, G. Volckaert and R. Swennen. 1995b.<br />

Transient gene expression in transformed banana (Musa cv. ‘Bluggoe’) protoplasts and<br />

embryogenic cell suspensions. Euphytica 85:89-95.<br />

Sági L., B. Panis, S. Remy, B.P.A. Cammue and R. Swennen. 1995c. Stable and transient genetic<br />

transformation <strong>of</strong> banana (Musa spp. cv. ‘Bluggoe’, ABB-group) using cells and protoplasts.<br />

Pp. 85-102.in Proceedings <strong>of</strong> the 11th ACORBAT Meeting (Morales Soto V. ed.). San José,<br />

Costa Rica, 13-18 February 1994<br />

Sági L., G.D. May, S. Remy and R. Swennen. 1998a. Recent developments in biotechnological<br />

research <strong>of</strong> banana (Musa spp.). Pp. 313-327 in Biotechnology and Genetic Engineering<br />

Reviews (M.P. Tombs, ed.). Intercept Ltd, Andover, England.<br />

235


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Sági L., S. Remy and R. Swennen. 1998b. Genetic transformation for the improvement <strong>of</strong><br />

<strong>bananas</strong> - A critical assessment. Focus Paper II. Pp. 33-35 in Networking Banana and<br />

Plantain: INIBAP Annual Report 1997. INIBAP, Montpellier, France.<br />

Sági L., S. Remy, J.B. Pérez Hernández, B.P.A. Cammue and R. Swennen. 2000. Transgenic<br />

banana (Musa Species). pp. 255-268 in Biotechnology in Agriculture and Forestry,<br />

Transgenic Crops II (Y. P. S. Bajaj, ed.). Vol. 47. Springer, Berlin, Heidelberg, New York.<br />

Schenk P.M., L. Sági, T. Remans, R.G. Dietzgen, M.J. Bernard, M.W. Graham and J.M. Manners.<br />

1999. A promoter from sugarcane bacilliform badnavirus drives transgene expression in<br />

banana and other monocot and dicot plants. Plant Molecular Biology 39:1221-1230.<br />

Schenk P.M., T. Remans, L. Sági, A.R. Elliott, R.G. Dietzgen, R. Swennen, P. Ebert, C.P.L. Gr<strong>of</strong><br />

and J.M. Manners. 2001. Promoters for pregenomic RNA <strong>of</strong> banana streak badnavirus are<br />

active for transgene expression in monocot and dicot plants. Plant Molecular Biology<br />

47:399-412.<br />

Scho<strong>of</strong>s H. 1997. The origin <strong>of</strong> embryogenic cells in Musa. PhD thesis No. 330, Katholieke<br />

Universiteit Leuven, Belgium. 257p.<br />

Scho<strong>of</strong>s H., B. Panis., H. Strosse, A. Mayo Mosqueda, J. Lopez Torres, N. Roux, J. Dolozel<br />

and R. Swennen. 1999. Bottlenecks in the generation and maintenance <strong>of</strong> morphogenic<br />

banana cell suspensions and plant regeneration via somatic embryogenesis therefrom.<br />

INFOMUSA 8(2):3-7.<br />

Sharma H.C., K.K. Sharma, N. Seetharama and R. Ortiz. 2000. Prospects for using transgenic<br />

resistance to insects in crop improvement. Electronic Journal <strong>of</strong> Biotechnology 3(2).<br />

http://www.ejb.org/content/vol3/issue2/full/3/<br />

Sherf B.A. and K.V. Wood. 1994. Firefly luciferase engineered for improved genetic reporting.<br />

Promega Notes 49:14-21.<br />

Shirasu K., H. Nakajima, K. Rajashekar, R.A. Dixon and C. Lamb. 1997. Salicylic acid potentiates<br />

an agonist-dependent gain control that amplifies pathogen signal in the activation <strong>of</strong><br />

defense mechanisms. Plant Cell 9:261-270.<br />

Sidorov V.A., D. Kasten, S.Z. Pang, P.T.J. Hajdukiewicz, J.M. Staub and N.S. Nehra. 1999.<br />

Technical advance: stable chloroplast transformation in potato: use <strong>of</strong> green fluorescent<br />

protein as a plastid marker. Plant J. 19:209-216.<br />

Strosse H., I. Van den houwe and B. Panis. in press. Banana cell and tissue culture review.<br />

In Banana Improvement: Cellular, Molecular Biology and Induced Mutations (S.M. Jain<br />

and R. Swennen, eds). Kluwer.<br />

Suprasanna P., L. Sági and R. Swennen. 2002. Positive selectable marker genes for routine<br />

plant transformation. In Vitro Cell Dev Biol Plant 38:125-128.<br />

Swennen R. 1990. Limits <strong>of</strong> morphotaxonomy: names and synonyms <strong>of</strong> plantain in Africa<br />

and elsewhere. Pp. 172-210 in The identification <strong>of</strong> genetic diversity in the genus Musa.<br />

Proceedings <strong>of</strong> an International Workshop (R. L. Jarret, ed.). Los Baños, Philippines, 5-<br />

10 September 1988. INIBAP, Montpellier, France.<br />

Swennen R. 1994. De veredeling van de banaan voor resistentie tegen de bladschimmel<br />

<strong>Mycosphaerella</strong> fijiensis. Mededeling der Zittingen Koninklijke Academie voor Overzeese<br />

Wetenschappen 39:567-576.<br />

Swennen R., I. Van den houwe, S. Remy, L. Sági and H. Scho<strong>of</strong>s. 1998. Biotechnological<br />

approaches for the improvement <strong>of</strong> Cavendish <strong>bananas</strong>, Acta Horticultura 490:415-423.<br />

Swennen R. and D. Vuylsteke. 1993. Breeding black Sigatoka resistant plantains with a wild<br />

banana. Tropical Agriculture (Trinidad) 70(1):74-77.<br />

Tailor R.A., D.P. Acland, S. Attenborough, B.P.A. Cammue, I.J. Evans, R.W. Osborn, J. Ray,<br />

S.B. Rees and W.F. Broekaert. 1997. A novel family <strong>of</strong> small cysteine-rich antimicrobial<br />

236


Session 4<br />

R. Swennen et al.<br />

peptides from seeds <strong>of</strong> Impatiens balsamina is derived from a single precursor protein. J.<br />

Biol. Chem. 272:24480-22487.<br />

Terras F.R.G., I.J. Goderis, F. Van Leuven, J. Vanderleyden, B.P.A. Cammue and W.F.<br />

Broekaert. 1992a. In vitro antifungal activity <strong>of</strong> a radish (Raphanus sativus L.) seed protein<br />

homologous to nonspecific lipid transfer proteins. Plant Physiol. 100:1055-1058.<br />

Terras F.R.G., H. Scho<strong>of</strong>s, M.F.C. De Bolle, F. Van Leuven, S.B. Rees, J. Vanderleyden, B.P.A.<br />

Cammue and W.F. Broekaert. 1992b. Analysis <strong>of</strong> two novel classes <strong>of</strong> plant antifungal<br />

proteins from radish (Raphanus sativus L.) seeds. J. Biol. Chem. 267:15301-15309.<br />

Trewavas A. 2000. GM is the best option we have. Agbioview. http://agbiovieuw.listbot.com/<br />

UNDP. 2001. Human Development Report 2001. Making new technologies work for human<br />

development. http://www.undp.org/hdr2001/<br />

Van den houwe I., J. Guns and R. Swennen. 1998. Bacterial contamination in Musa shoot<br />

tip cultures. Acta Horticultura 490:485-492.<br />

Van den houwe I. and R. Swennen. 2000. Characterization and control <strong>of</strong> bacterial<br />

contaminants in in vitro cultures <strong>of</strong> banana (Musa spp.). Acta Horticulturae 530:69-79.<br />

Van den houwe I., B. Panis and R. Swennen. 2000. The in vitro germplasm collection at the<br />

Musa INIBAP Transit Centre and the importance <strong>of</strong> cryopreservation. Pp.255-260 in<br />

Cryopreservation <strong>of</strong> tropical plant germplasm, Current research progress and applications<br />

(F. Engelmann and H. Takagi, eds). IPGRI, Rome, Italy .<br />

Van der Hoorn R.A.L., P.J.G.M De Wit and M.H.A.J Joosten. 2002. Balancing selection favors<br />

guarding resistance proteins. Trends Plant Science 7:67-71.<br />

Van Loon L.C., Y.A.M. Gerritsen and C.E. Ritter. 1987. Identification, purification and<br />

characterization <strong>of</strong> pathogenesis-related proteins from virus-infected Samsun NN tobacco<br />

leaves. Plant Mol. Biol. 9:593-609.<br />

Vance C.P., T.H. Kirk and R.T. Sherwood. 1980. Lignification as a mechanism <strong>of</strong> disease<br />

resistance. Annu. Rev. Phytopathol. 18:259-288.<br />

Vasil I.K. 1987. Developing cell and tissue culture systems for the improvement <strong>of</strong> cereal and<br />

grass crops. Journal <strong>of</strong> Plant Physiology 128:193-218.<br />

Vasil V. and I.K. Vasil. 1986. Plant regeneration from friable embryogenic callus and cell<br />

suspension cultures <strong>of</strong> Zea mays L. Journal <strong>of</strong> Plant Physiology 124:399-408.<br />

Vuylsteke D. and R. Swennen. 1993. Genetic improvement <strong>of</strong> plantains: the potential <strong>of</strong><br />

conventional approaches and the interface with in vitro culture and biotechnology.<br />

Pp. 169-176 in Biotechnology applications for banana and plantain improvement. San<br />

José, Costa Rica, 27-31 January 1992.<br />

Vuylsteke D., R. Swennen and R. Ortiz 1993a. Registration <strong>of</strong> 14 improved tropical Musa<br />

plantain hybrids with black Sigatoka resistance. HortScience 28(9):957-959.<br />

Vuylsteke D., R. Swennen and R. Ortiz. 1993b. Development and performance <strong>of</strong><br />

black Sigatoka-resistant tetraploid hybrids <strong>of</strong> plantain (Musa spp., AAB group). Euphytica<br />

65:33-42.<br />

Vuylsteke D., R. Ortiz and S. Ferris. 1993c. Genetic and agronomic improvement for<br />

sustainable production <strong>of</strong> plantain and banana in Sub-saharan Africa. African Crop Science<br />

Journal 1(1):1-8.<br />

Vuylsteke D., R. Ortiz and R. Swennen. 1994. Breeding plantain hybrids for resistance to Black<br />

Sigatoka. IITA Research 8:9-14.<br />

Vuylsteke D., R. Ortiz, S. Ferris and R. Swennen. 1995. ‘PITA-9’: a black-sigatoka-resistant<br />

triploid hybrid from the ‘False Horn’ plantain gene pool. HortScience 30(2):395-397.<br />

Wiame I., R. Swennen and L. Sági. 2000. PCR-based cloning <strong>of</strong> candidate disease resistance<br />

genes from banana (Musa acuminata). Acta Horticulturae 521:51–57.<br />

237


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Wong H. C., Y. Ting, H.-C. Lin, F. Reichert, K. Myambo, K. Watt, P.L. Toy, and R.J. Drummond.<br />

1991. Genetic organization and regulation <strong>of</strong> the xylose degradation genes in Streptomyces<br />

rubiginosus. J. Bacteriol. 173:6849-6858.<br />

Zhu Y., H. Chen, J. Fan, Y. Wang, Y. Li, J. Chen, J. Fan, S. Yang, L. Hu, H. Leung, T.W. Mew,<br />

P.S. Teng, Z. Wang and C.C. Mundt. 2000. Genetic diversity and disease control in rice.<br />

Nature 406:718-722.<br />

238


Session 4<br />

N. Roux et al.<br />

Mutagenesis and somaclonal<br />

variation to develop new resistance<br />

to <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

N. Roux 1 ,A.Toloza 1 ,J.P.Busogoro 2 ,B.Panis 3 ,<br />

H. Strosse 3 ,P.Lepoivre 2 ,R.Swennen 3 and F. J. Zapata-Arias 1<br />

Abstract<br />

<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> can reduce fruit yield by up to 50%. Chemical strategies exist<br />

to combat these <strong>diseases</strong>, but they are environmentally unsound, hazardous and very expensive<br />

for many farmers. The only sustainable means to reduce the use <strong>of</strong> pesticides is breeding for<br />

tolerant cultivars. Whereas breeders are looking for genetic variability to develop new varieties,<br />

edible Musa cultivars are multiplied vegetatively, polyploid and sterile. Spontaneous somatic<br />

mutation has already contributed largely to obtaining new cultivars <strong>of</strong> Musa.Nevertheless, the<br />

rate <strong>of</strong> occurrence is too low to satisfy practical breeding needs. Mutations can be induced by<br />

physical or chemical mutagens. With the development <strong>of</strong> tissue culture techniques, in vitro<br />

mutagenesis and somaclonal variation raised hopes in the 1980-1990s. In spite <strong>of</strong> this, very few<br />

useful and stable mutants/somaclones were obtained.The multicellular structure <strong>of</strong> meristems<br />

which leads to chimerism is certainly an impeding factor. Additionally, the random process in<br />

mutation induction calls for the screening <strong>of</strong> several thousand plants after treatment. Recently,<br />

following a five-year FAO/IAEA/DGIC coordinated research project, it has been possible to<br />

overcome these two barriers by mutagenic treatment <strong>of</strong> embryogenic cell suspensions and by<br />

establishing an early mass screening method resting on infiltration <strong>of</strong> juglone, a toxic metabolite<br />

<strong>of</strong> <strong>Mycosphaerella</strong> fijiensis. After screening approximately 4000 plants, 15 putative mutants<br />

showed tolerance to this metabolite. These plants must be evaluated for their resistance to<br />

M. fijiensis infection under controlled conditions and field experiment.<br />

Resumen - Mutagénesis y variación somaclonal para desarrollar nueva resistencia a<br />

las enfermedades de la mancha foliar por <strong>Mycosphaerella</strong><br />

Las enfermedades de las manchas foliares causadas por <strong>Mycosphaerella</strong> spp. afectan<br />

significativamente el cultivo bananero y puede reducir el rendimiento de la fruta en hasta un<br />

1<br />

International Atomic Energy Agency Laboratories, Seibersdorf, Austria<br />

2<br />

Agricultural University Gembloux, Gembloux, Belgium<br />

3<br />

Katholieke Universiteit Leuven,Leuven, Belgium<br />

239


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

50%. Existen estrategias de control químico para combatir estas enfermedades, pero estas causan<br />

daños al ambiente, son peligrosas y muy costosas para muchos agricultores. El único medio<br />

sostenible para reducir el uso de plaguicidas es el mejoramiento de los cultivares tolerantes. Los<br />

mejoradores que están investigando la variabilidad genética para desarrollar nuevas variedades,<br />

toman en cuenta que las variedades comestibles de Musa se multiplican vegetativamente, son<br />

poliploides y estériles. La mutación somática espontánea es una fuente de variación que ya ha<br />

contribuido en gran medida en la obtención de las nuevas variedades en Musa spp. Sin embargo,<br />

la tasa de ocurrencia es muy baja para satisfacer las necesidades prácticas de mejoramiento. Las<br />

mutaciones pueden ser inducidas por mutágenos físicos o químicos. Con el desarrollo de las<br />

técnicas del cultivo de tejidos, la mutagénesis y variación somaclonal in vitro elevaron las<br />

esperanzas en las décadas de los 80 y 90. A pesar de esto, se obtuvieron pocos mutantes o<br />

somaclones útiles y estables. La estructura multicelular de los meristemas que conduce al<br />

quimerismo es ciertamente un factor de impedimento. Adicionalmente, el proceso aleatorio para<br />

inducir la mutación requiere el cribado de varios miles de plantas después del tratamiento.<br />

Recientemente, después de realizar un proyecto de investigación de cinco años coordinado por<br />

FAO/IAEA/DGCI, fue posible vencer estas dos barreras mediante un tratamiento mutagénico de<br />

las suspensiones de células embriogénicas y el establecimiento de un método de cribado<br />

masivo temprano que se basa en la infiltración de juglone, un metabolito tóxico de <strong>Mycosphaerella</strong><br />

fijiensis.. Después de realizar el cribado de aproximadamente 4000 plantas, 15 mutantes putativos<br />

mostraron tolerancia a este metabolito.<br />

Résumé – Mutagénèse et variation somaclonale pour développer de nouvelles<br />

résistances aux maladies foliaires causées par <strong>Mycosphaerella</strong> spp.<br />

Les maladies foliaires causées par <strong>Mycosphaerella</strong> spp. peuvent réduire le rendement de jusqu’à<br />

50%. Des moyens chimiques existent pour lutter contre ces maladies, mais ils sont nocifs pour<br />

l’environnement, dangereux et très coûteux pour nombre de fermiers. La seule manière durable<br />

de réduire le recours aux insecticides est de créer des cultivars tolérants. Sauf que pour y arriver,<br />

les sélectionneurs ont besoin de variabilité génétique et que les cultivars de bananiers sont<br />

polyploïdes, stériles et multipliés végétativement. Les mutations somatiques spontanées ont<br />

déjà passablement contribuées à l’obtention de nouveaux cultivars, mais leur fréquence est trop<br />

faible pour satisfaire les besoins des sélectionneurs. Des agents mutagènes chimiques et<br />

physiques peuvent provoquer des mutations. Dans les années 1980-1990, le développement des<br />

techniques de culture de tissus, mutagénèse in vitro et variation somaclonale ont soulevé bien<br />

des espoirs. Malgré cela, peu de mutants/somaclones ont été obtenus. La structure multicellulaire<br />

des méristèmes, qui produit des chimères, est sans doute un facteur limitant. De plus,<br />

l’induction de mutations est un processus aléatoire qui nécessite le criblage de plusieurs milliers<br />

de plants. Récemment, suite à un projet de recherche de cinq ans coordonné par FAO/IAEA/DGIC,<br />

il a été possible de contourner ces deux obstacles en faisant subir un traitement mutagène à<br />

des suspensions de cellules embryogéniques et en mettant au point une méthode de criblage<br />

précoce basée sur l’infiltration de juglone, un métabolite toxique de <strong>Mycosphaerella</strong> fijiensis.<br />

Après avoir criblé 4000 plants, 15 mutants présomptifs ont montré une tolérance à ce<br />

métabolite. Ces plants doivent être évalués pour leur résistance à M. fijiensis sous des conditions<br />

contrôlées et en champ.<br />

Introduction<br />

Bananas and plantains (Musa spp.) are a staple food <strong>of</strong> millions <strong>of</strong> people and rank<br />

among the top five food commodities. However, <strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

can reduce fruit yield by 50% (Mourichon et al., 1997). Chemical control stategies<br />

exist to combat these <strong>diseases</strong>, but they are environmentally unsound, hazardous<br />

and very expensive for many farmers (Persley and George, 1999). Breeding <strong>of</strong> resistant<br />

240


Session 4<br />

N. Roux et al.<br />

cultivars is the only sustainable mean to reduce the use <strong>of</strong> pesticides. But edible Musa<br />

are difficult to breed since they are polyploid and sterile.<br />

Spontaneous mutations are at the origin <strong>of</strong> almost all <strong>of</strong> the edible banana and<br />

plantain cultivars (Buddenhagen, 1987). The best example is the spontaneous banana<br />

mutant ‘Cavendish’ from Vietnam, which is resistant to Fusarium wilt (race 1) and<br />

which replaced ‘Gros Michel’ in the 1950s and 1960s (Ploetz, 1994). The discovery<br />

<strong>of</strong> this banana mutant saved the banana industry from collapsing but as a<br />

consequence, the export trade relies on a very narrow genetic base since only one<br />

or two triploid cultivars <strong>of</strong> the subgroup Cavendish dominate the export market<br />

(Risède and Tézenas du Montcel, 1997).<br />

Banana is probably the best example in the history <strong>of</strong> agriculture <strong>of</strong> the<br />

pathological perils <strong>of</strong> monoclone culture. Indeed, without clonal diversification, the<br />

trade can hardly be expected to survive indefinitely and the generation and use <strong>of</strong><br />

genetic variability may be the only remaining option. Spontaneous somatic mutation<br />

has already contributed largely in obtaining new cultivars in Musa spp. Nevertheless,<br />

the rate <strong>of</strong> occurrence is too low to satisfy practical breeding needs. Mutations can<br />

be induced by tissue culture (somaclonal variation) and/or by physical or chemical<br />

mutagens (induced mutants). With the development <strong>of</strong> tissue culture techniques, in<br />

vitro mutagenesis and somaclonal variation raised hopes in the 1980 and 1990s. So<br />

far, however, very few useful and stable mutants/somaclones have been obtained.<br />

The multicellular structure <strong>of</strong> meristems, which leads to chimerism is certainly an<br />

impeding factor. Additionally, the random process <strong>of</strong> mutation induction calls for<br />

the screening <strong>of</strong> several thousand plants after treatment. Recently, following a fiveyear<br />

FAO/IAEA/DGIC coordinated research project, it has been possible to overcome<br />

these two barriers. This paper presents the improved methodology and the potential<br />

use <strong>of</strong> mutants in genetic improvement programmes.<br />

Advantages and limitations <strong>of</strong> induced mutation<br />

and somaclonal variation<br />

Somaclonal variants<br />

In vitro propagated plants are not necessarily true to type. Off-type plants might<br />

differ permanently (i.e. somaclonal variation) or temporarily from the source plant<br />

as a result <strong>of</strong> an epigenetic or physiological effect. The term ‘somaclonal variation’<br />

was introduced to describe the genetic variation in plants regenerated from any form<br />

<strong>of</strong> cell culture. Larkin and Scowcr<strong>of</strong>t (1981) advocate the view that somaclonal<br />

variation represents a new source <strong>of</strong> variability and therefore constitutes a powerful<br />

tool to the breeder. Nevertheless somaclonal variation from micropropagated banana<br />

and plantain should not be overestimated as a source <strong>of</strong> novel variability for use in<br />

genetic improvement (Vuylsteke et al., 1991). A narrow spectrum <strong>of</strong> variants has<br />

been obtained through somaclonal variation. It is becoming increasingly clear that<br />

somaclonal variation is usually undesirable (Vuylsteke et al., 1996).<br />

Some <strong>of</strong>f-types have improved agronomical traits, such as the higher yield <strong>of</strong><br />

the ‘French reversion’ variant plantain and the short stature <strong>of</strong> dwarfs (Vuylsteke et<br />

al., 1996). Regarding disease resistance, a ‘Cavendish’ banana was recovered in Taiwan<br />

241


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

for its tolerance to Fusarium wilt (Hwang et al., 1993). More recently Trujillo et al.<br />

(1996) reported the selection <strong>of</strong> a somaclonal variant CIEN BTA-03 tolerant to<br />

Sigatoka disease. But cytogenetic and molecular characterization revealed that CIEN<br />

BTA-03 was in fact a tetraploid clone and was not part <strong>of</strong> the Cavendish subgroup<br />

to which the parental ‘Williams’ (AAA) belongs (Gimenez et al., 2001).<br />

Induced mutants<br />

The Musa mutation induction system based on in vitro techniques to recover mutant<br />

plants and micropropagate desirable mutants was developed by Novák et al. (1990)<br />

in the FAO/IAEA Laboratories. It is now applied in a few Musa breeding programmes.<br />

Gamma irradiation is the main physical mutagen used to induce genetic variation.<br />

More recently, Roux (1997) standardized the methodology to provide guidelines to<br />

mutation induction programmes in Musa spp. Shoot-tips excised from clones<br />

representing the different genomic constitutions <strong>of</strong> the genus Musa were treated with<br />

10 doses from 10 to 100 Gy <strong>of</strong> a 60 Co gamma irradiation source at a dose rate <strong>of</strong> 44<br />

Gy/min. For each Musa accession, 200 explants were treated for sensitivity testing<br />

and 20 non-irradiated explants were used as control. Radiation sensitivity and postirradiation<br />

recovery were assessed by measuring the survival rate, the propagation<br />

rate, the shoot height and the fresh weight.<br />

The different Musa accessions showed different responses according to their ploidy<br />

level and genomic constitution. The following ranges <strong>of</strong> doses are recommended:<br />

• 10 to 20 Gy <strong>of</strong> gamma irradiation for diploid clones ‘Calcutta 4’ (AA) and<br />

‘Tani’ (BB);<br />

• 30 to 40 Gy <strong>of</strong> gamma irradiation for the triploids ‘Three hand planty’ (AAB),<br />

‘Grande naine’ (AAA), ‘Williams’ and ‘Kamaramasenge’ (triploid, formerly<br />

classified as AB);<br />

• 40 to 50 Gy <strong>of</strong> gamma irradiation for the triploid ‘Cachaco’ (ABB).<br />

From the FAO/IAEA mutant varieties database, two banana accessions were<br />

registered as improved mutant varieties: ‘Novaria’ for early flowering and ‘Klue hom<br />

thong KU1’ for its bunch size and cylindrical shape from which larger <strong>bananas</strong> can<br />

be selected.<br />

Other desirable variants/putative mutants have been identified for release or<br />

further confirmation trials. Examples are shown in Table 1.<br />

Most <strong>of</strong> the improved characteristics are agronomic features. Disease resistance<br />

seems to be difficult to obtain through mutation induction techniques. Consequently,<br />

Smith et al. (1995) used an original strategy. Instead <strong>of</strong> irradiating an agronomically<br />

superior but susceptible genotype, they irradiated ‘Dwarf parfitt’, an extra dwarf<br />

Cavendish banana that has shown a high level <strong>of</strong> resistance to race 4 <strong>of</strong> Fusarium<br />

wilt. Following radiation, 35 M 1<br />

V 3<br />

(M: Mutagenic treatment; V: vegetative generation)<br />

out <strong>of</strong> 500 explants irradiated at 20 Gy were recovered that possessed improved<br />

agronomic characteristics (taller plant size, increased yield and no choking). Most<br />

importantly these selections appeared to retain the resistance to race 4 derived from<br />

the mother plant ‘Dwarf parfitt’.<br />

242


Session 4<br />

N. Roux et al.<br />

Bhagwat and Duncan (1998) used gamma irradiation (8 to 20 Gy) and chemical<br />

mutagens to make ‘Highgate’ (AAA) tolerant to Fusarium oxysporum f.sp. cubense.<br />

Twelve weeks after inoculation in the greenhouse, 0.3 to 0.9% <strong>of</strong> the regenerated<br />

plants derived from irradiated explants and 1.9 to 6.1% <strong>of</strong> chemically treated explants<br />

had less than 10% vascular invasion in their corms with no external symptoms. These<br />

plants were considered tolerant to the fungus and were multiplied, ex vitro, for field<br />

screening.<br />

Table 1. List <strong>of</strong> putative mutants obtained.<br />

Country Parent clone Selected clone Selected traits Technique Place <strong>of</strong> induction<br />

Cuba SH-3436 SH-3436-L9 Reduced height Gamma rays Cuba<br />

Parecido al Rey 6.44 Reduced height Gamma rays IAEA<br />

Malaysia Pisang rastali Mutiara FOC* tolerance Somaclonal variation Malaysia<br />

Novaria FOC tolerance Somaclonal variation Malaysia<br />

Philippines Lakatan LK-40 Reduced height Gamma rays IAEA<br />

Latundan LT-3 Larger fruit size Gamma rays IAEA<br />

Sri Lanka Embul Embul-35 Gy Earliness Gamma rays Sri Lanka<br />

Austria Grande naine GN35-I to Tolerance to toxin Gamma rays IAEA<br />

(IAEA**) GN35-VIII from <strong>Mycosphaerella</strong><br />

fijiensis<br />

* FOC: Fusarium oxysporum f.sp. cubense<br />

**IAEA: International Atomic Energy Agency.<br />

Even though the traditional shoot tip mutation induction technique has permitted<br />

to obtain useful mutants, the following limitations are impeding its wider use:<br />

• The treatment <strong>of</strong> shoot-tips with mutagenic agents (physical or chemical) results<br />

in a high degree <strong>of</strong> chimerism. This is a serious obstacle to mutation techniques<br />

since it is not yet possible to distinguish mutated cells from none mutated cells.<br />

• Since mutation induction is a random process, efficiency requires the need to<br />

treat and screen as many plants as possible. However, a bottleneck occurs due<br />

to the time spent on field screening. The current methods <strong>of</strong> field screening are<br />

also site-specific and involve considerable resources: large numbers <strong>of</strong> technicians,<br />

hours <strong>of</strong> work, fertilizer, logistic support and high cost.<br />

Recent technical achievements<br />

Origin <strong>of</strong> embryogenic cell suspensions<br />

In order to screen efficiently for characters such as disease resistance, an efficient<br />

method to overcome chimerism after mutagenic treatment is needed (Roux et al.,<br />

2001). Considering this, somatic embryogenesis is the most promising method since<br />

somatic embryos are assumed to be <strong>of</strong> single cell origin (Halperin, 1966). In some<br />

species, histological studies confirmed the single cell origin <strong>of</strong> somatic embryos. They<br />

develop either directly from an explant or as secondary embryos at the surface <strong>of</strong><br />

older somatic embryos (Litz and Gray, 1992). Grapin et al. (1998) stipulated from<br />

cytological studies on somatic embryo ontogenesis in Musa that a unicellular origin<br />

was more than likely. However such studies can only be performed on few somatic<br />

243


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

embryos and thus the extrapolation <strong>of</strong> these findings to a large number <strong>of</strong> somatic<br />

embryos is risky.<br />

Colchicine treatment and ploidy analysis with flow cytometry, used by Roux et<br />

al. (2001) to monitor the efficiency <strong>of</strong> 3 different micropropagation techniques in<br />

dissociating chimeras, was applied to verify the unicellular origin <strong>of</strong> somatic embryos<br />

from ECS. After treating cell suspensions with colchicine (Figure 1), the embryos<br />

were subsequently transferred on a regeneration medium in test tubes. As soon as<br />

green plantlets with shoot and roots were obtained, leaves pieces <strong>of</strong> 0.5 cm 2 were<br />

excised and their ploidy measured through flow cytometry before acclimatization<br />

(Table 2). The majority <strong>of</strong> the regenerated plants were triploid. Among the treated<br />

cells, the proportion <strong>of</strong> regenerated hexaploid plants (5.3%) remained very low<br />

compared to triploid plants. We think that triploid cells have a comparative<br />

advantage over induced hexaploid cells during culture. In contrast to shoot-tip<br />

cultures that were treated with colchicine (Roux et al., 2001), no mixoploids among<br />

the regenerated plants were observed, which confirms the single cell origin <strong>of</strong><br />

embryos. Thus embryogenic cell suspensions seem to be the material <strong>of</strong> choice for<br />

mutagenic treatments.<br />

100<br />

Ploidy frequency distribution (% <strong>of</strong> nuclei)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0%<br />

0.05% 0.1%<br />

0.2%<br />

3x<br />

6x<br />

Colchicine concentration<br />

Figure 1. Ploidy frequency distribution <strong>of</strong> ‘Williams’ suspension cells (cell-line 124T) with triploid (3x) and<br />

hexaploid (6x) nuclear DNA content observed 15 days after colchicine treatment (w/v).<br />

244


Session 4<br />

N. Roux et al.<br />

Table 2. Ploidy distribution <strong>of</strong> regenerated plants from colchicines-treated ECS <strong>of</strong> the triploid cell line 124T.<br />

Colchicine Number <strong>of</strong> regenerated Ploidy<br />

concentration (%)<br />

plants<br />

3n 6n Chimeric mixoploidy (3n+6n)<br />

0 108 108 0 0<br />

0.05 63 58 5 0<br />

0.1 37 36 1 0<br />

0.2 88 84 4 0<br />

296 286 10 0<br />

Mutagenic treatment <strong>of</strong> embryogenic cell suspensions<br />

Two cell lines <strong>of</strong> the cultivars ‘Williams’ and ‘Three hand planty’ were sieved using a<br />

1000-mm pore size mesh to obtain a fine suspension. The ECS were then subcultured<br />

in the maintenance medium (ZZ) in 100-ml flasks at a concentration <strong>of</strong> about 3% <strong>of</strong><br />

settled cell volume (SCV). Three days after subculture 0.5 ml <strong>of</strong> cells were transferred<br />

to a sterile Petri dish and the medium was removed. The cells aggregates were then<br />

irradiated at doses ranging from 0 to 250 Gy with 25 Gy intervals using a 60 Co gamma<br />

source at a dose rate <strong>of</strong> 30 Gy/min. After irradiation, the cells were resuspended in<br />

maintenance ZZ medium in centrifuge tubes and transferred to 100-ml Erlenmeyer flasks<br />

at different quantities according to the parameter to be analyzed. To study the effect<br />

<strong>of</strong> gamma radiation on the growth <strong>of</strong> ECS, fresh weight gain and regeneration capacity<br />

were determined. The results were expressed in percentage <strong>of</strong> the control (non-irradiated<br />

cells) at all doses.<br />

The two radiosensitivity curves for ‘Williams’ and ‘Three hand planty’ are quite similar.<br />

After 28 days, at 75 Gy, the cells’ weight gain (CWG 75<br />

=0,84g) was 50% <strong>of</strong> the control<br />

(CWG 0<br />

= 1,68g) (Figure 2A).<br />

To measure the regeneration capacity, green plantlets were counted and transferred<br />

to Magenta GA7 boxes containing semi-solid R 3<br />

regeneration medium for further growth<br />

before acclimatization. The radiosensitivity curve was obtained by comparing the number<br />

<strong>of</strong> regenerated plantlets for each dose with the control plants (from non-irradiated ECS)<br />

(Figure 2B). Radiation at a low level seems to stimulate the regeneration capacity<br />

especially in ‘Williams’. We must, however, take into account that in control plants, the<br />

density <strong>of</strong> embryos in the temporary immersion system vessels was too high and hence,<br />

a considerable number <strong>of</strong> small plantlets could not develop. In both genotypes no plants<br />

regenerated above 200Gy. The number <strong>of</strong> regenerated plants drops drastically above<br />

50 Gy for ‘Williams’ whereas for ‘Three Hand Planty’ the number <strong>of</strong> regenerated plantlets<br />

decreases less dramatically. The regenerated plants were transferred to the greenhouse<br />

for early mass screening for tolerance to black <strong>leaf</strong> streak disease.<br />

Establishment <strong>of</strong> an early mass screening method<br />

Genetic variability is a prerequisite before selecting for disease resistance. A technique<br />

that can reliably identify resistant plants is then adopted to screen the populations<br />

(Lepoivre et al., 1993). The Plant Pathology Unit at the Faculté Universitaire des Sciences<br />

245


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

140<br />

Fresh weight gain (% <strong>of</strong> control)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

WIL<br />

THP<br />

0<br />

0 50 100 150 200 250<br />

Dose (Gy)<br />

Regeneration capacity (% <strong>of</strong> control)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

WIL<br />

THP<br />

0<br />

0 50 100 150 200 250<br />

Dose (Gy)<br />

Figure 2. Effect <strong>of</strong> gamma radiation on : (A) fresh weight gain 28 days after irradiation and (B) regeneration<br />

capacity at 60 to 90 days after irradiation. All parameters are expressed as a percentage <strong>of</strong> the non-irradiated<br />

ECS (control). Cell suspension cultures <strong>of</strong> the cultivars ‘Willliams’ (WIL) and ‘Three hand planty’ (THP) were<br />

used. For each treatment the equivalent <strong>of</strong> four Petri dishes were measured.<br />

Agronomiques de Gembloux (FUSAGx) developed a rapid and early screening protocol<br />

for tolerance to black <strong>leaf</strong> streak disease in banana and plantain. The method is based<br />

on the infiltration <strong>of</strong> juglone (5-hydroxy-1,4-naphthoquinone), a toxic metabolite <strong>of</strong><br />

<strong>Mycosphaerella</strong> fijiensis. After developing different bioassays, it was concluded that slow<br />

lesion development in cultivars exhibiting a partial resistance to black <strong>leaf</strong> streak disease<br />

was correlated with lower sensitivity to M. fijiensis toxins (Lepoivre, 1995). The toxin<br />

is probably not involved in the initiation <strong>of</strong> the infection but could serve as a secondary<br />

determinant <strong>of</strong> the pathogenecity, contributing to the lesion expansion in cultivars<br />

exhibiting partial resistance to black <strong>leaf</strong> streak disease (Harelimana et al., 1997). The<br />

prospect <strong>of</strong> utilizing plant tissue cultures to generate and identify novel genetic variants<br />

has sparked the interest <strong>of</strong> researchers for many years (Dix, 1996). Nevertheless in Musa,<br />

in vitro heterotrophic tissues are not suitable targets to perform the screening with such<br />

toxin (Harelimana et al., 1997). Our first goal was thus, under greenhouse conditions,<br />

to determine the lowest concentration <strong>of</strong> juglone which enables the differentiation<br />

between the susceptible cultivar ‘Grande naine’ and the partially resistant cultivar<br />

‘Fougamou’. After performing 10 assays with 4 to 8 replicates we concluded that 25<br />

ppm was the most suitable concentration <strong>of</strong> juglone to distinguish between a partially<br />

resistant and a susceptible plant. This dose was thus further used to screen the plants<br />

regenerated from irradiated shoot tips.<br />

Four batches (100 meristems/batch) <strong>of</strong> ‘Grande naine’ were irradiated at 35 Gy and<br />

propagated over four subcultures. The plants were then acclimatized in the greenhouse.<br />

The early mass screening method can be divided in three steps:<br />

1. Plant preparation: The acclimatized plants which reached the six <strong>leaf</strong> stage are<br />

maintained at 90 to 100% relative humidity for 48 hours to open the stomata.<br />

246


Session 4<br />

N. Roux et al.<br />

2. Infiltration: The second open <strong>leaf</strong> <strong>of</strong> each plant was infiltrated on its under surface<br />

with juglone (25 ppm), crude extract (500 ppm) and 10% methanol (control). The<br />

amount <strong>of</strong> each infiltration was 20µl.<br />

3. Necrose observation: 48 hours after infiltration, the plants were observed<br />

for necrosis and compared with control plants (non-irradiated ‘Grande naine’<br />

= positive control; non-irradiated ‘Fougamou’ = negative control).<br />

To date, from a total <strong>of</strong> 3 728 ‘Grande naine’ plants screened, 15 putative mutants<br />

(0.4%) were selected for their tolerance to 25 ppm <strong>of</strong> juglone (Table 3).<br />

Table 3. Inventory <strong>of</strong> irradiated meristems from the cultivar ‘Grande naine’, multiplied, regenerated into plants, screened<br />

and retained for their partial resistance to 25ppm <strong>of</strong> juglone.<br />

Batch Irradiated Number. <strong>of</strong> Number <strong>of</strong> plants Number <strong>of</strong><br />

meristems shoots in M 1<br />

V 1<br />

screened in M 1<br />

V 4<br />

plants retained<br />

A(ST) 100 142 780 8<br />

B (ST) 100 105 512 0<br />

C (MA) 100 110 1351 4<br />

D (MA) 100 140 1085 3<br />

TOTAL 400 497 3728 15<br />

ST: propagated by shoot tip culture; MA: propagated by the multi-apexing culture technique<br />

M x<br />

: Mutagenic treatment; V x<br />

: Vegetative generation.<br />

The two first batches were propagated by shoot-tip culture and the two second<br />

batches were propagated by the multi-apexing technique to dissociate more efficiently<br />

chimeras (Table 3). Among the young banana plants selected for their tolerance to<br />

juglone, some were showing an increased content <strong>of</strong> anthocyanin (Figure 3).<br />

Even though we may not have directly obtained mutant genes controlling for<br />

resistance, genes responsible for anthocyanin biosynthesis may have been activated<br />

and indirectly provided tolerance to black <strong>leaf</strong> streak disease. Atanassova et al. (2001)<br />

studied the effect <strong>of</strong> mutations affecting anthocyanin biosynthesis during tomato and<br />

pepper development under stress conditions. They found four genes, which had a kind<br />

<strong>of</strong> universal effect on tomato and pepper germination as they increased the germination<br />

potential <strong>of</strong> the individual accession under a relatively large range <strong>of</strong> stresses.<br />

Figure 3.<br />

Regenerated plants from irradiated<br />

shoot tips <strong>of</strong> the cultivar ‘Grande naine’:<br />

A) susceptible to 25 ppm <strong>of</strong> juglone;<br />

B) putative mutant, tolerant to 25 ppm<br />

<strong>of</strong> juglone.<br />

247


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Thus the establishment <strong>of</strong> morphological marker correlated with tolerance to stress<br />

even at one plant growth stage, could be useful. The pre-selected plants and their<br />

suckers were transferred to the FUSAGx for further screening by inoculation under<br />

control conditions.<br />

Conclusion and future directions<br />

Over the last five years, considerable efforts have been assembled to speed up<br />

the in vitro mutagenesis process and make it more efficient. Mutation induction<br />

techniques as well as other genetic improvement strategies will benefit from the<br />

use <strong>of</strong> embryogenic cell suspensions and the establishment <strong>of</strong> early mass<br />

screening techniques. Tolerance to <strong>Mycosphaerella</strong> fijiensis obtained after<br />

mutagenic treatment still needs to be confirmed under field conditions. Such<br />

mutants may not have the required agronomic characters but they would be very<br />

useful in genetic studies and may help in discovering genes responsible for<br />

resistance or susceptibility.<br />

Mutation induction should no longer be seen as an independent genetic<br />

improvement strategy but more as a tool which can contribute to functional<br />

genomics and genetic improvement programmes based on cross-breeding or<br />

genetic transformation. For example, a disease resistant mutant from a diploid<br />

Musa, could be used as a parent line in cross-breeding programmes and also help<br />

in understanding the mechanism <strong>of</strong> resistance and permit the isolation <strong>of</strong> genes<br />

to be used in genetic transformation.<br />

Acknowledgements<br />

The authors wish to thank Ms. Ines Van den Houwe, (INIBAP) for providing the vegetative<br />

clones <strong>of</strong> Musa. This work was supported by a Joint FAO/IAEA/GDIC (Belgian General<br />

Direction for International Cooperation) Coordinated Research Project. The study was<br />

undertaken as part <strong>of</strong> the Global Programme for Musa Improvement (PROMUSA).<br />

References<br />

Atanassova B., S. Daskalov, L. Shtereva and E. Balatcheva. 2001. Anthocyanin mutations<br />

improving tomato and pepper tolerance to adverse climatic conditions. Euphytica<br />

120:357-365.<br />

Bhagwat B. and E.J. Duncan. 1998. Mutation breeding <strong>of</strong> Highgate (Musa acuminata, AAA)<br />

for tolerance to Fusarium oxysporum f.sp. cubense using gamma irradiation. Euphytica<br />

101:143-150.<br />

Buddenhagen I.W. 1987. Disease susceptibility and genetics in relation to breeding <strong>of</strong> <strong>bananas</strong><br />

and plantains. Pp. 95-109 in Banana and Plantain Breeding Strategies (Persley G.J. and<br />

E.A. De Langhe , eds). ACIAR Proceedings 21, Canberra, Australia.<br />

Gimenez C., E. Garcia, N.X. Enrech and I. Blanca. 2001. Somaclonal variation in banana:<br />

Cytogenetic and molecular characterization <strong>of</strong> the somaclonal variant CIEN BTA-03.<br />

In Vitro Cell. Dev. Biol.-Plant 37:217-222.<br />

248


Session 4<br />

N. Roux et al.<br />

Grapin A., J.L Ortiz, R. Domergue, J. Babeau, S. Monmarson, J.V. Escalant, C. Teisson and F.<br />

Côte. 1998. Establishment <strong>of</strong> embryogenic callus and initiation and regeneration <strong>of</strong><br />

embryogenic cell suspensions from female and male immature flowers <strong>of</strong> Musa.<br />

INFOMUSA 7(1):13-15.<br />

Halperin W. 1966. Alternative morphogenetic events in cell suspensions. American Journal<br />

<strong>of</strong> Botany 53:443-453.<br />

Harelimana G., P. Lepoivre, H. Jijakli and X. Mourichon. 1997. Use <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

toxins for the selection <strong>of</strong> banana cultivars resistant to Black Leaf Streak. Euphytica 96:125-<br />

128.<br />

Hwang S.C., W.H. Ko and C.P. Chao. 1993. GCTCV-215-1: a promising Cavendish clone resistant<br />

to race 4 <strong>of</strong> Fusarium oxysporum f.sp. cubense. Pp. 62-74 in Recent Developments in<br />

Banana Cultivation Technology (R.V. Valmayor, S.C. Hwang, R. Ploetz, S.W. Lee and<br />

V.N. Roa, eds). Proceedings, International Symposium, Chiuju, Pingtung, Taiwan, 14-18<br />

December 1992. INIBAP, Los Banos, Philippines.<br />

Larkin P.J. and W.R. Scowcr<strong>of</strong>t. 1981. Somaclonal variation - a novel source <strong>of</strong> variability<br />

from cell culture for plant improvement. Theoretical and Applied Genetics 60:197-214.<br />

Lepoivre P. 1995. Development <strong>of</strong> screening procedures for resistance to black <strong>leaf</strong> streak<br />

disease in banana and plantain. End <strong>of</strong> mission report, IAEA, Vienna 16pp.<br />

Lepoivre P., C.P. Acuna and A.S. Riveros. 1993. Screening procedures for improving resistance<br />

to banana black <strong>leaf</strong> streak disease. Pp. 213-220 in Breeding banana and Plantain for<br />

Resistance to Diseases and Pests (J. Ganry, ed.). CIRAD, Montpellier, France.<br />

Litz R.E. and D.J. Gray. 1992. Organogenesis and Somatic Embryogenesis. Pp. 3-34 in<br />

Biotechnology <strong>of</strong> perennial Fruit Crops (Hammerschlag F.A. and Litz R.E. eds). CAB<br />

International, Wallingford, Oxon. U.K.<br />

Mourichon X., J. Carlier and E. Fouré. 1997. Les cercosporioses. Musa Disease Fact Sheet<br />

n o 8, INIBAP, Montpellier, France.<br />

Novak F.J., R. Afza, M. van Duren and M.S. Omar. 1990. Mutation induction by gamma<br />

irradiation <strong>of</strong> in vitro cultured shoot-tips <strong>of</strong> banana and plantain (Musa cvs). Tropical<br />

Agriculture (Trinidad) 67(1):21-28.<br />

Persley G.J. and P. George. 1999. Commodity Advances through Banana Improvement<br />

Research, 1994-1998. Environmentally and socially sustainable development, agricultural<br />

research and extension group series, The World Bank, Washington, D.C. 62pp.<br />

Ploetz R.C. 1994. Panama disease: Return <strong>of</strong> the first banana menace. International Journal<br />

<strong>of</strong> Pest Management 40:326-336.<br />

Risede J.M. and H. Tezenas du Montcel. 1997. Banana monocultures and environmental<br />

protection: assessment and perspectives. Fruits 52(4):225-232.<br />

Roux N. 1997. Improved methods to increase diversity in Musa using mutation and tissue<br />

culture techniques. Pp. 49-56 in Report <strong>of</strong> the second Research Co-ordination Meeting<br />

<strong>of</strong> FAO/IAEA/BADC Co-ordinated Research Project, Kuala Lumpur. IAEA, Vienna, Austria.<br />

Roux N.S., J. Dolezel, R. Swennen and F.J. Zapata-Arias. 2001. Effectiveness <strong>of</strong> three<br />

micropropagation techniques to dissociate cytochimeras in Musa sp. Plant Cell, Tissue<br />

and Organ Culture 66:189-197.<br />

Sharrock S. and E. Frison. 1999. Musa production around the world - trends, varieties and<br />

regional importance. Pp. 42-47 in INIBAP Annual Report 1998, focus paper 2. INIBAP,<br />

Montpellier, France.<br />

Smith M.K., S.D. Hamill, P.W. Langdon and Pegg, K.G. 1995. In vitro mutation breeding for<br />

the development <strong>of</strong> <strong>bananas</strong> with resistance to race 4, fusarium wilt (Fusarium oxysporum<br />

f.sp. cubense). Pp. 37-44 in Final reports <strong>of</strong> FAO/IAEA Co-ordinated research programme,<br />

TECDOC-800, IAEA, Vienna, Austria.<br />

249


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Trujillo I. and E. Garcia. 1996. Strategies for obtaining somaclonal variants resistant to yellow<br />

sigatoka (<strong>Mycosphaerella</strong> musicola). INFOMUSA 5:12-13.<br />

Vuylsteke D., R. Swennen and E. De Langhe. 1991. Somaclonal variation in plantains (Musa<br />

spp., AAB group) derived from shoot-tip culture. Fruits 46:429-439.<br />

Vuylsteke D., R. Swennen and E. De Langhe. 1996. Field performance <strong>of</strong> somaclonal variants<br />

<strong>of</strong> plantain (Musa spp., AAB group). Journal <strong>of</strong> the American Society for Horticultural<br />

Science 121(1):42-46.<br />

250


Session 4<br />

M. de J.B. Cavalcante et al.<br />

Reaction <strong>of</strong> banana genotypes<br />

to black <strong>leaf</strong> streak disease<br />

in the State <strong>of</strong> Acre in Brazil<br />

M. de J. B. Cavalcante 1 ,A.da S. Ledo 1 ,<br />

F. H. S. Costa 1 ,Z.J.M.Cordeiro 2 ,A.P.Matos 2<br />

Abstract<br />

Black <strong>leaf</strong> streak disease (caused by <strong>Mycosphaerella</strong> fijiensis Morelet) is the most severe disease<br />

affecting commercial varieties <strong>of</strong> banana <strong>of</strong> economic importance in the world. Its occurrence<br />

was verified in Brazil in 1998, in the State <strong>of</strong> Amazonas, and it has spread in the plantations through<br />

out the State <strong>of</strong> the Acre, severely attacking the varieties <strong>of</strong> the Terra subgroup. The objective<br />

<strong>of</strong> this study was to evaluate seven banana genotypes from the Embrapa Mandioca e Fruticultura,<br />

under two cultivation systems (with and without cultural practices) with the goal <strong>of</strong> obtaining<br />

low environmental impact alternatives to control the disease. The evaluations regarding severity<br />

were accomplished on a monthly basis, in ten plants <strong>of</strong> each genotype, using a descriptive scale.<br />

‘FHIA-01’, ‘FHIA-02’, ‘Caipira’, ‘FHIA-21’, ‘PV 42-85’ and ‘Thap maeo’ presented resistance to black<br />

<strong>leaf</strong> streak disease whereas ‘SH 36-40’ proved to be susceptible. There was no significant effect<br />

<strong>of</strong> the cultivation system on the severity <strong>of</strong> black <strong>leaf</strong> streak disease.<br />

Resumen - Reacción de los genotipos de banano a la Sigatoka negra en el estado de<br />

Acre, Brasil<br />

La Sigatoka negra (causada por <strong>Mycosphaerella</strong> fijiensis Morelet) es la enfermedad más severa<br />

que afecta las variedades comerciales de banano de importancia económica en todo el mundo.<br />

La enfermedad se confirmó en Brasil en 1998, en el Estado de Amazonas, y luego se propagó a<br />

las plantaciones a través del Estado de Acre, atacando severamente las variedades del subgrupo<br />

Terra. El objetivo de este estudio consistió en evaluar siete genotipos de banano de Embrapa<br />

Mandioca e Fruticultura, en dos sistemas de cultivo (con y sin empleo de prácticas culturales),<br />

con el fin de obtener alternativas de control para la enfermedad con un bajo impacto ambiental.<br />

Las evaluaciones con respecto a la severidad se realizaron mensualmente en diez plantas de<br />

cada genotipo, utilizando una escala descriptiva. Los resultados mostraron que los genotipos<br />

‘FHIA-01’,‘FHIA-02’,‘Caipira’,‘FHIA-21’,‘PV 42-85’ y ‘Thap maeo’ resultaron resistentes mientras que<br />

el genotipo ‘SH36-40’ fue susceptible a la Sigatoka negra en los dos sistemas de cultivo.<br />

1<br />

Embrapa, Rio Branco, Brazil<br />

2<br />

Embrapa Mandioca e Fruticultura, Cruz das Almas, Brazil<br />

251


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Résumé – Réaction de génotypes de bananiers à la maladie des raies noires dans l’Etat<br />

de Acre, Brésil<br />

La maladie des raies noires (causée par <strong>Mycosphaerella</strong> fijiensis Morelet) est la plus sévère des<br />

maladies qui affectent les cultivars commerciaux économiquement importants. Sa présence au<br />

Brésil a été confirmée en 1998, dans l’Etat d’Amazonas, d’où elle s’est propagée à l’Etat de Acre,<br />

affectant sévèrement les variétés du sous-groupe Terra. L’objectif de cette étude a été d’évaluer<br />

sept génotypes de bananiers de l’Embrapa Mandioca e Fruticultura sous deux systèmes de culture<br />

(avec ou sans pratiques culturales) dans le but de développer des méthodes de lutte contre la<br />

maladie dont l’impact sur l’environnement serait faible. Les criblages furent réalisés sur une base<br />

mensuelle, 10 plants par génotype, et utilisant une échelle descriptive. ‘FHIA-01’, ‘FHIA-02’,<br />

‘Caipira’, ‘FHIA-21’, ‘PV 42-85’ et ‘Thap maeo’ ont présenté une résistance à la maladie des raies<br />

noires, tandis que ‘SH 36-40’ s’est avéré susceptible. Il n’y a pas eu d’effet significatif du système<br />

de culture sur la sévérité de la maladie des raies noires.<br />

Introduction<br />

Black <strong>leaf</strong> streak disease is the most devastating disease <strong>of</strong> banana worldwide. It<br />

can cause losses up to 100% if no control measure is taken (Cordeiro et al., 1998).<br />

In the State <strong>of</strong> Acre, banana is the most consumed fruit and is considered a staple<br />

food among the poor populations. It is also exported to other States. The disease<br />

was observed for the first time in Brazil in early 1998 (Cordeiro et al.,1998), in the<br />

municipalities <strong>of</strong> Tabatinga and Benjamim Constant, and in Rio Branco and<br />

Acrelândia at the end <strong>of</strong> 1998 (Ritzinger et al., 1999; Cavalcante et al., 1999).<br />

Resistant varieties are not only less expensive to control pathogens than<br />

fungicides, they are also preferable from an environmental point <strong>of</strong> view (Pereira et<br />

al., 1999).<br />

The present work aims to evaluate the resistance to black <strong>leaf</strong> streak disease <strong>of</strong><br />

banana cultivars under the weather and soil conditions found in Acre and using two<br />

cultivation systems.<br />

Material and methods<br />

The research was conducted in Embrapa’s experimental station in Acre, Rio Branco.<br />

Seven genotypes (‘PV-4285’, ‘FHIA-21’, ‘Caipira’, ‘FHIA-01’, ‘FHIA-02’, ‘SH-3640’ and<br />

‘Thap maeo’) were evaluated for their resistance to black <strong>leaf</strong> streak disease under<br />

two systems <strong>of</strong> cultivation: a traditional system (weeding) and a more intensive system<br />

(weeding, trimming, shedding and fertilization).<br />

A randomized complete block design (7 genotypes x 2 cultivation systems) with<br />

5 replications was used and data were recorded on 10 leaves/plant.<br />

Disease severity was assessed on a monthly basis, starting from the sixth month<br />

after planting. The disease was observed in individual leaves, using the following<br />

scale (Stover, 1971 modified by Gauhl, 1994):<br />

0 = absence <strong>of</strong> symptoms<br />

1 = less than 1% <strong>of</strong> lamina with symptom (presence <strong>of</strong> streaks and/or > 10 <strong>spot</strong>s)<br />

2 = 1 to 5% <strong>of</strong> lamina with symptoms<br />

3 = 6 to 15% <strong>of</strong> lamina with symptoms<br />

4 = 16 to 33% <strong>of</strong> lamina with symptoms<br />

252


Session 4<br />

M. de J.B. Cavalcante et al.<br />

5 = 34 to 50% <strong>of</strong> lamina with symptoms and<br />

6 = 51 to 100% <strong>of</strong> lamina with symptoms<br />

The variables were submitted to an analysis <strong>of</strong> variance (F test) and the averages<br />

were compared using Scott & Knott’s test (1974) at 1% <strong>of</strong> significance.<br />

Results<br />

According to the analysis <strong>of</strong> variance (Table 1) the genotype had the most significant<br />

effect on disease severity <strong>of</strong> black <strong>leaf</strong> streak disease. There was no interaction<br />

between the cultivation system and the genotypes nor isolated effects <strong>of</strong> the<br />

cultivation system on the severity <strong>of</strong> the disease.<br />

Table 1. Analysis <strong>of</strong> variance <strong>of</strong> disease severity <strong>of</strong> black <strong>leaf</strong> streak disease in seven genotypes grown under two<br />

cultivation systems.<br />

Source <strong>of</strong> Variation DF Average Square<br />

Block 4 51.2714<br />

Genotypes (G) 6 1356.9333*<br />

Cultivation system (CS) 1 66.0571ns<br />

Interaction (G X CS) 6 10.3238ns<br />

Error 52 24.8868<br />

VC (%) 15.30<br />

*Statistically significant at probability 0.01.<br />

As presented in Figure 1, disease severity was highest on ‘SH-3640’ (55,10%),<br />

followed by ‘Thap maeo’ (39%) and ‘FHIA-21’ (33,3%). The genotypes ‘Caipira’,<br />

‘FHIA-01’ and ‘FHIA-02’ were similar whereas the hybrid ‘PV-4285’ presented the<br />

lowest severity (19,70%) (Figure 2).<br />

60<br />

55,1a<br />

50<br />

39b<br />

Disease severity (%)<br />

40<br />

30<br />

20<br />

19,7e<br />

25,8d 26,1d<br />

29,3d<br />

33,5c<br />

10<br />

0<br />

PV-4285<br />

FHIA-02 FHIA-01 Caipira FHIA-21 Thap maeo SH 36-40<br />

Figure 1. Average disease severity <strong>of</strong> black <strong>leaf</strong> streak disease in seven genotypes.<br />

253


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Figure 2.<br />

‘PV 42-85’free <strong>of</strong> black <strong>leaf</strong> streak<br />

disease.<br />

Figure 3.<br />

‘SH 36-40’ showing signs <strong>of</strong> black <strong>leaf</strong><br />

streak disease.<br />

The cultivation system (weeding, shedding, trimming and fertilization) did not<br />

influence the severity <strong>of</strong> black <strong>leaf</strong> streak disease in ‘PV-4285’, ‘Caipira’, ‘FHIA-01’,<br />

‘FHIA-02’, ‘Thap maeo’ and ‘SH-3640’.<br />

For the data recording shooting, it was observed that the genotypes ‘FHIA-01’,<br />

‘FHIA-02’, ‘Caipira’, ‘FHIA-21’, ‘PV 42-85’ and ‘Thap maeo’ were more resistant to<br />

black <strong>leaf</strong> streak disease than ‘SH 36-40’, which was the most susceptible to the<br />

disease (Figure 3).<br />

254


Session 4<br />

M. de J.B. Cavalcante et al.<br />

Conclusion<br />

The genotypes ‘FHIA-01’, ‘FHIA-02’, ‘Caipira’, ‘FHIA-21’, ‘PV 42-85’ and ‘Thap maeo’<br />

presented resistance to black <strong>leaf</strong> streak disease. The genotype ‘SH 36-40’ is<br />

susceptible to black <strong>leaf</strong> streak disease. There was no significant effect <strong>of</strong> the<br />

cultivation system on the severity <strong>of</strong> black <strong>leaf</strong> streak disease.<br />

References<br />

Cavalcante M.J.B., T.M.S Gondim, Z.J.M. Cordeiro, A.P. Matos, J.L. Hessel and F.R.V. Sampaio.<br />

1999. Ocorrência da sigatoka-negra em dez municípios do Estado do Acre. Rio Branco:<br />

EMBRAPA-CPAF/AC. p.1-2. (EMBRAPA-CPAF/AC. Comunicado Técnico 107).<br />

Cordeiro Z.J.M., A.P. Matos and S. de O Silva. 1998. La Sigatoka negra en Brasil. INFOMUSA<br />

7(1):30-31.<br />

Gauhl F. 1994. Epidemiology and ecology <strong>of</strong> black Sigatoka (<strong>Mycosphaerella</strong> fijiensis Morelet)<br />

on plantain and banana (Musa spp) in Costa Rica, Central América. INIBAP, Montpellier,<br />

120pp.<br />

Pereira L.V., Z.J.M. Cordeiro, A. Figueira, R. H. Hinz and A.P. Matos. 1999. Doenças da<br />

bananeira. Informe Agropecuário, Belo Horizonte 20(196):37-47.<br />

Ritzinger C.H.S.P, R. Ritzinger, Z.J.M. Cordeiro and M.J.B. Cavalcante. 1999. Ocorrência de<br />

sigatoka negra da bananeira em Rio Branco, AC, Brasil. Fitopatologia Brasileira, v.24<br />

(Suplemento), p.450.<br />

Stover R. H. 1971. A proposed international scale for estimating intensity <strong>of</strong> banana <strong>leaf</strong> <strong>spot</strong><br />

(<strong>Mycosphaerella</strong> musicola). Tropical Agriculture 48:185-196.<br />

255


Session 4<br />

J.V. Escalant<br />

The International Musa testing<br />

programme (IMTP): a worldwide<br />

programme to evaluate elite Musa<br />

cultivars<br />

J.V. Escalant<br />

Abstract<br />

The International Musa Testing Programme (IMTP) is a collaborative effort coordinated by INIBAP<br />

to evaluate, in suitable sites worldwide, elite Musa cultivars produced by breeding programmes<br />

and promising accessions from the INIBAP collection. Established in 1989, IMTP trials are<br />

designed to be replicated anywhere in the world and aim to evaluate elite clones for resistance<br />

and/or tolerance to black <strong>leaf</strong> streak disease, Sigatoka disease and Fusarium wilt. Phase II <strong>of</strong><br />

IMTP started in 1996 when 15 countries initiated their field plots using 9 elite clones from<br />

Honduras, Taiwan, Brazil and Cuba. This presentation summarizes the results. The analysis <strong>of</strong><br />

the results indicates that disease development time is not a reliable parameter for evaluating<br />

resistance levels and that the infection index is a more reliable parameter. Among the<br />

cultivars tested, ‘FHIA-23’ and ‘SH-3436-9’ displayed a good level <strong>of</strong> resistance using ‘Pisang<br />

Ceylan’ as a resistant reference. This conclusion is consistent with the youngest <strong>leaf</strong> <strong>spot</strong>ted<br />

score obtained in most countries. A correlation was found between the infection index at bunch<br />

emergence and the average finger weight across sites and genotypes. So far, 23 research<br />

institutes in 21 countries are participating in IMTP III.<br />

Resumen - Programa Internacional de Evaluación de Musa (IMTP): un programa<br />

mundial para evaluar las variedades elite de Musa<br />

El Programa Internacional de Evaluación de Musa (IMTP) es un esfuerzo colaborativo<br />

coordinado por INIBAP cuyo fin es evaluar las variedades elite de Musa en sitios apropiados<br />

alrededor de todo el mundo. Los ensayos del IMTP son diseñados para poder replicarlos en<br />

cualquier lugar del mundo. El IMTP primero se desarrolló para realizar evaluaciones detalladas<br />

de material nuevo con el fin de obtener información sobre su resistencia o tolerancia a las<br />

Sigatokas negra y amarilla y al marchitamiento por Fusarium. La fase II del IMTP empezó en<br />

1996 cuando 15 países diferentes establecieron sus parcelas en el campo utilizando<br />

INIBAP, Montpellier, France<br />

257


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

9 diferentes cultivares elite de Honduras, Taiwán, Brasil y Cuba. En esta presentación se brinda<br />

una apreciación global de los resultados. Un análisis de los resultados obtenidos indicó que<br />

el tiempo de desarrollo de esta enfermedad no es un parámetro confiable para evaluar los<br />

niveles de resistencia. Inversamente, los índices de infección parecen ser un parámetro más<br />

confiable. Dentro del material evaluado, ‘FHIA-23’ y ‘SH-3436-9’ mostraron un buen nivel de<br />

resistencia en comparación con la referencia resistente ‘Pisang Ceylan’. Esta conclusión es<br />

consistente con el conteo de la hoja más joven manchada obtenido en la mayoría de los países.<br />

Se observó una buena correlación entre el índice de infección durante la emergencia de<br />

racimos y el peso promedio del dedo por el sitio y por el genotipo. Actualmente, 23 institutos<br />

de investigaciones en 21 países están participando en la fase III del IMTP.<br />

Résumé – Le programme international d’évaluation des Musa (IMTP) : un programme<br />

international pour évaluer les cultivars d’élite Musa<br />

Le programme international d’évaluation des Musa (IMTP) est un effort de collaboration<br />

coordonné par l’INIBAP pour évaluer, dans des sites appropriés du monde entier, des cultivars de<br />

bananiers produits par les programmes d’amélioration et des accessions de la collection de l’INIBAP.<br />

Etablis en 1989, les essais IMTP sont conçus pour être répliqués n’importe où à travers le monde<br />

et vise à évaluer les clones d’élites pour leur résistance et/ou tolérance à la maladie des raies noires,<br />

la maladie de Sigatoka et la fusariose. La phase II de l’IMTP a débuté en 1996 lorsque 15 pays ont<br />

mis en place leurs parcelles expérimentales pour 9 clones d’élites provenant du Honduras,Taïwan,<br />

Brésil et Cuba. Cette présentation résume les résultats. L’analyse des résultats indique que le temps<br />

de développement de la maladie n’est pas un paramètre fiable pour évaluer les niveaux de<br />

résistance et que l’indice d’infection est plus fiable. Parmi les cultivars testés, ‘FHIA-23’ et ‘SH-<br />

3436-9’ ont démontré un bon niveau de résistance par comparaison au témoin ‘Pisang Ceylan’.<br />

Cette conclusion est consistante avec le score obtenu pour la plus jeune feuille nécrosé obtenu<br />

dans la plupart des pays. Une corrélation a été observée entre l’indice d’infection à l’émergence<br />

du régime et le poids moyen des doigts pour tous les sites et génotypes. Jusqu’à maintenant, 23<br />

instituts de recherche dans 21 pays participent à la phase III de l’IMTP.<br />

Introduction<br />

The International Musa Testing Programme (IMTP) is a collaborative effort<br />

coordinated by INIBAP to evaluate, in suitable sites worldwide, elite Musa cultivars<br />

produced by breeding programmes and promising accessions from the INIBAP<br />

collection. Taking into account local conditions, IMTP trials are designed to be<br />

replicated anywhere in the world. The programme was developed to evaluate new<br />

germplasm in order to obtain information on their resistance/tolerance to black <strong>leaf</strong><br />

streak disease and Sigatoka disease, caused respectively by <strong>Mycosphaerella</strong> fijiensis<br />

and M. musicola, and to Fusarium wilt, caused by Fusarium oxysporum fsp. cubense.<br />

IMTP trials can also be used to conduct basic research on the pathogen and its host,<br />

such as epidemiological studies, host-pathogen relationships <strong>of</strong> the different strains<br />

<strong>of</strong> a pathogen, and adaptability and productivity studies.<br />

Two protocols have been developed in response to the demand from national<br />

programmes to evaluate germplasm under local conditions while recognizing<br />

the need for more detailed research at a limited number <strong>of</strong> sites. The two types <strong>of</strong><br />

evaluation are:<br />

1) performance evaluations which use a simplified protocol to obtain data on cultivar<br />

or hybrid performance under local conditions and basic data on disease resistance<br />

or tolerance;<br />

258


Session 4<br />

J.V. Escalant<br />

2) and in-depth evaluations which are more complete disease resistance evaluations<br />

carried out at a smaller number <strong>of</strong> sites. These sites are used to screen new<br />

improved hybrids and, if requested by breeding programmes, parental breeding<br />

lines. They can also provide opportunities for basic research on host-pathogen<br />

interactions. A standard procedure for data management and statistical analysis<br />

has been developed and the guidelines, which have been revised in the wake <strong>of</strong><br />

phase II <strong>of</strong> the programme, are made available in English, French or Spanish to<br />

the participating programmes.<br />

Phase I<br />

The programme was established in 1989 as a partnership between National<br />

Agricultural Research Systems (NARS), INIBAP breeders and pathologists from several<br />

institutes. The objective was to use multilocation trials to identify banana and plantain<br />

hybrids meeting local requirements which small-scale farmers could use to replace<br />

susceptible cultivars. A second objective <strong>of</strong> IMTP was to stimulate Musa breeding<br />

programmes by providing information on the response <strong>of</strong> their improved cultivars<br />

to pathogens. A indirect effect <strong>of</strong> IMTP has been to increase the capacity <strong>of</strong> national<br />

organisations to carry out research on banana and plantain.<br />

The programme began by evaluating germplasm from the Fundación Hondureña<br />

de Investigación Agrícola (FHIA) for resistance to black <strong>leaf</strong> streak disease. Seven<br />

tetraploid hybrids from a wide variety <strong>of</strong> genetic backgrounds were tested along with<br />

several reference diploid clones (wild and edible) that represented the whole range<br />

<strong>of</strong> reactions to black <strong>leaf</strong> streak disease, i.e. from highly resistant to highly<br />

susceptible. The trials were conducted in six countries. Site managers were trained,<br />

and provided with technical guidelines and funding to carry out the trials. Four years<br />

later, the detailed results were published and three hybrids were recommended for<br />

distribution: ‘FHIA-01’ and ‘FHIA-02’, two dessert banana cultivars with outstanding<br />

performance and high resistance to black <strong>leaf</strong> streak disease, and ‘FHIA-03’, a cooking<br />

banana also with excellent performance and resistance to black <strong>leaf</strong> streak disease.<br />

Over the last ten years these clones have been distributed to more than 50 countries.<br />

In view <strong>of</strong> the success <strong>of</strong> the programme, INIBAP was asked to develop the programme<br />

further.<br />

Phase II<br />

In IMTP II, germplasm was assessed for resistance to M. fijiensis and M. musicola.<br />

Four breeding programmes contributed germplasm (Table 1) and the number <strong>of</strong> test<br />

sites was increased to 37, even though the trials were financed by the participating<br />

institutes. The majority <strong>of</strong> IMTP II trials were set up in 1996-1997. The complete<br />

report includes results on resistance to black <strong>leaf</strong> streak disease from sites in<br />

Cameroon, Costa Rica, Honduras, Nigeria, the Philippines,Tonga and Uganda and to<br />

Sigatoka disease from one site in Colombia. A complete analysis was not possible<br />

because <strong>of</strong> missing data, for reasons that include natural catastrophes, in places like<br />

Cameroon, Costa Rica, Tonga, Thailand, Cuba and India. This presentation summarises<br />

the IMTP II results.<br />

259


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Table 1. Improved cultivars included in the resistance trials <strong>of</strong> IMTP II.<br />

Clone Origin 1 Genome<br />

PV-03.44 EMBRAPA AAAB<br />

PA-03.22 EMBRAPA AAAB<br />

SH-3436-9 INIVIT AAAA<br />

FHIA-01 FHIA AAAB<br />

FHIA-03 FHIA AABB<br />

FHIA-17 FHIA AAAA<br />

FHIA-23 FHIA AAAA<br />

GCTCV-119 TBRI AAA<br />

GCTCV-215 TBRI AAA<br />

1<br />

Breeding programme<br />

EMBRAPA: Empresa Brasileira de Pesquisa Agropecuaria, Brazil; FHIA: Fundación Hondureña de Investigación Agrícola, Honduras;<br />

INIVIT: Instituto Nacional de Investigación en Viandas Tropicales, Cuba; TBRI: Taiwan Banana Research Institute, Taiwan.<br />

Resistance to black <strong>leaf</strong> streak disease<br />

The data presented come from the trials in Costa Rica, Cameroon (Figure 1) and Tonga<br />

which are considered as representative <strong>of</strong> Latin America, Africa and Pacific Asia<br />

respectively. The response <strong>of</strong> the clones to black <strong>leaf</strong> streak disease varied according<br />

to the biotic and abiotic factors present in each country.<br />

40 Infection Index<br />

YLS<br />

35<br />

12<br />

10<br />

30<br />

Infection Index<br />

25<br />

20<br />

15<br />

10<br />

5<br />

8<br />

6<br />

4<br />

2<br />

YLS<br />

0<br />

0<br />

Calcutta 4<br />

Yangambi Km5<br />

Pisang Ceylan<br />

FHIA 23<br />

SH3436-9<br />

Pisang Berlin<br />

Saba<br />

PA 03-22<br />

PV 02-44<br />

Local cultivar<br />

Figure 1. Infection index <strong>of</strong> black <strong>leaf</strong> streak disease and youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS) <strong>of</strong> different genotypes<br />

in Cameroon.<br />

260


Session 4<br />

J.V. Escalant<br />

‘FHIA-23’ and ‘SH-3436-9’ from Honduras, had an average infection index similar<br />

to that <strong>of</strong> the resistant reference cultivar ‘Pisang Ceylan’ suggesting that ‘FHIA-23’<br />

and ‘SH-3436-9’ are resistant to black <strong>leaf</strong> streak disease (Table 2). The conclusion<br />

is consistent with the youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS) score obtained in most countries.<br />

Table 2. Infection index <strong>of</strong> black <strong>leaf</strong> streak disease and youngest <strong>leaf</strong> <strong>spot</strong>ted (YLS) <strong>of</strong> various genotypes in Costa<br />

Rica and Tonga.<br />

Costa Rica<br />

Tonga<br />

Clone Infection Index YLS Clone Infection Index YLS<br />

PV-03.44 36.1 8.3 PV-03.44 29.9 7<br />

PA-03.22 34.4 8.2 PA-03.22 34.8 7<br />

SH-3436-9 24.6 7.4 SH-3436-9 13.5 11<br />

FHIA-23 18.4 7.8 FHIA-23 61 11<br />

References<br />

References<br />

Calcutta 4 10.6 7.9 Calcutta 4 1.8<br />

Yangambi Km5 19.1 7.9 Yangambi Km5 20 11<br />

Pisang Ceylan 18.7 8.5 Pisang Ceylan 19 10<br />

Saba 32.3 8 Saba 20.3 9<br />

Local cultivar 45.6 4.9 Local cultivar 30 5<br />

Agronomic performance<br />

Bunches <strong>of</strong> ‘FHIA-23’ and ‘SH 3436-9’ weighed on average 30.6 kg and 22.3 kg,<br />

with a maximum <strong>of</strong> 39.4 kg in Cameroon and 28.8 kg in Tonga. Although, the local<br />

cultivars differed between countries, their average bunch weight was 16.5 kg, with<br />

a maximum bunch weight <strong>of</strong> 22.8 kg in Tonga (Figure 2). This substantiates the<br />

improved performance <strong>of</strong> FHIA hybrids. However, FHIA and INIVIT hybrids had<br />

longer growth cycles than local reference cultivars with an average <strong>of</strong> 474 and 420<br />

days for ‘FHIA-23’ and ‘SH-3436-9’, respectively.<br />

Discussion<br />

Disease development time (DDT) was not a reliable measure <strong>of</strong> resistance possibly<br />

because <strong>of</strong> difficulties in interpreting <strong>leaf</strong> symptoms under certain conditions. For<br />

example, when disease pressure is high, stage 1 lesions may coalesce due to their<br />

high number and appear similar to a stage 6 necrotic lesion. The infection index<br />

seems to be a more reliable parameter. Besides being comparable between countries,<br />

the results can be used to classify the new hybrids.<br />

The same clone in different country can display a different tolerance to black<br />

<strong>leaf</strong> streak disease. Tolerance being influenced by many factors, e.g. management,<br />

soil fertility, pathogen pressure, presence <strong>of</strong> other pathogens and climatic conditions,<br />

it is not possible to generalize the results. The effect <strong>of</strong> these factors on yield is not<br />

easy to demonstrate or quantify. However, work at the Centre africain de recherches<br />

sur bananiers et plantains (CARBAP) in Cameroon has demonstrated an effect <strong>of</strong><br />

261


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

black <strong>leaf</strong> streak disease on bunch weight. A correlation between the infection index<br />

at bunch emergence and fruit weight averaged across sites and genotypes (r=-0.71)<br />

was found using IMTP data (Figure 3). The correlation suggests that the numbers <strong>of</strong><br />

characteristics to record could be reduced, thus simplifying data collection and<br />

management. It would also reduce the need for visual interpretation <strong>of</strong> symptoms<br />

in the field.<br />

40<br />

35<br />

30<br />

25<br />

Weight (Kg)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Bago Oshiro<br />

Cameroon<br />

Costa Rica Honduras Tonga<br />

Uganda<br />

Country<br />

FHIA-23 SH-3436-9 PV-03.44 PA-0.33-22 local<br />

Figure 2. Average bunch weight <strong>of</strong> various hybrids in different locations.<br />

Unexpectedly, in some sites, the highly resistant clones ‘Calcutta 4’ and ‘Yangambi<br />

km5’ had stage 6 necrotic lesions. Further investigations are needed to determine<br />

whether the effect was due to aggressive strains <strong>of</strong> M. fijiensis or to a new pathogen.<br />

Conclusion<br />

FHIA hybrids had consistently the best yields in trials. With few exceptions, their<br />

bunches were heavier than those from other improved and local cultivars. FHIA<br />

hybrids also responded well to careful management and to fertilizer. In summary,<br />

FHIA hybrids performed well under a range <strong>of</strong> conditions; the better the conditions<br />

the better their performance. The improved hybrid ‘FHIA-23’ had the best performance<br />

in all the trials to evaluate resistance to Sigatoka <strong>diseases</strong>.<br />

262


Session 4<br />

J.V. Escalant<br />

250<br />

Across clones<br />

200<br />

Finger Weight (g)<br />

150<br />

100<br />

50<br />

R 2 = 0,5064<br />

Figure 3. Correlation between the infection index at bunch emergence and fruit weight averaged across sites<br />

and genotypes.<br />

Cultivar ‘GCT CV 119’ deserves special attention as it had the lowest discoloration<br />

scores for two races <strong>of</strong> F. oxysporum f.sp. cubense and high bunch weight under<br />

conditions <strong>of</strong> good crop husbandry. It is important to stress that resistance alone is<br />

not useful. It needs to be combined with good production, and acceptable post-harvest<br />

and organoleptic traits. Improved banana varieties contribute not only to reducing<br />

disease incidence but also to improving food production. The complete data and<br />

statistical analysis will be published shortly 1 .<br />

IMTP II database<br />

All the information from IMTP II, including agronomic trains and host plant<br />

response, for all genotypes and all sites, is compiled in a database to facilitate<br />

access to the data on new Musa germplasm compiled throughout the world. The<br />

database is also included in the CD-ROM, ‘Evaluation <strong>of</strong> Musa germplasm for<br />

resistance to Sigatoka <strong>diseases</strong> and Fusarium wilt’. The CD-ROM also contains<br />

the technical guidelines, a comprehensive analysis <strong>of</strong> phase II results entitled<br />

‘Evaluating <strong>bananas</strong>: a global partnership’, a catalogue <strong>of</strong> candidate and reference<br />

clones, including those for IMTP III trials, and the transfer agreement to obtain<br />

genetic material. The CD-ROM contains all the information needed to participate<br />

in IMTP III.<br />

IMTP III<br />

0<br />

0 10 20 30 40 50 60 70<br />

Infection index<br />

In 2001, 450 consignments <strong>of</strong> germplasm accessions were sent from the INIBAP<br />

Transit Centre (ITC) to 23 institutes in 21 countries participating in IMTP III. Thirtyfive<br />

cultivars including plantains, cooking <strong>bananas</strong> and dessert <strong>bananas</strong> are available<br />

1<br />

To receive the final report please contact Jean-Vincent Escalant, the IMTP coordinator at INIBAP.<br />

263


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

for evaluation. Eleven institutes are carrying out “in-depth” studies to evaluate<br />

improved varieties, which will involve epidemiological and ecological research.<br />

Thirteen institutes will be evaluating the performance <strong>of</strong> varieties against local<br />

<strong>diseases</strong> and conditions following a more simplified format <strong>of</strong> data gathering (Table<br />

3). For the first time two private companies will be carrying out evaluations. The<br />

first results are expected in 2003.<br />

Table 3. Institutes and countries involved in IMTP III<br />

Country Institute Phase III<br />

Australia Queensland Horticultural Institute (QHI) Not confirmed<br />

Bangladesh Bangladesh Agricultural Research Institute (BARI) Performance<br />

Burundi Institut de recherches agronomique et zootechnique de la<br />

communauté économique des pays des grands lacs (IRAZ)<br />

Performance<br />

Cameroon Centre africain de recherches sur bananiers et plantains (CARBAP) In-depth<br />

China South China Agricultural University (SCAU) Performance<br />

Colombia Corporación Colombiana de Investigación Agropecuaria (CORPOICA) In-depth<br />

Costa Rica Corporación Bananera Nacional (CORBANA) In-depth<br />

Dominican Centro para el Desarrollo Agropecuario y Forestal (CEDAF) Performance<br />

Republic<br />

Haiti Institut Interaméricain de Coopération pour l’agriculture (IICA) Performance<br />

Honduras Fundación Hondureña de Investigación Agrícola (FHIA) Performance<br />

India National Research Center on Banana (NRCB) In-depth<br />

Indonesia Central Research Institute for Horticulture (CRIH) Performance<br />

Malaysia Malaysian Agricultural Research and Development Institute (MARDI) In-depth<br />

Mexico Instituto Nacional de Investigaciones Forestales y Agropecuarias (INIFAP) In-depth<br />

Nicaragua Universidad de León (UNAN León) Performance<br />

Peru Servicio Nacional de Sanidad Agraria (SENASA) Performance<br />

Philippines Bureau <strong>of</strong> Plant Industry – Davao National Crop Research In-depth<br />

and Development Center (BPI – DNCRDC)<br />

Philippines Dole Asia Research ; Stanfilco In-depth<br />

Philippines Lapanday Fruit Company In-depth<br />

Rwanda Institut des sciences agronomiques du Rwanda (ISAR) Performance<br />

Sri Lanka Agricultural Research Station (ARS) Performance<br />

Uganda National Agricultural Research Organization (NARO) In-depth/<br />

Performance<br />

Vietnam Vietnam Agricultural Science Institute (VASI) Performance<br />

All except two institutes asked to evaluate <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> demonstrating the<br />

worldwide impact <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> on banana production. Black <strong>leaf</strong> streak disease<br />

is the main <strong>leaf</strong> <strong>spot</strong> <strong>of</strong> banana in the world. Leaf <strong>spot</strong>s are also caused by M. musicola<br />

and M. eumusae, the latter recently discovered in Southeast Asia and easily confused<br />

with M. fijiensis. In the framework <strong>of</strong> IMTP, a training course for IMTP III<br />

participants was organized by INIBAP and CIRAD, and hosted by the Malaysian<br />

Agricultural Research and Development Institute (MARDI). The aims were to<br />

standardize methods <strong>of</strong> data collection and evaluation <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> and to<br />

264


Session 4<br />

J.V. Escalant<br />

train researchers to identify <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> by morphological differences in the<br />

anamorph stages <strong>of</strong> the fungi.<br />

Data entry forms<br />

265


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Recommendations <strong>of</strong> session 4<br />

Progress has been made towards the creation <strong>of</strong> new varieties resistant to black <strong>leaf</strong> streak<br />

disease, either through conventional and/or modern technologies. New tetraploid hybrids<br />

resistant to black <strong>leaf</strong> streak disease are already available and some <strong>of</strong> these are widely grown<br />

around the world. Good progress has also been made in the development <strong>of</strong> a molecular<br />

toolbox for <strong>bananas</strong> and plantains in the area <strong>of</strong> the genetic transformation.<br />

Musa balbisiana genome<br />

The presence <strong>of</strong> the activable form <strong>of</strong> the banana streak virus (BSV) in interspecific hybrids<br />

(AxB), hinders the production <strong>of</strong> a new generation <strong>of</strong> triploid hybrids. Access to all the<br />

balbisiana diversity is important because <strong>of</strong> the BSV related problem but also to get a better<br />

knowledge <strong>of</strong> the existing diversity in the B genome and its contribution in the interspecific<br />

hybrids.<br />

It is recommended to study the diversity <strong>of</strong> the Musa balbisiana genome with both<br />

morphological and molecular traits. It is also recommended to promote and facilitate<br />

new collecting missions.<br />

Breeding for resistance<br />

All possible sources <strong>of</strong> resistance to pests and <strong>diseases</strong> are needed to genetically improve<br />

Musa cultivars. Previous studies report on resistance to <strong>Mycosphaerella</strong> spp. <strong>of</strong> Musa<br />

schizocarpa (genome S) and Musa textilis (genome T). Cultivars containing T and S genomes<br />

have also been reported in Papua New Guinea as being highly resistant. Prospecting for new<br />

sources <strong>of</strong> resistance to <strong>Mycosphaerella</strong> spp., as well as other <strong>diseases</strong>, should be easier since<br />

all Musa species, except Musa textilis, are covered by the facilitated access provision in the<br />

recently signed International Treaty on Plant Genetic Resources for Food and Agriculture<br />

Treaty.<br />

It is recommended to anticipate the needs <strong>of</strong> genetic improvement programmes by<br />

screening the T and S genome species as well as subspecies <strong>of</strong> Musa acuminata and<br />

other Musa spp. to detect new possible sources <strong>of</strong> resistance to pests and <strong>diseases</strong>.<br />

Durability <strong>of</strong> resistance<br />

The erosion <strong>of</strong> resistance is a problem which should be addressed to ensure the durability <strong>of</strong><br />

resistant improved cultivars.<br />

The pathogen populations should be characterized in areas where the resistance <strong>of</strong> hybrids<br />

appears to be decreasing.<br />

266


Session 4<br />

Recommendations<br />

Computer modelling <strong>of</strong> black <strong>leaf</strong> streak disease epidemics is a way to predict the durability<br />

<strong>of</strong> resistance and to evaluate different disease management strategies. However, the<br />

development <strong>of</strong> such a model requires more quantitative parameters to describe disease<br />

epidemics and the evolution <strong>of</strong> the pathogen population in response to the selection pressure<br />

exerted by resistant hosts.<br />

It is recommended to study the structure <strong>of</strong> the population at field and regional level<br />

in mixed-cultivars plots combining vertical and horizontal resistances.<br />

Mutation induction<br />

Mutation induction techniques should no longer be seen as an independent genetic<br />

improvement strategy but more as a tool that can contribute to cross-breeding programmes<br />

by increasing genetic diversity in parental lines. For example, the barley MLO gene that confers<br />

complete resistance to powdery mildew was obtained by mutagenesis. Mutants could also<br />

help in understanding the mechanism <strong>of</strong> resistance. Induced deletion mutants and aneuploids<br />

in particular will be useful to map or locate genes <strong>of</strong> interest and molecular markers.<br />

Genetic transformation<br />

Triploid cultivars <strong>of</strong> banana are <strong>of</strong>ten pollen and seed sterile and as such they should benefit<br />

from simpler risk assessment protocols regarding geneflow.<br />

It is recommended to encourage the development <strong>of</strong> national legislation to allow field<br />

testing <strong>of</strong> transgenic Musa plants, to collect data, to guide further research and regulation.<br />

It is also recommended to continue the development <strong>of</strong> transgenic banana plants to allow<br />

in-depth studies on plant development and on plant-pathogen interactions and to increase<br />

resistance to <strong>Mycosphaerella</strong>.<br />

Genetic transformation is also recommended to identify and isolate genes <strong>of</strong> resistance<br />

using the ‘knockout’ strategy on resistant cultivars. This will also be very useful to study<br />

host-pathogen interaction.<br />

267


Session 5<br />

Integrated disease management


Session 5<br />

R. Peterson et al.<br />

Management <strong>of</strong> <strong>Mycosphaerella</strong><br />

<strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> in Australia<br />

R. Peterson 1 , K. Grice 1 and S. De La Rue 1<br />

Abstract<br />

<strong>Mycosphaerella</strong> musicola is the major <strong>leaf</strong> <strong>spot</strong> pathogen affecting banana in the tropics, whereas<br />

<strong>Mycosphaerella</strong> musae and M. musicola dominate in the sub-tropical areas <strong>of</strong> Australia. Control<br />

strategies include chemical and cultural practices. In tropical areas, sprays are applied at intervals<br />

<strong>of</strong> 10-14 days during the wet season and 21-28 days during the dry season. Annually, 18-24 protectant<br />

and systemic fungicide sprays are applied with petroleum oil (5L). Since the early 1980s, the banana<br />

industry has fought to exclude this disease from the production areas by monitoring and creating<br />

a barrier <strong>of</strong> resistant plant material. Black <strong>leaf</strong> streak disease has been detected in the Cape York<br />

area eight times in the past 20 years and was limited to a few plants and in all cases was successfully<br />

eradicated.<br />

In April 2001, black <strong>leaf</strong> streak disease was identified in the main production area <strong>of</strong> North<br />

Queensland. An intensive survey (April-June) indicated the disease was restricted to the Tully area.<br />

Black <strong>leaf</strong> streak disease was found on 14 commercial properties and on 11 clumps <strong>of</strong> unmanaged<br />

plants. An eradication programme commenced in September 2001 across approximately 4500 ha<br />

<strong>of</strong> banana plants and involved de<strong>leaf</strong>ing commercial plantings to zero disease, weekly fungicide<br />

applications for six months and the destruction <strong>of</strong> all non-managed plants. Black <strong>leaf</strong> streak disease<br />

has not been detected on commercial plantations for eight months and on non-managed plants<br />

for four months. It was not detected in 1550 samples assessed during January to April 2002.<br />

Manejo de las <strong>Mycosphaerella</strong> en Australia<br />

La Sigatoka amarilla (causada por <strong>Mycosphaerella</strong> musicola) es la principal enfermedad foliar que<br />

afecta a los bananos en los trópicos, mientras que la mancha foliar (<strong>Mycosphaerella</strong> musae) y la<br />

Sigatoka amarilla predominan en las áreas subtropicales de Australia. La estrategias de control<br />

incluyen la aplicación de los químicos y las prácticas culturales. En las áreas tropicales, los rociados<br />

se aplican a intervalos de 10-14 días durante la estación húmeda y de 21-28 días durante la estación<br />

seca. Anualmente, se aplican 18-24 rociados de fungicidas protectores y sistémicos mezclados con<br />

el aceite mineral (5L). Desde inicios de la década de los 80, la industria bananera ha luchado para<br />

erradicar esta enfermedad en las áreas de producción monitoreando y creando un barrera de material<br />

vegetal resistente. La Sigatoka negra fue detectada en el área de Cabo York ocho veces durante los<br />

últimos 20 años y fue limitada a unas pocas plantas y todos los casos fueron erradicados<br />

exitosamente.<br />

1<br />

Queensland Department <strong>of</strong> Primary Industries, Mareeba, Australia.<br />

271


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

En abril de 2001, la Sigatoka negra fue identificada en la principal área de producción de Queensland<br />

del Norte. Una encuesta intensiva (abril-junio) indicó que la enfermedad estaba restringida al área<br />

de Tully. La Sigatoka negra fue descubierta en 14 propiedades comerciales y en 11 grupos de plantas<br />

sin manejar. En septiembre de 2001 empezó un programa de erradicación en aproximadamente<br />

4500ha de bananos e incluyó el deshoje total de las siembras comerciales, aplicaciones semanales<br />

de funguicidas durante seis meses y la destrucción de todas las plantas sin manejo. La Sigatoka<br />

negra no fue detectada en las plantaciones comerciales durante ocho meses y en las plantas sin<br />

manejo durante cuatro meses. Tampoco se detectó la Sigatoka negra en 1550 muestras evaluadas<br />

durante enero-abril de 2002.<br />

Résumé - Gestion des <strong>Mycosphaerella</strong> en Australie<br />

<strong>Mycosphaerella</strong> musicola est le principal agent pathogène des maladies foliaires dans les tropiques,<br />

alors que <strong>Mycosphaerella</strong> musae et <strong>Mycosphaerella</strong> musicola dominent dans les régions subtropicales<br />

d’Australie. Les stratégies de contrôle incluent des pratiques chimiques et culturales. Dans<br />

les zones tropicales, les pulvérisations sont appliquées à des intervalles de 10-14 jours pendant la<br />

saison humide et de 21-28 jours pendant la saison sèche. Sur une année, 18-24 pulvérisations de<br />

fongicides protecteurs et systémiques sont appliquées avec de l’huile de pétrole (5L). Depuis le début<br />

des années 1980, l’industrie bananière a lutté pour exclure cette maladie des zones de production<br />

par une surveillance et la création d’une zone tampon de matériel résistant. La maladie des raies<br />

noires a été détectée huit fois au cours des 20 dernières années dans la zone de Cape York ; elle<br />

était limitée à quelques plantes et, chaque fois, elle a été éradiquée avec succès.<br />

En avril 2001, la maladie des raies noires a été identifiée dans la principale zone de production du<br />

North Queensland. Un inventaire intensif (avril-juin) a indiqué que la maladie était restreinte à la<br />

région de Tully. La maladie des raies noires a été trouvée dans 14 plantations commerciales et dans<br />

11 groupes de plantes n’ayant subi aucun traitement. Un programme d’éradication a commencé en<br />

septembre 2001 sur environ 4500 ha de bananeraies ; il comprenait la suppression des feuilles dans<br />

les plantations commerciales jusqu’au niveau zéro de la maladie, des applications hebdomadaires<br />

de fongicides pendant 6 mois et la destruction de toutes les plantes non traitées. La maladie des<br />

raies noires n’a pas été détectée dans les plantations commerciales pendant huit mois et pendant<br />

quatre mois sur les plantes non traitées. Elle n’a pas été détectée dans les 1550 échantillons analysés<br />

entre janvier et avril 2002.<br />

Introduction<br />

<strong>Mycosphaerella</strong> musicola, the cause <strong>of</strong> Sigatoka disease, was first recorded in Australia<br />

in 1924 (Benson, 1925) and is still the predominant <strong>leaf</strong> disease <strong>of</strong> banana in tropical<br />

Australia. In the sub-tropical areas <strong>of</strong> south Queensland and New South Wales, the<br />

major <strong>leaf</strong> disease is <strong>Mycosphaerella</strong> speckle (caused by <strong>Mycosphaerella</strong> musae) followed<br />

by Sigatoka disease. In the Pacific region Sigatoka disease was reported from the<br />

Sigatoka Valley in Fiji in 1912 (Philpott and Knowles, 1913). Black <strong>leaf</strong> streak disease,<br />

caused by <strong>Mycosphaerella</strong> fijiensis, was recorded later, in 1963 (Rhodes, 1964), but<br />

has since become the dominant <strong>leaf</strong> disease in the region.<br />

The integrated approach used to control <strong>Mycosphaerella</strong> pathogens involves the<br />

use <strong>of</strong> cultural practices and fungicides. In wet tropical areas, fungicide sprays are<br />

applied at 10-14 day intervals during the wet season (December–April) and 14-28 days<br />

during the remainder <strong>of</strong> the year, for a total <strong>of</strong> 18-24 applications per year. Mancozeb<br />

is the main protectant fungicide used, and the application rates and length <strong>of</strong><br />

withholding periods vary depending on the formulation used. Systemic fungicides are<br />

used predominately during the wet season when conditions are more conducive to<br />

272


Session 5<br />

R. Peterson et al.<br />

disease development. Current registrations include propiconazole (Tilt®, Bumper® and<br />

Aurora®), tebuconazole (Folicur®) and benomyl (Benlate®). Cultural practices used to<br />

minimize the risk <strong>of</strong> disease development include good drainage and air circulation,<br />

location (not adjacent to a permanent body <strong>of</strong> water, buildings or rainforest) and most<br />

importantly inoculum reduction (de<strong>leaf</strong>ing).<br />

Loss <strong>of</strong> sensitivity to fungicides<br />

Sensitivity <strong>of</strong> M. musicola populations to systemic fungicides was assessed using a<br />

technique based on the Fungicide Resistance Action Committee (FRAC) guidelines, but<br />

using conidia instead <strong>of</strong> ascospores. Conidia are produced at an earlier stage <strong>of</strong> disease<br />

development than ascospores and are readily available during the winter and spring<br />

months. Ascospores are produced in summer and autumn and are not readily available<br />

during the remainder <strong>of</strong> the year due to phytosanitary regulations and de<strong>leaf</strong>ing practices.<br />

For the triazole fungicides (propiconazole and tebuconazole) germtube growth after<br />

72 hours was expressed on a sensitivity graph (germtube growth at 4 fungicide<br />

concentrations in relation to growth in the absence <strong>of</strong> the fungicide). For benomyl,<br />

percent germination was recorded after 48 hours. Sensitivity graphs were used to establish<br />

the effective concentration required to give a 50% reduction in germtube growth (EC50).<br />

Baseline data was established for each fungicide by averaging the EC50 <strong>of</strong> more<br />

than 20 populations <strong>of</strong> M. musicola. These were selected from unsprayed plants at least<br />

25 km away from the nearest commercial/sprayed block <strong>of</strong> <strong>bananas</strong>. The sensitivity <strong>of</strong><br />

populations from fungicide-treated plots was assessed by comparing the graph and the<br />

EC50 with the baseline average. An 8 to 16-fold increase in the EC50 is considered a<br />

moderate shift while a more than 16-fold increase is a severe shift in sensitivity.<br />

A shift in sensitivity was first detected with propiconazole in 1995. In limited surveys<br />

over the last 2 to 3 years, moderate to severe shifts in sensitivity to triazole fungicides<br />

were detected:<br />

• Shifts in sensitivity to tebuconazole were detected in populations that had not been<br />

sprayed with tebuconazole, but with propiconazole;<br />

• Cross resistance from propiconazole to tebuconazole is 100%, but the reverse is<br />

variable;<br />

• Shifts in sensitivity to triazoles occurred in isolated plantations and was linked to<br />

excessive use (8-12 applications), the number <strong>of</strong> consecutive applications (3-5) and<br />

applications to heavily diseased plants (no de<strong>leaf</strong>ing);<br />

• Shifts in sensitivity were reversed when the product or similar products were withheld<br />

for 6-12 months;<br />

• Resistance to benomyl was high across most populations.<br />

Strategies to minimise the risk <strong>of</strong> resistance to fungicides based on FRAC guidelines<br />

were developed and implemented in 1996 in consultation with growers and government<br />

representatives. Some <strong>of</strong> the cultural and chemical related strategies included:<br />

• Regular de<strong>leaf</strong>ing to remove heavily infested <strong>leaf</strong> material prior to the application<br />

<strong>of</strong> a systemic fungicide;<br />

273


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

•A maximum <strong>of</strong> 6 triazole fungicides to be used in any one season;<br />

•A maximum <strong>of</strong> 2 consecutive applications <strong>of</strong> triazoles;<br />

•A 4-month triazole-free period from July to October.<br />

Black <strong>leaf</strong> streak disease<br />

Black <strong>leaf</strong> streak disease was first identified in Australia in 1981 at Bamaga on the tip<br />

<strong>of</strong> the Cape York Peninsula and throughout the Torres Strait area. It has since been<br />

found at 7 other locations in the sparsely populated Cape York area. Areas affected<br />

have ranged from only a few plants to an entire residential area (Weipa, population<br />

3000), and to a 36-ha commercial plantation <strong>of</strong> organically grown <strong>bananas</strong>. In all cases<br />

the outbreaks were eradicated by destroying all plant tissue at the site <strong>of</strong> the infection<br />

and within a surrounding buffer area. Plants were destroyed by burying, ploughing<br />

or burning. In most cases the destroyed plants were replaced with resistant/tolerant<br />

banana cultivars. Regular intense surveys have failed to detect black <strong>leaf</strong> streak disease<br />

at any <strong>of</strong> these sites except at Bamaga. Bamaga was the first attempt at eradication<br />

and the program has been refined considerably over the past 15-20 years.<br />

In April 2001, M. fijiensis was detected in the Tully area where approximately 55%<br />

<strong>of</strong> northern Queensland <strong>bananas</strong> are produced. An intensive survey indicated the disease<br />

was restricted to the Tully Valley area and was only a recent introduction (6-12 months).<br />

The identification <strong>of</strong> the organism was initially complicated by the lack <strong>of</strong> conidia<br />

and sporodochia due to heavy rainfall in the area. Drying and wetting <strong>of</strong> <strong>leaf</strong> material<br />

failed to produce the sporulating structures required for microscopic identification. All<br />

samples were therefore visually assessed and suspicious samples were analysed by using<br />

the PCR (Polymerase Chain Reaction) method. Dr Juliane Henderson from the<br />

Cooperative Research Centre for Tropical Plant Protection in Brisbane modified and<br />

refined the PCR protocol based on methods published by the Natural Resource Institute<br />

(Johanson, 1997). In the 12 months to 31 st March 2002, more than 8600 banana <strong>leaf</strong><br />

samples have been assessed and 2979 samples have been PCR tested. Black <strong>leaf</strong> streak<br />

disease has been positively identified on only 13 commercial properties and on 12<br />

clumps <strong>of</strong> unmanaged banana plants.<br />

An eradication programme was started in September 2001 and was based on:<br />

• M. fijiensis having no dormant structures;<br />

• M. fijiensis only affects Musa species. There are no alternative hosts;<br />

• M. fijiensis survives in <strong>leaf</strong> tissue:<br />

– >20 weeks in <strong>leaf</strong> in canopy (Gauhl, 1994);<br />

– 4-8 weeks in <strong>leaf</strong> tissue in contact with the ground (Peterson et al., 1998);<br />

• De<strong>leaf</strong>ing and placing leaves in piles reduces by about 80% the potential <strong>of</strong> inoculum<br />

production.<br />

Programme consisted <strong>of</strong> three components:<br />

• Maintaining all commercial banana plantations (4500 ha) at zero visible disease levels<br />

for 6-8 months to ensure all inoculum is destroyed;<br />

274


Session 5<br />

R. Peterson et al.<br />

•Weekly spray programme (systemic and protectant) for at least 6 months to prevent<br />

any ascospore release from plant remains establishing new infections;<br />

• Eradication/destruction <strong>of</strong> all non-managed plants (not sprayed or de<strong>leaf</strong>ed) in the<br />

entire area.<br />

Results<br />

In commercial plantations:<br />

•All blocks were monitored (every 2 nd row) at 4 to 6-week intervals;<br />

• >95% <strong>of</strong> the area have had zero visible disease since December 2001;<br />

• Over 8600 samples have been collected and examined:<br />

– No black <strong>leaf</strong> streak disease has been recorded since August 2001.<br />

In un-managed plants:<br />

•> 28 000 stems and > 22 000 suckers have been destroyed;<br />

• No black <strong>leaf</strong> streak disease has been recorded since November 2001;<br />

•All land plots (8500) have been visited and revisited between January to March<br />

2002, and 6382 stems and 6610 suckers have been destroyed;<br />

•No black <strong>leaf</strong> streak disease has been found in the 6382 stems.<br />

Spray programme:<br />

• 13 systemic and 14 protectant fungicide sprays have been applied from August<br />

2001 to February 2002;<br />

•Oil caused damage when applied in hot dry conditions to plants that were waterstressed;<br />

•Trifloxystrobin (temporary registration) caused considerable damage when applied<br />

in hot conditions. Damage was reduced when applied early morning or late<br />

afternoon.<br />

The success <strong>of</strong> this programme can be attributed to the full participation <strong>of</strong> growers<br />

in combination with the unseasonable dry weather conditions experienced between<br />

August 2001 and April 2002. Conditions during this period were not conducive to<br />

the development <strong>of</strong> Sigatoka <strong>diseases</strong>.<br />

Future plan – establish disease-free areas<br />

Sentinel plant network:<br />

• >130 sites <strong>of</strong> 8-10 plants have been established throughout the area on a 1-km<br />

grid near sites where black <strong>leaf</strong> streak disease has been confirmed. Spacing was<br />

increased to a 10-km grid when >15 km from a known black <strong>leaf</strong> streak disease<br />

site. All plants are inspected and sampled for disease identification at 4-week<br />

intervals.<br />

275


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Commercial plantations:<br />

• Monitoring <strong>of</strong> commercial plantations is to be conducted at 6 to 8-week intervals<br />

and all diseased leaves are to be sampled;<br />

•Sites where un-managed plants have been destroyed are to be inspected for<br />

regrowth. All diseased tissue is to be collected for identification;<br />

•A period <strong>of</strong> 12 months, including an average wet season, without an outbreak<br />

should demonstrate an area free <strong>of</strong> black <strong>leaf</strong> streak disease.<br />

References<br />

Benson A.H. 1925. Leaf <strong>spot</strong> disease <strong>of</strong> <strong>bananas</strong>. Qld Agric. J. 24:392-393.<br />

Gauhl F. 1994. Epidemiology and ecology <strong>of</strong> black Sigatoka (<strong>Mycosphaerella</strong> fijiensis Morelet)<br />

on plantain and banana (Musa spp.) in Costa Rica, Central America. INIBAP, Montpellier,<br />

France, 120pp.<br />

Johanson A. 1997. Detection <strong>of</strong> Sigatoka <strong>leaf</strong> <strong>spot</strong> pathogens <strong>of</strong> banana by the Polymerase chain<br />

reaction. Natural Resource Institute, Chatham, UK.<br />

Peterson R.A., K.R.E. Grice and A. Wunsch (eds). 1998. Survival <strong>of</strong> M. musicola in <strong>leaf</strong> tissue.<br />

Report, Department <strong>of</strong> Primary Industries, Mareeba, Australia.<br />

Philpot J. and C.H. Knowles. 1913. Report on a visit to Sigatoka. Pamphlet Dep. Agric. Fiji 3.<br />

Rhodes P.L. 1964. A new banana disease in Fiji. Commonw. Phytopath. News 10:38-41.<br />

276


Session 5<br />

P.E. Jorge and T. Polanco<br />

Spread and management<br />

<strong>of</strong> black <strong>leaf</strong> streak disease<br />

in the Dominican Republic<br />

P.E. Jorge and T. Polanco<br />

Abstract<br />

In the Dominican Republic, black <strong>leaf</strong> streak disease, caused by <strong>Mycosphaerella</strong> fijiensis,was<br />

first identified in 1996 in the province <strong>of</strong> Montecristi, on the Northwest side <strong>of</strong> the country,<br />

a region with prevalent dry conditions. It was then identified in Hato Mayor on the<br />

southeast side on July, 1998. The disease appeared in 1999 in the Provinces <strong>of</strong> Sánchez<br />

Ramírez, Samaná, Dajabón, Santiago Rodríguez and Monte Plata, located on the Northwest<br />

and East sides, suggesting spread to these areas from the first two identified regions. In<br />

the year 2000, it was first identified in the Southwest, in the Provinces <strong>of</strong> Azua and San Juan<br />

de la Maguana. Today, black <strong>leaf</strong> streak disease continues to spread, moving to the central<br />

portion <strong>of</strong> the country, the largest plantain growing area, where favorable environmental<br />

conditions are common.<br />

Previous to the appearance <strong>of</strong> black <strong>leaf</strong> streak disease, measures regarding management<br />

<strong>of</strong> Sigatoka disease were not very intensive. After the appearance <strong>of</strong> black <strong>leaf</strong> streak disease<br />

in 1996 growers have changed their practices to compensate for the presence <strong>of</strong> the disease.<br />

Currently, the management methods adopted by growers include de<strong>leaf</strong>ing, application <strong>of</strong><br />

fungicides, fertilization and, to a lesser extent, planting tolerant-resistant materials, mostly<br />

FHIA hybrids. De<strong>leaf</strong>ing and application <strong>of</strong> fungicides are done on a routine basis, determined<br />

by climatic conditions. De<strong>leaf</strong>ing is usually done weekly during the rainy seasons and monthly<br />

otherwise. A similar pattern is observed for the application <strong>of</strong> fungicides. Fertilization is<br />

implemented to assure a prompt recovery from the stress induced by the disease. Few<br />

growers rely on climatic/biological disease forecasting strategies.<br />

Resumen - Propagación y manejo de la Sigatoka negra en República Dominicana<br />

En República Dominicana, la Sigatoka negra, causada por <strong>Mycosphaerella</strong> fijiensis, fue<br />

identificada por primera vez en 1996 en la provincia de Montecristi, en el noroeste del país,<br />

una región con condiciones secas prevalecientes. Se identificó en Hato Mayor en la parte<br />

sudeste en julio de 1998. En 1999, la enfermedad apareció en las provincias de Sánchez<br />

Instituto Dominicano de Investigaciones Agropecuarias y Forestales, Santo Domingo, Dominican Republic<br />

277


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Ramírez, Samaná, Dajabón, Santiago Rodríguez y Monte Plata, localizadas en el noroeste y<br />

oriente, sugiriendo su propagación a estas áreas de las dos regiones identificadas<br />

inicialmente. En el año 2000, la enfermedad fue identificada por primera vez en el sudoeste,<br />

en las provincias de Azua y San Juan de la Maguana. Actualmente, la Sigatoka negra continua<br />

propagándose, moviéndose a la parte central del país, la principal área bajo el cultivo de<br />

plátano, donde existen condiciones ambientales favorables.<br />

Previas a la aparición de la Sigatoka negra, las medidas con respecto al manejo de la Sigatoka<br />

amarilla no fueron muy intensivas. Después de la aparición de la Sigatoka negra en 1996<br />

los productores han cambiado sus prácticas de cultivo para compensar la presencia de la<br />

enfermedad. Actualmente, los paquetes de manejo de la enfermedad adoptados por los<br />

productores consisten en el deshoje, aplicación de fungicidas, fertilización y, en un menor<br />

grado, siembra de materiales tolerantes o resistentes, principalmente los híbridos de la FHIA.<br />

El deshoje y la aplicación de funguicidas son actividades que se realizan sobre una base<br />

habitual determinada por las condiciones climatológicas. Los deshojes se realizan<br />

aproximadamente cada semana durante las estaciones lluviosas, y mensualmente durante<br />

las estaciones secas. Un comportamiento similar se observa para la aplicación de los<br />

fungicidas. La fertilización se implementa para asegurar una rápida respuesta y recuperación<br />

del estrés inducido por la enfermedad. Pocos productores confían en estrategias de preaviso<br />

biológico y climatológico de la enfermedad.<br />

Résumé - Propagation et gestion de la maladie des raies noires en République<br />

Dominicaine<br />

En République Dominicaine, la maladie des raies noires, causée par <strong>Mycosphaerella</strong> fijiensis,<br />

a été identifiée pour la première fois en 1996 dans la province de Montecristi, dans la partie<br />

nord-ouest du pays, une région où les conditions de sécheresse prévalent. Elle a été ensuite<br />

identifiée à Hato Mayor sur la côte sud-est en juillet 1998. La maladie est apparue en 1999<br />

dans les provinces de Sánchez Ramírez, Samaná, Dajabón, Santiago Rodríguez et Monte Plata,<br />

situées dans les parties nord-ouest et est, ce qui suggère une propagation vers ces zones<br />

depuis les deux premières régions identifiées. En 2000, la maladie a d’abord été identifiée<br />

dans le sud-ouest, dans les provinces d’Azua et San Juan de la Maguana. Aujourd’hui, la<br />

maladie des raies noires continue de s’étendre, en se dirigeant vers la partie centrale du pays,<br />

la plus grande zone de culture des bananes plantain, dans laquelle des conditions<br />

d’environnement favorables sont fréquemment rencontrées.<br />

Avant l’apparition de la maladie des raies noires, les mesures visant à contrôler la maladie<br />

de Sigatoka n’étaient pas très intensives. Après l’apparition de la maladie des raies noires<br />

en 1996, les planteurs ont modifié leurs pratiques pour compenser la présence de la<br />

maladie. Aujourd’hui, les méthodes de gestion adoptées par les planteurs incluent<br />

l’effeuillage, l’application de fongicides l’application d’engrais et, dans une moindre mesure,<br />

l’utilisation de plants résistants/tolérants, principalement des hybrides de la FHIA. L’effeuillage<br />

et l’application de fongicides sont réalisés de manière routinière, en fonction des conditions<br />

climatiques. L’effeuillage est effectué de façon hebdomadaire pendant la saison des pluies,<br />

et mensuellement le reste du temps. La même périodicité est utilisée pour l’application des<br />

fongicides. La fourniture d’engrais est réalisée afin d’assurer une récupération rapide du stress<br />

induit par la maladie. Seul un petit nombre de planteurs ont recours à des stratégies de<br />

prévision du développement de la maladie basées sur des facteurs climatiques/biologiques.<br />

Introduction<br />

The Dominican Republic occupies the eastern two thirds <strong>of</strong> the Caribbean island <strong>of</strong><br />

Hispaniola and has an area <strong>of</strong> 48 422 km 2 . The Republic <strong>of</strong> Haiti occupies the<br />

remainder <strong>of</strong> the island. Both countries are divided by the Cordillera Mountains which<br />

278


Session 5<br />

P.E. Jorge and T. Polanco<br />

reach 3175 meters and separate the southern and northern parts <strong>of</strong> the island. The<br />

mountains are important barriers for the natural movement <strong>of</strong> inoculum <strong>of</strong> black<br />

<strong>leaf</strong> streak disease (teleomorph: <strong>Mycosphaerella</strong> fijiensis Morelet; anamorph:<br />

Paracercospora fijiensis (Morelet) Deighton), and the movement <strong>of</strong> plant material<br />

between the two regions. Banana and plantain production is mainly in the<br />

southwestern and northern parts <strong>of</strong> the country. Thus, production <strong>of</strong> Musa species<br />

is isolated and growers from either side <strong>of</strong> the mountains do not exchange planting<br />

material and there is little commercial exchange <strong>of</strong> fruits.<br />

The country is divided into eight regions by the Ministry <strong>of</strong> Agriculture and Musa<br />

production occurs in all regions. Banana production is mainly in the Southwest and<br />

the Northwest and plantain production is mainly in the Cibao Central, including the<br />

Northeast, North and North Central regions, where a large proportion <strong>of</strong> the<br />

population lives. Plantain is the third most important vegetable crop, after rice and<br />

beans. Plantain is consumed within the country, whereas a large part <strong>of</strong> the <strong>bananas</strong>,<br />

whether grown by using conventional or organic methods, are exported.<br />

The average rainfall from 1961 to 1990 (Secretaría de Estado de Agricultura, 1998)<br />

from the different regions are shown on Table 1. Full expression <strong>of</strong> black <strong>leaf</strong> streak<br />

disease in the Dominican Republic is dependent on rainfall and humidity, as<br />

temperatures in banana and plantain areas do not limit the development <strong>of</strong> the<br />

disease. Figure 1 shows symptoms <strong>of</strong> the disease in dry and humid regions. Generally,<br />

the disease is limited to the first stages in dry regions and fully expressed in humid<br />

regions.<br />

Humid regions<br />

Dry regions<br />

Figure 1. Symptoms <strong>of</strong> black <strong>leaf</strong> streak disease in the dry and humid regions <strong>of</strong> the Dominican Republic.<br />

Table 1. Average rainfall in the Dominican Republic by region for the period between 1961 and 1990.<br />

Region<br />

Rainfall (mm/year)<br />

North central 1200-1350<br />

Northeast 1200-1350<br />

North 1200-1350<br />

East 1342-1583<br />

Central 1400-1800<br />

Northwest < 600<br />

South < 600<br />

Southwest 665<br />

Source: Secretaría de Estado de Agricultura, 1998. Anuario estadistico de planificación sectorial agropecuaria.<br />

279


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

The distribution <strong>of</strong> plantain and banana crops is shown in Table 2 (Secretaría de<br />

Estado de Agricultura, 1998). More information on farm size and number <strong>of</strong> farmers<br />

can be found in Perez (2000), Departamento de Sanidad Vegetal (2000) and<br />

Secretaría de Estado de Agricultura (2000).<br />

Table 2. Area planted with plantain and banana in the Dominican Republic.<br />

Region Plantain (ha) Banana (ha)<br />

South 12 500* 3105<br />

Southwest 5764 5722<br />

Central 3750 -<br />

East 625b -<br />

North Central 14 333 2062<br />

Northeast 3139 4280<br />

North 9000 3585<br />

Northwest 2876 6201<br />

Total 51 987 24 955<br />

Source: Secretaría de Estado de Agricultura, 1998<br />

* Personal communication, Secretaría de Estado de Agricultura.<br />

Pathogen identification<br />

The spread <strong>of</strong> black <strong>leaf</strong> streak disease in the Americas, and the appearance <strong>of</strong> the<br />

disease in Cuba in 1990 (Vidal, 1992) and in Jamaica in 1994-1995 alerted the<br />

Dominican authorities. In 1995, the Department <strong>of</strong> Plant Protection, Secretaría de<br />

Estado de Agricultura de la República Dominicana, started a project to monitor the<br />

disease, in collaboration with Haiti’s phytosanitary authorities and the Animal and<br />

Plant Health Inspection Service <strong>of</strong> the United States Department <strong>of</strong> Agriculture<br />

(APHIS/USDA).<br />

Black <strong>leaf</strong> streak disease was first recorded in 1996 in Guayubín, province <strong>of</strong><br />

Montecristi, at a farm <strong>of</strong> 125 ha planted with banana and plantain. Distribution within<br />

the farm was limited mainly to plantain. At the time, M. fijiensis was identified by<br />

the Plant Pathology Laboratory at CIRAD/AMIS, Montpellier, France. In May 1997,<br />

the pathogen was also identified, with the assistance <strong>of</strong> Mary E. Palm from the United<br />

States Department <strong>of</strong> Agriculture, as P. fijiensis based on the morphology <strong>of</strong> the<br />

conidia and conidiogenous cells in her report to the Secretaría de Estado de<br />

Agricultura. This was the first <strong>of</strong>ficial diagnosis <strong>of</strong> the causal agent <strong>of</strong> black <strong>leaf</strong><br />

streak disease in the Dominican Republic. Also in 1997, Fouré surveyed the different<br />

production areas to determine the distribution and incidence <strong>of</strong> black <strong>leaf</strong> streak<br />

disease in the country (Fouré, 1997).<br />

In 1999, T. Polanco, in collaboration with Jean Carlier and Marie Zapater from<br />

CIRAD, isolated M. fijiensis from the northwestern, northeastern and central eastern<br />

regions <strong>of</strong> the Dominican Republic (Polanco, 1999). Rivas et al. (2001) compared the<br />

isolates with those from South America, Central America and the Caribbean. The<br />

isolate from the Dominican Republic was found to be closely related to the Cuban<br />

isolate.<br />

280


Session 5<br />

P.E. Jorge and T. Polanco<br />

Spread <strong>of</strong> black <strong>leaf</strong> streak disease<br />

The spread <strong>of</strong> black <strong>leaf</strong> streak disease in the Dominican Republic up to May 2002<br />

is shown in Figure 2.<br />

Infected zones<br />

Figure 2. Spread <strong>of</strong> black <strong>leaf</strong> streak disease in the Dominican Republic up to May 2002.<br />

Northwestern region<br />

The northwest region has approximately 6250 ha <strong>of</strong> banana and 2876 ha <strong>of</strong> plantain<br />

(Table 2). Bananas are mostly grown for export to Europe, where there is a market<br />

for traditional and organic fruits.<br />

The northwestern region <strong>of</strong> the Dominican Republic where black <strong>leaf</strong> streak disease<br />

was first identified in 1996, is mostly characterized by environmental conditions<br />

that do not favour the development <strong>of</strong> the disease. The rainy-humid conditions that<br />

favour disease development prevail for only short periods, usually in May and<br />

November-December. There were less than 600 mm <strong>of</strong> rain/year for the period 1961-<br />

1990, usually distributed over 70 days per year (Secretaría de Estado de Agricultura,<br />

1998). Thus, most <strong>of</strong> the region is considered to be low risk for black <strong>leaf</strong> streak<br />

disease because <strong>of</strong> the climate.<br />

In June 1999, black <strong>leaf</strong> streak disease was identified in the Provinces <strong>of</strong> Valverde,<br />

Dajabón, Santiago Rodriguez and other communities <strong>of</strong> Montecristi (Polanco 1999).<br />

Eastern region<br />

The eastern region includes the provinces <strong>of</strong> Hato Mayor, La Altagracia, El Seibo,<br />

La Romana and Higuey. Rainfall is an average <strong>of</strong> 1342-1583 mm per year distributed<br />

281


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

over 122-137 days (Secretaría de Estado de Agricultura, 1998). Because <strong>of</strong> the<br />

favourable conditions, black <strong>leaf</strong> streak disease has caused considerable damage in<br />

the region. Many growers have changed from plantain to another crop and the<br />

plantains harvested are very small.<br />

Black <strong>leaf</strong> streak disease was identified in the eastern region in the Provinces <strong>of</strong><br />

Hato Mayor and El Seibo in the communities <strong>of</strong> Hato Mayor, Sabana de la Mar and<br />

El Valle, respectively (Figure 2). The disease was identified in 1998, i.e. two years<br />

after the original identification in the northwest and it appears that the disease had<br />

been introduced through planting material.<br />

Before the occurrence <strong>of</strong> black <strong>leaf</strong> streak disease in 1998, approximately 1875<br />

ha <strong>of</strong> plantain were grown in Hato Mayor and El Seibo but today, only 625 ha are<br />

planted with the crop because <strong>of</strong> the disease. Even though the area and the production<br />

are small, plantain is <strong>of</strong> economic and social importance for the region. Previously<br />

the crop supplied the regional market and provided food and economic support to<br />

several thousand small-scale farmers. Today, demand for plantain is satisfied by<br />

supply from other producing areas.<br />

The hybrid ‘FHIA-21’ is being introduced to this region as an alternative, but<br />

acceptance has been limited.<br />

Central region<br />

The central region includes the provinces <strong>of</strong> Monte Plata, San Cristóbal and Santo<br />

Domingo-Distrito Nacional and production <strong>of</strong> plantain is approximately 3750 ha<br />

(Table 2). Rainfall is an average <strong>of</strong> 1400-1800 mm/year (Table 1).<br />

The disease appeared in 1999 in the three provinces, and affected the communities<br />

<strong>of</strong> Bayaguana, Monte Plata, San Cristobal, Yamasá and Villa Mella. The occurrence<br />

<strong>of</strong> the disease appeared to have been caused by hurricane George in September 1998.<br />

Hurricane George moved from the southeast to the southwest and winds blew strongly<br />

into the middle <strong>of</strong> the island, suggesting that introduction was from the south region,<br />

which was already affected.<br />

Southwestern region<br />

The southwestern region includes the provinces <strong>of</strong> Azua and San Juan de la Maguana.<br />

Approximately 5722 ha <strong>of</strong> banana, mostly organic (Azua) and 5764 ha <strong>of</strong> plantain<br />

are grown in the region (Secretaría de Estado de Agricultura, 1998; Departamento<br />

de Sanidad Vegetal, 2000).<br />

Annual rainfall for the region was an average <strong>of</strong> 665 mm in 1961-1990<br />

(Table 1) and was mostly distributed over 62 days, mostly in November-December.<br />

Thus, the region is at low risk from outbreaks <strong>of</strong> black <strong>leaf</strong> streak disease (Secretaría<br />

de Estado de Agricultura, 1998).<br />

Black <strong>leaf</strong> streak disease was reported on plantain from a small locality in the<br />

province <strong>of</strong> San Juan de la Maguana in January 2000; however, production in the<br />

province is not important. The disease was identified on organic <strong>bananas</strong> in Azua,<br />

in February 2000. Spread <strong>of</strong> the disease in Azua has been limited by the weather<br />

conditions; in contrast to Sigatoka disease which is moderately prevalent.<br />

282


Session 5<br />

P.E. Jorge and T. Polanco<br />

North, north-central and northeastern regions (Cibao Central)<br />

The Cibao Central includes the north-central, northeastern and north regions, and<br />

includes the provinces <strong>of</strong> La Vega, Espaillat, Sánchez, Samaná, Puerto Plata and others.<br />

Production <strong>of</strong> plantain is 26 472 ha (Table 2) and <strong>of</strong> banana 9927 ha. Both crops are<br />

mostly produced without irrigation, and are dependent on rainfall for the water supply.<br />

These regions have an average rainfall <strong>of</strong> 1200-1350 mm/year, with 122-173 days<br />

<strong>of</strong> rain per year (Secretaría de Estado de Agricultura, 1998). These conditions suggest<br />

that black <strong>leaf</strong> streak disease has the potential to cause considerable damage to plantain.<br />

In August 1998, black <strong>leaf</strong> streak disease was found on plantain in the northeastern<br />

region in the provinces <strong>of</strong> Sánchez and Samaná but not in the north-central and<br />

northern regions. In 2001, symptoms were recognized on plantain in the community<br />

<strong>of</strong> La Isabela, province <strong>of</strong> Puerto Plata in the northern region. The disease has caused<br />

considerable damage to the point where crops have been abandoned, or de<strong>leaf</strong>ing and<br />

fungicide treatments implemented to ensure production, as was the case in La Isabela.<br />

In March 2002, black <strong>leaf</strong> streak disease was identified in the north-central region<br />

in the community Hoya del Camú, province <strong>of</strong> La Vega, and in April 2002 in the north<br />

in the community <strong>of</strong> Moca, province <strong>of</strong> Espaillat. In both instances the disease affected<br />

plantain. In Moca, symptoms were diagnosed at a farm <strong>of</strong> 1.5 ha. Information from<br />

the growers suggests that the disease was introduced through plant material from<br />

infected areas.<br />

Southern region<br />

In the provinces <strong>of</strong> Barahona and Neiba, most <strong>of</strong> the area planted with plantains is<br />

irrigated. In this region, 5807 hectares <strong>of</strong> plantain are irrigated. Rainfall in 1961-1990<br />

was on average less than 600 mm/year and was distributed over 48-69 days<br />

(Secretaría de Estado de Agricultura, 1998). As a dry region, there is little likelihood<br />

<strong>of</strong> serious damage from black <strong>leaf</strong> streak disease, unless favourable conditions were<br />

to occur for prolonged periods.<br />

Management <strong>of</strong> black <strong>leaf</strong> streak disease<br />

In general, management <strong>of</strong> black <strong>leaf</strong> streak disease in the Dominican Republic depends<br />

on the pressure exerted by the disease. The type <strong>of</strong> crop (banana or plantain), the size<br />

<strong>of</strong> the production unit, whether production is conventional or organic and the prevailing<br />

environmental conditions determine the techniques used to control black <strong>leaf</strong> streak<br />

disease. A technical package was only adopted in response to the spread <strong>of</strong> the disease<br />

within the country. Limited preventive measures had been taken when the disease was<br />

not present.<br />

Cultural practices<br />

Banana growers and, to a lesser extent, plantain growers have adopted the practice<br />

<strong>of</strong> de<strong>leaf</strong>ing. This is done every 7-10 days during the rainy season when the disease<br />

pressure is high, and every 30 days in the dry season when disease pressure is low.<br />

283


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

In the eastern region, where environmental conditions favour the disease and<br />

the pathogen is present, only growers who have adopted de<strong>leaf</strong>ing and biological<br />

monitoring <strong>of</strong> the diseased have survived. Production <strong>of</strong> plantain in the region<br />

declined from 1875 ha before the appearance <strong>of</strong> black <strong>leaf</strong> streak disease in 1998<br />

to 625 ha today.<br />

Chemical control<br />

Conventional banana production is dependent on the use <strong>of</strong> fungicides, e.g.<br />

triazoles, dithiocarbamates, benzimidazoles, mineral oils and to a lesser extent,<br />

strobilurins (Polanco, 1998). Organic banana growers use mineral and vegetable oils<br />

and organic products, e.g. citric acid and waxes. Preliminary data obtained by Polanco<br />

(personal communication) showed that there was no difference between mineral oil<br />

alone and mixtures <strong>of</strong> oil with the organic products. This work is to be repeated.<br />

Conventional and organic banana growers spray between 7 to 10 times a year,<br />

and sometimes up to 15 times, depending on rainfall. Sometimes treatments are not<br />

effective or may have been unnecessary because <strong>of</strong> faulty forecast or monitoring.<br />

Large-scale banana plantations, conventional or organic, are usually sprayed by<br />

plane in the northwestern and south-western regions. Small-scale growers use<br />

motorized high-pressure pumps. In general, plantain growers do not use fungicides.<br />

Resistant hybrids<br />

FHIA hybrids, especially ‘FHIA-21’ and to a lesser extent ‘FHIA-20’, were introduced<br />

before the identification <strong>of</strong> the disease in the country. At present, there are 600 000<br />

to 800 000 ‘FHIA-21’ plants in affected and not affected areas, representing less than<br />

one percent <strong>of</strong> the total planted area. Production is for the fresh and industrial markets.<br />

Research plots have been established to compare native plantains, ‘FHIA-21’ and<br />

‘Rulo’ for their response to the technical package needed for a successful crop <strong>of</strong> ‘FHIA-<br />

21’ and to different management practices including de<strong>leaf</strong>ing and fertilizer treatment.<br />

When black <strong>leaf</strong> streak disease is absent, cultural practices and chemical control are<br />

not implemented, unless if there is a high incidence <strong>of</strong> Sigatoka disease.<br />

As a preventive measure, the replacement or planting <strong>of</strong> new areas with the resistant<br />

clone ‘FHIA-21’ has recently been recommended and plants are available from the<br />

Ministry <strong>of</strong> Agriculture. Acceptance <strong>of</strong> ‘FHIA-21’ by growers has been limited mainly<br />

because the most important plantain production areas do not have the disease or the<br />

spread is very limited, and consumers prefer the fruit <strong>of</strong> local cultivars.<br />

Little has been done to diversify the gene pool <strong>of</strong> banana, and production is entirely<br />

with ‘Cavendish’ cultivars. This is mainly because <strong>of</strong> the requirements <strong>of</strong> the banana<br />

market and the prevalence <strong>of</strong> dry conditions in the banana production areas.<br />

References<br />

Departamento de Sanidad Vegetal. 2000. Proyecto manejo integrado de la Sigatoka negra<br />

(<strong>Mycosphaerella</strong> fijiensis Morelet) en los cultivos de musáceas de la República Dominicana.<br />

Secretaría de Estado de Agricultura, Santo Domingo, República Dominicana. 42pp.<br />

284


Session 5<br />

P.E. Jorge and T. Polanco<br />

Fouré E. 1997. La maladie des raies noires des bananiers et plantains en République<br />

Dominicaine. Distribution, incidence et méthodes de contrôle. CIRAD-FHLOR. Rapport de<br />

mission en République Dominicaine du 28 août au 5 septembre 1997.<br />

Palm M.E. 1997. Trip report to the Dominican Republic.<br />

Pérez Vicente L. 2000. Informe de la misión del consultor en manejo integrado de Sigatoka<br />

negra, Dr Luis Pérez Vicente realizada en la Secretaría de Agricultura (SEA) de República<br />

Dominicana del 28 de agosto al 29 de septiembre del 2000. FAO. 32pp.<br />

Polanco T. 1998. La Sigatoka negra del plátano y guineo: reconocimiento y manejo.<br />

Departamento de Sanidad Vegetal, Secretaría de Estado de Agricultura, Santo Domingo,<br />

Republica Dominicana. 14pp.<br />

Polanco T. 1999. Informe de estancia, Laboratorio de fitopatología CIRAD-AMIS, Montpellier,<br />

Francia. Departamento de Sanidad Vegetal, Secretaría de Estado de Agricultura, Santo<br />

Domingo, Dominican Republic. 14pp.<br />

Rivas G.G., M.F. Zapater and J. Carlier. 2001. Estructura de poblaciones de <strong>Mycosphaerella</strong><br />

fijiensis en América Latina. Congreso Internacional de Sigatoka. SERBANA. San José, Costa<br />

Rica, Abril 2001.<br />

Secretaría de Estado de Agricultura. 1998. Anuario estadístico de planificación sectorial<br />

agropecuaria. Subsecretaría de Estado de Planificación Sectorial Agropecuaria, Santo<br />

Domingo, República Dominicana. 126pp.<br />

Secretaría de Estado de Agricultura. 2000. Registro Nacional de Productores Agropecuarios,<br />

Santo Domingo, República Dominicana.<br />

Vidal A. 1992. Sigatoka negra en Cuba. En nuevos focos de plagas y enfermedades. Boletin<br />

Fitosanitario de la FAO 40:1-2.<br />

285


Session 5<br />

A.S. Riveros et al.<br />

Microbiological control <strong>of</strong> black <strong>leaf</strong><br />

streak disease<br />

A. S. Riveros, C. I. Giraldo and A. Gamboa<br />

Abstract<br />

Isolates <strong>of</strong> bacteria from the phyllosphere <strong>of</strong> tomato and banana were obtained from CATIE: Bacillus<br />

cereus, Bacillus sp., Serratia marcescens, Serratia entomophila, unidentified strains with glucanolytic<br />

and chitinolytic capacity (GS2-GS3-GC1-GBC2, SE/PO 2<br />

,White) and one isolate <strong>of</strong> <strong>Mycosphaerella</strong> fijiensis<br />

collected recently in Turrialba, Costa Rica. Crude culture filtrates <strong>of</strong> some microorganisms inhibited<br />

ascospore germination and the growth in vitro <strong>of</strong> M. fijiensis colonies. The two filtrates with the<br />

greatest effect resulted in changes to the ultrastructure <strong>of</strong> M. fijiensis hyphae, when examined under<br />

a scanning electronic microscope, in comparison with untreated tissue.<br />

Resumen - Control microbiológico de la Sigatoka negra<br />

Los aislados de la colección del CATIE obtenidos de la filosfera de tomate y hojas de banano: Bacillus<br />

cereus, Bacillus sp., Serratia marcescens, Serratia entomophila y cepas no identificadas con<br />

capacidad glucanolítica y quitinolítica (GS2-GS3-GC1-GBC2, SE/PO 2<br />

,White) y un aislado de<br />

<strong>Mycosphaerella</strong> fijiensis recolectado recientemente en Turrialba, Costa Rica, fueron utilizados para<br />

preparar filtrados de los cultivos líquidos. Los filtrados crudos de estos microorganismos se<br />

evaluaron bajo condiciones in vitro con el fin de determinar la germinación de las ascosporas y<br />

el crecimiento de las colonias de M. fijiensis (agente causal de la Sigatoka negra en banano y<br />

plátano). Los resultados muestran un efecto inhibitorio importante de algunos de estos filtrados<br />

en diferentes etapas de desarrollo de <strong>Mycosphaerella</strong>. La observación, bajo un microscopio<br />

electrónico de barrido, de las estructuras del hongo tratado con dos filtrados prometedores,<br />

muestra claras alteraciones de ultraestructura en el tejido tratado en comparación con el<br />

testigo sin tratamiento.<br />

Résumé – Lutte microbiologique contre la maladie des raies noires<br />

Des isolats de bactéries de la phyllosphère de tomates et de bananiers ont été obtenus du CATIE :<br />

Bacillus cereus, Bacillus sp., Serratia marcescens, Serratia entomophila, des souches non identifiées<br />

ayant une capacité glucanolytique et chitinolytique (GS2-GS3-GC1-GBC2, SE/PO 2<br />

,White) et un<br />

isolat de <strong>Mycosphaerella</strong> fijiensis collecté récemment à Turrialba, au Costa Rica. Les filtrats bruts<br />

de certains micro-organismes ont inhibé la germination des ascospores et la germination in vitro<br />

de colonies de M. fijiensis.Les deux filtrats qui ont eu le plus d’effet ont induit des changements<br />

de l’ultrastructure des hyphes de M. fijiensis,en microscopie électronique à balayage, par rapport<br />

aux tissus non traités.<br />

CATIE, Turrialba, Costa Rica.<br />

287


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Introduction<br />

During the seventies, agriculture was characterized by an undiscriminated use <strong>of</strong><br />

agrochemicals. The situation has changed little since then, but international and national<br />

regulations have imposed changes aimed at reducing pollution and making agriculture<br />

sustainable. There are two types <strong>of</strong> agriculture: 1) high input agriculture characterized<br />

by high productivity that is limited by environmental factors and 2) low input agriculture<br />

with production that, in addition to environmental factors, is limited by pests, <strong>diseases</strong><br />

and weeds.<br />

Changes in the market and the influence <strong>of</strong> globalization are forcing a reconsideration<br />

<strong>of</strong> research in agriculture. The preferences <strong>of</strong> consumers play an increasingly important<br />

role that affects the work <strong>of</strong> multidisciplinary teams made up <strong>of</strong> researchers, ecologists,<br />

producers and industry.<br />

The Tropical Agricultural Research and Higher Education Center (CATIE) has<br />

defined one <strong>of</strong> its research objectives as the implementation <strong>of</strong> methodologies focused<br />

on the biological control <strong>of</strong> the most common <strong>diseases</strong> and pests affecting economically<br />

important tropical crops, such as those belonging to the Musaceae family.<br />

CATIE started studies on Musa and M. fijiensis with a project financed by AID/ ROCA<br />

(USA) the first phase <strong>of</strong> which started in July 1984. Other projects which followed were<br />

financed by RENARM (USA), CIRAD (France), INIBAP, INCO-Musa, Natural Resource<br />

Institute (NRI; UK), CINVESTAV (Mexico) and FONTAGRO.<br />

These collaborative projects not only spurred research on biological control but also<br />

on the control <strong>of</strong> black <strong>leaf</strong> streak disease, a disease that was already threatening banana<br />

and plantain production. Other outcomes were the development <strong>of</strong> systems for somatic<br />

embryogenesis, cell suspensions, plant pathology, cryopreservation, genetic transformation<br />

and the genetics <strong>of</strong> M. fijiensis populations.<br />

Since then, the Plant Protection Unit at CATIE has developed integrated pest<br />

management (IPM) practices for black <strong>leaf</strong> streak disease based on the preservation <strong>of</strong><br />

the environment, reduced risks to farmers, the rural population and consumers, and<br />

the sustainability <strong>of</strong> traditional agriculture. Countries included within the CATIE mandate<br />

have a rich biodiversity which may contain materials or products, e.g. genes <strong>of</strong> wild<br />

plants or biopesticides, that might be useful in IPM programmes.<br />

Research into the biological control <strong>of</strong> black <strong>leaf</strong> streak disease at CATIE encouraged<br />

researchers involved in the AID/ROCAP-USA project, e.g. Dr Elkin Bustamante and his<br />

team who were the first to work on the project. After a careful study <strong>of</strong> the different<br />

aspects <strong>of</strong> the parasitic relationship between Musa and M. fijiensis: the biology and<br />

morphology <strong>of</strong> the pathogen, the phenology and physiology <strong>of</strong> the plant, the<br />

phyllosphere, soil (importance <strong>of</strong> rhizobacteria, endophytic fungi and organic<br />

amendments), and methods <strong>of</strong> internal and external inoculation, a research programme<br />

was constructed to better study these aspects (Figure 1).<br />

Step 1. Identification <strong>of</strong> antagonistic microorganisms<br />

The purpose was to isolate microorganisms antagonistic to M. fijiensis and to evaluate<br />

their effectiveness under greenhouse and field conditions. One hundred and twenty<br />

isolates with chitinolytic activity were obtained from plants <strong>of</strong> cv. ‘Grande naine’ coming<br />

288


Session 5<br />

A.S. Riveros, C.I. Giraldo and A. Gamboa<br />

Identification<br />

<strong>of</strong> antagonistic<br />

microorganisms<br />

Effect <strong>of</strong> substrate<br />

on antagonistic<br />

microorganisms<br />

External vs internal<br />

application<br />

<strong>of</strong> antagonistic<br />

microorganisms<br />

1 2<br />

3<br />

Induced resistance “ISR”,<br />

substrates and promotion<br />

<strong>of</strong> plant growth<br />

Induced resistance and<br />

promotion <strong>of</strong> plant growth<br />

5 4<br />

Figure 1. Steps involved in research on biological control at CATIE (1986-2001).<br />

from two different locations: an area <strong>of</strong> high incidence <strong>of</strong> black <strong>leaf</strong> streak disease and<br />

an area <strong>of</strong> low incidence <strong>of</strong> the disease. The ‘low incidence site’ provided the highest<br />

population <strong>of</strong> microorganisms, which was evaluated on chitin agar. Thirteen bacterial<br />

strains were selected on the basis <strong>of</strong> their chitinolytic activity (Serratia marcescens,<br />

Serratia entomophyla and Bacillus spp.). Under greenhouse and field conditions the level<br />

<strong>of</strong> control <strong>of</strong> M. fijiensis was 40% in comparison with a level <strong>of</strong> 60% using fungicides<br />

(González, 1994; González et al. 1996).<br />

Regarding glucanolytic activity, 196 strains <strong>of</strong> bacteria were collected from plants <strong>of</strong><br />

cv. ‘Grande naine’ <strong>of</strong> which 37 belonged to the genus Bacillus. The microorganisms were<br />

purified and evaluated on glucan-agar and glucan nutrient-agar media. Seven strains<br />

with glucanolytic activity were selected. GS2, GBC2, BS3 and BC1 showed antagonistic<br />

effects, inhibiting germination <strong>of</strong> M. fijiensis ascospores in 25% <strong>of</strong> the cases and reducing<br />

germination tube length in 47%. Four strains were tested in the presence and absence<br />

<strong>of</strong> glucan. Commercial glucan being expensive, a common source <strong>of</strong> glucan from<br />

agricultural waste was used (Talavera-Sevilla, 1996; Talavera et al., 1998a, b).<br />

Step 2. Effect <strong>of</strong> substrate on antagonistic bacteria<br />

The effect <strong>of</strong> different substrates on the growth and survival <strong>of</strong> antagonistic bacteria<br />

were investigated. The aim was to modify the physical and nutritional conditions in<br />

order to inhibit germination and establishment <strong>of</strong> the pathogen and favour antagonistic<br />

organisms.<br />

The bacterial strains used were: Serratia marcescens R1, Serratia entomophila A100<br />

and Bacillus cereus A30. The substrates tested, singly or in combination, were <strong>leaf</strong> extract,<br />

milk, foliar fertilizers, molasses, cassava starch, glucan and chitin. The highest recovery<br />

level <strong>of</strong> bacteria was observed in molasses which had positive effects on antagonistic<br />

microorganisms. A combination <strong>of</strong> milk and molasses increased multiplication and<br />

survival <strong>of</strong> R1 and A30 (Ruiz-Silvera et al., 1997a, b).<br />

Plants treated with a combination <strong>of</strong> chitin, yeast and calcium nitrate alternated<br />

with commonly used fungicides, reduced the number <strong>of</strong> fungicide treatments by 40%<br />

in comparison with fungicides alone (Arango-Ospina, 2000).<br />

289


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Step 3. Internal vs external application <strong>of</strong> antagonistic<br />

microorganisms<br />

The objectives <strong>of</strong> the study were to evaluate the effects <strong>of</strong> R1 and A30 applied<br />

externally in combination with Silwet L-77, Nu-Film 17 and mineral oil, and to<br />

evaluate an endophytic inoculation method via the roots. R1 was compatible with<br />

stickers, mainly mineral oil. Mineral oil in combination with antagonistic<br />

microorganisms reduced disease severity in comparison with the controls. The best<br />

colonization <strong>of</strong> internal tissues was when A30 was applied directly inside the plant<br />

(Miranda, 1996).<br />

Step 4. The phenomenom <strong>of</strong> induced resistance and promotion<br />

<strong>of</strong> plant growth<br />

Stimulation by pathogens, non-pathogenic microorganisms, and by substances<br />

<strong>of</strong> biological or non-biological origin can induce resistance in susceptible plants<br />

Induced resistance to disease and growth promotion have potential for controlling<br />

disease (Figure 2).<br />

Four bacterial and one fungal suspensions were applied to the rhizosphere;<br />

KH 2<br />

PO 4<br />

and K 2<br />

HPO 4<br />

solutions were applied to the leaves as abiotic exogenous<br />

inducers. In a second experiment, microorganisms were evaluated with the addition<br />

<strong>of</strong> sugarcane pulp, sugarcane filter press and c<strong>of</strong>fee husks to the rhizosphere.<br />

Pseudomonas fluorescens (PRA25), P. cepacia (AMMD) and Trichoderma harzianum<br />

(Th) plus substrates increased growth the most and reduced disease in comparison<br />

with the controls (water and substrates). However, the lowest percentage was<br />

obtained with propiconazole (Tilt®). There was a significant and negative<br />

correlation between them (Gutiérrez, 1996).<br />

KH2PO4<br />

Chemical inductors<br />

{<br />

Serratia marcescens R1<br />

Bacillus cereus (A30)<br />

Pseudomonas fluorescens (PRA25)<br />

Pseudomonas cepacia (AMMD)<br />

Trichoderma harzianum (Th)<br />

R1 - A30<br />

PRA25 - AMMD<br />

Th<br />

+<br />

Substrates<br />

sugarcane pulp<br />

sugarcane filter press<br />

c<strong>of</strong>fee busk<br />

Figure 2. Illustration <strong>of</strong> the concept <strong>of</strong> induced resistance to control black <strong>leaf</strong> streak disease (© S. Belalcázar,<br />

2002).<br />

290


Session 5<br />

A.S. Riveros, C.I. Giraldo and A. Gamboa<br />

Step 5. Induced systemic resistance (ISR)<br />

Investigations improved the understanding <strong>of</strong> the use <strong>of</strong> organic amendments,<br />

antagonistic microorganisms, substrates as energy sources in combination with<br />

mycorrhizal fungi and organic extracts known as efficient microorganisms, EMs<br />

(Okumoto et al., 2001; Ayuso, 2000; Sanchez Garita et al., 1998; Okumoto, 1992).<br />

Induced systemic resistance (ISR) to disease results from the inoculation <strong>of</strong> lower<br />

leaves, or roots, with restrictive pathogens, non-pathogenic races <strong>of</strong> pathogens,<br />

non-pathogens, products <strong>of</strong> pathogen or non-pathogens, and organic or inorganic<br />

chemicals. ISR is also referred to in the scientific literature as SIR or SAR (Ku,?<br />

2001).<br />

The objective <strong>of</strong> the research was to evaluate, under greenhouse conditions,<br />

the response <strong>of</strong> cv. ‘Grande naine’, as an example <strong>of</strong> a banana cultivar susceptible<br />

to black <strong>leaf</strong> streak disease, and <strong>of</strong> ‘FHIA-23’ as an example <strong>of</strong> a clone resistance<br />

to the disease, in the presence <strong>of</strong> 3 resistance inducers and one foliar substrate as<br />

an energy source. The inducers were PRA25 bacteria and culture filtrate from<br />

germinating spores <strong>of</strong> M. fijiensis strains according to Riveros and Lepoivre (1998),<br />

and Acilbenzolar-S-metil (BION ® ,), a synthetic inducer provided by Syngenta. ISR<br />

was higher in ‘FHIA-23’ than in cv. ‘Grande naine’.<br />

Vermicompost increased ISR. BION ® resulted in high ISR in both cultivars.<br />

Rhizobacteria and M. fijiensis filtrate induced resistance only in ‘FHIA-23’ when in<br />

the presence <strong>of</strong> an energy source. In the field, BION ® reduced disease incidence in<br />

cv. ‘Grande naine’ in comparison with conventional control measures (Patiño, 2001).<br />

Massive applications <strong>of</strong> antagonistic bacteria or fungi on crops could have<br />

unforeseen effects on the environment. The objectives <strong>of</strong> the study were to evaluate<br />

the in vitro biological activity <strong>of</strong> microbiological filtrates on a M. fijiensis ascospore<br />

preparation, the growth <strong>of</strong> M. fijiensis colonies, and the effects <strong>of</strong> two filtrates on<br />

the cell structure <strong>of</strong> M. fijiensis. Emphasis was put on the isolation <strong>of</strong> strains with<br />

glucanolytic and chitinolytic activity.<br />

Materials and methods<br />

Bacterial strains used in this study were obtained from CATIE’s Plant Protection Unit<br />

Collection:<br />

•Bacillus cereus (A30), isolated from tomato leaves (Lycopersicon pimpinelli),<br />

Turrialba, Costa Rica (Okumoto, 1992).<br />

• Serratia marcescens (R1), isolated from banana leaves (Musa sp.), Limon province,<br />

Costa Rica (González, 1994).<br />

• Serratia entomophila (SE), isolated from Canterbury Valley (New Zealand) from<br />

the digestive tract <strong>of</strong> the scarabid Costelytra zealandica (donated by Trevor Jackson<br />

from the AgResearch Lincoln Laboratory in 1994).<br />

• GS2, GS3, GC1, GBC2, bacteria with glucanolytic activity isolated from banana<br />

leaves, Indiana farm, Siquirres, Costa Rica (Talavera-Sevilla, 1996).<br />

• SE/PO2, isolated from deep well water, Carmen de Siquirres, Costa Rica (Gamboa,<br />

personal communication).<br />

291


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

• White, chitinolytic bacteria, isolated from plantain leaves, La Montaña farm, CATIE,<br />

Costa Rica (Arango-Ospina, 2000).<br />

• Extracts <strong>of</strong> the fluid obtained from suspensions <strong>of</strong> the conidia <strong>of</strong> M. fijiensis isolated<br />

from La Montaña farm, Turrialba, Costa Rica.<br />

Microbiological extracts were prepared on Petri dishes containing nutrient agar<br />

medium (DIFCO). Two boxes per bacterial strain were inoculated with 20 ml bacterial<br />

solution previously maintained at 4°C. The bacterial suspension was uniformly<br />

distributed over the medium using a glass handle and the boxes were incubated at 30°C<br />

for 2 days. Once the bacteria started growing, sterile distilled water was added to the<br />

medium and the bacterial suspension removed with a scalpel; approximately 20 ml <strong>of</strong><br />

each bacterial suspension was transfered to sterile vials. The absorbency at 600 nm <strong>of</strong><br />

a 3-ml solution was measured with a spectrophotometer. One ml <strong>of</strong> each suspension<br />

with an optical density <strong>of</strong> 1.2 was transferred to 200 ml <strong>of</strong> sterile nutrient medium<br />

(DIFCO) and incubated for 12 hours at 30°C and 150 rpm. Absorbency was measured<br />

again at 600 nm and gave values <strong>of</strong> 1.2 ±0.02 after 12 hours<br />

Suspensions were adjusted to an absorbency <strong>of</strong> 1.1 by adding sterile nutrient medium.<br />

Cell-free extracts were obtained by centrifugation at 5000 rpm for 40 minutes followed<br />

by vacuum filtration on 0.22 mm membranes. Extracts were kept at 4°C in sterile flasks<br />

and protected from the light.<br />

A suspension <strong>of</strong> 2x10 5 conidia/ml <strong>of</strong> M. fijiensis in 700 ml <strong>of</strong> sterile distilled water<br />

was agitated at 100 cycles/min for 48 hours in darkness and then filtered using ethamine<br />

and Whatman paper. The residue was lyophilized to obtain 0.341 mg <strong>of</strong> powder, which<br />

was diluted in 70 ml sterile distilled water to obtain a final 20x concentration and filtered<br />

through 0.22 µm Nalgene Disposable Filterware filters. The filtrate was protected from<br />

light and kept at 4°C until its utilization.<br />

Samples <strong>of</strong> plantain leaves with black <strong>leaf</strong> streak disease were transferred to La<br />

Montaña farm, which belongs to CATIE. Using a magnifying glass, fragments <strong>of</strong> viable<br />

perithecia were removed and transferred to the laboratory in paper bags. The samples<br />

were checked using a stereomicroscope, and sections with abundant sporulating lesions<br />

were cut into 2x2 cm pieces. Two to four <strong>of</strong> these pieces were stapled to pieces <strong>of</strong> paper<br />

and incubated in a humid chamber for 24 hours at room temperature. They were then<br />

transferred to sterile distilled water for 5 minutes to hydrate the perithecia. Ascospores<br />

discharged on water agar (4% w/v). Treatments were 0.5, 0.1 and 0.01 ppm dilutions<br />

<strong>of</strong> microbiological filtrates and the controls were without microbiological filtrate or with<br />

fungicide.<br />

Diluted culture filtrates were mixed with 15 ml <strong>of</strong> V8 medium, with constant agitation<br />

and then transferred to Petri dishes; there were 3 replicates per treatment. Seven-dayold<br />

sub-cultures <strong>of</strong> a strain <strong>of</strong> M. fijiensis isolated from La Montaña were used. Twenty<br />

colonies <strong>of</strong> 1–1.5 cm in diameter were excised with a scalpel and transferred to an assay<br />

tube with 3 ml <strong>of</strong> 0.05% (v/v) Tween 20 and agitated in a vortex. The assay tubes were<br />

left to rest for 10 minutes and 10 drops <strong>of</strong> liquid were transferred to each Petri dish<br />

and spread with a glass handle. Dishes were sealed, and incubated in darkness for 5<br />

days at 26°C.<br />

After 5 days <strong>of</strong> incubation, the diameters <strong>of</strong> M. fijiensis colonies were measured<br />

using a micrometer in the 4x field <strong>of</strong> a microscope. Thirty readings were taken,<br />

292


Session 5<br />

A.S. Riveros, C.I. Giraldo and A. Gamboa<br />

frequency intervals <strong>of</strong> amplitude 10 were done and only the 10 data from the<br />

interval with higher frequency were registered to conduct the analysis <strong>of</strong> the<br />

inhibiting effect <strong>of</strong> the microbiological filtrates.<br />

The longest germination tubes were measured using a micrometer in the 10x<br />

or 40x field <strong>of</strong> a microscope. One hundred readings were taken per treatment.<br />

Since bacteria were cultured in nutrient medium, bioassays included a treatment<br />

with this medium.<br />

Initial data were multiplied by a correction factor to transform the values to<br />

microns, 10 and 2.5 for 10x and 40x, respectively. The difference in germ tube<br />

length between the treatment and the control was used to calculate the inhibiting<br />

effect <strong>of</strong> the microbiological filtrates.<br />

Results and discussion<br />

Figure 3A shows preliminary results obtained for growth inhibition <strong>of</strong> M. fijiensis<br />

ascospores discharged onto different concentrations <strong>of</strong> microorganism filtrates. The<br />

percentage <strong>of</strong> inhibition was almost 50% and in some cases higher when the medium<br />

included filtrates <strong>of</strong> bacteria with glucanolytic activity (GBC2) and chitinolytic activity<br />

(SE/PO2 and White).<br />

Regarding colony diameter (Figure 3B) the general tendency remained similar<br />

except that another bacterium with glucanolytic activity (GC1) showed a higher<br />

inhibition which was not fully revealed during the sexual phase <strong>of</strong> ascospore<br />

development.<br />

The crude M. fijiensis filtrate (FCMf) also revealed, a clear inhibiting effect on<br />

ascospore and colony growth at a concentration <strong>of</strong> 0.5 ppm but not at the other<br />

concentrations. This poses the question as to whether the toxin(s) produced by<br />

M. fijiensis spores during the germination process can inhibit the pathogen in a<br />

“suicidal” type <strong>of</strong> action.<br />

After five days <strong>of</strong> incubation, physical growth had stopped in the cultures with<br />

the crude filtrates <strong>of</strong> GBC2 and SE/PO2 at the 0.1 ppm concentration in comparison<br />

to the absolute control (water). Electronic transmission microscopy revealed<br />

modifications at the level <strong>of</strong> cell organelles with a strong presence <strong>of</strong> electrodense<br />

osmophylic globules that were not found neither in the control nor in the transversal<br />

longitudinal cut (Figure 4).<br />

The preliminary results suggest that liquid culture filtrates <strong>of</strong> four bacterial strains<br />

with glucanolytic or chitinolytic activity, and the liquid filtrate <strong>of</strong> germinating spores<br />

<strong>of</strong> a Costa Rican strain <strong>of</strong> M. fijiensis inhibited the growth <strong>of</strong> M. fijiensis germ tubes<br />

and colonies.<br />

Crude liquid preparations diluted from antibiotic(s) or toxin(s) and without<br />

bacterial or fungal cells had similar effects as the fungicide Tilt®.<br />

The promising microbiological filtrates need to be evaluated under greenhouse<br />

and field conditions with or without adjuvant applications.<br />

Acknowledgements<br />

The authors thank INIBAP and INIBAP-LAC for the financial support to conduct this<br />

research.<br />

293


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

140.0<br />

100.0<br />

A<br />

% <strong>of</strong> inhibition<br />

60.0<br />

20.0<br />

- 20.0<br />

- 60.0<br />

- 100.0<br />

- 140.0<br />

A30<br />

R1<br />

SE<br />

GS2<br />

GS3<br />

GC1<br />

GBC2<br />

GBC2<br />

GBC2<br />

White<br />

Tilt<br />

Microbiological filtrates<br />

0.5 ppm 0.1 ppm 0.01 ppm<br />

120.0<br />

100.0<br />

B<br />

% <strong>of</strong> inhibition<br />

80.0<br />

60.0<br />

40.0<br />

20.0<br />

0.0<br />

- 20.0<br />

A30<br />

R1<br />

SE<br />

GS2<br />

GS3<br />

GC1<br />

GBC2<br />

GBC2<br />

GBC2<br />

Microbiological filtrates<br />

White<br />

Tilt<br />

A30:Bacillus cereus; R1:Serratia marcescens; SE: Serratia entomophila;GS2,GS3,GC1 and GBC2:<br />

bacteria with glucanolytic activity; SEPO2 and White: bacteria with chitinolytic activity; FCMf:<br />

conidial fluid; Tilt: triazole fungicide.<br />

Figure 3. Effect <strong>of</strong> three dilutions <strong>of</strong> nine antagonistic microbiological filtrates on A) germ tube growth<br />

and B) colony diameter <strong>of</strong> M. fijiensis .Arrows indicate where crude filtrates affected growth.<br />

294


Session 5<br />

A.S. Riveros et al.<br />

GBC2<br />

0.1 ppm<br />

Control<br />

(water)<br />

SE/P0 2<br />

0.1 ppm<br />

Figure 4. Cytological changes revealed by transmission electron microscopy <strong>of</strong> M. fijiensis hyphae tissues<br />

treated with crude filtrates <strong>of</strong> bacteria with glucanolytic (CBC2) and chitinolytic (SE/PO2) activity.<br />

References<br />

Arango-Ospina M.E. 2000. Manejo de sustratos para el control biológico de Sigatoka negra<br />

(<strong>Mycosphaerella</strong> fijiensis) en el cultivo de banano. Tesis Mag. Sc. Biblioteca ORTON-CATIE,<br />

Turrialba, Costa Rica. 102pp.<br />

Ayuso F. 2000. Influencia de enmiendas orgánicas y un hongo endomicorricico sobre el<br />

nematodo Radopholus similis, en banano Musa (AAA). Tesis Mag. Sc. Biblioteca ORTON-<br />

CATIE, Turrialba, Costa Rica. 114pp.<br />

González R. 1994. Efecto de microorganismos quitinolíticos en el desarrollo de Sigatoka negra<br />

(<strong>Mycosphaerella</strong> fijiensis) en banano. Tesis Mag. Sc. Biblioteca ORTON-CATIE, Turrialba,<br />

Costa Rica. 97pp.<br />

González R., E. Bustamante, Ph. Shannon, S. Okumoto and G. Leandro. 1996. Selección de<br />

microorganismos quitinolíticos en el control de Sigatoka negra (<strong>Mycosphaerella</strong> fijiensis)<br />

en banano. Manejo Integrado de Plagas (Costa Rica) 40:6-11.<br />

Gutiérrez FA. 1996. Estudio de factores en la inducción de resistencia a <strong>Mycosphaerella</strong> fijiensis<br />

y promoción de crecimiento en plantas de banano. Tesis Mag. Sc. Biblioteca ORTON-CATIE,<br />

Turrialba, Costa Rica. 91pp.<br />

Ku J. 2001. Concepts and direction <strong>of</strong> induced systemic resistance in plants and its<br />

application. European Journal <strong>of</strong> Plant Pathology107:7-12.<br />

295


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Miranda J.E. 1996. Evaluación de microorganismos antagonistas al hongo <strong>Mycosphaerella</strong><br />

fijiensis Morelet, colocados en el interior y exterior de la planta de banano. Tesis Mag.<br />

Sc. Biblioteca ORTON-CATIE, Turrialba, Costa Rica. 101pp.<br />

Okumoto S. 1992. Efecto de enmiendas sobre bacterias antagónicas a Alternaria solani en<br />

tomate (Lycopersicon esculentum Mill). Tesis Mag. Sc. Biblioteca ORTON-CATIE, Turrialba,<br />

Costa Rica. 114pp.<br />

Okumoto S., E. Bustamante and A. Gamboa. 2001. Actividad de cepas de bacterias<br />

quitinolíticas antagonistas a Alternaria solani in vitro. Manejo Integrado de Plagas (Costa<br />

Rica) 59:58-62.<br />

Patiño L.F. 2001. Efecto de una fuente de energía, tres inductores de resistencia y un sustrato<br />

foliar sobre Sigatoka negra en banano. Tesis Mag. Sc. Biblioteca ORTON-CATIE, Turrialba,<br />

Costa Rica. 91pp.<br />

Riveros A.S. and P. Lepoivre. 1998. Alternativas bioquímicas para el control indirecto de<br />

Sigatoka en Musáceas. Pp. 436-447. Resúmenes. XIII Reunión ACORBAT, Guayaquil,<br />

Ecuador.<br />

Ruiz-Silvera C., E. Bustamante, F. Jimenez, J.L. Saunders, S. Okumoto and R. Gonzalez. 1997a.<br />

Efecto de sustratos sobre crecimiento y supervivencia de bacterias antagonistas a<br />

<strong>Mycosphaerella</strong> fijiensis. Manejo Integrado de Plagas (Costa Rica) 45:1-8.<br />

Ruiz-Silvera C., E. Bustamante, F. Jimenez, J.L. Saunders, S. Okumoto and R. Gonzalez. 1997b.<br />

Sustratos y bacterias antagonistas para el manejo de <strong>Mycosphaerella</strong> fijiensis en banano.<br />

Manejo Integrado de Plagas (Costa Rica) 45:9-17.<br />

Sánchez Garita V., E. Bustamante and R. Shattock. 1998. Selección de antagonistas para el<br />

control biológico de Phytophthora infestans en tomate. Manejo Integrado de Plagas (Costa<br />

Rica) 48:25-34.<br />

Talavera-Sevilla M.E. 1996. Determinación de ß-glucano en subproductos agrícolas y<br />

evaluación del efecto de microorganismos glucanoliticos sobre <strong>Mycosphaerella</strong> fijiensis<br />

en banano. Selección de antagonistas. Tesis Mag. Sc. Biblioteca ORTON-CATIE, Turrialba,<br />

Costa Rica. 80pp.<br />

Talavera M., E. Bustamante, R. González and V. Sanchez. 1998a. Selección y evaluación en<br />

laboratorio y campo de microorganismos glucanoliticos antagonistas a <strong>Mycosphaerella</strong><br />

fijiensis. Manejo Integrado de Plagas (Costa Rica) 47:24-30.<br />

Talavera M., F. Lopez, E. Bustamante and R. González. 1998b. Extracción y cuantificación<br />

de beta-glucano a partir de sustratos comunes en el trópico. Manejo Integrado de Plagas<br />

(Costa Rica) 47:31-36.<br />

296


Session 5<br />

E. Spaans and L. Quiros<br />

Precision agriculture to improve<br />

management decisions and field<br />

research<br />

E. Spaans and L. Quiros<br />

Abstract<br />

Precision agriculture helps farmers to improve management decisions. Since standard<br />

management practices do not take into account variability in the environment, resources may<br />

be wasted at certain stages <strong>of</strong> production and insufficient in others. Instead <strong>of</strong> calculating an<br />

average over the whole farm, there is a need for a detailed pr<strong>of</strong>it analysis <strong>of</strong> agricultural<br />

enterprises. It should be done at a high spatial resolution in order to make decisions that take<br />

into account local conditions within a field. At the Commercial Farm <strong>of</strong> EARTH University, we are<br />

mapping harvests at a spatial resolution <strong>of</strong> 4 ha within a 110 ha plantation.The network <strong>of</strong> railways<br />

in the plantation is used to georeference the origin <strong>of</strong> the fruit which, together with the weight<br />

<strong>of</strong> the fruits, is stored in a database. The costs are divided into fixed and variable costs, and<br />

subtracted from the income to produce a map <strong>of</strong> pr<strong>of</strong>its. The spatial variability <strong>of</strong> the harvest<br />

was enormous, with a greater than 300% difference within the field. To investigate the possible<br />

causes <strong>of</strong> the variability, we are monitoring parameters in the field that affect the growth <strong>of</strong><br />

banana, e.g. soil fertility, plant nutrition, functional roots and age <strong>of</strong> plantation. Correlation<br />

coefficients between production and each <strong>of</strong> the parameters were calculated. The coefficients<br />

can be used to decide exactly what needs to be done in the areas not generating sufficient pr<strong>of</strong>it.<br />

This scheme <strong>of</strong> intensive and systematic data acquisition, and interpretation provides<br />

opportunities for field research and hence to improve management practices including the control<br />

<strong>of</strong> Sigatoka disease. The project is in progress and we present the preliminary results we have<br />

obtained.<br />

Resumen - Agricultura de precisión como base para mejorar las decisiones de manejo<br />

e investigación en la finca<br />

La agricultura de precisión es una herramienta eficaz para ayudar a los agricultores a tomar<br />

mejores decisiones de manejo. Debido a que las prácticas de manejo normalizadas no reconocen<br />

la variabilidad del ambiente, se gastan recursos en algunas áreas, mientras que en otras no se<br />

invierte lo suficiente. Este hecho requiere realizar un análisis de ganancias detallado de nuestra<br />

empresa agrícola; de esta manera, no es el promedio global de toda la finca, sino una resolución<br />

Escuela de Agricultura de la Región Tropical Húmedo (EARTH University), San José, Costa Rica<br />

297


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

espacial alta, la que permite hacer tomas de decisiones que respondan a las condiciones locales<br />

en el campo. Los ingresos provienen de la cosecha, así, que hacemos un mapa de la cosecha con<br />

una resolución espacial de 4 ha dentro de una plantación de 110 ha en la Finca Comercial de la<br />

Universidad EARTH. La red de ferrocarril en la plantación sirve para hacer referencia geográfica<br />

al origen de la fruta, que, junto con su peso, medido en la planta empacadora, se almacena en<br />

una base de datos. Los costos se dividen en costos fijos y variables, y se restan del ingreso para<br />

producir un mapa de ganancias. Se descubrió que la variabilidad espacial de la cosecha fue enorme,<br />

con más de 300% de diferencias dentro del campo. Para investigar las causas posibles de esta<br />

variabilidad, empezamos a monitorear varios parámetros en el campo que afectan el crecimiento<br />

de los bananos, como la fertilidad del suelo, nutrición de la planta, raíces funcionales y la edad<br />

de la plantación, y luego calculamos los coeficientes de correlación entre la productividad por<br />

un lado y cualesquiera de los parámetros del campo por otro. Este coeficiente puede ser utilizado<br />

como una guía para decidir las necesidades exactas que deben ser cumplidas en cada una de<br />

las áreas que no generan ganancias suficientes. Este esquema de adquisición de datos intensivo<br />

y sistemático y su interpretación también permiten que las investigaciones en el campo mejoren<br />

eficazmente las prácticas de manejo, incluyendo el control de la Sigatoka. Este proyecto se<br />

encuentra en progreso y se discutirán los resultados preliminares.<br />

Résumé - L’agriculture de précision pour améliorer les décisions de gestion et la<br />

recherche en champ<br />

L’agriculture de précision aide les agriculteurs à améliorer leurs décisions de gestion. Les pratiques<br />

de gestion standard ne prennent pas en compte la variabilité de l’environnement, et des ressources<br />

pourraient donc être dilapidées à certains stades de la production mais être insuffisantes à d’autres.<br />

Au lieu de calculer une moyenne sur toute l’exploitation,une analyse détaillée du pr<strong>of</strong>it est nécessaire<br />

pour les entreprises agricoles. Elle devrait être réalisée avec une résolution spatiale élevée afin de<br />

prendre des décisions qui tiennent compte des conditions locales au sein du champ. Dans la ferme<br />

commerciale de l’Université EARTH, nous cartographions les récoltes avec une résolution spatiale de<br />

4 ha sur une plantation de 110 ha. Le réseau des voies ferrées est utilisé pour géoréférencer l’origine<br />

des fruits qui, avec le poids des fruits, est stocké dans une base de données. Les coûts sont divisés en<br />

coûts fixes et variables, et soustraits du revenu pour produire une cartographie des bénéfices. La<br />

variabilité spatiale de la récolte s’est avérée énorme, avec une différence de plus de 300% à l’intérieur<br />

d’un même champ. Afin de rechercher les causes possibles de cette variabilité, nous faisons le suivi<br />

des paramètres en champ qui affectent la croissance des bananiers, tels que la fertilité du sol, la<br />

nutrition des plants, les racines fonctionnelles et l’âge de la plantation. Les coefficients de corrélation<br />

entre la production et chacun des paramètres ont été calculés. Les coefficients peuvent être utilisés<br />

pour décider exactement ce qui doit être fait dans les zones qui ne génèrent pas de bénéfices suffisants.<br />

Ce schéma intensif et systématique d’acquisition de données, ainsi que son interprétation, <strong>of</strong>frent<br />

des occasions de recherche en champ qui pourraient permettre d’améliorer les pratiques de gestion,<br />

y compris la lutte contre les cercosporioses. Le projet est en cours et nous présentons les premiers<br />

résultats que nous avons obtenus.<br />

Introduction<br />

Globalization has opened up markets and increased competition between agricultural<br />

producers worldwide. In order to continue being competitive, farmers must improve<br />

the efficiency <strong>of</strong> their production systems, i. e. reduce costs while maintaining or<br />

even improving the production as well as the social and environmental impact <strong>of</strong><br />

their agricultural enterprise.<br />

Productivity has increased considerably over the last century due to technological<br />

advances (e.g. in plant nutrition and fertilization, genetic improvement, mechaniza-<br />

298


Session 5<br />

E. Spaans and L. Quiros<br />

tion and pest management) that have allowed the agricultural community to respond<br />

to the increasing demands for agricultural products. These advances and the<br />

improved understanding <strong>of</strong> plant-environment interactions have resulted in the<br />

development <strong>of</strong> technological packages for major crops. A technological package is<br />

a recommended set <strong>of</strong> agricultural practices that describe all stages <strong>of</strong> production,<br />

e.g. soil preparation, crop planting, pest and disease management, soil fertilization,<br />

and harvesting and packing methods. These generalized recommendations were<br />

developed for average conditions and do not take into account specific conditions<br />

encountered in the field. Inevitably, resources are wasted in some areas, hence raising<br />

production costs and increasing the risk <strong>of</strong> environmental contamination. Similarly,<br />

not enough resources may be invested in certain areas, resulting in suboptimal growth<br />

and loss <strong>of</strong> income.<br />

The reason is that resources (soil, weather, water, etc.) are not homogenous<br />

throughout the farm and over time. Moreover, the socioeconomic conditions <strong>of</strong> the<br />

enterprise (the nature <strong>of</strong> the market, prices, policies and standards <strong>of</strong> certification,<br />

amongst others) are also in a constant state <strong>of</strong> change.<br />

This calls for a more entrepreneurial approach to agriculture. We need to improve<br />

decision making, to take more precise ones to fine-tune the management <strong>of</strong> the<br />

resources to what is really needed. That is precisely what precision agriculture is all<br />

about: doing the right thing, at the right time and at the right place. To answer the<br />

fundamental question <strong>of</strong> what is the right thing to do, we need to acquire detailed<br />

information about the production system at the adequate spatial resolution.<br />

Traditional applications <strong>of</strong> precision agriculture involve high-tech data acquisition:<br />

a Global Positioning System (GPS) for georeferencing the data and monitoring sensors<br />

mounted on harvest equipment to gather high resolution data while harvesting. A<br />

Geographic Information System (GIS) is used to manage the large amount <strong>of</strong> data<br />

and map them. These requirements have hindered the implementation <strong>of</strong> precision<br />

agriculture in many Latin American production systems where harvest is <strong>of</strong>ten done<br />

manually and access to and support for technology is sometimes limited. We think<br />

it is more useful to reflect on the basic principle <strong>of</strong> precision agriculture and then<br />

creatively develop the implementation, considering the specific conditions <strong>of</strong> each<br />

production system.<br />

This paper presents our interpretation <strong>of</strong> precision agriculture and its<br />

implementation at the commercial banana farm <strong>of</strong> EARTH University. Its relevance<br />

to this workshop is that it improves decision-making and provides a unique<br />

opportunity for farm research, including the study <strong>of</strong> the effectiveness <strong>of</strong> agricultural<br />

practices to control the <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> caused by <strong>Mycosphaerella</strong>.<br />

Implementation <strong>of</strong> precision agriculture<br />

Being competitive means optimizing pr<strong>of</strong>its, not harvest. Considering the spatial and<br />

temporal variability <strong>of</strong> resources throughout the farm, it is to be expected that pr<strong>of</strong>its<br />

vary as well, particularly if the same technological package is used over the entire<br />

farm. Thus an overall financial analysis <strong>of</strong> the farm is not sufficient; we need a<br />

detailed analysis <strong>of</strong> costs and benefit throughout the entire farm, in order to allow<br />

for precise management.<br />

299


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

As an example we describe banana production on a 110 ha farm located on the<br />

campus <strong>of</strong> EARTH University in the Humid Tropics <strong>of</strong> the Atlantic Zone <strong>of</strong> Costa<br />

Rica. Since the financial benefit results from the harvested banana, a precise<br />

monitoring <strong>of</strong> the harvest is needed to precisely monitor income. The farm was divided<br />

into blocks <strong>of</strong> 4 ha (100 m wide by 400 m long). The railway that transports the<br />

fruits runs through the center <strong>of</strong> each block, and fruits are collected on a distance<br />

<strong>of</strong> 50 m on each side <strong>of</strong> the tracks. Each block has an identification, and printed<br />

labels are provided to the harvesters to identify the geographical origin <strong>of</strong> every<br />

bunch. At the packing plant, each bunch is weighed and the information, together<br />

with the geographical origin (spatial component) and the day <strong>of</strong> harvest (temporal<br />

component), is stored in a database. This mapping system is inexpensive and lowtech,<br />

and does not require GPS since the railway network serves as the geographical<br />

reference within the plantation.<br />

The <strong>bananas</strong> are still attached to the stalk when they are weighed. To obtain the<br />

weight <strong>of</strong> the <strong>bananas</strong>, the weight <strong>of</strong> the bunch is reduced by 10% to account for<br />

the weight <strong>of</strong> the stalk. Finally, a general value <strong>of</strong> 12% is used to account for the<br />

rejected banana (because <strong>of</strong> inadequate quality), and the resulting weight is divided<br />

by 18.14 to obtain the number <strong>of</strong> boxes exported. Data were collected for the period<br />

February to October <strong>of</strong> the year 2001.<br />

The costs <strong>of</strong> banana production were divided into fixed and variable costs. Fixed<br />

costs were equally divided over the entire productive surface area <strong>of</strong> the plantation,<br />

while the variable costs were equally divided over the weight <strong>of</strong> the bunches.<br />

Costs are currently recorded for the entire farm, thus we do not have the same<br />

spatial resolution as for the production costs. The quantity <strong>of</strong> rejected <strong>bananas</strong> is<br />

also an average value reported by the packing plant. These factors limit data analysis,<br />

but the situation shows how in practice the implementation <strong>of</strong> precision agriculture<br />

is a process <strong>of</strong> gradual improvements in data acquisition and interpretation as the<br />

entire team familiarizes itself with the process. This year, for example, we started<br />

recording production cost per block.<br />

The cost and harvest data are used to determine the pr<strong>of</strong>it <strong>of</strong> every block within<br />

the plantation. In order to explain and eventually reduce the variability in harvest,<br />

we also monitor field parameters that we know affect plant growth, like soil fertility,<br />

plant nutritional status, nematode infestation (as expressed in % <strong>of</strong> functional roots),<br />

and age <strong>of</strong> plantation. This last parameter was included because the plantation was<br />

planted about 40 years ago and recently some areas have been renewed. To determine<br />

the degree to which each field parameter affects the production <strong>of</strong> <strong>bananas</strong>,<br />

correlation coefficients were calculated between the harvest and each <strong>of</strong> the field<br />

parameters.<br />

Results and discussion<br />

The productivity <strong>of</strong> <strong>bananas</strong> varied widely over the entire farm, ranging from<br />

943 to 3040 boxes/ha/year, with an average <strong>of</strong> 1538 boxes/ha/year. These numbers<br />

translated into a loss <strong>of</strong> US$2655/ha/year in the least productive block, and a pr<strong>of</strong>it<br />

<strong>of</strong> US$2316/ha/year in the most productive block. Overall, the financial loss <strong>of</strong> the<br />

plantation was US$1245/ha/yea. This shows the usefulness <strong>of</strong> precision agriculture,<br />

300


Session 5<br />

E. Spaans and L. Quiros<br />

since we can now locate the problem areas and quantify their impact on the financial<br />

return <strong>of</strong> the farm.<br />

So what should we do with this information, and where and when? To answer<br />

these questions we analysed the data statistically. Although we found a range <strong>of</strong><br />

soil fertility, plant nutritional status and nematode infestation, none <strong>of</strong> these<br />

parameters were significantly correlated (P


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

We propose that this type <strong>of</strong> on-farm research also be used to identify farmspecific<br />

optimum values or tolerance levels, which could lead to a better interpretation<br />

<strong>of</strong> field data. Furthermore, this implementation <strong>of</strong> precision agriculture could be used<br />

to rapidly and effectively evaluate alternative practices, by selecting those blocks<br />

that meet certain criteria and applying different treatments to them. The monitoring<br />

system provides us with continuous and timely feedback about the plant response<br />

to the different treatments. We are now including the monitoring <strong>of</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong><br />

in the plantation, which will allow us to investigate the interaction between the<br />

disease and other field parameters, and determine the tolerance levels for the desired<br />

production.<br />

Conclusion<br />

In this paper we presented the principles <strong>of</strong> precision agriculture, and discussed a<br />

case where we implemented precision agriculture on a commercial banana plantation.<br />

The implementation is a gradual process, and we are still far away from a precisely<br />

managed farm. However, we have shown that the system has already provided<br />

valuable information which we have been able to use to take effective management<br />

decisions. The continuous and precise data collection inherent to precision agriculture<br />

should make it possible to rapidly and effectively investigate alternative practices.<br />

302


Session 5<br />

S. Knight et al.<br />

Poster<br />

The role <strong>of</strong> managing resistance<br />

to fungicides in maintaining<br />

the effectiveness <strong>of</strong> integrated<br />

strategies to control<br />

black <strong>leaf</strong> streak disease<br />

S. Knight 1 ,M.Wirz 1 ,A.Amil 2 ,A.Hall 2 and M. Shaw 3<br />

Abstract<br />

Fungicide programmes are essential for commercial production <strong>of</strong> banana in all regions where<br />

<strong>Mycosphaerella</strong> fijiensis is prevalent. A key factor influencing the design <strong>of</strong> fungicide programmes is<br />

the importance <strong>of</strong> following resistance management principles, in order to preserve long-term<br />

effectiveness. Resistance management is based on the appropriate limitation and alternation <strong>of</strong><br />

products that have a site-specific mode <strong>of</strong> action. The introduction in 1997 <strong>of</strong> the first strobilurin<br />

fungicide, azoxystrobin, represented a significant step forward in the integrated control <strong>of</strong> black <strong>leaf</strong><br />

streak disease, because <strong>of</strong> its efficacy and favourable environmental and toxicological pr<strong>of</strong>ile. In<br />

recognition <strong>of</strong> its site-specific mode <strong>of</strong> action,anti-resistance management guidelines were developed<br />

by the Fungicide Resistance Action Committee (FRAC) before its commercial introduction. The first<br />

strobilurin-resistant individuals were documented in 2000 in Costa Rica,and resistance to strobilurin<br />

has reached high levels on some farms in the main banana production zones <strong>of</strong> Costa Rica.<br />

Molecular characterization <strong>of</strong> resistant isolates has identified the cause <strong>of</strong> resistance as a single point<br />

mutation in the fungal target protein, cytochrome b. A large-scale population dynamics study is in<br />

progress to examine the evolution <strong>of</strong> resistance in the field, using molecular techniques (PCR). The<br />

factors that influence this evolution (disease pressure, climate, fungicide spray programme) will be<br />

examined, and the extent <strong>of</strong> migration <strong>of</strong> the resistant population will be estimated. The study will<br />

enable recommendations to be validated or improved,and will support efforts to limit the proliferation<br />

<strong>of</strong> resistance to this important group <strong>of</strong> fungicides.<br />

1<br />

Syngenta Costa Rica SA, San José, Costa Rica<br />

2<br />

Jealott’s Hill International Research Centre, Bracknell, UK<br />

3<br />

University <strong>of</strong> Reading, Reading, UK<br />

303


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Resumen - Papel del manejo de la resistencia a funguicidas en el mantenimiento de la<br />

eficacia de las estrategias integradas para el control de la Sigatoka negra<br />

Los programas de funguicidas siguen siendo esenciales para la producción bananera comercial en<br />

todas las regiones donde prevalece <strong>Mycosphaerella</strong> fijiensis.Un factor clave que influye sobre el diseño<br />

de los programas de aplicación de funguicidas es la importancia de seguir los principios del manejo<br />

de la resistencia, con el fin de preservar la eficacia a largo plazo. El manejo de la resistencia se basa<br />

en la limitación y alternación apropiadas de los productos que tienen un modo de acción específico<br />

en el sitio. La introducción en 1997 del primer funguicida strobilurina, azoxistrobina, representó un<br />

significativo paso hacia adelante en el control integrado de la Sigatoka negra, debido a su excelente<br />

eficacia y perfil ambiental y toxicológico favorable.En reconocimiento de este modo de acción específico<br />

del sitio, se desarrollaron guías de manejo anti-resistencia por el Comité de Acción de Resistencia a<br />

los Fungicidas (Fungicide Resistance Action Committee (FRAC) antes de su introducción comercial. Los<br />

primeros individuos resistentes a la strobilurina fueron documentados en 2000 en Costa Rica, y la<br />

resistencia a la strobilurina ha alcanzado niveles altos en algunas fincas en las principales zonas<br />

productoras de banano de Costa Rica. La caracterización molecular de los aislados resistentes ha<br />

identificado como la causa de la resistencia una mutación puntual individual en la proteína diana<br />

fungosa, citochromo b. Un estudio a gran escala de las dinámicas de la población se está llevando a<br />

cabo actualmente para examinar la evolución de la resistencia en el campo, utilizando técnicas<br />

moleculares (PCR). Se examinarán los factores que influyen sobre esta evolución (presión de la<br />

enfermedad, clima, programa de rociado de fungicidas), y se estimará la extensión de migración de<br />

la población resistente. El estudio permitirá validar o mejorar las recomendaciones para el manejo<br />

de la resistencia, y apoyará los esfuerzos que se realizan para limitar la proliferación de la resistencia<br />

en este importante grupo de fungicidas.<br />

Résumé – Le rôle de la gestion de la résistance aux fongicides dans le maintien de<br />

l’efficacité des stratégies de lutte intégrée contre la maladie des raies noires<br />

Les plantations commerciales de bananes sont extrêmement dépendantes des applications de<br />

fongicides partout où <strong>Mycosphaerella</strong> fijiensis prévaut. Un facteur important qui influence la<br />

conception des programmes d’arrosage est le respect des principes de gestion de la résistance aux<br />

fongicides afin de préserver leur efficacité. La gestion de la résistance repose sur la restriction de l’usage<br />

et l’alternance de produits qui ont un mode d’action spécifique. L’introduction en 1997 du premier<br />

fongicide à base de strobilurine, l’azoxystrobine, représente une avancée importante dans la lutte à<br />

la maladie des raies noires étant donné son efficacité et son pr<strong>of</strong>il environnemental et toxicologique<br />

favorable. En reconnaissance de son mode d’action spécifique, des lignes directrices pour empêcher<br />

le développement de résistance ont été élaborées par le Fungicide Resistance Action Committee (FRAC)<br />

avant sa distribution commerciale. Les premiers cas de résistance à la strobilurine ont été documentés<br />

en 2000 au Costa Rica,et les niveaux de résistance ont atteint des taux très élevés sur certaines fermes<br />

des principales zones de production bananière au Costa Rica. La caractérisation d’isolats résistants<br />

a permis de remonter à la source de la résistance, une mutation isolée dans une protéine ciblée du<br />

champignon, le cytochrome b. Une étude de population à grande échelle utilisant des techniques<br />

moléculaires (PCR) est en cours afin de suivre l’évolution de la résistance en champ. Les facteurs qui<br />

influencent cette évolution (la pression de la maladie, le climat, le programme d’application des<br />

fongicides) seront examinés et l’étendue de la migration de la population résistante estimée. L’étude<br />

permettra de valider ou améliorer les recommandations et participera à limiter la prolifération de la<br />

résistance à ce groupe important de fongicides.<br />

Introduction<br />

A key factor influencing the design <strong>of</strong> programmes to control <strong>Mycosphaerella</strong> fijiensis<br />

is the importance <strong>of</strong> following resistance management guidelines to ensure the long<br />

term effectiveness <strong>of</strong> site-specific fungicides.<br />

304


Session 5<br />

S. Knight et al.<br />

The introduction in 1997 <strong>of</strong> the first strobilurin fungicide, Bankit<br />

(azoxystrobin), represented a significant step forward in the integrated control<br />

<strong>of</strong> black <strong>leaf</strong> streak disease, due to its excellent efficacy and favourable<br />

environmental and toxicological pr<strong>of</strong>ile. It has a site-specific mode <strong>of</strong> action, the<br />

inhibition <strong>of</strong> electron transport at the Qo site <strong>of</strong> cytochrome bc 1<br />

. Anti-resistance<br />

management guidelines were developed by the Fungicide Resistance Action<br />

Committee (FRAC) before commercial introduction, and a global sensitivity<br />

monitoring programme was initiated.<br />

The first strobilurin-resistant individuals were documented in 2000 in Costa<br />

Rica, and resistance to strobilurin has reached high levels on some farms in the<br />

main banana production zones <strong>of</strong> Costa Rica. Molecular characterization <strong>of</strong><br />

resistant isolates has identified the cause <strong>of</strong> resistance as a single point mutation<br />

in the fungal target protein, cytochrome b, known as G 143<br />

A. Resistant isolates can<br />

be detected via quantitative polymerase chain reaction analysis.<br />

A large-scale population dynamics study was initiated to examine the<br />

evolution <strong>of</strong> resistance in the field by using molecular analysis. The factors that<br />

influence this evolution (disease pressure, climate, fungicide spray programme)<br />

will be examined, and the extent <strong>of</strong> migration <strong>of</strong> the resistant population will be<br />

estimated.<br />

Materials and methods<br />

In vitro bioassay<br />

Sensitivity <strong>of</strong> M. fijiensis to azoxystrobin, was evaluated using the in vitro<br />

methodology recommended by the FRAC.<br />

Sporulating tissue collected from the field was allowed to discharge onto<br />

fungicide-amended agar, and elongation <strong>of</strong> ascospore germ tubes was measured<br />

at 1 ppm and 10 ppm <strong>of</strong> azoxystrobin. A 75% growth or more at the discriminating<br />

rate <strong>of</strong> 10 ppm relative to the control indicated resistance to strobilurin.<br />

Molecular detection <strong>of</strong> strobilurin resistance<br />

The presence <strong>of</strong> the G 143<br />

A mutation was detected using a diagnostic primer for<br />

G 143<br />

A. This primer is extended only when the G 143<br />

A mutation is present in the<br />

sample.<br />

Field population dynamics in Costa Rica<br />

Repeated sampling <strong>of</strong> infected foliar tissue by using a multilayered hierarchical<br />

sampling structure is in progress in Costa Rica. Plantations were selected to meet<br />

the following criteria:<br />

•Presence/absence <strong>of</strong> strobilurin selection pressure;<br />

•Presence/absence <strong>of</strong> resistance mutation;<br />

•High/low levels <strong>of</strong> relative disease pressure.<br />

305


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Results and discussion<br />

Global resistance monitoring results<br />

The results <strong>of</strong> the global sensitivity monitoring for azoxystrobin are shown in Table 1.<br />

Table 1. Prevalence <strong>of</strong> resistance to azoxystrobin (measured as >75% growth at 10 ppm relative to control) in Central<br />

and South America (Number <strong>of</strong> subpopulations tested given in brackets).<br />

Country<br />

Prevalence <strong>of</strong> resistance (% <strong>of</strong> resistant spores)<br />

1999 2000 2001<br />

Mexico - 0.0 (2) -<br />

Belize - - -<br />

Honduras 0.0 (2) 0.0 (3) 0.0 (30)<br />

Guatemala 0.0 (4) 0.0 (4) 0.0 (25) *<br />

Nicaragua - - 0.0 (3)<br />

Costa Rica - 11.3 (33) 11.1 (78)<br />

Panama - 0.0 (5) 3.3 (15)<br />

Colombia - 0.0 (4) 0.03 (25)<br />

Ecuador - - 0.01 (28)**<br />

Cameroon - - 0.0 (12)<br />

*Some resistant individuals were detected in two farms through PCR analysis.<br />

**False positive (PCR analysis demonstrated that resistance was not present).<br />

Molecular analysis<br />

A high degree <strong>of</strong> correlation (P≤0.01) was observed between extensive germ tube<br />

growth at 10 ppm <strong>of</strong> zoxystrobin and detection <strong>of</strong> the G 143<br />

A resistance mutation<br />

(Figure 1). The correlation between the PCR data and bioassay data at 1 ppm was<br />

lower (0.8 compared to 0.94). In other words, extensive germ tube growth, normally<br />

associated with resistance, is occasionally detected on 1 ppm agar in the absence<br />

<strong>of</strong> G 143<br />

A. This may indicate an alternative mechanism <strong>of</strong> resistance, and studies are<br />

in progress to address this question.<br />

Table 2 gives the prevalence <strong>of</strong> resistance in populations <strong>of</strong> M. fijiensis sampled<br />

from 16 different sites in Costa Rica. Ten samples showed varying levels <strong>of</strong> G 143<br />

A<br />

detection, whilst 6 sample populations remained sensitive.<br />

Conclusion<br />

The ongoing population dynamics study will enable the validation or improvement<br />

<strong>of</strong> the management guidelines regarding resistance to strobilurin, and will support<br />

efforts to limit the proliferation <strong>of</strong> resistance to this important group <strong>of</strong> fungicides.<br />

Fungicide programmes are expected to remain indispensable for commercial<br />

banana production for the foreseeable future. A significant increase in the number<br />

<strong>of</strong> alternative chemical modes <strong>of</strong> action to control <strong>of</strong> Sigatoka <strong>diseases</strong> is unlikely<br />

within the next 5-10 years. Therefore resistance management will remain a prime<br />

consideration in the design <strong>of</strong> control programmes, underpinned by appropriate<br />

monitoring efforts and a judicious revision <strong>of</strong> the guidelines.<br />

306


Session 5<br />

S. Knight et al.<br />

100<br />

Baseline<br />

Farm 1<br />

Farm 2<br />

Farm 3<br />

% <strong>of</strong> ascopores<br />

75<br />

50<br />

PCR<br />


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Recommendations <strong>of</strong> session 5<br />

Integrated disease management<br />

Strategies to control black <strong>leaf</strong> streak disease and Sigatoka disease can, according to the<br />

country and the scale <strong>of</strong> production, include not only chemical and cultural practices but<br />

also the use <strong>of</strong> mixed crops or resistant clones. The important inhibitory effect <strong>of</strong> some natural<br />

substances derived from microorganisms antagonistic to fungi, have also been reported as<br />

effective in reducing, in vitro, the development <strong>of</strong> M. fijiensis.<br />

It was recommended to integrate working groups from different disciplines to develop<br />

an achievable IPM approach to manage Sigatoka <strong>diseases</strong>.<br />

Chemical strategies and/or the use <strong>of</strong> improved hybrids should always be used jointly<br />

with adequate agricultural practices to maximize yield and efficacy <strong>of</strong> the<br />

management practices.<br />

It was recommended to study natural/synthetic substances capable <strong>of</strong> promoting or<br />

activating systemic acquired resistance in the broad sense.<br />

It was recommended to evaluate the feasibility <strong>of</strong> precision agriculture farming to<br />

optimize disease management.<br />

It was recommended to assess different crop systems with potential positive impact<br />

on disease management.<br />

It was recommended to include the FRAC’s Banana Working Group Guidelines<br />

(http://www.gcpf.org/frac) for fungicide resistance management in order to broaden<br />

the knowledge <strong>of</strong> such guidelines.<br />

It was recommended to develop alternative/improved methods/equipments for groundbased<br />

applications that can be used by smallholders.<br />

308


List <strong>of</strong> participants<br />

Catherine Abadie*<br />

CARBAP<br />

BP 832 Douala<br />

Cameroon<br />

* Current address<br />

CIRAD-FLHOR<br />

Avenue Agropolis TA 40/02<br />

34398 Montpellier Cedex 5<br />

France<br />

Tel.: +33 467616529<br />

Fax: +33 467615793<br />

E-mail: catherine.abadie@cirad.fr<br />

Juan Fernando Aguilar<br />

FHIA<br />

P.O. Box 2067<br />

San Pedro Sula<br />

Honduras<br />

Tel.: +504 6 682470<br />

Fax: +504 6 682313<br />

E-mail: jaguilar@fhia.org.hn<br />

Maria Elena Aguilar<br />

CATIE<br />

Lab. de Biotecnología<br />

Apdo 7170, Turrialba<br />

Costa Rica<br />

Tel.: +506 5566455<br />

Fax: +506 5562626<br />

E-mail: aguilarm@catie.ac.cr<br />

Yelenis Alvarado Capó<br />

Instituto de Biotecnología de las Plantas<br />

Carretera a Camajuaní km 5<br />

Santa Clara, Villa Clara<br />

Cuba<br />

Tel.: +53 42 281257<br />

Fax: +53 42 281329<br />

E-mail: yalvarado@uclv.edu.cu<br />

Sergio Mauricio Aponte<br />

CORPOICA<br />

Km 14 via Mosquera<br />

A.A. 240142<br />

Las Palmas, Bogotá D.C.<br />

Colombia<br />

Tel.: +57 1 3443000<br />

Fax: +57 1 3441435<br />

E-mail: apontesergio@yahoo.com<br />

Maria J. Barbosa Cavalcante<br />

EMBRAPA<br />

Rod. BR-364 Km 14<br />

Caixa Postal 321<br />

Rio Branco<br />

Acre<br />

Brazil<br />

Tel.: +55 68 2123200<br />

Fax: +55 68 2123200<br />

E-mail: maju@cpafac.embrapa.br<br />

Silvio Belalcázar Caravajal<br />

INIBAP Honorary Research Fellow<br />

Latin America and Caribbean Regional Office<br />

C/o CATIE<br />

7170 Turrialba<br />

Costa Rica<br />

Tel.: +506 5562431<br />

Fax: +506 5562431<br />

E-mail: belcar@armenia.multi.net.co<br />

Peter Burt<br />

Natural Resources Institute<br />

University <strong>of</strong> Greenwich<br />

Chatham, Kent ME4 4TB<br />

United Kingdom<br />

Tel.: +44 634 883231<br />

Fax: +44 634 880066/77<br />

E-mail: P.J.A.Burt@greenwich.ac.uk<br />

Russell Caid<br />

Consultant<br />

Chiquita Brands International<br />

Kentucky<br />

USA<br />

E-mail: RussCaid@aol.com<br />

Jean Carlier<br />

CIRAD-AMIS<br />

Avenue Agropolis TA 40/02<br />

34398 Montpellier Cedex 5<br />

France<br />

Tel.: +33 467616529<br />

Fax: +33 467615793<br />

E-mail: jean.carlier@cirad.fr<br />

311


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Arllen Carpio Brenes<br />

EARTH<br />

Apdo Postal 4442-1000<br />

San José<br />

Costa Rica<br />

Tel.: +506 7 130000<br />

Fax: +506 7 130001<br />

E-mail: afcarpio@earth.ac.cr<br />

Laura Conde<br />

CICY<br />

Calle 43<br />

Col. Churburna de Hidalgo<br />

CP 97200 Merida<br />

Mexico<br />

Tel.: +52 99 813914<br />

Fax: + 52 99 813900<br />

E-mail: laura@cicy.mx<br />

Zilton Cordeiro<br />

EMBRAPA<br />

Rua EMBRAPA s/n<br />

Caixa Postal 007<br />

Cruz das Almas<br />

44380-000 Bahia<br />

Brazil<br />

Tel.: +55 021 756212120<br />

Fax: +55 021 756211118<br />

E-mail: zilton@cnpmf.embrapa.br<br />

Pedro Crous<br />

Department <strong>of</strong> Plant Pathology<br />

University <strong>of</strong> Stellenbosch<br />

P.Bag X1<br />

Matieland 7602<br />

South Africa<br />

Tel.: +27 218 084 796<br />

Fax: +27 218 084 956<br />

E-mail: PWC@sun.ac.za<br />

Fritz Elango<br />

EARTH<br />

Apartado 4442-1000<br />

San José<br />

Costa Rica<br />

Tel.: +506 7 130000<br />

Fax: +506 7 130133<br />

E-mail: felango@earth.ac.cr<br />

Jean-Vincent Escalant<br />

INIBAP<br />

Parc Scientifique Agropolis II<br />

34397 Montpellier Cedex 5<br />

France<br />

Tel.: +33 467611302<br />

Fax: +33 467610334<br />

E-mail: j.escalant@cgiar.org<br />

Emily Fabregar<br />

Lapanday Agricultural & Development<br />

Corporation<br />

Maryknoll Road<br />

Bo Pampanga<br />

Davao City<br />

The Philippines<br />

Tel.: +63 82 2352551<br />

Fax: +63 82 2342359<br />

E-mail: egf@skyinet.net<br />

Henry Fagan<br />

WIBDECO<br />

Manoel Street, Compton Bldg<br />

PO Box 115<br />

Castrie, Saint Lucia<br />

W.I.<br />

Tel.: +1 758 4522411<br />

Fax: +1 758 4514601<br />

Pedro Ferreira<br />

CATIE<br />

7170 Turrialba<br />

Costa Rica<br />

Tel.: +506 5566431<br />

Fax: +506 5561533<br />

E-mail: CATIE@catie.ac.cr<br />

Emile Frison<br />

INIBAP<br />

Parc Scientifique Agropolis II<br />

Montpellier 34397 Cedex 5<br />

France<br />

Tel.: +33 467611302<br />

Fax: +33 467610334<br />

E-mail: e.frison@cgiar.org<br />

José Garza<br />

INIFAP<br />

Campo Exp. Tecoman<br />

Km 35 Carretera Colima Manzanillo<br />

Colima, Mexico<br />

Tel.: +31 33 240133<br />

E-mail:<br />

tecoman@cirpac.inifap.conacy.mx<br />

312


List <strong>of</strong> participants<br />

Friedhelm Gauhl<br />

Chiquita<br />

PO Box 025216-1582<br />

Miami, Florida<br />

33102-5216<br />

USA<br />

Tel.: +1 506 2042001<br />

Fax: +1 506 2042397<br />

E-mail: fgauhl@chiquita.com<br />

Kathy Grice<br />

Centre for Tropical Agriculture<br />

28 Peters St<br />

Mareeba Qld 4880<br />

Australia<br />

Tel.: +61 740 928555<br />

Fax: +61 740 923593<br />

E-mail: Kathy.Grice@dpi.qld.gov.au<br />

Mauricio Guzman<br />

CORBANA SA<br />

La Rita de Pococi<br />

Limón<br />

Costa Rica<br />

Tel.: +506 7 633176<br />

Fax: +506 7 633055<br />

E-mail: mguzman@corbana.co.cr<br />

Juliane Henderson<br />

CRCTPP<br />

Molecular Diversity<br />

et Diagnostics Research Laboratory<br />

80 Meiers Rd<br />

Indooroopilly Qld 4068<br />

Australia<br />

Tel.: +61 738 969341<br />

Fax: +61 738 969533<br />

E-mail:<br />

juliane.henderson@dpi.qld.gov.au<br />

Erwan Jade<br />

Total Fina Elf<br />

Special fluids<br />

51, Esplanade du Général de Gaulle<br />

La Défense 10<br />

92907 Paris La Défense Cedex<br />

France<br />

Tel.: +33 141356123<br />

Fax: +33 141355134<br />

Email: Erwan.Jade@TotalFinaElf.com<br />

Andrew James<br />

Unidad de Biotecnología, División Biología<br />

Vegetal<br />

CICY<br />

Calle 43 No. 130<br />

Colonia Churburná de Hidalgo<br />

CP 97200, Merida<br />

Mexico<br />

Tel.: +52 99 813914<br />

Fax: +52 99 813900<br />

E-mail: andyj007@cicy.mx<br />

Ramiro Jaramillo<br />

Honorary guest<br />

Former INIBAP regional coordinator<br />

Apdo 4824-1000<br />

San José<br />

Costa Rica<br />

Christophe Jenny<br />

CIRAD-FHLOR<br />

Station de Neufchâteau<br />

Sainte Marie<br />

F-97130 Capesterre Belle-Eau<br />

Guadeloupe<br />

Tel.: +33 590861768<br />

Fax: +33 590868077<br />

E-mail: christophe.jenny@cirad.fr<br />

Luis Jacome<br />

PATHOTEC, S.A.<br />

Local # 5 Interior, Hotel Lima<br />

La Lima, Cortes<br />

Honduras<br />

Tel.: +504 6685610<br />

Fax: +504 6685613<br />

E-mail: lhjacome@infovia.hn<br />

David Jones<br />

12 Charlotte Brontë Drive<br />

Droitwich Spa<br />

Worcestershire WR9 7HU<br />

United Kingdom<br />

Tel.: +44 190 4462098<br />

Fax: +44 190 4462250<br />

E-mail: bananadoctor@email.msn.com<br />

313


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Pedro E. Jorge<br />

IDIAF<br />

Rafael Augusto Sanchez #89<br />

Santo Domingo<br />

Dominican Republic<br />

Tel.: +1 809 5678999<br />

Fax: +1 809 5679199<br />

E-mail: pejorge@hotmail.com<br />

Dieter Kaemmer<br />

CICY<br />

Calle 43 No. 130<br />

Colonia Churburná de Hidalgo<br />

CP 97200<br />

Mexico<br />

Tel.: +52 99 813914<br />

Fax.: +52 99 813900<br />

E-mail: dieter@cicy.mx<br />

Gert H.J. Kema<br />

IPO-DLO<br />

P.O.Box 9060<br />

6700 GW Wageningen<br />

The Netherlands<br />

Tel.: +31 317 476149<br />

Fax: +31 317 410113<br />

E-mail: G.H.J.Kema@IPO.DLO.NL<br />

Susan Knight<br />

Syngenta<br />

Centro de Ciencia y Tecnología Ultrapark<br />

Edificio 7B Segundo Piso<br />

La Aurora de Heredia<br />

Heredia<br />

Costa Rica<br />

Tel.: +506 2 939500<br />

Fax: +506 2 931628<br />

E-mail: susan.knight@syngenta.com<br />

Marc Henri Lebrun<br />

UMR 1932 CNRS INRA<br />

Aventis Cropscience<br />

14-20 rue Pierre Baizet<br />

B.P. 9163<br />

69263 Lyon, Cedex 09<br />

France<br />

Tel.: +33 472852481<br />

Fax: +33 472852297<br />

E-mail: marc-henri.lebrun@aventis.com<br />

Philippe Lepoivre<br />

Unité de Phytopathologie<br />

University <strong>of</strong> Gembloux<br />

Passage des Déportés, 2<br />

B-5030 Gembloux<br />

Belgium<br />

Tel.: +32 81 622437<br />

Fax: +32 81 610126<br />

E-mail: lepoivre.p@fsagx.ac.be<br />

Ronald Madrigal Barrantes<br />

EARTH University<br />

Las Mercedes<br />

Guácimo<br />

Limón<br />

Costa Rica<br />

Tel.: +506 7 130000<br />

Fax: +506 7 130005<br />

E-mail: rmadriga@ns.earth.ac.cr<br />

Douglas Marin<br />

Delmonte<br />

Apdo 4084-1000<br />

San José<br />

Costa Rica<br />

Tel.: +506 7103674<br />

Fax: +506 7103627<br />

E-mail: marin.douglas@freshdelmonte.co.cr<br />

Sergia Milagrosa<br />

Del Monte Fresh<br />

Davao City<br />

Philippines<br />

Tel.: +63 82 2331838<br />

Fax: +63 82 2340438<br />

E-mail: SMilagrosa@freshdelmonte.com.ph<br />

Gus Molina<br />

INIBAP<br />

Asia-Pacific Regional Office<br />

c/o IRRI Rm 31<br />

Khush Hall<br />

Los Baños, Laguna 4031<br />

Philippines<br />

Tel.: +63 2 8450563<br />

Fax: +63 2 8911292<br />

E-mail: a.molina@cgiar.org<br />

314


List <strong>of</strong> participants<br />

Xavier Mourichon<br />

CIRAD-AMIS<br />

Av. Agropolis TA 40/02<br />

34398 Montpellier Cedex 5<br />

France<br />

Tel.: +33 467615869<br />

Fax: +33 467615581<br />

E-mail: xavier.mourichon@cirad.fr<br />

Juan Luis Ortiz<br />

CATIE<br />

7170 Turrialba<br />

Cartago<br />

Costa Rica<br />

Tel.: +506 5 566455<br />

E-mail: jortiz@catie.ac.cr<br />

Rodomiro Ortiz<br />

IITA<br />

Carolyn House 26<br />

Dingwall Road<br />

Croydon CR9 3EE<br />

Nigeria<br />

Tel.: +234 2 2412626<br />

Fax: +234 2 2412221<br />

E-mail: r.ortiz@cgiar.org<br />

Luis Fernando Patiño<br />

AUGURA<br />

CENIBANANO<br />

Calle 3 Sur No. 41-65, Edif. Banco de Occidente<br />

Medellin<br />

Colombia<br />

Tel.: +57 4 3211333<br />

Fax: +57 4 8236606<br />

E-mail: lpatino@augura.com.co<br />

Leticia Peraza<br />

CICY<br />

Calle 43 No. 130<br />

Colonia Churburná de Hidalgo<br />

CP 97200<br />

Mexico<br />

Tel.: +52 99 813966<br />

Fax: +52 99 813900<br />

E-mail: latyperaza@yahoo.com<br />

Luis Pérez Vicente<br />

INISAV<br />

Gaveta 634<br />

11300, Playa<br />

Ciudad Habana<br />

Cuba<br />

Tel.: +537 8 782 420<br />

Fax: +537 2 405 35<br />

E-mail: lperez@inisav.cu<br />

Claudine Picq<br />

INIBAP<br />

Parc Scientifique Agropolis II<br />

Montpellier 34397 Cedex 5<br />

France<br />

Tel.: +33 467611302<br />

Fax: +33 467610334<br />

E-mail: c.picq@cgiar.org<br />

Luis Pocasangre<br />

INIBAP<br />

Latin America and Caribbean Regional Office<br />

C/o CATIE<br />

7170 Turrialba<br />

Costa Rica<br />

Tel.: +506 5562431<br />

Fax: +506 5562431<br />

E-mail: lpoca@catie.ac.cr<br />

Tania Polanco<br />

IDIAF<br />

Rafael Augusto Sanchez #89<br />

Santo Domingo<br />

Dominican Republic<br />

Tel.: +809 5678999<br />

Fax: +809 5679199<br />

E-mail: tpolanco@hotmail.com<br />

Vivencio L. Quiñon<br />

TADECO<br />

A.O. Floriendo<br />

Papabo City<br />

Davao<br />

Philippines<br />

Tel.: +63 84 8220541<br />

Edwin Raros<br />

Dole Asia Research<br />

c/o Stanfilco<br />

Davao City<br />

Philippines<br />

315


<strong>Mycosphaerella</strong> <strong>leaf</strong> <strong>spot</strong> <strong>diseases</strong> <strong>of</strong> <strong>bananas</strong>: present status and outlook<br />

Galileo Rivas<br />

CATIE<br />

7170 Turrialba<br />

Costa Rica<br />

Tel.: +506 5560232<br />

Fax: +506 5566480<br />

E-mail: grivas@catie.ac.cr<br />

Mauricio Rivera Canales<br />

FHIA<br />

Apdo Postal 2067<br />

San Pedro Sula<br />

Honduras<br />

Tel.: +504 6682470<br />

Fax: +504 6682313<br />

E-mail: mrivera@fhia.org.hn<br />

Alba Stella Riveros Angarita<br />

UTOLIMA - CATIE<br />

Unidad de Fitoprotección<br />

7170 Turrialba<br />

Costa Rica<br />

Tel.: +506 5566021<br />

Fax: +506 5560606<br />

E-mail: asrivero@catie.ac.cr<br />

Ronald Romero<br />

Chiquita<br />

Building D, FORUM<br />

Santa Ana<br />

Costa Rica<br />

Tel.: +506 2 042 001<br />

Fax: +506 2 042 399<br />

E-mail: rromero@chiquita.com<br />

Franklin Rosales<br />

INIBAP<br />

Latin America and Caribbean<br />

Regional Office<br />

C/o CATIE<br />

7170 Turrialba<br />

Costa Rica<br />

Tel.: +506 5562431<br />

Fax: +506 5562431<br />

E-mail: frosales@catie.ac.cr<br />

Nicolas Roux*<br />

IAEA Plant Breeding Unit<br />

A-2444 Seibersdorf<br />

Austria<br />

*Current address<br />

INIBAP<br />

Parc Scientifique Agropolis II<br />

34397 Montpellier Cedex 5<br />

France<br />

Tel.: +33 467611302<br />

Fax: +33 467610334<br />

E-mail: n.roux@cgiar.org<br />

Laszlo Sági<br />

KULeuven<br />

Kasteelpark Arenberg 13<br />

B-3001 Leuven<br />

Belgium<br />

Tel.: +32 16 321681<br />

Fax: +32 16 321993<br />

E-mail: laszlo.sagi@agr.kuleuven.ac.be<br />

Jorge Sandoval<br />

CORBANA<br />

Apdo 390-7210<br />

Guápiles, Limón<br />

Costa Rica<br />

Tel.: +506 7633176<br />

Fax: +506 7633055<br />

E-mail: jsandoval@corbana.co.cr<br />

Hans Sauter<br />

Edificio Del Monte<br />

200 Metros al Este del Periódico La Republica<br />

Barrio Tournon<br />

San Jose<br />

Costa Rica<br />

Fax: +506 222-9769<br />

E-mail:<br />

sauter.hans@freshdelmonte.co.cr<br />

R. Selvarajan<br />

NRCB<br />

44 Ramalinga Nagar<br />

Vayalur Road<br />

Trichy 620 017<br />

India<br />

Tel.: +91 431 771299<br />

Fax: +91 431 770564<br />

E-mail: selvarama@yahoo.com<br />

316


List <strong>of</strong> participants<br />

Moises Soto Ballestero<br />

Corporación S y M SA<br />

Apartado Postal 479-1007<br />

San José<br />

Costa Rica<br />

Tel.: +506 336670<br />

Fax: +506 2904941<br />

E-mail: agbenig@racsa.co.cr<br />

Egbert Spaans<br />

EARTH<br />

Apdo 4442-1000<br />

San Jose,<br />

Costa Rica<br />

Tel.: +506 7 130000<br />

Fax: +506 7 130001<br />

E-mail: espaans@earth.ac.cr<br />

Rony Swennen<br />

KULeuven<br />

Laboratory <strong>of</strong> Tropical Crop Improvement<br />

Kasteelpark Arenberg 13<br />

3001 Leuven<br />

Belgium<br />

Tel.: +32 16 321420<br />

Fax: +32 16 321993<br />

E-mail: Rony.Swennen@agr.kuleuven.ac.be<br />

Ana Cecilia Tapia<br />

CATIE<br />

Apdo 7170, Turrialba<br />

Costa Rica<br />

Tel.: +506 5560232<br />

Fax: +506 556 6480<br />

E-mail: tapiaa@catie.ac.cr<br />

Ronald Vargas<br />

CORBANA<br />

Diagonal a la Casa Presidencial<br />

A.A. 6504-1000<br />

San José<br />

Costa Rica<br />

Tel.: +506 7633176<br />

Fax: +506 7633055<br />

E-mail: rvargas@corbana.co.cr<br />

Altus Viljoen<br />

FABI<br />

University <strong>of</strong> Pretoria<br />

Pretoria 0002<br />

South Africa<br />

Tel.: +27 12 4203856<br />

Fax: +27 12 4203960<br />

E-mail: aviljoen@postino.up.ac.za<br />

Manuel Wirz<br />

Syngenta<br />

Centro de Ciencia y Tecnología Ultrapark<br />

Edificio 7B Segundo Piso<br />

La Aurora de Heredia<br />

Heredia<br />

Costa Rica<br />

Tel.: +506 2939500<br />

Fax: +506 2931628<br />

E-mail: manuel.wirz@syngenta.com<br />

317


isbn 2-910810-57-7

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!