24.12.2012 Views

Septoria and Stagonospora Diseases of Cereals - CIMMYT ...

Septoria and Stagonospora Diseases of Cereals - CIMMYT ...

Septoria and Stagonospora Diseases of Cereals - CIMMYT ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

<strong>Septoria</strong> <strong>and</strong><br />

<strong>Stagonospora</strong><br />

<strong>Diseases</strong> <strong>of</strong> <strong>Cereals</strong>:<br />

A Compilation <strong>of</strong> Global Research<br />

M. van Ginkel,<br />

A. McNab,<br />

<strong>and</strong> J. Krupinsky,<br />

editors


<strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong><br />

<strong>Diseases</strong> <strong>of</strong> <strong>Cereals</strong>:<br />

A Compilation<br />

<strong>of</strong> Global Research<br />

Proceedings <strong>of</strong> the Fifth<br />

International <strong>Septoria</strong> Workshop<br />

September 20-24, 1999<br />

<strong>CIMMYT</strong>, Mexico<br />

M. van Ginkel, A. McNab, <strong>and</strong> J. Krupinsky, editors<br />

Dedicated to the memory <strong>of</strong><br />

Dr. Zahir Eyal


ii<br />

The Organizing Committee expresses it sincere thanks to the Workshop Sponsors:<br />

Bayer de México, S.A., <strong>and</strong> Zeneca Mexicana, S.A.<br />

<strong>CIMMYT</strong> (www.cimmyt.mx or www.cimmyt.cgiar.org) is an internationally funded, nonpr<strong>of</strong>it<br />

scientific research <strong>and</strong> training organization. Headquartered in Mexico, the Center works with<br />

agricultural research institutions worldwide to improve the productivity, pr<strong>of</strong>itability, <strong>and</strong><br />

sustainability <strong>of</strong> maize <strong>and</strong> wheat systems for poor farmers in developing countries. It is one <strong>of</strong> 16<br />

similar centers supported by the Consultative Group on International Agricultural Research<br />

(CGIAR). The CGIAR comprises over 55 partner countries, international <strong>and</strong> regional<br />

organizations, <strong>and</strong> private foundations. It is co-sponsored by the Food <strong>and</strong> Agriculture<br />

Organization (FAO) <strong>of</strong> the United Nations, the International Bank for Reconstruction <strong>and</strong><br />

Development (World Bank), the United Nations Development Programme (UNDP), <strong>and</strong> the<br />

United Nations Environment Programme (UNEP). Financial support for <strong>CIMMYT</strong>’s research<br />

agenda also comes from many other sources, including foundations, development banks, <strong>and</strong><br />

public <strong>and</strong> private agencies.<br />

<strong>CIMMYT</strong> supports Future Harvest, a public awareness campaign that builds underst<strong>and</strong>ing about<br />

the importance <strong>of</strong> agricultural issues <strong>and</strong> international agricultural research. Future Harvest links<br />

respected research institutions, influential public figures, <strong>and</strong> leading agricultural scientists to<br />

underscore the wider social benefits <strong>of</strong> improved agriculture—peace, prosperity, environmental<br />

renewal, health, <strong>and</strong> the alleviation <strong>of</strong> human suffering (www.futureharvest.org).<br />

© International Maize <strong>and</strong> Wheat Improvement Center (<strong>CIMMYT</strong>) 1999. Responsibility for this<br />

publication rests solely with <strong>CIMMYT</strong>. The designations employed in the presentation <strong>of</strong> material<br />

in this publication do not imply the expressions <strong>of</strong> any opinion whatsoever on the part <strong>of</strong><br />

<strong>CIMMYT</strong> or contributory organizations concerning the legal status <strong>of</strong> any country, territory, city, or<br />

area, or <strong>of</strong> its authorities, or concerning the delimitation <strong>of</strong> its frontiers or boundaries.<br />

Printed in Mexico.<br />

Correct citation: van Ginkel, M., A. McNab, <strong>and</strong> J. Krupinsky, eds. 1999. <strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong><br />

<strong>Diseases</strong> <strong>of</strong> <strong>Cereals</strong>: A Compilation <strong>of</strong> Global Research. Mexico, D.F.: <strong>CIMMYT</strong>.<br />

ISBN: 970-648-035-8<br />

AGROVOC descriptors: Wheats; Triticum; Triticum aestivum; S<strong>of</strong>t wheat; Triticum durum; Hard<br />

wheat; Winter crops; Plant diseases; Fungal diseases; <strong>Septoria</strong>; <strong>Stagonospora</strong>; Blotches;<br />

Mycosphaerella; Epidemiology; Plant breeding; Selection; Disease resistance; Genetic control; Gene<br />

location; Cultural control; Plant response; Research projects<br />

Additional Keywords: Triticum tauschii<br />

AGRIS category codes: H20 Plant <strong>Diseases</strong><br />

F30 Plant Genetics <strong>and</strong> Breeding<br />

Dewey decimal classification: 632.4<br />

Additional information on <strong>CIMMYT</strong> is available on the WorldWideWeb at: www.cimmyt.cgiar.org.


Table <strong>of</strong> Contents<br />

vi In Memoriam, Dr. Zahir Eyal<br />

vii Foreword<br />

1 Opening remarks<br />

1 Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program<br />

S. Rajaram<br />

19 Session 1: Pathogen Biology<br />

19 Biology <strong>of</strong> the <strong>Septoria</strong>/<strong>Stagonospora</strong> Pathogens: An Overview<br />

A.L. Scharen<br />

23 Molecular Analysis <strong>of</strong> a DNA Fingerprint Probe from Mycosphaerella graminicola<br />

S.B. Goodwin <strong>and</strong> J.R. Cavaletto<br />

26 Characterization <strong>of</strong> <strong>Septoria</strong> tritici Variants <strong>and</strong> PCR Assay for Detecting <strong>Stagonospora</strong> nodorum <strong>and</strong><br />

<strong>Septoria</strong> tritici in Wheat<br />

S. Hamza, M. Medini, T. Sassi, S. Abdennour, M. Rouassi, A.B. Salah, M. Cherif, R. Strange, <strong>and</strong> M.<br />

Harrabi<br />

32 Populations <strong>of</strong> <strong>Septoria</strong> spp. Affecting Winter Wheat in the Forest-Steppe Zone <strong>of</strong> the Ukraine<br />

S. Kolomiets<br />

34 <strong>Septoria</strong> passerinii Closely Related to the Wheat Pathogen Mycosphaerella graminicola<br />

S.B. Goodwin <strong>and</strong> V.L. Zismann<br />

37 <strong>Septoria</strong>/<strong>Stagonospora</strong> Leaf Spot <strong>Diseases</strong> on Barley in North Dakota, USA<br />

J.M. Krupinsky <strong>and</strong> B.J. Steffenson (poster)<br />

39 Interrelations among <strong>Septoria</strong> tritici Isolates <strong>of</strong> Varying Virulence<br />

S. Ezrati, S. Schuster, A. Eshel, <strong>and</strong> Z. Eyal (poster)<br />

41 Session 2: The Infection Process<br />

41 <strong>Stagonospora</strong> <strong>and</strong> <strong>Septoria</strong> Pathogens <strong>of</strong> <strong>Cereals</strong>: The Infection Process<br />

B.M. Cunfer<br />

46 Aggressiveness <strong>of</strong> Phaeosphaeria nodorum Isolates <strong>and</strong> Their In Vitro Secretion <strong>of</strong> Cell-Wall-<br />

Degrading Enzymes<br />

P. Halama, F. Lalaoui, V. Dumortier, <strong>and</strong> B. Paul<br />

50 Growth <strong>of</strong> <strong>Stagonospora</strong> nodorum Lesions<br />

A.M. Djurle (poster)<br />

51 Session 3A: Host-Parasite Interactions<br />

51 Genetic Control <strong>of</strong> Avirulence in Mycosphaerella graminicola (Anamorph <strong>Septoria</strong> tritici)<br />

G.H.J. Kema <strong>and</strong> E.C.P. Verstappen<br />

53 Cytogenetics <strong>of</strong> Resistance <strong>of</strong> Wheat to <strong>Septoria</strong> Tritici Leaf Blotch<br />

L.S. Arraiano, A.J. Worl<strong>and</strong>, <strong>and</strong> J.K.M. Brown<br />

54 A Possible Gene-for-Gene Relationship for <strong>Septoria</strong> Tritici Leaf Blotch Resistance in Wheat<br />

P.A. Brading, G.H.J. Kema, <strong>and</strong> J.K.M. Brown<br />

56 Diallel Analysis <strong>of</strong> <strong>Septoria</strong> Tritici Blotch Resistance in Winter Wheat<br />

X. Zhang, S.D. Haley, <strong>and</strong> Y. Jin<br />

59 Analysis <strong>of</strong> the <strong>Septoria</strong> Monitoring Nursery<br />

L. Gilchrist, C. Velazquez, <strong>and</strong> J. Crossa<br />

63 Session 3B: Host Parasite Interactions<br />

63 Host – Parasite Interactions: <strong>Stagonospora</strong> nodorum<br />

E. Arseniuk <strong>and</strong> P.C. Czembor<br />

71 Identification <strong>of</strong> a Molecular Marker Linked to <strong>Septoria</strong> Nodorum Blotch Resistance in Triticum<br />

tauschii Using F2 Bulked Segregant<br />

N.E.A. Murphy, R. Loughman, R. Wilson, E.S. Lagudah, R. Appels, <strong>and</strong> M.G.K. Jones<br />

74 Inheritance <strong>of</strong> <strong>Septoria</strong> Nodorum Blotch Resistance in a Triticum tauschii Accession Controlled by a<br />

Single Gene<br />

N.E.A. Murphy, R. Loughman, R. Wilson, E.S. Lagudah, R. Appels, <strong>and</strong> M.G.K. Jones<br />

77 Session 4: Population Dynamics<br />

77 Population Genetics <strong>of</strong> Mycosphaerella graminicola <strong>and</strong> Phaeosphaeria nodorum<br />

B.A. McDonald, C.C. Mundt, <strong>and</strong> J. Zhan<br />

83 Characterization <strong>of</strong> Less Aggressive <strong>Stagonospora</strong> nodorum Isolates from Wheat<br />

E. Arseniuk, H.S. Tsang, J.M. Krupinsky, <strong>and</strong> P.P. Ueng<br />

85 A Vertically Resistant Wheat Selects for Specifically Adapted Mycosphaerella graminicola Strains<br />

C. Cowger, C.C. Mundt, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

iii


iv<br />

87 Genetic Variability in a Collection <strong>of</strong> <strong>Stagonospora</strong> nodorum Isolates from Western Australia<br />

N.E.A. Murphy, R. Loughman, E.S. Lagudah, R. Appels, <strong>and</strong> M.G.K. Jones<br />

90 Mating Type-Specific PCR Primers for <strong>Stagonospora</strong> nodorum Field Studies<br />

R.S. Bennett, S.-H. Yun, T.Y. Lee, B.G. Turgeon, B. Cunfer, E. Arseniuk, <strong>and</strong> G.C. Bergstrom (poster)<br />

93 Session 5: Epidemiology<br />

93 Epidemiology <strong>of</strong> Mycosphaerella graminicola <strong>and</strong> Phaeosphaeria nodorum: An Overview<br />

M.W. Shaw<br />

98 Spore Dispersal <strong>of</strong> Leaf Blotch Pathogens <strong>of</strong> Wheat (Mycosphaerella graminicola <strong>and</strong> <strong>Septoria</strong> tritici)<br />

C.A. Cordo, M.R. Simón, A.E. Perelló, <strong>and</strong> H.E. Alippi<br />

102 Epidemiology <strong>of</strong> Seedborne <strong>Stagonospora</strong> nodorum: A Case Study on New York Winter Wheat<br />

D.A. Shah <strong>and</strong> G.C. Bergstrom<br />

108 Sessions 6A <strong>and</strong> 6B: Cultural Practices <strong>and</strong> Disease Management<br />

108 Influence <strong>of</strong> Cultural Practices on <strong>Septoria</strong>/<strong>Stagonospora</strong> <strong>Diseases</strong><br />

J.M. Krupinsky<br />

111 Disease Management Using Varietal Mixtures<br />

C.C. Mundt, C. Cowger, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

117 Session 6C: Breeding for Disease Resistance<br />

117 Breeding for Resistance to the <strong>Septoria</strong>/<strong>Stagonospora</strong> Blights <strong>of</strong> Wheat<br />

M. van Ginkel <strong>and</strong> S. Rajaram<br />

127 Breeding for Resistance to <strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong> Blotches in Winter Wheat in the United States<br />

G. Shaner<br />

131 <strong>Septoria</strong> tritici Resistance <strong>of</strong> Wheat Cultivars at Different Growth Stages<br />

M. Díaz de Ackermann, M.M. Kohli, <strong>and</strong> V. Ibañez<br />

134 <strong>Septoria</strong> tritici Resistance Sources <strong>and</strong> Breeding Progress at <strong>CIMMYT</strong>, 1970-99<br />

L. Gilchrist, B. Gomez, R. Gonzalez, S. Fuentes, A. Mujeeb-Kazi, W. Pfeiffer, S. Rajaram, R.<br />

Rodriguez, B. Skovm<strong>and</strong>, M. van Ginkel, <strong>and</strong> C. Velazquez (Field presentation)<br />

140 Selecting Wheat for Resistance to <strong>Septoria</strong>/<strong>Stagonospora</strong> in Patzcuaro, Michoacan, Mexico<br />

R.M. Gonzalez I., S. Rajaram, <strong>and</strong> M. van Ginkel<br />

145 Varieties <strong>and</strong> Advanced Lines Resistant to <strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat in Western Australia<br />

R. Loughman, R.E. Wilson, I.M. Goss, D.T. Foster, <strong>and</strong> N.E.A. Murphy<br />

148 Field Resistance <strong>of</strong> Wheat to <strong>Septoria</strong> Tritici Leaf Blotch, <strong>and</strong> Interactions with Mycosphaerella<br />

graminicola Isolates<br />

J.K.M. Brown, G.H.J. Kema, H.-R. Forrer, E.C.P. Verstappen, L.S. Arraiano, P.A. Brading, E.M. Foster,<br />

A. Hecker, <strong>and</strong> E. Jenny<br />

150 Using Precise Genetic Stocks to Investigate the Control <strong>of</strong> <strong>Stagonospora</strong> nodorum Resistance in Wheat<br />

C.M. Ellerbrook, V. Korzun, <strong>and</strong> A.J. Worl<strong>and</strong> (poster)<br />

154 Evaluating Triticum durum x Triticum tauschii Germplasm for Resistance to <strong>Stagonospora</strong> nodorum<br />

L.R. Nelson <strong>and</strong> M.E. Sorrells (poster)<br />

156 Sources <strong>of</strong> Resistance to <strong>Septoria</strong> passerinii in Hordeum vulgare <strong>and</strong> H. vulgare subsp. spontaneum<br />

H. Toubia-Rahme <strong>and</strong> B.J. Steffenson (poster)<br />

159 S<strong>of</strong>t Red Winter Wheat with Resistance to <strong>Stagonospora</strong> nodorum <strong>and</strong> Other Foliar Pathogens<br />

B.M. Cunfer <strong>and</strong> J.W. Johnson (poster)<br />

160 Partial Resistance to <strong>Stagonospora</strong> nodorum in Wheat<br />

C.G. Du, L.R. Nelson, <strong>and</strong> M.E. McDaniel (poster)<br />

163 Comparison <strong>of</strong> Methods <strong>of</strong> Screening for <strong>Stagonospora</strong> nodorum Resistance in Winter Wheat<br />

D.E. Fraser, J.P. Murphy, <strong>and</strong> S. Leath (poster)<br />

167 Response <strong>of</strong> Winter Wheat Genotypes to Artificial Inoculation with Several <strong>Septoria</strong> tritici<br />

Populations<br />

M. Mincu (poster)<br />

170 Comparison <strong>of</strong> Greenhouse <strong>and</strong> Field Levels <strong>of</strong> Resistance to <strong>Stagonospora</strong> nodorum<br />

S.L. Walker, S. Leath, <strong>and</strong> J.P. Murphy (poster)<br />

173 Session 6D: Chemical Control<br />

173 Adjusting Thresholds for <strong>Septoria</strong> Control in Winter Wheat Using Strobilurins<br />

L.N. Jørgensen, K.E. Henriksen, <strong>and</strong> G.C. Nielsen<br />

177 Concluding Remarks<br />

177 The <strong>Septoria</strong>/<strong>Stagonospora</strong> Blotch <strong>Diseases</strong> <strong>of</strong> Wheat: Past, Present, <strong>and</strong> Future<br />

Z. Eyal (paper presented by A.L. Scharen)<br />

183 List <strong>of</strong> Participants


In Memoriam, Dr. Zahir Eyal<br />

Our friend <strong>and</strong> colleague, Pr<strong>of</strong>essor Zahir Eyal, died Friday, July 30, 1999. Zahir was<br />

intimately involved in all <strong>of</strong> the International <strong>Septoria</strong> Workshops, from the first in 1976 held in<br />

Griffin, Georgia, USA, until the fifth, <strong>and</strong> present, one held in <strong>CIMMYT</strong>, Mexico. He put forward<br />

his many ideas for program <strong>and</strong> participants in a forceful, but thoughtful way, <strong>and</strong> was able to<br />

settle disputes with good humor <strong>and</strong> a smile. Until the last few days <strong>of</strong> his life, Dr. Eyal<br />

continued to work on plans for this <strong>Septoria</strong>/<strong>Stagonospora</strong> workshop.<br />

After finishing agricultural high school in Israel, Zahir went to the USA, where he earned his<br />

B.Sc. degree from Oklahoma State University <strong>and</strong> his Ph.D. from Rutgers. This was followed by a<br />

post-doctoral term at Purdue, where he studied with Jack Schafer <strong>and</strong> the late Ralph Caldwell.<br />

His work on septoria <strong>of</strong> cereals began when he joined the Department <strong>of</strong> Botany at Tel Aviv<br />

University in 1967. He developed an integrated program <strong>of</strong> fundamental <strong>and</strong> applied research<br />

aimed at minimizing the economic impact <strong>of</strong> cereal pathogens, particularly <strong>Septoria</strong> tritici, on<br />

production. He reached out to colleagues in many countries <strong>and</strong> to international organizations,<br />

most especially <strong>CIMMYT</strong>, for cooperation. Over the years, Zahir contributed greatly to those<br />

programs. During his tenure at Tel Aviv University, Pr<strong>of</strong>essor Eyal served two terms as Head <strong>of</strong><br />

the Department <strong>of</strong> Botany (1984-87 <strong>and</strong> 1992-94).<br />

Pr<strong>of</strong>essor Eyal was an enthusiastic teacher, well-loved by students, both undergraduate <strong>and</strong><br />

graduate. He taught in English or in Hebrew with equal facility, sharing his knowledge <strong>and</strong><br />

insights with students <strong>and</strong> faculty during two sabbaticals at Montana State University <strong>and</strong> at<br />

several other institutions. The numerous publications he authored with his students <strong>and</strong> the<br />

important posts those students occupy today attest to the excellence <strong>of</strong> his teaching abilities.<br />

Zahir’s research <strong>and</strong> outreach programs incorporated ideas that were new to his country;<br />

they were solidly anchored in basic science <strong>and</strong> innovative to the end. These programs not only<br />

improved wheat production in Israel but had positive effects on cereal improvement programs<br />

throughout the world. At the time <strong>of</strong> his death, Pr<strong>of</strong>essor Eyal was Director <strong>of</strong> the Institute for<br />

Cereal Crop Improvement at Tel Aviv University, where germplasm <strong>of</strong> wild ancestors <strong>of</strong><br />

cultivated small grains are being preserved, characterized, <strong>and</strong> utilized in breeding improved<br />

cultivars.<br />

Dr. Eyal’s contributions to research, teaching, university administration, <strong>and</strong> international<br />

agriculture are many <strong>and</strong> far reaching. He received the Hazera Seed Co. Melamed Award in 1968,<br />

the A.C. Cohen Award in 1978, <strong>and</strong> in 1995 was made a Fellow <strong>of</strong> the American<br />

Phytopathological Society. Pr<strong>of</strong>essor Eyal served as President <strong>of</strong> the Israeli Phytopathological<br />

Society from 1979 to 1982. He will be fondly remembered <strong>and</strong> sadly missed by his multitude <strong>of</strong><br />

friends, colleagues, <strong>and</strong> students throughout the world.<br />

v


vi<br />

Foreword<br />

In the mid-1970s the idea <strong>of</strong> holding a septoria workshop began to take hold among a small<br />

group <strong>of</strong> scientists in the USA. They were interested in exchanging ideas <strong>and</strong> finding ways to<br />

manage the septoria diseases that affect wheat <strong>and</strong> other cereals all over the world. The first<br />

workshop was organized in a matter <strong>of</strong> a few months <strong>and</strong> held in Griffin, Georgia, in 1976.<br />

Among the 50 scientists who attended were a few researchers from outside the US. The<br />

enthusiasm <strong>of</strong> that first workshop led to the development <strong>of</strong> the second, which was a truly<br />

international meeting attended by more than 100 scientists from many countries, held in<br />

Bozeman, Montana, in 1983.<br />

Since then, international septoria workshops have been held about every five years: in<br />

Zurich, Switzerl<strong>and</strong>, in 1989; in Radzikow, Pol<strong>and</strong>, in 1994; <strong>and</strong> this year at <strong>CIMMYT</strong> in El<br />

Batan, Mexico. Each workshop has exp<strong>and</strong>ed the network <strong>of</strong> scientists who share their<br />

knowledge <strong>and</strong> pose the many questions that remain to be solved about these diseases <strong>and</strong> their<br />

management.<br />

The Zurich workshop had increased participation by workers from Europe <strong>and</strong> Africa. The<br />

Radzikow workshop brought increased participation from scientists in eastern Europe. The early<br />

workshops focused on the biology <strong>of</strong> the pathogens <strong>and</strong> breeding strategies, subjects in which<br />

there remain many unanswered questions. The 1994 workshop <strong>and</strong> the current one emphasize<br />

molecular approaches to the genetics <strong>of</strong> the pathogens.<br />

The Fifth International Workshop provides another opportunity to focus on the <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases, but also to see them in the context <strong>of</strong> the worldwide programs <strong>of</strong><br />

<strong>CIMMYT</strong>, which emphasize collaboration with developing countries with the aim <strong>of</strong> developing<br />

stable high yielding wheat varieties that possess durable resistance to the diseases.<br />

This workshop also gives us the opportunity to remember our friend <strong>and</strong> colleague, Zahir<br />

Eyal, who passed away not long ago. An integral part <strong>of</strong> the program development process <strong>and</strong><br />

the discussions at each workshop, he organized the scientific program for this workshop as well.<br />

Zahir Eyal was an enthusiastic supporter <strong>of</strong> the septoria workshops <strong>and</strong> the international<br />

exchange <strong>of</strong> ideas. He will be missed.<br />

We would like to express our appreciation for the efforts <strong>of</strong> Ravi Singh, Maarten van Ginkel,<br />

<strong>and</strong> Linda Ainsworth, who organized the workshop. We wish to thank Diana Godínez, María<br />

Luisa Varela, <strong>and</strong> Laura Rodríguez for managing the logistical support. We also recognize the<br />

efforts <strong>of</strong> Arnoldo Amaya, María Garay, Lucy Gilchrist, Monique Henry, Gilberto Hernández,<br />

Reynaldo Villareal, Juan José Joven, Marcelo Ortíz, Eliot Sánchez, Kelly Cassaday, Miguel<br />

Mellado, Wenceslao Almazán, <strong>and</strong> Antonio Luna, as well as many other members <strong>of</strong> <strong>CIMMYT</strong><br />

staff who contributed to the success <strong>of</strong> this event.<br />

The International Organizing Committee<br />

September 21, 1999


Opening Remarks<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an<br />

International Wheat Program<br />

S. Rajaram<br />

Wheat Program, <strong>CIMMYT</strong>, El Batan, Mexico<br />

I am immensely honored <strong>and</strong> grateful to the organizing committee <strong>of</strong> the 5 th International <strong>Septoria</strong> Workshop for asking<br />

me to deliver this lecture in the opening session. Even though my presentation is very broad <strong>and</strong> covers many issues, I assure<br />

you that I have been involved in breeding for resistance to septoria tritici blotch for at least 25 years, with some remarkable<br />

success. In this attempt, I would like to recognize the contribution <strong>of</strong> Pr<strong>of</strong>. Zahir Eyal, who served as a consultant on septoria<br />

issues to <strong>CIMMYT</strong> in the 1970s <strong>and</strong> 1980s. Indeed, he <strong>and</strong> I have some common intellectual roots through Pr<strong>of</strong>. Ralph<br />

Caldwell <strong>of</strong> Purdue University. Pr<strong>of</strong>. Eyal’s untimely death <strong>and</strong> departure from the scientific community is a loss to us all. I<br />

dedicate this opening lecture to him.<br />

Wheat is the most widely<br />

grown <strong>and</strong> consumed food crop in<br />

the world. It is the staple food <strong>of</strong><br />

nearly 35% <strong>of</strong> the world<br />

population, <strong>and</strong> dem<strong>and</strong> for wheat<br />

will grow faster than for any other<br />

major crop. The forecasted global<br />

dem<strong>and</strong> for wheat in the year 2020<br />

varies between 840 (Rosegrant et<br />

al., 1995) to 1050 million tons<br />

(Kronstad, 1998). To reach this<br />

target, global production will need<br />

to increase 1.6 to 2.6% annually<br />

from the present production level<br />

<strong>of</strong> 560 million tons. Increases in<br />

realized grain yield have provided<br />

about 90% <strong>of</strong> the growth in world<br />

cereal production since 1950<br />

(Mitchell et al., 1997) <strong>and</strong> by the<br />

first decade <strong>of</strong> the next century<br />

most <strong>of</strong> the increase needed in<br />

world food production must come<br />

from higher absolute yields<br />

(Ruttan, 1993). For wheat, the<br />

global average grain yield must<br />

increase from the current 2.5 t/ha<br />

to 3.8 t/ha. In 1995, only 18<br />

countries world worldwide had<br />

average wheat grain yields <strong>of</strong> more<br />

than 3.8 t/ha, the majority located<br />

in Northern Europe (<strong>CIMMYT</strong>,<br />

1996).<br />

The formidable challenge to<br />

meet this dem<strong>and</strong> is not new to<br />

agricultural scientists who have<br />

been involved in the development<br />

<strong>of</strong> improved wheat production<br />

technologies for the past half<br />

century. For all developing<br />

countries, wheat yields have grown<br />

at an average annual rate <strong>of</strong> over<br />

2% between 1961 <strong>and</strong> 1994<br />

(<strong>CIMMYT</strong>, 1996). In Western<br />

Europe <strong>and</strong> North America the<br />

annual rate <strong>of</strong> growth for wheat<br />

yield was 2.7% from 1977 to 1985,<br />

falling to 1.5% from 1986 to 1995.<br />

Recent data have indicated a<br />

decrease in the productivity gains<br />

being achieved by major wheat<br />

producing countries (Brown, 1997).<br />

In Western Europe, where the<br />

highest average wheat grain yield<br />

is obtained in the Netherl<strong>and</strong>s<br />

(8.6 t/ha), yield increased from 5 to<br />

6 t/ha in five years, but it took<br />

more than a decade to raise yields<br />

from 6 to 7 t/ha. Worldwide,<br />

annual wheat grain yield growth<br />

decreased from 3.0% between 1977-<br />

1985, to 1.6% from 1986-1995,<br />

excluding the USSR (<strong>CIMMYT</strong>,<br />

1996). Degradation <strong>of</strong> the l<strong>and</strong><br />

resource base, together with a<br />

slackening <strong>of</strong> research investment<br />

<strong>and</strong> infrastructure, have<br />

1<br />

contributed to this decrease<br />

(Pingali <strong>and</strong> Heisey, 1997). Whether<br />

production constraints are affected<br />

by physiological or genetic limits is<br />

hotly debated, but future increases<br />

in food productivity will require<br />

substantial research <strong>and</strong><br />

development investment to<br />

improve the pr<strong>of</strong>itability <strong>of</strong> wheat<br />

production systems through<br />

enhancing input efficiencies. Due to<br />

a continuing necessity for multidisciplinary<br />

team efforts in plant<br />

breeding, <strong>and</strong> the rapidly changing<br />

development <strong>of</strong> technologies, three<br />

overlapping avenues can be<br />

considered for raising the yield<br />

frontier in wheat: continued<br />

investments in “conventional<br />

breeding” methods; use <strong>of</strong> current<br />

<strong>and</strong> exp<strong>and</strong>ed genetic diversity;<br />

<strong>and</strong> investigation <strong>and</strong><br />

implementation <strong>of</strong> biotechnology<br />

assisted plant breeding.<br />

Conventional Wheat<br />

Breeding<br />

It is likely that gains to be<br />

achieved from conventional<br />

breeding will continue to be<br />

significant for the next two decades<br />

or more (Duvick, 1996), but these


2 Opening Remarks — S. Rajaram<br />

are likely to come at a higher<br />

research cost than in the past. In<br />

recent surveys <strong>of</strong> wheat breeders<br />

(Braun et al., 1998; Rejesus et al.,<br />

1996), more than 80% <strong>of</strong><br />

respondents expressed concern that<br />

plant variety protection (PVP) <strong>and</strong><br />

plant or gene patents will restrict<br />

access to germplasm. This may<br />

have deleterious consequences for<br />

future breeding success.<br />

Rasmusson (1996) has stated that<br />

nearly half <strong>of</strong> the progress made by<br />

breeders in the past can be<br />

attributed to germplasm exchange.<br />

Regional <strong>and</strong> international<br />

nurseries have been an efficient<br />

means <strong>of</strong> gathering data from<br />

varied environments <strong>and</strong> exposing<br />

germplasm to diverse pathogen<br />

selection pressures, while<br />

providing access <strong>and</strong> exchange <strong>of</strong><br />

germplasm. Breeders utilize these<br />

cooperative nurseries extensively<br />

in their crossing programs (Braun<br />

et al., 1998). However, the number<br />

<strong>of</strong> cooperatively distributed wheat<br />

yield <strong>and</strong> screening nurseries has<br />

been greatly reduced during the<br />

past decade.<br />

Investments needed for<br />

breeding efforts will increase with<br />

increasing yield levels. Further,<br />

progress to develop higher yielding<br />

cultivars is reduced with every<br />

objective added to a breeding<br />

program. Though the list <strong>of</strong><br />

important traits may get longer <strong>and</strong><br />

longer, little if any assistance has<br />

been provided by economists to<br />

prioritize breeding objectives.<br />

Considering that a wheat breeding<br />

program like <strong>CIMMYT</strong> allocates<br />

around 60% <strong>of</strong> its resources to<br />

durable resistance breeding, the<br />

need for research in this field is<br />

obvious. Due to high costs, we see<br />

durable resistance breeding as one<br />

<strong>of</strong> the first fields where<br />

transformation should be applied<br />

by breeders through introgression<br />

<strong>of</strong> one or more genes controlling<br />

disease resistance.<br />

Adoption <strong>of</strong> <strong>CIMMYT</strong>-<br />

Based Germplasm<br />

<strong>CIMMYT</strong>’s breeding<br />

methodology is tailored to develop<br />

widely adapted, disease resistant<br />

germplasm with high <strong>and</strong> stable<br />

yield across a wide range <strong>of</strong><br />

environments. The impact <strong>of</strong> this<br />

approach has been significant.<br />

The total spring bread wheat<br />

(Triticum aestivum L.) area in<br />

developing countries, excluding<br />

China, is around 63 million ha, <strong>of</strong><br />

which 36 million ha or 58% are<br />

planted to varieties derived from<br />

<strong>CIMMYT</strong> germplasm (Table 1)<br />

(Byerlee <strong>and</strong> Moya, 1993;<br />

Rajaram, 1995). During the 1966-90<br />

period, 1317 bread wheat cultivars<br />

were released by developing<br />

countries, <strong>of</strong> which 70% were either<br />

direct releases from <strong>CIMMYT</strong><br />

advanced lines or had at least one<br />

<strong>CIMMYT</strong> parent (Byerlee <strong>and</strong><br />

Moya, 1993). For the 1986-90<br />

period, 84% <strong>of</strong> all bread wheat<br />

cultivars released in developing<br />

countries had <strong>CIMMYT</strong> germplasm<br />

in their pedigrees. Simultaneously<br />

the use <strong>of</strong> dwarfing genes has<br />

continued to increase over time.<br />

Today, regardless <strong>of</strong> the type <strong>of</strong><br />

wheat, more than 90% <strong>of</strong> all wheat<br />

varieties released in developing<br />

countries are semidwarfs, which<br />

covered 70% <strong>of</strong> the total wheat area<br />

in developing countries by the end<br />

<strong>of</strong> 1990 (Byerlee <strong>and</strong> Moya, 1993).<br />

The continuous adoption <strong>of</strong><br />

semidwarf spring wheat cultivars<br />

in the post-Green Revolution<br />

period (1977-90) resulted in about<br />

15.5 million tons <strong>of</strong> additional<br />

wheat production in 1990, valued<br />

at about US$ 3 billion, <strong>of</strong> which<br />

50%, or US$ 1.5 billion, is attributed<br />

to the adoption <strong>of</strong> new Mexican<br />

semidwarf wheat cultivars (Byerlee<br />

<strong>and</strong> Moya, 1993). In 1990, an<br />

estimated 93% <strong>of</strong> the total spring<br />

bread wheat production in<br />

developing countries, excluding<br />

China, came from semidwarf<br />

spring wheats, which covered<br />

about 83% <strong>of</strong> the total spring bread<br />

wheat area in developing countries<br />

(Byerlee <strong>and</strong> Moya, 1993).<br />

Table 1. Origin <strong>of</strong> spring bread wheat varieties in<br />

developing countries.<br />

<strong>CIMMYT</strong> <strong>CIMMYT</strong><br />

NARS cross<br />

<strong>CIMMYT</strong> No<br />

cross parents ancestor <strong>CIMMYT</strong><br />

1966-90 45% 28% 3% 24% *<br />

1991-97 58% 30% 3% 9%<br />

* Estimated.<br />

Note: Excluding China. NARS=national agricultural<br />

research system.<br />

The cornerstone <strong>of</strong> <strong>CIMMYT</strong>’s<br />

breeding methodology is targeted<br />

breeding for the megaenvironments,<br />

the use <strong>of</strong> a diverse<br />

gene pool for crossing, shuttle<br />

breeding, selection for yield under<br />

optimum conditions, <strong>and</strong> multilocational<br />

testing to identify<br />

superior germplasm with good<br />

disease resistance. In this paper we<br />

would like to present some recent<br />

developments in <strong>CIMMYT</strong>’s Wheat<br />

Program.<br />

Targeted breeding: The<br />

mega-environment concept<br />

To address the needs <strong>of</strong> diverse<br />

wheat growing areas, <strong>CIMMYT</strong><br />

introduced in 1988 the concept <strong>of</strong><br />

mega-environments (ME) (Rajaram<br />

et al., 1994). Mega-environments<br />

are defined as a broad, not<br />

necessarily contiguous areas,<br />

occurring in more than one country<br />

<strong>and</strong> frequently transcontinental,<br />

defined by similar biotic <strong>and</strong><br />

abiotic stresses, cropping system<br />

requirements, consumer<br />

preferences, <strong>and</strong>, for convenience,


y volume <strong>of</strong> production.<br />

Germplasm generated for a given<br />

ME is useful throughout that<br />

environment, accommodating<br />

major stresses but perhaps not all<br />

the significant secondary stresses.<br />

Within an ME, millions <strong>of</strong> hectares<br />

are addressed with a certain degree<br />

<strong>of</strong> homogeneity as relates to wheat.<br />

By 1993, 12 ME had been defined, 6<br />

for spring wheats (ME1-ME6), 3 for<br />

facultative wheats (ME7-ME9), <strong>and</strong><br />

3 for winter wheats (ME9-ME12).<br />

Details <strong>of</strong> each ME are given in<br />

Table 2.<br />

Table 2. Classification <strong>of</strong> megaenvironments (MEs) used by the <strong>CIMMYT</strong> Wheat Program.<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 3<br />

Use <strong>of</strong> diverse genepools to<br />

maintain genetic diversity<br />

Recent surveys conducted by<br />

the <strong>CIMMYT</strong> Economics Program<br />

have found that 58% <strong>of</strong> all wheat<br />

varieties in developing countries<br />

derive from <strong>CIMMYT</strong> germplasm;<br />

this percentage rises to more than<br />

80%, if varieties with parents <strong>of</strong><br />

<strong>CIMMYT</strong> origin are also included<br />

(Table 1). This spectacular success<br />

puts an enormous burden on<br />

<strong>CIMMYT</strong> to continually diversify<br />

its germplasm base for resistance<br />

<strong>and</strong> stability parameters.<br />

Broad-based plant germplasm<br />

resources are imperative for a<br />

sound <strong>and</strong> successful breeding<br />

program. Utmost attention is given<br />

to the genetic diversity within<br />

<strong>CIMMYT</strong> germplasm to minimize<br />

the risk <strong>of</strong> genetic vulnerability,<br />

since it is grown on large areas <strong>and</strong><br />

is widely used by national<br />

programs. I believe that the use <strong>of</strong><br />

genetically diverse material is<br />

m<strong>and</strong>atory to increase yield<br />

potential <strong>and</strong> yield stability in the<br />

future. In any year 500-800 parental<br />

lines are considered for crossing.<br />

Major Representative<br />

Year<br />

breeding<br />

Area Moisture Temperature Growth breeding locations/ began at<br />

ME Latitudea (m ha) b regimec regimed habit Sowne objectivesf, g regions <strong>CIMMYT</strong><br />

SPRING WHEAT<br />

1 Low 32.0 Low rainfall, Temperate Spring A Resistance to Yaqui Valley, 1945<br />

irrigated lodging, SR, LR, YR Mexico; Indus<br />

Valley, Pakistan;<br />

Gangetic Valley,<br />

India; Nile Valley,<br />

Egypt<br />

2 Low 10.0 High rainfall Temperate Spring A As for ME1 + North African 1972<br />

resistance to Coast, Highl<strong>and</strong>s <strong>of</strong><br />

YR, <strong>Septoria</strong> East Africa, Andes,<br />

spp., sprouting <strong>and</strong> Mexico<br />

3 Low 1.7 High rainfall Temperate Spring A As for ME2 + Passo Fundo, 1974<br />

acid soil<br />

tolerance<br />

Brazil<br />

4A Low 10.0 Low rainfall, Temperate Spring A Resistance to Aleppo, Syria; 1974<br />

winter drought, Settat, Morocco<br />

dominant <strong>Septoria</strong> spp., YR<br />

4B Low 5.8 Low rainfall, Temperate Spring A Resistance to Marcos Juárez, 1974<br />

summer drought, Argentina<br />

dominant <strong>Septoria</strong> spp.,<br />

Fusarium spp.,<br />

LR, SR<br />

4C Low 5.8 Mostly Hot Spring A Resistance to Indore, India 1974<br />

residual drought, <strong>and</strong><br />

moisture heat in seedling<br />

stage<br />

5A Low 3.9 High rainfall Hot Spring A Resistance to Joydepur, 1981<br />

/ irrigated, heat, Helmin- Bangladesh;<br />

humid thosporium<br />

spp., Fusarium<br />

spp., sprouting<br />

Londrina, Brazil<br />

5B Low 3.2 Irrigated , Hot Spring A Resistance to Gezira, Sudan; 1975<br />

low humidity heat <strong>and</strong> SR Kano, Nigeria<br />

6 High 5.4 Moderate Temperate Spring S Resistance to Harbin, China 1989<br />

rainfall/ SR, LR, Helminsummer<br />

thosporium spp.,<br />

dominant Fusarium spp.,<br />

sprouting,<br />

photoperiod<br />

sensitivity<br />

(cont’d.)


4 Opening Remarks — S. Rajaram<br />

Table 2. Continued.<br />

Major Representative<br />

Year<br />

breeding<br />

Area Moisture Temperature Growth breeding locations/ began at<br />

ME Latitudea (m ha) b regimec regimed habit Sowne objectivesf, g regions <strong>CIMMYT</strong><br />

WINTER/FACULTATIVE WHEAT<br />

7 High – Irrigated Moderate Facultative A Rapid grain Zhenzhou, China 1986<br />

cold fill, resistance<br />

to cold, YR, PM,<br />

BYD<br />

8A High – High rainfall Moderate Facultative A Resistance to Chillan, Chile 1986<br />

/ irrigated, cold cold, YR, <strong>Septoria</strong><br />

long season spp.<br />

8B High – High rainfall Moderate Facultative A Resistance to Edirne, Turkey 1986<br />

/ irrigated, cold <strong>Septoria</strong> spp.,<br />

short season YR, PM, Fusarium<br />

spp., sprouting<br />

9 High – Low rainfall Moderate Facultative A Resistance to Diyarbakir, Turkey 1986<br />

cold cold, drought<br />

10 High – Irrigated Severe cold Winter A Resistance to<br />

winterkill, YR,<br />

LR, PM, BYD<br />

Beijing, China 1986<br />

11A High – High rainfall Moderate Winter A Resistance to Temuco, Chile 1986<br />

/irrigated, cold <strong>Septoria</strong> spp.,<br />

long season Fusarium spp.,<br />

YR, LR, PM<br />

11B High – High rainfall Severe cold Winter A Resistance to Lovrin, Romania 1986<br />

/ irrigated, LR, SR, PM,<br />

short season winterkill,<br />

sprouting<br />

12 High – Low rainfall Severe cold Winter A Resistance to<br />

winterkill,<br />

drought, YR,<br />

bunts<br />

Ankara, Turkey 1986<br />

Source: Adapted from Rajaram et. al. (1995).<br />

a Low = less than about 35-40 degrees.<br />

b Exact area distribution for winter/facultative wheat is not available.<br />

c Refers to rainfall just before <strong>and</strong> during the crop cycle. High = >500mm; low = 17.5°C; cold =


wheat to increase grain size. The<br />

six highest yielding lines derived<br />

from this program outyielded their<br />

breadwheat parent by 5-20% in<br />

yield trials in Cd. Obregon, Mexico.<br />

Shuttle breeding within<br />

Mexico<br />

Young <strong>and</strong> Frey (1994) provide<br />

two factors that influence the<br />

success <strong>of</strong> a shuttle program: a) the<br />

use <strong>of</strong> a germplasm pool<br />

encompassing genotypes with<br />

broad adaptation, <strong>and</strong> b) the use <strong>of</strong><br />

selection environments eliciting<br />

different responses from plant<br />

types. They also state that the<br />

wheat breeding program <strong>of</strong> N.E.<br />

Borlaug met these conditions.<br />

When Borlaug started the shuttle<br />

breeding approach in 1945, his only<br />

objective was to speed up breeding<br />

for stem rust resistance. Since then,<br />

segregating populations have been<br />

shuttled 100 times between the two<br />

environmentally contrasting sites<br />

in Mexico, Cd. Obregon <strong>and</strong> Toluca<br />

(Braun et al., 1992).<br />

Some <strong>of</strong> the salient points <strong>of</strong><br />

this shuttle breeding program are:<br />

• Cd. Obregon is situated at 28 o N<br />

at 40 masl, in the sunny, fertile,<br />

<strong>and</strong> irrigated Yaqui Valley <strong>of</strong><br />

Sonora. Wheats are planted in<br />

November when temperatures<br />

are low <strong>and</strong> harvested in April/<br />

May when temperatures are<br />

high. The yield potential <strong>of</strong><br />

location is high (±10 t/ha);<br />

wheat diseases are limited to<br />

only leaf rust, Karnal bunt, <strong>and</strong><br />

black point.<br />

• The Toluca location is<br />

characterized by high humidity<br />

(precipitation: ±1000 mm). The<br />

nursery is planted in May/June<br />

when temperatures are high <strong>and</strong><br />

harvested in October when they<br />

are low. High humidity causes<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 5<br />

incidence <strong>of</strong> many diseases<br />

including rust, septorias, BYD,<br />

<strong>and</strong> fusarium.<br />

An important result <strong>of</strong> shuttle<br />

breeding was the selection <strong>of</strong><br />

photo-insensitive wheat genotypes.<br />

Initially, selection for photoperiod<br />

insensitivity was unconscious, but<br />

only this trait permitted the wide<br />

spread <strong>of</strong> the Mexican semidwarfs<br />

(Borlaug, 1995). Today, this trait has<br />

been incorporated into basically all<br />

spring wheat cultivars grown<br />

below 48 o latitude <strong>and</strong> is now also<br />

spreading to wheat areas above 48 o<br />

N (Worl<strong>and</strong> et al., 1994).<br />

Multi-locational testing<br />

<strong>and</strong> wide adaptation<br />

About 1500 sets <strong>of</strong> yield trials<br />

<strong>and</strong> screening nurseries consisting<br />

<strong>of</strong> around 4000 advanced bread<br />

wheat lines are annually sent to<br />

more than 200 locations. Multilocational<br />

testing plays a key role in<br />

identifying the best performing<br />

entries for crossing. Since the<br />

shuttle program permits two full<br />

breeding cycles a year, it takes<br />

around five to six years from<br />

crossing to international<br />

distribution <strong>of</strong> advanced lines to<br />

cooperators. This “recurrent<br />

selection program” ensures<br />

continuous <strong>and</strong> rapid pyramiding<br />

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

Ceccarelli (1989) pointed out<br />

that the widespread cultivation <strong>of</strong><br />

some wheat cultivars should not be<br />

taken as a demonstration <strong>of</strong> wide<br />

adaptation, since a large fraction <strong>of</strong><br />

these areas are similar or made<br />

similar by use <strong>of</strong> irrigation <strong>and</strong>/or<br />

fertilizer. Therefore, the term wide<br />

adaptation has been used mainly to<br />

describe geographical rather than<br />

environmental differences. If this is<br />

true, the genotypic variation<br />

should be considerably higher than<br />

GxE interaction in ANOVAs <strong>of</strong><br />

<strong>CIMMYT</strong> trials. Braun et al. (1992)<br />

showed that this is not the case.<br />

When subsets <strong>of</strong> locations were<br />

grouped by geographical <strong>and</strong>/or<br />

environmental similarities, GxE<br />

interaction was mostly greater than<br />

the genotypic variance. The<br />

environmental diversity <strong>of</strong> sites<br />

where <strong>CIMMYT</strong>’s 21st<br />

International Bread Wheat<br />

Screening Nursery was grown <strong>and</strong><br />

the diversity among genotypes in<br />

this nursery were demonstrated by<br />

Bull et al. (1994). They classified<br />

similarities among environments<br />

by forming subsets <strong>of</strong> genotypes<br />

from the total dataset <strong>and</strong><br />

comparing them with the<br />

classification based on the<br />

remaining genotypes. Using this<br />

procedure they concluded that it<br />

was not possible to form a stable<br />

grouping <strong>of</strong> environments, because<br />

little or no relationship existed<br />

among them.<br />

Conclusions drawn from trials<br />

carried out on research stations are<br />

always open to critics who argue<br />

that these results do not necessarily<br />

reflect conditions in farmers’ fields.<br />

However, the wide acceptance <strong>of</strong><br />

<strong>CIMMYT</strong> germplasm by farmers in<br />

MEs 1-5 does not support the view<br />

that the wide adaptation <strong>of</strong><br />

<strong>CIMMYT</strong> germplasm is based on<br />

geographical rather than<br />

environmental differences.<br />

Breeding for High Yield<br />

Potential <strong>and</strong> Enhanced<br />

Stability<br />

Selection <strong>of</strong> segregating<br />

populations <strong>and</strong> consequent yield<br />

testing <strong>of</strong> advanced lines are<br />

paramount for identifying high<br />

yielding <strong>and</strong> input responsive<br />

wheat genotypes. The increase in<br />

yield potential <strong>of</strong> <strong>CIMMYT</strong>


6 Opening Remarks — S. Rajaram<br />

cultivars developed since the 1960s<br />

is shown in Figure 1 (Rees et al.,<br />

1993). The average increase per<br />

year was 0.9%, <strong>and</strong> there is no<br />

evidence that a yield plateau has<br />

been reached. This progress in<br />

increasing genetic yield potential is<br />

closely associated with an increase<br />

in photosynthetic activity (Rees et<br />

al., 1993.). Both photosynthetic<br />

activity <strong>and</strong> yield potential<br />

increased over the 30-year period<br />

by some 25%. These findings may<br />

have major implications for<br />

<strong>CIMMYT</strong>’s future selection strategy<br />

since there is evidence that wheat<br />

genotypes with a higher<br />

photosynthetic rate have lower<br />

canopy temperature, which can be<br />

easily, quickly, <strong>and</strong> cheaply<br />

measured using a h<strong>and</strong>-held<br />

thermometer. If this is verified in<br />

future trials, this trait may be used<br />

by breeders to increase selection<br />

efficiency for yield potential. This<br />

technique may be particularly<br />

useful in selecting wheat genotypes<br />

adapted to environments where<br />

heat is a production constraint.<br />

Yield per se is closely associated<br />

with input responsiveness.<br />

Increasing the input efficiency at<br />

low production levels can shift<br />

crossover points, provided they<br />

exist, <strong>and</strong> enhance residual effects<br />

<strong>of</strong> high genetic yield potential.<br />

Furthermore, combining input<br />

efficiency with high yield potential<br />

will allow farmers to benefit from<br />

such cultivars over a wide range <strong>of</strong><br />

input levels. The increase in<br />

nitrogen use efficiency is shown in<br />

Figure 2 (Ortiz-Monasterio et al.,<br />

1995).<br />

<strong>CIMMYT</strong>’s breeding strategy<br />

has resulted in the development <strong>of</strong><br />

widely grown varieties, such as<br />

Siete Cerros, Anza, Sonalika, <strong>and</strong><br />

Seri 82, which at their peak were<br />

grown on several million<br />

hectares. Seri 82 was released for<br />

irrigated as well as rainfed<br />

environments. Reynolds et al.<br />

(1994) reported that Seri 82 was<br />

the highest yielding entry in the<br />

1st <strong>and</strong> 2nd International Heat<br />

Stress Genotype Experiment.<br />

Seri 82 can be considered as the<br />

first wheat genotype truly<br />

adapted to several MEs,<br />

particularly to ME1, ME2, ME4,<br />

<strong>and</strong> ME5. A comparison between<br />

Seri 82 <strong>and</strong> Pastor, a recently<br />

developed <strong>CIMMYT</strong> cultivar,<br />

demonstrates the progress made<br />

in widening adaptation during<br />

the last ten years. Figure 3 shows<br />

the performance <strong>of</strong> Pastor<br />

(Pfau/Seri//Bow) in <strong>CIMMYT</strong>’s<br />

13th Elite Spring Wheat Yield<br />

Nursery. In 50 trials grown in all<br />

6 MEs, Pastor yielded<br />

significantly (P=0.01) lower than<br />

the highest yielding entry only<br />

in eight trials. This figure also<br />

demonstrates that Pastor has no<br />

tendency to crossover at any<br />

yield level. While we do not<br />

reject that crossover may exist<br />

for some cultivars, Pastor <strong>and</strong><br />

Seri 82 are clear examples that it<br />

is possible to combine abiotic<br />

stress tolerance with high yield<br />

potential. Figure 4 shows the<br />

yield difference between Seri 82<br />

<strong>and</strong> Pastor. Only in 16 out <strong>of</strong> 50<br />

trials did Seri outyield Pastor.<br />

The latter cultivar proves that<br />

breeding for wide adaptation<br />

has not yet reached its limit.<br />

Apart from the physiological<br />

basis <strong>of</strong> yield potential, the yield<br />

gains in <strong>CIMMYT</strong> wheats are<br />

due to the utilization <strong>of</strong> certain<br />

genetic resources. The<br />

germplasm has been paramount<br />

to increase yield in <strong>CIMMYT</strong>’s<br />

Wheat Program <strong>and</strong> in<br />

Yield in kg/ha<br />

8000<br />

Y=-104607+56.56x<br />

7500<br />

7000<br />

R=0.96***<br />

Gain/year 0.9%<br />

6500<br />

6000<br />

1960<br />

1965 1970 1975 1980 1985 1990<br />

Variety year <strong>of</strong> release<br />

Figure 1. Mean grain yields for the historical series<br />

<strong>of</strong> bread wheat varieties for the year 1990-93 at Cd.<br />

Obregon, Mexico (data from Rees et al., 1993).<br />

Grain yield in kg/ha<br />

8000<br />

7000<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

50<br />

300 kg N<br />

ON<br />

55 60 65 70 75 80 85<br />

Genotype year <strong>of</strong> release<br />

Figure 2. Grain yield <strong>of</strong> the historical series <strong>of</strong> bread<br />

wheats at Cd. Obregon, Mexico, at 0 <strong>and</strong> 300 kg/ha N<br />

application (data from J.I. Ortiz-Monasterio et al.,<br />

1995).<br />

Yield (kg/ha)<br />

12000<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

0<br />

2000 4000 6000 8000 10000<br />

Location mean yield (kg/ha)<br />

Maximum yield<br />

Pastor significantly different<br />

from highest yield in entry<br />

Mean<br />

Figure 3. Yield <strong>of</strong> Pastor at 50 locations <strong>of</strong> the 13th<br />

ESWYT.<br />

Yield (kg/ha)<br />

4000<br />

3000 Pastor<br />

2000<br />

1000<br />

0<br />

-1000<br />

Seri 82<br />

-2000<br />

-3000<br />

600<br />

1000<br />

2000<br />

3000<br />

4000<br />

5000<br />

6000<br />

7000<br />

8000<br />

Location mean yield (kg/ha)<br />

9000<br />

10000<br />

Figure 4. Yield difference between Pastor <strong>and</strong> Seri 82<br />

at 50 locations <strong>of</strong> the 13th ESWYT.


Minnesota’s barley program<br />

(Rasmusson, 1996). Some examples<br />

are listed next.<br />

• The incorporation <strong>of</strong> Norin 10 x<br />

Brevor germplasm not only<br />

produced dwarf wheats, but also<br />

simultaneously gave high yield.<br />

• Spring <strong>and</strong> winter crosses<br />

involving the variety Kavkaz<br />

resulted in Veerys, representing<br />

high yield potential <strong>and</strong><br />

enhanced yield stability<br />

(Figure 5).<br />

• The incorporation <strong>of</strong> the Lr19<br />

gene <strong>and</strong> Aegilops squarrosaderived<br />

synthetic wheats has<br />

further increased yield potential.<br />

The variety Super Seri has the<br />

Lr19 gene (Figure 6) <strong>and</strong> a<br />

derivative <strong>of</strong> Ae. squarrosa is<br />

given in Table 3.<br />

Breeding for durable disease<br />

resistance<br />

From the beginning,<br />

incorporating durable, non-specific<br />

disease resistance into <strong>CIMMYT</strong><br />

germplasm was a high priority,<br />

since breeding widely adapted<br />

germplasm with stable yields<br />

without adequate resistance against<br />

the major diseases would be<br />

impossible. The concept goes back<br />

to Niederhauser et al. (1954),<br />

Borlaug (1966), <strong>and</strong> Caldwell<br />

(1968), who advocated developing<br />

general resistance in the <strong>CIMMYT</strong><br />

Yield (t/ha)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

ISWYN 15<br />

2 3 5<br />

Environments (t/ha)<br />

7 9<br />

Figure 5. Performance <strong>of</strong> Veery in 73 global<br />

environments (ISWYN 15).<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 7<br />

program versus the specific or<br />

hypersensitive type. Very diverse<br />

sources <strong>of</strong> resistance for rusts <strong>and</strong><br />

other diseases are intentionally<br />

used in the crossing program. The<br />

major sources are germplasm from<br />

national programs, advanced<br />

<strong>CIMMYT</strong> lines, germplasm<br />

received from the <strong>CIMMYT</strong> or<br />

other genebanks, <strong>and</strong> <strong>CIMMYT</strong>’s<br />

wide crossing program.<br />

<strong>CIMMYT</strong>’s strategy in the case<br />

<strong>of</strong> cereal rusts is to breed for<br />

general resistance (slow rusting)<br />

based on historically proven stable<br />

genes. This non-specific resistance<br />

can be further diversified by<br />

accumulating several minor genes<br />

<strong>and</strong> combining them with different<br />

specific genes to provide a certain<br />

degree <strong>of</strong> additional genetic<br />

diversity. This concept is also<br />

applied to other diseases like<br />

septoria leaf blotch,<br />

helminthosporium spot blotch, <strong>and</strong><br />

fusarium head scab. The present<br />

situation <strong>of</strong> <strong>CIMMYT</strong> germplasm<br />

regarding resistance to major<br />

diseases may be summarized as<br />

follows:<br />

• Stem rust (Puccinia graminis f.sp.<br />

tritici) resistance has been stable<br />

after 40 years <strong>of</strong> utilization <strong>of</strong> the<br />

genes derived from the variety<br />

Hope. Losses due to stem rust<br />

Veery<br />

Average<br />

yield<br />

have been negligible since the<br />

late 1960s. The resistance is<br />

based on the gene complex Sr2,<br />

which actually consists <strong>of</strong> Sr2<br />

plus 4-5 minor genes pyramided<br />

into three to four gene<br />

combinations (Rajaram et al.,<br />

1988). Sr 2 alone behaves as a<br />

slow rusting gene. Since there<br />

has been no major stem rust<br />

epidemic in areas where<br />

<strong>CIMMYT</strong> germplasm is grown,<br />

the resistance seems to be<br />

durable.<br />

• Leaf rust (Puccinia recondita f.sp.<br />

tritici) resistance has been<br />

stabilized by using genes<br />

derived from many sources, in<br />

particular the Brazilian cultivar<br />

Frontana (Singh <strong>and</strong> Rajaram,<br />

1992). No major epidemic has<br />

been observed in almost 20<br />

years. Four partial resistance<br />

genes, including Lr 34, give a<br />

slow rusting response <strong>and</strong> have<br />

been the reason for the<br />

containment <strong>of</strong> leaf rust<br />

epidemics in the developing<br />

world during the last 15 years.<br />

About 60% <strong>of</strong> <strong>CIMMYT</strong><br />

germplasm carries one to four <strong>of</strong><br />

these partial resistance genes.<br />

Lr34 is linked to Yr18 as well as<br />

to a morphological marker (leaf<br />

tip necrosis) that makes the gene<br />

particularly attractive for<br />

8500<br />

Super Seri<br />

8000<br />

Bacanora 88<br />

7500<br />

Seri 82 Oasis 86<br />

7000 Yecora 70<br />

Ciano 79<br />

Nacozari 76<br />

6500<br />

6000<br />

1960<br />

Siete Cerros 66<br />

Pitic 62<br />

1970 1980<br />

Year <strong>of</strong> release<br />

1990 2000<br />

Figure 6. Increase in grain yield potential <strong>of</strong><br />

<strong>CIMMYT</strong>-derived wheats as a function <strong>of</strong> year <strong>of</strong><br />

release.<br />

Grain yield in kg/ha


8 Opening Remarks — S. Rajaram<br />

breeders (Singh, 1992a, b).<br />

<strong>CIMMYT</strong> continues to look for<br />

new sources <strong>of</strong> partial resistance.<br />

• Stripe rust (Puccinia striiformis):<br />

Slow rusting genes like Yr18<br />

have been identified (Singh,<br />

1992b); however, their<br />

interaction is less additive than<br />

for leaf <strong>and</strong> stem rust. More<br />

basic research is needed to<br />

underst<strong>and</strong> the status <strong>of</strong> durable<br />

resistance in high yielding<br />

germplasm. The breakdown <strong>of</strong><br />

Yr9 in West Asia <strong>and</strong> North<br />

Africa <strong>and</strong> the present yellow<br />

rust epidemics underline the<br />

need for the release <strong>of</strong> cultivars<br />

with accumulated durable<br />

resistance.<br />

• <strong>Septoria</strong> tritici: Initially all<br />

semidwarf cultivars developed<br />

for irrigated conditions were<br />

susceptible. Today more than<br />

eight genes have been identified<br />

in <strong>CIMMYT</strong> germplasm <strong>and</strong> two<br />

to three genes in combination<br />

provide acceptable resistance.<br />

Future activities will concentrate<br />

on pyramiding these genes <strong>and</strong><br />

spreading them more widely<br />

within <strong>CIMMYT</strong> germplasm<br />

(Jlibene, 1992; Matus-Tejos,<br />

1993).<br />

• Karnal bunt (Tilletia indica):<br />

More than five genes have been<br />

identified <strong>and</strong> most <strong>of</strong> them are<br />

partially dominant. Genes<br />

providing resistance to Karnal<br />

bunt have been incorporated<br />

into high yielding lines (Singh et<br />

al., 1995).<br />

• Powdery mildew (Erysiphe<br />

graminis f.sp. tritici): <strong>CIMMYT</strong><br />

germplasm is considered<br />

vulnerable to this disease. The<br />

disease is absent in Mexico <strong>and</strong><br />

the responsibility to transfer<br />

resistance genes has been<br />

delegated to <strong>CIMMYT</strong>’s regional<br />

breeder in South America.<br />

Breeding for<br />

Drought Tolerance<br />

There has been a large<br />

transformation in the productivity<br />

<strong>of</strong> wheat due to the application <strong>of</strong><br />

Green Revolution technology. This<br />

has resulted in a doubling <strong>and</strong><br />

tripling <strong>of</strong> wheat production in<br />

many environments, but especially<br />

in irrigated areas. High yielding<br />

semidwarf wheats have<br />

continuously replaced the older tall<br />

types at a rate <strong>of</strong> 2 million ha per<br />

year since 1977 (Byerlee <strong>and</strong> Moya,<br />

1993).<br />

There is a growing recognition<br />

that the dissemination, application,<br />

<strong>and</strong> adoption <strong>of</strong> this technology<br />

has, however, been slower in<br />

marginal environments, especially<br />

in the semiarid environments<br />

affected by poor distribution <strong>of</strong><br />

water <strong>and</strong> drought. The annual<br />

gain in genetic yield potential in<br />

drought environments is only<br />

about half that obtained in<br />

irrigated, optimum conditions.<br />

Many investigators have attempted<br />

to produce wheat varieties adapted<br />

to these semiarid environments<br />

with limited success. Others have<br />

criticized the Green Revolution<br />

technology (Ceccarelli et al., 1987)<br />

for failing to adequately address<br />

productivity constraints in<br />

semiarid environments, although<br />

their own recommended<br />

technology has had limited impact,<br />

in particular in farmers’ fields. This<br />

criticism is in clear contrast to the<br />

actual acceptance <strong>of</strong> semidwarf<br />

wheat cultivars in rainfed areas,<br />

since most <strong>of</strong> the 16 million ha<br />

increase in the area sown to<br />

Mexican semidwarf wheats in the<br />

mid-1980s occurred in rainfed<br />

areas; in 1990, more than 60% <strong>of</strong> the<br />

dryl<strong>and</strong> area in developing<br />

countries were planted with<br />

semidwarfs (Byerlee <strong>and</strong> Moya,<br />

1993).<br />

Definition <strong>of</strong> semiarid<br />

environments <strong>and</strong><br />

description<br />

<strong>of</strong> drought patterns<br />

In Table 2, the major global<br />

drought patterns observed in<br />

wheat production are presented<br />

(Rajaram et al., 1994, Edmeades et<br />

al., 1989). Through respectively<br />

dealing with spring (ME4A),<br />

facultative (ME9), <strong>and</strong> winter<br />

wheat (ME12), these three MEs are<br />

characterized by sufficient rainfall<br />

prior to anthesis, followed by<br />

drought during the grain-filling<br />

period. In South America, the<br />

Southern Cone type <strong>of</strong> drought<br />

(ME4B) is characterized by<br />

moisture stress early in the crop<br />

season, with rainfall occurring<br />

during the post-anthesis phase. In<br />

the Indian Subcontinent type <strong>of</strong><br />

drought stress (ME4C), the wheat<br />

crop utilizes water reserves left<br />

from the monsoon rains during the<br />

previous summer season. In the<br />

Subcontinent the irrigated wheat<br />

crop (ME1) may also suffer drought<br />

due to a reduced or less than<br />

optimum number <strong>of</strong> irrigations.<br />

The traditional methodology<br />

for breeding for drought stress is<br />

typified by h<strong>and</strong>ling all segregating<br />

populations under target<br />

conditions <strong>of</strong> drought, <strong>and</strong> the use<br />

<strong>of</strong> local l<strong>and</strong>races is recommended<br />

in the breeding process (Ceccarelli<br />

et al., 1987). The methodology rests


on the assumption that the agroecological<br />

situation facing the<br />

farmer does not vary in its<br />

expression over time <strong>and</strong> that<br />

responsiveness <strong>of</strong> varieties to<br />

improved growing conditions will<br />

not be needed. It also assumes that<br />

crossover will always occur below<br />

a certain yield level under dry<br />

conditions, where modern high<br />

yielding varieties <strong>of</strong> a responsive<br />

nature would always yield less<br />

than traditional l<strong>and</strong>race-based<br />

genotypes. Such crossovers may<br />

occur for selected genotypes, <strong>and</strong><br />

one should always be open to the<br />

possibility that there are real<br />

“drought tolerance” traits<br />

operating at the 1 t/ha <strong>and</strong> below<br />

yield level that adversely affect<br />

high yield potential at the 4 t/ha<br />

<strong>and</strong> higher yield levels. So far such<br />

traits have not been identified at<br />

<strong>CIMMYT</strong>. In any case, crossover<br />

would be restricted to such harsh<br />

conditions, where in fact farmers<br />

choose—rightfully so—not to grow<br />

wheat at all, but to produce other,<br />

more drought tolerant crops such<br />

as barley or sorghum, or resort to<br />

grazing (van Ginkel et al., 1998).<br />

At <strong>CIMMYT</strong> we advocate an<br />

“open-ended system” <strong>of</strong> breeding<br />

in which yield responsiveness is<br />

combined with adaptation to<br />

drought conditions. Most semiarid<br />

environments differ significantly<br />

across years in their water<br />

availability <strong>and</strong> distribution<br />

pattern. Hence it is prudent to<br />

construct a genetic system in which<br />

plant responsiveness provides a<br />

bonus whenever conditions<br />

improve due to higher rainfall.<br />

With such a system, improved<br />

moisture conditions immediately<br />

translate into greater gains for the<br />

farmer.<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 9<br />

The Veerys<br />

In the early 1980s, when<br />

advanced lines derived from the<br />

spring x winter cross Kavkaz/<br />

Buho//KAL/BB (CM33027) were<br />

tested in 73 global environments <strong>of</strong><br />

the 15th International Wheat Yield<br />

Nursery (15th ISWYN) (Figure 5),<br />

their performance was quite<br />

untypical compared to any<br />

previously known high yielding<br />

varieties. In later tests, we found<br />

that these lines, called Veerys, carry<br />

the 1B.1R translocation from rye<br />

<strong>and</strong> that the general performance<br />

<strong>of</strong> such germplasm was superior<br />

not only in high yielding<br />

environments but particularly<br />

under drought conditions (Villareal<br />

et al., 1995; Table 4). From the Veery<br />

cross 43 varieties were released,<br />

excluding those released in Europe.<br />

The Veerys represent a genetic<br />

system in which high yield<br />

performance in favorable<br />

environments <strong>and</strong> adaptation to<br />

drought could be combined in one<br />

genotype. The two genetic systems<br />

are apparently not always<br />

incompatible, although others have<br />

claimed that their combination<br />

would not be possible. However, it<br />

is possible to hypothesize a plant<br />

system in which efficient input use<br />

<strong>and</strong> responsiveness to improved<br />

levels <strong>of</strong> external inputs (in this<br />

case, available water) can be<br />

combined to produce germplasm<br />

for marginal (in this case, semiarid)<br />

environments that at least<br />

maintains minimum traditional<br />

yields <strong>and</strong> expresses dramatic<br />

increases whenever conditions<br />

improve. The impacts described<br />

below support the utilization <strong>of</strong><br />

this methodology.<br />

• By the mid-1980s <strong>CIMMYT</strong><br />

germplasm occupied 45% <strong>of</strong> the<br />

semiarid wheat areas with 300-<br />

500 mm <strong>of</strong> rainfall, <strong>and</strong> 21% <strong>of</strong><br />

the area less than 300 mm<br />

(Morris et al., 1991), including<br />

large tracts in West Asia/North<br />

Africa (WANA). By 1990 63% <strong>of</strong><br />

the dryl<strong>and</strong> areas, especially in<br />

ME4A <strong>and</strong> ME4B, was planted<br />

with semidwarf wheats (Byerlee<br />

<strong>and</strong> Moya, 1993), many carrying<br />

the 1B/1R translocation.<br />

• To support the above<br />

assumptions, an experiment was<br />

conducted (Calhoun et al., 1994;<br />

Tables 5 <strong>and</strong> 6) to determine<br />

how the most modern <strong>and</strong><br />

widely (spatially) adapted<br />

germplasm compared to<br />

commercial germplasm from<br />

countries representing the<br />

Mediterranean region (ME4A),<br />

the Southern Cone <strong>of</strong> South<br />

America (ME4B), <strong>and</strong> the Indian<br />

Table 4. Effect <strong>of</strong> the 1BL.1RS translocation on yield characteristics <strong>of</strong> 28 r<strong>and</strong>om F2-derived F6<br />

lines from the cross Nacozari 76/Seri 82 under reduced irrigated conditions.<br />

Plant characteristics 1BL.1RS 1B Mean diff.<br />

Grain yield 4945 4743 202 *<br />

Above-ground biomass at maturity (t/ha) 12600 12100 500 *<br />

Grains/m2 14074 13922 152NS<br />

Grains/spike 43.5 40.6 2.9 *<br />

1000-grain weight (g) 37.1 36.5 0.5 *<br />

Source: Villareal et al. (1995).<br />

Note: NS: not significatnt, * : significant at the 0.05 level.


10 Opening Remarks — S. Rajaram<br />

Table 5. Wheat genotypes representing adaptation to different moisture environments.<br />

ME1 ME4A ME4B ME4 Irrigation<br />

(Mediterranean)<br />

(Southern Cone)<br />

(Subcontinent)<br />

Super Kauz, Pavon 76, Genaro 81, Opata 85<br />

Almansor, Nesser, Sitta, Siete Cerros<br />

Cruz Alta, Prointa Don Alberto, LAP1376, PSN/BOW CM69560<br />

C306, Sonalika, Punjab 81, Barani<br />

Source: Calhoun et al. (1994).<br />

Table 6. Grain yields <strong>of</strong> selected wheat genotypes grouped by adaptation <strong>and</strong> tested under<br />

moisture regimes in the Yaqui Valley, Mexico, 1989-90 <strong>and</strong> 1990-91<br />

Full Late<br />

Adaptation group<br />

Early Residual<br />

irrigation1 drought2 drought3 moisture4 ME1 Irrigation 6636 a * ME4C ME4B ME4C Mediterranean<br />

Southern Cone<br />

Subcontinent<br />

6342 b<br />

5028 c<br />

4778 c<br />

4198 a<br />

3990 ab<br />

3148 bc<br />

3245 bc<br />

4576 a<br />

4390 b<br />

4224 b<br />

3657 c<br />

3032 a<br />

2883 b<br />

2359 c<br />

2704 b<br />

Source: Calhoun, et al. 1994<br />

1 Received 5 irrigations; 2 received 2 irrigations early before heading; 3 received one irrigation for<br />

germination <strong>and</strong> two post heading; 4 received one irrigation for germination only.<br />

* Means in the same column followed by the same letter are not significantly different at P=0.05.<br />

Subcontinent (ME4C), under<br />

conditions artificially simulating<br />

those three MEs. The most<br />

widely adapted <strong>CIMMYT</strong> lines<br />

outyielded the commercial<br />

varieties in all artificially<br />

simulated environments.<br />

• Nesser is an advanced line with<br />

superior performance in<br />

drought conditions bred at<br />

<strong>CIMMYT</strong>/Mexico <strong>and</strong> identified<br />

at ICARDA/Syria. The cross<br />

combines a high yielding<br />

<strong>CIMMYT</strong> variety Jupateco <strong>and</strong> a<br />

drought tolerant Australian<br />

variety W3918A. The<br />

performance <strong>of</strong> Nesser in<br />

WANA’s ME4A environments<br />

has been widely publicized<br />

(ICARDA, 1993), <strong>and</strong> the line is<br />

considered by ICARDA to<br />

represent a uniquely drought<br />

tolerant genotype. However, it<br />

was selected at <strong>CIMMYT</strong>/<br />

Mexico under favorable<br />

environments, <strong>and</strong> carries a<br />

combination <strong>of</strong> input efficiency<br />

<strong>and</strong> high yield responsiveness. It<br />

performs similarly to the Veery<br />

lines in the absence <strong>of</strong> rust.<br />

The breeding scheme<br />

The breeding scheme<br />

described below is used to<br />

combine the two genetic systems.<br />

Two contrasting selection<br />

environments are alternated,<br />

allowing alternate selection for<br />

input efficiency <strong>and</strong> input<br />

responsiveness.<br />

F1 Crosses involving spatially<br />

widely adapted germplasm<br />

representing yield stability <strong>and</strong><br />

yield potential, with lines with<br />

proven drought tolerance in<br />

the specific setting <strong>of</strong> either<br />

ME4A, ME4B or ME4C. Winter<br />

wheats <strong>and</strong> synthetic<br />

germplasm are emphasized.<br />

F2 Individual plants are raised<br />

under irrigated <strong>and</strong> optimally<br />

fertilized conditions <strong>and</strong><br />

inoculated with a wide<br />

spectrum <strong>of</strong> rust virulence.<br />

Only robust <strong>and</strong> (horizontally)<br />

resistant plants are selected.<br />

These may represent<br />

adaptation to favorable<br />

environments.<br />

F3, F4 The selected F2 plants are<br />

evaluated using a modified<br />

pedigree/bulk breeding system<br />

(Rajaram <strong>and</strong> van Ginkel, 1995)<br />

under rainfed conditions or very<br />

low water availability. The<br />

selection is based on individual<br />

lines rather than individual<br />

plants. The progenies are<br />

selected based on such criteria as<br />

spike density, biomass/vigor,<br />

grains/m 2 , <strong>and</strong> others (van<br />

Ginkel et al., 1998) (Table 7). This<br />

index helps identify lines which<br />

may adapt to low water<br />

situations.<br />

F5, F6 The selected lines from F4<br />

are further evaluated under<br />

optimum conditions.<br />

Table 7. Genotypic correlation (rg) between<br />

agronomic traits <strong>and</strong> final grain yield, for<br />

optimum environment (full irrigations) <strong>and</strong><br />

reduced water regime (late drought,<br />

Mediterranean type) in wheat.<br />

Moisture regime<br />

Full Late<br />

Trait irrigation drought<br />

Days to heading 0.40 0.19<br />

Days to maturity 0.29 0.27<br />

Grain fill period -0.32 0.36<br />

Height -0.39 0.05<br />

Peduncle length<br />

Relative peduncle<br />

-0.46 0.22<br />

extrusion -0.51* 0.25<br />

Spike length -0.28 -0.50*<br />

Spike/m2 -0.12 0.64**<br />

Grains/spike 0.62* -0.42<br />

Grains/m2 0.74** 0.68**<br />

Yield/spike 0.55* -0.64**<br />

1000 grain weight 0.08 -0.45<br />

Test weight 0.13 0.05<br />

Harvest index 0.83** -0.39<br />

Biomass 0.90** 0.94**<br />

Straw yield 0.52* 0.86**<br />

Yield / day (planting) 0.99** 0.57*<br />

Yield / day (heading) 0.94** 0.44<br />

Biomass / day (planting) 0.86** 0.69**<br />

Biomass / day (heading) 0.74** 0.63**<br />

Vegetative growth rate 0.32 0.63**<br />

Spike growth rate 0.62** -0.58*<br />

Grain growth rate 0.17 -0.44<br />

*, ** indicate significance at the 0.05 <strong>and</strong><br />

0.01probability level, respectively.<br />

Source: van Ginkel et al. (1998).


F7, F8 Simultaneous evaluations<br />

under optimum <strong>and</strong> low water<br />

environments. Selection <strong>of</strong> lines<br />

showing outst<strong>and</strong>ing<br />

performance under both<br />

conditions. Further evaluation in<br />

international environments is<br />

carried out for purposes <strong>of</strong><br />

verification.<br />

The proposed breeding<br />

methodology is supported by<br />

research published in recent years<br />

by others, not only on wheat<br />

(Bramel-Cox et al., 1991; Cooper et<br />

al., 1994; Duvick, 1990, 1992;<br />

Ehdaie et al., 1988; Uddin et al.,<br />

1992; Zavala-Garcia et al., 1992),<br />

where the importance <strong>of</strong> testing<br />

<strong>and</strong> selecting in a range <strong>of</strong><br />

environments, including wellirrigated<br />

ones, has shown to<br />

identify superior genotypes for<br />

stressed conditions. The<br />

methodology aims at combining<br />

input efficiency with input<br />

responsiveness by alternating<br />

selection environments during the<br />

breeding process.<br />

Future Research<br />

Directions<br />

Yield stability <strong>and</strong> yield<br />

potential<br />

Traxler et al. (1995) analyzed<br />

grain yield increases <strong>and</strong> yield<br />

stability <strong>of</strong> bread wheat cultivars<br />

released during the last 45 years. In<br />

the early period <strong>of</strong> the Green<br />

Revolution, when rapid yield<br />

increases occurred, variance for<br />

yield concomitantly increased.<br />

Since the early 1970s, yield stability<br />

has increased at the cost <strong>of</strong><br />

increases in yield. However, steady<br />

progress was made in developing<br />

varieties with improved stability,<br />

grain yield or both. For the<br />

developing world, yield stability<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 11<br />

increased since the beginning <strong>of</strong> the<br />

Green Revolution (Smale <strong>and</strong><br />

McBride, 1996). While price policy,<br />

input supplies, <strong>and</strong> environmental<br />

variation contribute more to yield<br />

stability than the genotype, the<br />

increasing yield stability reflects<br />

the emphasis given by breeders to<br />

develop germplasm with tolerance<br />

to a wider range <strong>of</strong> diseases <strong>and</strong><br />

abiotic stresses. Sayre et al. (1997)<br />

concluded that from 1964 to 1990,<br />

yield potential in <strong>CIMMYT</strong>-derived<br />

cultivars increased at a rate <strong>of</strong> 67<br />

kg/ha/yr or 0.88% per year. The<br />

data did not suggest that a yield<br />

plateau had been reached <strong>and</strong> the<br />

performance <strong>of</strong> recently released<br />

lines, such as Attilla (pb343) <strong>and</strong><br />

Babax (Baviacora M92) indicates<br />

that yield potential has been<br />

further enhanced. Improvements<br />

made by breeding for yield stability<br />

<strong>and</strong> adaptation may be illustrated<br />

by data for the advanced line<br />

Pastor, which outyielded the<br />

hallmark check cultivar Seri 82 in<br />

34 out <strong>of</strong> 50 locations where the 13 th<br />

Elite Spring Wheat Yield Nursery<br />

was grown (Figure 4). Discussion<br />

on how to increase the yield<br />

potential <strong>of</strong> wheat <strong>of</strong>ten still centers<br />

around traits that contributed to<br />

the success <strong>of</strong> the Green Revolution<br />

varieties more than 30 years ago,<br />

e.g., photoperiod <strong>and</strong> dwarfing<br />

genes (Worl<strong>and</strong> et al., 1998; Sears,<br />

1998).<br />

<strong>CIMMYT</strong> has made a modest<br />

investment in restructuring <strong>and</strong><br />

creation <strong>of</strong> a new plant type<br />

characterized by robust stem, broad<br />

leaf, long spike (30 cm), <strong>and</strong> large<br />

numbers <strong>of</strong> grain per spike. The<br />

new plant type still suffers due to<br />

diseases <strong>and</strong> is deficient in quality<br />

<strong>and</strong> certain agronomic<br />

characteristics. In 1994, we<br />

launched a dynamic breeding<br />

program to correct these<br />

deficiencies.<br />

Plant nutrition<br />

Selecting for yield potential <strong>and</strong><br />

yield stability under medium to<br />

high levels <strong>of</strong> nitrogen has<br />

indirectly increased efficiency for<br />

nutrient uptake. Recently released<br />

<strong>CIMMYT</strong> bread wheat cultivars<br />

require less nitrogen to produce a<br />

unit amount <strong>of</strong> grain than cultivars<br />

released in the previous decades<br />

(Ortiz Monasterio et al., 1997).<br />

Under low N levels in the soil, N<br />

use efficiency increased mainly due<br />

to a higher N uptake efficiency—<br />

the ability <strong>of</strong> plants to absorb N<br />

from the soil—whereas under high<br />

N levels, the N utilization<br />

efficiency—the capacity <strong>of</strong> plants to<br />

convert the absorbed N into grain<br />

yield—increased. In spite <strong>of</strong> the<br />

increased N use efficiency <strong>of</strong><br />

recently released wheat cultivars,<br />

the response to nitrogen <strong>of</strong> wheat<br />

production systems has been<br />

observed to be declining in many<br />

areas <strong>of</strong> Southeast Asia. In Turkey,<br />

where zinc deficient soils are<br />

common, recently released winter<br />

bread wheat cultivars have<br />

improved Zn uptake <strong>and</strong><br />

consequently higher grain yield<br />

than local l<strong>and</strong>races (M. Kalayci,<br />

pers. comm.).<br />

Physiology<br />

A recent survey <strong>of</strong> wheat<br />

breeders suggested that research in<br />

plant physiology has had a limited<br />

impact on wheat improvement<br />

(Jackson et al., 1996). A strong body<br />

<strong>of</strong> evidence now indicates that<br />

physiological traits may have real<br />

potential for complementing early<br />

generation phenotypic selection in<br />

wheat. One <strong>of</strong> the more promising<br />

traits identified is canopy


12 Opening Remarks — S. Rajaram<br />

temperature depression (CTD).<br />

CTD refers to the cooling effect<br />

exhibited by a leaf as transpiration<br />

occurs. While soil water status has<br />

a major influence on CTD, there are<br />

strong genotypic effects under<br />

well-watered, heat-stressed, or<br />

drought-stressed conditions. CTD<br />

gives an indirect estimate <strong>of</strong><br />

stomatal conductance <strong>and</strong> is a<br />

highly integrative trait affected by<br />

several major physiological<br />

processes, including photosynthetic<br />

metabolism, evapotranspiration,<br />

<strong>and</strong> plant nutrition. CTD <strong>and</strong><br />

stomatal conductance, measured<br />

on sunny days during grain filling,<br />

showed a strong association with<br />

the yield <strong>of</strong> semidwarf wheat lines<br />

grown under irrigation, in both<br />

temperate (Fischer et al., 1998) <strong>and</strong><br />

subtropical environments<br />

(Reynolds et al., 1994). In addition,<br />

CTD, as measured on large<br />

numbers <strong>of</strong> advanced breeding<br />

lines in irrigated yield trials, was a<br />

powerful predictor <strong>of</strong> performance<br />

not only at the selection site but<br />

also for yield averaged across 15<br />

international sites. CTD has been<br />

shown to be associated with yield<br />

differences between homozygous<br />

lines, indicating a potential for<br />

genetic gains in yield, in response<br />

to selection for CTD (Reynolds et<br />

al., 1998).<br />

Germplasm is paramount<br />

Three-quarters <strong>of</strong> recently<br />

surveyed wheat breeders felt that a<br />

lack <strong>of</strong> genetic diversity would<br />

limit future breeding advances<br />

(Rejesus et al., 1996), though<br />

genetic diversity was not<br />

considered an immediately limiting<br />

factor in most programs. This<br />

concern was greater from breeders<br />

in developing <strong>and</strong> former USSR<br />

countries (>80%) than from higher<br />

income countries (59%).<br />

Furthermore, in countries where<br />

privatization <strong>of</strong> wheat breeding<br />

programs has occurred,<br />

investments in strategic germplasm<br />

development that may be risky or<br />

important only in the long term<br />

have declined (McGuire, 1997).<br />

A wide range <strong>of</strong> opinion has<br />

been expressed concerning the<br />

abundance or availability <strong>of</strong><br />

usefully exploitable genetic<br />

variability. Allard (1996)<br />

emphasized that the most readily<br />

useful genetic resources were<br />

modern elite cultivars, since these<br />

lines possessed relatively high<br />

frequencies <strong>of</strong> favorable alleles.<br />

Rasmusson <strong>and</strong> Phillips (1997)<br />

have shown that the assumption<br />

that all genetic variability is a result<br />

<strong>of</strong> the inherent exclusive<br />

contribution <strong>of</strong> two parents, per se,<br />

is not necessarily true, considering<br />

results from molecular analysis.<br />

They discuss mechanisms by which<br />

induction <strong>of</strong> genetic variability may<br />

involve altering the expression <strong>of</strong><br />

genes, the possible mechanisms <strong>of</strong><br />

single allele change, intragenic<br />

recombination, unequal crossover,<br />

element transpositions, DNA<br />

methylation, paramutation, or gene<br />

amplification. They also stressed<br />

the possible importance <strong>of</strong> epistasis<br />

effects which may have been<br />

underestimated in the past.<br />

Introduction <strong>of</strong> genetic<br />

variability from distantly related<br />

wheat cultivars, or related or alien<br />

species, has <strong>of</strong>ten been specifically<br />

aimed at the introduction <strong>of</strong> simply<br />

inherited traits (e.g., genes for<br />

disease resistance), but it has<br />

appeared to be <strong>of</strong> limited value in<br />

quantitative trait improvement<br />

(Cox et al., 1997). incorporated<br />

genes for leaf rust resistance from<br />

T. tauschii into bread wheat. With<br />

two backcrosses to the recurrent<br />

wheat parent, leaf rust resistant<br />

winter wheat advanced lines with<br />

acceptable quality <strong>and</strong> equal in<br />

yield to the highest yielding<br />

commercially grown cultivars were<br />

identified. In addition, it has been<br />

postulated that since recombination<br />

between the D genomes <strong>of</strong> T.<br />

aestivum <strong>and</strong> T. tauschii occurred at<br />

a level similar to that in an<br />

intraspecific cross (Fritz et al.,<br />

1995), T. tauschii could be<br />

considered another primary source<br />

<strong>of</strong> genes for wheat improvement.<br />

The number <strong>of</strong> wheat/rye<br />

translocations that have had a<br />

significant impact on wheat<br />

improvement are actually few in<br />

number. The majority <strong>of</strong> the<br />

1BL.1RS translocations occurring in<br />

more than 300 cultivars worldwide<br />

can be traced to one German source<br />

<strong>and</strong> all 1AL.1RS translocations,<br />

widely present in bread wheat<br />

cultivars grown in the Great Plains<br />

<strong>of</strong> the US, trace to one source,<br />

“Amigo” (Schlegel, 1997a,b;<br />

Rabinovich, 1998). Other<br />

translocations carry genes for<br />

copper efficiency (4BL.5R) <strong>and</strong><br />

Hessian fly resistance (2RL.2BS,<br />

6RL.6B, 6RL.4B, 6RL.4A; McIntosh,<br />

1993). Chromosomes 2R <strong>and</strong> 7R<br />

enhance zinc efficiency in wheatrye<br />

addition lines (Cakmak <strong>and</strong><br />

Braun, unpublished). Considering<br />

the impacts which have come from<br />

the use <strong>of</strong> wheat/rye<br />

translocations, further exploitation<br />

<strong>of</strong> these translocations may be<br />

warranted.<br />

While there have been reports<br />

indicating a positive effect <strong>of</strong><br />

1BL.1RS translocations on yield<br />

performance <strong>and</strong> adaptation<br />

(Rajaram et al., 1990), Singh et al.<br />

(1998) determined that with Seri 82,<br />

replacing the translocation with<br />

1BL from cv. Oasis resulted in a<br />

yield increase <strong>of</strong> 3.4 <strong>and</strong> 5.0% in


irrigated <strong>and</strong> moisture stress<br />

conditions, respectively. A further<br />

increase in grain yield in disease<br />

free conditions <strong>of</strong> about 5% was<br />

observed in the irrigated trials<br />

through the introgression <strong>of</strong><br />

7DL.7Ag translocation carrying the<br />

Lr19 gene (from Agropyron<br />

elongatum). This yield increase was<br />

attributed to higher rate <strong>of</strong> biomass<br />

production in the 7DL.7Ag lines.<br />

However, under moisture stress<br />

conditions 7DL.7Ag lines were<br />

associated with a 16% yield<br />

reduction, possibly due to<br />

excessive biomass production in<br />

early growth stages. This would<br />

suggest that the effect <strong>of</strong> the<br />

1BL.1RS translocation is genotype<br />

specific <strong>and</strong> that 7DL.7Ag could be<br />

a useful translocation for<br />

enhancing yield potential at least in<br />

irrigated conditions.<br />

Recent efforts to generate newly<br />

accessible genetic diversity have<br />

involved the reconstitution <strong>of</strong><br />

hexaploid wheat by producing<br />

“synthetic wheat” by crossing<br />

durum wheat (T. turgidum), the<br />

donor <strong>of</strong> the A <strong>and</strong> B genomes,<br />

with T. tauschii, the donor <strong>of</strong> the D<br />

genome (Mujeeb-Kazi et al., 1996).<br />

Villareal (1995) <strong>and</strong> Villareal et al.<br />

(1997) showed that lines derived<br />

after two backcrosses to T. aestivum<br />

showed increased morphoagronomic<br />

variation, <strong>and</strong> resistance<br />

to Karnal bunt <strong>and</strong> head scab.<br />

Under full irrigation in<br />

northwestern Mexico, the yield <strong>of</strong><br />

this material was nearly 8 t/ha.<br />

When tested under drought<br />

conditions for two years, nearly all<br />

<strong>of</strong> the synthetic derivatives had<br />

significantly higher 1000-kernel<br />

weight, with grain yield varying<br />

between 84 to 114%, when<br />

compared with the bread wheat<br />

checks.<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 13<br />

The more focused the breeding<br />

objective, the more restricted a<br />

breeder is in the choice <strong>of</strong> suitable<br />

parents. The use <strong>of</strong> genetically<br />

diverse material will continue to be<br />

a prime genetic source for<br />

increasing yield potential, a<br />

complex trait still not well<br />

understood genetically or<br />

physiologically. As long as breeders<br />

have no other readily accessible<br />

tools, genetic diversity <strong>and</strong> the<br />

opportunity for its recombination<br />

through crossing will be important<br />

to break undesired linkages <strong>and</strong><br />

increase the frequency <strong>of</strong> desirable<br />

alleles. Future breakthroughs in<br />

yield potential will likely come<br />

from such genetically diverse<br />

crosses.<br />

Hybrid wheat<br />

Pickett (1993) <strong>and</strong> Picket <strong>and</strong><br />

Galwey (1997) evaluated 40 years<br />

<strong>of</strong> wheat hybrid development <strong>and</strong><br />

concluded that hybrid wheat<br />

production is not economically<br />

feasible because <strong>of</strong> a) limited<br />

heterotic advantage; b) lack <strong>of</strong><br />

advantage in terms <strong>of</strong> agronomic,<br />

quality, or disease resistance traits;<br />

c) higher seed costs; <strong>and</strong> d)<br />

probably most importantly,<br />

heterosis could be “fixed” in<br />

polyploid plants <strong>and</strong> consequently<br />

hybrids would have no advantage<br />

over inbred lines.<br />

The use <strong>of</strong> hybrid crops is<br />

usually targeted to higher yield<br />

potential environments. Results<br />

from South Africa (Jordaan, 1996),<br />

however, show that hybrids<br />

outyield inbred lines by 15% at a 2<br />

t/ha mean production potential<br />

when narrow row spacing <strong>and</strong> low<br />

seeding rates (


14 Opening Remarks — S. Rajaram<br />

Conventional plant breeders<br />

adopt breeding methods that<br />

increase their breeding efficiency<br />

but are conservative when making<br />

methodological changes. A small<br />

survey <strong>of</strong> wheat programs having<br />

unrestricted access to new<br />

biotechnological methods found<br />

that few research programs, <strong>and</strong> no<br />

main-line wheat breeding<br />

programs, routinely used MAS or<br />

quantitatively inherited trait loci<br />

(QTL). Limitation in use is caused<br />

by lack <strong>of</strong> markers for traits <strong>of</strong><br />

interest, population specificity <strong>of</strong> a<br />

given marker, or markers’relatively<br />

high costs when compared with<br />

conventional selection techniques.<br />

These limitations may lessen in the<br />

next decade.<br />

Modern cultivars are the<br />

product <strong>of</strong> recombinations among<br />

the high number <strong>of</strong> l<strong>and</strong>races in<br />

their pedigrees (Smale <strong>and</strong><br />

McBride, 1996). In contemporary<br />

breeding programs, however,<br />

l<strong>and</strong>races are used directly only as<br />

a source <strong>of</strong> qualitatively inherited<br />

traits. Tanksley <strong>and</strong> McCouch<br />

(1997) argued that the lack <strong>of</strong><br />

success from crosses involving<br />

l<strong>and</strong>races for the improvement <strong>of</strong><br />

grain yield was mainly due to<br />

evaluation on a phenotypic basis,<br />

an imprecise indicator <strong>of</strong> genetic<br />

potential. Analyses <strong>of</strong> QTLs have<br />

revealed that loci controlling a<br />

quantitatively inherited trait do not<br />

contribute equally to the observed<br />

variation for the trait, <strong>and</strong> <strong>of</strong>ten a<br />

few QTLs explain most <strong>of</strong> the<br />

observed variation. In rice, QTLs<br />

for yield were identified in a wild,<br />

low yielding relative. After<br />

introgression into modern hybrid<br />

rice cultivars, yield increases <strong>of</strong><br />

17% compared to the original<br />

hybrid were observed. Based on<br />

the observed gains, Tanksley <strong>and</strong><br />

McCouch (1997) identified the need<br />

to more thoroughly evaluate exotic<br />

germplasm. Those accessions most<br />

distinct from modern cultivars may<br />

contain the highest number <strong>of</strong><br />

unexploited, potentially useful<br />

alleles.<br />

The comparative genetic<br />

mapping <strong>of</strong> cereal genomes has<br />

identified a vast amount <strong>of</strong><br />

conserved linearity <strong>of</strong> gene order<br />

(Devos <strong>and</strong> Gale, 1997). This<br />

observation will likely accelerate<br />

the application <strong>of</strong> QTLs in wheat,<br />

as well as aid in the identification<br />

<strong>of</strong> genes required for introgression<br />

from alien species. Considering the<br />

low number <strong>of</strong> loci tagged today in<br />

wheat, the problems related to<br />

developing a high density map for<br />

wheat (Snape, 1998), <strong>and</strong><br />

consequently the limited progress<br />

to identify QTLs for yield in wheat,<br />

we believe that the impact from<br />

this linearity on wheat<br />

improvement will be significant.<br />

Wheat has been successfully<br />

transformed for herbicide<br />

resistance <strong>and</strong> high molecular<br />

weight (HMW) glutenins, using<br />

both the ballistic <strong>and</strong> Agrobacterium<br />

tumefaciens systems (Cheng et al.,<br />

1997). Barro et al. (1997) inserted<br />

two additional HMW glutenin<br />

subunits, 1Ax1 <strong>and</strong> 1Dx5, <strong>and</strong><br />

observed a stepwise improvement<br />

<strong>of</strong> dough strength. Altpeter et al.<br />

(1996) introduced 1Ax1 into<br />

Bobwhite <strong>and</strong> increased total<br />

HMW glutenin subunit protein by<br />

71% over Bobwhite. However, the<br />

effects <strong>of</strong> transformation are not<br />

necessary additive as was shown<br />

by Blechl et al. (1998), who<br />

identified trangenics for HMW<br />

glutenins that also exhibited<br />

decreased accumulation due to<br />

transgene-mediated suppression.<br />

Conclusions<br />

The challenge to annually<br />

produce one billion tons <strong>of</strong> wheat<br />

within the next 25 years is<br />

formidable <strong>and</strong> can be met only by<br />

a concerted action <strong>of</strong> scientists<br />

involved in diverse disciplines<br />

(agronomy, pathology, physiology,<br />

biotechnology, breeding), as well as<br />

economics <strong>and</strong> politics. I am<br />

optimistic that this target will be<br />

met. Today, funds are directed from<br />

breeding towards biotechnology,<br />

<strong>of</strong>ten due simply to the novelty<br />

required for publication.<br />

Eventually, transformation may be<br />

a valuable technique to alter the<br />

performance <strong>of</strong> a genotype;<br />

however, at least during the next<br />

decade, the simple decision <strong>of</strong> a<br />

breeder in the field to “keep or<br />

discard” will contribute more to<br />

yield increase than any other<br />

approach. In conclusion, I agree<br />

with Ruttan (1993) who stated that<br />

“at least for the next two decades to<br />

come, progress through<br />

conventional breeding will remain<br />

the primary source <strong>of</strong> growth in<br />

crop <strong>and</strong> animal production.”<br />

References<br />

Allard, R.W. 1996. Genetic basis <strong>of</strong> the<br />

evolution <strong>of</strong> adaptedness in plants.<br />

Euphytica 92: 1-11.<br />

Altpeter, F.V., V. Vasil, V. Srivastava,<br />

<strong>and</strong> I.K. Vasil. 1996. Integration<br />

<strong>and</strong> expression <strong>of</strong> the highmolecular-weight<br />

glutenin subunit<br />

1Ax1 gene into wheat. Nature<br />

Biotechnology 14: 1155-1159.<br />

Barro, F., L. Rooke, F. Bekes, P. Gras,<br />

A.S. Tatham, R. Fido, P.A. Lazzeri,<br />

P.R. Shewry, <strong>and</strong> P. Barcelo. 1997.<br />

Transformation <strong>of</strong> wheat with high<br />

molecular weight subunit genes<br />

results in improved functional<br />

properties. Nature Biotechnology<br />

15: 1295-1299.


Blechl, A.E., B.S.B. Altenbach, H.Q.<br />

Le, P.W. Gras, F. Bekes <strong>and</strong> O. D.<br />

Anderson. 1998. Genetic<br />

transformation can be used to<br />

either increase or decrease levels <strong>of</strong><br />

wheat HMW-glutenin subunits.<br />

Abstracts <strong>of</strong> 21 st HRWWWW, Jan.<br />

28-30, Denver, CO. USDA-ARS,<br />

University <strong>of</strong> Nebraska, Lincoln,<br />

NE.<br />

Borlaug, N.E. 1995. Wheat Breeding<br />

at <strong>CIMMYT</strong>. Commemorating 50<br />

years <strong>of</strong> research in Mexico for<br />

global wheat improvement. Wheat<br />

Special Report No 29. pp. 4-6.<br />

Borlaug, N.E. 1966. Basic concepts<br />

which influence the choice <strong>of</strong><br />

methods for use in breeding for<br />

diverse resistance in cross<br />

pollinated <strong>and</strong> self pollinated crop<br />

plants. In: H.D. Gerold et al., eds.,<br />

Breeding Pest Resistant Trees.<br />

Pergamon Press, Oxford.<br />

Bramel-Cox, P.J., T. Barker, F. Zavala-<br />

Garcia, <strong>and</strong> J.D. Eastin. 1991.<br />

Selection <strong>and</strong> testing environments<br />

for improved performance under<br />

reduced-input conditions. In Plant<br />

breeding <strong>and</strong> sustainable<br />

agriculture: Considerations for<br />

objectives <strong>and</strong> methods. CSSA<br />

Special Publication No. 18. CSSA<br />

<strong>and</strong> ASA, Madison, USA.<br />

Braun, H.-J., W.H. Pfeiffer, <strong>and</strong> W.G.<br />

Pollmer. 1992. Environments for<br />

selecting widely adapted spring<br />

wheat. Crop Sci. 32 (6): 1420-1427.<br />

Braun, H.-J., S. Rajaram. <strong>and</strong> M. van<br />

Ginkel. 1996. <strong>CIMMYT</strong>’s approach<br />

to breeding for wide adaptation.<br />

Euphytica 92: 147-153.<br />

Braun, H.-J., H. Ekiz, V. Eser, M.<br />

Keser, H. Ketata, G. Marcucci, A.I.<br />

Morgounov, <strong>and</strong> N. Zencirci. 1998.<br />

Breeding priorities <strong>of</strong> winter wheat<br />

programs. In H.-J. Braun, F. Altay,<br />

W.E. Kronstad, S.P.S. Beniwal, <strong>and</strong><br />

A. McNab (eds.). Wheat: Prospects<br />

for Global Improvement. Proc. <strong>of</strong><br />

the 5 th Int. Wheat Conf., Ankara,<br />

Turkey. Developments in Plant<br />

Breeding, v. 6. Kluwer Academic<br />

Publishers. Dordrecht. p. 553-560.<br />

Brown, L.R. 1997. Can we raise grain<br />

yields fast enough? World Watch<br />

Jul/Aug 1997: 8-18.<br />

Bruns, R., <strong>and</strong> C.J. Peterson. 1998.<br />

Yield <strong>and</strong> stability factors<br />

associated with hybrid wheat.<br />

Euphytica 100: 1-5.<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 15<br />

Bull, J.K., M. Cooper, <strong>and</strong> K.E.<br />

Basford. 1994. A procedure for<br />

investigating the number <strong>of</strong><br />

genotypes required to provide a<br />

stable classification <strong>of</strong><br />

environments. Field Crops Res. 38:<br />

47-56.<br />

Byerlee, D., <strong>and</strong> P. Moya. 1993.<br />

Impacts <strong>of</strong> international wheat<br />

breeding research in the<br />

developing world, 1969-90.<br />

Mexico, D.F.: <strong>CIMMYT</strong>. 135pp.<br />

Caldwell, R.M., 1968. Breeding for<br />

general <strong>and</strong>/or specific plant<br />

disease resistance. In Proc. 3rd Int.<br />

Wheat Genetics Symp. Canberra,<br />

Australia.<br />

Calhoun, D.S., G. Gebeyehu, A.<br />

Mir<strong>and</strong>a, S. Rajaram, <strong>and</strong> M. van<br />

Ginkel. 1994. Choosing evaluation<br />

environments to increase wheat<br />

grain yield under drought<br />

conditions. Crop Sci. 34:673-678.<br />

Ceccarelli, S., M.M. Nachit, G.O.<br />

Ferrara, M.S.Mekni, M.Tahir, J.Van<br />

Leur, <strong>and</strong> J.P. Srivastava. 1987.<br />

Breeding strategies for improving<br />

cereal yield <strong>and</strong> stability under<br />

drought. In J.P. Srivastava, E.<br />

Porceddu, E. Acevedo, <strong>and</strong> S.<br />

Varma (eds.). Drought tolerance in<br />

winter cereals. John Wiley & Sons.<br />

New York. Pp. 101-114.<br />

Ceccarelli, S. 1989. Wide adaptation:<br />

How wide? Euphytica 40: 197-205.<br />

Cheng, M., J.E. Fry, S. Pang, H. Zhou,<br />

C.M. Hironaka, D.R. Duncan, T.W.<br />

Conner, <strong>and</strong> Y. Wan. 1997. Genetic<br />

transformation <strong>of</strong> wheat mediated<br />

by Agrobacterium tumefaciens. Plant<br />

Physiology 115: 971-980.<br />

<strong>CIMMYT</strong>. 1996. <strong>CIMMYT</strong> 1995/96<br />

World Wheat Facts <strong>and</strong> Trends.<br />

Mexico, D.F.<br />

Cooper, M., D.E. Byth, <strong>and</strong> D.R.<br />

Woodruff. 1994. An investigation<br />

<strong>of</strong> the grain yield adaptation <strong>of</strong><br />

<strong>CIMMYT</strong> wheat lines to water<br />

stress environments in<br />

Queensl<strong>and</strong>. II. Classification<br />

analysis. Agric. Res. 45:985-1002.<br />

Cox, T.S., R.K. Bequette, R.L. Bowden,<br />

<strong>and</strong> R.G. Sears. 1997. Grain yield<br />

<strong>and</strong> bread making quality <strong>of</strong><br />

wheat lines with the leaf rust<br />

resistance gene Lr41. Crop Science<br />

37: 154-161.<br />

Devos, K.M., <strong>and</strong> M.D. Gale. 1997.<br />

Comparative genetics in the<br />

grasses. Plant Molecular Biology<br />

35: 3-15.<br />

Duvick, D.N. 1990. Ideotype<br />

evolution <strong>of</strong> hybrid maize in the<br />

USA, 1930-1990. In ATTI<br />

Proceedings, Vol. II, II National<br />

maize conference; research,<br />

economy, environment. Grado<br />

(GO)- Italy September 19-20-21,<br />

1990. centro regionale per la<br />

sperimentazione agraria, pozzuolo<br />

del friuli edagrocole s.p.a.,<br />

Bologna, Italy. Pp. 557-570.<br />

Duvick, D.N. 1992. Genetic<br />

contributions to advances in yield<br />

<strong>of</strong> U.S. maize, Maydica 37:69-79.<br />

Duvick, D.N. 1996. Plant breeding, an<br />

evolutionary concept. Crop<br />

Science 36: 539-548.<br />

Eavis, R.M., S.E. Batchelor, F. Murray,<br />

<strong>and</strong> K.C. Walker. 1996. Hybrid<br />

breeding <strong>of</strong> wheat, barley <strong>and</strong> rye:<br />

Development to date <strong>and</strong> future<br />

prospects. Res. Rev. no. 35. Home-<br />

Grown <strong>Cereals</strong> Authority, London.<br />

Edmeades, G.O., J. Bolanos, H.R.<br />

Lafitte, S. Rajaram, W. Pfeiffer, <strong>and</strong><br />

R.A. Fischer. 1989. Traditional<br />

approaches to breeding for<br />

drought resistance in cereals. In:<br />

F.W.G. Baker (ed.) Drought<br />

resistance in cereals. ICSU, Paris,<br />

<strong>and</strong> CABI Publishing, Wallingford,<br />

UK. Pp. 27-52.<br />

Ehdaie, B., J.G. Waines, <strong>and</strong> A.E. Hall.<br />

1988. Differential responses <strong>of</strong><br />

l<strong>and</strong>race <strong>and</strong> improved spring<br />

wheat genotypes to stress<br />

environments. Crop Sci. 28:838-<br />

842.<br />

Fischer, R.A., D. Rees, K.D. Sayre, Z.<br />

Lu, A.G. Condon, A. Larque-<br />

Saavedra, <strong>and</strong> E. Zeiger. 1998.<br />

Wheat yield progress associated<br />

with higher stomatal conductance<br />

<strong>and</strong> higher photosynthetic rate,<br />

<strong>and</strong> cooler canopies. Crop Science<br />

38(6):1467-1475.<br />

Fritz, A.K., T.S. Cox, B.S. Gill, <strong>and</strong><br />

R.G. Sears. 1995. Molecular<br />

marker-facilitated analysis <strong>of</strong><br />

introgression in winter wheat x<br />

Triticum tauschii populations. Crop<br />

Science 35: 1691-1695.<br />

Gallais, A. 1989. Lines versus hybrids<br />

– The choice for the optimum type<br />

<strong>of</strong> variety. Vorträge f. Pflanzenz.<br />

16: 69-80. Paul Parey, Berlin.<br />

Hewstone M., C. 1997. Los cambios<br />

geneticos y agronomicos que<br />

incrementaron el rendimiento de<br />

trigo en Chile. Explor<strong>and</strong>o altos


16 Opening Remarks — S. Rajaram<br />

rendimientos de trigo. Oct. 20-23,<br />

1997. INIA <strong>and</strong> <strong>CIMMYT</strong>, La<br />

Estanzuela, Uruguay.<br />

ICARDA. 1993. Cereal Program.<br />

Annual Report for 1992.<br />

Jackson, P., M. Robertson, M. Cooper,<br />

<strong>and</strong> G. Hammer. 1996. The role <strong>of</strong><br />

physiological underst<strong>and</strong>ing in<br />

plant breeding: from a breeding<br />

perspective. Field Crops Research<br />

49: 11-37.<br />

Jlibene M., J.P. Gustafson, <strong>and</strong> S.<br />

Rajaram. 1992. A field disease<br />

evaluation method for selecting<br />

wheat resistant to Mycosphaerella<br />

graminicola. Plant Breeding 108, 26-<br />

32.<br />

Jordaan, J.P. 1996. Hybrid wheat.<br />

Advances <strong>and</strong> challenges. In M.P.<br />

Reynolds, S. Rajaram, <strong>and</strong> A.<br />

McNab (eds). Increasing Yield<br />

Potential in Wheat: Breaking the<br />

Barriers. Mexico, D.F.: <strong>CIMMYT</strong>. P.<br />

66-75.<br />

Kronstad, W.E. 1998. Agricultural<br />

development <strong>and</strong> wheat breeding<br />

in the 20 th century. In H.-J. Braun,<br />

F. Altay, W.E. Kronstad, S.P.S.<br />

Beniwal, <strong>and</strong> A. McNab (eds).<br />

Wheat: Prospects for Global<br />

Improvement. Proc. <strong>of</strong> the 5 th Int.<br />

Wheat Conf., Ankara, Turkey.<br />

Developments in Plant Breeding, v.<br />

6. Kluwer Academic Publishers.<br />

Dordrecht. p. 1-10.<br />

Matus-Tejos, I.A., 1993. Genetica de la<br />

resistencia a <strong>Septoria</strong> tritici en<br />

trigos harineros. Tesis de Maestria<br />

en Ciencia, Montecillo, Mexico.<br />

McGuire, S. 1997. The effects <strong>of</strong><br />

privatization on winter-wheat<br />

breeding in the UK. Biotechnology<br />

<strong>and</strong> Development Monitor 33<br />

(Dec.): 8-11.<br />

McIntosh, R.A. 1993. Catalogue <strong>of</strong><br />

gene symbols for wheat. In Z.S. Li<br />

<strong>and</strong> Z.Y. Xin (eds.). Proc. 8 th IWGS,<br />

Beijing. China Agric. Scientech<br />

Press, Beijing. p 1333-1500.<br />

Mitchell, D.O., M.D. Onco, <strong>and</strong> R.D.<br />

Duncan. 1997. The world food<br />

outlook. Cambridge University<br />

Press.<br />

Morris, M.L., A. Belaid, <strong>and</strong> D.<br />

Byerlee. 1991. Wheat <strong>and</strong> barley<br />

production in rainfed marginal<br />

environments <strong>of</strong> developing word.<br />

Part I <strong>of</strong> 1990-91 <strong>CIMMYT</strong> world<br />

wheat facts <strong>and</strong> trends. Mexico,<br />

D.F.: <strong>CIMMYT</strong>. 51 pp.<br />

Mujeeb-Kazi, A., V. Rosas, <strong>and</strong> S.<br />

Roldan. 1996. Conservation <strong>of</strong> the<br />

genetic variation <strong>of</strong> Triticum<br />

tauschii (Coss.) Schmalh. (Aegilops<br />

squarrosa auct. non L.) in synthetic<br />

hexaploid wheats (T. turgidum L. s.<br />

lat. x T. tauschii; 2n=6x=42,<br />

AABBDD) <strong>and</strong> its potential<br />

utilization for wheat improvement.<br />

Genetic Resources <strong>and</strong> Crop<br />

Evolution 43: 129-134.<br />

Niederhauser, J.S., J. Servantes, <strong>and</strong> L.<br />

Servin. 1954. Late blight in Mexico<br />

<strong>and</strong> its implications. Phytopath.<br />

44:406-408.<br />

Ortiz-Monasterio, J.I., K.D. Sayre, S.<br />

Rajaram, <strong>and</strong> M. McMahon. 1995.<br />

Genetic progress <strong>of</strong> <strong>CIMMYT</strong>’s<br />

bread wheat germplasm under<br />

different levels <strong>of</strong> notrogen. I.<br />

Grain yield <strong>and</strong> nitrogen use<br />

eficiency.<br />

Ortiz-Monasterio, J.I., K.D. Sayre, S.<br />

Rajaram <strong>and</strong> M. McMahon. 1997.<br />

Genetic progress in wheat yield<br />

<strong>and</strong> nitrogen use efficiency under<br />

four nitrogen rates. Crop Science<br />

37: 898-904.<br />

Peterson, C.J., J.M. M<strong>of</strong>fatt, <strong>and</strong> J.R.<br />

Erickson. 1997. Yield stability <strong>of</strong><br />

hybrid vs. pureline hard red<br />

winter wheats in regional<br />

performance trials. Crop Science<br />

37: 116-120.<br />

Pickett, A.A. 1993. Hybrid wheat –<br />

results <strong>and</strong> problems. Advances in<br />

Plant Breeding. No. 15. Paul Parey,<br />

Scientific Publishers. Berlin.<br />

Pickett, A.A., <strong>and</strong> N.W. Galwey. 1997.<br />

A further evaluation <strong>of</strong> hybrid<br />

wheat. Plant Varieties <strong>and</strong> Seeds<br />

10: 15-32.<br />

Pingali, P.L., <strong>and</strong> P.W. Heisey. 1997.<br />

Cereal crop productivity in<br />

developing countries: Past trends<br />

<strong>and</strong> future prospects. In Global<br />

Agric. Sci. Policy for the 21 st<br />

Century. 26-28 August, 1996.<br />

Melbourne, Australia.<br />

Rabinovich, S.V. 1998. Importance <strong>of</strong><br />

wheat-rye translocations for<br />

breeding modern cultivars <strong>of</strong><br />

Triticum aestivum L. Euphytica 100:<br />

323-340.<br />

Rajaram, S., R.P. Singh, <strong>and</strong> E. Torres.<br />

1988. Current <strong>CIMMYT</strong><br />

approaches in breeding wheat for<br />

rust resistance. In: S. Rajaram <strong>and</strong><br />

N. W. Simmonds, eds., Breeding<br />

strategies for resistance to the rusts<br />

<strong>of</strong> wheat. Mexico D.F.: <strong>CIMMYT</strong>.<br />

pp. 101-118.<br />

Rajaram, S., R. Villareal, <strong>and</strong> A.<br />

Mujeeb-Kazi. 1990. Global impact<br />

<strong>of</strong> 1B/1R spring wheats.<br />

Agronomy Abstracts, ASA,<br />

Madison, USA.<br />

Rajaram, S., M. van Ginkel, <strong>and</strong> R.A.<br />

Fischer. 1994. <strong>CIMMYT</strong>’s wheat<br />

breeding mega-environments<br />

(ME). In: Proceedings <strong>of</strong> the 8th<br />

International wheat genetic<br />

symposium, July 19-24, 1993.<br />

Beijing, China.<br />

Rajaram, S., <strong>and</strong> M. van Ginkel. 1995.<br />

Wheat breeding methodology:<br />

International perspectives. In:<br />

Hard Red Winter Wheat Workers<br />

Conference, 20. Oklahoma City,<br />

OK, USA. 25-25 Jan., 1995. Guenzi,<br />

A.C., Kuehn, M., <strong>and</strong> Hunger,<br />

R.M. (eds.). Oklahoma State<br />

University. pp. 1-15.<br />

Rajaram, S. 1995. Yield stability <strong>and</strong><br />

avoiding genetic vulnerability in<br />

bread wheat. In: S. Rajaram <strong>and</strong> G.<br />

Hettel, eds.Wheat Breeding at<br />

<strong>CIMMYT</strong>. Commemorating 50<br />

years <strong>of</strong> research in Mexico for<br />

global wheat improvement. Wheat<br />

Special Report No 29. Mexico, D.F.:<br />

<strong>CIMMYT</strong>. pp. 11-15.<br />

Rasmusson, D.C. 1996. Germplasm is<br />

paramount. In M.P. Reynolds, S.<br />

Rajaram, <strong>and</strong> A. McNab (eds.).<br />

Increasing Yield Potential in<br />

Wheat: Breaking the Barriers.<br />

Mexico, D.F.: <strong>CIMMYT</strong>. pp. 28-35.<br />

Rasmusson, D.C., <strong>and</strong> R.L. Phillips.<br />

1997. Plant breeding progress <strong>and</strong><br />

genetic diversity from de novo<br />

variation <strong>and</strong> elevated epistasis.<br />

Crop Science 37: 303-310.<br />

Rees, D., K.D. Sayre, E. Acevedo, T.N.<br />

Sanchez, Z. Lu, E. Zeiger, <strong>and</strong> L.<br />

Limon. 1993. Canopy temperatures<br />

<strong>of</strong> wheat: Relationship with yield<br />

<strong>and</strong> potential as a technique for<br />

early generation selection. Wheat<br />

Special Report No. 10. Mexico,<br />

D.F.: <strong>CIMMYT</strong>.<br />

Rejesus, R.M., M van Ginkel, <strong>and</strong> M.<br />

Smale. 1996. Wheat Breeder’s<br />

Perspectives <strong>of</strong> Genetic Diversity<br />

<strong>and</strong> Germplasm Use. Wheat<br />

Special Report 40. Mexico D.F.:<br />

<strong>CIMMYT</strong>.<br />

Reynolds, M.P., M. Balota, M.I.B.<br />

Delgado, I. Amani, <strong>and</strong> R.A.<br />

Fischer. 1994. Physiological <strong>and</strong><br />

morphological traits associated<br />

with spring wheat yield under hot,<br />

irrigated conditions. Australian J.<br />

<strong>of</strong> Plant Physiology 21:717-730.


Reynolds, M.P., R.P. Singh, A.<br />

Ibrahim, O.A.A. Ageeb, A. Larque-<br />

Saavedra, <strong>and</strong> J.S. Quick. 1998.<br />

Evaluating physiological traits to<br />

complement empirical selection<br />

for wheat in warm environments.<br />

Euphytica 100: 85-94.<br />

Rosegrant, M.W., A. Agcaoili-<br />

Sombilla, <strong>and</strong> N. Perez. 1995.<br />

Global Food Projections to 2020.<br />

Discussion paper 5. Washington,<br />

D.C.: IFPRI.<br />

Ruttan, V.W. 1993. Research to meet<br />

crop production needs: Into the<br />

21 st century. In D.R. Buxton et al.<br />

(eds.) International Crop Science<br />

Congress I. CSSA, Madison,<br />

Wisconsin, USA. pp. 3-10.<br />

Sayre, K.D., S. Rajaram, <strong>and</strong> R.A.<br />

Fischer. 1997. Yield potential<br />

progress in short bread wheat in<br />

Northwest Mexico. Crop Science<br />

37: 36-42.<br />

Schlegel, R. 1997a. Currant list <strong>of</strong><br />

wheats with rye introgressions <strong>of</strong><br />

homoeologous group 1. Wheat<br />

Information Service 84: 64-69.<br />

Schlegel, R. 1997b. About the origin <strong>of</strong><br />

1RS.1BL wheat rye chromosome<br />

translocations from Germany.<br />

Plant Breeding 116: 537-540.<br />

Sears, R.G. 1998. Strategies for<br />

improving wheat grain yield. In<br />

H.-J. Braun, F. Altay, W.E.<br />

Kronstad, S.P.S. Beniwal <strong>and</strong> A.<br />

McNab (eds.). Wheat: Prospects<br />

for Global Improvement. Proc. <strong>of</strong><br />

the 5 th Int. Wheat Conf., Ankara,<br />

Turkey. Developments in Plant<br />

Breeding, v. 6. Kluwer Academic<br />

Publishers. Dordrecht. pp. 17-22.<br />

Singh, R.P., <strong>and</strong> S. Rajaram, 1992.<br />

Genetics <strong>of</strong> adult plant resistance<br />

to leaf rust in “Frontana” <strong>and</strong> three<br />

<strong>CIMMYT</strong> wheats. Genome 35: 24-<br />

31.<br />

Singh, R.P. 1992a. Association<br />

between gene Lr34 for leaf rust<br />

resistance <strong>and</strong> leaf tip necrosis in<br />

wheat. Crop Sci. 32: 874-878.<br />

Historical Aspects <strong>and</strong> Future Challenges <strong>of</strong> an International Wheat Program 17<br />

Singh, R.P. 1992b. Genetic association<br />

<strong>of</strong> leaf rust resistance gene Lr34<br />

with adult plant resistance to<br />

stripe rust in bread wheat.<br />

Phytopathology 82: 835-838.<br />

Singh, R.P., S. Rajaram, J. Montoya,<br />

<strong>and</strong> G. Fuentes-Davila, 1995.<br />

Genetic analysis <strong>of</strong> resistance to<br />

Karnal bunt (Tilletia indica Mitra)<br />

in bread wheat. Euphytica 81:117-<br />

120.<br />

Singh, R.P., J. Huerta-Espino, S.<br />

Rajaram, <strong>and</strong> J. Crossa. 1998.<br />

Agronomic effects from<br />

chromosome translocations<br />

7DL.7Ag <strong>and</strong> 1BL.1RS in spring<br />

wheat. Crop Science 38: 27-33.<br />

Smale, M., <strong>and</strong> T. McBride, 1996.<br />

Underst<strong>and</strong>ing global trends in the<br />

use <strong>of</strong> wheat diversity <strong>and</strong><br />

international flows <strong>of</strong> wheat<br />

genetic resources. Part 1: <strong>CIMMYT</strong><br />

1995/96 World Wheat Facts <strong>and</strong><br />

Trends. Mexico D.F.: <strong>CIMMYT</strong>.<br />

Tanksley, S.D., <strong>and</strong> S.R. McCouch.<br />

1997. Seed banks <strong>and</strong> molecular<br />

maps: Unlocking genetic potential<br />

for the wild. Science 277: 1063-<br />

1066.<br />

Traxler, G., J. Falck-Zepeda, J.I. Ortiz<br />

Monasterio, <strong>and</strong> K.D. Sayre. 1995.<br />

Production risk <strong>and</strong> the evolution<br />

<strong>of</strong> varietal technology. American<br />

Journal Agricultural Economics 77:<br />

1-7.<br />

Uddin, N., B.F. Carver, <strong>and</strong> A.C.<br />

Clutter. 1992. Genetic analysis <strong>and</strong><br />

selection for wheat yield in<br />

drought-stressed <strong>and</strong> irrigated<br />

environments. Euphytica 62:89-96.<br />

Van Ginkel, M., D.S. Calhoun, G.<br />

Gebeyehou, A.Mir<strong>and</strong>a, C. Tianyou,<br />

R. Pargas Lara, R.M.<br />

Trethowan, K.D. Sayre, J. Crossa,<br />

<strong>and</strong> S. Rajaram. 1998. Plant traits<br />

related to yield <strong>of</strong> wheat in early,<br />

late, or continuous drought<br />

conditions. Euphytica 100:109-121.<br />

Villareal, R.L., G. Fuentes-Davila, <strong>and</strong><br />

A. Mujeeb-Kazi. 1995. Synthetic<br />

hexaploids x Triticum aestivum<br />

advanced derivatives resistant to<br />

Karnal bunt (Tilletia indica Mitra).<br />

Cer. Res. Commun. 23:127-132.<br />

Villareal, R.L. 1995. Exp<strong>and</strong>ing the<br />

genetic base <strong>of</strong> <strong>CIMMYT</strong> bread<br />

wheat germplasm. In: Rajaram S.<br />

<strong>and</strong> G. Hettel, eds.Wheat Breeding<br />

at <strong>CIMMYT</strong>. Commemorating 50<br />

years <strong>of</strong> research in Mexico for<br />

global wheat improvement. Wheat<br />

Special Report No 29. Mexico, D.F.:<br />

<strong>CIMMYT</strong>. pp. 16-21.<br />

Villareal. R.L., O. Bañuelos, J. Borja,<br />

A. Mujeeb-Kazi, <strong>and</strong> S. Rajaram.<br />

1997. Agronomic performance <strong>of</strong><br />

some advanced derivatives <strong>of</strong><br />

synthetic hexaploids (Triticum<br />

turgidum x T. tauschii). Annual<br />

Wheat Newsletter 43: 175-176.<br />

Villareal, R.L., E. Del Toro, A. Mujeeb-<br />

Kazi, <strong>and</strong> S. Rajaram. 1995. The<br />

1BL/1RS chromosome<br />

translocation effect on yield<br />

characterization in a Triticum<br />

aestivum L. cross. Plant Breeding<br />

114:497-500.<br />

Worl<strong>and</strong>, A.J., M.L. Appendino, <strong>and</strong><br />

E.J. Sayers, 1994. The distribution,<br />

in European winter wheats, <strong>of</strong><br />

genes that influence ecoclimatic<br />

adaptability whilst determining<br />

photoperiodic insensitivity <strong>and</strong><br />

plant height. Euphytica 80(3):219-<br />

228.<br />

Worl<strong>and</strong>, A.J., A. Börner, V. Koyrun,<br />

W.M. Li, S. Petrovic, <strong>and</strong> E.J.<br />

Sayers. 1998. The influence <strong>of</strong><br />

photoperiod genes on the<br />

adaptability <strong>of</strong> European winter<br />

wheats. Euphytica 100: 385-394.<br />

Young C., <strong>and</strong> .J. Frey, 1994. Grainyield<br />

characteristics <strong>of</strong> oat lines<br />

surviving uniform <strong>and</strong> shuttle<br />

selection strategies. Euphytica<br />

76:63-71.<br />

Zavala-Garcia, F., P.J. Bramel-Cox,<br />

J.D.Eastin, M.D. Witt, <strong>and</strong> D.J.<br />

Andrews. 1992. Increasing the<br />

efficiency <strong>of</strong> crop selection for<br />

unpredictable environments. Crop<br />

Sci. 32:51-57.


Session 1: Pathogen Biology<br />

Biology <strong>of</strong> the <strong>Septoria</strong>/<strong>Stagonospora</strong> Pathogens: An<br />

Overview<br />

A.L. Scharen<br />

Department <strong>of</strong> Plant Sciences, Montana State University, Bozeman, Montana, USA<br />

Abstract<br />

More than 2000 form-species <strong>of</strong> fungi, mostly plant parasites, comprise the genera <strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong>. The two<br />

most important pathogens in wheat production are Mycosphaerella graminicola (<strong>Septoria</strong> tritici) <strong>and</strong> Phaeosphaeria<br />

nodorum (<strong>Stagonospora</strong> nodorum). Primary inoculum for the wheat diseases caused by these pathogens is most <strong>of</strong>ten airborne<br />

ascospores, but may also be wind- <strong>and</strong> rain-borne conidia. Field diagnoses may be augmented <strong>and</strong> made more exact by<br />

use <strong>of</strong> rapid immunological tests <strong>and</strong> molecular genetic methods. Infection processes <strong>of</strong> S. tritici <strong>and</strong> S. nodorum are similar,<br />

but penetration by S. tritici is known only via stomata. Patterns <strong>of</strong> occurrence <strong>of</strong> the two pathogens have changed<br />

dramatically in recent years. <strong>Septoria</strong> tritici has become more important in northern Europe, <strong>and</strong> S. nodorum incidence has<br />

increased in parts <strong>of</strong> North Africa. Changes in cultivars <strong>and</strong> cultural practices are thought to be responsible for the shifts in<br />

pathogens <strong>and</strong> diseases. Genetics <strong>of</strong> resistance in the wheat host <strong>and</strong> virulence in the pathogen populations continues to be<br />

unclear. Some gene-for-gene interactions have been shown, but in field situations resistance is generally observed as nonspecific<br />

<strong>and</strong> pathogen populations vary most in aggressiveness.<br />

“Scientists build on foundations<br />

laid by their predecessors…, but<br />

they show great reluctance to<br />

inspect those foundations”<br />

(Ainsworth, 1965). Such reluctance<br />

was evident for many years in the<br />

case <strong>of</strong> the <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

diseases <strong>of</strong> cereal grains.<br />

Nineteenth <strong>and</strong> early twentieth<br />

century workers in Europe <strong>and</strong><br />

North America described the<br />

pathogens <strong>and</strong> the diseases they<br />

cause in considerable detail. But,<br />

only in the years since the advent<br />

<strong>of</strong> cultivars bred for dwarf stature<br />

<strong>and</strong> high yield under conditions <strong>of</strong><br />

intensive culture have the <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases been<br />

recognized as having major<br />

impacts upon yield <strong>and</strong> quality.<br />

The causal organisms are still<br />

called by several names, as are the<br />

diseases. A concise summary <strong>of</strong> the<br />

latest information on taxonomy<br />

<strong>and</strong> nomenclature was published<br />

in 1997 (Cunfer, 1997) <strong>and</strong> 1999<br />

(Cunfer <strong>and</strong> Ueng, 1999). I believe<br />

that we should urge from this<br />

venue, as did those attending the<br />

Fourth International Workshop on<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong> (July 4-7, 1994,<br />

IHAR, Radzikow, Pol<strong>and</strong>), that all<br />

workers accept <strong>and</strong> use the latest<br />

nomenclature.<br />

More than 2000 form-species <strong>of</strong><br />

fungi, most <strong>of</strong> them plant parasites,<br />

have been assigned to the genera<br />

<strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong>. Two <strong>of</strong><br />

these that attack wheat are <strong>of</strong> most<br />

concern to us. Losses <strong>of</strong> potential<br />

yield world-wide from just these<br />

two are estimated in the millions <strong>of</strong><br />

metric tons <strong>of</strong> grain <strong>and</strong> billions <strong>of</strong><br />

U.S. dollars each year. The common<br />

names <strong>of</strong> the diseases are septoria<br />

tritici blotch <strong>and</strong> stagonospora<br />

nodorum blotch. Classification <strong>and</strong><br />

19<br />

nomenclature <strong>of</strong> these <strong>and</strong> other<br />

pathogens <strong>of</strong> cereals are given in<br />

Table 1 (Eyal, 1999). The <strong>Septoria</strong><br />

anamorphic states have a<br />

teleomorphic state assigned to the<br />

genus Mycosphaerella, while the<br />

<strong>Stagonospora</strong> states have<br />

teleomorphic forms in the genus<br />

Phaeosphaeria. The teleomorphs are<br />

very important because they<br />

furnish the primary inoculum for<br />

disease development under specific<br />

environmental conditions.<br />

Several periods <strong>of</strong> ascospore<br />

deposition (both Mycosphaerella <strong>and</strong><br />

Phaeosphaeria) from the atmosphere<br />

during the months <strong>of</strong> August to<br />

October in the northern<br />

hemisphere <strong>and</strong> February to April<br />

in the southern hemisphere have<br />

recently been shown to have critical<br />

importance in epidemic<br />

establishment (Arseniuk et al.,


20<br />

Session 1 — A.L. Scharen<br />

Table 1. Classification <strong>and</strong> nomenclature <strong>of</strong> the <strong>Septoria</strong> spp. <strong>and</strong> <strong>Stagonospora</strong> spp. fungi on small grain cereals. a<br />

Genus Teleomorph Anamorph Common name Host<br />

<strong>Septoria</strong> spp. Mycosphaerella <strong>Septoria</strong> tritici <strong>Septoria</strong> tritici Wheat<br />

graminicola blotch <strong>of</strong> wheat<br />

1998; Shaw <strong>and</strong> Royle, 1989). The<br />

anamorphic conidia, also called<br />

pycnidiospores, are most important<br />

as secondary inoculum locally as<br />

the crop is growing <strong>and</strong> are<br />

disseminated mainly by rain<br />

splash.<br />

Sources <strong>of</strong> primary inoculum in<br />

areas where the teleomorph is not<br />

known remain a matter <strong>of</strong><br />

controversy <strong>and</strong> speculation,<br />

particularly in the case <strong>of</strong> septoria<br />

tritici blotch. Both S. nodorum <strong>and</strong><br />

S. tritici are found parasitizing a<br />

wide range <strong>of</strong> graminaceous hosts.<br />

(Krupinsky, 1994; Sprague, 1950).<br />

Several species <strong>of</strong> grasses have<br />

been suspected as alternative hosts<br />

<strong>and</strong> inoculum sources, but the<br />

question is yet unresolved. Conidia<br />

from plant debris may act as<br />

primary inoculum for disease<br />

development in some cases. The<br />

fact remains to puzzle us that no<br />

case has been reported in which<br />

septoria tritici blotch <strong>and</strong>/or<br />

stagonospora nodorum blotch<br />

failed to appear because <strong>of</strong> a lack <strong>of</strong><br />

primary inoculum when a<br />

- b <strong>Septoria</strong> tritici f. avenae Oats<br />

- <strong>Septoria</strong> tritici f. holci Holcus<br />

- <strong>Septoria</strong> tritici f. lolicola Lolium<br />

- <strong>Septoria</strong> passerinii Speckled leaf blotch<br />

<strong>of</strong> barley<br />

Barley<br />

- <strong>Septoria</strong> secalis Leaf spot <strong>of</strong> rye Rye<br />

<strong>Stagonospora</strong> spp. Phaeosphaeria <strong>Stagonospora</strong> nodorum <strong>Stagonospora</strong> nodorum Wheat<br />

nodorum blotch <strong>of</strong> wheat<br />

Phaeosphaeria <strong>Stagonospora</strong> avenae Oats<br />

avenaria f. sp. avenae<br />

Phaeosphaeria <strong>Stagonospora</strong> avenae Oats, wheat<br />

a (7)<br />

avenaria<br />

f. sp. triticea<br />

f. sp. triticea <strong>and</strong> triticale<br />

b Teleomorphic stages not found.<br />

susceptible crop <strong>and</strong> favorable<br />

environmental conditions<br />

prevailed.<br />

As late as the 1960s, diagnosis<br />

<strong>of</strong> <strong>Septoria</strong> diseases on wheat was<br />

not well developed even among<br />

plant pathologists <strong>and</strong> plant<br />

breeders. Leaf chlorosis <strong>and</strong><br />

necrosis was <strong>of</strong>ten viewed as part<br />

<strong>of</strong> the natural process <strong>of</strong><br />

maturation. Any <strong>of</strong> several leaf<br />

spotting pathogens could have<br />

been present <strong>and</strong> contributing to<br />

leaf death. When <strong>Septoria</strong> was<br />

recognized, it was <strong>of</strong>ten called<br />

“head septoria” which we know<br />

now as stagonospora nodorum<br />

blotch <strong>and</strong> “leaf septoria” which<br />

we know as septoria tritici blotch.<br />

Both <strong>of</strong>ten occur together <strong>and</strong> with<br />

other pathogens, <strong>and</strong> both can<br />

infect <strong>and</strong> cause symptoms on all<br />

parts <strong>of</strong> the wheat plant.<br />

Field diagnosis on the basis <strong>of</strong><br />

symptomology is regularly done,<br />

but <strong>of</strong>ten laboratory backup, with<br />

microscopic examination <strong>of</strong> spores,<br />

is necessary for accurate diagnosis.<br />

As can be seen in Table 2 (Eyal,<br />

1997), spore size <strong>and</strong> appearance<br />

may overlap, so some doubts may<br />

remain even after microscopic<br />

examination. Rapid tests have been<br />

developed that use immunological<br />

techniques mainly for early<br />

diagnosis in intensive production<br />

areas where chemical control is<br />

commonly used. Molecular genetic<br />

methods have been used recently<br />

to show similarities <strong>and</strong> differences<br />

between species <strong>and</strong> forma speciales<br />

(Arseniuk et al., 1997; Ueng et al.,<br />

1998.). The karyotype <strong>of</strong> S. nodorum<br />

suggests 14-19 chromosomes<br />

having approximately 0.5-3.5<br />

megabase pairs (Cooley <strong>and</strong> Caten,<br />

1991). McDonald <strong>and</strong> Martinez<br />

(1991) determined 14-16 b<strong>and</strong>s<br />

believed to correspond to<br />

chromosomes <strong>of</strong> 0.33-3.5 megabase<br />

pairs in S. tritici.<br />

Regardless <strong>of</strong> the fact that<br />

morphologic <strong>and</strong> genetic<br />

differences <strong>of</strong> considerable<br />

magnitude exist between S.<br />

nodorum <strong>and</strong> S. tritici, histological<br />

studies have shown that the


Table 2. The <strong>Septoria</strong> tritici/<strong>Stagonospora</strong> nodorum pathogens <strong>of</strong> wheat.<br />

Pycnidium Pycnidiospore Number Chromosome<br />

Asexual state (µm) (µm) <strong>of</strong> septa number Lesion<br />

<strong>Septoria</strong> tritici 60-200 35-98 x 1-3 3-5 14-19 a Irregular to<br />

rectangular,<br />

elongated<br />

between veins<br />

<strong>Stagonospora</strong> 160-210 15-32 x 2-4 0-3 14-16b Lens shaped,<br />

nodorum with chlorotic<br />

border<br />

Pseudothecium Ascospore Number<br />

Sexual state (µm) (µm) <strong>of</strong> cells<br />

Mycosphaerella<br />

graminicola 70-100 10-15 x 2-3 2<br />

Phaeosphaeria<br />

nodorum 120-200 23-32 x 4-6 4<br />

a McDonald <strong>and</strong> Martinez (1991).<br />

b Cooley <strong>and</strong> Caten (1991).<br />

infection process <strong>and</strong> the<br />

production <strong>of</strong> pycnidia in the host<br />

leaf bear some remarkable<br />

similarities (Karjalainen <strong>and</strong><br />

Lounatmaa, 1986; Kema et al.,<br />

1996b; Straley, 1979). Usually all<br />

observed conidia <strong>of</strong> S. tritici on leaf<br />

surfaces <strong>of</strong> either resistant or<br />

susceptible cultivars germinated<br />

<strong>and</strong> produced germ tubes. The<br />

same can be said in the case <strong>of</strong> S.<br />

nodorum, but in one case (T.<br />

aestivum Manitou, CI 13775<br />

resistant spring wheat), significant<br />

differences were found in<br />

germination <strong>of</strong> conidia on leaves <strong>of</strong><br />

susceptible <strong>and</strong> resistant cultivars<br />

(Straley, 1979). Hyphae <strong>of</strong> both<br />

species grow extensively over leaf<br />

surfaces, branching <strong>and</strong> <strong>of</strong>ten<br />

crossing stomata <strong>and</strong> guard cells<br />

without entering them.<br />

According to Kema et al.<br />

(1996b), penetrations <strong>of</strong> S. tritici<br />

were strictly stomatal <strong>and</strong> no direct<br />

penetrations <strong>of</strong> the cuticle were<br />

observed. Appressorial structures<br />

were seen in both pathogens, not<br />

associated with particular<br />

anatomical features, but most <strong>of</strong>ten<br />

in epidermal cracks. In the case <strong>of</strong><br />

S. nodorum, direct epidermal cell<br />

penetrations under appressorial<br />

structures could not be observed<br />

with the light microscope, but were<br />

confirmed with the scanning<br />

electron microscope (SEM).<br />

Stomatal penetration by S. nodorum<br />

was readily observed 72 h after<br />

inoculation. Stomata were<br />

penetrated in either the open or<br />

closed condition, with or without<br />

appressoria (Harrower, 1978;<br />

Straley, 1979).<br />

Reported patterns <strong>of</strong> occurrence<br />

<strong>of</strong> S. nodorum <strong>and</strong> S. tritici have<br />

changed dramatically during the<br />

past 15 years. In the early 1980s, S.<br />

nodorum was considered most<br />

important in northern Europe,<br />

eastern USA <strong>and</strong> Western Australia,<br />

while S. tritici was most prevalent<br />

in Mediterranean climates <strong>and</strong> the<br />

Great Plains <strong>of</strong> North America.<br />

Shifts to a greater importance <strong>of</strong> S.<br />

tritici have occurred in the UK <strong>and</strong><br />

are continuing to progress in<br />

central Europe. Some <strong>of</strong> the<br />

Biology <strong>of</strong> the <strong>Septoria</strong>/<strong>Stagonospora</strong> Pathogens: An Overview 21<br />

contributing factors in the UK may<br />

have been host susceptibility,<br />

earlier sowing, increased N<br />

fertilization <strong>and</strong> high summer<br />

rainfall as well as resistance to<br />

certain fungicides (Bayles, 1991).<br />

An interesting situation that<br />

could benefit from a systematic<br />

region-wide study has been<br />

reported from the Maghreb in<br />

North Africa. Populations <strong>of</strong> S.<br />

tritici particularly adapted to durum<br />

wheats have evolved, along with<br />

populations <strong>of</strong> S. nodorum more<br />

adapted to aestivum wheats. This<br />

phenomonon has become<br />

particularly obvious in Morocco,<br />

where a decided increase in<br />

plantings <strong>of</strong> bread wheats has<br />

occurred in recent years (Jlibene et<br />

al., 1995; Saadoui, 1987).<br />

Where sexual reproduction<br />

plays a role in epidemics, new<br />

virulence combinations can be<br />

selected by host virulence genes,<br />

<strong>and</strong> recombination <strong>of</strong> virulence<br />

genes will overcome “pyramids” <strong>of</strong><br />

resistance genes (McDonald, 1997).<br />

Ahmed et al. (1996) concluded that<br />

S. tritici populations adapt to host<br />

cultivars <strong>and</strong> that susceptible<br />

cultivars tend to select higher levels<br />

<strong>of</strong> pathogen aggressiveness in the<br />

field. Although evidence has been<br />

accumulated over a long period for<br />

specific gene-for-gene relationships<br />

in the S. tritici – wheat system (Eyal<br />

et al., 1985; Kema et al., 1996a), the<br />

question remains whether<br />

distinctive races can be identified<br />

with specific virulence for<br />

particular cultivars <strong>and</strong> whether<br />

the interaction indicates a gene-forgene<br />

system in the S. tritici — <strong>and</strong><br />

S. nodorum – wheat pathosystem<br />

(Johnson, 1992).


22<br />

Session 1 — A.L. Scharen<br />

The bulk <strong>of</strong> evidence continues<br />

to indicate that in S. nodorum —<br />

<strong>and</strong> S. tritici – wheat systems, genefor-gene<br />

interactions probably<br />

occur when selected cultures are<br />

inoculated on specific host plants<br />

under controlled conditions, but<br />

when population confronts<br />

population in the field, resistance<br />

to infection in wheat is generally<br />

non-specific <strong>and</strong> aggressiveness is<br />

the principal attribute <strong>of</strong> pathogen<br />

populations.<br />

References<br />

Ahmed, H.U., Mundt, C.C., H<strong>of</strong>fer,<br />

M.E., <strong>and</strong> Coakley, S.M. 1996.<br />

Selective influence <strong>of</strong> wheat<br />

cultivars on pathogenicity <strong>of</strong><br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici).<br />

Phytopathology 86:454-458.<br />

Ainsworth, G.C. 1965. Historical<br />

introduction to mycology. In: The<br />

Fungi, an Advanced Treatise. G.C.<br />

Ainsworth <strong>and</strong> A.S. Sussman<br />

(eds.). New York, Academic Press.<br />

748 pp.<br />

Arseniuk, E., Cunfer, B.M., Mitchell,<br />

S., <strong>and</strong> Kresovich, S. 1997.<br />

Characterization <strong>of</strong> genetic<br />

similarities among isolates <strong>of</strong><br />

<strong>Stagonospora</strong> spp. <strong>and</strong> <strong>Septoria</strong><br />

tritici by AFLP analysis.<br />

Phytopathology 87 (Suppl) S5.<br />

Arseniuk, E., Goral, T., <strong>and</strong> Scharen,<br />

A.L. 1998. Seasonal patterns <strong>of</strong><br />

spore dispersal <strong>of</strong> Phaeosphaeria<br />

spp. <strong>and</strong> <strong>Stagonospora</strong> spp. Plant<br />

Disease 82:187-194.<br />

Bayles, R.A. 1991. Varietal resistance<br />

as a factor contributing to the<br />

increased importance <strong>of</strong> <strong>Septoria</strong><br />

tritici Rob. <strong>and</strong> Desm. In the UK<br />

wheat crop. Plant Varieties <strong>and</strong><br />

Seed 4:177-183.<br />

Cooley, R.N., <strong>and</strong> Caten, C.E. 1991.<br />

Variation in electrophoretic<br />

karyotype between strains <strong>of</strong><br />

<strong>Septoria</strong> nodorum. Molecular <strong>and</strong><br />

General Genetics 228:17-23.<br />

Cunfer, B.M. 1997. Taxonomy <strong>and</strong><br />

nomenclature <strong>of</strong> <strong>Septoria</strong> <strong>and</strong><br />

<strong>Stagonospora</strong> species on small grain<br />

cereals. Plant Disease 81:427-428.<br />

Cunfer, B.M., <strong>and</strong> Ueng, P.P. 1999.<br />

Taxonomy <strong>and</strong> identification <strong>of</strong><br />

<strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong> species<br />

on small grain cereals. Ann. Rev. <strong>of</strong><br />

Phytopathology (in press).<br />

Eyal, Z., Scharen, A.L., Huffman,<br />

M.D., <strong>and</strong> Prescott, J.M. 1985.<br />

Global insights into virulence<br />

frequencies <strong>of</strong> Mycosphaerella<br />

graminicola. Phytopathology<br />

75:1456-1462.<br />

Eyal, Z. 1999. <strong>Septoria</strong> <strong>and</strong><br />

<strong>Stagonospora</strong> diseases <strong>of</strong> cereals: A<br />

comparative perspective.<br />

Proceedings <strong>of</strong> the 15 th Long<br />

Ashton International Symposium –<br />

Underst<strong>and</strong>ing Pathosystems: A<br />

Focus on <strong>Septoria</strong>. 15-17 September,<br />

1997. Long Ashton, UK. pp. 1-25.<br />

Harrower, K.M. 1978. Some aspects <strong>of</strong><br />

the infection process <strong>and</strong><br />

sporogenesis <strong>of</strong> <strong>Septoria</strong> nodorum<br />

<strong>and</strong> <strong>Septoria</strong> tritici. Proc.<br />

Australasian <strong>Septoria</strong> Workshop,<br />

20. Christchurch, New Zeal<strong>and</strong>.<br />

Johnson, R. 1992. Past, present <strong>and</strong><br />

future opportunities in breeding<br />

for disease resistance, with<br />

examples from wheat. Euphytica<br />

63:3-22.<br />

Jlibene, M., Mazouz, H., Farih, A., <strong>and</strong><br />

Saadoui, E.M. 1995. Host-pathogen<br />

interaction <strong>of</strong> wheat (Triticum<br />

aestivum) <strong>and</strong> <strong>Septoria</strong> tritici in<br />

Morocco. In Proceedings <strong>of</strong> the<br />

<strong>Septoria</strong> tritici workshop. Gilchrist,<br />

L., van Ginkel, M., McNab, A., <strong>and</strong><br />

Kema, G.H.J. (eds.) Mexico, D.F.:<br />

<strong>CIMMYT</strong>. pp. 34-40.<br />

Karjalainen, R., <strong>and</strong> Lounatmaa, K.<br />

1986. Ultrastructure <strong>of</strong> penetration<br />

<strong>and</strong> colonization <strong>of</strong> wheat leaves<br />

by <strong>Septoria</strong> nodorum. Physiological<br />

Molecular Plant Pathology 29:263-<br />

270.<br />

Kema, G.H.J., Sayoud, R., Annone,<br />

J.G., <strong>and</strong> van Silfout, C.H. 1996a.<br />

Genetic variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem II. Analysis <strong>of</strong><br />

interaactions between pathogen<br />

isolates <strong>and</strong> host cultivars.<br />

Phytopathology 86:213-220.<br />

Kema, G.H.J., Yu, D.Z., Rijkenberg,<br />

F.H.J., Shaw, M.W., <strong>and</strong> Baayen,<br />

R.P. 1996b. Histology <strong>of</strong> the<br />

pathogenesis <strong>of</strong> Mycosphaerella<br />

graminicola in wheat.<br />

Phytopathology 86:777-786.<br />

Krupinsky, J.M. 1994. Aggressiveness<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum isolates<br />

from alternative hosts after<br />

passage through wheat.<br />

Proceedings <strong>of</strong> the 4 th International<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong> Workshop. In<br />

Arseniuk, E., Goral, T., <strong>and</strong><br />

Czembor, P. (eds.) IHAR,<br />

Radzikow, Pol<strong>and</strong>. pp. 123-126.<br />

McDonald, B.A., <strong>and</strong> Martinez, J.P.<br />

1991. Chromosome length<br />

polymorphisms in <strong>Septoria</strong> tritici<br />

populations. Current Genetics<br />

19:265-271.<br />

McDonald, B.A. 1997. The population<br />

genetics <strong>of</strong> fungi: Tools <strong>and</strong><br />

Techniques. Phytopathology<br />

87:448-453.<br />

Saadoui, E.M. 1987. Physiologic<br />

specialization <strong>of</strong> <strong>Septoria</strong> tritici in<br />

Morocco. Plant Disease 71:153-155.<br />

Shaw, M.W., <strong>and</strong> Royle, D.J. 1989.<br />

Airborne inoculum as a major<br />

source <strong>of</strong> <strong>Septoria</strong> tritici<br />

(Mycosphaerella graminicola)<br />

infections in winter wheat crops in<br />

the UK. Plant Pathology 38:35-43.<br />

Sprague, R. 1950. <strong>Diseases</strong> <strong>of</strong> cereals<br />

<strong>and</strong> grasses in North America.<br />

Ronald Press, New York. 538 pp.<br />

Straley, M.L. 1979. Pathogenesis <strong>of</strong><br />

<strong>Septoria</strong> nodorum (Berk.) Berk. on<br />

wheat cultivars varying in<br />

resistance to glume blotch. Ph.D.<br />

Thesis, Montana State University,<br />

Bozeman.<br />

Ueng, P.P., Subramaniam, K., Chen,<br />

W., Arseniuk, E., Lixin, W.,<br />

Cheung, A.M., H<strong>of</strong>fmann, G.M.,<br />

<strong>and</strong> Bergstrom, G.C. 1998.<br />

Intraspecific genetic variation <strong>of</strong><br />

<strong>Stagonospora</strong> avenae <strong>and</strong> its<br />

differentiation from S. nodorum.<br />

Mycol. Res. 102(5):607-614.


Molecular Analysis <strong>of</strong> a DNA Fingerprint Probe from<br />

Mycosphaerella graminicola<br />

S.B. Goodwin <strong>and</strong> J.R. Cavaletto<br />

USDA-ARS, Department <strong>of</strong> Botany <strong>and</strong> Plant Pathology, Purdue University, West Lafayette, IN, USA<br />

Abstract<br />

Clones hybridizing to the Mycosphaerella graminicola DNA fingerprint probe pSTL70 were identified from<br />

subgenomic libraries <strong>and</strong> sequenced. Analyses <strong>of</strong> the DNA sequences <strong>of</strong> these clones plus the original pSTL70 clone revealed<br />

that pSTL70 contains part <strong>of</strong> the open reading frame for a probable homologue <strong>of</strong> an osmosensing histidine kinase gene from<br />

yeast. The remaining portion <strong>of</strong> the clone contained a partial reverse transcriptase gene sequence <strong>and</strong> a 29 base pair direct<br />

repeat, which could mean that the clone is a transposable element. Methods for converting transposable elements into<br />

improved DNA fingerprinting techniques are discussed.<br />

DNA fingerprinting is a<br />

powerful tool for analyzing the<br />

genetic structure <strong>of</strong> fungal<br />

populations. Several fingerprinting<br />

strategies have been employed,<br />

including those based on the<br />

polymerase chain reaction (PCR)<br />

such as r<strong>and</strong>om amplified<br />

polymorphic DNA (RAPD),<br />

amplified fragment length<br />

polymorphism (AFLP) <strong>and</strong> DNA<br />

amplification fingerprinting (DAF).<br />

These techniques can provide<br />

information on hundreds <strong>of</strong><br />

potential genetic loci in a very short<br />

time. However, each method has<br />

problems that can limit its utility.<br />

The RAPD technique relies on<br />

annealing <strong>of</strong> short (only 10 base)<br />

primers at low temperatures. This<br />

leads to high variability <strong>and</strong> low<br />

transportability to other labs.<br />

AFLPs require a reasonably high<br />

degree <strong>of</strong> sophistication in<br />

expertise <strong>and</strong> facilities, <strong>and</strong> also can<br />

suffer from problems with<br />

repeatability. DAF is operationally<br />

simple but the large number <strong>of</strong><br />

b<strong>and</strong>s produced can be difficult to<br />

separate <strong>and</strong> interpret.<br />

The most widely used DNA<br />

fingerprint technique is restriction<br />

fragment length polymorphism<br />

analysis using small pieces <strong>of</strong><br />

repetitive genomic DNA as probes.<br />

This technique has been used<br />

extensively to analyze the<br />

population biology <strong>of</strong> the<br />

ascomycetes Magnaporthe grisea<br />

(Hamer et al., 1989) <strong>and</strong><br />

Mycosphaerella graminicola<br />

(McDonald <strong>and</strong> Martinez, 1991),<br />

<strong>and</strong> the oomycete Phytophthora<br />

infestans (Goodwin et al., 1992).<br />

Thous<strong>and</strong>s <strong>of</strong> isolates <strong>of</strong> each<br />

species have been analyzed. The M.<br />

grisea repeat (MGR) 586 probe<br />

contains part <strong>of</strong> an inverted repeat<br />

transposon (Farman et al., 1996).<br />

However, the nature <strong>of</strong> the<br />

repeating elements in the M.<br />

graminicola pSTL70 <strong>and</strong> P. infestans<br />

RG57 probes has not been<br />

determined.<br />

This study was initiated to test<br />

whether the M. graminicola pSTL70<br />

probe was part <strong>of</strong> a transposable<br />

element. The long-term goal is to<br />

clone individual DNA fingerprint<br />

loci <strong>and</strong> convert them to a PCR-<br />

23<br />

based system by designing specific<br />

primers to unique regions at each<br />

genetic locus.<br />

Materials <strong>and</strong> Methods<br />

Subgenomic libraries were<br />

constructed from isolates IPO 323<br />

<strong>and</strong> 94269. These are the parent<br />

isolates <strong>of</strong> the M. graminicola<br />

mapping population (Kema et al.,<br />

1996). Approximately 2 µg <strong>of</strong><br />

genomic DNA from each isolate<br />

was digested to completion with<br />

the restriction enzyme Pst I. DNA<br />

fragments from 0.5-3 kb <strong>and</strong> from<br />

3-9 kb for each isolate were excised<br />

from gels <strong>and</strong> purified using<br />

Wizard PCR Prep (Promega,<br />

Madison, WI). The DNA fragments<br />

were then ligated into pBluescript<br />

vector <strong>and</strong> transformed into<br />

competent cells <strong>of</strong> E. coli strain<br />

INValphaF’. White colonies were<br />

transferred into 200 µL LB+amp<br />

medium in 96-well Microtest tissue<br />

culture plates <strong>and</strong> grown at 37°C<br />

overnight. The 96 cultures from<br />

each plate were transferred onto<br />

large (150 x 15 mm) LB+amp agar


24<br />

Session 1 — S.B. Goodwin <strong>and</strong> J.R. Cavaletto<br />

plates using a replica plater, <strong>and</strong><br />

incubated upside down at 37°C<br />

overnight.<br />

Colony lifts were made onto 8 x<br />

12 cm pieces <strong>of</strong> Zeta Probe<br />

(BioRad) membranes by briefly<br />

laying the membrane pieces over<br />

the 96 colonies <strong>and</strong> lifting to pick<br />

up the bacteria. Membranes were<br />

placed colony side up onto blotting<br />

paper soaked with 10% SDS for 3<br />

min, then 0.5 M NaOH/1.5 M NaCl<br />

for 5 min, 0.5 M Tris, pH 8.0/1.5 M<br />

NaCl for 5 min, <strong>and</strong> 6x SSC for 5<br />

min. Finally, a UV Stratalinker<br />

(Stratagene) was used to crosslink<br />

the plasmid DNA to the membrane.<br />

For Southern analysis, pSTL70<br />

DNA was labeled using the<br />

R<strong>and</strong>om Primer Fluorescein<br />

labeling kit with antifluorescein-<br />

HRP (DuPont NEN) <strong>and</strong><br />

hybridized according to the<br />

manufacturer’s instructions.<br />

Approximately 4,000 clones (2,000<br />

from each isolate) were screened.<br />

Positive hybridizations were<br />

verified by digesting each positive<br />

clone with Pst I to release the insert,<br />

separating the fragments on<br />

agarose gels, blotting <strong>and</strong> probing<br />

as described above. Clones that<br />

hybridized after two rounds <strong>of</strong><br />

screening were sequenced using a<br />

Pharmacia ALF automated DNA<br />

sequencer.<br />

Results<br />

Among 4,000 clones screened,<br />

five hybridized strongly to pSTL70.<br />

Complete sequences have been<br />

obtained for the original pSTL70<br />

clone <strong>and</strong> three others. Clones<br />

2E11, 9A5 <strong>and</strong> 11E8 each had large<br />

regions <strong>of</strong> near identity with<br />

pSTL70 (Table 1). However, the<br />

three clones shared no similarity<br />

with each other. BLASTX searches<br />

<strong>of</strong> the putative translation products<br />

identified a number <strong>of</strong> GenBank<br />

accessions with similarity to clones<br />

pSTL70, 2E11 <strong>and</strong> 11E8 (Table 1).<br />

The original pSTL70 clone had high<br />

similarity to a two-component<br />

regulator gene from yeast, Sln1,<br />

<strong>and</strong> a similar gene from C<strong>and</strong>ida<br />

albicans.<br />

Clone 2E11 had even higher<br />

similarity to the same genes. This<br />

clone corresponded to <strong>and</strong><br />

extended the 5’ end <strong>of</strong> pSTL70<br />

(Figure 1). Clone 11E8<br />

corresponded to <strong>and</strong> extended the<br />

3’ end <strong>of</strong> pSTL70 (Figure 1). This<br />

sequence may code for a reverse<br />

transcriptase. The final 239 bases <strong>of</strong><br />

pSTL70 contained part <strong>of</strong> the<br />

putative reverse transcriptase<br />

coding sequence, but not enough <strong>of</strong><br />

the gene to obtain positive BLAST<br />

hits in GenBank.<br />

Table 1. Analysis <strong>of</strong> the Mycosphaerella graminicola DNA fingerprint clone pSTL70 <strong>and</strong> three<br />

additional clones that hybridized to pSTL70 in Southern analysis.<br />

Size Overlap Blastx E<br />

Clone (bp) with pSTL70 results value Comments<br />

pSTL70 2860 N/A Sln1 8e-14 29 bp repeat<br />

2E11 3073 1278 Sln1 2e-15 Gene sequence<br />

9A5 2640 1103 No hits N/A 29 bp repeat<br />

11E8 1636 417 bp Reverse 0.006 Transposable<br />

transcriptase element (?)<br />

pSTL70<br />

2E11<br />

9A5<br />

11E8<br />

The first 917 bases <strong>of</strong> 11E8 are<br />

not related to the sequence <strong>of</strong><br />

pSTL70. Clone 9A5 had no<br />

similarity to any sequences in<br />

GenBank. The first 1103 bases <strong>of</strong><br />

this clone were identical to pSTL70,<br />

but the final 1537 bases were<br />

unrelated (Figure 1). Clones<br />

pSTL70 <strong>and</strong> 9A5 both contained a<br />

29 bp sequence that was t<strong>and</strong>emly<br />

repeated approximately 3.5 times<br />

(Figure 1). Southern analysis <strong>of</strong><br />

genomic DNA from the parents<br />

<strong>and</strong> several progeny isolates <strong>of</strong> the<br />

M. graminicola mapping population<br />

revealed that 9A5 also gave a DNA<br />

fingerprint pattern, while 2E11 <strong>and</strong><br />

16D7 did not (data not shown).<br />

Discussion<br />

The M. graminicola fingerprint<br />

probe pSTL70 (McDonald <strong>and</strong><br />

Martinez, 1991) contains part <strong>of</strong> the<br />

open reading frame for the yeast<br />

two-component regulator gene<br />

Sln1. This gene has been shown to<br />

function as an osmosensing<br />

Sln 1 R<br />

RT<br />

Figure 1. Relationships among the Mycosphaerella graminicola DNA fingerprint probe pSTL70<br />

<strong>and</strong> related clones. Regions <strong>of</strong> overlap are indicated by open boxes, unique regions by single<br />

lines. Clones are labeled on the left. Approximate locations <strong>of</strong> the Sln1 <strong>and</strong> reverse<br />

transcriptase (RT) open reading frames are indicated by arrows. The t<strong>and</strong>em repeats are<br />

indicated by R.


histidine kinase in yeast, although<br />

its function in M. graminicola<br />

remains unknown. Clone 2E11<br />

includes an even larger portion <strong>of</strong><br />

this gene containing the first 500<br />

amino acids <strong>of</strong> the putative yeast<br />

homologue.<br />

Other parts <strong>of</strong> the pSTL70 clone<br />

contain a 29 bp t<strong>and</strong>em repeat <strong>and</strong><br />

a partial coding sequence for a<br />

reverse transcriptase. These<br />

indicate that pSTL70 may contain<br />

part <strong>of</strong> a transposable element.<br />

Hybridization analysis confirmed<br />

that this region <strong>of</strong> the clone was<br />

responsible for the DNA<br />

fingerprint pattern, which<br />

strengthens the transposable<br />

element hypothesis. The MGR586<br />

probe <strong>of</strong> M. grisea also has been<br />

shown to contain a transposable<br />

element flanked by 16 bp t<strong>and</strong>em<br />

repeats <strong>and</strong> 9 bp inverted repeat<br />

sequences (Farman et al., 1996). If<br />

pSTL70 is a transposable element, it<br />

is truncated partway through the<br />

reverse-transcriptase sequence.<br />

Cloning <strong>and</strong> analysis <strong>of</strong> the<br />

remaining portion <strong>of</strong> the reversetranscriptase<br />

gene will be necessary<br />

to test whether it is flanked by<br />

direct <strong>and</strong>/or inverted repeats.<br />

It is surprising that the three<br />

additional clones identified using<br />

pSTL70 as a probe did not overlap<br />

with each other. Two <strong>of</strong> the clones,<br />

9A5 <strong>and</strong> 11E8, contained unique<br />

regions that did not overlap with<br />

pSTL70. There are two likely<br />

explanations for this result. One is<br />

that these clones were from<br />

different loci in the M. graminicola<br />

Molecular Analysis <strong>of</strong> a DNA Fingerprint Probe from Mycosphaerella graminicola 25<br />

genome that shared segments due<br />

to independent movements <strong>of</strong> the<br />

putative transposable element. The<br />

other possibility is that one or more<br />

<strong>of</strong> these clones is a chimera with<br />

multiple inserts. The chimera<br />

hypothesis is being tested by<br />

additional analyses <strong>of</strong> the cloned<br />

sequences.<br />

If the DNA fingerprint pattern<br />

<strong>of</strong> pSTL70 is due to a transposable<br />

element, it may be possible to make<br />

specific PCR primers to amplify<br />

single DNA fingerprint loci. This<br />

could be accomplished using two<br />

different strategies. If the element is<br />

relatively small, primers could be<br />

made in the single-copy regions<br />

flanking the element. These should<br />

give a large amplification product<br />

when the element is present <strong>and</strong> a<br />

small one when it is not. Such<br />

differences could be resolved easily<br />

on agarose gels without the need<br />

for autoradiography, <strong>and</strong> should be<br />

easily transportable to other<br />

laboratories.<br />

The other approach would be to<br />

design one primer within the<br />

transposable element <strong>and</strong> the other<br />

in the flanking region. This would<br />

give a plus/minus polymorphism<br />

at each locus: plus when the<br />

transposable element is present <strong>and</strong><br />

minus when it is absent. The<br />

advantage <strong>of</strong> this approach is that<br />

the primers could be designed so<br />

that the sizes <strong>of</strong> the PCR products<br />

would not overlap. It then may be<br />

possible to use multiplex PCR to<br />

perform DNA fingerprinting <strong>of</strong><br />

individual genetic loci from small<br />

quantities <strong>of</strong> starting genomic DNA<br />

in single PCR reactions. This would<br />

be much faster <strong>and</strong> simpler than<br />

Southern analysis <strong>and</strong> would avoid<br />

the repeatability, transportability,<br />

<strong>and</strong> interpretation problems <strong>of</strong><br />

other PCR-based methods for<br />

fungal population genetic analyses.<br />

Acknowledgments<br />

We thank Bruce McDonald for<br />

providing the original pSTL70<br />

DNA fingerprint clone <strong>and</strong> for<br />

encouragement during the course<br />

<strong>of</strong> this project.<br />

References<br />

Farman, M.L., S. Taura, <strong>and</strong> S.A.<br />

Leong. 1996. The Magnaporthe<br />

grisea DNA fingerprinting probe<br />

MGR586 contains the 3’ end <strong>of</strong> an<br />

inverted repeat transposon. Mol.<br />

Gen. Genet. 251:675-681.<br />

Goodwin, S.B., A. Drenth, <strong>and</strong> W.E.<br />

Fry. 1992. Cloning <strong>and</strong> genetic<br />

analyses <strong>of</strong> two highly<br />

polymorphic, moderately<br />

repetitive nuclear DNAs from<br />

Phytophthora infestans. Curr. Genet.<br />

22:107–115.<br />

Hamer, J.E., L. Farrall, M.J. Orbach, B.<br />

Valent, <strong>and</strong> F.G. Chumley. 1989.<br />

Host species-specific conservation<br />

<strong>of</strong> a family <strong>of</strong> repeated DNA<br />

sequences in the genome <strong>of</strong> a<br />

fungal plant pathogen. Proc. Natl.<br />

Acad. Sci., USA 86:9981-9985.<br />

Kema, G.H.J., E.C.P. Verstappen, M.<br />

Todorova, <strong>and</strong> C. Waalwijk. 1996.<br />

Successful crosses <strong>and</strong> molecular<br />

tetrad <strong>and</strong> progeny analyses<br />

demonstrate heterothallism in<br />

Mycosphaerella graminicola. Curr.<br />

Genet. 30:251-258.<br />

McDonald, B.A., <strong>and</strong> J.P. Martinez.<br />

1991. DNA fingerprinting <strong>of</strong> the<br />

plant pathogenic fungus<br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici). Exp.<br />

Mycol. 15:146-158.


26<br />

Characterization <strong>of</strong> <strong>Septoria</strong> tritici Variants <strong>and</strong> PCR<br />

Assay for Detecting <strong>Stagonospora</strong> nodorum <strong>and</strong> <strong>Septoria</strong><br />

tritici in Wheat<br />

S. Hamza, 1 M. Medini, 1 T. Sassi, 1 S. Abdennour, 1 M. Rouassi, 1<br />

A.B. Salah, 1 M. Cherif, 1 R. Strange, 2 <strong>and</strong> M. Harrabi1 1 Laboratoire de Génétique, Institut National Agronomique de Tunisie, Tunisia<br />

2 Plant Pathology, UCL London, United Kingdom<br />

Abstract<br />

Pathogenic specialization <strong>of</strong> <strong>Septoria</strong> tritici was studied by inoculating 14 isolates <strong>of</strong> the fungus, <strong>of</strong> which seven were<br />

obtained from durum wheat <strong>and</strong> seven were isolated from bread wheat. Isolates obtained from durum wheat were more<br />

virulent on durum wheat, while those isolated from bread wheat were more severe on bread wheat, which revealed<br />

physiologic specialization <strong>of</strong> S. tritici to either bread or durum wheat. The sequence coding for the nuclear 5.8S rDNA <strong>and</strong><br />

the internal transcribed spacer (ITS1 <strong>and</strong> ITS2) were amplified by polymerase chain reaction <strong>and</strong> sequenced for five isolates<br />

adapted to bread wheat <strong>and</strong> five isolates adapted to durum wheat. These sequences were identical between both variants,<br />

resulting in the absence <strong>of</strong> divergence inside tritici species. <strong>Septoria</strong> tritici <strong>and</strong> <strong>Stagonospora</strong> nodorum isolates collected<br />

from Tunisia were tested for amplification with specific primers to these pathogens. This revealed the primer’s efficiency to<br />

distinguish <strong>Septoria</strong> species <strong>and</strong> detect <strong>Stagonospora</strong> nodorum DNA with as little as 2 pg <strong>of</strong> DNA. Time course<br />

quantification <strong>of</strong> S. tritici mycelia after inoculation <strong>of</strong> resistant <strong>and</strong> susceptible cultivars distinguished those cultivars by the<br />

earliest date <strong>of</strong> apparition <strong>of</strong> a PCR product.<br />

Extensive genetic variation for<br />

virulence in <strong>Septoria</strong> tritici or its<br />

telomorph Mycosphaerella<br />

graminicola, characterized by<br />

differential interaction between<br />

host <strong>and</strong> pathogen genotypes,<br />

suggested the involvement <strong>of</strong><br />

specific factors for virulence <strong>and</strong><br />

resistance in the pathosystem<br />

(Kema et al., 1996a). Specific<br />

interaction between M. graminicola<br />

<strong>and</strong> wheat was well demonstrated<br />

by statistical evidence (Kema et al.,<br />

1996b), which suggested gene-forgene<br />

interaction between resistance<br />

<strong>and</strong> virulence in host <strong>and</strong><br />

pathogen, respectively.<br />

Mycosphaerella graminicola<br />

specialization is much more<br />

pronounced on bread wheat <strong>and</strong><br />

durum wheat than differential<br />

specificity on a particular cultivar.<br />

In particular when considering the<br />

presence <strong>of</strong> pycnidia as a disease<br />

parameter, bread wheat <strong>and</strong> durum<br />

wheat isolates were particularly<br />

virulent on bread <strong>and</strong> durum<br />

wheat, respectively. Therefore<br />

bread wheat <strong>and</strong> durum wheat<br />

variants in M. graminicola were<br />

considered. The identical sequence<br />

<strong>of</strong> the internal transcribed spacer<br />

ribosomal DNA (ITS rDNA)<br />

observed by Kema et al. (1996a)<br />

showed that both variants could<br />

not be distinguished using their<br />

ribosomal DNA <strong>and</strong> that both were<br />

therefore from a similar taxonomic<br />

rank.<br />

PCR has been shown to be a<br />

powerful technique to detect small<br />

amounts <strong>of</strong> fungal plant pathogens.<br />

The technique is now commonly<br />

used by field pathologists <strong>and</strong><br />

growers for diagnosing several<br />

plant pathogens. The early<br />

detection <strong>of</strong> plant disease allows<br />

the judicious use <strong>of</strong> agricultural<br />

fungicides; early treatment is <strong>of</strong>ten<br />

more effective because the<br />

pathogen is less well established in<br />

its host. PCR amplification using<br />

specific primers for S. tritici <strong>and</strong><br />

<strong>Stagonospora</strong> nodorum or <strong>Septoria</strong><br />

nodorum was able to distinguish<br />

these pathogens <strong>and</strong> detect a few<br />

fungal cells in infested wheat<br />

tissues well before disease<br />

symptoms were apparent (Beck<br />

<strong>and</strong> Ligon, 1995). Since both<br />

pathogens may infect the same<br />

plant <strong>and</strong> visually distinguishing<br />

them is difficult, the specific<br />

detection <strong>of</strong> S. tritici <strong>and</strong> S. nodorum<br />

is well circumvented using the PCR<br />

technique <strong>and</strong> facilitates the<br />

decision <strong>of</strong> whether or not to treat<br />

with fungicides <strong>and</strong> which<br />

fungicides to use.


According to the previous<br />

observations, this paper describes<br />

physiologic specialization <strong>of</strong> bread<br />

<strong>and</strong> durum wheat isolates collected<br />

from Morocco <strong>and</strong> Tunisia,<br />

respectively, <strong>and</strong> the sequence <strong>of</strong><br />

their ITS rDNA. Specific primers<br />

determined by Beck <strong>and</strong> Ligon<br />

(1995) for PCR detection <strong>of</strong> S. tritici<br />

<strong>and</strong> S. nodorum were used to<br />

differentiate those pathogens from<br />

other fungal pathogens <strong>of</strong> wheat<br />

<strong>and</strong> to detect their presence in<br />

asymptomatic wheat plants. Time<br />

course PCR quantification <strong>of</strong> S.<br />

tritici in infected wheat was<br />

assessed to analyze the evolution <strong>of</strong><br />

mycelia in resistant <strong>and</strong> susceptible<br />

cultivars.<br />

Materials <strong>and</strong> Methods<br />

Plant material <strong>and</strong><br />

experimental design<br />

Thirty-six durum<br />

wheat <strong>and</strong> eight bread<br />

wheat cultivars<br />

(Table 1) were<br />

employed to study<br />

genetic variation for<br />

virulence. Seeds <strong>of</strong> the<br />

cultivars were sown in<br />

trays with 228 alveolus<br />

with 3 cm x 3 cm<br />

surface <strong>and</strong> 5 cm<br />

height. Each alveolus<br />

contained four seeds.<br />

The experiment was<br />

conducted in an<br />

arbitrary complete<br />

block design with three<br />

replicates for each<br />

isolate. Replicates were<br />

blocked in the same<br />

tray, separated by two<br />

rows <strong>of</strong> alveolus.<br />

Characterization <strong>of</strong> <strong>Septoria</strong> tritici Variants <strong>and</strong> PCR Assay for Detecting <strong>Stagonospora</strong> nodorum <strong>and</strong> <strong>Septoria</strong> tritici in Wheat 27<br />

Experimental procedure<br />

Seven isolates <strong>of</strong> S. tritici taken<br />

from durum wheat <strong>and</strong> seven<br />

isolates from bread wheat collected<br />

in different regions <strong>of</strong> Tunisia <strong>and</strong><br />

Morocco, respectively, were used to<br />

inoculate durum <strong>and</strong> bread wheat<br />

cultivars (Table 2). Eleven-day-old<br />

seedlings with emerging second<br />

leaves were inoculated with a<br />

monosporal suspension till run-<strong>of</strong>f.<br />

The inoculum was prepared from<br />

monosporal culture cultivated on<br />

potato dextrose agar (PDA)<br />

medium. The spores were scraped<br />

from the agar, re-suspended in<br />

distilled water, filtered, <strong>and</strong><br />

adjusted to 106-107 spores/ml. After<br />

inoculation the trays were<br />

incubated in a humid chamber for<br />

72 hours <strong>and</strong> returned to a growth<br />

chamber at temperature 20-25°C<br />

during the day, 17°C during the<br />

night, <strong>and</strong> 12 hours photoperiod.<br />

Table 1. List <strong>of</strong> durum <strong>and</strong> bread wheat cultivars used to study<br />

genetic variation for virulence in Mycosphaerella<br />

graminicola.<br />

Code Cultivars Code Cultivars<br />

01 Florence aurore* 23 Medea AC3<br />

02 Florence Aurore x PUSA* 24 Medea AC4<br />

03 Derbessi x Biskri 25 Medea AP1<br />

04 Guelma* 26 Medea AP2<br />

05 EAPC6326127* 27 Allorea*<br />

06 Tunis9* 28 Richelle*<br />

07 Tunis23* 29 Sebei glabre<br />

08 Abdelkader 30 Sebei pubescent<br />

09 Agili pubescent AC1 31 Souri AC8<br />

10 Bidi AP4 32 Souri AC9<br />

11 Bidi AP1 33 Medea AC1<br />

12 Bidi AP3 34 Medea AC2<br />

13 Bidi AP 35 Medea AP6<br />

14 Biskri glabre 36 Medea AP10<br />

15 Biskri glabre AP3 37 Medea RP1<br />

16 Derbessi AC1 38 Biancullida ICM27<br />

17 Derbessi AP1 39 Mahmoudi ICM28<br />

18 Derbessi AP2 40 Mahmoudi ICM67<br />

19 Hamira AC2 41 Khotifa x ICM75<br />

20 Hamira AC3 42 ICM313<br />

21 Hamira AC4 43 ICM314<br />

22 Hamira AC5 44 Hamira<br />

* Bread wheat cultivars.<br />

Disease evaluation<br />

<strong>and</strong> statistical analysis<br />

Disease severity was evaluated<br />

at 21 days after inoculation on the<br />

second leaf <strong>of</strong> the plant <strong>and</strong> using<br />

the presence <strong>of</strong> pycnidia as the<br />

disease parameter (Kema et al.,<br />

1996a). The collected data were<br />

treated with SAS s<strong>of</strong>tware (1993<br />

version). The presence <strong>of</strong> pycnidia<br />

was subjected to analysis <strong>of</strong><br />

variance. The tables <strong>of</strong> means <strong>of</strong> the<br />

14 isolates were subjected to<br />

hierarchical clustering using the<br />

statistical analysis s<strong>of</strong>tware CSTAT.<br />

DNA extraction from fungal<br />

spores <strong>and</strong> infected plants<br />

Spores <strong>of</strong> S. tritici <strong>and</strong> S. nodorum<br />

cultivated on PDA medium were<br />

subcultured in 100 ml flasks <strong>of</strong> yeast<br />

extract <strong>and</strong> glucose liquid medium.<br />

Five-day spore cultures were<br />

centrifuged <strong>and</strong> DNA extraction <strong>of</strong><br />

fungal spores was performed using<br />

CTAB extraction buffer according to<br />

the protocol described by Morjane et<br />

al. (1995). DNA extraction from three<br />

infected leaves was performed<br />

according to the protocol described<br />

by Möller et al. (1992). DNA pellet<br />

after extraction was re-suspended in<br />

50 µl Tris-HCl; 10 mM EDTA; 1 mM<br />

(TE) buffer.<br />

PCR amplification<br />

<strong>and</strong> sequencing<br />

The sequence <strong>of</strong> ITS rDNA<br />

concerned 5 isolates (MAR 1, 2, 3, 4,<br />

5) collected from Morocco (Table 2)<br />

<strong>and</strong> 5 isolates (TUN 1, 2, 3, 4, 5)<br />

collected from Tunisia (Table 2). ITS<br />

amplification was performed with<br />

ITS1 (5’-<br />

TCCGTAGGTGAACCTGCGG-3’)<br />

<strong>and</strong> ITS 4 (5’-<br />

TCCTCCGCTTATTGATATGC-3’)<br />

universal primers. PCR reactions


28<br />

Session 1 — S. Hamza, M. Medini, T. Sassi, S. Abdennour, M. Rouassi, A.B. Salah, M. Cherif, R. Strange, <strong>and</strong> M. Harrabi<br />

Table 2. Experimental code <strong>and</strong> origin <strong>of</strong> Mycosphaerella graminicola collected from Morocco<br />

<strong>and</strong> Tunisia to study genetic variation for virulence.<br />

Isolates collected from Tunisia Isolates collected from Morocco<br />

Origin Origin<br />

Code Location Wheat species Code Location Wheat species<br />

TUN1 Beja DW MAR1 Jemaa Sham BT<br />

TUN2 Mateur DW MAR2 Douyet BT<br />

TUN3 Mornag DW MAR3 Agadir BT<br />

TUN4 El Agba DW MAR4 Ain Orma BT<br />

TUN5 Zaghouane DW MAR5 Azrou BT<br />

TUN6 Siliana DW MAR6 Essaouira BT<br />

TUN7 Mjez El Bab DW MAR7 Tetouan BT<br />

were performed in a volume <strong>of</strong> 50<br />

µl containing 50 mM KCl, 10 mM<br />

Tris-HCl, pH8.3, 1.5 mM MgCl2,<br />

200 µM <strong>of</strong> dNTP, 50 pmole <strong>of</strong><br />

primer, 2.5 units <strong>of</strong> Taq<br />

polymerase (Boerhringer-<br />

Manheim) <strong>and</strong> 25 ng <strong>of</strong> genomic<br />

DNA. Reactions were run in a<br />

Thermolyne, Temptronic model<br />

thermal cycler for 30 cycles, each<br />

consisting <strong>of</strong> 15s at 94°C, 15 s at<br />

50°C, <strong>and</strong> 45 s at 72°C. An<br />

additional 5 min polymerization<br />

step at 72°C ended the<br />

amplification. The products were<br />

analyzed by electrophoresis <strong>of</strong> 20<br />

µl aliquot <strong>of</strong> each PCR sample on<br />

0.8% agarose gel.<br />

Specific DNA amplifications<br />

were performed as determined by<br />

Beck <strong>and</strong> Ligon (1995). ITS1 <strong>and</strong><br />

JB446 (5’-<br />

CGAGGCTGGAGTGGTGT-3’)<br />

primers were used for specific<br />

amplification <strong>of</strong> S. tritici DNA <strong>and</strong><br />

JB433 (5’-<br />

ACACTCAGTAGTTTACTACT-3’)<br />

<strong>and</strong> JB434 (5’-<br />

TGTGCTGCGTTCAATA-3’) were<br />

used to amplify S. nodorum DNA.<br />

PCR was performed as described<br />

above, except the annealing<br />

temperature was 57°C.<br />

ITS sequencing was performed<br />

using 90 ng <strong>of</strong> 500 bp amplified<br />

product with ITS1 <strong>and</strong> ITS4. The<br />

sequence was determined by the<br />

dideoxynucleotide chain<br />

termination method using an<br />

automated sequencer (Perkin<br />

Elmer) at the Darwin Building <strong>of</strong><br />

University College London<br />

(London, United Kingdom).<br />

Fungal DNA amplification in<br />

resistant <strong>and</strong> susceptible<br />

cultivars<br />

To evaluate the competition <strong>of</strong><br />

plant DNA with fungal DNA using<br />

ITS1 <strong>and</strong> ITS4 primers,<br />

amplification was performed using<br />

1 µl <strong>and</strong> 5 µl <strong>of</strong> plant DNA extract<br />

mixed with several amounts (<br />

0.001, 0.01; 0.1; 1; 10; 100 <strong>and</strong> 1000<br />

ng) <strong>of</strong> fungal DNA. The PCR<br />

conditions were the same as<br />

described before with a 57°C<br />

annealing temperature.<br />

The time course amplification<br />

using ITS1 <strong>and</strong> ITS4 primers was<br />

realized using 2 µl <strong>of</strong> DNA extract<br />

from one infected leaf at 3, 6, 9, 14,<br />

18, <strong>and</strong> 22 days after inoculation.<br />

The PCR conditions were the same<br />

as described before, with a 57°C<br />

annealing temperature.<br />

Results <strong>and</strong> Discussion<br />

Bread wheat derived isolates<br />

almost exclusively produced<br />

pycnidia in the bread cultivars,<br />

whereas pycnidial production by<br />

durum wheat derived isolates was<br />

almost entirely restricted to durum<br />

wheat isolates. The analysis <strong>of</strong><br />

variance (Table 3) shows highly<br />

significant differences (p


cultivars. The presence <strong>of</strong> pycnidia<br />

<strong>of</strong> bread wheat derived isolates<br />

does not exceed 15% <strong>of</strong> the leaf<br />

surface (data not shown). An<br />

explanation <strong>of</strong> the observed weak<br />

virulence would be long time<br />

conservation <strong>of</strong> these isolates in<br />

glycerol. Aggressivity <strong>of</strong> these<br />

isolates would be recovered after<br />

their direct isolation from infected<br />

leaves <strong>and</strong> inoculation from fresh<br />

spore culture.<br />

Cluster analysis <strong>of</strong> the 14<br />

isolates according to the presence<br />

<strong>of</strong> pycnidia resulted in nine<br />

different clusters. Bread wheat<br />

derived isolates are clearly<br />

separated from durum wheat<br />

derived isolates (Figure 1). Durum<br />

wheat isolates are clustered into six<br />

groups. Two isolates (TUN7 <strong>and</strong><br />

TUN6) originated from distant<br />

regions (Mjez El Bab <strong>and</strong> Siliana,<br />

respectively) belong to the same<br />

group, whereas isolates from the<br />

same regions belong to different<br />

216.7<br />

150<br />

100<br />

Figure 1. Hierarchical classification <strong>of</strong> 14<br />

<strong>Septoria</strong> tritici isolates as determined by<br />

CSTAT s<strong>of</strong>tware.<br />

Characterization <strong>of</strong> <strong>Septoria</strong> tritici Variants <strong>and</strong> PCR Assay for Detecting <strong>Stagonospora</strong> nodorum <strong>and</strong> <strong>Septoria</strong> tritici in Wheat 29<br />

50<br />

MAR1<br />

MAR2<br />

MAR3<br />

MAR5<br />

MAR4<br />

MAR6<br />

MAR7<br />

TUN7<br />

TUN6<br />

TUN4<br />

TUN5<br />

TUN3<br />

TUN2<br />

TUN1<br />

0.0<br />

groups, indicating genetic variation<br />

for virulence within local<br />

populations.<br />

ITS sequence analysis<br />

Physiological specialization <strong>of</strong><br />

S. tritici to bread wheat <strong>and</strong> durum<br />

wheat led to the hypothesis <strong>of</strong> the<br />

existence <strong>of</strong> two subspecies inside<br />

tritici. For that purpose sequencing<br />

ITS rDNA would provide evidence<br />

about the taxonomy <strong>of</strong> those<br />

variants. The comparison <strong>of</strong> the<br />

consensus sequence <strong>of</strong> the 550 pb<br />

ITS region fragments obtained from<br />

five bread wheat derived isolates<br />

<strong>and</strong> five durum wheat derived<br />

isolates shows 100% homology<br />

(Figure 2). Therefore S. tritici<br />

variants belong to the same<br />

taxonomic rank <strong>and</strong> cannot be<br />

considered as two different<br />

subspecies <strong>of</strong> tritici. Since both<br />

variants coexist in the same field<br />

(McDonald, personal<br />

communication), permanent<br />

Figure 2. Nucleotide sequence <strong>of</strong> the internal transcribed spacer region derived from bread<br />

wheat (MAR) <strong>and</strong> durum wheat (TUN) derived isolates <strong>of</strong> <strong>Septoria</strong> tritici. The sequence includes<br />

the 5.8S rDNA gene <strong>and</strong> the internal transcribed spacer (ITS1 <strong>and</strong> ITS2). The consensus<br />

sequences were deduced from bread wheat <strong>and</strong> durum wheat derived isolates each.


30<br />

Session 1 — S. Hamza, M. Medini, T. Sassi, S. Abdennour, M. Rouassi, A.B. Salah, M. Cherif, R. Strange, <strong>and</strong> M. Harrabi<br />

Figure 2. cont.<br />

conversion <strong>of</strong> one variant to the<br />

other by genetic exchange <strong>of</strong><br />

virulence genes may occur through<br />

sexual reproduction.<br />

Diagnosis <strong>of</strong> S. tritici<br />

<strong>and</strong> S. nodorum<br />

In order to investigate the<br />

specificity <strong>of</strong> S. tritici <strong>and</strong> S.<br />

nodorum specific primers<br />

determined by Beck <strong>and</strong> Ligon<br />

(1995), we examined whether they<br />

produce PCR products from other<br />

fungal pathogens <strong>of</strong> wheat, such as<br />

Fusarium graminearum, Ustilago<br />

maydis, <strong>and</strong> Puccinia graminis. The<br />

S. tritici specific primers JB446 <strong>and</strong><br />

ITS1, <strong>and</strong> S. nodorum specific<br />

primers JB433 <strong>and</strong> JB434 did<br />

produce PCR products only<br />

from those pathogens (Figure<br />

3) revealing that these<br />

primers can be used to<br />

diagnose S. tritici <strong>and</strong> S.<br />

nodorum isolates collected<br />

from Tunisia.<br />

Amounts ranging from 2<br />

µg to 2 pg <strong>of</strong> S. nodorum were<br />

tested by PCR amplification with<br />

their respective specific primers<br />

(JB433, JB434) to determine the<br />

DNA threshold that can be<br />

detected. PCR amplification with<br />

550 bp<br />

JB446 <strong>and</strong> ITS1 JB433 <strong>and</strong> JB434<br />

Labda HindIII<br />

Healthy wheat<br />

<strong>Septoria</strong> tritici<br />

<strong>Stagonospora</strong> nodorum<br />

Fusarium graminearum<br />

Puccinia graminis<br />

Ustilago maydis<br />

Lambda HindIII<br />

Healty wheat<br />

<strong>Septoria</strong> tritici<br />

<strong>Stagonospora</strong> nodorum<br />

Fusarium graminearum<br />

Puccinia graminis<br />

Ustilago maydis<br />

Lambda HindIII<br />

Figure 3. Ethidium bromide stained agarose gel <strong>of</strong><br />

polymerase chain reaction amplification product<br />

using <strong>Septoria</strong> tritici specific primers (JB446 <strong>and</strong><br />

ITS1) <strong>and</strong> <strong>Stagonospora</strong> nodorum specific primers<br />

(JB433 <strong>and</strong> JB434) with healthy wheat DNA <strong>and</strong><br />

fungal DNA <strong>of</strong> wheat pathogen.<br />

Lambda DNA HindIII<br />

2 µg<br />

200 ng<br />

20 ng<br />

2 ng<br />

200 pg<br />

20 pg<br />

2 pg<br />

Lambda DNA HindIII<br />

Figure 4. Ethidium bromide stained agarose gel <strong>of</strong><br />

polymerase chain reaction products from<br />

amplification <strong>of</strong> 0.2 pg to 2 µg <strong>of</strong> genomic DNA <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum DNA using S. nodorum<br />

specific primers JB433 <strong>and</strong> JB434.<br />

JB433 <strong>and</strong> JB434 was able to detect<br />

2 pg <strong>of</strong> S. nodorum DNA (Figure 4),<br />

which is almost the same amount<br />

detected by Beck <strong>and</strong> Ligon (1995).


Provided that S. nodorum fungal<br />

genomes do not exceed 100 Mbp in<br />

those conditions, PCR is able to<br />

detect as little as 20 fungal cells.<br />

Fungal DNA amplification in<br />

resistant <strong>and</strong> susceptible<br />

cultivars<br />

Different fragment sizes 600 bp<br />

<strong>and</strong> 550 bp amplification products<br />

were obtained from plant <strong>and</strong><br />

fungal DNA, respectively, using<br />

ITS1 <strong>and</strong> ITS4 primers (Figure 5).<br />

PCR amplification with ITS primers<br />

<strong>and</strong> using a mixture <strong>of</strong> both fungal<br />

<strong>and</strong> plant DNA templates resulted<br />

in the amplification <strong>of</strong> both<br />

fragments. However, the intensity<br />

<strong>of</strong> the fungal PCR product is<br />

relative to the amount <strong>of</strong> fungal<br />

DNA (Figure 5). These result<br />

suggest that plant DNA could be<br />

used as competitor <strong>of</strong> fungal DNA<br />

amplification with ITS primers.<br />

Time course PCR amplification<br />

using ITS1 <strong>and</strong> ITS4 primers at 0, 3,<br />

6, 9, 14, 18, <strong>and</strong> 22 days after<br />

inoculation <strong>of</strong> susceptible (Karim)<br />

<strong>and</strong> resistant (Jennah Khotifa)<br />

cultivars with S. tritici is shown in<br />

Figure 6. ITS rDNA amplification<br />

products from fungal DNA (550 bp)<br />

appeared at 3 <strong>and</strong> 18 days from<br />

infected susceptible <strong>and</strong> resistant<br />

cultivars, respectively.<br />

This result showed that the<br />

multiplication <strong>of</strong> S. tritici inside the<br />

plant is delayed in the resistant<br />

cultivar. Therefore, the time<br />

interval corresponding to the<br />

beginning <strong>of</strong> PCR product<br />

apparition may be an indicator <strong>of</strong><br />

resistance. Furthermore, the<br />

intensity <strong>of</strong> the PCR product<br />

increased with the days postinoculation<br />

(Figure 6), indicating a<br />

quantitative response <strong>of</strong> the PCR<br />

Characterization <strong>of</strong> <strong>Septoria</strong> tritici Variants <strong>and</strong> PCR Assay for Detecting <strong>Stagonospora</strong> nodorum <strong>and</strong> <strong>Septoria</strong> tritici in Wheat 31<br />

600 bp<br />

550 bp<br />

1000 ng<br />

100 ng<br />

10 ng<br />

1 ng<br />

0.1 ng<br />

0.01 ng<br />

0.001 ng<br />

Figure 5. Ethidium bromide stained agarose<br />

gel <strong>of</strong> polymerase chain reaction products<br />

using ITS1 <strong>and</strong> ITS4 primers with 0.0001 to<br />

1000 ng fungal DNA <strong>of</strong> <strong>Septoria</strong> tritici <strong>and</strong> 1 µl<br />

plant DNA extract from healthy wheat leaves.<br />

550 bp <strong>and</strong> 600 bp are ITS rDNA amplification<br />

products from <strong>Septoria</strong> tritici <strong>and</strong> wheat,<br />

respectively.<br />

Days after inoculation<br />

0 3 6 9 14 18 22<br />

600 bp<br />

550 bp<br />

Karim<br />

600 bp Jennah<br />

550 bp Khotifa<br />

Figure 6. Ethidium bromide stained agarose<br />

gel <strong>of</strong> time course PCR amplification<br />

products using ITS1 <strong>and</strong> ITS4 primers with<br />

<strong>Septoria</strong> tritici infected wheat (susceptible;<br />

Karim <strong>and</strong> resistant; Jennah Khotifa) DNA.<br />

550 bp <strong>and</strong> 600 bp are ITS rDNA amplification<br />

products from S. tritici <strong>and</strong> wheat,<br />

respectively.<br />

assay. These observations are<br />

preliminary results towards a<br />

quantification <strong>of</strong> fungal<br />

proliferation within tissue to<br />

rapidly characterize varying levels<br />

<strong>of</strong> host resistance. Characterization<br />

<strong>of</strong> host resistance by measuring the<br />

fungal biomass inside wheat leaves<br />

was tested using ELISA (Kema et<br />

al., 1996c). This technique did not<br />

detect the fungal antigen within the<br />

48 h-8 dpi interval, although a<br />

slight increase <strong>of</strong> the fungal tissue<br />

was observed microscopically during<br />

that interval. Thus the ability <strong>of</strong> PCR<br />

to detect fungal mycelia in a<br />

susceptible cultivar at 3 dpi better<br />

reflects the colonization <strong>of</strong> wheat<br />

cells by the fungus. These<br />

preliminary results will also help<br />

establish an efficient protocol for<br />

fungicide utilization. Effective<br />

utilization <strong>of</strong> fungicides would<br />

decrease or maintain the intensity <strong>of</strong><br />

the PCR product a long time after its<br />

application.<br />

References<br />

Beck, J.J., <strong>and</strong> J.M. Ligon. 1995.<br />

Polymerase chain reaction assays for<br />

the detection <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum <strong>and</strong> <strong>Septoria</strong> tritici in wheat.<br />

Phytopathology 85:319-324.<br />

Kema, G.H.J., J.G. Annone, R. Sayoud,<br />

C.H. van Silfhout, M. van Ginkel,<br />

<strong>and</strong> J. Bree. 1996a. Genetic variation<br />

for virulence <strong>and</strong> resistance in the<br />

wheat-Mycosphaerella graminicola<br />

pathosystem I. Interaction between<br />

pathogen isolates <strong>and</strong> host cultivars.<br />

Phytopathology 86:200-211.<br />

Kema, G.H.J., R. Sayoud, J.G. Annone,<br />

<strong>and</strong> C.H. van Silfhout. 1996b.<br />

Genetic variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem II. Analysis <strong>of</strong><br />

interactions between isolates <strong>and</strong><br />

host cultivars. Phytopathology<br />

86:213-219.<br />

Kema, G.H.J., D. Yu, F.H.J. Rijkenberg,<br />

M.W. Shaw, <strong>and</strong> R.P. Baayen. 1996c.<br />

Histology <strong>of</strong> the pathogenesis <strong>of</strong><br />

Mycosphaerella graminicola in wheat.<br />

Phytopathology 86:777-786.<br />

Morjane, H., J. Geistlinger, M. Harrabi,<br />

K. Weising, <strong>and</strong> G. Kahl. 1995.<br />

Oligonucleotide fingerprinting<br />

detects genetic diversity among<br />

Ascochyta rabiei isolates from a single<br />

chickpea field in Tunisia. Curr.<br />

Genet. 26:191-197.<br />

Möller, E.M., G. Bahnweg, H.<br />

S<strong>and</strong>erman, <strong>and</strong> H.H. Geiger. 1992.<br />

A simple <strong>and</strong> efficient protocol for<br />

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

DNA from filamentous fungi, fruit<br />

bodies <strong>and</strong> infected plant tissue.<br />

Nucleic Acids Res. 20:6115-6116.


32<br />

Populations <strong>of</strong> <strong>Septoria</strong> spp. Affecting Winter Wheat in<br />

the Forest-Steppe Zone <strong>of</strong> the Ukraine<br />

S. Kolomiets*<br />

Institute <strong>of</strong> Plant Protection, Ukrainian Academy <strong>of</strong> Agrarian Sciences, Kiev, Ukraine<br />

Abstract<br />

Data in the literature <strong>and</strong> results <strong>of</strong> our investigations indicate that septoria leaf blotch <strong>of</strong> winter wheat has been reported<br />

annually in the forest-steppe zone <strong>of</strong> the Ukraine. <strong>Septoria</strong> tritici is the predominant species among the causal agents <strong>of</strong><br />

septoria leaf blotch <strong>of</strong> winter wheat. However, the portion caused by <strong>Stagonospora</strong> nodorum increased to 23-44% in recent<br />

years.<br />

<strong>Septoria</strong> tritici, the pathogen that<br />

causes leaf blotch, <strong>and</strong> <strong>Stagonospora</strong><br />

nodorum, the pathogen that induces<br />

spike <strong>and</strong> leaf blotch, are the most<br />

widespread pathogens <strong>of</strong> winter<br />

wheat in the forest-steppe zone <strong>of</strong><br />

the Ukraine.<br />

<strong>Septoria</strong> spp. induce a decrease<br />

in assimilation surface,<br />

developmental retardation,<br />

premature leaf desiccation, <strong>and</strong><br />

1000-grain weight. In epidemic<br />

years, yield losses may reach 30–<br />

50%. Total yield losses caused by<br />

these pathogens all over the world<br />

are estimated at 9 million tons.<br />

Developing cultivars resistant<br />

to the pathogens <strong>and</strong> establishing<br />

their cultivation is impossible<br />

without investigating the<br />

composition <strong>of</strong> pathogenic species<br />

in a given area <strong>and</strong> systematically<br />

recording its changes. Climatic<br />

conditions, the composition <strong>of</strong><br />

biocenoses, <strong>and</strong> the substrate<br />

where pathogens develop<br />

significantly affect the ratio<br />

between species. In the literature<br />

data on the areas occupied by the<br />

pathogens are quite limited, but<br />

can be found in articles published by<br />

Kovalenko (1975), Vasetskaja et al.<br />

(1983), Dyak (1990), <strong>and</strong> Sanina <strong>and</strong><br />

Antsiferova (1991).<br />

In 1995–97, we investigated the<br />

composition <strong>of</strong> <strong>Septoria</strong> pathogens in<br />

the forest-steppe zone <strong>of</strong> the Ukraine<br />

(Kyiv, Cherkasy, Vinnytsya,<br />

Khmelnytskyy, Ternopil, Zhytomyr,<br />

<strong>and</strong> Poltava regions). Studies were<br />

conducted on promising <strong>and</strong><br />

cultivated cultivars <strong>of</strong> winter wheat<br />

using routine methodologies. The<br />

species were identified via<br />

evaluation <strong>of</strong> stable traits: form,<br />

length, <strong>and</strong> width <strong>of</strong> conidia <strong>and</strong><br />

ends.<br />

Our research demonstrated that<br />

the forest-steppe zone <strong>of</strong> the Ukraine<br />

is occupied by both <strong>Septoria</strong><br />

pathogens, but S. tritici<br />

predominated in the 1970s, while S.<br />

nodorum was found only in certain<br />

years (Kovalenko, 1975). In the 1980s<br />

S. nodorum was reported in Kyiv (30-<br />

36%) <strong>and</strong> Ternopil (13-23%) regions<br />

(Dyak, 1990). In Cherkasy region<br />

(6.6%) the pathogen was reported<br />

only in certain years <strong>and</strong> was not<br />

detected at all in Vinnytsya region.<br />

* Author prevented from attending workshop by unforeseen travel problems.<br />

In the mid 1990s, the proportion<br />

<strong>of</strong> S. nodorum increased among the<br />

species. <strong>Septoria</strong> tritici dominated in<br />

all investigated regions: the highest<br />

percentage was reported in<br />

Vinnytsya region (76%), while the<br />

lowest was reported in Ternopil<br />

region (53%). In Kyiv, Cherkasy,<br />

Poltava, Khmelnytskyy, <strong>and</strong><br />

Zhytomyr regions, it reached 68,<br />

65, 69, 72, <strong>and</strong> 72%, respectively<br />

(Figure 1). Data from the literature<br />

<strong>and</strong> our own research results<br />

indicate an increase in the area<br />

occupied by S. nodorum, as well as<br />

in its ratio <strong>of</strong> <strong>Septoria</strong> pathogens<br />

present.<br />

Changes in the ratio <strong>of</strong> the<br />

pathogens may be explained by<br />

changes in climatic conditions, the<br />

range <strong>of</strong> cultivars used <strong>and</strong>,<br />

possibly, by the spread <strong>of</strong> S.<br />

nodorum infection through seeds,<br />

especially in recent years, when<br />

seeds were not properly treated.<br />

Thus the spread <strong>of</strong> S. nodorum–the<br />

most aggressive <strong>and</strong> damaging<br />

<strong>Septoria</strong> pathogen–increased under<br />

the above conditions.


;; yy<br />

;; yy<br />

6<br />

28%<br />

1<br />

32%<br />

;y; ; y y ; ; y y 28%<br />

31%<br />

68%<br />

72%<br />

2<br />

72%<br />

; y 35% 69% ;y<br />

44% 23% 65%<br />

3<br />

77% ;y ;y 4;<br />

y<br />

56%<br />

5<br />

<strong>Stagonospora</strong> nodorum<br />

<strong>Septoria</strong> tritici<br />

7<br />

Figure 1. The proportion between leaf septoriose pathogens <strong>of</strong> winter wheats in the foreststeppe<br />

zone <strong>of</strong> the Ukraine.<br />

1-Kyiv, 2-Poltava, 3-Cherkasy, 4-Vinnytsya, 5-Ternopil, 6-Khmelnytskyy, 7-Zhytomyr regions.<br />

Populations <strong>of</strong> <strong>Septoria</strong> spp. Affecting Winter Wheat in the Forest-Steppe Zone <strong>of</strong> the Ukraine 33<br />

References<br />

Dyak, U.P. 1990. Distribution <strong>of</strong> the<br />

basic inducers <strong>of</strong> septoriosis on a<br />

winter wheat in Ukrainian SSR.<br />

Zaschitz rasteniy 3:7-9.<br />

Kovalenko, S.N. 1975. <strong>Septoria</strong> leaf<br />

blotch <strong>of</strong> a winter wheat in<br />

requirements forest-steppe zone <strong>of</strong><br />

Ukraine SSR. Avtoreferat<br />

diss…k<strong>and</strong>. biol. nauk. Kiev, 21 p.<br />

Sanina, A.A., <strong>and</strong> Antsiferova, L.V.<br />

1991. <strong>Septoria</strong> Sacc. species on<br />

wheat in the European part <strong>of</strong> the<br />

USSR. Mikilogia i fitopatologia<br />

25(3):250-252.<br />

Vasetskaja, M.N., Kulikova, G.N., <strong>and</strong><br />

Borzionova, T.I. 1983. <strong>Septoria</strong><br />

species <strong>of</strong> fungi distributed on<br />

grades <strong>of</strong> wheat in the USSR.<br />

Mykologia i phytopatologia<br />

17(3):210-213.


34<br />

<strong>Septoria</strong> passerinii Closely Related to the Wheat<br />

Pathogen Mycosphaerella graminicola<br />

S.B. Goodwin <strong>and</strong> V.L. Zismann<br />

USDA-ARS, Department <strong>of</strong> Botany <strong>and</strong> Plant Pathology, Purdue University, West Lafayette, IN, USA<br />

Abstract<br />

<strong>Septoria</strong> passerinii is known only from its anamorphic <strong>Septoria</strong> state; no teleomorph has been identified. In culture, S.<br />

passerinii looks very similar to Mycosphaerella graminicola from wheat. Comparisons <strong>of</strong> the nucleotide sequences <strong>of</strong> the<br />

internal transcribed spacer (ITS) regions <strong>of</strong> the ribosomal DNA <strong>of</strong> both species <strong>and</strong> <strong>of</strong> many other fungi in the Dothideales<br />

revealed that the ITS regions <strong>of</strong> S. passerinii <strong>and</strong> M. graminicola differed by only 10 bases out <strong>of</strong> 571 total. Therefore, these<br />

species are very closely related. Phylogenetic analysis showed that S. passerinii <strong>and</strong> M. graminicola grouped together within<br />

a large cluster <strong>of</strong> Mycosphaerella species. Thus, S. passerinii almost certainly has a Mycosphaerella teleomorph.<br />

<strong>Septoria</strong> passerinii causes<br />

speckled leaf blotch on barley. The<br />

colony <strong>and</strong> conidial morphologies<br />

<strong>of</strong> S. passerinii <strong>and</strong> the wheat<br />

pathogen Mycosphaerella graminicola<br />

are very similar. Both are<br />

pathogens <strong>of</strong> cereals, <strong>and</strong> they may<br />

be closely related. However,<br />

nothing is known about the<br />

evolutionary relationships <strong>of</strong> S.<br />

passerinii (Cunfer <strong>and</strong> Ueng, 1999);<br />

no fruiting structures have been<br />

found <strong>and</strong> its teleomorph remains<br />

unknown.<br />

Analyses <strong>of</strong> the internal<br />

transcribed spacer (ITS) region <strong>of</strong><br />

ribosomal DNA have been very<br />

useful for elucidating the<br />

phylogenetic relationships <strong>of</strong> fungi.<br />

This region contains the highly<br />

variable ITS1 <strong>and</strong> ITS2, separated<br />

by the conserved 5.8S ribosomal<br />

RNA gene, <strong>and</strong> is bounded by the<br />

highly conserved 18 <strong>and</strong> 28S<br />

ribosomal RNA genes. This greatly<br />

facilitates analysis, because<br />

polymerase chain reaction (PCR)<br />

primers within the conserved<br />

regions <strong>of</strong> the 18 <strong>and</strong> 28S genes<br />

amplify the intervening ITS region<br />

<strong>of</strong> approximately 600 bp.<br />

Furthermore, many fungal ITS<br />

sequences are available in<br />

GenBank, which exp<strong>and</strong>s the<br />

number <strong>of</strong> species available for<br />

comparison.<br />

The objective <strong>of</strong> this research<br />

was to assemble a database <strong>of</strong> ITS<br />

sequences to test the hypothesis<br />

that S. passerinii is a close relative <strong>of</strong><br />

M. graminicola. The alternative<br />

hypothesis is that S. passerinii could<br />

be more closely related to other<br />

cereal pathogens, such as the<br />

septoria nodorum pathogen <strong>of</strong><br />

wheat <strong>and</strong> barley, Phaeosphaeria<br />

nodorum.<br />

Materials <strong>and</strong> Methods<br />

Isolates <strong>of</strong> species <strong>of</strong><br />

Mycosphaerella, Leptosphaeria, <strong>and</strong><br />

Phaeosphaeria were obtained from<br />

various sources (Table 1). DNA was<br />

extracted from lyophilized tissue<br />

<strong>and</strong> the ITS regions were amplified<br />

using primers ITS4 <strong>and</strong> ITS5.<br />

Amplification was with the<br />

following cycling parameters: 94 C<br />

for 2 min, 30 cycles <strong>of</strong> 93 C for 30 s,<br />

53 C for 2 min, 72 C for 2 min, <strong>and</strong><br />

a final extension <strong>of</strong> 10 min at 72 C.<br />

Amplification products were<br />

purified <strong>and</strong> cloned. Each clone<br />

was sequenced in both directions<br />

on an automated DNA sequencer<br />

<strong>and</strong> several clones were sequenced<br />

per isolate.<br />

DNA sequence alignment,<br />

genetic distance calculation,<br />

bootstrap analysis (1000<br />

replications), <strong>and</strong> a neighborjoining<br />

tree was prepared using<br />

Clustal X (http://www-igbmc.ustrasbg.fr/BioInfo/ClustalX/<br />

Top.html). The final tree was<br />

printed using NJplot.<br />

Results<br />

The ITS regions <strong>of</strong> all S.<br />

passerinii isolates were 571 bases<br />

long, including the primer regions,<br />

<strong>and</strong> were essentially identical. A<br />

BLAST search on GenBank<br />

identified M. graminicola as the<br />

closest match to S. passerinii.<br />

Sequences <strong>of</strong> other species showing<br />

good matches with S. passerinii in<br />

the BLAST search were<br />

downloaded from GenBank, along<br />

with those <strong>of</strong> species <strong>of</strong><br />

Leptosphaeria, Phaeosphaeria, <strong>and</strong><br />

other fungi in the Dothideales<br />

(Table 1). A multiple alignment<br />

revealed that the ITS sequences <strong>of</strong><br />

S. passerinii <strong>and</strong> M. graminicola<br />

differed by only 10 bases (data not


Table 1. Sources <strong>of</strong> isolates or DNA sequences for the internal transcribed spacer database <strong>of</strong><br />

<strong>Septoria</strong> passerinii, Mycosphaerella graminicola, <strong>and</strong> other fungi in the Dothideales.<br />

Species Isolate Growth medium Source<br />

Cladosporium caryigenum LCF — GenBank<br />

Cladosporium herbarum MZKI B-1002 — GenBank<br />

Cladosporium sphaerospermum MZKI B-1005 — GenBank<br />

Dothidea hippophaeos CBS 186.58 — GenBank<br />

Dothidea insculpta CBS 189.58 — GenBank<br />

Hormonema dematioides — — GenBank<br />

Leptosphaeria bicolor — — GenBank<br />

Leptosphaeria conteca — Malt broth ATCC<br />

Leptosphaeria doliolum — — Genbank<br />

Leptosphaeria maculans — — Genbank<br />

Leptosphaeria microscopica — — Genbank<br />

Mycosphaerella citri — Yeast malt broth Tobin Peever<br />

Mycosphaerella fijiensis — PDB ATCC<br />

Mycosphaerella graminicola — — GenBank<br />

Mycosphaerella graminicola IPO323 Yeast malt broth Gert Kema<br />

Mycosphaerella graminicola T1 Yeast malt broth Minnesota<br />

Mycosphaerella graminicola T48 Yeast malt broth Indiana<br />

Mycosphaerella musicola — PDB ATCC<br />

Mycosphaerella pini — — GenBank<br />

Mycosphaerella tassiana CBS 111.82 — GenBank<br />

Ophiosphaerella herpotricha — — GenBank<br />

Ophiosphaerella korrae — — GenBank<br />

Phaeosphaeria avenaria — — GenBank<br />

Phaeosphaeria halima — Emerson YpSs ATCC<br />

Phaeosphaeria nodorum — — GenBank<br />

Phaeosphaeria spartinae — Emerson YpSs ATCC<br />

Phaeosphaeria typharum — Emerson YpSs ATCC<br />

Phaeotheca triangularis MZKI B-950 — GenBank<br />

Rhizopycnis vagum — — GenBank<br />

<strong>Septoria</strong> passerinii 22585 Yeast malt broth ATCC<br />

<strong>Septoria</strong> passerinii 26516 Yeast malt broth ATCC<br />

<strong>Septoria</strong> passerinii P70 Yeast malt broth North Dakota<br />

<strong>Septoria</strong> passerinii P71 Yeast malt broth North Dakota<br />

<strong>Septoria</strong> passerinii P78 Yeast malt broth Minnesota<br />

<strong>Septoria</strong> passerinii P83 Yeast malt broth North Dakota<br />

Trimmatostroma salinum MZKI B-962 — GenBank<br />

shown). The neighbor-joining tree<br />

showed that S. passerinii <strong>and</strong> M.<br />

graminicola clustered together along<br />

with most <strong>of</strong> the other<br />

Mycosphaerella species (Figure 1).<br />

Phaeosphaeria nodorum was in a very<br />

different cluster that mostly<br />

contained species <strong>of</strong> Phaeosphaeria,<br />

Leptosphaeria, <strong>and</strong> Ophiosphaerella<br />

(Figure 1). Bootstrap analysis<br />

indicated strong support for all <strong>of</strong><br />

the major clusters (Figure 1).<br />

Discussion<br />

These data clearly indicate that<br />

S. passerinii <strong>and</strong> M. graminicola are<br />

very closely related. Because these<br />

two taxa were deeply embedded<br />

within a large cluster <strong>of</strong><br />

<strong>Septoria</strong> passerinii Closely Related to the Wheat Pathogen Mycosphaerella graminicola 35<br />

Mycosphaerella species, it seems<br />

certain that S. passerinii has, or was<br />

derived from a progenitor that had,<br />

a Mycosphaerella teleomorph. This<br />

must be confirmed by searching for<br />

the Mycosphaerella stage on infected<br />

barley tissue.<br />

In addition to elucidating the<br />

evolutionary relationships <strong>of</strong> S.<br />

passerinii, the cluster analysis raises<br />

a number <strong>of</strong> interesting questions.<br />

For example, Phaeosphaeria,<br />

Leptosphaeria, <strong>and</strong> Ophiosphaerella<br />

form a large cluster. However, it is<br />

not clear whether Phaeosphaeria<br />

warrants a separate generic<br />

designation. Clarifying the<br />

relationships among the mitosporic<br />

Dothideales (those with no known<br />

sexual stage) will require a larger<br />

number <strong>of</strong> species <strong>and</strong> a more<br />

comprehensive analysis. Aligning<br />

the 5.8S <strong>and</strong> ITS2 regions was fairly<br />

straightforward. However, ITS1<br />

varied greatly in length, which<br />

made alignment difficult. The<br />

alignment used for the cluster<br />

analysis presented here has not<br />

been optimized <strong>and</strong>, thus, should<br />

be considered preliminary.<br />

However, the close relationship<br />

between S. passerinii <strong>and</strong> M.<br />

graminicola is certain.<br />

Reference<br />

Cunfer, B.M., <strong>and</strong> Ueng, P.P. 1999.<br />

Taxonomy <strong>and</strong> identification <strong>of</strong><br />

<strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong> on small<br />

grain cereals. Annual Review <strong>of</strong><br />

Phytopathology (in press).


36<br />

Session 1 — S.B. Goodwin <strong>and</strong> V.L. Zismann<br />

0.05<br />

100<br />

96<br />

100<br />

100<br />

97.4<br />

100<br />

100<br />

100<br />

100<br />

100<br />

100<br />

100<br />

100<br />

99.9<br />

99.6<br />

Leptosphaeria bicolor<br />

Leptosphaeria conteca<br />

Ophiosphaerella korrae<br />

Ophiosphaerellaherpotricha<br />

Phaeosphaeria typharum<br />

Phaeosphaeria halima<br />

Phaeosphaeria avenaria<br />

Phaeosphaeria nodorum<br />

Phaeosphaeria spartinae<br />

Leptosphaeria microscopica<br />

Leptosphaeria maculans<br />

Leptosphaeria doliolum<br />

Rhizopycnis vagum<br />

<strong>Septoria</strong> passerinii P78<br />

<strong>Septoria</strong> passerinii P26516<br />

<strong>Septoria</strong> passerinii P71<br />

<strong>Septoria</strong> passerinii P83<br />

<strong>Septoria</strong> passerinii P22585<br />

<strong>Septoria</strong> passerinii P70<br />

Mycosphaerella graminicola T1<br />

Mycosphaerella graminicola<br />

Mycosphaerella graminicola IPO<br />

Mycosphaerella graminicola T48<br />

Mycosphaerella citri<br />

Mycosphaerella musicola<br />

Mycosphaerella fijiensis<br />

Mycosphaerella pini<br />

100 Dothidea hippophaeos<br />

Dothidea insculpta<br />

Hormonema dematioides<br />

Mycosphaerella tassiana<br />

Cladosporium herbarum<br />

Cladosporium sphaerospermum<br />

Trimmatostroma salinum<br />

Phaeotheca triangularis<br />

Cladosporium caryigenum<br />

Figure 1. Phylogenetic analysis <strong>of</strong> sequences <strong>of</strong> the internal transcribed spacer region <strong>of</strong> the ribosomal DNA <strong>of</strong><br />

<strong>Septoria</strong> passerinii, Mycosphaerella graminicola <strong>and</strong> other fungi in the Dothideales. All bootstrap values<br />

above 90% are indicated.


<strong>Septoria</strong>/<strong>Stagonospora</strong> Leaf Spot <strong>Diseases</strong> on Barley in<br />

North Dakota, USA<br />

J.M. Krupinsky1 <strong>and</strong> B.J. Steffenson2 (Poster)<br />

1 USDA, Agricultural Research Service, Northern Great Plains Research Laboratory, M<strong>and</strong>an, ND, USA<br />

2 Dept. <strong>of</strong> Plant Pathology, North Dakota State Univ., Fargo, ND, USA<br />

Abstract<br />

Diseased barley leaves were collected from fields in North Dakota in 1998. The most common <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

diseases were septoria speckled leaf blotch (<strong>Septoria</strong> passerinii) <strong>and</strong> stagonospora avenae leaf blotch (<strong>Stagonospora</strong><br />

avenae f. sp. triticea). Net blotch (Drechslera teres) <strong>and</strong> spot blotch (Bipolaris sorokiniana) were also commonly<br />

present. <strong>Stagonospora</strong> nodorum blotch (<strong>Stagonospora</strong> nodorum) <strong>and</strong> tan spot (Drechslera tritici-repentis), which are<br />

major diseases on wheat in the area, were detected on barley but at rather low levels. When isolates <strong>of</strong> S. nodorum from<br />

barley were tested on wheat <strong>and</strong> barley in greenhouse inoculations, higher symptom severity ratings were obtained on wheat<br />

compared to barley, indicating that the isolates obtained from barley were probably wheat-type isolates. Because <strong>of</strong> the<br />

importance <strong>of</strong> both <strong>Septoria</strong> passerinii <strong>and</strong> <strong>Stagonospora</strong> avenae f. sp. triticea on barley, selecting barley for <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> resistance will require screening with both organisms.<br />

Barley (Hordeum vulgare L.) can<br />

be affected by a number <strong>of</strong> plant<br />

diseases that can cause economic<br />

losses in yield <strong>and</strong> quality. The<br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> diseases<br />

common on barley are septoria<br />

speckled leaf blotch (<strong>Septoria</strong><br />

passerinii Sacc.), stagonospora<br />

avenae leaf blotch (<strong>Stagonospora</strong><br />

avenae Bissett f. sp. triticea T.<br />

Johnson [syn. <strong>Septoria</strong> avenae A.B.<br />

Frank f. sp. triticea T. Johnson]), <strong>and</strong><br />

stagonospora nodorum blotch<br />

(<strong>Stagonospora</strong> nodorum [Berk.]<br />

Castellani & E.G. Germano [syn.<br />

<strong>Septoria</strong> nodorum {Berk.} Berk. in<br />

Berk & Broome]) (Kiesling, 1985;<br />

Mathre, 1997). Barley-type isolates<br />

<strong>of</strong> S. nodorum have also been<br />

identified (Cunfer <strong>and</strong> Youmans,<br />

1983; Smedegard-Peterson, 1974). A<br />

cooperative survey was undertaken<br />

to determine the importance <strong>of</strong><br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> leaf spot<br />

diseases common on barley in<br />

North Dakota.<br />

Materials <strong>and</strong> Methods<br />

Diseased barley leaves (green<br />

leaves with leaf spots) were<br />

gathered from naturally-infected<br />

barley plants in the field in 1998.<br />

Leaves were collected from 70<br />

fields located in the southwestern,<br />

central, northeastern, <strong>and</strong> eastern<br />

areas <strong>of</strong> North Dakota. Collected<br />

leaves were dried <strong>and</strong> stored dry at<br />

4C in a refrigerator until processed.<br />

Leaf sections, 2 cm long, from<br />

approximately 8 leaves per<br />

collection, were processed. Leaf<br />

sections were surface-sterilized for<br />

3 min in a 1% sodium hypochlorite<br />

solution containing a surfactant,<br />

rinsed in sterile distilled water,<br />

plated on 2% water agar in plastic<br />

Petri dishes, <strong>and</strong> incubated under a<br />

12-h photoperiod (cool-white<br />

fluorescent tubes) at 21C. After 7<br />

days, leaf sections were examined<br />

for fungi. Pycnidiospores from<br />

pycnidia on the leaf sections were<br />

identified microscopically. The<br />

presence <strong>of</strong> Drechslera teres (Sacc.)<br />

Shoemaker, Bipolaris sorokiniana<br />

(Sacc.) Shoemaker, <strong>and</strong> D. tritici-<br />

37<br />

repentis (Died.) Shoemaker was also<br />

noted. Two to four fungal species<br />

were present on some leaf sections.<br />

Isolates were obtained, grown,<br />

stored, <strong>and</strong> inoculum was prepared<br />

as described by Krupinsky (1997).<br />

Nine isolates <strong>of</strong> S. nodorum<br />

obtained from barley were<br />

compared in glasshouse<br />

inoculations <strong>of</strong> wheat <strong>and</strong> barley<br />

seedling plants. Two wheat<br />

cultivars, Eureka <strong>and</strong> Fortuna, <strong>and</strong><br />

two barley cultivars, Bowman <strong>and</strong><br />

Hector, were used. In a glasshouse,<br />

seedlings were planted, grown,<br />

fertilized, inoculated, incubated,<br />

<strong>and</strong> assessed for percentage<br />

necrosis <strong>of</strong> the first leaf as<br />

previously reported (Krupinsky<br />

1997).<br />

Results <strong>and</strong> Discussion<br />

Of the 531 leaf sections<br />

processed, 45% were infected with<br />

S. passerinii, 37% with S. avenae f.<br />

sp. triticea, <strong>and</strong> 14% with S.<br />

nodorum. Drechslera teres, B.<br />

sorokiniana, <strong>and</strong> D. tritici-repentis


38<br />

Session 1 — J.M. Krupinsky <strong>and</strong> B.J. Steffenson<br />

were identified on 55, 51, <strong>and</strong> 10%<br />

<strong>of</strong> the leaf sections, respectively.<br />

Using the number <strong>of</strong> leaf sections<br />

infected with a particular fungus as<br />

an indicator <strong>of</strong> the relative<br />

importance <strong>of</strong> a fungus, the most<br />

common diseases making up the<br />

leaf-spot disease complex on barley<br />

in North Dakota were in<br />

descending order <strong>of</strong> importance net<br />

blotch, spot blotch, septoria<br />

speckled leaf blotch, <strong>and</strong><br />

stagonospora avenae leaf blotch.<br />

This ranking was similar to those in<br />

Saskatchewan, Canada, north <strong>of</strong><br />

western North Dakota, where D.<br />

teres was considered the most<br />

common leaf spotting pathogen,<br />

followed by B. sorokiniana, which<br />

was more common than <strong>Septoria</strong><br />

spp. (Fern<strong>and</strong>ez et al., 1999). In<br />

Manitoba, Canada, north <strong>of</strong> eastern<br />

North Dakota, leaf spot diseases<br />

were considered to cause minimal<br />

damage in 1998. Drechslera teres <strong>and</strong><br />

B. sorokiniana were found in 90-93%<br />

<strong>of</strong> the barley fields <strong>and</strong> S. passerinii<br />

was recovered in 22% <strong>of</strong> the fields<br />

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

Two <strong>of</strong> these diseases, spot<br />

blotch <strong>and</strong> stagonospora avenae<br />

leaf blotch, are usually minor<br />

diseases on wheat in central North<br />

Dakota. <strong>Stagonospora</strong> nodorum<br />

blotch <strong>and</strong> tan spot, although major<br />

diseases on wheat in North Dakota,<br />

were only isolated at low levels<br />

from barley <strong>and</strong> are not considered<br />

to be serious problems on barley at<br />

the present time. Because <strong>of</strong> the<br />

importance <strong>of</strong> both S. passerinii <strong>and</strong><br />

S. avenae f. sp. triticea on barley in<br />

North Dakota, selecting barley for<br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> resistance will<br />

require screening with both<br />

pathogens.<br />

The nine isolates <strong>of</strong> S. nodorum<br />

obtained from barley consistently<br />

produced higher symptom severity,<br />

as measured by percentage<br />

necrosis, on wheat than on barley.<br />

Overall, Fortuna averaged 28%<br />

necrosis <strong>and</strong> Eureka averaged 29%<br />

necrosis, compared to less than 1%<br />

for Bowman <strong>and</strong> Hector.<br />

Acknowledgments<br />

The authors thank D. Wetch for<br />

technical assistance, as well as D.E.<br />

Mathre <strong>and</strong> M. Babadoost for their<br />

reviews <strong>and</strong> constructive<br />

comments.<br />

Mention <strong>of</strong> a trademark,<br />

proprietary product, or company<br />

by USDA personnel is intended for<br />

explicit description only <strong>and</strong> does<br />

not constitute a guarantee or<br />

warranty <strong>of</strong> the product by the<br />

USDA <strong>and</strong> does not imply its<br />

approval to the exclusion <strong>of</strong> other<br />

products that may also be suitable.<br />

USDA-ARS, Northern Plains Area,<br />

is an equal opportunity/<br />

affirmative action employer <strong>and</strong> all<br />

agency services are available<br />

without discrimination.<br />

References<br />

Cunfer, B.M., <strong>and</strong> Youmans, J. 1983.<br />

<strong>Septoria</strong> nodorum on barley <strong>and</strong><br />

relationships among isolates from<br />

several hosts. Phytopathology<br />

73:911-914.<br />

Fern<strong>and</strong>ez, M.R., Celetti, M.J., <strong>and</strong><br />

Hughes, G. 1999. Leaf diseases <strong>of</strong><br />

barley <strong>and</strong> oat in Saskatchewan in<br />

1998. Can. Plant Dis. Survey 79:75-<br />

77.<br />

Kiesling, R.L. 1985. The diseases <strong>of</strong><br />

barley. Chapter 10, pages 269-312.<br />

In: Barley. D.C. Rasmusson, ed.<br />

American Soc. <strong>of</strong> Agronomy.<br />

Madison, WI. 522 pages.<br />

Krupinsky, J.M. 1997. Aggressiveness<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum isolates<br />

obtained from wheat in the<br />

Northern Great Plains. Plant<br />

Disease 81:1027-1031.<br />

Mathre, D.E., 1997. Compendium <strong>of</strong><br />

Barley <strong>Diseases</strong>. Second edition.<br />

The American Phytopathological<br />

Society, APS Press, St. Paul, MN.<br />

90 pages.<br />

Smedegard-Peterson, V. 1974.<br />

Leptosphaeria nodorum (<strong>Septoria</strong><br />

nodorum), a new pathogen on<br />

barley in Denmark, <strong>and</strong> its<br />

physiologic specialization on<br />

barley <strong>and</strong> wheat. Friesia 10:251-<br />

264.<br />

Tekauz, A., McCallum, B., Gilbert, J.,<br />

Mueller, E., Idris, M., Stulzer, M.,<br />

Beyene, M., <strong>and</strong> Parmentier, M.<br />

1999. Foliar diseases <strong>of</strong> barley in<br />

Manitoba in 1998. Can. Plant Dis.<br />

Survey 79:74.


Interrelations among <strong>Septoria</strong> tritici Isolates <strong>of</strong><br />

Varying Virulence<br />

S. Ezrati, S. Schuster, A. Eshel, <strong>and</strong> Z. Eyal (Poster)<br />

Department <strong>of</strong> Plant Science, Tel Aviv University, Israel<br />

Reduction in pycnidial<br />

coverage on Seri 82 was found by<br />

Halperin et al. (1996) after<br />

inoculation with two <strong>Septoria</strong> tritici<br />

isolates: ISR398 (avirulent) <strong>and</strong><br />

ISR8036 (virulent). Cross protection<br />

was suggested as a mechanism for<br />

the suppression phenomenon,<br />

triggered by the avirulent isolate.<br />

The aim <strong>of</strong> the current study was to<br />

investigate possible interactions<br />

between new isolates <strong>of</strong> S. tritici on<br />

Seri 82 <strong>and</strong> other wheat cultivars.<br />

Co-inoculations with mixtures<br />

<strong>of</strong> isolates were conducted on<br />

seedlings <strong>and</strong> maturing plants <strong>of</strong><br />

various cultivars. The following S.<br />

tritici isolates <strong>and</strong> wheat cultivars<br />

were used:<br />

• ISR398 - avirulent on Seri 82,<br />

KK, Bobwhite “S” <strong>and</strong> IAS20;<br />

virulent on Shafir.<br />

• ISR8036 - avirulent on KK,<br />

Bobwhite “S” <strong>and</strong> IAS20;<br />

virulent on Shafir <strong>and</strong> Seri 82.<br />

• ISR9812 - avirulent on IAS20;<br />

virulent on Shafir, Seri 82, KK <strong>and</strong><br />

Bobwhite “S”.<br />

• ISR9840 - avirulent on KK <strong>and</strong><br />

Bobwhite “S”; virulent on Shafir,<br />

Seri 82 <strong>and</strong> IAS20.<br />

Pairs <strong>of</strong> isolates were used for<br />

inoculation with 1:1 mixtures on<br />

each <strong>of</strong> the five cultivars. Pycnidial<br />

coverage by the virulent isolate <strong>of</strong><br />

each pair was used as a reference.<br />

The results are shown in Table 1.<br />

Significant suppression <strong>of</strong> pycnidial<br />

coverage was observed when<br />

seedlings were inoculated with<br />

mixtures composed <strong>of</strong> a virulent <strong>and</strong><br />

an avirulent isolate. However, no<br />

suppression was found when isolate<br />

ISR9840 was used as the virulent<br />

component. Since all isolates were<br />

virulent on Shafir, no suppression<br />

was found on this cultivar.<br />

Suppression <strong>of</strong> pycnidial<br />

coverage was observed on maturing<br />

plants <strong>of</strong> Seri 82 in the field<br />

following co-inoculation at GS 37<br />

with a 1:1 isolate mixture <strong>of</strong><br />

Table 1. Reduction in pycnidial coverage on seedlings <strong>of</strong> the wheat cultivars Shafir (SHF), Seri<br />

82 (SER), IAS 20, Bobwhite “S” (BOW) <strong>and</strong> Kavkaz/K4500 (KK), inoculated with 1:1 mixture <strong>of</strong><br />

<strong>Septoria</strong> tritici. Data are percent reduction as compared to pycnidial coverage by the more<br />

virulent isolate in each case.<br />

Wheat cultivars<br />

Isolate mixture SHF SER IAS20 BOW KK<br />

ISR398 + ISR8036 52.2 98.7 - a - -<br />

ISR398 + ISR9812 38.3 78.0 - 99.5 99.7<br />

ISR398 + ISR9840 11.0 7.0 21.6 21.9 -<br />

ISR8036 + ISR9812 16.3 28.9 - - 71.8<br />

ISR8036 + ISR9840 26.7 24.3 33.8 27.4 -<br />

ISR9812 + ISR9840 26.1 24.3 32.8 63.4 100.0<br />

a Cultivar resistant to both isolates.<br />

39<br />

ISR398+ISR8036. The other five<br />

isolate mixtures on all wheat<br />

cultivar combinations are currently<br />

being tested in field trials.<br />

The identity <strong>of</strong> pycnidia<br />

developed on leaves <strong>of</strong> Seri 82 <strong>and</strong><br />

Shafir, co-inoculated with<br />

ISR398+ISR8036, was verified using<br />

specific PCR primers. Isolate<br />

ISR8036 dominated (>70%) the<br />

pycnidial population on seedlings<br />

<strong>and</strong> from the onset <strong>of</strong> the epidemics<br />

in field plots <strong>of</strong> both Shafir <strong>and</strong> Seri<br />

82. The low fraction <strong>of</strong> ISR398 on<br />

the susceptible cultivar Shafir is<br />

explained by competition. In the<br />

mixture ISR398+ISR8036, isolate<br />

ISR8036 has a competitive<br />

advantage over ISR398, attributed<br />

to its higher aggressiveness. The<br />

same experimental approach will be<br />

adapted to the other isolate<br />

mixtures under investigation, once<br />

specific PCR primers are obtained<br />

for ISR9812 <strong>and</strong> ISR9840.<br />

The structure <strong>of</strong> natural<br />

populations <strong>of</strong> this pathogen is<br />

determined by both cultivar-isolate<br />

interactions <strong>and</strong> isolate-isolate<br />

interactions.<br />

References<br />

Halperin, T., Schuster, S., Pnini-Cohen,<br />

S., Zilberstein, A., <strong>and</strong> Eyal, Z. 1996.<br />

The suppression <strong>of</strong> pycnidial<br />

production on wheat seedlings<br />

following sequential inoculation by<br />

isolates <strong>of</strong> <strong>Septoria</strong> tritici.<br />

Phytopathology 86:728-732.


Session 2: The Infection Process<br />

<strong>Stagonospora</strong> <strong>and</strong> <strong>Septoria</strong> Pathogens <strong>of</strong> <strong>Cereals</strong>: The<br />

Infection Process<br />

B.M. Cunfer<br />

Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Georgia, Griffin, GA<br />

Abstract<br />

Definitive information on the infection process has been reported for <strong>Stagonospora</strong> nodorum, <strong>Septoria</strong> tritici, <strong>and</strong><br />

<strong>Septoria</strong> passerinii. Like other necrotrophic pathogens, they do not elicit the hypersensitive reaction. A significant difference<br />

in the infection process between <strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong> pathogens is that spore germination <strong>and</strong> penetration proceed<br />

much faster for S. nodorum than for S. tritici <strong>and</strong> S. passerinii. The <strong>Septoria</strong> pathogens penetrate the leaf primarily<br />

through stomata, whereas S. nodorum penetrates both directly <strong>and</strong> through stomata. <strong>Stagonospora</strong> nodorum kills the<br />

epidermal cells quickly, but S. tritici <strong>and</strong> S. passerinii do not kill epidermal cells until hyphae have ramified through the leaf<br />

mesophyll <strong>and</strong> rapid necrosis begins. Resistance slows host colonization but has no appreciable effect on the process <strong>of</strong> lesion<br />

development. The mechanisms controlling host response are still unclear. The infection process for ascospores is probably very<br />

similar to that for pycnidiospores. Ascospores <strong>of</strong> Phaeosphaeria nodorum germinate over a wide range <strong>of</strong> temperatures <strong>and</strong><br />

their germ tubes penetrate the leaf directly. However, unlike pycnidiospores, the ascospores do not germinate in free water.<br />

The infection process has been<br />

studied most intensely for<br />

<strong>Stagonospora</strong> nodorum <strong>and</strong> <strong>Septoria</strong><br />

tritici. One in-depth study on<br />

<strong>Septoria</strong> passerinii is available.<br />

Nearly all <strong>of</strong> the information<br />

reported is for infection by<br />

pycnidiospores. However, the<br />

infection process for other spore<br />

forms is quite similar. The<br />

information presented is mostly for<br />

infection <strong>of</strong> leaves under optimum<br />

conditions. Some studies were<br />

done with intact seedling plants,<br />

whereas others were conducted<br />

with detached leaves. Infection <strong>of</strong><br />

the wheat coleoptile <strong>and</strong> seedling<br />

by S. nodorum was described in<br />

detail by Baker (1971) <strong>and</strong><br />

reviewed by Cunfer (1983).<br />

Although no precise comparisons<br />

have been made, it appears that the<br />

infection process has many<br />

similarities in each host-parasite<br />

system <strong>and</strong> is typical <strong>of</strong> many<br />

necrotrophic pathogens.<br />

Information on factors influencing<br />

symptom development <strong>and</strong> disease<br />

expression are excluded but have<br />

been reviewed by other authors<br />

(Eyal et al., 1987; King et al., 1983;<br />

Shipton et al., 1971). A summary <strong>of</strong><br />

factors affecting spore longevity on<br />

the leaf surface is included.<br />

Role <strong>of</strong> the Cirrus <strong>and</strong><br />

Spore Survival on the<br />

Leaf Surface<br />

The most detailed information<br />

on the function <strong>of</strong> the cirrus<br />

encasing the pycnidiospores exuded<br />

from the pycnidium is for S.<br />

nodorum. The cirrus is a gel<br />

composed <strong>of</strong> proteinatous <strong>and</strong><br />

saccharide compounds. Its<br />

composition <strong>and</strong> function are<br />

similar to that <strong>of</strong> other fungi in the<br />

Sphaeropsidales (Fournet, 1969;<br />

Fournet et al., 1970; Griffiths <strong>and</strong><br />

Peverett, 1980).<br />

41<br />

The primary roles <strong>of</strong> cirrus<br />

components are protection <strong>of</strong><br />

pycnidiospores from dessication<br />

<strong>and</strong> prevention <strong>of</strong> premature<br />

germination. The cirrus protects the<br />

pycnidiospores so that some<br />

remain viable at least 28 days<br />

(Fournet, 1969). When the cirrus<br />

was diluted with water, if the<br />

concentration <strong>of</strong> cirrus solution was<br />

>20%, less that 10% <strong>of</strong><br />

pycnidiospores germinated. At a<br />

lower concentration, the<br />

components provide nutrients that<br />

stimulate spore germination <strong>and</strong><br />

elongation <strong>of</strong> germ tubes. Germ<br />

tube length increased up to 15%<br />

cirrus concentration, then declined<br />

moderately at higher<br />

concentrations (Harrower, 1976).<br />

Brennan et al. (1986) reported<br />

greater germination in dilute cirrus<br />

fluid. Cirrus components reduced<br />

germination at 10-60% relative<br />

humidity. Once spores are


42<br />

Session 2 — B.M. Cunfer<br />

dispersed, the stimulatory effects <strong>of</strong><br />

the cirrus fluid are probably<br />

negligible (Griffiths <strong>and</strong> Peverett,<br />

1980). At 35-45% relative humidity,<br />

spores <strong>of</strong> S. tritici in cirri remained<br />

viable at least 60 days (Gough <strong>and</strong><br />

Lee, 1985). The cirrus components<br />

may act as an inhibitor <strong>of</strong> spore<br />

germination, or the high osmotic<br />

potential <strong>of</strong> the cirrus may prevent<br />

germination.<br />

Pycnidiospores <strong>of</strong> S. nodorum<br />

did not survive for 24 hours at<br />

relative humidity above 80% at 20<br />

C. Spores survived two weeks or<br />

more at


<strong>Stagonospora</strong> nodorum releases a<br />

wide range <strong>of</strong> cell wall degrading<br />

enzymes including amylase, pectin<br />

methyl esterase,<br />

polygalacturonases, xylanases, <strong>and</strong><br />

cellulase in vitro <strong>and</strong> during<br />

infection <strong>of</strong> wheat leaves (Baker,<br />

1969; Lehtinen, 1993; Magro, 1984).<br />

The information related to cell wall<br />

degradation by enzymes agrees<br />

with histological observations.<br />

These enzymes may act in<br />

conjunction with toxins. Enzyme<br />

sensitivity may be related to<br />

resistance <strong>and</strong> rate <strong>of</strong> fungal<br />

colonization (Magro, 1984).<br />

Like many necrotrophs,<br />

<strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong><br />

pathogens produce phytotoxic<br />

compounds in vitro. Cell<br />

deterioration <strong>and</strong> death in advance<br />

<strong>of</strong> hyphal growth into mesophyll<br />

tissue (Bird <strong>and</strong> Ride, 1981) is<br />

consistent with toxin production.<br />

However, a definitive role for<br />

toxins in the infection process <strong>and</strong><br />

their relation to host resistance has<br />

not been established (Bethenod et<br />

al, 1982; Bousquet et al, 1980; Essad<br />

<strong>and</strong> Bousquet, 1981; King et al,<br />

1983). Differences in host range<br />

between wheat <strong>and</strong> barleyadapted<br />

strains <strong>of</strong> S. nodorum may<br />

be related to toxin production<br />

(Bousquet <strong>and</strong> Kollmann, 1998).<br />

Initiation <strong>of</strong> spore germination<br />

<strong>and</strong> percentage <strong>of</strong> spores<br />

germinated are not influenced by<br />

host susceptibility (Bird <strong>and</strong> Ride,<br />

1981; Morgan 1974; Straley, 1979;<br />

Straley <strong>and</strong> Scharen, 1979; Baker<br />

<strong>and</strong> Smith, 1978). Bird <strong>and</strong> Ride<br />

(1981) reported that extension <strong>of</strong><br />

germ tubes on the leaf surface was<br />

slower on resistant than on<br />

susceptible cultivars. This<br />

mechanism, expressed at least 48<br />

hours after spore deposition,<br />

indicates pre-penetration resistance<br />

to elongation <strong>of</strong> germ tubes. There<br />

were fewer successful penetrations<br />

in resistant cultivars, <strong>and</strong><br />

penetration proceeded more slowly<br />

on resistant cultivars (Baker <strong>and</strong><br />

Smith, 1978; Bird <strong>and</strong> Ride, 1981).<br />

Lignification was proposed to<br />

limit infection in both resistant <strong>and</strong><br />

susceptible cultivars, but other<br />

factors slowed fungal development<br />

in resistant lines. In susceptible<br />

lines, faster growing hyphae may<br />

escape lignification <strong>of</strong> host cells.<br />

Four days after inoculation <strong>of</strong><br />

barley with a wheat biotype isolate<br />

<strong>of</strong> S. nodorum, hyphae grew through<br />

the cuticle <strong>and</strong> sometimes in outer<br />

cellulose layers <strong>of</strong> epidermal cell<br />

walls. Thick papillae were<br />

deposited beneath the penetration<br />

hyphae <strong>and</strong> the cells were not<br />

penetrated (Keon <strong>and</strong> Hargreaves,<br />

1984).<br />

Infection by <strong>Septoria</strong><br />

tritici<br />

Pycnidiospores <strong>of</strong> S. tritici<br />

germinate in free water from both<br />

ends <strong>of</strong> the spore or from<br />

intercalary cells (Weber, 1922).<br />

Spore germination does not begin<br />

until about 12 hours after contact<br />

with the leaf. Germ tubes grow<br />

r<strong>and</strong>omly over the leaf surface.<br />

Weber (1922) observed only direct<br />

penetration between epidermal<br />

cells, but others concluded that<br />

penetration through both open <strong>and</strong><br />

closed stomata is the primary<br />

means <strong>of</strong> host penetration<br />

(Benedict, 1971; Cohen <strong>and</strong> Eyal,<br />

1993; Hilu <strong>and</strong> Bever, 1957). Kema<br />

<strong>Stagonospora</strong> <strong>and</strong> <strong>Septoria</strong> Pathogens <strong>of</strong> <strong>Cereals</strong>: The Infection Process 43<br />

et al. (1996) observed only stomatal<br />

penetration. Hyphae growing<br />

through stomata become<br />

constricted to about 1 µm diameter,<br />

then become wider after reaching<br />

the substomatal cavity.<br />

Hyphae grow parallel to the<br />

leaf surface under epidermal cells,<br />

then through the mesophyll to cells<br />

<strong>of</strong> lower the epidermis, but not into<br />

the epidermis. No haustoria are<br />

formed <strong>and</strong> hyphal growth is<br />

limited by sclerenchyma cells<br />

around the vascular bundles,<br />

except when hyphae are very<br />

dense. Vascular bundles are not<br />

invaded. Hyphae grow<br />

intercellularly along cell walls<br />

through the mesophyll, branching<br />

at a septum or middle <strong>of</strong> a cell. No<br />

macroscopic symptoms appear for<br />

about 9 days except for an<br />

occasional dead cell, but mesophyll<br />

cells die rapidly after 11 days.<br />

Pycnidia develop in substomatal<br />

chambers. Hyphae seldom grow<br />

into host cells (Hilu <strong>and</strong> Bever,<br />

1957; Kema et al, 1996;<br />

Weber, 1922).<br />

Successful infection only occurs<br />

after at least 20 hours <strong>of</strong> high<br />

humidity. Only a few brown flecks<br />

developed if leaves remained wet<br />

for 5-10 hours after spore<br />

deposition (Holmes <strong>and</strong> Colhoun,<br />

1974) or up to 24 hours (Kema et<br />

al., 1996). Host-parasite relations<br />

are the same on resistant or<br />

susceptible wheats. Spore<br />

germination on the leaf surface is<br />

the same regardless <strong>of</strong><br />

susceptibility. The number <strong>of</strong><br />

successful penetrations is about the<br />

same, but hyphal growth is faster<br />

in susceptible cultivars, resulting in<br />

more lesions. Hyphae extend


44<br />

Session 2 — B.M. Cunfer<br />

beyond the necrotic area in all<br />

cultivars. A toxin may play a role in<br />

pathogenesis (Cohen <strong>and</strong> Eyal,<br />

1993; Hilu <strong>and</strong> Bever, 1957). In<br />

contrast, colonization was greatly<br />

reduced on a resistant line (Kema et<br />

al., 1996).<br />

Infection by <strong>Septoria</strong><br />

passerinii<br />

Green <strong>and</strong> Dickson (1957)<br />

present a detailed description <strong>of</strong><br />

the infection process <strong>of</strong> S. passerinii<br />

on barley. The infection process is<br />

similar to S. tritici. Like S. tritici, the<br />

length <strong>of</strong> time required for leaf<br />

penetration is considerably longer<br />

than for S. nodorum.<br />

Germ tubes branch <strong>and</strong> grow<br />

over the leaf surface at r<strong>and</strong>om, but<br />

sometimes along depressions<br />

between epidermal cells. Leaf<br />

penetration is almost exclusively<br />

through stomata. Germination<br />

hyphae become swollen, <strong>and</strong> if<br />

penetration is unsuccessful, hyphae<br />

continue to elongate. No<br />

penetration occurs 48 hours after<br />

spore deposition. After 72 hours,<br />

germ tubes thicken over stomata,<br />

grow between guard cells <strong>and</strong> on<br />

surfaces <strong>of</strong> accessary cells <strong>and</strong> into<br />

the substomatal cavities. Direct<br />

penetration between epidermal<br />

cells is seen only rarely.<br />

Spore germination <strong>and</strong> host<br />

penetration are the same on<br />

resistant <strong>and</strong> susceptible cultivars.<br />

There is much less extension <strong>of</strong><br />

hyphae within leaves on resistant<br />

cultivars <strong>and</strong> papillae are observed<br />

on many but not all cell walls.<br />

Hyphae grow beneath the<br />

epidermis from one stoma to<br />

another, but do not penetrate<br />

between epidermal cells. The<br />

mesophyll is colonized, but no<br />

haustoria form. After the<br />

mesophyll cells become necrotic,<br />

epidermal cells collapse. Mycelial<br />

development in the leaf is sparse<br />

<strong>and</strong> usually blocked by vascular<br />

bundles. In younger leaves, if the<br />

vascular sheath is less developed,<br />

hyphae pass between the bundle<br />

<strong>and</strong> the epidermis. Pycnidia form<br />

in substomatal cavities, mostly on<br />

the upper leaf surface (Green <strong>and</strong><br />

Dickson, 1957).<br />

Factors Affecting Spore<br />

Longevity on the Leaf<br />

Surface<br />

Among the <strong>Stagonospora</strong> <strong>and</strong><br />

<strong>Septoria</strong> pathogens <strong>of</strong> cereals,<br />

definitive information on the<br />

infection process has been reported<br />

only for S. nodorum, S. tritici, <strong>and</strong> S.<br />

passerinii. Like many other<br />

necrotrophic pathogens, neither<br />

group <strong>of</strong> pathogens elicit the<br />

hypersensitive reaction. A<br />

significant difference in the<br />

infection process between <strong>Septoria</strong><br />

<strong>and</strong> <strong>Stagonospora</strong> pathogens is that<br />

spore germination <strong>and</strong> penetration<br />

proceeds much faster for S.<br />

nodorum than for S. tritici <strong>and</strong> S.<br />

passerinii. This has a significant<br />

influence on disease epidemiology.<br />

The <strong>Septoria</strong> pathogens penetrate<br />

the plant primarily through<br />

stomata, whereas S. nodorum<br />

penetrates both directly <strong>and</strong><br />

through stomata. S. nodorum<br />

penetrates <strong>and</strong> kills the epidermal<br />

cells quickly, but S. tritici <strong>and</strong> S.<br />

passerinii do not kill epidermal cells<br />

until hyphae have ramified<br />

through the leaf mesophyll <strong>and</strong><br />

rapid necrosis begins.<br />

Histological studies <strong>of</strong> fungal<br />

growth following host penetration<br />

match the data generated from<br />

epidemiological studies <strong>of</strong> host<br />

resistance. Resistance slows the rate<br />

<strong>of</strong> host colonization but has no<br />

appreciable effect on the process <strong>of</strong><br />

lesion development. The<br />

mechanisms controlling host<br />

response, whether related to<br />

enzymes <strong>and</strong> toxins or other<br />

metabolites released by the<br />

pathogens during infection, are still<br />

unclear.<br />

There is little information about<br />

infection by ascospores. The<br />

infection process is probably very<br />

similar to that for pycnidiospores.<br />

Ascospores <strong>of</strong> Phaeosphaeria nodorum<br />

germinate over a wide range <strong>of</strong><br />

temperatures, <strong>and</strong> their germ tubes<br />

penetrate the leaf directly. However,<br />

according to Rapilly et al. (1973),<br />

ascospores, unlike pycnidiospores,<br />

do not germinate in free water.<br />

References<br />

Baker, C.J. 1971. Morphology <strong>of</strong><br />

seedling infection by Leptosphaeria<br />

nodorum. Trans. Brit. Mycol. Soc.<br />

56:306-309.<br />

Baker, C.J. 1969. Studies on<br />

Leptosphaeria nodorum Müller <strong>and</strong><br />

<strong>Septoria</strong> tritici Desm. on wheat. Ph.<br />

D. thesis. University <strong>of</strong> Exeter.<br />

Baker E.A., <strong>and</strong> Smith, I.M. 1978.<br />

Development <strong>of</strong> the resistant <strong>and</strong><br />

susceptible reactions in wheat on<br />

inoculation with <strong>Septoria</strong> nodorum.<br />

Trans. Brit. Mycol. Soc. 71:475-482.<br />

Benedict, W.G. 1971. Differential effect<br />

<strong>of</strong> light intensity on the infection <strong>of</strong><br />

wheat by <strong>Septoria</strong> tritici Desm.<br />

under controlled environmental<br />

conditions. Physiol. Pl. Pathol. 1:55-<br />

56.<br />

Bird, P.M., <strong>and</strong> Ride, J.P. 1981. The<br />

resistance <strong>of</strong> wheat to <strong>Septoria</strong><br />

nodorum: fungal development in<br />

relation to host lignification.<br />

Physiol. Pl. Pathol. 19:289-299.


Bethenod, O., Bousquet, J., Laffray, D.,<br />

<strong>and</strong> Louguet, P. 1982. Reexamen<br />

des modalités d’action de<br />

l’ochracine sur la conductance<br />

stomatique des feuilles de<br />

plantules de blé, Triticum aestivum<br />

L., cv “Etoile de Choisy”.<br />

Agronomie 2:99-102.<br />

Bousquet, J.F., <strong>and</strong> Kollman, A. 1998.<br />

Variation in metabolite production<br />

by <strong>Septoria</strong> nodorum isolates<br />

adapted to wheat or to barley. J.<br />

Phytopathol. 146:273-277.<br />

Bousquet, J.F., Belhomme de<br />

Franqueville, H., Kollmann, A.,<br />

<strong>and</strong> Fritz, R. 1980. Action de la<br />

septorine, phytotoxine synthétisée<br />

par <strong>Septoria</strong> nodorum, sur la<br />

phosphorylation oxydative dans<br />

les mitochondries isolées de<br />

coléoptiles de blé. Can. J. Bot.<br />

58:2575-2580.<br />

Brennan, R.M., Fitt, B.D.L., Colhoun,<br />

J., <strong>and</strong> Taylor, G.S. 1986. Factors<br />

affecting the germination <strong>of</strong><br />

<strong>Septoria</strong> nodorum pycnidiospores. J.<br />

Phytopathol. 117:49-53.<br />

Brönnimann, A., Sally, B.K., <strong>and</strong><br />

Sharp, E.L. 1972. Investigations on<br />

<strong>Septoria</strong> nodorum in spring wheat<br />

in Montana. Plant Dis. Reptr.<br />

56:188-191.<br />

Cohen, L., <strong>and</strong> Eyal, Z. 1993. The<br />

histology <strong>of</strong> processes associated<br />

with the infection <strong>of</strong> resistant <strong>and</strong><br />

susceptible wheat cultivars with<br />

<strong>Septoria</strong> tritici. Plant Pathol. 42:737-<br />

743.<br />

Cunfer, B.M. 1983. Epidemiology <strong>and</strong><br />

control <strong>of</strong> seed-borne <strong>Septoria</strong><br />

nodorum. Seed Sci. Tecnol. 11:707-<br />

718.<br />

Essad, S., <strong>and</strong> Bousquet, J.F. 1981.<br />

Action de l’ochracine, phytotoxine<br />

de <strong>Septoria</strong> nodorum Berk., sur le<br />

cycle mitotique de Triticum<br />

aestivum L. Agronomie 1:689-694.<br />

Eyal, Z., Scharen, A.L., Prescott, J.M.,<br />

<strong>and</strong> van Ginkel, M. 1987. The<br />

<strong>Septoria</strong> diseases <strong>of</strong> wheat:<br />

Concepts <strong>and</strong> methods <strong>of</strong> disease<br />

management. <strong>CIMMYT</strong>. Mexico<br />

D.F. 46 pp.<br />

Fern<strong>and</strong>es, J.M.C., <strong>and</strong> Hendrix, J.W.<br />

1986a. Viability <strong>of</strong> conidia <strong>of</strong><br />

<strong>Septoria</strong> nodorum exposed to<br />

natural conditions in Washington.<br />

Fitopathol. Bras. 11:705-710.<br />

Fern<strong>and</strong>es, J.M.C., <strong>and</strong> Hendrix, J.W.<br />

1986b. Leaf wetness <strong>and</strong><br />

temperature effects on the hyphal<br />

growth <strong>and</strong> symptoms<br />

development on wheat leaves<br />

infected by <strong>Septoria</strong> nodorum.<br />

Fitopathol. Bras. 11:835-845.<br />

Fournet, J. 1969. Properiétés et role du<br />

cirrhe du <strong>Septoria</strong> nodorum Berk.<br />

Ann. Phytopathol. 1:87-94.<br />

Fournet, J., Pauvert, P., <strong>and</strong> Rapilly, F.<br />

1970. Propriétés des gelées<br />

sporifères de quelques<br />

Sphaeropsidales et Melanconiales.<br />

Ann. Phytopathol. 2:31-41.<br />

Gough, F.J., <strong>and</strong> Lee, T.S. 1985.<br />

Moisture effects on the discharge<br />

<strong>and</strong> survival <strong>of</strong> conidia <strong>of</strong> <strong>Septoria</strong><br />

tritici. Phytopathology 75:180-182.<br />

Green, G.J., <strong>and</strong> Dickson, J.G. 1957.<br />

Pathological histology <strong>and</strong> varietal<br />

reactions in <strong>Septoria</strong> leaf blotch <strong>of</strong><br />

barley. Phytopathology 47:73-79.<br />

Griffiths, E., <strong>and</strong> Peverett, H. 1980.<br />

Effects <strong>of</strong> humidity <strong>and</strong> cirrhus<br />

extract on survival <strong>of</strong> <strong>Septoria</strong><br />

nodorum spores. Trans. Brit. Mycol.<br />

Soc. 75:147-150.<br />

Harrower, K.M. 1976. Cirrus function<br />

in Leptosphaeria nodorum. APPS<br />

Newsletter 5:20-21.<br />

Hilu, H.M., <strong>and</strong> Bever, W.M. 1957.<br />

Inoculation, oversummering, <strong>and</strong><br />

suscept-pathogen relationship <strong>of</strong><br />

on Triticum species.<br />

Phytopathology 47:474-480.<br />

Holmes, S.J.I., <strong>and</strong> Colhoun, J. 1974.<br />

Infection <strong>of</strong> wheat by <strong>Septoria</strong><br />

nodorum <strong>and</strong> S. tritici. Trans. Brit.<br />

Mycol. Soc. 63:329-338.<br />

Jenkins, P.D. 1978. Interaction in<br />

cereal diseases. Ph. D. Thesis.<br />

University <strong>of</strong> Wales.<br />

Karjalainen, R., <strong>and</strong> Lounatmaa, K.<br />

1986. Ultrastructure <strong>of</strong> penetration<br />

<strong>and</strong> colonization <strong>of</strong> wheat leaves<br />

by <strong>Septoria</strong> nodorum. Physiol. Mol.<br />

Pl. Pathol. 29:263-270.<br />

Kema, G.H.J., DaZhao, Y., Rijkenberg,<br />

F.H.J., Shaw, M.W., <strong>and</strong> Baayen,<br />

R.P. 1996. Histology <strong>of</strong> the<br />

pathogenesis <strong>of</strong> Mycosphaerella<br />

graminicola in wheat.<br />

Phytopathology 86:777-786.<br />

Keon, J.P.R., <strong>and</strong> Hargreaves, J.A.<br />

1984. The response <strong>of</strong> barley leaf<br />

epidermal cells to infection by<br />

<strong>Septoria</strong> nodorum. New Phytol.<br />

98:387-398.<br />

<strong>Stagonospora</strong> <strong>and</strong> <strong>Septoria</strong> Pathogens <strong>of</strong> <strong>Cereals</strong>: The Infection Process 45<br />

King, J.E., Cook, R.J., <strong>and</strong> Melville,<br />

S.C. 1983. A review <strong>of</strong> <strong>Septoria</strong><br />

diseases <strong>of</strong> wheat <strong>and</strong> barley. Ann.<br />

Appl. Biol. 103:345-373.<br />

Krupinsky, J.M., Scharen, A.L., <strong>and</strong><br />

Schillinger, J.A. 1973. Pathogen<br />

variation in <strong>Septoria</strong> nodorum Berk.,<br />

in relation to organ specificity,<br />

apparent photosynthetic rate <strong>and</strong><br />

yield <strong>of</strong> wheat. Physiol. Pl. Pathol.<br />

3:187-194.<br />

Lehtinen, U. 1993. Plant cell wall<br />

degrading enzymes <strong>of</strong> <strong>Septoria</strong><br />

nodorum. Physiol. Mol. Pl. Pathol.<br />

43:121-134.<br />

Magro, P. 1984. Production <strong>of</strong><br />

polysacchride-degrading enzymes<br />

in culture <strong>and</strong> during<br />

pathogenesis. Plant Sci. Letters<br />

37:63-68.<br />

Morgan, W.M. 1974. Physiological<br />

studies <strong>of</strong> diseases <strong>of</strong> wheat<br />

caused by <strong>Septoria</strong> spp. <strong>and</strong><br />

Fusarium culmorum. Ph. D. thesis.<br />

University <strong>of</strong> London.<br />

O’Reilly, P., <strong>and</strong> Downes, M.J. 1986.<br />

Form <strong>of</strong> survival <strong>of</strong> <strong>Septoria</strong><br />

nodorum on symptomless winter<br />

wheat. Trans. Brit. Mycol. Soc.<br />

86:381-385.<br />

Rapilly, F., Foucault, B., <strong>and</strong><br />

Lacazedieux, J. 1973. Études sur<br />

l’inoculum de <strong>Septoria</strong> nodorum<br />

Berk. (Leptosphaeria nodorum<br />

Müller) agent de la septoriose du<br />

blé I. Les ascospores. Ann.<br />

Phytopathol. 5:131-141.<br />

Rasanayagam, M.S., Paul, N.D.,<br />

Royle, D.J., <strong>and</strong> Ayres, P.G. 1995.<br />

Variation in responses <strong>of</strong> <strong>Septoria</strong><br />

tritici <strong>and</strong> S. nodorum to UV-B<br />

irradiation in vitro. Mycol. Res.<br />

99:1371-1377.<br />

Shipton, W.A., Boyd, W.R.J., Rosielle,<br />

A.A., <strong>and</strong> Shearer, B.L. 1971. The<br />

common <strong>Septoria</strong> diseases <strong>of</strong><br />

wheat. Bot. Rev. 37:231-262.<br />

Straley, M. L. 1979. Pathogenesis <strong>of</strong><br />

<strong>Septoria</strong> nodorum (Berk.) Berk. On<br />

wheat cultivars varying in<br />

resistance to glume blotch. Ph. D.<br />

thesis. Montana State University.<br />

99 pp.<br />

Straley, M.L., <strong>and</strong> Scharen, A.L. 1979.<br />

Development <strong>of</strong> <strong>Septoria</strong> nodorum<br />

in resistant <strong>and</strong> susceptible wheat<br />

leaves. Phytopathology 69:920-921.<br />

Weber, G.F. 1922. <strong>Septoria</strong> diseases <strong>of</strong><br />

cereals. Phytopathology 12:537-<br />

585.


46<br />

Aggressiveness <strong>of</strong> Phaeosphaeria nodorum Isolates <strong>and</strong><br />

Their In Vitro Secretion <strong>of</strong> Cell-Wall-Degrading Enzymes<br />

P. Halama, a F. Lalaoui, a V. Dumortier, a <strong>and</strong> B. Paul b<br />

a Institut Supérieur d’Agriculture, Université Catholique de Lille, Lille, France<br />

b Laboratoire des Sciences de la Vigne, Institut Jules Guyot, Université de Bourgogne, Dijon, France<br />

Abstract<br />

The relationship between the in vitro production <strong>of</strong> cell-wall-degrading enzymes <strong>and</strong> the aggressiveness <strong>of</strong> three<br />

Phaeosphaeria nodorum isolates was studied. When grown in liquid medium containing 1% cell wall from wheat leaves as<br />

the only carbon source, the isolates secreted xylanase, α-arabinosidase, β-xylosidase, polygalacturonase, β-galactosidase,<br />

cellulase, β-1.3-glucanase, β-glucosidase, acetyl-esterase, <strong>and</strong> butyrate-esterase. Time course experiments showed different<br />

levels <strong>of</strong> enzyme production <strong>and</strong> different kinetics between isolates. Xylanase, cellulase, polygalacturonase, <strong>and</strong> butyrateesterase<br />

were positively correlated with isolate aggressiveness. The most aggressive isolate produced a higher proportion <strong>of</strong><br />

xylanase than the other two isolates, suggesting the role <strong>of</strong> this enzyme in the pathogenesis <strong>of</strong> P. nodorum.<br />

Studies have revealed that<br />

Phaeosphaeria nodorum populations<br />

have significant variability in<br />

pathogenicity that has been<br />

measured by the ability <strong>of</strong> isolates<br />

to cause symptoms on wheat<br />

(Griffith <strong>and</strong> Ao, 1980; Allinghan<br />

<strong>and</strong> Jackson, 1981;Yang <strong>and</strong><br />

Hughes, 1986; Krupinsky, 1997).<br />

As other plant-pathogenic<br />

fungi, P. nodorum produces a range<br />

<strong>of</strong> cell-wall-degrading enzymes<br />

(CWDE) that enables it to penetrate<br />

<strong>and</strong> infect host tissues (Lehtinen,<br />

1993; Magro, 1984). Very little is<br />

known about the variability <strong>of</strong><br />

enzyme production in P. nodorum<br />

populations; there is, however, a<br />

need to assess the diversity <strong>of</strong> these<br />

wall-degrading enzymes <strong>and</strong>, in<br />

particular, the extent to which this<br />

diversity could lead to differences<br />

in pathogenicity.<br />

The objective <strong>of</strong> this study was<br />

to compare enzymatic variation<br />

between two wild isolates <strong>and</strong> one<br />

mutant isolate <strong>of</strong> P. nodorum by<br />

examining CWDEs, <strong>and</strong> to<br />

establish the possible relationship<br />

between enzymes associated to<br />

pathogenesis <strong>and</strong> the<br />

aggressiveness <strong>of</strong> fungal strains.<br />

Materials <strong>and</strong> Methods<br />

Fungal strains<br />

Two wild strains (A/5 <strong>and</strong> 6/T)<br />

that differed significantly in their<br />

pathogenic behavior were isolated<br />

as described by Rapilly et al. (1992).<br />

The other isolate (300.2) is a<br />

carbendazim (MBC) resistant strain<br />

that was obtained using the<br />

protocol described in a previous<br />

study (Halama et al., 1999).<br />

Growth conditions <strong>and</strong><br />

preparation <strong>of</strong> fungal<br />

extracts<br />

The pycniospores used for<br />

inoculation were produced on<br />

synthetic (S) medium (Halama <strong>and</strong><br />

Lacoste, 1992). For enzyme<br />

production, isolates <strong>of</strong> S. nodorum<br />

were grown in liquid medium<br />

(Lehtinen, 1993) in which sucrose<br />

was replaced by 1% wheat cell<br />

walls. Wheat cell wall material was<br />

taken from 1-week-old seedlings <strong>of</strong><br />

cv. ‘Soissons’ <strong>and</strong> prepared<br />

according to Lehtinen’s modified<br />

procedure (Lehtinen, 1993). After<br />

incubating cultures at 24°C under<br />

constant shaking (72 rpm), liquid<br />

filtrates were centrifuged at 8000 g<br />

for 15 min at 4°C. Supernatants were<br />

used for enzyme activity assays, <strong>and</strong><br />

each culture experiment was<br />

replicated three times.<br />

Enzyme assays<br />

Cell-wall-degrading enzymes <strong>of</strong><br />

P. nodorum were assayed from the<br />

crude culture filtrates using<br />

different methods. The<br />

dinitrosalicylic acid (DNS) modified<br />

method, described by Miller (1959)<br />

for determining the reducing group,<br />

was used for the endo-enzyme<br />

assay. Xylanase, endopolygalacturonase,<br />

cellulase, <strong>and</strong> β-<br />

1.3-glucanase activities were<br />

measured respectively on oat spelts<br />

xylan, galacturonic acid,<br />

carboxymethyl cellulose, <strong>and</strong><br />

laminarin used as substrates.<br />

Absorbances were measured at 540<br />

nm. The reference sugar was<br />

glucose.<br />

Glycosidase activities were<br />

measured using p-nitrophenyl<br />

glycoside as a substrate (Poutanen,<br />

1988). β-glucosidase, β-


galactosidase, β-xylosidase, <strong>and</strong> αarabinosidase<br />

activities were<br />

assayed using p-nitrophenyl-β-Dglucopyranoside,<br />

p-nitrophenyl-β-<br />

D-galactopyranoside, pnitrophenyl-β-D-xylopyranoside,<br />

<strong>and</strong> p-nitrophenyl-α-Larabinopyranoside,<br />

respectively, as<br />

substrates.<br />

Esterase acetate activity was<br />

tested as described by Biely et al.<br />

(1988) using p-nitrophenylacetate.<br />

Degradation <strong>of</strong> pnitrophenylbutyrate<br />

was measured<br />

using p-nitrophenylbutyrate<br />

according to the procedure <strong>of</strong><br />

Kolattukudy et al. (1981). In each<br />

case, p-nitrophenol was used for<br />

the st<strong>and</strong>ard microtitre plate. For<br />

all enzyme assays, one unit <strong>of</strong><br />

enzyme was defined as the amount<br />

<strong>of</strong> enzyme that hydrolyzed 1 mmol<br />

<strong>of</strong> appropriate substrate, <strong>and</strong> each<br />

activity was expressed in mU.ml -1 .<br />

Pathogenicity test<br />

Isolates were tested for their<br />

relative pathogenicity on detached<br />

seedling leaves <strong>of</strong> the cultivar<br />

‘Soissons’ as previously described<br />

by Halama et al. (1999).<br />

Statistical analysis<br />

The analysis <strong>of</strong> variance was<br />

performed with the aid <strong>of</strong> the<br />

STAT-ITCF statistical s<strong>of</strong>tware.<br />

Differences between the isolates<br />

were assessed using the Newman-<br />

Keuls test. Correlations between<br />

data for aggressiveness rates <strong>and</strong><br />

for enzyme secretion were<br />

examined.<br />

Results<br />

The maximum secretions <strong>of</strong><br />

CWDEs are presented in Table 1,<br />

which shows that P. nodorum<br />

isolates were able to produce a<br />

wide range <strong>of</strong> enzymes on<br />

Aggressiveness <strong>of</strong> Phaeosphaeria nodorum Isolates <strong>and</strong> Their In Vitro Secretion <strong>of</strong> Cell-Wall-Degrading Enzymes 47<br />

synthetic medium supplemented<br />

with isolated cell wall. The secreted<br />

enzymes were determined every<br />

second day over the growth period<br />

(14 days).<br />

Time course for production<br />

<strong>of</strong> CWDEs<br />

The A/5 isolate produced, in<br />

order <strong>of</strong> decreasing activity,<br />

xylanase, butyrate-esterase, βglucosidase,<br />

β-1.3-glucanase,<br />

acetyl-esterase, cellulase,<br />

polygalacturonase, β-galactosidase,<br />

<strong>and</strong> β-xylosidase (Table 1). Culture<br />

filtrates <strong>of</strong> the A/5 isolate had<br />

significantly higher levels <strong>of</strong><br />

xylanase, butyrate-esterase,<br />

cellulase, <strong>and</strong> polygalacturonase<br />

activities than those <strong>of</strong> 6/T <strong>and</strong><br />

300.2 isolates. Levels <strong>of</strong> enzyme<br />

activity between 6/T <strong>and</strong> 300.2<br />

isolates were significantly different<br />

mainly for two peaks <strong>of</strong> β-1.3glucanase<br />

<strong>and</strong> acetyl-esterase<br />

activity on days 10 <strong>and</strong> 14<br />

produced respectively by the 6/T<br />

<strong>and</strong> the 300.2 isolates. Xylanase<br />

activity by the A/5 isolate on day<br />

10 was approximately seven times<br />

higher (1574 mU.m1 -1 ) than that <strong>of</strong><br />

the two other isolates.<br />

No α-arabinosidase activity<br />

could be detected in the<br />

filtrate <strong>of</strong> A/5 isolate,<br />

<strong>and</strong> very low levels were<br />

detected for 6/T <strong>and</strong><br />

300.2 isolates. Very low<br />

β-xylosidase activity was<br />

detected both by A/5<br />

<strong>and</strong> 6/T isolates.<br />

The polygalacturonase activities<br />

were highest on day 6 for A/5<br />

isolate (56 mU.ml –1 ), after which<br />

the activity decreased. For 6/T <strong>and</strong><br />

300.2 isolates, the highest activities<br />

were 27 <strong>and</strong> 12 mU.ml -1<br />

respectively, showing differences in<br />

their kinetics. As early as the<br />

second day, isolate 6/T produced a<br />

large amount <strong>of</strong> polygalacturonase,<br />

although isolate 300.2 secreted<br />

maximum polygalacturonase only<br />

after 12 days.<br />

β-1.3-glucanase assay was<br />

detected mainly for A/5 <strong>and</strong> 6/T<br />

isolates. β-glucosidase activity was<br />

produced by the three isolates with<br />

maximum activity on day 14 for the<br />

A/5 <strong>and</strong> 300.2 isolates <strong>and</strong> on day<br />

10 for the 6/T isolate. However, A/<br />

5 <strong>and</strong> 300.2 isolates may not have<br />

reached their peaks <strong>of</strong> βglucosidase<br />

activity at the end <strong>of</strong><br />

the time course.<br />

Where detected, the cellulase<br />

activity secreted by P. nodorum<br />

occurred transiently, <strong>and</strong> the<br />

highest activity took place during<br />

the early stages <strong>of</strong> fungal culture<br />

for A/5 isolate. The butyrateesterase<br />

activity was observed for<br />

the three isolates with a maximum<br />

Table 1. Maximum enzyme activity expressed by three<br />

isolates <strong>of</strong> Phaeosphaeria nodorum (A/5, 6/T, <strong>and</strong> 300.2) on<br />

medium supplemented with wheat cell walls.<br />

Isolate<br />

Enzyme A/5 6/T 300.2<br />

Xylanase 1574 1 (10) 2 213 (6) 200 (14)<br />

α-arabinosidase 0 5 (8) 7 (14)<br />

β-xylosidase 6 (10) 16 (14) 8 (14)<br />

Polygalacturonase 56 (6) 27(2) 12 (12)<br />

β-galactosidase 17 (12) 21 (10) 20 (14)<br />

Cellulase 106 (6) 0 3.5 (10)<br />

β-1,3 – glucanase 145 (12) 165 (10) 12 (12)<br />

β-glucosidase 153 (14) 158 (10) 118 (14)<br />

Acetyl esterase 126 (10) 44 (6) 124 (14)<br />

Butyrate esterase 527 (10) 409 (6) 379 (14)<br />

1 Data are expressed as mU.ml –1 culture medium filtrate.<br />

2 Enzyme activity was measured on the day given in<br />

parentheses.


48<br />

Session 2 — P. Halama, F. Lalaoui, V. Dumortier, <strong>and</strong> B. Paul<br />

at day 10, 6 <strong>and</strong> 14 for A/5, 6/T<br />

<strong>and</strong> 300.2 isolate respectively.<br />

Acetyl-esterase <strong>and</strong> β-galactosidase<br />

activities were also detectable for<br />

all three isolates.<br />

Aggressiveness <strong>of</strong> P.<br />

nodorum isolates<br />

The pathogenicity test allowed<br />

us to distinguish two groups <strong>of</strong><br />

isolates according to their<br />

aggressiveness (Table 2). Significant<br />

differences in the average <strong>of</strong><br />

necrosis symptoms were conclusive<br />

for differentiating the A/5 isolate<br />

from isolates 6/T <strong>and</strong> 300.2. The<br />

A/5 isolate <strong>of</strong> P. nodorum produced<br />

longer lesions on detached leaves<br />

than did the other two isolates 6/T<br />

<strong>and</strong> 300.2. Furthermore, there was<br />

partial correlation between the<br />

incubation period <strong>and</strong> length <strong>of</strong><br />

necrotic lesions; the longer lesion <strong>of</strong><br />

A/5 was related to a short latent<br />

period, but the similar lengths <strong>of</strong><br />

necrotic lesions caused by 6/T <strong>and</strong><br />

300.2 were unrelated to their<br />

incubation periods.<br />

Discussion<br />

Results reported here constitute<br />

the first report on the relationship<br />

between aggressiveness <strong>and</strong> in vitro<br />

production <strong>of</strong> CWDEs by P.<br />

nodorum. Our study indicated that<br />

Table 2. Lesion length <strong>and</strong> incubation period<br />

on detached leaves <strong>of</strong> wheat (cv. ‘Soissons’)<br />

caused by isolates <strong>of</strong> Phaeosphaeria nodorum<br />

10 days after inoculation.<br />

Lesion Incubation<br />

length period 1<br />

Isolate (mm) (hours)<br />

A/5 9.78 a 2 43<br />

6/T 3.60 b 63<br />

300.2 3.53 b 41<br />

1 The incubation period was determined by<br />

measuring the period between inoculation <strong>and</strong><br />

the moment when 50% <strong>of</strong> the detached leaves<br />

presented visible lesions.<br />

2 Within a column, means followed by different<br />

letters are significantly different according to t<br />

test (P


References<br />

Allighan, E.A., <strong>and</strong> Jackson, L.F. 1981.<br />

Variation in pathogenicity,<br />

virulence, <strong>and</strong> aggressiveness <strong>of</strong><br />

<strong>Septoria</strong> nodorum in Florida.<br />

Phytopathology 71:1080-1085.<br />

Biely, P., Mackenzie, C.R., Schneider,<br />

H. 1988. Acetylxylan esterase <strong>of</strong><br />

Schizopyllum commune. Methods in<br />

Enzymology 160:700-707.<br />

Carder, J.H., Hignett, R.C., <strong>and</strong><br />

Swinburne, T.R. 1987. Relationship<br />

between the virulence <strong>of</strong> hop<br />

isolates <strong>of</strong> Verticillium albo-atrum<br />

<strong>and</strong> their in vitro secretion <strong>of</strong> cellwall<br />

degrading enzymes. Physiol.<br />

Mol. Plant Pathol. 31:441-452.<br />

Chan, Y.H., <strong>and</strong> Sackston, W.E. 1972.<br />

Production <strong>of</strong> pectolytic <strong>and</strong><br />

cellulotytic enzymes by virulent<br />

<strong>and</strong> avirulent isolates <strong>of</strong> Sclerotium<br />

bataticola during disease<br />

development in sunflowers. Can. J.<br />

Bot. 50:2449-2453.<br />

Cooper, R.M. 1977. Regulation <strong>of</strong><br />

synthesis <strong>of</strong> cell-degrading<br />

enzymes <strong>of</strong> plant pathogens. In:<br />

Cell Wall Biochemistry Related to<br />

Specificity in Host-Plant Pathogen<br />

Interactions. Solheim, B. <strong>and</strong> Raa,<br />

J. (eds.), pp. 163-211.<br />

Universitetsforlaget, Oslo.<br />

Cooper, R.M., Longman D.,<br />

Campdell, A., Henry, M., <strong>and</strong> Lees,<br />

P.E. 1988. Enzymic adaptation <strong>of</strong><br />

cereal pathogens to the<br />

monocotyledonous primary wall.<br />

Physiol. Mol. Plant Pathol. 32:37-<br />

47.<br />

Aggressiveness <strong>of</strong> Phaeosphaeria nodorum Isolates <strong>and</strong> Their In Vitro Secretion <strong>of</strong> Cell-Wall-Degrading Enzymes 49<br />

Griffith, E., <strong>and</strong> Ao, H.C. 1980.<br />

Variation in <strong>Septoria</strong> nodorum. Ann.<br />

Appl. Biol. 94:294-296.<br />

Halama, P., <strong>and</strong> Lacoste, L. 1992. Etude<br />

des conditions optimales<br />

permettant la pycniogénèse de<br />

Phaeosphaeria (Leptosphaeria)<br />

nodorum, agent de la septoriose du<br />

blé. Agronomie 12:705-710.<br />

Halama, P., Skajennik<strong>of</strong>f, M., <strong>and</strong><br />

Dehorter, B. 1999. Tredad analysis<br />

<strong>of</strong> mating type, mutations <strong>and</strong><br />

aggressiveness in Phaeosphaeria<br />

nodorum. Mycol. Res. 1:43-49.<br />

Howel, H.E. 1975. Correlation <strong>of</strong><br />

virulence with secretion in vitro <strong>of</strong><br />

three wall-degrading enzymes in<br />

isolates <strong>of</strong> Sclerotium fructigena<br />

obtained after mutagen treatment. J.<br />

Gen. Microb. 90:32-40.<br />

Keon, J.P., Byrde, R.J.W., <strong>and</strong> Cooper,<br />

R.M. 1987. Some aspects <strong>of</strong> fungal<br />

enzymes that degrade plant cell<br />

walls. In: Fungal Infection <strong>of</strong> Plants.<br />

Pegg, G.F.<strong>and</strong> Ayres, P.G., (eds.),<br />

pp.133-157. Cambridge University<br />

Press.<br />

Kolattukudy, P.E., Purdy, R.E., <strong>and</strong><br />

Maiti, I.B. 1981. Cutinases from<br />

fungi <strong>and</strong> pollen. Methods in<br />

Enzymology 71:652-664.<br />

Krupinsky, J.M. 1997. Aggressiveness<br />

<strong>of</strong> Stagnospora nodonum isolates<br />

obtained from wheat in the<br />

northern great plains. Plant Dis.<br />

81:1027-1031.<br />

Lehtinen, U. 1993. Plant cell wall<br />

degrading enzymes <strong>of</strong> <strong>Septoria</strong><br />

nodorum. Physiol. Mol. Plant Pathol.<br />

43:121-134.<br />

Magro, P. 1984. Production <strong>of</strong><br />

polysaccharide-degrading<br />

enzymes by <strong>Septoria</strong> nodorum in<br />

culture <strong>and</strong> during pathogenesis.<br />

Plant Sc. Let. 37:63-68.<br />

Martinez, M.J., Alconada, T., Guillen,<br />

F., Vasquez, C., <strong>and</strong> Reyes, F. 1991.<br />

Pectic activities from Fusarium<br />

oxysporum f. sp. melonis:<br />

purification <strong>and</strong> characterization<br />

<strong>of</strong> an exopolygalacturonase. FEMS<br />

Microb. Lett. 81:145-150.<br />

Miller, G.L. 1959. Use <strong>of</strong><br />

dinitrosalicylic acid reagent for<br />

determination <strong>of</strong> reducing sugar.<br />

Analytical Chemistry 31:426-428.<br />

Poutanen, K. 1988. Characterization<br />

<strong>of</strong> xylanolytic enzymes for<br />

potential applications. Technical<br />

Research Centre <strong>of</strong> Finl<strong>and</strong>,<br />

Publication 47, pp. 1-59.<br />

Rapilly, F., Skajennik<strong>of</strong>f, M., Halama,<br />

P., <strong>and</strong> Touraud, G. 1992. La<br />

reproduction sexuée et<br />

l’aggressivité de Phaeosphaeria<br />

nodorum Hedj (<strong>Septoria</strong> nodorum<br />

Berk). Agronomie 12:639-649.<br />

Senna, L.A., <strong>and</strong> Goodwin P.H. 1996.<br />

Comparison <strong>of</strong> cell wall-degrading<br />

enzymes produced by highly <strong>and</strong><br />

weakly virulent isolates <strong>of</strong><br />

Leptosphaeria maculans in culture.<br />

Microbiol. Res. 151:401-406.<br />

Yang, R.C., <strong>and</strong> Hughes, G.R. 1986.<br />

Pathogenic variation among<br />

isolates <strong>of</strong> <strong>Septoria</strong> nodorum. Can. J.<br />

Plant Pathol. 8:356.


50<br />

Growth <strong>of</strong> <strong>Stagonospora</strong> nodorum Lesions<br />

A.M. Djurle (Poster)<br />

Department <strong>of</strong> Ecology <strong>and</strong> Crop Production Science/Plant Pathology, Swedish University <strong>of</strong> Agricultural<br />

Sciences, Uppsala, Sweden<br />

Disease severity is <strong>of</strong>ten<br />

expressed as a percentage or<br />

proportion <strong>of</strong> leaf area showing<br />

symptoms. Increase in disease<br />

severity <strong>of</strong> stagonospora nodorum<br />

glume blotch is caused by either an<br />

increase in the number <strong>of</strong> lesions as<br />

a result <strong>of</strong> new infection or by<br />

expansion <strong>of</strong> existing lesions. While<br />

new infections are still in the<br />

incubation period, existing lesions<br />

have time to exp<strong>and</strong> considerably<br />

(Lannou et al., 1994). Berger et al.<br />

(1997) suggested that lesion<br />

expansion should become the sixth<br />

component in describing polycyclic<br />

epidemics, <strong>and</strong> should be added to<br />

the already existing “epidemic<br />

quintet.” Lesion expansion has, for<br />

a long time, been considered as one<br />

<strong>of</strong> the most important factors for<br />

stagonospora nodorum disease<br />

increase (Rapilly et al., 1977;<br />

Rapilly et al., 1982).<br />

Materials <strong>and</strong> Methods<br />

In a field experiment with<br />

winter wheat inoculated with<br />

<strong>Stagonospora</strong> nodorum, length <strong>and</strong><br />

width <strong>of</strong> glume blotch lesions were<br />

measured repeatedly on the three<br />

uppermost leaves <strong>of</strong> tagged plants.<br />

The area <strong>of</strong> lesions was calculated<br />

assuming that they had an<br />

ellipsoidal shape. Lesion growth<br />

rates were analyzed in relation to<br />

current weather, lesion size, when a<br />

lesion was formed (first observed),<br />

<strong>and</strong> lesion age.<br />

Results <strong>and</strong> Discussion<br />

The results show that lesions<br />

followed a pattern <strong>of</strong> exponential<br />

growth <strong>and</strong> thus the rates were not<br />

constant over time. Among<br />

weather factors, mean temperature<br />

was the only factor that had a small<br />

effect on lesion growth rate, aside<br />

from the actual size <strong>of</strong> the lesions.<br />

Lesions that were formed (first<br />

observed) before or at heading<br />

(DC


Session 3A: Host-Parasite Interactions<br />

Genetic Control <strong>of</strong> Avirulence in Mycosphaerella<br />

graminicola (Anamorph <strong>Septoria</strong> tritici)<br />

G.H.J. Kema <strong>and</strong> E.C.P. Verstappen<br />

DLO-Research Institute for Plant Protection (IPO-DLO), Wageningen, The Netherl<strong>and</strong>s<br />

Mycosphaerella graminicola is a<br />

plant pathogenic bipolar<br />

heterothallic ascomycete (Kema et<br />

al., 1996c) that causes septoria<br />

tritici leaf blotch <strong>of</strong> wheat. The<br />

disease is currently considered the<br />

major threat to European wheat<br />

crops. Our main interest is to<br />

underst<strong>and</strong> the genetic control <strong>of</strong><br />

specificity in mating <strong>and</strong><br />

pathogenicity in this pathosystem.<br />

Therefore, we previously studied<br />

the histopathology <strong>of</strong> specific<br />

interactions <strong>and</strong> the genetic<br />

variation for virulence <strong>and</strong><br />

resistance in pathogen isolates <strong>and</strong><br />

host species <strong>and</strong> cultivars,<br />

respectively. We concluded that<br />

resistance to M. graminicola in the<br />

tested wheat cultivars is most<br />

probably not based on<br />

hypersensitivity (Kema et al.,<br />

1996d).<br />

Like others we observed<br />

numerous host-pathogen<br />

interactions, which were confirmed<br />

in repeated field experiments (Eyal<br />

<strong>and</strong> Levy, 1987; Saadaoui, 1987; van<br />

Ginkel <strong>and</strong> Scharen, 1988; Ahmed<br />

et al., 1995; Kema et al., 1996a, b;<br />

Kema <strong>and</strong> van Silfhout, 1997). The<br />

next step was to underst<strong>and</strong> the<br />

genetic control <strong>of</strong> these<br />

phenomena. Therefore, we had to<br />

determine the mating system <strong>of</strong> M.<br />

graminicola <strong>and</strong> find a way to cross<br />

particular isolates <strong>of</strong> the fungus,<br />

which eventually succeeded (Kema<br />

et al., 1996c).<br />

Since then, our group continues<br />

to study genome plasticity in M.<br />

graminicola <strong>and</strong> is exploring the<br />

possibility <strong>of</strong> dissecting the genetic<br />

factors that control the interaction<br />

between host <strong>and</strong> pathogen, as well<br />

as mating between pathogen<br />

strains (Kema et al., 1999). The<br />

current status <strong>and</strong> future prospects<br />

<strong>of</strong> this work will be discussed.<br />

Materials <strong>and</strong> Methods<br />

Wheat cultivars. Cultivars Shafir,<br />

Veranopolis, <strong>and</strong> Kavkaz were<br />

used as differentials <strong>and</strong> cvs.<br />

Taichung 29 <strong>and</strong> Kavkaz/K4500<br />

were used as susceptible <strong>and</strong><br />

resistant checks, respectively, in<br />

seedling experiments. In field<br />

experiments cvs. Vivant <strong>and</strong><br />

Hereward <strong>and</strong> the breeding lines<br />

NSL92-5719, RU51A, <strong>and</strong> CH76337<br />

were used.<br />

Table 1. Responses <strong>of</strong> wheat cultivars to the parental Mycosphaerella graminicola strains<br />

IPO323 <strong>and</strong> IPO94269 in seedling <strong>and</strong> adult plant experiments conducted in growth rooms <strong>and</strong> in<br />

the field, respectively. a<br />

Seedling Adult plant<br />

Cultivars IPO323 IPO94269 IPO323 IPO94269<br />

Taichung 29 + +<br />

Kavkaz/K4500 - -<br />

Kavkaz - +<br />

Veranopolis - +<br />

Shafir - + - +<br />

Hereward - +<br />

Vivant - +<br />

NSL92-5719 - +<br />

RU51A +/- -<br />

CH76337 +/- -<br />

a Pycnidial coverage: +=susceptible, -=resistant.<br />

51<br />

Fungal isolates. Isolate IPO323<br />

[MAT1-1] <strong>and</strong> IPO94269 [MAT1-2]<br />

were crossed to generate an F1<br />

population. Individual F1 progeny<br />

isolates were 1) backcrossed to both<br />

parents for mating type<br />

identification <strong>and</strong> to generate BC1<br />

populations, <strong>and</strong> 2) intercrossed to<br />

generate F2 populations. The<br />

avirulence/virulence <strong>of</strong> these<br />

parental isolates to the<br />

aforementioned cultivars is shown<br />

in Table 1.<br />

Experiments. F1, BC1, <strong>and</strong> F2<br />

populations were tested at the<br />

seedling stage in growth rooms <strong>and</strong><br />

repeated twice. Individual progeny<br />

isolates have been tested several<br />

times on the differential cultivars in<br />

the course <strong>of</strong> additional<br />

experiments. The F1 mapping<br />

population was also evaluated on<br />

differential wheat cultivars in the<br />

field. Each plot, with r<strong>and</strong>omized


52<br />

Session 3A — G.H.J. Kema <strong>and</strong> E.C.P. Verstappen<br />

cultivars, was replicated twice <strong>and</strong><br />

was inoculated with an individual<br />

F1 isolate.<br />

Results <strong>and</strong> Discussion<br />

In the seedling experiments all<br />

progeny isolates were virulent on<br />

cv. Taichung. None <strong>of</strong> these isolates<br />

carried virulence for cv. Kavkaz/<br />

K4500, indicating that the parental<br />

isolates carry the same avirulence<br />

factor(s) for this cultivar.<br />

Avirulence for each <strong>of</strong> the<br />

differentiating cultivars is inherited<br />

as a single gene. However, these<br />

avirulences co-segregated. Thus the<br />

entire F1, BC1, <strong>and</strong> F2 progenies<br />

showed the parental types. This<br />

suggests that the avirulences for<br />

these cultivars are tightly linked.<br />

There are two reasons for this<br />

hypothesis: 1) additional data<br />

indicate that cvs. Shafir,<br />

Veranopolis, <strong>and</strong> Kavkaz carry<br />

different resistance factors (Kema et<br />

al., 1996a, b) <strong>and</strong> 2) identification <strong>of</strong><br />

recombinant F1 progeny isolates in<br />

another cross<br />

(USDA50*IPO323.69.1) with<br />

avirulence for either Kavkaz or<br />

Veranopolis (unpublished data).<br />

Still, two alternative hypotheses<br />

cannot be excluded: 1) IPO323<br />

carries an avirulence factor for a<br />

common resistance factor in the<br />

tested differentials or 2) an<br />

avirulence gene product from<br />

IPO323 is recognized by different<br />

resistance factors in these cultivars.<br />

In order to investigate whether<br />

the avirulence in IPO323 for other<br />

wheat cultivars (Table 1) would<br />

segregate independently from the<br />

identified locus, a field experiment<br />

was conducted. Again, avirulence<br />

for the individual cultivars was<br />

controlled by a single locus. The<br />

avirulences for cvs. Vivant,<br />

Hereward, NSL92-5719, <strong>and</strong> Shafir<br />

co-segregated, indicating that they<br />

map to the same position as the<br />

locus identified in the seedling<br />

experiments. However, in addition<br />

to that, the avirulences for the<br />

breeding lines RU51A <strong>and</strong> CH76337<br />

inherited independently from that<br />

locus. Hence, recombinant progeny<br />

isolates were identified that were<br />

entirely virulent or avirulent on all<br />

cultivars.<br />

Hence, we hypothesize the<br />

presence <strong>of</strong> a complex major effect<br />

locus <strong>of</strong> tightly linked avirulence<br />

genes in M. graminicola IPO323 <strong>and</strong><br />

two additional independent loci<br />

carrying minor effect avirulence<br />

factors in IPO94269. The former<br />

locus was mapped on the M.<br />

graminicola genome <strong>and</strong> is currently<br />

being isolated following map-based<br />

cloning strategies.<br />

Acknowledgments<br />

We thank Jos Koeken, Suzanne<br />

Verhaegh, <strong>and</strong> the IPO-DLO<br />

technical <strong>and</strong> experimental farm<br />

staff for contributing to the<br />

presented data. Stimulating<br />

discussions with the “Wageningen”<br />

Mycosphaerella group, the “John<br />

Innes” Mycosphaerella group, <strong>and</strong><br />

the EU-SIS consortium are kindly<br />

acknowledged. Part <strong>of</strong> this project is<br />

funded by EU-BIOTECH PL096-352.<br />

References<br />

Ahmed, H.U., Mundt, C.C., <strong>and</strong><br />

Coakley, S.M. 1995. Host-pathogen<br />

relationship <strong>of</strong> geographically<br />

diverse isolates <strong>of</strong> <strong>Septoria</strong> tritici<br />

<strong>and</strong> wheat cultivars. Plant<br />

Pathology 44:838-847.<br />

Eyal, Z., <strong>and</strong> Levy, E. 1987. Variations<br />

in pathogenicity patterns <strong>of</strong><br />

Mycosphaerella graminicola within<br />

Triticum spp. in Israel. Euphytica<br />

36:237-250.<br />

Kema, G.H.J., Annone, J.G., Sayoud,<br />

R., van Silfhout, C.H., van Ginkel,<br />

M., <strong>and</strong> De Bree, J. 1996a. Genetic<br />

variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem. I. Interactions<br />

between pathogen isolates <strong>and</strong><br />

host cultivars. Phytopathology<br />

86:200-212.<br />

Kema, G.H.J., Sayoud, R., Annone,<br />

J.G., <strong>and</strong> van Silfhout, C.H. 1996b.<br />

Genetic variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem. II. Analysis <strong>of</strong><br />

interactions between pathogen<br />

isolates <strong>and</strong> host cultivars.<br />

Phytopathology 86:213-220.<br />

Kema, G.H.J., <strong>and</strong> van Silfhout, C.H.<br />

1997. Genetic variation for<br />

virulence <strong>and</strong> resistance in the<br />

wheat-Mycosphaerella graminicola<br />

pathosystem III. Comparative<br />

seedling <strong>and</strong> adult plant<br />

experiments. Phytopathology<br />

87:266-272.<br />

Kema, G.H.J., Verstappen, E.C.P.,<br />

Todorova, M., <strong>and</strong> Waalwijk, C.<br />

1996c. Successful crosses <strong>and</strong><br />

molecular tetrad <strong>and</strong> progeny<br />

analyses demonstrate<br />

heterothallism in Mycosphaerella<br />

graminicola. Current Genetics<br />

30:251-258.<br />

Kema, G.H.J., Verstappen, E.C.P.,<br />

Waalwijk, C., Bonants, P.J.M., De<br />

Koning, J.R.A., Hagenaar de<br />

Weerdt, M., Hamza, S., Koeken,<br />

J.G.P., <strong>and</strong> van der Lee, Th.A.J.<br />

1999. Genetics <strong>of</strong> biological <strong>and</strong><br />

molecular markers in<br />

Mycosphaerella graminicola, the<br />

cause <strong>of</strong> septoria tritici leaf blotch<br />

<strong>of</strong> wheat. In Lucas, J.A., Bowyer, P.,<br />

<strong>and</strong> Anderson, H.M. (eds.).<br />

<strong>Septoria</strong> on cereals: a study <strong>of</strong><br />

pathosystems. CABI Publishing,<br />

New York, NY, USA, pp. 161-180.<br />

Kema, G.H.J., Yu, D.Z., Rijkenberg,<br />

F.H.J., Shaw, M.W., <strong>and</strong> Baayen,<br />

R.P. 1996d. Histology <strong>of</strong> the<br />

pathogenesis <strong>of</strong> Mycosphaerella<br />

graminicola in wheat.<br />

Phytopathology 86:777-786.<br />

Saadaoui, E.M. 1987. Physiologic<br />

specialization <strong>of</strong> <strong>Septoria</strong> tritici in<br />

Morocco. Plant Disease 71:153-155.<br />

van Ginkel, M., <strong>and</strong> Scharen, A.L.<br />

1988. Host-pathogen relationships<br />

<strong>of</strong> wheat <strong>and</strong> <strong>Septoria</strong> tritici.<br />

Phytopathology 78:762-766.


Cytogenetics <strong>of</strong> Resistance <strong>of</strong> Wheat to <strong>Septoria</strong> Tritici<br />

Leaf Blotch<br />

L.S. Arraiano, A.J. Worl<strong>and</strong>, <strong>and</strong> J.K.M. Brown<br />

<strong>Cereals</strong> Research Department, John Innes Centre, Norwich, UK<br />

The John Innes Centre holds the<br />

world’s most comprehensive<br />

collection <strong>of</strong> precise genetic stocks<br />

<strong>of</strong> wheat. Elements <strong>of</strong> this<br />

collection, including intervarietal<br />

substitution <strong>and</strong> alien addition<br />

lines, are ideal for detecting<br />

chromosomes carrying<br />

agronomically important genes.<br />

These lines are being used in<br />

research on the genetics <strong>of</strong><br />

resistance to septoria tritici leaf<br />

blotch caused by Mycosphaerella<br />

graminicola (anamorph <strong>Septoria</strong><br />

tritici).<br />

A synthetic hexaploid wheat<br />

(Synthetic) was developed from a<br />

cross between Triticum dicoccoides<br />

(AABB) <strong>and</strong> Aegilops squarrosa (DD)<br />

(Sears, 1976). Synthetic carries<br />

genes for resistance to <strong>Stagonospora</strong><br />

nodorum (Nicholson et al., 1993),<br />

<strong>and</strong> a complete set <strong>of</strong> substitution<br />

lines <strong>of</strong> Synthetic chromosomes<br />

into a susceptible wheat variety,<br />

Chinese Spring (CS), has been<br />

developed (Worl<strong>and</strong> et al., 1996).<br />

To study resistance to M.<br />

graminicola, a detached seedling<br />

leaf technique (Arraiano et al.,<br />

1999) was used to test both Chinese<br />

Spring <strong>and</strong> Synthetic. Individual<br />

Dutch <strong>and</strong> Portuguese M.<br />

graminicola isolates supplied by<br />

IPO-DLO (Netherl<strong>and</strong>s) were<br />

tested. Synthetic was completely<br />

resistant to all isolates, except for<br />

IPO92006, whereas Chinese Spring<br />

was susceptible to all isolates<br />

except IPO323, to which it was<br />

moderately resistant. The results<br />

for Chinese Spring are consistent<br />

with field trial data (Brown et al.,<br />

1999). Baldus <strong>and</strong> Longbow, the<br />

susceptible controls, were very<br />

susceptible to all isolates.<br />

Based on these results, CS<br />

(Synthetic) substitution lines were<br />

tested using the detached leaf<br />

technique. Leaves were inoculated<br />

with the Dutch isolates IPO323 <strong>and</strong><br />

IPO94269, to identify the Synthetic<br />

chromosomes that carry genes for<br />

resistance to M. graminicola. The<br />

line carrying chromosome 7D (i.e.,<br />

CS background with the 7D<br />

chromosome substituted by that <strong>of</strong><br />

Synthetic) showed complete<br />

resistance to both isolates. Onehundred<br />

7D single-chromosome<br />

recombinant lines are now being<br />

tested to locate Synthetic’s<br />

resistance gene more precisely in<br />

relation to microsatellite <strong>and</strong> RFLP<br />

markers.<br />

Bezostaya 1 has been identified<br />

as a source <strong>of</strong> resistance to septoria<br />

tritici leaf blotch (Danon et al.,<br />

1982), <strong>and</strong> field trials showed it to<br />

be specifically resistant to IPO323<br />

(Brown et al., 1999). Substitution<br />

lines <strong>of</strong> Bezostaya 1 into Dwarf A, a<br />

susceptible line, were tested as<br />

adult plants, <strong>and</strong> chromosome 3A<br />

was identified as carrying genes for<br />

resistance to IPO323.<br />

Acknowledgment<br />

This research was supported by<br />

PRAXIS XXI – Fundação para a<br />

Ciência e a Tecnologia, Portugal,<br />

<strong>and</strong> the Ministry <strong>of</strong> Agriculture,<br />

Fisheries <strong>and</strong> Food, UK.<br />

References<br />

53<br />

Arraiano, L.S., P.A. Brading, <strong>and</strong><br />

J.K.M. Brown. 1999. A detached<br />

seedling leaf technique to screen<br />

for resistance to Mycosphaerella<br />

graminicola in field trials. In<br />

preparation.<br />

Brown, J.K.M., G.H.J. Kema, E.C.P.<br />

Verstappen, H.R. Forrer, L.S.<br />

Arraiano, P.A. Brading, E.M.<br />

Foster, A. Hecker, <strong>and</strong> E. Jenny.<br />

1999. Resistance to septoria tritici<br />

leaf blotch caused by isolates <strong>of</strong><br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici) in wheat<br />

varieties. In preparation.<br />

Danon, T., J.M. Sacks, <strong>and</strong> Z. Eyal.<br />

1982. The relationship among<br />

plant stature, maturity class <strong>and</strong><br />

susceptibility to <strong>Septoria</strong> leaf<br />

blotch <strong>of</strong> wheat. Phytopathology<br />

72:1037-1042.<br />

Nicholson, P., H.N. Rezanoor, <strong>and</strong> A.J.<br />

Worl<strong>and</strong>. 1993. Chromosomal<br />

location <strong>of</strong> resistance to <strong>Septoria</strong><br />

nodorum in a synthetic hexaploid<br />

wheat determined by the study <strong>of</strong><br />

chromosomal substitution lines in<br />

‘Chinese Spring’ wheat. Plant<br />

Breeding 110:177-184.<br />

Sears, E.R. 1976. A synthetic<br />

hexaploid wheat with fragile<br />

rachis. Wheat Information Service<br />

41:31-32.<br />

Worl<strong>and</strong>, A.J., S. Lewis, <strong>and</strong> C.<br />

Ellerbrook. 1996. <strong>Septoria</strong><br />

resistance in wheat. In John Innes<br />

Centre <strong>and</strong> Sainsbury Laboratory<br />

Annual Report 1995/96. Norwich,<br />

UK. p. 46.


54<br />

A Possible Gene-for-Gene Relationship for <strong>Septoria</strong> Tritici<br />

Leaf Blotch Resistance in Wheat<br />

P.A. Brading, 1 G.H.J. Kema, 2 <strong>and</strong> J.K.M. Brown 1<br />

1 <strong>Cereals</strong> Research Department, John Innes Centre, Norwich, UK<br />

2 DLO-Research Institute for Plant Protection (IPO-DLO), Wageningen, The Netherl<strong>and</strong>s<br />

Abstract<br />

In an F2 population <strong>of</strong> a cross between the resistant variety Flame <strong>and</strong> the susceptible variety Longbow it was shown that<br />

Flame carried one partially recessive gene for resistance to <strong>Septoria</strong> tritici isolate IPO323, as measured in a detached leaf<br />

assay. Work is underway to determine whether this resistance gene has a gene-for-gene relationship with the single avirulence<br />

locus identified in the IPO323 isolate.<br />

<strong>Septoria</strong> tritici leaf blotch,<br />

caused by the fungus<br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici), is one <strong>of</strong><br />

the most serious foliar pathogens <strong>of</strong><br />

wheat in Europe. In recent field<br />

trials using individual M.<br />

graminicola isolates, specific variety<br />

x isolate interactions have been<br />

observed (Kema <strong>and</strong> van Silfhout,<br />

1997). A number <strong>of</strong> wheat varieties<br />

show specific resistance to the<br />

Dutch isolate IPO323, including the<br />

British varieties Flame <strong>and</strong><br />

Hereward (Brown et al., 1999). In<br />

Holl<strong>and</strong>, crosses have been made<br />

between IPO323 <strong>and</strong> another<br />

isolate, IPO94269. Avirulence <strong>of</strong><br />

IPO323 on three specifically<br />

resistant cultivars, Shafir, Kavkaz,<br />

<strong>and</strong> Veranopolis, appears to be<br />

controlled at a single locus with<br />

simple inheritance (Kema et al.,<br />

1999). To investigate whether the<br />

specific resistances <strong>of</strong> Flame <strong>and</strong><br />

Hereward to IPO323 conform to a<br />

gene-for-gene relationship <strong>and</strong> also<br />

whether Flame <strong>and</strong> Hereward<br />

share the same resistance gene,<br />

crosses between Flame, Hereward,<br />

<strong>and</strong> a susceptible variety, Longbow,<br />

have been made.<br />

Eighty F 2 seedlings from a cross<br />

between Flame <strong>and</strong> Longbow were<br />

tested in a detached leaf assay<br />

(Arraiano et al., 1999) in which<br />

primary <strong>and</strong> secondary leaves were<br />

inoculated with IPO323. Leaf<br />

sections were scored according to<br />

the percentage leaf area covered<br />

with lesions bearing pycnidia 15-28<br />

days after inoculation. Each F 2<br />

plant was represented by four leaf<br />

sections tested in separate boxes.<br />

The parental varieties, Flame <strong>and</strong><br />

Longbow, were also included in<br />

each box. Comparison <strong>of</strong> infection<br />

levels on all four replicate leaf<br />

sections allowed individual plants<br />

to be rated as resistant, susceptible<br />

or intermediate. Flame was always<br />

resistant <strong>and</strong> Longbow always<br />

susceptible. Of the 80 progeny<br />

tested 23 were scored as resistant<br />

(less than 10% infection on all<br />

leaves). The other 57 progeny were<br />

scored as either susceptible (all leaf<br />

sections heavily infected) or<br />

intermediate (variable levels <strong>of</strong><br />

infection). The 1:3 ratio <strong>of</strong><br />

resistant:susceptible/intermediate<br />

is consistent with a single, partially<br />

recessive resistance gene in Flame.<br />

F 3 families generated from the<br />

80 F 2 individuals are being tested<br />

as seedlings (GS 25) in a polytunnel<br />

to test the prediction <strong>of</strong> a single<br />

gene for resistance to IPO323 in<br />

Flame. If this prediction is correct,<br />

F 3 families from resistant or<br />

susceptible F 2 individuals are<br />

expected to be uniformly resistant<br />

or susceptible, whereas<br />

intermediate F 2 individuals should<br />

produce F 3 families segregating 1:3<br />

for IPO 323<br />

resistance:susceptibility.<br />

To investigate whether<br />

Hereward carries the same IPO323<br />

resistance gene as Flame, F 2<br />

progeny <strong>of</strong> crosses between<br />

Hereward <strong>and</strong> Longbow <strong>and</strong><br />

between Hereward <strong>and</strong> Flame will<br />

be tested as detached leaves. The<br />

Hereward x Longbow cross will<br />

determine whether Hereward’s<br />

resistance is controlled by a single<br />

gene. The Hereward x Flame cross<br />

will test for allelism. The results<br />

from both tests will be presented.<br />

As IPO323 avirulence is<br />

controlled at a single locus (Kema<br />

et al., 1999) <strong>and</strong> Flame’s resistance<br />

to this isolate may be due to a<br />

single gene, we hypothesize that a<br />

gene-for-gene relationship exists<br />

between Flame <strong>and</strong> IPO323. To<br />

confirm this hypothesis, we are<br />

testing 61 progeny isolates from the<br />

IPO323 x IPO94269 cross on Flame<br />

<strong>and</strong> Longbow as detached leaves.


Acknowledgments<br />

This work was funded by the<br />

EU Framework 4 Biotechnology<br />

program <strong>and</strong> the Ministry <strong>of</strong><br />

Agriculture Fisheries <strong>and</strong> Foods<br />

(MAFF).<br />

A Possible Gene-for-Gene Relationship for <strong>Septoria</strong> Tritici Leaf Blotch Resistance in Wheat 55<br />

References<br />

Arraiano, L.S., P.A. Brading, <strong>and</strong><br />

J.K.M. Brown. 1999. A detached<br />

seedling leaf technique to screen<br />

for resistance to Mycosphaerella<br />

graminicola (anamorph <strong>Septoria</strong><br />

tritici) in wheat varieties. In<br />

preparation.<br />

Brown, J.K.M., G.H.J. Kema, H.R.<br />

Forrer, E.C.P. Verstappen, L.S.<br />

Arraiano, P.A. Brading <strong>and</strong> E.M.<br />

Foster. 1999. Resistance <strong>of</strong> wheat<br />

varieties to septoria tritici leaf<br />

blotch caused by isolates <strong>of</strong><br />

Mycosphaerella graminicola in field<br />

trials. In preparation.<br />

Kema, G.H.J., <strong>and</strong> C.H. Van Silfhout.<br />

1997. Genetic variation for<br />

virulence <strong>and</strong> resistance in the<br />

wheat-Mycosphaerella graminicola<br />

pathosystem III. Comparative<br />

seedling <strong>and</strong> adult plant<br />

experiments. Phytopathology<br />

87:266-272.<br />

Kema, G.H.J., E.C.P. Verstappen, C.<br />

Waalwijk, P.J.M. Bonants, J.R.A. de<br />

Koning, M. Hagenaar-de Weerdt,<br />

S. Hamza, J.G.P. Koeken, <strong>and</strong> T.A.J.<br />

van der Lee. 1999. Genetics <strong>of</strong><br />

biological <strong>and</strong> molecular markers<br />

in Mycosphaerella graminicola, the<br />

cause <strong>of</strong> septoria tritici leaf blotch<br />

<strong>of</strong> wheat. In: In <strong>Septoria</strong> on cereals:<br />

a Study <strong>of</strong> Pathosystems. Lucas,<br />

J.A., P. Bowyer, <strong>and</strong> H.M.<br />

Anderson (eds.). CABI Publishing,<br />

New York, NY, USA. pp.161-180.


56<br />

Diallel Analysis <strong>of</strong> <strong>Septoria</strong> Tritici Blotch Resistance in<br />

Winter Wheat<br />

X. Zhang, 1 S.D. Haley, 2 <strong>and</strong> Y. Jin 1 *<br />

1 Plant Science Department, South Dakota State University, Brookings, SD, USA<br />

2 Department <strong>of</strong> Soil <strong>and</strong> Crop Sciences, Colorado State University, Fort Collins, CO, USA<br />

Abstract<br />

In the winter wheat area <strong>of</strong> the northern Great Plains, leaf spot complex has been problematic in the past decade. In years<br />

with high precipitation from late April to July, septoria tritici blotch (STB), caused by <strong>Septoria</strong> tritici, is most prevalent. As<br />

part <strong>of</strong> our effort to improve STB resistance, inheritance <strong>of</strong> STB resistance was investigated by an eight-parent full diallel<br />

scheme. Parents, F 1 , <strong>and</strong> reciprocal F 1 were planted on three different dates. Within each planting date, three to five seeds <strong>of</strong><br />

each experimental unit were planted in the greenhouse. Materials were arranged in a r<strong>and</strong>omized complete block design<br />

(RCBD) with three replicates. Plants at the second-leaf stage were inoculated with a bulk <strong>of</strong> six S. tritici isolates. Significant<br />

general combining ability (GCA), specific combining ability (SCA), <strong>and</strong> reciprocal effects were observed in the analysis <strong>of</strong><br />

variance. The ratio <strong>of</strong> GCA sum <strong>of</strong> squares relative to SCA sum <strong>of</strong> squares suggested that GCA was more important than<br />

SCA. Additive effects played the major role in host response to STB, while non-additive effects were also detected. General<br />

combining ability effects <strong>of</strong> individual genotypes were in close agreement with parental performance. KS94U338, a genotype<br />

with resistance derived from Triticum tauschii, had the lowest STB score <strong>and</strong> the highest general combining ability. This<br />

result indicates that this genotype, possessing resistance distinct from other known sources, should prove useful in breeding<br />

efforts to improve STB resistance in wheat.<br />

Winter injury is the most<br />

adverse factor for winter wheat<br />

production in the northern Great<br />

Plains <strong>of</strong> the USA. Adoption <strong>of</strong><br />

conservation tillage practices, with<br />

winter wheat planting into spring<br />

wheat stubble, has increased<br />

steadily over the past decade.<br />

While this practice improves winter<br />

survival, accumulation <strong>of</strong> wheat<br />

residue on the soil surface has<br />

promoted the development <strong>of</strong> leaf<br />

spot diseases in this region. Severe<br />

epidemics <strong>of</strong> leaf spot complex on<br />

winter wheat have occurred during<br />

the last decade. <strong>Septoria</strong> tritici<br />

blotch has been predominant in<br />

years with above average<br />

precipitation from late April to July.<br />

Improvement <strong>of</strong> STB resistance was<br />

intensified in our project beginning<br />

in 1995. As part <strong>of</strong> this effort, we<br />

initiated studies to investigate<br />

combining ability <strong>of</strong> STB resistance<br />

* First author prevented from attending<br />

workshop by unforeseen travel<br />

problems.<br />

in known resistant genotypes. A<br />

better underst<strong>and</strong>ing <strong>of</strong> resistance<br />

could lead to more efficient<br />

deployment <strong>of</strong> germplasm<br />

resources.<br />

Materials <strong>and</strong> Methods<br />

Eight winter wheat genotypes<br />

(Table 1) were selected based on the<br />

level <strong>of</strong> resistance (field <strong>and</strong><br />

greenhouse) <strong>and</strong> their diverse<br />

origins <strong>of</strong> resistance. A total <strong>of</strong> 64<br />

genotypes (parents, F 1 , <strong>and</strong><br />

reciprocal F 1 ) were included in the<br />

test. The study consisted <strong>of</strong> three<br />

planting dates at three-day<br />

intervals. In each planting, three to<br />

five seeds <strong>of</strong> each experimental<br />

unit were planted in plastic<br />

conetainers filled with a peat-moss<br />

<strong>and</strong> perlite mixture. Each planting<br />

Table 1. Parental means, general combining ability (GCA), specific combining ability (SCA), <strong>and</strong><br />

reciprocal effects <strong>of</strong> septoria tritici blotch scores on the second leaf in an eight-parent diallel.<br />

Resistant<br />

Moderately<br />

resistant Susceptible<br />

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

1 KS94U338 —† 0.5 0.1 0.3 -1.9** 0.9** 0.6* -0.6*<br />

2 Jagger 0.8* — -0.5 -1.0** -2.0** 1.0** 1.1** 2.0**<br />

3 KS91W005-1-4 -0.9* -0.9* — -2.0* 1.6** 0.8** 1.0** 0.7*<br />

4 KS91W0935-29-1 1.3** -0.6 0.0 — 1.3** 0.4 0.5 -0.6*<br />

5 KS87822-2-1 0.3 0.1 -0.6 -0.5 — 0.1 0.3 0.5<br />

6 SD93493 -0.1 -0.4 -0.2 0.1 0.2 — -0.8* -1.3**<br />

7 T<strong>and</strong>em -1.1* 0.3 0.2 -0.2 0.1 -0.2 — -0.9*<br />

8 SD93500 -1.2** -0.3 -0.1 -1.1** -0.2 -0.2 -0.1 —<br />

Parental means 1.4 1.4 1.9 5.8 5.8 8.0 8.3 8.7<br />

Parental GCA -2.2** -1.6** -1.1** -0.5** 0.0 1.8** 1.6** 1.8**<br />

*, ** Significantly different from zero at 0.05 <strong>and</strong> 0.01 probability levels, respectively.<br />

† SCA effects are above the diagonal line, <strong>and</strong> reciprocal effects are below.


consisted <strong>of</strong> three replicates<br />

arranged in a r<strong>and</strong>omized complete<br />

block design (RCBD).<br />

Six isolates were used<br />

throughout this study. Fresh<br />

inoculum was prepared each time<br />

before inoculation. Conidia <strong>of</strong> each<br />

isolate were produced on acidic<br />

potato dextrose agar petri plates.<br />

The plates were incubated on a<br />

laboratory bench for 12 h under<br />

cool white fluorescent lights at<br />

room temperature. When the edge<br />

<strong>of</strong> pink colonies tended to darken,<br />

conidia were harvested by flooding<br />

the plates with doubled-distilled<br />

water <strong>and</strong> gently scraping the<br />

colonies with a rubber spatula<br />

attached to a glass rod. The conidial<br />

suspensions from the six isolates<br />

were combined, filtered through<br />

two layers <strong>of</strong> cheesecloth, <strong>and</strong><br />

adjusted to approximately 5-10 x<br />

10 6 conidia ml -1 .<br />

Plants were inoculated when<br />

the second leaf was fully exp<strong>and</strong>ed.<br />

A volume <strong>of</strong> 100 ml <strong>of</strong> the conidial<br />

suspension was sprayed evenly<br />

onto 300 plants using an atomizer.<br />

Plants were maintained in a mist<br />

chamber for 96 h at 100% relative<br />

humidity <strong>and</strong> 21±1°C under a 12-h<br />

photoperiod. To avoid leaf<br />

senescence, the mist chamber was<br />

opened 30 min every 24 h while the<br />

plants were kept wet. Then the<br />

Table 2. <strong>Septoria</strong> tritici blotch reading scales used in this study.<br />

plants were incubated in a growth<br />

chamber with a 12-h photoperiod at<br />

21±1°C.<br />

Twenty-one days after<br />

inoculation, disease ratings <strong>of</strong> the<br />

second leaf were taken based on a 1<br />

to 9 scale (Table 2). The average<br />

disease scores <strong>of</strong> resistant<br />

genotypes (R) were 1.0 to 4.9; <strong>of</strong><br />

moderate resistant genotypes (MR)<br />

from 5.0 to 6.9; <strong>of</strong> moderate<br />

susceptible genotypes (MS) from<br />

7.0 to 7.9; <strong>and</strong> <strong>of</strong> susceptible<br />

genotypes from 8.0 to 9.0. The data<br />

were analyzed with the “Diallel<br />

analysis <strong>and</strong> simulation” program<br />

designed by Burow <strong>and</strong> Coors<br />

(1994) according to Griffing’s<br />

Model 1 (fixed model), Method 1<br />

(Griffing, 1956). Because these<br />

experiments were conducted in one<br />

controlled environment, the<br />

genotype by environment<br />

interaction was theoretically<br />

negligible. Thus, sets (planting<br />

dates) were treated as replicates,<br />

<strong>and</strong> the mean value <strong>of</strong> the three<br />

observations within each set<br />

represented the estimated STB score<br />

<strong>of</strong> the replicate.<br />

A t-test was used to test<br />

whether the GCA, SCA, <strong>and</strong><br />

reciprocal effects were significantly<br />

different from zero. The degrees <strong>of</strong><br />

freedom for estimated GCA effects<br />

(g i ) were (p-1), where p = number<br />

Rating Symptom description<br />

1. No visible symptoms are observed, <strong>and</strong> the leaf remains green <strong>and</strong> healthy.<br />

2. A few chlorotic lesions are present, <strong>and</strong> the infection site is a tan-colored spot.<br />

3. Extensive chlorotic lesions are present. Lesions occasionally have necrosis at the infection<br />

sites.<br />

6. Extensive chlorotic lesions merge with each other, <strong>and</strong> individual lesions are identified by<br />

initial infection sites. Necrosis <strong>and</strong> chlorosis both exist on the leaf.<br />

7. Lesions fully merge, <strong>and</strong> more than half <strong>of</strong> the leaf is desiccated by necrosis.<br />

8. The entire leaf is desiccated, <strong>and</strong> water-soaked lesions occupy the entire leaf.<br />

9. A few pycnidia are visible on the infected sites, <strong>and</strong> less than 30% <strong>of</strong> the leaf is occupied by<br />

pycnidia covered lesions that remain dry <strong>and</strong> green.<br />

10. Pycnidia occupy 50 to 70% <strong>of</strong> the leaf, <strong>and</strong> infected sites are dry yet still green.<br />

11. The entire leaf is covered by pycnidial lesions, <strong>and</strong> the leaf is dry yet still green.<br />

Diallel Analysis <strong>of</strong> <strong>Septoria</strong> Tritici Blotch Resistance in Winter Wheat 57<br />

<strong>of</strong> parents. The degrees <strong>of</strong> freedom<br />

for estimated SCA effects (s ij ) <strong>and</strong><br />

estimated reciprocal effects (r ij )<br />

were p(p-1) (Kang, 1994).<br />

Results <strong>and</strong> Discussion<br />

No significant differences<br />

between replicates (different<br />

planting dates) were detected (Table<br />

3). However, highly significant<br />

effects (P


58<br />

Session 3A — X. Zhang, S.D. Haley, <strong>and</strong> Y. Jin<br />

with resistance derived from T.<br />

tauschii, should prove useful in<br />

breeding efforts to improve STB<br />

resistance in wheat. The other two<br />

resistant genotypes, ‘Jagger’ <strong>and</strong><br />

‘KS91W005-1-4’, had large,<br />

negative, highly significant GCA<br />

effects. Of the moderate resistant<br />

parents, ‘KS91W0935-29-1’ <strong>and</strong><br />

‘KS87822-2-1’, only KS91W0935-29-<br />

1 had significant negative GCA<br />

effects. Each <strong>of</strong> the susceptible<br />

parents (‘SD93493’, ‘SD93500’, <strong>and</strong><br />

‘T<strong>and</strong>em’) had high GCA values.<br />

Specific combining ability<br />

effects were observed in 20 <strong>of</strong> the 28<br />

possible combinations, indicating<br />

that non-additive effects may exist<br />

for STB resistance. Reciprocal<br />

effects were present in five <strong>of</strong> the<br />

eight combinations involving<br />

KS94U338, crosses <strong>of</strong> KS91W005-1-<br />

4 with Jagger, <strong>and</strong> SD93500 with<br />

KS91W0935-29-1. When crossed<br />

with KS94U338 as female, the<br />

hybrids <strong>of</strong> KS91W005-1-4 <strong>and</strong> the<br />

two susceptible parents (SD93500<br />

<strong>and</strong> T<strong>and</strong>em) had lower disease<br />

score than those <strong>of</strong> the reciprocal<br />

crosses (with reciprocal effects <strong>of</strong> -<br />

0.9, -1.1, <strong>and</strong> -1.2). Apparently<br />

KS91W005-1-4, T<strong>and</strong>em, <strong>and</strong><br />

SD93500 could contribute<br />

additional resistance when used as<br />

female. The maternal effects <strong>of</strong><br />

SD93500 were also observed when<br />

it was crossed with KS91W0935-29-<br />

1. The maternal effects in<br />

combinations <strong>of</strong> KS94U338 with<br />

Jagger <strong>and</strong> KS94U338 with<br />

KS91W0935-29-1 are ascribed to<br />

KS94U338, due to the positive<br />

reciprocal effects between crosses<br />

<strong>of</strong> KS94U338 used as male <strong>and</strong><br />

female. The existence <strong>of</strong> reciprocal<br />

effects in resistance to STB was also<br />

observed by Jlibene et al. (1994).<br />

With the observation <strong>of</strong><br />

predominant GCA effects <strong>and</strong><br />

reciprocal effects for enhanced<br />

resistance, improvement <strong>of</strong> STB<br />

resistance can be achieved by<br />

crossing two parents having good<br />

resistance, while selecting resistant<br />

progeny from particular crosses<br />

based on the direction <strong>of</strong> the<br />

crosses is also predictable.<br />

Acknowledgments<br />

The authors gratefully<br />

acknowledge Drs. Rollin Sears, Joe<br />

Martin, <strong>and</strong> Stan Cox (Kansas State<br />

University, USDA-ARS) for<br />

providing some <strong>of</strong> the germplasm<br />

used in our studies.<br />

References<br />

Burow, M.D., <strong>and</strong> J.G. Coors. 1994.<br />

DIALLEL: A microcomputer<br />

program for the simulation <strong>and</strong><br />

analysis <strong>of</strong> diallel crosses. Agron. J.<br />

86:154-159.<br />

Griffing, B. 1956. Concept <strong>of</strong> general<br />

<strong>and</strong> specific combining ability in<br />

relation to diallel crossing systems.<br />

Aust. J. Biol. Sci. 9:463-493.<br />

Jlibene, M., J.P. Gustafson, <strong>and</strong> S.<br />

Rajaram. 1994. Inheritance <strong>of</strong><br />

resistance to Mycosphaerella<br />

graminicola in hexaploid wheat.<br />

Plant Breed. 112:301-310.<br />

Kang, M.S. 1994. Applied quantitative<br />

genetics. Kang M.S. Publisher,<br />

Baton Rouge, LA.


Analysis <strong>of</strong> the <strong>Septoria</strong> Monitoring Nursery<br />

L. Gilchrist, C. Velazquez, <strong>and</strong> J. Crossa<br />

<strong>CIMMYT</strong> Wheat Program, El Batan, Mexico<br />

Abstract<br />

The <strong>Septoria</strong> Monitoring Nursery (SMN) was divided into two sections: 1) the best sources <strong>of</strong> resistance identified by<br />

<strong>CIMMYT</strong> <strong>and</strong> national agricultural research breeding programs (renewed every three years) <strong>and</strong> 2) a tentative group <strong>of</strong><br />

differentials. The nursery included two susceptible durum varieties <strong>and</strong> a resistant one, as well as two susceptible bread wheat<br />

varieties <strong>and</strong> a resistant one. Readings <strong>of</strong> septoria leaf blotch were taken using the double digit modified scale, <strong>and</strong> analyses<br />

were done separately by digits. The variance component estimate <strong>and</strong> the F test for the first <strong>and</strong> the second digits were highly<br />

significant for site, as well as for the site x genotype interaction component for the 1 st <strong>and</strong> 2 nd SMN. An analysis <strong>of</strong> the group<br />

<strong>of</strong> tentative differentials across the seven years <strong>and</strong> all sites also gave significant differences between sites <strong>and</strong> between sites<br />

<strong>and</strong> genotypes. The contrast between durum <strong>and</strong> bread wheat for the 1 st SMN was not significant, which indicates that at<br />

least at those locations there were no host-specific pathogen populations. The contrast between durum <strong>and</strong> bread wheat was<br />

significant in the 2 nd SMN, indicating the existence <strong>of</strong> specific pathogen populations for each <strong>of</strong> the two crops in some sites.<br />

The most resistant lines across sites for the 1 st SMN were Trap#1/Bow <strong>and</strong> #1959, <strong>of</strong> Chinese origin, <strong>and</strong> for the 2 nd<br />

SMN, Eg-AH567.71// 4*Eg-A/3/2*CMH79.243, Cal/NH/H567.71/3/2*Ning 7840/4/CMH83.2277/5/Bow/2*Ning 7840//<br />

CMH83.2277, <strong>and</strong> Sha5/Bow. A more detailed analysis <strong>of</strong> genotype x site interaction will be carried out to identify the<br />

differential response <strong>of</strong> some lines to the pathogen populations in different sites.<br />

The <strong>Septoria</strong> Monitoring<br />

Nursery was initiated in 1992<br />

(Gilchrist, 1994). Based on virulence<br />

studies on seedlings involving<br />

<strong>Septoria</strong> tritici isolates obtained from<br />

hot spot locations in many countries<br />

around the world, Kema et al.<br />

(1996; 1997) found significant<br />

interaction between pathogen<br />

isolates <strong>and</strong> host cultivars. The<br />

same authors also confirmed the<br />

specificity <strong>of</strong> the relationship<br />

between bread wheat genotypes<br />

<strong>and</strong> isolates taken from bread<br />

wheat, <strong>and</strong> between durum wheat<br />

genotypes <strong>and</strong> isolates collected<br />

from durum wheat, as<br />

demonstrated by Eyal et al. (1973),<br />

Saadaoui (1987), <strong>and</strong> van Ginkel<br />

<strong>and</strong> Scharen (1988).<br />

A better underst<strong>and</strong>ing <strong>of</strong><br />

specificity in the <strong>Septoria</strong>-wheat<br />

pathosystem may facilitate<br />

quantifying its magnitude <strong>and</strong><br />

underst<strong>and</strong>ing its role in providing<br />

protection at the crop level rather<br />

than under controlled conditions<br />

(Eyal, 1999). According to Eyal<br />

(1999), “the International<br />

Monitoring Nursery philosophy<br />

executed by <strong>CIMMYT</strong> can provide<br />

information on pathogen x cultivar<br />

interaction in the regional, national<br />

<strong>and</strong> international domain.<br />

Futhermore, it may serve as a<br />

common basis for broad-base<br />

evaluation <strong>of</strong> germplasm to diverse<br />

pathogen populations under<br />

variable environmental<br />

conditions.”<br />

The information generated by<br />

this nursery could be extremely<br />

useful to <strong>CIMMYT</strong>’s wheat<br />

breeding program <strong>and</strong> those <strong>of</strong><br />

national agricultural research<br />

systems (NARS) in their attempts to<br />

increase stable resistance over time.<br />

Material <strong>and</strong> Methods<br />

The <strong>Septoria</strong> Monitoring<br />

Nursery was divided into two<br />

sections: 1) one (renewed every<br />

three years) with the best sources <strong>of</strong><br />

59<br />

resistance identified by <strong>CIMMYT</strong><br />

<strong>and</strong> NARS breeding programs <strong>and</strong><br />

2) a tentative group <strong>of</strong> differentials<br />

proposed by Dr. Z. Eyal, Tel Aviv<br />

University, as a fixed group. They<br />

were Ald/Pvn, Enkoy (K=4500),<br />

Colotana, IAS 20, Bow CM33203-K-<br />

9M-2Y-1M-1Y-2M-0Y-1Ptz, Don<br />

Ernesto (Bow) CM33203-K-9M-33Y-<br />

1M-500Y-0M-OJ-1J, Seri 82, Kvz-<br />

K500.L.6.A.4, Beth Lehem, Lakhish,<br />

Kauz, Penjamo 62, Etit 38 (D), Inbar<br />

(D), Glennson 81, <strong>and</strong> Barrigon<br />

Yaqui (D).<br />

The same nursery with two<br />

replications was sent to each site<br />

during three years to avoid<br />

environmental <strong>and</strong> year effects,<br />

<strong>and</strong> to obtain more reliable data.<br />

The first <strong>Septoria</strong> Monitoring<br />

Nursery was sent to 23 locations.<br />

The following group <strong>of</strong> resistant<br />

lines selected under Patzcuaro<br />

(Michoacan state) <strong>and</strong> Toluca<br />

(Mexico state) conditions were<br />

included:


60<br />

Session 3A — L. Gilchrist, C. Velazquez, <strong>and</strong> J. Crossa<br />

1. Trap #1/Bow<br />

2. Thb//Ias 20/H567.71<br />

3. PF70354/Bow<br />

4. Lfn/1158.57//Prl/3/Hahn<br />

5. Ias 58/4/Kal/Bb//Cj/3/Ald/5/Bow<br />

6. Br 14*2/Sumai 3<br />

7. Sushoe #6//Ald/Pav<br />

8. Br 14*2//Nobre*2/Tp<br />

9. M2A/Cml//Nyubay/3/CMH 72A .576/Mrg<br />

10. M2A/Cml//2* Nyubay<br />

11. CMH80A.253/Sx<br />

12. CMH80.278/3/Ssfm/H567.71//2*Ssmf<br />

13. #1959<br />

14. Sumai#3<br />

15. CS/Th. curv.//Glenn81/3/Ald/Pvn<br />

The second <strong>Septoria</strong> Monitoring<br />

Nursery was sent to 22 locations<br />

<strong>and</strong> included the following resistant<br />

lines:<br />

1. Bobwhite (CM 33203-K-10M-7Y-3M-2Y-1M-<br />

0M)<br />

2. Cno79/4/CS//Th. curv.//Glenn81/3/Ald/Pvn<br />

3. TIA.2/4/CS/Th. curv.//Glenn81/3/Ald/Pvn<br />

4. CS/Th. curv.//Glenn81/3/ Ald/Pvn/4/ CS/Le.<br />

Rac//*…<br />

5. Chirya 3<br />

6. Chirya 1<br />

7. Chirya 4<br />

8. CS/Th. curv.//Glenn81/3/Ald/Pvn/4/ Suz8<br />

9. CS/Th.curv.//Glenn81/3/Ald/Pvn/4/Nanjing<br />

8401<br />

10. Eg-A/H567.71//4*Eg-A/3/2*CMH79.243<br />

11. Cal/NH//H567.71/3/2*Ning 7840/4/…<br />

12. CMH72A.576/Mrng//CMH78.443/3/<br />

CMH79.243/…<br />

13. CMH77.308/CMH82.205<br />

14. Ald/Pvn//Ymi #6<br />

15. Sha 5/Bow<br />

The nursery included two<br />

susceptible durum varieties <strong>and</strong> a<br />

resistant one, as well as two<br />

susceptible bread wheat varieties<br />

<strong>and</strong> a resistant one. Readings <strong>of</strong><br />

septoria tritici leaf blotch were<br />

taken using the double digit<br />

modified scale (Eyal et al., 1987),<br />

<strong>and</strong> the analysis was done<br />

separately by digits. Height <strong>and</strong><br />

spike emergence measurements<br />

were used as covariants in the<br />

analysis.<br />

Results <strong>and</strong> Discussion<br />

Seven <strong>of</strong> the 23 locations were<br />

included in the analysis <strong>of</strong> first<br />

<strong>Septoria</strong> Monitoring Nursery.<br />

Practically no infection was<br />

detected at the other locations due<br />

to low moisture conditions. The<br />

sites where weather conditions<br />

produced medium to good<br />

epidemics were Argentina (Balcarce<br />

<strong>and</strong> Pergamino), Chile (Hidango<br />

<strong>and</strong> Chillan), Guatemala<br />

(Quetzaltenango), Mexico (Toluca,<br />

four years), <strong>and</strong> Uruguay (La<br />

Estanzuela). Some sites were<br />

eliminated because there was<br />

strong interference from other<br />

diseases. Toluca was the only site<br />

that had no such interference, <strong>and</strong> it<br />

allowed four diseases evaluations.<br />

The variance component<br />

estimate <strong>and</strong> the F test for the first<br />

<strong>and</strong> the second digits were highly<br />

significant for the factor site, as was<br />

the site x genotype interaction<br />

component, although the latter was<br />

much smaller than the former<br />

(Table 1).<br />

Infection averages per country<br />

<strong>and</strong> site are presented in Table 2.<br />

Results indicate that La Estanzuela<br />

(Uruguay) <strong>and</strong> Toluca (Mexico) had<br />

the heaviest epidemic compared<br />

with the other sites.<br />

Table 1. Variance <strong>and</strong> F test for digits 1<br />

<strong>and</strong> 2 <strong>of</strong> the response to infection by<br />

<strong>Septoria</strong> tritici for site <strong>and</strong> genotype x site<br />

interaction (1 st SMN).<br />

Variance<br />

component F test<br />

Digit 1<br />

Site 2.68 82.98***<br />

Site x genotype 1.12 1.95***<br />

Digit 2<br />

Site 1.34 40.53***<br />

Site x genotype 1.64 2.44***<br />

The responses <strong>of</strong> durum <strong>and</strong><br />

bread wheat lines in the first<br />

<strong>Septoria</strong> Monitoring Nursery were<br />

not significantly different, which<br />

indicates that at least at those<br />

locations there were no host-specific<br />

pathogen populations. There were<br />

no significant differences between<br />

the susceptible bread wheat checks<br />

(Seri 82 <strong>and</strong> Lakish) <strong>and</strong> the<br />

resistant check (Beth Lehem). This<br />

suggests that the resistant entries<br />

were not always resistant, <strong>and</strong> that<br />

our universal resistant check was<br />

not always resistant. In the case <strong>of</strong><br />

durum wheat, this difference was<br />

significant. The resistant variety Etit<br />

38 was always resistant, while<br />

susceptible varieties Inbar <strong>and</strong><br />

Barrigon Yaqui were always<br />

susceptible.<br />

The mean infection (first <strong>and</strong><br />

second digit) <strong>of</strong> the varieties across<br />

sites <strong>and</strong> years are shown in Table 3.<br />

The most resistant varieties across<br />

sites were Trap#1/Bow <strong>and</strong> #1959,<br />

<strong>of</strong> Chinese origin.<br />

Fourteen <strong>of</strong> a total <strong>of</strong> twentytwo<br />

sites sampled were included in<br />

the second <strong>Septoria</strong> Monitoring<br />

Nursery data analysis. The sites not<br />

included had problems with<br />

interference by rust infection or<br />

experienced low moisture<br />

conditions. The countries <strong>and</strong><br />

Table 2. <strong>Septoria</strong> tritici infection average per<br />

country <strong>and</strong> site, <strong>and</strong> disease evaluation for<br />

digits 1 <strong>and</strong> 2 (1 st SMN).<br />

Infection average<br />

Country <strong>and</strong> site Digit 1 Digit 2<br />

Toluca (Mexico) 7.25 6.63<br />

La Estanzuela (Uruguay) 6.06 6.96<br />

Chillan (Chile) 6.91 5.32<br />

SNA-Hidango (Chile) 3.08 5.67<br />

Balcarce (Argentina) 5.64 4.32<br />

Pergamino (Argentina) 6.07 4.92<br />

Quetzaltenango (Guatemala) 2.45 4.07


Table. 3. Infection averages (digits 1 <strong>and</strong> 2)<br />

across sites <strong>and</strong> years (1 st SMN).<br />

Variety Digit 1 Digit 2<br />

Trap #1/Bow 3.56 1.49<br />

Thb//Ias 20/H567.71 4.78 2.70<br />

PF70354/Bow 4.62 2.73<br />

Lfn/1158.57//Prl/3/Hahn 4.97 2.61<br />

Ias 58/4/Kal/Bb//Cj/3/Ald/5/Bow 5.81 3.25<br />

Br 14*2/Sumai 3 4.25 2.06<br />

Sushoe #6//Ald/Pvn 4.42 2.92<br />

Br 14*2//Nobre*2/Tp<br />

M2A/Cml//Nyubay/3/<br />

4.72 2.70<br />

CMH 72A.576/Mrg 5.36 3.75<br />

M2A/Cml//2* Nyubay 5.39 3.34<br />

CMH80A.253/Sx<br />

CMH80.278/3/Ssfm/<br />

6.24 4.79<br />

H567.71//2*Ssmf 5.51 3.93<br />

#1959 4.03 1.94<br />

Sumai#3<br />

CS/Th. curv. //Glenn81/3/<br />

5.25 4.20<br />

Ald/Pvn<br />

Tentative differentials<br />

5.37 3.55<br />

Ald/Pvn 5.03 3.72<br />

Enkoy (K=4500) 5.18 3.09<br />

Colotana 3.06 1.56<br />

IAS 20<br />

Bow CM33203-K-9M-2Y-1M-<br />

3.67 2.31<br />

1Y-2M-0Y-1Ptz<br />

Don Ernesto(Bow)CM33203-<br />

6.29 4.52<br />

K-9M-33Y-1M-500Y-0M-OJ-1J 6.62 5.50<br />

Seri 82 7.07 6.33<br />

Kvz-K500.L6.A.4 6.16 4.48<br />

Beth Lehem 7.17 6.77<br />

Lakhish 7.03 6.78<br />

Kauz 6.92 6.50<br />

Penjamo 62 7.28 6.75<br />

Etit 38 (D) 4.04 3.06<br />

Inbar (D) 4.37 3.15<br />

Glennson 81 6.44 5.53<br />

Barrigon Yaqui (D) 7.00 5.74<br />

locations selected were Argentina<br />

(Tres Arroyos, La Plata), Chile<br />

(Temuco), Ethiopia (Holetta),<br />

Mexico (Toluca, Patzcuaro),<br />

Portugal (Elvas), Russia<br />

(Krasnoda), Switzerl<strong>and</strong> (Zurich),<br />

<strong>and</strong> Uruguay (La Estanzuela,<br />

Tarariras). Some <strong>of</strong> them were<br />

included for more than one year.<br />

The variance component<br />

estimate <strong>and</strong> the F test for digits 1<br />

<strong>and</strong> 2 S. tritici evaluation were<br />

highly significant for the factor site,<br />

as was the genotype x site<br />

interaction component (Table 4).<br />

Again, the interaction component<br />

was much smaller than the main<br />

factor component.<br />

The S. tritici infection averages<br />

per country <strong>and</strong> site for digits 1<br />

<strong>and</strong> 2 are presented in Table 5.<br />

Mexico (Toluca <strong>and</strong> Patzcuaro) <strong>and</strong><br />

Uruguay (La Estanzuela) had the<br />

lowest infection averages. This was<br />

to be expected, given that data for<br />

these three locations were<br />

considered in selecting resistant<br />

lines.<br />

The contrast between durum<br />

<strong>and</strong> bread wheats was significant,<br />

which indicated that most sites<br />

included in the second nursery had<br />

specific pathogen populations for<br />

every crop (Table 6). As in the first<br />

nursery, there were no differences<br />

between the susceptible bread<br />

wheat checks (Seri 82 <strong>and</strong> Lakish)<br />

Table 4. Variance <strong>and</strong> F test for digits 1<br />

<strong>and</strong> 2 <strong>of</strong> the response to infection by<br />

<strong>Septoria</strong> tritici for site <strong>and</strong> genotype x site<br />

interaction (2 nd SMN).<br />

Variance<br />

component F test<br />

Digit 1<br />

Site 1.12 45.07***<br />

Site x genotype<br />

Digit 2<br />

1.20 2.64***<br />

Site 0.78 22.55***<br />

Site x genotype 1.18 2.11***<br />

Table 5. <strong>Septoria</strong> tritici infection averages per<br />

country <strong>and</strong> site for digits 1 <strong>and</strong> 2 (2 nd SMN).<br />

Infection average<br />

Country <strong>and</strong> site Digit 1 Digit 2<br />

Argentina Tres Arroyos 7.81 4.05<br />

La Plata 6.45 3.46<br />

Chile Temuco 6.43 5.11<br />

Ethiopia Holetta 7.64 5.04<br />

Mexico Toluca 4.87 3.52<br />

Patzcuaro 6.75 4.90<br />

Portugal Elvas 7.40 3.94<br />

Russia Krasnovar 7.36 5.10<br />

Switzerl<strong>and</strong> Zurich 7.45 3.23<br />

Uruguay La Estanzuela 6.45 4.47<br />

Tarariras 7.37 4.98<br />

Analysis <strong>of</strong> the <strong>Septoria</strong> Monitoring Nursery 61<br />

<strong>and</strong> the resistant check (Beth<br />

Lehem). This confirms that the<br />

resistant check was not always<br />

resistant, <strong>and</strong> that our resistant<br />

check is not universally resistant<br />

<strong>and</strong> should be changed. Results<br />

confirmed that the resistant durum<br />

variety Etit 38 was always resistant,<br />

<strong>and</strong> varieties Inbar <strong>and</strong> Barrigon<br />

Yaqui were always<br />

susceptible(Table 6).<br />

Infection averages (digits 1 <strong>and</strong><br />

2) <strong>of</strong> the varieties in the second<br />

<strong>Septoria</strong> Monitoring Nursery<br />

across sites <strong>and</strong> years are shown in<br />

Table 7. The most resistant lines<br />

across sites were: 1) Eg-AH567.71//<br />

4*Eg-A/3/2*CMH79.243, 2) Cal/<br />

NH/H567.71/3/2*Ning 7840/4/<br />

CMH83.2277/5/Bow/2*Ning<br />

7840//CMH83.2277, <strong>and</strong> 3) Sha 5/<br />

Bow.<br />

Table 6. Differences between average durum<br />

<strong>and</strong> bread wheat <strong>and</strong> resistant (R) <strong>and</strong><br />

susceptible (S) checks for each crop for<br />

the 1 st SMN, 2 nd SMN, <strong>and</strong> the combination <strong>of</strong><br />

all sites for the 1 st <strong>and</strong> the 2 nd SMN.<br />

Significance<br />

levels<br />

Source<br />

1<br />

Digit 1 Digit 2<br />

st SMN<br />

Bread <strong>and</strong> durum wheat<br />

Bread wheat (S) <strong>and</strong><br />

1.93 NS 2.64 NS<br />

bread wheat (R)<br />

Durum wheat (S) <strong>and</strong><br />

0.04 NS -0.22 NS<br />

durum wheat (R)<br />

2<br />

1.64 ** 1.38 *<br />

nd SMN<br />

Bread <strong>and</strong> durum wheat<br />

Bread wheat (S) <strong>and</strong><br />

1.77 *** 1.89 NS<br />

bread wheat (R)<br />

Durum wheat (S) <strong>and</strong><br />

-0.07 NS 0.25 NS<br />

durum wheat (R)<br />

1<br />

2.02 *** 0.99 **<br />

st <strong>and</strong> 2nd SMN combined<br />

Bread <strong>and</strong> durum wheat<br />

Bread wheat (S) <strong>and</strong><br />

1.83 *** 2.16 **<br />

bread wheat (R)<br />

Durum wheat (S) <strong>and</strong><br />

-0.08 NS 0.23 NS<br />

durum wheat (R)<br />

NS: Not significant.<br />

* Significant at the 5% level.<br />

1.91 *** 1.11 ***<br />

** Significant at the 1% level.<br />

*** Significant at the 0.1% level.


62<br />

Session 3A — L. Gilchrist, C. Velazquez, <strong>and</strong> J. Crossa<br />

An analysis <strong>of</strong> the differential<br />

group across the seven years <strong>and</strong><br />

all sites also gave significant<br />

differences between sites <strong>and</strong> a<br />

significant sites x varieties value.<br />

The values <strong>of</strong> the S. tritici infection<br />

averages for digits 1 <strong>and</strong> 2 for each<br />

variety are in Table 8.<br />

Table 7. Infection averages (digits 1 <strong>and</strong> 2)<br />

across sites <strong>and</strong> years (2 nd SMN).<br />

Variety<br />

Bow CM 33203-K-10M-7Y-<br />

Digit 1 Digit 2<br />

3M-2Y-1M-0M<br />

Cno79/4/CS/Th. curv.//<br />

6.10 3.26<br />

Glenn81/3/Ald/Pvn<br />

TIA.2/4/CS/Th. curv.//Glenn81/<br />

6.61 3.62<br />

3/Ald/Pvn<br />

CS/Th. curv//Glenn81/3/Ald/<br />

6.81 3.92<br />

Pvn/4/CS/Le.Rac//*… 6.64 3.56<br />

Chirya 3 6.87 3.75<br />

Chirya 1 6.82 4.27<br />

Chirya 4<br />

CS/Th. curv. //Glenn81/3/Ald/<br />

6.94 3.86<br />

Pvn/4/ Suz8<br />

CS/Th.curv. //Glenn81/3/Ald/<br />

6.26 3.37<br />

Pvn/4/Nanjing 8401<br />

Eg-A/H567.71//4*EG-A/3/<br />

6.74 4.08<br />

2*CMH79.243<br />

Cal/NH//H567.71/3/<br />

5.25 2.60<br />

2*Ning 7840/4/…<br />

CMH72A.576/Mrng//CMH78.443/<br />

5.47 3.38<br />

3/CMH79.243/… 7.04 5.53<br />

CMH77.308/CMH82.205 6.67 3.78<br />

Ald/Pvn//Ymi #6 6.27 3.12<br />

Sha 5/Bow<br />

Tentative differentials<br />

5.96 3.17<br />

Ald/Pvn 6.10 3.34<br />

Enkoy (K=4500) 5.37 2.52<br />

Colotana 3.74 2.16<br />

IAS 20<br />

Bow CM33203-K-9M-2Y-1M-<br />

4.51 2.34<br />

1Y-2M-0Y-1Ptz<br />

Don Ernesto (Bow) CM33203-K<br />

6.35 3.87<br />

-9M-33Y-1M-500Y-0M-OJ-1J 6.85 4.45<br />

Seri 82 7.27 5.39<br />

Kvz-K500.L.6.A.4 6.60 3.67<br />

Beth Lehem 7.61 5.62<br />

Lakhish 7.82 5.35<br />

Kauz 7.27 4.74<br />

Penjamo 62 7.78 5.60<br />

Etit 38 (D) 4.44 2.90<br />

Inbar (D) 5.89 3.35<br />

Glennson 81 6.74 4.75<br />

Barrigon Yaqui (D) 7.42 4.44<br />

A more detailed analysis <strong>of</strong><br />

variety x site interactions will be<br />

carried out in the near future to<br />

identify the differential response <strong>of</strong><br />

some lines to the pathogen<br />

populations in different sites.<br />

Acknowledgments<br />

The authors fondly remember<br />

Dr. Zahir Eyal <strong>and</strong> recognize his<br />

positive influence, his wise advice<br />

to follow this line <strong>of</strong> research, <strong>and</strong><br />

his keen visualization <strong>of</strong> its<br />

potential impact. They also thank<br />

the NARS cooperators who sent in<br />

the data that made it possible to<br />

obtain the reported results. Last,<br />

but not least, the authors wish to<br />

thank Dr. Jesse Dubin for<br />

contributing his well-considered<br />

ideas on the data analyses.<br />

Table 8. <strong>Septoria</strong> tritici infection averages for<br />

differential varieties across the seven years<br />

<strong>of</strong> evaluation covering 33 locations.<br />

Variety number Digit 1 Digit 2<br />

Ald/Pvn 5.71 3.43<br />

Enkoy (K=4500) 5.29 2.69<br />

Colotana 3.51 1.96<br />

IAS 20<br />

Bow CM33203-K-9M-2Y-1M-<br />

4.23 2.33<br />

1Y-2M-0Y-1Ptz<br />

Don Ernesto (Bow) CM33203-K-<br />

6.30 4.08<br />

9M-33Y-1M-500Y-0M-OJ-1J 6.76 4.78<br />

Seri 82 7.19 5.70<br />

Kvz-K500.L.6.A.4 6.44 3.90<br />

Beth Lehem 7.45 5.98<br />

Lakhish 7.55 5.81<br />

Kauz 7.15 5.30<br />

Penjamo 62 7.61 5.96<br />

Etit 38 (D) 4.28 2.93<br />

Inbar (D) 5.37 3.26<br />

Glennson 81 6.59 4.96<br />

Barrigon Yaqui (D) 7.01 4.83<br />

References<br />

Eyal, Z., Amiri, Z., <strong>and</strong> Wahl, I. 1973.<br />

Physiologic specialization <strong>of</strong><br />

<strong>Septoria</strong> tritici. Phytopathology<br />

63:1087-1091.<br />

Eyal, Z., Sharen, A.L., Prescott, J.M.,<br />

<strong>and</strong> van Ginkel, M. 1987. The<br />

<strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Concepts <strong>and</strong> Methods <strong>of</strong> Disease<br />

Management. Mexico, D.F.:<br />

<strong>CIMMYT</strong>. 46 pp.<br />

Eyal, Z. 1999. Breeding for resistance<br />

to <strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong>. In:<br />

<strong>Septoria</strong> on <strong>Cereals</strong>: A Study <strong>of</strong><br />

Pathosystems. Lucas, J.A., Bowyer,<br />

P. <strong>and</strong> Anderson, H.M. (eds.).<br />

CABI Publishing. Wallingford, UK.<br />

pp. 1-25.<br />

Gilchrist, L. 1994. New <strong>Septoria</strong> tritici<br />

resistance sources in <strong>CIMMYT</strong><br />

germplasm <strong>and</strong> its incorporation<br />

in the <strong>Septoria</strong> Monitoring<br />

Nursery. In: Proceedings <strong>of</strong> the 4 th<br />

International Workshop on:<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>. July 4-7, 1994.<br />

Arseniuk, E., Goral, T., <strong>and</strong><br />

Czembor, P. (eds.). Ihar Radzikow,<br />

Pol<strong>and</strong>. pp. 187-190.<br />

Kema, G.H.J., Annone, G.J., Sayoud,<br />

R., van Silfhout, C.H., van Ginkel,<br />

M.,<strong>and</strong> de Bree, J. 1996. Genetic<br />

variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem I. Interaction between<br />

pathogen isolates <strong>and</strong> host<br />

cultivars Phytopathology 86:200-<br />

212.<br />

Kema, G.H.J., <strong>and</strong> van Silfhout, C.H.<br />

1997. Genetic variation for<br />

virulence <strong>and</strong> resistance in the<br />

wheat-Mycosphaerella graminicola<br />

pathosystem III. Comparative<br />

seeedling <strong>and</strong> adult plant<br />

experiments. Phytopathology<br />

87:266-272.<br />

Saadaoui, E.M. 1987. Physiologic<br />

specialization <strong>of</strong> <strong>Septoria</strong> tritici in<br />

Morocco. Plant Disease 71:153-155.<br />

Van Ginkel, M., <strong>and</strong> A.L. Scharen.<br />

1988. Host-pathogen relationships<br />

<strong>of</strong> wheat <strong>and</strong> <strong>Septoria</strong> tritici.<br />

Phytopathology 78(6):762-766.


Session 3B: Host-Parasite Interactions<br />

Host – Parasite Interactions: <strong>Stagonospora</strong> nodorum<br />

E. Arseniuk <strong>and</strong> P.C. Czembor<br />

Plant Breeding <strong>and</strong> Acclimatization Institute, Radzików, Pol<strong>and</strong><br />

Abstract<br />

Relative virulence <strong>of</strong> 15 <strong>Stagonospora</strong> (<strong>Septoria</strong>) nodorum (SNB) monopycnidiospore isolates <strong>and</strong> their mixture was<br />

studied under field conditions on a differential set <strong>of</strong> triticale <strong>and</strong> bread wheat cultivars. The isolates originated from diseased<br />

triticale <strong>and</strong> wheat plants sampled in diverse geographical regions <strong>of</strong> Pol<strong>and</strong>. Significant effects <strong>of</strong> isolates, cultivars, <strong>and</strong><br />

cultivar by isolate interactions were detected. The interaction between S. nodorum isolates <strong>and</strong> wheat <strong>and</strong> triticale<br />

genotypes appeared to be to a great extent a statistical interaction, rather than a well defined physiological specialization.<br />

The detection <strong>of</strong> the interaction was influenced by the amount <strong>of</strong> disease inflicted by the pathogen isolates on selected host<br />

genotypes <strong>and</strong> the effect <strong>of</strong> environmental conditions on disease development, as well as by the tools <strong>and</strong> precision with<br />

which the host was examined. Nonetheless, since the isolate by cultivar interaction was significant in most cases, the term<br />

virulence is used. The mixture <strong>of</strong> isolates was classified among isolates showing intermediate virulence on the cultivars<br />

tested. In comparison to single isolates, the differential capacity <strong>of</strong> the isolate mixture was also reduced. Relationships<br />

between mean isolate virulences <strong>and</strong> cultivar variances <strong>of</strong> SNB reactions were not significant for leaves (moderate SNB<br />

level) <strong>and</strong> significant for heads (low SNB level). A single pathogenic isolate used for screening breeding materials may<br />

provide reliable information on their SNB resistance.<br />

<strong>Stagonospora</strong> nodorum (Berk.)<br />

Castellani et E.G. Germano<br />

(=<strong>Septoria</strong> nodorum (Berk. Berk. in<br />

Berk. & Broome) [teleomorph<br />

Phaeosphaeria nodorum (E. Müller)<br />

Hedjaroude (=Leptosphaeria nodorum<br />

E. Müller) is a necrotrophic<br />

pathogen <strong>of</strong> graminaceous plant<br />

species. The pathogen, together<br />

with S. avenae <strong>and</strong> <strong>Septoria</strong> tritici,<br />

belongs to a group <strong>of</strong> cereal<br />

pathogens sometimes called<br />

<strong>Septoria</strong> complex. Among these<br />

pathogens special forms have been<br />

recognized only in S. avenae: f.sp.<br />

avenae <strong>and</strong> f.sp. triticea. In the other<br />

two pathogens, i.e. S. nodorum <strong>and</strong><br />

S. tritici, host specificity is generally<br />

low, but more pronounced in S.<br />

tritici (Eyal et al., 1987; Skajennik<strong>of</strong>f<br />

<strong>and</strong> Rapilly, 1983).<br />

The presence <strong>of</strong> specificity in<br />

these pathogen populations is<br />

based on the analyses <strong>of</strong> the<br />

statistical interaction terms between<br />

cultivars <strong>and</strong> isolates. The isolate<br />

by cultivar interaction terms in S.<br />

nodorum are usually <strong>of</strong> a smaller<br />

magnitude than for S. tritici<br />

(Yechilevich-Auster et al., 1983;<br />

Scharen <strong>and</strong> Eyal, 1983; Scharen et<br />

al., 1985). Therefore, distinguishing<br />

specificity (virulence) from<br />

aggressiveness is more difficult for<br />

S. nodorum. Since specificity in the<br />

S. nodorum/cereal systems at the<br />

cultivar level is not so clear, it<br />

might be called genome specificity.<br />

Isolate by genotype interactions<br />

between S. nodorum isolates <strong>and</strong><br />

cereal species have been reported<br />

by a number <strong>of</strong> authors under<br />

controlled environment as well as<br />

63<br />

field conditions (Arseniuk et. al.,<br />

1994; Krupinsky, 1986, 1994,<br />

1997a,b,c). Information on the<br />

magnitude <strong>of</strong> such interaction is<br />

important for inoculum<br />

composition <strong>and</strong> resistance<br />

management in breeding programs.<br />

Although genetic protection is<br />

considered the main pillar <strong>of</strong> host<br />

defense, the majority <strong>of</strong> commercial<br />

wheat cultivars do not have<br />

suitable protection.<br />

The lack <strong>of</strong> resistant germplasm,<br />

high pathogen variability, <strong>and</strong> low<br />

specificity have all contributed to<br />

the slow progress in breeding for<br />

resistance <strong>and</strong> the deficient genetic<br />

protection in wheat <strong>and</strong> triticale.<br />

Therefore, the objective <strong>of</strong> this<br />

study was to compare the screening<br />

efficiency <strong>of</strong> triticale <strong>and</strong> wheat<br />

cultivars for stagonospora


64<br />

Session 3B — E. Arseniuk <strong>and</strong> P.C. Czembor<br />

nodorum blotch (SNB) resistance<br />

with single isolates <strong>and</strong> their<br />

mixture under field conditions.<br />

Concurrently, the relative virulence<br />

<strong>of</strong> S. nodorum isolates, the<br />

magnitude <strong>of</strong> cultivar by isolate<br />

interaction, <strong>and</strong> the relative<br />

resistance <strong>of</strong> the cereal genotypes<br />

used in the study were examined.<br />

Materials <strong>and</strong> Methods<br />

The relative virulence <strong>of</strong> 15<br />

single pycnidiospore isolates <strong>of</strong><br />

<strong>Stagonospora</strong> (= <strong>Septoria</strong>) nodorum<br />

<strong>and</strong> their mixture was studied in<br />

1993-1995 under field conditions on<br />

both winter <strong>and</strong> spring triticale <strong>and</strong><br />

bread wheat. For the purpose 11<br />

winter triticale <strong>and</strong> 4 winter wheat<br />

cultivars <strong>and</strong> germplasm lines<br />

(Tables 1 <strong>and</strong> 2) <strong>and</strong> 5 spring<br />

triticale <strong>and</strong> 3 spring wheat<br />

cultivars (Tables 3 <strong>and</strong> 4) were<br />

used. The genotypes had been<br />

preselected earlier based on their<br />

capacity to differentiate S. nodorum<br />

isolates in regard to their<br />

virulence/aggressiveness<br />

(pathogenicity). (The term<br />

virulence will be used from this<br />

point on in this paper, since the<br />

interaction <strong>of</strong> isolates by cultivars<br />

was significant.)<br />

Fourteen isolates used for the<br />

study originated from different<br />

geographical regions <strong>of</strong> Pol<strong>and</strong> <strong>and</strong><br />

one from Switzerl<strong>and</strong> (87TS-1-1).<br />

As for host association, <strong>of</strong> those 15<br />

isolates, 14 were derived from<br />

diseased hexaploid triticale plants<br />

<strong>and</strong> one (Wen-1-1) from a hexaploid<br />

bread wheat plant cv. ‘Weneda’<br />

(Table 5). No record was made <strong>of</strong><br />

the specific triticale genotypes from<br />

which the isolates were derived. All<br />

isolates used in the study were<br />

derived from single pycnidiospores<br />

Table 1. Analysis <strong>of</strong> variance for the amount <strong>of</strong> SNB on leaves <strong>and</strong> heads <strong>of</strong> 11 winter triticale<br />

<strong>and</strong> 4 winter wheat cultivars scored on a 0-9 digit scale (0 = resistant, 9 = susceptible) under<br />

field conditions.<br />

Leaves Heads<br />

Source <strong>of</strong> Mean F Proba- Mean F Probavariation<br />

D. f. Square value bility D. f. Square value bility<br />

Years (Y) 2 427.5 21.8 0.01 2 1957.8 2696.1 0.00<br />

Blocks (B)<br />

Error 1 (Y × B) 3 19.6 3 0.7<br />

Scoring dates (S.d.) 3 4217.7 866.7 0.00 2 650.3 4.1 0.07<br />

Y × S.d. 6 77.7 16.0 0.02 4 55.0 0.3 0.83<br />

Error 2 (Y×B×S.d.) 9 4.9 6 159.7<br />

Cultivars (C) 14 113.1 321.3 0.00 14 139.3 346.4 0.00<br />

Isolates (I) 15 157.3 446.7 0.00 15 131.4 326.8 0.00<br />

C × I 210 1.4 3.9 0.00 210 1.1 2.8 0.00<br />

Y × C 28 17.1 48.6 0.00 28 20.2 50.2 0.00<br />

Y × I 30 26.5 75.4 0.00 28 3.4 8.5 0.00<br />

S.d. × C 42 4.5 12.7 0.00 30 46.3 115.1 0.00<br />

S.d. × I 45 2.2 6.3 0.00 30 3.7 9.1 0.00<br />

Y × C × I 420 0.7 1.9 0.00 56 1.2 3.1 0.00<br />

Y × S.d. × I 90 1.4 3.9 0.00 60 1.8 4.6 0.00<br />

Y × S.d. × C 84 1.8 5.2 0.00 420 0.3 0.7 1.00<br />

S.d. × C × I 630 0.3 0.8 1.00 420 0.7 1.8 0.00<br />

Y × S.d.× C × I) 1260 0.3 0.7 1.00 840 0.3 0.7 1.00<br />

Error(Y×B×S.d.×C×I) 2868 0.4 2151 0.4<br />

Table 2. The expression <strong>of</strong> virulence by 15 <strong>Stagonospora</strong> nodorum isolates <strong>and</strong> their mixture (a composite isolate) on leaves <strong>and</strong> heads <strong>of</strong> 11<br />

winter triticale (t) <strong>and</strong> 4 winter wheat (w) cultivars under field conditions.<br />

Almo Ugo Malno Lasko Grado Dagro Bolero Presto Largo LAD- DAD- Alba Jawa Begra Liwilla<br />

No. Isolate (t) (t) (t) (t) (t) (t) (t) (t) (t) 285 (t) 187 (t) (w) (w) (w) (w) Mean<br />

1 Lb-2-1 3.9* 4.1 3.7 2.9 4.3 4.5 3.0 3.1 3.5 4.4 3.8 3.5 4.2 4.9 3.4 3.8<br />

2 Lb-3-1 4.8 4.7 4.1 3.6 4.8 5.0 4.0 3.8 4.2 5.1 4.7 3.9 4.7 5.3 3.5 4.4<br />

3 Lb-4-1 4.8 4.6 4.3 3.5 4.9 4.9 3.8 3.8 4.3 5.0 4.5 3.9 4.4 5.6 3.6 4.4<br />

4 Z-20-1 3.4 3.5 3.4 2.6 3.8 3.9 2.9 2.8 3.3 3.6 3.5 3.3 4.1 4.9 3.1 3.5<br />

5 Z-22-1 4.1 4.2 4.2 3.6 4.9 5.0 3.3 4.1 4.6 5.4 4.8 3.8 4.7 5.6 3.5 4.4<br />

6 Gw-3-1 5.9 5.8 5.5 4.9 6.2 5.9 4.6 5.3 5.3 5.9 5.6 4.5 5.4 6.1 4.3 5.4<br />

7 TS87-1-1 5.4 5.4 5.3 4.8 5.8 5.9 4.3 5.2 5.0 5.7 5.1 4.7 5.6 6.1 4.7 5.3<br />

8 Wen-1-1 5.1 5.3 5.3 4.3 6.0 5.9 4.5 5.3 5.1 5.2 5.4 4.3 5.2 5.6 4.1 5.1<br />

9 Ml-2-1 3.5 3.3 3.4 2.9 3.9 4.0 2.8 2.8 3.7 3.7 3.5 3.2 4.3 5.0 3.1 3.5<br />

10 Ch-9-1 3.2 3.5 3.3 2.8 3.7 3.7 2.6 2.8 3.2 3.3 3.3 2.8 4.0 4.8 3.3 3.3<br />

11 Wr-6-1 4.4 4.2 4.0 3.5 4.5 4.5 3.5 3.6 4.2 4.5 4.4 3.8 4.6 4.9 3.8 4.2<br />

12 Ch90-3-1 5.1 4.9 4.8 4.1 5.8 5.9 4.1 4.8 4.8 5.6 8.6 3.8 5.1 5.4 3.8 5.1<br />

13 Rd93-1-1 4.0 3.8 3.7 3.0 4.3 4.3 3.1 3.2 3.5 4.0 4.0 3.4 4.3 5.4 3.7 3.8<br />

14 T91-1-4 3.4 3.9 3.3 2.8 3.9 4.1 2.8 2.7 3.5 4.1 4.0 3.3 4.3 4.9 3.0 3.6<br />

15 Lu90-1-1 4.3 4.5 4.3 3.4 5.3 5.3 3.4 4.4 4.7 4.8 4.3 3.8 4.8 5.2 3.6 4.4<br />

16 Mixture 5.0 4.6 4.5 3.9 5.4 5.3 3.7 4.3 5.0 5.4 4.8 4.0 4.9 5.4 3.5 4.6<br />

Mean 4.4 4.4 4.2 3.5 4.8 4.9 3.5 3.9 4.2 4.7 4.6 3.7 4.7 5.3 3.6 4.3<br />

LSD 0.05 = 0.114 for virulence <strong>of</strong> isolates on leaves; LSD 0.01 = 0.110 for reaction <strong>of</strong> cultivars on leaves.<br />

* Mean score on a 0-9 scale <strong>of</strong> 4 SNB scorings per season × 6 blocks over 3 growing seasons.


<strong>and</strong>, as determined after the<br />

research was completed, all were <strong>of</strong><br />

the wheat biotype.<br />

The isolates used for inoculation<br />

were increased on bread wheat<br />

grain. The composite isolate (a<br />

volumetric mixture <strong>of</strong> isolates) was<br />

prepared by mixing equal volumes<br />

<strong>of</strong> spore suspensions <strong>of</strong> all isolates<br />

adjusted to equal concentrations <strong>of</strong><br />

spores/ml. Plants were inoculated<br />

at the late boot stage (5 × 106 spores/ml) <strong>and</strong> after heading (2 ×<br />

106 spores/ml). SNB was rated<br />

visually on a 0 (resistant) - 9<br />

(susceptible) scale at about weekly<br />

intervals. Separate notes were taken<br />

for leaves <strong>and</strong> heads four <strong>and</strong> three<br />

times, respectively, on winter types,<br />

<strong>and</strong> four <strong>and</strong> two times,<br />

respectively, on spring types. All<br />

Table 3.The expression <strong>of</strong> virulence by 15 <strong>Stagonospora</strong> nodorum isolates <strong>and</strong> their mixture (a<br />

composite isolate) on leaves <strong>and</strong> heads <strong>of</strong> 5 spring triticale <strong>and</strong> 3 spring wheat cultivars un-der<br />

field conditions.<br />

Jago Maja Gabo Grego Migo Henika Eta Sigma<br />

No. Isolate (t) (t) (t) (t) (t) (w) (w) (w) Mean<br />

1 Lb-2-1 2.8* 2.6 2.2 2.9 2.8 3.6 2.4 2.8 2.8<br />

2 Lb-3-1 2.7 2.8 2.4 3.2 2.7 3.5 3.0 2.8 2.9<br />

3 Lb-4-1 2.8 2.1 2.3 2.8 2.8 3.4 2.6 2.6 2.7<br />

4 Z-20-1 1.9 2.1 2.2 2.5 2.7 3.2 2.2 2.3 2.4<br />

5 Z-22-1 2.2 2.2 2.3 2.3 2.5 2.8 2.2 2.1 2.3<br />

6 Gw-3-1 3.3 3.3 3.6 3.4 3.7 4.2 3.2 3.0 3.4<br />

7 TS87-1-1 4.4 4.3 4.4 4.3 4.9 4.9 4.1 4.1 4.4<br />

8 Wen-1-1 3.8 3.8 3.9 4.2 4.0 4.8 3.9 3.8 4.0<br />

9 Ml-2-1 2.7 2.8 2.5 2.5 3.2 2.9 2.7 2.9 2.8<br />

10 Ch-9-1 2.8 2.9 2.7 2.5 3.2 2.7 3.1 2.9 2.8<br />

11 Wr-6-1 2.7 3.1 2.6 2.3 2.9 2.9 2.6 2.7 2.7<br />

12 Ch-90-3 2.2 2.4 2.2 2.0 2.4 2.6 2.3 2.1 2.3<br />

13 Rd91-1-3 2.1 2.5 2.1 2.0 2.3 2.2 2.0 2.2 2.2<br />

14 T91-1-4 2.9 3.6 3.4 2.9 3.9 3.8 3.3 3.6 3.4<br />

15 Lu90-1-1 3.8 4.5 4.1 4.3 4.9 4.8 4.2 4.4 4.4<br />

16 Mixture 3.6 4.1 3.7 3.6 4.3 4.6 3.7 3.9 3.9<br />

Mean 2.9 3.1 2.9 3.0 3.3 3.6 3.0 3.0 3.1<br />

LSD0.01 = 0.243 - for virulence <strong>of</strong> isolates on leaves; LSD0.01 = 0.172 - for reactions <strong>of</strong> cultivars on leaves.<br />

* Mean score on a 0-9 scale <strong>of</strong> 4 SNB scorings per season x 5 blocks over 3 growing seasons.<br />

Table 4. The expression <strong>of</strong> virulence by 15 <strong>Stagonospora</strong> nodorum isolates <strong>and</strong> their mixture (a<br />

composite isolate) on heads <strong>of</strong> 5 spring triticale <strong>and</strong> 3 spring wheat cultivars under field<br />

conditions.<br />

Jago Maja Gabo Grego Migo Henika Eta Sigma<br />

No. Isolate (t) (t) (t) (t) (t) (w) (w) (w) Mean<br />

1 Lb-2-1 0.2 0.0 0.1 0.3 0.1 1.3 2.0 2.0 0.7<br />

2 Lb-3-1 0.1 0.1 0.3 0.2 0.0 2.2 2.0 2.0 0.9<br />

3 Lb-4-1 0.1 0.0 0.3 0.0 0.0 2.6 3.1 2.0 1.0<br />

4 Z-20-1 0.0 0.5 0.1 0.4 0.0 2.6 2.8 2.0 1.0<br />

5 Z-22-1 0.3 0.2 0.1 0.3 0.1 1.7 2.9 2.0 0.9<br />

6 Gw-3-1 0.7 0.2 0.0 1.0 0.8 2.9 3.5 2.0 1.4<br />

7 TS87-1-1 0.3 0.2 0.3 0.3 0.1 2.2 3.3 3.5 1.3<br />

8 Wen-1-1 0.2 0.2 0.6 0.4 0.3 2.9 3.2 3.5 1.4<br />

9 Ml-2-1 0.3 0.0 0.1 0.1 0.1 1.8 2.2 2.5 0.9<br />

10 Ch-9-1 0.0 0.1 0.1 0.0 0.0 1.6 1.6 2.0 0.7<br />

11 Wr-6-1 0.1 0.2 0.1 0.1 0.0 1.6 1.8 1.5 0.7<br />

12 Ch-90-3 0.4 0.4 0.2 0.3 0.3 1.9 1.9 2.5 1.0<br />

13 Rd91-1-3 0.0 0.0 0.0 0.6 0.0 1.8 2.0 2.5 0.9<br />

14 T91-1-4 0.1 0.2 0.2 0.0 0.0 2.2 1.9 2.0 0.8<br />

15 Lu-90-1 0.1 0.1 0.1 0.0 0.2 2.1 1.8 2.0 0.8<br />

16 Mixture 0.0 0.3 0.0 0.3 0.2 2.0 2.4 3.0 1.0<br />

Mean 0.2 0.2 0.2 0.3 0.1 2.1 2.4 2.3 1.0<br />

LSD 0.01 = 0.268 - for virulence <strong>of</strong> isolates on heads; LSD 0.01 = 0.189 - for reactions <strong>of</strong> cultivars on heads.<br />

* Mean score on 0 - 9 scale <strong>of</strong> 2 SNB scorings per season ¥ 5 blocks over 3 growing seasons.<br />

Host – Parasite Interactions: <strong>Stagonospora</strong> nodorum 65<br />

the tabulated data (Tables 2, 6, 3, 4)<br />

are averages <strong>of</strong> the numbers <strong>of</strong><br />

replications <strong>and</strong> disease scoring<br />

dates.<br />

A split-plot experimental design<br />

was used. One experimental block<br />

comprised a number <strong>of</strong> main plots<br />

equal to the number <strong>of</strong> individual S.<br />

nodorum isolates studied. The main<br />

plots (rows <strong>of</strong> parcels) contained 1m2<br />

parcels (subplots) planted to<br />

assigned cereal cultivars <strong>and</strong><br />

germplasm lines. Individual isolates<br />

<strong>of</strong> the pathogen, considered as the<br />

main plots, were r<strong>and</strong>omly<br />

assigned to the sets <strong>of</strong> subplots. The<br />

number <strong>of</strong> subplots within the main<br />

plot was equal to the number <strong>of</strong><br />

cereal genotypes studied (15: 11<br />

triticales + 4 wheat genotypes) <strong>and</strong><br />

8 (5 triticales + 3 wheat genotypes)<br />

for both winter <strong>and</strong> spring types,<br />

respectively. Each isolate (main<br />

plot) had its own control, treated<br />

with a fungicide (Tilt 250 EC, 0.1%<br />

a.i., 500 l/ha). In fact, the disease<br />

scores were actually differences<br />

between disease levels on cereal<br />

subpolts inoculated with individual<br />

isolates <strong>and</strong> its mirror images<br />

protected chemically. Isolate main<br />

plots containing the cereal subplots<br />

(cultivars/lines) were replicated<br />

twice for both winter <strong>and</strong> spring<br />

types (in 1994 only once).<br />

Data were subjected to an<br />

analysis <strong>of</strong> variance using an<br />

MSUSTAT computer program,<br />

Version 5.25 (developed by R. Lund,<br />

Montana State University, MT<br />

59715-002, USA) for a split-plot<br />

experiment. The variance among<br />

cereal cultivars/germplasm lines<br />

for disease reaction to individual S.<br />

nodorum isolates was regressed


66<br />

Session 3B — E. Arseniuk <strong>and</strong> P.C. Czembor<br />

Table 5. Origin <strong>of</strong> <strong>Stagonospora</strong> nodorum isolates.<br />

against mean virulence <strong>of</strong> the<br />

No. Isolate name Host Geographic origin (location)<br />

isolates. Similarly, the variance<br />

1 Lb-2-1 triticale Lubinicko, woj. zielonogórskie<br />

among isolates for virulence to<br />

2<br />

3<br />

4<br />

Lb-3-1<br />

Lb-4-1<br />

Z-20-1<br />

triticale<br />

triticale<br />

triticale<br />

Lubinicko, woj. zielonogórskie<br />

Lubinicko, woj. zielonogórskie<br />

Zadabrowie. woj. przemyskie<br />

individual cultivars/lines was<br />

regressed against mean resistance<br />

5<br />

6<br />

7<br />

Z-22-1<br />

Gw-3-1<br />

TS87-1-1<br />

triticale<br />

triticale<br />

triticale<br />

Zadabrowie. woj. przemyskie<br />

Grodkowice, woj. krakowskie<br />

Z¸rich, Switzerl<strong>and</strong><br />

<strong>of</strong> the cultivars/lines expressed on<br />

the 0-9 scale.<br />

8 Wen-1-1 winter wheat Weneda Elblag, woj. torunskie<br />

9<br />

10<br />

Ml-2-1<br />

Ch-9-1<br />

triticale<br />

triticale<br />

Malyszyn, woj. gorzowskie<br />

Choryn, woj. poznanskie Results<br />

11 Wr-6-1 triticale Wrocikowo, woj. olsztynskie<br />

12 Ch90-3-1 triticale Chelm, woj. chelmskie<br />

13 Rd93-1-1 triticale Radzików, woj. warszawskie<br />

All isolates <strong>of</strong> S. nodorum used<br />

14<br />

15<br />

T91-1-4<br />

Lu90-1-1<br />

triticale<br />

triticale<br />

Torun, woj. torunskie<br />

Lucmierz, woj. lódzkie<br />

in the study were pathogenic to<br />

leaves <strong>and</strong> to heads <strong>of</strong> all triticale<br />

<strong>and</strong> bread wheat cultivars (Tables<br />

2, 6, 3, 4). On average, the disease<br />

Table 6. The expression <strong>of</strong> virulence by 15 <strong>Stagonospora</strong> nodorum isolates <strong>and</strong> their mixture (a composite<br />

isolate) on heads <strong>of</strong> 11 winter triticale (t) <strong>and</strong> 4 winter wheat (w) cultivars under field conditions.<br />

on leaves was 2-3 times<br />

more severe than on<br />

No. Isolate<br />

Almo Ugo<br />

(t) (t)<br />

Malno Lasko Grado Dagro Bolero Presto Largo<br />

(t) (t) (t) (t) (t) (t) (t)<br />

LAD-<br />

285 (t)<br />

DAD-<br />

187 (t)<br />

Alba Jawa Begra Liwilla<br />

(w) (w) (w) (w) Mean<br />

heads. Some <strong>of</strong> the<br />

isolates did not cause<br />

1 Lb-2-1 0.7* 0.8 1.5 1.4 1.3 1.4 0.7 0.8 0.9 1.6 1.1 1.8 2.9 2.7 1.8 1.4<br />

disease on heads <strong>of</strong><br />

2 Lb-3-1 1.2 1.0 1.7 1.7 1.7 2.2 1.2 1.5 1.5 2.5 1.6 2.7 3.6 3.6 1.9 2.0 spring triticale genotypes<br />

3<br />

4<br />

5<br />

Lb-4-1<br />

Z-20-1<br />

Z-22-1<br />

1.3<br />

0.5<br />

0.9<br />

1.4<br />

0.6<br />

0.8<br />

2.2<br />

0.7<br />

1.1<br />

2.1<br />

0.6<br />

1.2<br />

2.5<br />

1.9<br />

1.2<br />

2.8<br />

1.1<br />

1.8<br />

1.8<br />

0.7<br />

0.8<br />

1.8<br />

0.8<br />

1.3<br />

1.8<br />

0.9<br />

1.0<br />

3.1<br />

1.2<br />

2.1<br />

1.8<br />

0.9<br />

1.6<br />

2.3<br />

1.6<br />

2.3<br />

3.6<br />

2.3<br />

3.2<br />

3.2<br />

2.0<br />

2.9<br />

2.3<br />

1.6<br />

1.9<br />

2.3<br />

1.2<br />

1.6<br />

(Table 4). The lack <strong>of</strong><br />

disease was due to low<br />

6 Gw-3-1 2.1 2.1 3.2 2.7 2.5 3.6 2.1 2.8 2.0 3.5 2.6 4.2 5.2 4.4 3.8 3.1 rainfall <strong>and</strong> drought<br />

7<br />

8<br />

TS87-1-1<br />

Wen-1-1<br />

1.8<br />

1.6<br />

2.1<br />

2.2<br />

3.2<br />

2.7<br />

2.8<br />

2.6<br />

3.3<br />

3.3<br />

3.7<br />

3.4<br />

2.2<br />

1.8<br />

2.8<br />

2.3<br />

2.6<br />

2.2<br />

4.0<br />

3.6<br />

2.5<br />

2.6<br />

3.3<br />

3.4<br />

4.9<br />

4.5<br />

4.1<br />

3.9<br />

3.7<br />

2.7<br />

3.1<br />

2.9<br />

during the summers. In<br />

9 Ml-2-1 0.7 0.6 0.9 0.9 2.3 1.3 0.6 0.6 0.8 1.8 1.0 1.7 2.7 2.2 1.7 1.3 spite <strong>of</strong> the adverse<br />

10<br />

11<br />

12<br />

Ch-9-1<br />

Wr-6-1<br />

Ch90-3-1<br />

0.7<br />

1.4<br />

1.4<br />

0.5<br />

1.3<br />

1.3<br />

0.6<br />

1.8<br />

2.2<br />

0.7<br />

1.6<br />

2.3<br />

0.9<br />

1.5<br />

2.8<br />

0.9<br />

2.4<br />

3.2<br />

0.7<br />

1.3<br />

1.5<br />

0.7<br />

1.7<br />

1.8<br />

0.6<br />

1.4<br />

1.6<br />

1.3<br />

2.4<br />

3.1<br />

1.0<br />

1.7<br />

2.1<br />

1.2<br />

2.4<br />

2.7<br />

2.1<br />

3.5<br />

4.3<br />

1.9<br />

3.2<br />

3.3<br />

2.9<br />

2.8<br />

2.6<br />

1.1<br />

2.0<br />

2.4<br />

weather, the same<br />

isolates were able to<br />

13 Rd93-1-1 0.8 0.9 0.8 1.1 2.3 1.8 0.8 1.2 1.1 1.9 1.4 2.1 2.6 2.4 3.6 1.7 cause the disease on<br />

14<br />

15<br />

T91-1-4 0.6<br />

Lu90-1-1 1.0<br />

0.5<br />

0.9<br />

1.0<br />

1.6<br />

1.4<br />

1.5<br />

1.6<br />

1.7<br />

1.7<br />

2.1<br />

0.8<br />

1.0<br />

1.0<br />

1.4<br />

1.1<br />

1.6<br />

1.8<br />

2.6<br />

1.2<br />

1.4<br />

2.1<br />

2.4<br />

2.8<br />

3.5<br />

2.6<br />

2.7<br />

1.8<br />

2.4<br />

1.5<br />

1.9<br />

heads <strong>of</strong> more SNB<br />

16 Mixture 1.1 1.3 2.4 2.3 2.4 3.0 1.6 1.7 2.1 2.6 1.7 2.7 4.0 3.3 2.8 2.3 susceptible wheat<br />

Mean 1.1 1.1 1.7 1.7 2.1 2.3 1.2 1.5 1.4 2.4 1.6 2.4 3.5 3.0 2.5 2.0 cultivars.<br />

LSD0.01 = 0.141 - for virulence <strong>of</strong> isolates on heads; LSD0.01 = 0.136 for reaction <strong>of</strong> cultivars on heads.<br />

* Mean score on a 0-9 scale <strong>of</strong> 3 SNB scorings per season ¥ 6 blocks over 3 growing seasons.<br />

Due to a large<br />

number degrees <strong>of</strong> freedom <strong>and</strong><br />

Table 7. Analysis <strong>of</strong> variance for the amount <strong>of</strong> SNB on leaves <strong>and</strong> heads <strong>of</strong> 5 spring triticale low error values, most <strong>of</strong> the<br />

<strong>and</strong> 3 spring wheat cultivars scored on a 0-9 digit scale (0 = resistant, 9 = susceptible) under<br />

field conditions.<br />

Leaves Heads<br />

Source <strong>of</strong> Mean F Proba- Mean F Probavariation<br />

D. f. Square value bility D. f. Square value bility<br />

sources <strong>of</strong> variation included in the<br />

analysis <strong>of</strong> variance (Tables 1 <strong>and</strong> 7)<br />

proved to be significant. The main<br />

effects <strong>of</strong> cultivars <strong>and</strong> isolates<br />

Blocks (B) 4 844.03 18.13 0.01 4 13.17 1.84 0.28<br />

were highly significant. It is<br />

Scoring dates (R.d.) 3 1553.60 33.37 0.00 1 39.55 5.52 0.08 noticeable that the effect <strong>of</strong> isolates<br />

Error (B × S.d.)<br />

Cultivars (C)<br />

Isolates (I)<br />

12<br />

7<br />

15<br />

46.56<br />

193.32<br />

11.3<br />

272.50<br />

15.65<br />

0.00<br />

0.00<br />

4<br />

7<br />

15<br />

7.17<br />

238.49<br />

4.58<br />

552.86<br />

10.62<br />

0.00<br />

0.00<br />

was smaller than that <strong>of</strong> cultivars.<br />

This would indicate that the<br />

C × I<br />

S.d. × C<br />

S.d. × I<br />

105<br />

21<br />

45<br />

0.636<br />

4.95<br />

0.693<br />

0.90<br />

6.97<br />

0.98<br />

0.76<br />

0.00<br />

0.51<br />

105<br />

7<br />

15<br />

0.71<br />

2.55<br />

0.167<br />

1.65<br />

5.91<br />

0.39<br />

0.01<br />

0.00<br />

0.98<br />

reaction <strong>of</strong> cultivars to inoculation<br />

with different isolates <strong>of</strong> S. nodorum<br />

S.d. × I × C<br />

Error (B×S.d.×C×I)<br />

315<br />

2032<br />

0.239<br />

0.709<br />

0.34 0.00 105<br />

1016<br />

0.129<br />

0.431<br />

0.30 1.00 was determined mainly by the


cultivar genotype. Nonetheless, the<br />

differences in virulence among<br />

isolates were large enough to make<br />

cultivar response range from<br />

resistant to susceptible, depending<br />

on the isolate used (Tables 2, 6, 3, 4).<br />

The effect <strong>of</strong> cultivar by isolate<br />

interaction was not significant only<br />

for SNB assessment on leaves <strong>of</strong><br />

spring genotypes (Table 7). In all<br />

other cases, the latter effect, though<br />

highly significant, contributed only<br />

slightly to the total variance (Tables<br />

1 <strong>and</strong> 7).<br />

The significance <strong>of</strong> cultivar by<br />

isolate interaction indicated that<br />

there is some degree <strong>of</strong> specificity<br />

in the host-pathogen relationship.<br />

The low specificity in the<br />

pathosystem was confirmed<br />

through Spearman rank<br />

correlations. They revealed that<br />

rankings <strong>of</strong> the genotypes on their<br />

disease reaction to inoculation with<br />

individual S. nodorum isolates, <strong>and</strong><br />

conversely, the rankings <strong>of</strong> isolates<br />

on their virulence expressed on the<br />

genotypes were correlated.<br />

The analyses <strong>of</strong> variance <strong>of</strong> the<br />

disease scores revealed significant<br />

differences between disease scoring<br />

dates, primarily on leaves. This<br />

means that the disease progressed<br />

with the passing <strong>of</strong> time. For SNB<br />

on heads, scoring dates were<br />

bordering on significance for both<br />

winter <strong>and</strong> spring genotypes. This<br />

indicates that, due to adverse<br />

environmental conditions<br />

(primarily summer droughts),<br />

disease progress over time was<br />

slower on heads than on leaves<br />

(Tables 1 <strong>and</strong> 7). The highly<br />

significant interactions between<br />

disease scoring dates <strong>and</strong> other<br />

components <strong>of</strong> the analysis <strong>of</strong><br />

variance indicate that the disease<br />

response progressed at different<br />

rates (Tables 1 <strong>and</strong> 7) on individual<br />

genotypes (S.d. by C) between<br />

consecutive scoring dates, <strong>and</strong> that<br />

the isolates (S.d. by I) caused more<br />

disease during the intervals<br />

between consecutive scoring dates.<br />

This may have something to do<br />

with the susceptibility <strong>of</strong> plant<br />

tissue at different growth stages.<br />

Older tissue may be more<br />

susceptible. The significance <strong>of</strong><br />

second <strong>and</strong> third degree<br />

interactions among years, scoring<br />

dates, isolates, <strong>and</strong> cultivars<br />

indicated that disease progress was<br />

influenced by all factors involved,<br />

i.e. host <strong>and</strong> pathogen genotypes,<br />

as well as the environment.<br />

It should be pointed out that the<br />

interactions <strong>of</strong> years with isolates<br />

<strong>and</strong> cultivars (Table 1 <strong>and</strong> 7),<br />

although significant, were low <strong>and</strong><br />

contributed only slightly to total<br />

variation. Thus rankings <strong>of</strong><br />

cultivars based on their SNB<br />

reaction <strong>and</strong> isolates on their<br />

Cultivar variances<br />

4.6<br />

4.2<br />

3.8<br />

3.4<br />

3.0<br />

2.6<br />

Cv. variance = 1.306 + 0.488 (mean isolate virulence)<br />

r = 0.682, p


68<br />

Cultivar variances<br />

Cultivar variances<br />

Isolate variances<br />

5.6<br />

5.4<br />

5.2<br />

7<br />

6<br />

5<br />

4<br />

3.2<br />

2.8<br />

Session 3B — E. Arseniuk <strong>and</strong> P.C. Czembor<br />

Mixture<br />

5.0<br />

Gw-3-1 TS87-1-1<br />

4.8<br />

4.6<br />

Lb-4-1<br />

Ch90-3-1<br />

Z-22-1<br />

4.4 T91-1-4<br />

Lu90-1-1<br />

Lb-3-1<br />

4.2<br />

4.0<br />

3.8<br />

2.6<br />

Wr-6-1<br />

Ch-9-1<br />

2.8<br />

MI-2-1<br />

Lb-2-1<br />

Z-20-1<br />

3.0<br />

Rd93-1-1<br />

3.2<br />

Cv. variance = 1.203 + 1.094 (mean isolate virulence)<br />

r = 0.607, p


Isolate variances<br />

Isolate variances<br />

4.6<br />

4.0<br />

3.4<br />

2.8<br />

2.2<br />

1.6<br />

3.2 3.6 4.0 4.4 4.8 5.2 5.6<br />

Mean cultivars reactions on leaves (0-9 scale)<br />

Figure 5. Relationship between the mean reactions <strong>of</strong> 11 winter triticale (t) <strong>and</strong> 4 winter wheat<br />

(w) cultivars <strong>and</strong> variances <strong>of</strong> SNB induction on leaves <strong>of</strong> the cultivars by 15 <strong>Stagonospora</strong><br />

nodorum isolates <strong>and</strong> their mixture (3 years’ data).<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

r = 0.044, n.s.<br />

Lasko (t)<br />

Bolero (t)<br />

Presto (t)<br />

Largo (t)<br />

Bolero (t)<br />

Ugo (t)<br />

Almo (t)<br />

Presto (t)<br />

Liwilla (w)<br />

Alba (w)<br />

Lasko (t)<br />

Malno (t)<br />

DAD-187 (t)<br />

Malno (t)<br />

Almo (t)<br />

Largo (t)<br />

Grado (t)<br />

Ugo (t)<br />

DAD-187 (t)<br />

Dagro (t)<br />

LAD-285 (t)<br />

Alba (w)<br />

Liwilla (t)<br />

LAD-285 (t)<br />

Dagro (t)<br />

Jawa (w)<br />

Grado (t)<br />

Begra (w)<br />

Begra (w)<br />

Jawa (w)<br />

Isolate variance = 0.285 + 11.022<br />

(mean cultivar reaction)<br />

r = 0.624, p < 0.03<br />

1.0<br />

0.8 1.4 2.0 2.6 3.2 3.8<br />

Mean cultivars reaction on heads (0-9 scale)<br />

Figure 6. Relationship between the mean reactions <strong>of</strong> 11 winter triticale (t) <strong>and</strong> 4 winter wheat<br />

(w) cultivars <strong>and</strong> variances <strong>of</strong> SNB induction on heads <strong>of</strong> the cultivars by 15 <strong>Stagonospora</strong><br />

nodorum isolates <strong>and</strong> their mixture (3 years’ data).<br />

Cultivar variances<br />

7.0<br />

6.5<br />

6.0<br />

5.5<br />

r = 0.577, n.s.<br />

Sigma (w)<br />

5.0<br />

Henika (w)<br />

4.5<br />

4.0<br />

Gabo (t)<br />

Eta (w)<br />

3.5<br />

Jago (t) Maja (t)<br />

3.0 Grego (t)<br />

Migo (t)<br />

2.5<br />

2.0 2.4 2.8 3.2 3.6 4.0 4.4 4.8<br />

Mean cultivar reaction on leaves (0-9 scale)<br />

Figure 7. Relationship between mean reactions <strong>of</strong> 5 spring triticale (t) <strong>and</strong> 3 spring wheat (w)<br />

cultivars <strong>and</strong> variances <strong>of</strong> SNB induction on leaves <strong>of</strong> the cultivars by 15 <strong>Stagonospora</strong> nodorum<br />

isolates <strong>and</strong> their mixture (3 years’ data).<br />

Host – Parasite Interactions: <strong>Stagonospora</strong> nodorum 69<br />

showed two distinct resistance<br />

groups (Figures 7 <strong>and</strong> 8).<br />

Correlations between mean host<br />

reactions on leaves <strong>and</strong> heads <strong>and</strong><br />

the variances in SNB induction<br />

among isolates were significant<br />

only for heads (Figures 5 <strong>and</strong> 7) <strong>and</strong><br />

not for leaves (Figures 6 <strong>and</strong> 8). It<br />

was found that a highly susceptible<br />

genotype is not necessarily efficient<br />

in differentiating virulence among<br />

isolates (Figures 5, 4, 7, 8).<br />

Conclusions<br />

The results produced in the<br />

study do not seem to support that<br />

using a mixture <strong>of</strong> S. nodorum<br />

isolates is the most efficient way to<br />

differentiate cereal breeding<br />

materials on their SNB resistance.<br />

The finding for S. nodorum is in<br />

accordance with results obtained by<br />

Zelikovitch <strong>and</strong> Eyal (1991) that<br />

inoculation <strong>of</strong> wheat seedlings <strong>and</strong><br />

adult plants with a mixture <strong>of</strong> S.<br />

tritici isolates resulted in significant<br />

suppression <strong>of</strong> symptoms (necrosis<br />

<strong>and</strong> pycnidial coverage) compared<br />

to the most virulent single isolate in<br />

the mixture.<br />

1. Screening for SNB resistance<br />

with a few virulent isolates that<br />

also show good differential<br />

capacity seems most appropriate.<br />

Screening breeding materials<br />

with a single highly virulent<br />

isolate may provide reliable<br />

information on their SNB<br />

resistance.<br />

2. Sets <strong>of</strong> selected isolates <strong>and</strong><br />

cultivars differentiated each<br />

other sufficiently well, but<br />

according to cluster analyses<br />

similar results could have been<br />

obtained with a lower number <strong>of</strong><br />

both isolates <strong>and</strong> germplasm<br />

lines.


70<br />

Session 3B — E. Arseniuk <strong>and</strong> P.C. Czembor<br />

3. The reported data indicate that<br />

the interaction between S.<br />

nodorum isolates <strong>of</strong> the wheat<br />

biotype <strong>and</strong> wheat <strong>and</strong> triticale<br />

genotypes to a great extent is<br />

more a matter <strong>of</strong> statistics than a<br />

well defined physiological<br />

specialization. Whether the<br />

interaction is detected or not<br />

depends upon the amount <strong>of</strong><br />

disease inflicted by the pathogen<br />

isolates on selected host<br />

genotypes, the effect <strong>of</strong><br />

environmental conditions on<br />

disease development, <strong>and</strong> the<br />

tools <strong>and</strong> precision with which<br />

the host has been examined.<br />

Acknowledgments<br />

This study was financially<br />

supported by the ICD-RSE-FAS-<br />

USDA (Project no. PL-AES-241,<br />

Grant no. MR-USDA-93-139) <strong>and</strong><br />

by the Plant Breeding <strong>and</strong><br />

Acclimatization Institute. The<br />

technical assistance <strong>of</strong> T. Bajor, E.<br />

Matyska, <strong>and</strong> E. Pawióska is<br />

gratefully acknowledged.<br />

Isolate variances<br />

2.4<br />

2.0<br />

1.6<br />

1.2<br />

0.8<br />

Isolate variance = 0.048 + 0.503 (mean cultivar reaction)<br />

r = 0.928, p = 0.01<br />

References<br />

Arseniuk, E., A L. Scharen, P.M. Fried,<br />

<strong>and</strong> H.J. Czembor. 1994.<br />

Pathogenic interactions in X<br />

Triticosecale - <strong>Septoria</strong> spp. <strong>and</strong><br />

Triticum aestivum - <strong>Septoria</strong> spp.<br />

systems. Hod. Rol. Aklim., Nasien.<br />

(Special Edition) 38(3/4):101-108.<br />

Eyal, Z., A.L. Scharen, J.M. Prescott,<br />

<strong>and</strong> M. van Ginkel. 1987. The<br />

<strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Concepts <strong>and</strong> Methods <strong>of</strong> Disease<br />

Management. Mexico D.F.:<br />

<strong>CIMMYT</strong>.<br />

Eyal, Z. 1992. The response <strong>of</strong> fieldinoculated<br />

wheat cultivars to<br />

mixtures <strong>of</strong> <strong>Septoria</strong> tritici isolates.<br />

Euphytica 61:25-35.<br />

Krupinsky, J.M. 1986. Virulence on<br />

wheat <strong>of</strong> Leptosphaeria nodorum<br />

isolates from Bromus inermis. Can.<br />

J. Plant Path. 8:201-207.<br />

Krupinsky, J.M. 1994. Aggressiveness<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum from<br />

alternative hosts after passage<br />

through wheat. Hod Rol. Aklim.,<br />

Nasien. (Special Edition) 38(3/<br />

4):123-126.<br />

Krupinsky, J.M. 1997a.<br />

Aggressiveness <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum isolates obtained from<br />

wheat in the Northern Great<br />

Plains. Plant Dis. 81:1027-1031.<br />

Henika (t)<br />

Eta (w)<br />

Sigma (w)<br />

0.4<br />

Migo (t)<br />

Jago (t) Maja (t)<br />

Gabo (t) Grego (t)<br />

0.0<br />

-0.2 0.4 1.0 1.6 2.2 2.8 3.4<br />

Mean cultivar reaction on heads (0-9 scale)<br />

Figure 8. Relationship between mean reactions <strong>of</strong> 5 spring triticale (t) <strong>and</strong> 3 spring wheat (w)<br />

cultivars <strong>and</strong> variances <strong>of</strong> SNB induction on heads <strong>of</strong> the cultivars by 15 <strong>Stagonospora</strong> nodorum<br />

isolates <strong>and</strong> their mixture (3 years’ data).<br />

Krupinsky, J.M. 1997b. Stability <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum isolates from<br />

perennial grass hosts after passage<br />

through wheat. Plant Dis. 81:1037-<br />

1041.<br />

Krupinsky, J.M. 1997c. Aggressiveness<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum isolates<br />

from perennial grasses on wheat.<br />

Plant Dis. 81:1032-1036.<br />

Scharen, L.A., <strong>and</strong> Z. Eyal. 1983.<br />

Analysis <strong>of</strong> symptoms on spring<br />

<strong>and</strong> winter wheat cultivars<br />

inoculated with different isolates<br />

<strong>of</strong> <strong>Septoria</strong> nodorum.<br />

Phytopathology 73:143-147.<br />

Scharen, L.A., Z. Eyal, M.D. Huffman,<br />

<strong>and</strong> J.M. Prescott. 1985. The<br />

distribution <strong>and</strong> frequency <strong>of</strong><br />

virulence genes in geographically<br />

separated populations <strong>of</strong><br />

Leptosphaeria nodorum.<br />

Phytopathology 75:1463-1468.<br />

Skajennik<strong>of</strong>f, M., <strong>and</strong> F. Rapilly. 1983.<br />

Aggressivness <strong>of</strong> <strong>Septoria</strong> nodorum<br />

on wheat <strong>and</strong> triticale. Effects <strong>of</strong><br />

the host <strong>and</strong> infected organs.<br />

Agronomie 3:131-140.<br />

Yechilevich-Auster, M., E. Levi, <strong>and</strong><br />

Z. Eyal. 1983. Assessment <strong>of</strong><br />

interactions between cultivated<br />

<strong>and</strong> wild wheats <strong>and</strong> <strong>Septoria</strong><br />

tritici. Phytopathology 73:1077-<br />

1083.<br />

Zelikovitch, N., <strong>and</strong> Z. Eyal. 1991.<br />

Reduction in pycnidial coverage<br />

after inoculation <strong>of</strong> wheat with<br />

mixtures <strong>of</strong> isolates <strong>of</strong> <strong>Septoria</strong><br />

tritici. Plant Dis. 75:907-910.


Identification <strong>of</strong> a Molecular Marker Linked to <strong>Septoria</strong><br />

Nodorum Blotch Resistance in Triticum tauschii Using<br />

F2 Bulked Segregant<br />

N.E.A. Murphy, 1 R. Loughman, 2 R. Wilson, 2 E.S. Lagudah, 3<br />

R. Appels, 3 <strong>and</strong> M.G.K. Jones1 1 WA State Agriculture Biotechnology Centre, Division <strong>of</strong> Science <strong>and</strong> Engineering, Murdoch University,<br />

Murdoch, Australia<br />

2 Agriculture Western Australia, Bentley Delivery Centre, Western Australia<br />

3 Plant Industries, CSIRO, Canberra, Australia<br />

Abstract<br />

A search was conducted for a molecular marker linked to a gene for resistance to septoria nodorum blotch in the<br />

Triticum tauschii accession RL5271. DNA was extracted from leaves <strong>of</strong> F2 plants which were progeny tested to identify<br />

homozygous resistant <strong>and</strong> homozygous susceptible F2 plants. The DNA from the homozygous resistant plants was pooled<br />

together <strong>and</strong> the low copy sequences were enriched using renaturation kinetics <strong>and</strong> hydroxyapatite to remove the repetitive<br />

DNA sequences. The pooled DNA from the homozygous susceptible plants was treated in the same manner. The pooled<br />

DNA <strong>and</strong> the parental DNA were screened using RAPD primers. Two markers present in the pooled resistant DNA <strong>and</strong><br />

in the parental DNA were identified <strong>and</strong> cloned. These markers were verified using RFLPs with the cloned marker as a<br />

probe. One <strong>of</strong> the probes did not produce any polymorphism as a RFLP, even when a number <strong>of</strong> different restriction<br />

enzymes were used. The second marker was polymorphic between the two parents when the DNA was restricted with<br />

HindIII <strong>and</strong> then MseI. In the 14 homozygous F2 plants tested, the marker was completely linked to the resistance gene.<br />

This marker may be valuable in introgressing the resistance gene from T. tauschii into a commercial bread wheat cultivar.<br />

<strong>Septoria</strong> nodorum blotch is the<br />

major disease <strong>of</strong> wheat (Triticum<br />

aestivum L.) in the Western<br />

Australian wheat belt. It is caused<br />

by <strong>Stagonospora</strong> nodorum (Berk.)<br />

Castellani & E.G. Germano<br />

(teleomorph Phaeosphaeria nodorum<br />

(E. Müller) Hedjaroude). The<br />

preferred method <strong>of</strong> control is the<br />

use <strong>of</strong> resistant cultivars, but this<br />

has been limited in the past by the<br />

lack <strong>of</strong> major genes for resistance.<br />

In bread wheat inheritance <strong>of</strong><br />

resistance to septoria nodorum<br />

blotch is frequently complex<br />

(Mullaney et al., 1982; Scharen <strong>and</strong><br />

Krupinsky, 1978).<br />

The complex <strong>and</strong> additive<br />

nature <strong>of</strong> septoria nodorum<br />

resistance in bread wheats has<br />

made the selection for resistance<br />

difficult. Relatively small<br />

improvements in the overall<br />

resistance can be difficult to<br />

identify, as they can be masked by<br />

strong environmental effects.<br />

Selection <strong>of</strong> resistant plants is<br />

complicated by the association <strong>of</strong><br />

resistance <strong>and</strong> both plant height<br />

<strong>and</strong> maturity (Rosielle <strong>and</strong> Brown,<br />

1980; Scott et al., 1982). The use <strong>of</strong> a<br />

molecular marker would help<br />

overcome these problems by<br />

avoiding interpretation <strong>of</strong><br />

phenotype under environmental<br />

influences (Allen, 1994).<br />

Furthermore, if the molecular<br />

marker is co-dominant, it will make<br />

it possible to identify an individual<br />

plant as homozygous for resistance<br />

without further progeny testing.<br />

This ability to identify homozygous<br />

71<br />

plants would be particularly useful<br />

in any program where extensive<br />

backcrossing would be required,<br />

such as the introgression <strong>of</strong> a<br />

resistance gene from wild wheats.<br />

The aim <strong>of</strong> this work was to<br />

identify a molecular marker that is<br />

linked to a single gene conferring<br />

resistance to septoria nodorum<br />

blotch identified in the accession<br />

RL5271 <strong>of</strong> the wild wheat Triticum<br />

tauschii.<br />

Materials <strong>and</strong> Methods<br />

A cross between the resistant<br />

accession RL5271 <strong>and</strong> the<br />

susceptible accession CPI110889<br />

was progeny tested. The F2 plants<br />

were progeny tested using F3


72<br />

Session 3B — N.E.A. Murphy, R. Loughman, R. Wilson, E.S. Lagudah, R. Appels, <strong>and</strong> M.G.K. Jones<br />

families to identify homozygous F2<br />

plants. The DNA <strong>of</strong> eight<br />

homozygous resistant F2 plants<br />

<strong>and</strong> 14 homozygous susceptible F2<br />

plants was selected for the analysis.<br />

Equal amounts <strong>of</strong> DNA from<br />

the homozygous resistant F2 plants<br />

were bulked together as was the<br />

DNA from the homozygous<br />

susceptible F2 plants. Each bulked<br />

sample <strong>and</strong> the DNA from each<br />

parent were divided into two<br />

aliquots. One aliquot was left<br />

untreated. The second aliquot was<br />

treated to enrich the proportion <strong>of</strong><br />

low copy sequence DNA by<br />

removing the repetitive DNA<br />

sequences using hydroxyapatite<br />

after selective renaturation. The<br />

technique followed Eastwood et al.<br />

(1994). The double-str<strong>and</strong>ed DNA<br />

was removed from the samples<br />

after a Cot value <strong>of</strong> 120 had been<br />

achieved (Smith <strong>and</strong> Flavell, 1975).<br />

The DNA was re-suspended in<br />

0.1xTE buffer <strong>and</strong> diluted to a final<br />

concentration <strong>of</strong> 5ng/ml <strong>of</strong> DNA.<br />

RAPD reactions were carried<br />

out on a Perkin-Elmer PE9600. The<br />

r<strong>and</strong>om primers were produced by<br />

Operon Technologies (Alameda,<br />

Ca.). The RAPD products were<br />

electrophoresed on polyacrylamide<br />

gels, <strong>and</strong> the gels were stained<br />

using ethidium bromide with gel<br />

images recorded digitally using a<br />

Bio-Rad Gel Doc 1000 system. For<br />

each primer nine reactions were<br />

carried out: a negative control,<br />

sterile distilled water, was used<br />

instead <strong>of</strong> template DNA, the next<br />

four reactions used the total<br />

genomic DNA as template, <strong>and</strong> the<br />

final four reactions used the nonrepetitive<br />

DNA as template.<br />

If a polymorphism was<br />

identified, the polymorphic b<strong>and</strong><br />

was stabbed with a sterile syringe<br />

needle. The needle was washed in<br />

sterile distilled water which was<br />

used as template for a further<br />

RAPD reaction using the same<br />

primer <strong>and</strong> reaction conditions that<br />

had generated the polymorphism.<br />

The amplified fragment was ligated<br />

into the plasmid pGEM-T. The<br />

plasmids were transformed into E.<br />

coli cells by heat shock. The<br />

following day colonies were<br />

screened using PCR primers, <strong>and</strong><br />

permanent cultures were<br />

established from insert-containing<br />

clones.<br />

The cloned polymorphic b<strong>and</strong><br />

was used as an RFLP probe to<br />

verify <strong>and</strong> determine the linkage <strong>of</strong><br />

the potential marker to the<br />

resistance gene. Total genomic<br />

DNA <strong>of</strong> the parental accessions was<br />

digested with a number <strong>of</strong><br />

restriction enzymes, singly <strong>and</strong> in<br />

combination, to find a restriction<br />

enzyme that would produce a<br />

polymorphism with the marker<br />

between the resistant <strong>and</strong> the<br />

susceptible parent. The marker<br />

b<strong>and</strong> was prepared using an alkali-<br />

SDS minprep, double digested with<br />

NcoI <strong>and</strong> SacI, separated on an<br />

agarose gel <strong>and</strong> purified using<br />

agarase. The purified b<strong>and</strong> was<br />

labeled using r<strong>and</strong>om nonamer<br />

primers with 32P as dCTP. The<br />

membranes were hybridized<br />

overnight, washed the following<br />

morning, <strong>and</strong> exposed to<br />

autoradiograph film for up to one<br />

week.<br />

When a suitable restriction<br />

enzyme had been selected, the<br />

DNA from the selected seven<br />

homozygous resistant F2 plants<br />

<strong>and</strong> seven homozygous susceptible<br />

F2 plants was digested,<br />

electrophoresed <strong>and</strong> blotted using<br />

an alkali capillary blot. The<br />

membranes were autoradiographed<br />

for up to two weeks.<br />

Results<br />

Sixty r<strong>and</strong>om primers were<br />

used to screen the DNA samples.<br />

Two polymorphic b<strong>and</strong>s were<br />

identified by comparing the<br />

b<strong>and</strong>ing pattern <strong>of</strong> the resistant<br />

parent <strong>and</strong> the resistant bulk with<br />

the susceptible parent <strong>and</strong> the<br />

susceptible bulk. The first was<br />

amplified using primer OPJ6 <strong>and</strong><br />

was approximately 770bp long<br />

(OPJ6770 ). The second polymorphic<br />

b<strong>and</strong> was amplified using primer<br />

OPJ18 <strong>and</strong> the fragment was<br />

approximately 350bp (OPJ18350 ).<br />

Both b<strong>and</strong>s were ligated into<br />

pGEM-T.<br />

For the first marker, OPJ6770 , no<br />

restriction enzyme nor combination<br />

<strong>of</strong> enzymes tested was able to<br />

produce a polymorphism between<br />

the resistant <strong>and</strong> the susceptible<br />

parent. Consequently the marker<br />

could not be tested for its linkage to<br />

resistance. The second marker,<br />

OPJ18350 , produced a<br />

polymorphism between the<br />

resistant <strong>and</strong> susceptible parent<br />

when the DNA was double<br />

restricted with HindIII <strong>and</strong><br />

followed by MseI. The<br />

polymorphism was present as an<br />

extra b<strong>and</strong> in the resistant parent.<br />

All <strong>of</strong> the F2 plants tested were<br />

correctly classified as resistant or<br />

susceptible when the marker was<br />

used.


Discussion<br />

A marker linked to a resistance<br />

gene to septoria nodorum blotch in<br />

the Triticum tauschii accession<br />

RL5271 has been found. The marker<br />

<strong>and</strong> the resistance gene appear to<br />

be completely linked; however, a<br />

greater number <strong>of</strong> plants need to be<br />

tested to develop a more accurate<br />

estimate <strong>of</strong> the linkage. The<br />

multiple b<strong>and</strong>s produced when the<br />

marker was used as an RFLP probe<br />

indicate that regions <strong>of</strong> repetitive<br />

DNA exist within the marker,<br />

which is not uncommon (Eastwood<br />

et al., 1994).<br />

The RFLP probe is currently<br />

being sequenced <strong>and</strong> primers<br />

designed that will allow the<br />

development <strong>of</strong> a sequence<br />

characterized amplified region<br />

(SCAR). This will allow a simpler<br />

<strong>and</strong> faster test to be developed. The<br />

simpler assay will be required if it<br />

is to be used extensively in a<br />

marker-assisted selection program.<br />

If the designed primers do amplify<br />

a polymorphism between the<br />

Identification <strong>of</strong> a Molecular Marker Linked to <strong>Septoria</strong> Nodorum Blotch Resistance in Triticum tauschii 73<br />

resistant <strong>and</strong> the susceptible plants,<br />

it may be one <strong>of</strong> several types. The<br />

primers may amplify a region that<br />

is present in the resistant plants but<br />

not in the susceptible plants. This<br />

marker will be dominant <strong>and</strong> will<br />

not allow for the identification <strong>of</strong><br />

heterozygotes; however, this will<br />

enable simple alternatives to<br />

running the PCR product on a gel<br />

to determine if the b<strong>and</strong> is present<br />

or not to be used (Dedryver et al.,<br />

1996). This will remove one <strong>of</strong> the<br />

major time constraints <strong>and</strong> limiting<br />

factors in the screening <strong>of</strong> large<br />

numbers <strong>of</strong> plants. Alternatively,<br />

the SCAR primers could amplify<br />

different size fragments between<br />

the resistant <strong>and</strong> the susceptible<br />

plants. This type <strong>of</strong> polymorphism<br />

will allow heterozygotes to be<br />

identified (co-dominant), as both<br />

alleles will be amplified. This in<br />

turn will allow the selection <strong>of</strong><br />

homozygotes for use in a<br />

backcrossing population, where the<br />

direct <strong>and</strong> accurate selection <strong>of</strong><br />

homozygotes will be <strong>of</strong> greatest<br />

benefit.<br />

References<br />

Allen, F.L. 1994. Usefulness <strong>of</strong> plant<br />

genome mapping to plant<br />

breeding. In Plant Genome<br />

Analysis (G.M. Gressh<strong>of</strong>f, ed.).<br />

CRC Press, London.<br />

Dedryver, F., M.F. Jubier, J.<br />

Thouvenin, <strong>and</strong> H. Goyeau. 1996.<br />

Molecular markers linked to the<br />

leaf rust resistance gene Lr24 in<br />

different wheats. Genome 39:830-<br />

835.<br />

Eastwood, R.F., E.S. Lagudah, <strong>and</strong> R.<br />

Appels. 1994. A directed search for<br />

DNA sequences tightly linked to<br />

cereal cyst nematode (CCN)<br />

resistance in Triticum tauschii.<br />

Genome 37:311-319.<br />

Mullaney, J., M. Martin, <strong>and</strong> A.L.<br />

Scharen. 1982. Generation mean<br />

analysis to identify <strong>and</strong> partition<br />

the components <strong>of</strong> genetic<br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Euphytica 31:539-545.<br />

Rosielle, A.A., <strong>and</strong> A.G.P. Brown.<br />

1980. Selection for resistance to<br />

<strong>Septoria</strong> nodorum in wheat.<br />

Euphytica 29:337-346.<br />

Scharen, A.L., <strong>and</strong> J.M. Krupinsky.<br />

1978. Detection <strong>and</strong> manipulation<br />

<strong>of</strong> resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Phytopathology 68:245-248.<br />

Scott, P.R., P.W. Benedikz, <strong>and</strong> C.J.<br />

Cox. 1982. A genetic study <strong>of</strong> the<br />

relationship between height, time<br />

<strong>of</strong> ear emergence <strong>and</strong> resistance to<br />

<strong>Septoria</strong> nodorum in wheat. Plant<br />

Pathology 31:45-60.<br />

Smith, D.B., <strong>and</strong> R.B. Flavell. 1975.<br />

Characterisation <strong>of</strong> the wheat<br />

genome by renaturation kinetics.<br />

Chromosoma 50:223-242.


74<br />

Inheritance <strong>of</strong> <strong>Septoria</strong> Nodorum Blotch Resistance in a<br />

Triticum tauschii Accession Controlled by a Single Gene<br />

N.E.A. Murphy, 1 R. Loughman, 2 R. Wilson, 2 E.S. Lagudah, 3<br />

R. Appels, 3 <strong>and</strong> M.G.K. Jones1 1 WA State Agriculture Biotechnology Centre, Division <strong>of</strong> Science <strong>and</strong> Engineering, Murdoch University,<br />

Murdoch, Australia<br />

2 Agriculture Western Australia, Bentley Delivery Centre, Western Australia<br />

3 Plant Industries, CSIRO, Canberra, Australia<br />

Abstract<br />

A potentially useful source <strong>of</strong> resistance has been identified in an accession <strong>of</strong> Triticum tauschii. A cross was made<br />

between the resistant T. tauschii accession, RL5271, <strong>and</strong> a susceptible accession, CPI110889, to study the genetics <strong>of</strong><br />

resistance from this source. The parental accessions <strong>and</strong> the F1 <strong>and</strong> F3 progeny were screened in the glasshouse as seedlings.<br />

The resistant parent took significantly longer to develop symptoms, developed significantly fewer lesions, <strong>and</strong> expressed<br />

significantly lower levels <strong>of</strong> disease than the susceptible parent. The F1 mean response for disease severity indicated there was<br />

no complete dominance. The genotypic ratios generated by screening the F3 families were not significantly different from the<br />

genotypic ratio expected for a single gene. The effectiveness <strong>and</strong> simple genetic control <strong>of</strong> the resistance in the T. tauschii<br />

accession RL5271 may be useful as a resistance source in a bread wheat breeding program.<br />

<strong>Septoria</strong> nodorum blotch,<br />

caused by Phaeosphaeria nodorum (E.<br />

Müller) Hedjaroude (anamorph<br />

<strong>Stagonospora</strong> nodorum (Berk.)<br />

Castellani, <strong>and</strong> E.G. Germano), is<br />

the principal foliar disease <strong>of</strong> wheat<br />

(Triticum aestivum L.) in the cereal<br />

belt <strong>of</strong> Western Australia (Murray<br />

<strong>and</strong> Brown, 1987). The preferred<br />

method <strong>of</strong> control is the use <strong>of</strong><br />

resistant cultivars. In bread wheat<br />

inheritance <strong>of</strong> resistance to septoria<br />

nodorum blotch is reported to be<br />

complex, with an additive action <strong>of</strong><br />

the genes (Mullaney et al., 1982;<br />

Scharen <strong>and</strong> Krupinsky, 1978).<br />

Resistance is <strong>of</strong>ten linked to a<br />

number <strong>of</strong> traits such as height <strong>and</strong><br />

maturity (Rosielle <strong>and</strong> Brown, 1980;<br />

Scott et al., 1982), which makes<br />

breeding for resistance difficult.<br />

Promising levels <strong>of</strong> resistance to<br />

septoria nodorum blotch were<br />

found when a collection <strong>of</strong> T.<br />

tauschii accessions was investigated<br />

for resistance to septoria nodorum<br />

blotch. The ability to use these<br />

potential resistance sources in a<br />

wheat breeding program will be<br />

influenced by the number <strong>of</strong> genes<br />

controlling the resistance <strong>and</strong> its<br />

expression in a bread wheat<br />

background. The aim <strong>of</strong> this work<br />

was to investigate the differences<br />

between the parents in the<br />

components <strong>of</strong> resistance <strong>and</strong> to<br />

determine the number <strong>of</strong> genes<br />

controlling resistance to septoria<br />

nodorum blotch in the Triticum<br />

tauschii accession RL5271.<br />

Materials <strong>and</strong> Methods<br />

The resistant T. tauschii<br />

accession RL5271 was crossed to a<br />

susceptible accession CPI110889.<br />

The 17 F1 progenies were selfed,<br />

producing 300 F2s which were<br />

selfed to produce the F3 families.<br />

The F1 plants were screened for<br />

resistance to assess dominance, <strong>and</strong><br />

the F3 plants were screened to<br />

determine the genotypic ratio,<br />

which was used to estimate the<br />

number <strong>of</strong> genes controlling<br />

resistance.<br />

The T. tauschii seed was<br />

germinated on sterile filter papers<br />

in petri dishes <strong>and</strong> vernalized<br />

before being sown into pots in a<br />

glasshouse. In the F3 generation, a<br />

r<strong>and</strong>om selection <strong>of</strong> approximately<br />

200 F3 families was screened<br />

sequentially in two subpopulations.<br />

Each sub-population<br />

was grown on a single bench in the<br />

glasshouse, <strong>and</strong> each F3 family<br />

contained 10 plants.<br />

The plants were inoculated<br />

when the third leaf on the main<br />

stem had emerged. A single isolate<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum was used.<br />

The inoculum was produced using<br />

V8 Czapek-Dox agar (V8CDA) as a


modified method <strong>of</strong> Cooke <strong>and</strong><br />

Jones (1970). The plants were<br />

inoculated with 106 spores/mL,<br />

with a surfactant, until the point <strong>of</strong><br />

run-<strong>of</strong>f using an airless spraypainting<br />

gun while the pots were<br />

revolving on a turntable at 33 rpm.<br />

After inoculation the plants were<br />

placed in a humidity chamber for<br />

48 hours <strong>and</strong> then returned to the<br />

bench. The plants were rated using<br />

the youngest fully emerged leaf at<br />

the time <strong>of</strong> inoculation on the main<br />

tiller eight days after inoculation.<br />

The leaves were rated using disease<br />

score (DS), a 0 to 5 severity scale<br />

where 0 was an immune response<br />

<strong>and</strong> a score <strong>of</strong> 5 was given if the<br />

leaves were fully necrosed. The DS<br />

is based upon the number, size, <strong>and</strong><br />

type <strong>of</strong> lesions, as well as the area<br />

<strong>of</strong> the leaf covered by the lesions.<br />

The two parental accessions<br />

were assessed twice for two<br />

components <strong>of</strong> resistance, <strong>and</strong> the<br />

F1 plants were assessed once. The<br />

components were the incubation<br />

period (IP, the number <strong>of</strong> days from<br />

the time <strong>of</strong> inoculation till<br />

symptoms first appear) <strong>and</strong> the<br />

infection frequency (IF, the number<br />

<strong>of</strong> lesions per square centimeter <strong>of</strong><br />

leaf tissue at IP). In the first<br />

screening the IF had to be<br />

transformed using a natural<br />

logarithmic function, <strong>and</strong> during<br />

the second screening it was<br />

transformed using a square root<br />

function to remove the trends in the<br />

residuals.<br />

The F3 families were classified<br />

using the response <strong>of</strong> individual<br />

plants within an F3 family. During<br />

the rating it was evident from the<br />

parental accession’s response that a<br />

Inheritance <strong>of</strong> <strong>Septoria</strong> Nodorum Blotch Resistance in a Triticum tauschii Accession Controlled by a Single Gene 75<br />

score <strong>of</strong> less than 2.5 was resistant<br />

<strong>and</strong> a score <strong>of</strong> equal to or greater<br />

than 2.5 was susceptible. A F3<br />

family was classified as resistant if<br />

all its plants had a DS <strong>of</strong> less than<br />

2.5 <strong>and</strong> as susceptible if all the<br />

plants had a disease score greater<br />

than or equal to 2.5. Any family<br />

with plants with a DS <strong>of</strong> less than<br />

2.5 <strong>and</strong> greater than 2.5 were<br />

classified as segregating.<br />

Results<br />

In both tests the components <strong>of</strong><br />

resistance showed significant<br />

differences between the resistant<br />

<strong>and</strong> susceptible parents in their<br />

response to infection. The IP for the<br />

parents was not significantly<br />

different in the first screening but<br />

the resistant parent (RL5271) had a<br />

Table 1. The incubation period (IP), infection<br />

frequency (IF), <strong>and</strong> disease score (DS) <strong>of</strong> the<br />

F1 mean response during screening 1, <strong>and</strong> the<br />

parental accessions, RL5271 <strong>and</strong> CPI110889,<br />

during screenings 1 <strong>and</strong> 2.<br />

IF<br />

IP (lesions/<br />

(days) cm2 ) i DS<br />

Screening 1<br />

F1 8.5bii RL5271<br />

0.8a 1.3b<br />

(Resistant parent)<br />

CPI110889<br />

6.8a 0.8a 0.4a<br />

(Susceptible parent)<br />

Screening 2<br />

RL5271<br />

5.9a 2.0b 2.2c<br />

(Resistant parent)<br />

CPI110889<br />

5.5a 2.4a 0.9a<br />

(Susceptible parent) 3.5b 5.8b 3.3b<br />

i Back transformed values are presented for the<br />

IF in both screenings.<br />

ii Numbers from the same screening <strong>and</strong> for the<br />

same trait followed by different letters are<br />

significantly different from each other at p


76<br />

Session 3B — N.E.A. Murphy, R. Loughman, R. Wilson, E.S. Lagudah, R. Appels, <strong>and</strong> M.G.K. Jones<br />

different from the expected<br />

genotypic ratio for a single gene<br />

(X2 =0.81, p=0.67) (Table 2). When<br />

the genotypic ratios <strong>of</strong> the two subpopulations<br />

were compared using<br />

a Pearson’s chi-squared test for<br />

homogeneity, they were not<br />

significantly different (X2 =0.84,<br />

p=0.66). The combined ratio (45<br />

resistant F3 families, 97<br />

segregating, <strong>and</strong> 52 susceptible<br />

families) was not significantly<br />

different from the genotypic ratio<br />

expected for a single gene (X2 =0.50,<br />

p=0.78) (Table 2).<br />

Discussion<br />

Resistance to septoria nodorum<br />

blotch in the Triticum tauschii<br />

accession RL5271 is controlled by a<br />

single gene. From the F1 mean<br />

response there was no evidence <strong>of</strong><br />

complete dominance or<br />

recessiveness for resistance. The<br />

resistant parent took significantly<br />

longer to develop symptoms,<br />

developed significantly fewer<br />

infections <strong>and</strong> expressed<br />

significantly lower levels <strong>of</strong> disease<br />

than the susceptible parent. The<br />

variation in the resistance<br />

components between the two<br />

parents was consistent with<br />

previous studies reporting a<br />

resistance response. The IP has<br />

been shown to vary significantly<br />

among wheats with longer IP being<br />

associated with resistance<br />

(Wilkinson et al., 1990; Bruno <strong>and</strong><br />

Nelson, 1990). The IF was<br />

significantly lower for the resistant<br />

parent, which is consistent with<br />

previous work where a low IF has<br />

been associated with resistance<br />

(Wilkinson et al., 1990; Ma <strong>and</strong><br />

Hughes, 1993; Loughman et al.,<br />

1996). The inheritance <strong>of</strong> the IF has<br />

been found to be clearly dominant<br />

in the F1 (Ma <strong>and</strong> Hughes, 1993)<br />

which is consistent with this study.<br />

Resistance in Triticum tauschii<br />

accession RL5271 is the first singlegene<br />

resistance to septoria<br />

nodorum blotch identified in the D<br />

genome. Resistance to septoria<br />

nodorum blotch has been<br />

previously identified in another<br />

Triticum tauschii accession<br />

(=Aegilops squarrosa Tausch). It was<br />

found to be controlled by three<br />

genes, located on chromosomes 3D,<br />

5D, <strong>and</strong> 7D, with 3D being the most<br />

important <strong>of</strong> the three (Nicholson<br />

et al., 1993). The simplicity <strong>of</strong> the<br />

inheritance <strong>and</strong> strong expression<br />

<strong>of</strong> the resistance gene in T. tauschii<br />

warrants making an attempt at<br />

introgressing the resistance into a<br />

bread wheat background.<br />

Acknowledgment<br />

This work was funded by the<br />

Grains Research <strong>and</strong> Development<br />

Corporation <strong>of</strong> Australia.<br />

References<br />

Bruno, H.H., <strong>and</strong> L.R. Nelson. 1990.<br />

Partial resistance to septoria glume<br />

blotch analyzed in winter wheat<br />

seedlings. Crop Science 30:54-59.<br />

Cooke, B.M., <strong>and</strong> D.G. Jones. 1970.<br />

The effect <strong>of</strong> near-ultraviolet<br />

irradiation <strong>and</strong> agar medium on<br />

the sporulation <strong>of</strong> <strong>Septoria</strong> nodorum<br />

<strong>and</strong> <strong>Septoria</strong> tritici. Transactions <strong>of</strong><br />

the British Mycological Society<br />

54:221-226.<br />

Loughman, R., R.E. Wilson, <strong>and</strong> G.J.<br />

Thomas. 1996. Components <strong>of</strong><br />

resistance to Mycosphaerella<br />

graminicola <strong>and</strong> Phaeosphaeria<br />

nodorum in spring wheats.<br />

Euphytica 89:377-385.<br />

Ma, H., <strong>and</strong> G.R. Hughes. 1993.<br />

Resistance to septoria nodorum<br />

blotch in several Triticum species.<br />

Euphytica 70:151-157.<br />

Mullaney, J., M. Martin, <strong>and</strong> A.L.<br />

Scharen. 1982. Generation mean<br />

analysis to identify <strong>and</strong> partition<br />

the components <strong>of</strong> genetic<br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Euphytica 31:539-545.<br />

Murray, G.M., <strong>and</strong> J.F. Brown. 1987.<br />

The incidence <strong>and</strong> relative<br />

importance <strong>of</strong> wheat diseases in<br />

Australia. Australasian Plant<br />

Pathology 16:34-37.<br />

Nicholson, H., N. Rezanoor, <strong>and</strong> A.J.<br />

Worl<strong>and</strong>. 1993. Chromosomal<br />

location <strong>of</strong> resistance to <strong>Septoria</strong><br />

nodorum in a synthetic hexaploid<br />

wheat determined by the study <strong>of</strong><br />

chromosomal substitution lines in<br />

Chinese Spring wheat. Plant<br />

Breeding 110:177-184.<br />

Rosielle, A.A., <strong>and</strong> A.G.P. Brown.<br />

1980. Selection for resistance to<br />

<strong>Septoria</strong> nodorum in wheat.<br />

Euphytica 29:337-346.<br />

Scharen, A.L., <strong>and</strong> J.M. Krupinsky.<br />

1978. Detection <strong>and</strong> manipulation<br />

<strong>of</strong> resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Phytopathology 68:245-248.<br />

Scott, P.R., P.W. Benedikz, <strong>and</strong> C.J.<br />

Cox. 1982. A genetic study <strong>of</strong> the<br />

relationship between height, time<br />

<strong>of</strong> ear emergence <strong>and</strong> resistance to<br />

<strong>Septoria</strong> nodorum in wheat. Plant<br />

Pathology 31:45-60.<br />

Wilkinson, C.A., J.P. Murphy, <strong>and</strong><br />

R.C. Rufty. 1990. Diallel analysis <strong>of</strong><br />

components <strong>of</strong> partial resistance to<br />

<strong>Septoria</strong> nodorum. Plant Disease<br />

74:47-50.


Session 4: Population Dynamics<br />

Population Genetics <strong>of</strong> Mycosphaerella graminicola <strong>and</strong><br />

Phaeosphaeria nodorum<br />

B.A. McDonald, 1 C.C. Mundt, 2 <strong>and</strong> J. Zhan2 1 Institute <strong>of</strong> Plant Sciences, Phytopathology Group, Zürich, Switzerl<strong>and</strong><br />

2 Department <strong>of</strong> Botany <strong>and</strong> Plant Pathology, Oregon State University, Corvallis, OR, USA<br />

Abstract<br />

Restriction fragment length polymorphisms (RFLPs) in the nuclear (nu) <strong>and</strong> mitochondrial (mt) genomes were used to<br />

determine the genetic structure <strong>of</strong> populations <strong>of</strong> Mycosphaerella graminicola <strong>and</strong> Phaeosphaeria nodorum from<br />

around the world. Both fungi have genetic structures consistent with a regular sexual cycle <strong>and</strong> a high degree <strong>of</strong> gene flow<br />

occurring on a global scale. Gene as well as genotype diversity in the nuDNA are high for both fungi. There was no<br />

evidence for widespread clones within field populations <strong>of</strong> either fungus. While both fungi had less diversity in the mtDNA,<br />

M. graminicola exhibited significantly less diversity for the mtDNA compared to P. nodorum. Mycosphaerella<br />

graminicola populations from Patzcuaro, Mexico, <strong>and</strong> Australia exhibited significantly lower gene diversity, suggesting<br />

that these populations originated from a limited number <strong>of</strong> founders. Collections <strong>of</strong> M. graminicola taken from the same<br />

field between 1990 <strong>and</strong> 1995 showed that genetic drift is negligible, suggesting that effective population sizes are very large.<br />

A replicated field experiment showed that selection can cause significant changes in genotype frequencies during the course<br />

<strong>of</strong> a growing season, <strong>and</strong> that the contributions <strong>of</strong> immigration <strong>and</strong> recombination to genetic diversity in field populations<br />

can change over the growing season.<br />

Ten years ago, we began<br />

developing DNA-based markers as<br />

tools to learn about the population<br />

genetics <strong>of</strong> the wheat leaf blotch<br />

pathogen Mycosphaerella<br />

graminicola. One year later, we<br />

began parallel studies using the<br />

same genetic tools for the wheat<br />

glume blotch pathogen<br />

Phaeosphaeria nodorum. We began<br />

with two elementary questions<br />

regarding the population genetics<br />

<strong>of</strong> both fungi. How much genetic<br />

diversity is present within<br />

populations? How is genetic<br />

diversity distributed within <strong>and</strong><br />

among populations? As our<br />

knowledge <strong>of</strong> the genetic structure<br />

<strong>of</strong> both pathogens deepened, we<br />

addressed more complex questions.<br />

What are the relative contributions<br />

<strong>of</strong> sexual <strong>and</strong> asexual reproduction<br />

to the genetic structure <strong>of</strong><br />

populations? How stable are<br />

populations over time? Does<br />

selection for specific pathogen<br />

genotypes occur on particular host<br />

genotypes? Is there evidence for<br />

host specialization in these<br />

pathosystems?<br />

To address the latter questions,<br />

we utilized increasingly<br />

sophisticated field experiments to<br />

differentiate among the various<br />

evolutionary forces acting on<br />

populations <strong>of</strong> these fungi. In this<br />

manuscript, I will briefly review<br />

our underst<strong>and</strong>ing <strong>of</strong> the<br />

population genetics <strong>of</strong> both fungi at<br />

this point in time. The majority <strong>of</strong><br />

this manuscript was distilled from a<br />

review chapter written for the Long<br />

Ashton Symposium on <strong>Septoria</strong> in<br />

<strong>Cereals</strong> held in 1997. Detailed data<br />

to support our interpretations are<br />

presented in that chapter<br />

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

Materials <strong>and</strong> Methods<br />

77<br />

DNA markers<br />

The RFLP markers utilized for<br />

these studies were developed in the<br />

same way for both fungi utilizing<br />

the methods described in<br />

McDonald <strong>and</strong> Martinez (1990b).<br />

Single-locus probes were used to<br />

measure gene diversity for<br />

individual RFLP loci <strong>and</strong> to<br />

measure population subdivision<br />

<strong>and</strong> genetic similarity among<br />

populations (McDonald <strong>and</strong><br />

Martinez, 1990a; Boeger et al., 1993;<br />

McDonald et al., 1994; Keller et al.,<br />

1997a,b). Probes that hybridized to<br />

repetitive elements were used for


78<br />

Session 4 — B.A. McDonald, C.C. Mundt, <strong>and</strong> J. Zhan<br />

DNA fingerprinting <strong>and</strong> to make<br />

measures <strong>of</strong> genotype diversity<br />

(McDonald <strong>and</strong> Martinez, 1991;<br />

Keller et al., 1997a,b). To measure<br />

variation in the mitochondrial (mt)<br />

genome, we purified mtDNA <strong>and</strong><br />

used it as a probe to hybridize to<br />

the entire digested mitochondrial<br />

genome for each individual.<br />

Sampling methods<br />

Hierarchical sampling methods<br />

now are commonly used in plant<br />

pathology (Kohli et al., 1995;<br />

McDonald et al., 1995; McDonald,<br />

1997). For many <strong>of</strong> the collections<br />

described in this manuscript, a<br />

st<strong>and</strong>ardized six-site hierarchy was<br />

used to sample field populations<br />

(McDonald et al., 1999). For other<br />

field collections, long transects were<br />

arranged in a field <strong>and</strong> leaves were<br />

sampled at 1-2-meter intervals<br />

along each transect. For some<br />

collections, leaves were collected at<br />

r<strong>and</strong>om in a field. All isolates in a<br />

collection were taken from a single<br />

field at a single time point for the<br />

majority <strong>of</strong> collections. I refer to<br />

these collections as field<br />

populations. Four collections <strong>of</strong><br />

isolates came from the same<br />

country or geographical region, but<br />

did not originate from a single field.<br />

I refer to these as regional<br />

populations to distinguish them<br />

from field populations.<br />

For P. nodorum, the Arkansas<br />

population was made up <strong>of</strong> isolates<br />

collected from infected seed from<br />

two different seed lots distributed<br />

in Arkansas. The crested<br />

wheatgrass population consisted <strong>of</strong><br />

isolations made from the weed<br />

crested wheatgrass over a two-year<br />

period in the state <strong>of</strong> North Dakota.<br />

The goal for each collection was to<br />

have a large enough field sample to<br />

be confident that allele frequencies<br />

for individual RFLP loci were<br />

representative <strong>of</strong> the entire field<br />

population. Only populations<br />

having at least 25 isolates were<br />

included in the analysis. A<br />

summary <strong>of</strong> the M. graminicola <strong>and</strong><br />

P. nodorum isolates in our collection<br />

was presented previously<br />

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

Data analysis<br />

Methods used to measure gene<br />

<strong>and</strong> genotype diversity, genetic<br />

similarity, population subdivision,<br />

<strong>and</strong> gametic disequilibrium have<br />

been published elsewhere<br />

(McDonald <strong>and</strong> Martinez, 1990a;<br />

Boeger et al., 1993; Chen et al., 1994;<br />

Chen <strong>and</strong> McDonald, 1996). The<br />

DNA fingerprint probes were<br />

validated for both fungi in previous<br />

experiments (McDonald <strong>and</strong><br />

Martinez, 1991; Keller et al., 1997b).<br />

Here we will pay special attention<br />

to the collection <strong>of</strong> M. graminicola<br />

isolates from Mexico.<br />

Results <strong>and</strong> Discussion<br />

Genetic diversity <strong>and</strong><br />

mating/reproduction systems<br />

Both gene <strong>and</strong> genotype<br />

diversity are high for both fungi.<br />

On average, 18 alleles were present<br />

for 11 RFLP loci across ~2000<br />

isolates <strong>of</strong> M. graminicola, <strong>and</strong> five<br />

alleles were present for seven loci<br />

across ~950 isolates <strong>of</strong> P. nodorum.<br />

In each population, between 3-4<br />

alleles were present in ~90% <strong>of</strong> the<br />

isolates (McDonald et al., 1999).<br />

Nei’s measure <strong>of</strong> gene diversity<br />

across all populations averaged 0.44<br />

across eight RFLP loci for M.<br />

graminicola <strong>and</strong> 0.51 across seven<br />

RFLP loci for P. nodorum. Though<br />

both fungi exhibited high gene<br />

diversities, M. graminicola had a<br />

significantly greater number <strong>of</strong><br />

alleles per locus then P. nodorum<br />

<strong>and</strong> P. nodorum had a more even<br />

distribution <strong>of</strong> allele frequencies.<br />

Comparisons <strong>of</strong> gene diversities<br />

across populations revealed some<br />

interesting patterns. The Israeli<br />

population had significantly higher<br />

gene diversity than other<br />

populations around the world<br />

(McDonald et al., 1999). We<br />

interpret this as evidence that the<br />

Middle East is a center <strong>of</strong> diversity<br />

for M. graminicola, <strong>and</strong> the likely<br />

center <strong>of</strong> origin for this fungus. The<br />

Mexican population exhibited<br />

lower gene diversity than other<br />

populations. It was made from<br />

leaves collected by Dr. Lucy<br />

Gilchrist at a <strong>CIMMYT</strong> disease<br />

nursery in Patzcuaro, a location<br />

remote from other wheat fields (L.<br />

Gilchrist, personal communication).<br />

The Mexican population also<br />

exhibited a lower genotype<br />

diversity than expected (McDonald<br />

et al., 1999), having a higher<br />

incidence <strong>of</strong> a few widespread<br />

clones than any other population<br />

surveyed. These findings are<br />

consistent with a small founding<br />

population that has not undergone<br />

regular sexual reproduction. We<br />

hypothesized that these isolates<br />

represent an historic inoculation <strong>of</strong><br />

the Patzcuaro site with a few<br />

isolates that were not sexually<br />

compatible. We hope to collect<br />

another M. graminicola field<br />

population from wheat-growing<br />

areas <strong>of</strong> Mexico to obtain a more<br />

representative sample <strong>of</strong> this<br />

fungus in that region.


The gene diversity in the P.<br />

nodorum population was lowest in<br />

the collections from crested<br />

wheatgrass <strong>and</strong> from Mexico<br />

(McDonald et al., 1999). These were<br />

the only two collections that were<br />

fixed for one allele at an RFLP<br />

locus. These populations also<br />

exhibited substantial differences in<br />

allele frequencies for some RFLP<br />

loci when compared to other<br />

populations (McDonald et al.,<br />

1999). It is possible that the lower<br />

gene diversity <strong>and</strong> different allele<br />

frequencies in the crested<br />

wheatgrass population resulted<br />

from selection due to host<br />

specialization. However, the<br />

finding that common alleles are<br />

shared between isolates taken from<br />

crested wheatgrass <strong>and</strong> wheat<br />

suggests that crested wheatgrass<br />

may serve as a reservoir <strong>of</strong><br />

inoculum for P. nodorum that infects<br />

wheat. The differences in the<br />

Patzcuaro population may reflect<br />

founder events <strong>and</strong> the geographic<br />

isolation <strong>of</strong> this location as<br />

described previously for M.<br />

graminicola.<br />

Genotype diversity. Genotype<br />

diversity in the nuclear genome<br />

was high for both fungi (McDonald<br />

et al., 1999). In the majority <strong>of</strong><br />

populations surveyed, each leaf<br />

was colonized by a unique fungal<br />

genotype. When isolates shared the<br />

same DNA fingerprint, they usually<br />

were sampled from the same leaf.<br />

The only evidence for a widespread<br />

clone in M. graminicola was in the<br />

Patzcuaro population. The only<br />

evidence for a widespread clone in<br />

P. nodorum was from our first<br />

collection in a Texas field<br />

(McDonald et al., 1994).<br />

Population Genetics <strong>of</strong> Mycosphaerella graminicola <strong>and</strong> Phaeosphaeria nodorum 79<br />

It appears that the typical<br />

population <strong>of</strong> both fungi exhibits a<br />

low degree <strong>of</strong> clonality at the field<br />

level. Clones were usually<br />

distributed across an area <strong>of</strong><br />

approximately 1 square meter <strong>and</strong><br />

did not become widespread. This<br />

finding is consistent with limited<br />

spread <strong>of</strong> the splash-dispersed<br />

conidia for both fungi.<br />

Our interpretation <strong>of</strong> the high<br />

degree <strong>of</strong> genotype diversity is that<br />

populations <strong>of</strong> both fungi undergo<br />

regular sexual cycles <strong>and</strong> that the<br />

primary inoculum for both fungi is<br />

likely to be ascospores (Chen <strong>and</strong><br />

McDonald, 1996; Keller et al.,<br />

1997b). We have used tests for the<br />

r<strong>and</strong>omness <strong>of</strong> associations among<br />

RFLP loci to indicate the frequency<br />

<strong>of</strong> sexual recombination for both<br />

fungi (Chen <strong>and</strong> McDonald, 1996).<br />

Associations were r<strong>and</strong>om for both<br />

fungi in the cases where sample<br />

sizes were adequate to make robust<br />

tests for disequilibrium, supporting<br />

the hypothesis <strong>of</strong> r<strong>and</strong>om mating.<br />

Though the sample size was<br />

smaller (N=160), the Israel<br />

population also showed the<br />

gametic equilibrium <strong>and</strong> high<br />

genotypic diversity typical <strong>of</strong> a<br />

r<strong>and</strong>om mating fungus (McDonald<br />

et al., 1999), suggesting that the<br />

teleomorph is present in the Middle<br />

East.<br />

After a field experiment<br />

provided evidence that the<br />

teleomorph was contributing<br />

significantly to infection during the<br />

growing season (Zhan et al., 1998),<br />

we sampled lower leaves in the<br />

crop canopy in Oregon <strong>and</strong> found<br />

the teleomorph existing as a<br />

significant fraction <strong>of</strong> all fruiting<br />

bodies on wheat leaves (C. Cowger,<br />

C.C. Mundt, <strong>and</strong> B.A. McDonald,<br />

unpublished).<br />

mtDNA diversity. For both fungi,<br />

the mtDNA exhibited less diversity<br />

then the nuDNA. The mtDNA<br />

genome ranges from 48-62 kb in size<br />

for both fungi. Among 385 isolates<br />

<strong>of</strong> P. nodorum from Switzerl<strong>and</strong>,<br />

Texas, <strong>and</strong> Oregon, we found only<br />

46 different mtDNA haplotypes<br />

compared to over 300 nuDNA<br />

haplotypes. Fifteen <strong>of</strong> the mtDNA<br />

haplotypes were shared among<br />

these field populations (S. Keller<br />

<strong>and</strong> B. A. McDonald, unpublished<br />

data).<br />

The mtDNA <strong>of</strong> M. graminicola<br />

exhibited even lower diversity. We<br />

routinely found only 2-3 mtDNA<br />

haplotypes in field populations <strong>of</strong><br />

M. graminicola (McDonald et al.,<br />

1999). We sometimes found only<br />

one M. graminicola mtDNA<br />

haplotype in a field (McDonald et<br />

al., 1999). We picked the three most<br />

common mtDNA haplotypes from<br />

around the world <strong>and</strong> digested the<br />

mtDNA with 10 different restriction<br />

enzymes that sampled ~5% <strong>of</strong> the<br />

entire 48 kb <strong>of</strong> mtDNA sequence in<br />

an attempt to detect cryptic<br />

variation. Types 1, 2, <strong>and</strong> 3 mtDNA<br />

haplotypes produced identical<br />

digestion patterns with all enzymes<br />

in all populations tested, suggesting<br />

that these mtDNA haplotypes are<br />

identical globally (B.A. McDonald<br />

<strong>and</strong> K. Hogan, unpublished). We<br />

have proposed that the limited<br />

mtDNA diversity is due to a<br />

selective sweep that occurred as M.<br />

graminicola populations became<br />

specialized to infect bread wheat<br />

(McDonald et al., 1999).


80<br />

Session 4 — B.A. McDonald, C.C. Mundt, <strong>and</strong> J. Zhan<br />

Genetic drift<br />

The high nuclear gene diversity<br />

found in most populations for both<br />

fungi suggests that population<br />

sizes are very large <strong>and</strong> the effects<br />

<strong>of</strong> genetic drift are small. In a<br />

previous experiment, we showed<br />

that the population genetic<br />

structure <strong>of</strong> an M. graminicola<br />

population did not change over a<br />

three-year period <strong>of</strong> 1990-1992<br />

(Chen et al., 1994). We have<br />

extended data from this population<br />

to 1995 <strong>and</strong> find the same result.<br />

The frequencies <strong>of</strong> neutral RFLP<br />

alleles did not change significantly<br />

over time (Table 1), <strong>and</strong> the degree<br />

<strong>of</strong> population subdivision across<br />

years is less than 1% (Table 2),<br />

demonstrating that population<br />

sizes are large enough to make<br />

genetic drift insignificant as an<br />

evolutionary force in populations<br />

<strong>of</strong> this fungus.<br />

Population subdivision<br />

<strong>and</strong> gene flow<br />

The distribution <strong>of</strong> genetic<br />

diversity across populations was<br />

consistent with a significant level<br />

<strong>of</strong> gene flow for both fungi.<br />

Common alleles at individual<br />

RFLP loci were shared among<br />

nearly all populations (McDonald<br />

et al., 1999), <strong>and</strong> allele frequencies<br />

<strong>of</strong>ten were quite similar though<br />

populations were separated by<br />

thous<strong>and</strong>s <strong>of</strong> kilometers<br />

(McDonald et al., 1999). We have<br />

interpreted this similarity as<br />

indicating a significant degree <strong>of</strong><br />

gene flow among populations <strong>of</strong><br />

both fungi on an international scale<br />

(McDonald et al., 1995; Keller et al.,<br />

1997). Global estimates <strong>of</strong> GST were<br />

0.06 <strong>and</strong> 0.07 for M. graminicola <strong>and</strong><br />

P. nodorum, respectively. These<br />

values <strong>of</strong> GST are consistent with<br />

movement <strong>of</strong> 11 <strong>and</strong> 7 individuals<br />

among populations globally every<br />

generation.<br />

Table 1. Allele frequencies for seven RFLP loci in Mycosphaerella graminicola collections from<br />

Corvallis, Oregon, in 1990, 1991, 1992, <strong>and</strong> 1995. Alleles present at a frequency <strong>of</strong> less than 0.05 in<br />

all four populations were pooled into one category (allele “p”). Sample sizes (N) used to<br />

calculate allele frequencies are shown for each locus. The chi-square values <strong>and</strong> their degrees<br />

<strong>of</strong> freedom (in parentheses) for measurement <strong>of</strong> population differentiation are also included.<br />

Populations<br />

RFLP Locus Allele 1990 1991 1992 1995 c2 pSTS192-PstIA 1 0.943 0.877 0.911 0.953 6.36(6)<br />

2 0.030 0.088 0.054 0.047<br />

p 0.027 0.035 0.036 0.000<br />

N 401 57 56 43<br />

pSTS192-PstIB 1 0.980 1.000 0.981 1.000 1.99(3)<br />

P 0.020 0.000 0.019 0.000<br />

N 406 56 56 44<br />

pSTS14-PstI 1 0.808 0.825 0.821 0.818 0.85(6)<br />

2 0.187 0.175 0.179 0.182<br />

p 0.005 0.000 0.000 0.000<br />

N 407 57 56 44<br />

pSTS2-PstI 1 0.648 0.667 0.571 0.818 8.87(9)<br />

2 0.085 0.056 0.125 0.023<br />

3 0.203 0.204 0.214 0.136<br />

p 0.065 0.074 0.089 0.023<br />

N 400 54 56 44<br />

pSTL10-PstI 1 0.680 0.649 0.518 0.643 14.51(9)<br />

2 0.051 0.053 0.107 0.000<br />

3 0.220 0.281 0.357 0.286<br />

p 0.049 0.018 0.018 0.071<br />

N 409 57 56 43<br />

pSTL53-PstI 1 0.475 0.561 0.536 0.465 26.86(18)<br />

2 0.119 0.070 0.107 0.023<br />

3 0.161 0.158 0.232 0.070<br />

5 0.042 0.053 0.000 0.047<br />

6 0.119 0.140 0.107 0.209<br />

7 0.027 0.000 0.000 0.047<br />

p 0.057 0.018 0.018 0.140<br />

N 404 57 56 43<br />

pSTS197-PstI 1 0.632 0.649 0.589 0.591<br />

2 0.227 0.175 0.286 0.364 9.65(9)<br />

3 0.108 0.158 0.107 0.045<br />

p 0.033 0.018 0.018 0.000<br />

N 259 57 56 44<br />

Table 2. Nei’s measures <strong>of</strong> gene diversity <strong>and</strong> population subdivision for 7 RFLP loci in<br />

Mycosphaerella graminicola populations collected from Corvallis, Oregon, in 1990, 1991, 1992,<br />

<strong>and</strong> 1995.<br />

Hi a<br />

The present lack <strong>of</strong> subdivision<br />

in the nuclear genome probably<br />

reflects historic movement <strong>of</strong> both<br />

fungi around the world. It is<br />

possible that gene flow is not a<br />

RFLP Loci 1990 1991 1992 1995 HTb HSc Dstd Gste pSTS192A-PstI 0.110 0.222 0.166 0.091 0.126 0.125 0.001 0.008<br />

pSTS192B-PstI 0.039 0.000 0.104 0.000 0.039 0.039 0.000 0.007<br />

pSTS14-PstI 0.312 0.289 0.294 0.298 0.307 0.307 0.000 0.000<br />

pSTS2-PstI 0.531 0.508 0.610 0.323 0.524 0.520 0.004 0.007<br />

pSTL10-PstI 0.486 0.497 0.592 0.526 0.506 0.501 0.005 0.010<br />

pSTL53-PstI 0.716 0.633 0.636 0.722 0.704 0.700 0.004 0.006<br />

pSTS197-PstI 0.537 0.523 0.560 0.517 0.540 0.536 0.004 0.007<br />

Pooled 0.390 0.382 0.423 0.354 0.392 0.390 0.003 0.006<br />

Notes:a= gene diversity for each population; b= combined gene diversity across all collections; c= gene<br />

diversity within populations; d= gene diversity among populations; e= population differentiation


significant evolutionary force at<br />

present. However, we consider it<br />

more likely that some gene flow<br />

continues as a result <strong>of</strong> the global<br />

commerce in grain. The obvious<br />

mechanism for gene flow on a<br />

regional basis is air-dispersed<br />

ascospores. In the case <strong>of</strong> P.<br />

nodorum, the most likely<br />

mechanism for intercontinental<br />

dispersal is infected seed (King et<br />

al., 1983). Since it has been shown<br />

that M. graminicola can infect seed<br />

(Brokenshire, 1975) we consider it<br />

likely that this also is the<br />

mechanism for long distance gene<br />

flow in M. graminicola. Whatever<br />

the mechanism, the high degree <strong>of</strong><br />

similarity in populations around<br />

the world for both fungi suggests<br />

that they have been transported<br />

around the world by humans.<br />

Evidence for selection<br />

We recently completed an<br />

experiment to measure competition<br />

among 10 genotypes <strong>of</strong> M.<br />

graminicola in a field setting. The 10<br />

isolates were inoculated onto three<br />

host treatments consisting <strong>of</strong> a<br />

moderately resistant wheat variety<br />

(Madsen), a susceptible wheat<br />

variety (Stephens) <strong>and</strong> a 1:1<br />

mixture <strong>of</strong> these cultivars. Our<br />

most important finding in this<br />

experiment was that intense<br />

competition appeared to occur<br />

among the different genotypes.<br />

Significant changes in the<br />

frequencies <strong>of</strong> specific pathogen<br />

genotypes occurred over the season<br />

(McDonald et al., 1999). Some<br />

isolates showed evidence for<br />

adaptation to particular hosts. The<br />

results from this experiment<br />

provided our first direct evidence<br />

Population Genetics <strong>of</strong> Mycosphaerella graminicola <strong>and</strong> Phaeosphaeria nodorum 81<br />

that selection operates on specific<br />

M. graminicola pathogen genotypes<br />

in a field setting.<br />

We have not yet conducted<br />

similar replicated field experiments<br />

to measure selection in P. nodorum.<br />

But we have indirect evidence that<br />

selection does not result in<br />

widespread clones that are adapted<br />

to specific host genotypes. In an<br />

experiment conducted in<br />

Switzerl<strong>and</strong> in collaboration with<br />

Martin Wolfe’s group, we sampled<br />

50 isolates <strong>of</strong> P. nodorum from each<br />

<strong>of</strong> nine wheat fields near Zurich.<br />

Three different wheat varieties were<br />

represented three times each among<br />

the nine wheat fields. Though fields<br />

planted to the same variety used the<br />

same source <strong>of</strong> seed, no genotypes<br />

were shared among field<br />

populations. Only six pairs <strong>of</strong> clones<br />

were found among the 432 isolates<br />

that were assayed. Isolates with the<br />

same DNA fingerprints always<br />

came from the same site within a<br />

field (Keller et al., 1997b).<br />

Taken together, all <strong>of</strong> our<br />

experiments suggest that nuclear<br />

genotypes do not persist through<br />

time for either fungus. Instead, the<br />

genes are the units <strong>of</strong> selection that<br />

are carried forward across<br />

generations. Selection operates on<br />

the population instead <strong>of</strong> the<br />

individual. In order to gain a<br />

representative spectrum <strong>of</strong> the<br />

diversity for virulence in natural<br />

populations, plant breeders should<br />

include the widest possible<br />

diversity <strong>of</strong> strains when screening<br />

germplasm for resistance to these<br />

fungi. Similarly, chemical<br />

companies should include at least<br />

several hundred strains in their<br />

screens for resistance to fungicides.<br />

Conclusions<br />

Given our present data, we have<br />

drawn the following conclusions<br />

regarding the evolutionary forces<br />

that affect the population genetics<br />

<strong>of</strong> M. graminicola <strong>and</strong> P. nodorum:<br />

• For both fungi, the mating system<br />

includes both sexual <strong>and</strong> asexual<br />

reproduction. Asexual<br />

reproduction may have an<br />

important impact over an area <strong>of</strong><br />

a few square meters, but the<br />

sexual reproduction has much<br />

greater consequences for the<br />

evolutionary biology <strong>of</strong> both<br />

fungi. Genotypes are ephemeral<br />

but genes persist in populations<br />

through time.<br />

• Population sizes are large enough<br />

to make genetic drift negligible<br />

for both fungi. Large population<br />

sizes also ensure that ample<br />

mutations are present in every<br />

population to allow for a rapid<br />

response to selection, e.g.<br />

mutations from avirulence to<br />

virulence for major resistance<br />

genes. The population from<br />

Patzcuaro, Mexico, exhibits a<br />

genetic structure consistent with<br />

a founder effect.<br />

• Gene flow is sufficient to unite<br />

large geographical areas into a<br />

single genetic population. If gene<br />

flow is ongoing, then breeders<br />

should continue to test their<br />

resistant lines over the widest<br />

possible geographical area. If<br />

gene flow is episodic, continued<br />

vigilance is needed to limit the<br />

spread <strong>of</strong> new virulence genes<br />

<strong>and</strong> fungicide resistance genes.<br />

Quarantines in areas with low<br />

gene diversity, such as Australia,<br />

should be enforced to limit the<br />

evolutionary potential <strong>of</strong> these<br />

populations. If <strong>CIMMYT</strong><br />

continues to use Patzcuaro as a<br />

field site to screen for resistance


82<br />

Session 4 — B.A. McDonald, C.C. Mundt, <strong>and</strong> J. Zhan<br />

to M. graminicola <strong>and</strong> P. nodorum,<br />

it may want to consider<br />

introducing more diverse fungal<br />

populations from other parts <strong>of</strong><br />

Mexico into this disease nursery.<br />

• Selection appears to operate on<br />

both nuclear <strong>and</strong> mitochondrial<br />

genomes in M. graminicola.<br />

Though selection may increase<br />

the frequency <strong>of</strong> particular<br />

genotypes over the course <strong>of</strong> a<br />

growing season, it appears that<br />

particular genotypes are unlikely<br />

to reach high frequencies within<br />

field populations because <strong>of</strong> the<br />

limited dispersal potential for<br />

conidia. But the genes in the most<br />

fit individuals will persist <strong>and</strong> be<br />

recombined to create novel<br />

genotypes in the next growing<br />

season. Over the course <strong>of</strong> many<br />

growing seasons, selection will<br />

change the frequency <strong>of</strong> genes<br />

that affect adaptation to the<br />

wheat host, but new genotypes<br />

will appear each season. It is too<br />

early to say if selection operates<br />

the same way in P. nodorum.<br />

In summary, the population<br />

genetics <strong>of</strong> P. nodorum <strong>and</strong> M.<br />

graminicola are very similar. This<br />

probably reflects the similarity in<br />

their life histories. Both fungi<br />

produce airborne sexual ascospores<br />

<strong>and</strong> splash-dispersed asexual<br />

spores. They both infect aboveground<br />

plant parts on the same<br />

host <strong>and</strong> they both infect seeds that<br />

can be transported globally as part<br />

<strong>of</strong> the world commerce in wheat.<br />

Use <strong>of</strong> multi-allelic, neutral genetic<br />

markers combined with<br />

hierarchical sampling has allowed<br />

us to achieve a much greater<br />

underst<strong>and</strong>ing <strong>of</strong> the population<br />

biology <strong>of</strong> both fungi.<br />

Acknowledgments<br />

BAM gratefully acknowledges<br />

the many collectors <strong>and</strong><br />

collaborators around the world who<br />

responded to his request for<br />

infected leaf material. Funding for<br />

this project came from the USDA<br />

National Research Initiative<br />

Competitive Grants Program (Grant<br />

# 93-37303-9039), the National<br />

Science Foundation (Grant # DEB-<br />

9306377), the Texas Agricultural<br />

Experiment Station (Hatch project<br />

#6928), <strong>and</strong> the Swiss National<br />

Fund (Grant # 5002-38966).<br />

References<br />

Boeger, J.M., Chen, R.S., <strong>and</strong><br />

McDonald, B.A. 1993. Gene flow<br />

between geographic populations <strong>of</strong><br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici) detected<br />

with RFLP markers.<br />

Phytopathology 83:1148-1154.<br />

Brokenshire, T. 1975. Wheat seed<br />

infection by <strong>Septoria</strong> tritici.<br />

Transactions <strong>of</strong> the British<br />

Mycological Society 64:331-335.<br />

Chen, R.S., <strong>and</strong> McDonald, B.A. 1996.<br />

Sexual reproduction plays a major<br />

role in the genetic structure <strong>of</strong><br />

populations <strong>of</strong> the fungus<br />

Mycosphaerella graminicola. Genetics<br />

142:1119-1127.<br />

Chen, R.S., Boeger, J.M., <strong>and</strong><br />

McDonald, B.A. 1994. Genetic<br />

stability in a population <strong>of</strong> a plant<br />

pathogenic fungus over time.<br />

Molecular Ecology 3:209-218.<br />

Keller, S.M., McDermott, J.M.,<br />

Pettway, R.E., Wolfe, M.S., <strong>and</strong><br />

McDonald, B.A. 1997a. Gene flow<br />

<strong>and</strong> sexual reproduction in the<br />

wheat glume blotch pathogen<br />

Phaeosphaeria nodorum (anamorph<br />

<strong>Stagonospora</strong> nodorum).<br />

Phytopathology 87:353-358.<br />

Keller, S.M., Wolfe, M.S., McDermott,<br />

J.M., <strong>and</strong> McDonald, B.A. 1997b.<br />

High genetic similarity among<br />

populations <strong>of</strong> Phaeosphaeria<br />

nodorum across wheat cultivars <strong>and</strong><br />

regions in Switzerl<strong>and</strong>.<br />

Phytopathology 87:1134-1139.<br />

King, J.E., Cook, R.J., <strong>and</strong> Melville, S.C.<br />

1983. A review <strong>of</strong> <strong>Septoria</strong> diseases<br />

<strong>of</strong> wheat <strong>and</strong> barley. Annals<br />

Applied Biology 103:345-373.<br />

Kohli, Y., Brunner, L.J., Yoell, H.,<br />

Milgroom, M.G., Anderson, J.B.,<br />

Morrall, R.A.A., <strong>and</strong> Kohn, L.M.<br />

1995. Clonal dispersal <strong>and</strong> spatial<br />

mixing in populations <strong>of</strong> the plant<br />

pathogenic fungus, Sclerotinia<br />

sclerotiorum. Molecular Ecology<br />

4:69-77.<br />

McDonald, B.A. 1997. The population<br />

genetics <strong>of</strong> fungi: tools <strong>and</strong><br />

techniques. Phytopathology 87:448-<br />

453.<br />

McDonald, B.A., <strong>and</strong> Martinez, J.P.<br />

1990a. DNA restriction fragment<br />

length polymorphisms among<br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici) isolates<br />

collected from a single wheat field.<br />

Phytopathology 80:1368-1373.<br />

McDonald, B.A., <strong>and</strong> Martinez, J.P.<br />

1990b. Restriction fragment length<br />

polymorphisms in <strong>Septoria</strong> tritici<br />

occur at a high frequency. Current<br />

Genetics 17:133-138.<br />

McDonald, B.A., <strong>and</strong> Martinez, J.P.<br />

1991. DNA fingerprinting <strong>of</strong> the<br />

plant pathogenic fungus<br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici).<br />

Experimental Mycology 15:146-158.<br />

McDonald, B.A., Miles, J., Nelson, L.R.,<br />

<strong>and</strong> Pettway, R.E. 1994. Genetic<br />

variability in nuclear DNA in field<br />

populations <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum. Phytopathology 84:250-<br />

255.<br />

McDonald, B.A., Pettway, R.E., Chen,<br />

R.S., Boeger, J.M., <strong>and</strong> Martinez, J.P.<br />

1995. The population genetics <strong>of</strong><br />

<strong>Septoria</strong> tritici (teleomorph<br />

Mycosphaerella graminicola).<br />

Canadian Journal <strong>of</strong> Botany 73<br />

(supplement), S292-S301.<br />

McDonald, B.A., Zhan J., Yarden O.,<br />

Hogan K., Garton J., <strong>and</strong> Pettway<br />

R.E. 1999. The population genetics<br />

<strong>of</strong> Mycosphaerella graminicola <strong>and</strong><br />

Phaeosphaeria nodorum. pp. 44-69 In:<br />

<strong>Septoria</strong> in <strong>Cereals</strong>: a Study <strong>of</strong><br />

Pathosystems. Lucas, J.A., Bowyer,<br />

P., Anderson, H.M., eds. CABI<br />

Publishing, Wallingford, UK.<br />

Zhan, J., Mundt, C.C., <strong>and</strong> McDonald,<br />

B.A. 1998. Measuring immigration<br />

<strong>and</strong> sexual reproduction in field<br />

populations <strong>of</strong> Mycosphaerella<br />

graminicola. Phytopathology<br />

88:1330-1337.


Characterization <strong>of</strong> Less Aggressive <strong>Stagonospora</strong><br />

nodorum Isolates from Wheat<br />

E. Arseniuk, 1 H.S. Tsang, 2 J.M. Krupinsky, 3 <strong>and</strong> P.P. Ueng 4<br />

1 Plant Breeding <strong>and</strong> Acclimatization Institute, Radzików, Pol<strong>and</strong><br />

2 Department <strong>of</strong> Biochemistry, University <strong>of</strong> Maryl<strong>and</strong>, MD, USA<br />

3 USDA-ARS, Northern Great Plains Research Lab, M<strong>and</strong>an, ND, USA<br />

4 USDA-ARS, Molecular Plant Pathology Lab, Beltsville, MD, USA<br />

Abstract<br />

Two less aggressive <strong>Stagonospora</strong> nodorum isolates, 9074 <strong>and</strong> 9076, were characterized by inoculation tests, mating<br />

ability, <strong>and</strong> molecular tools. As expected, they caused mild symptom severity on wheat, triticale, <strong>and</strong> rye. They crossed<br />

with other S. nodorum isolates <strong>and</strong> had “+” mating type determinant(s). These two isolates also had the same restriction<br />

patterns <strong>and</strong> sequences in the internal transcribed spacer (ITS) region <strong>of</strong> rDNA similar to other S. nodorum isolates.<br />

With AFLP analysis, it was concluded that these less aggressive S. nodorum isolates were closely related <strong>and</strong> genetically<br />

different from highly aggressive ones.<br />

<strong>Stagonospora</strong> nodorum (Berk.)<br />

Castellani & E.G. Germano<br />

(teleomorph Phaeosphaeria nodorum<br />

(E. Müller) Hedjaroude) causes<br />

septoria nodorum blotch on wheat,<br />

barley, <strong>and</strong> other small grains<br />

(Sprague, 1950). Genetic variation<br />

<strong>of</strong> S. nodorum isolates has been<br />

reported (Allingham <strong>and</strong> Jackson,<br />

1981; Krupinsky, 1997a,b; Scharen et<br />

al., 1985). Underst<strong>and</strong>ing this<br />

genetic variation would help in the<br />

selection <strong>and</strong>/or development <strong>of</strong><br />

resistant germplasm. This study<br />

was undertaken to determine<br />

whether less aggressive isolates<br />

could be characterized <strong>and</strong> how<br />

they relate to highly aggressive<br />

isolates.<br />

Materials <strong>and</strong> Methods<br />

Inoculation tests<br />

<strong>Stagonospora</strong> nodorum isolates<br />

were obtained from lesions on<br />

infected wheat (Sn26-1, 9074, <strong>and</strong><br />

9076) <strong>and</strong> rye (Sn48-1) in Pol<strong>and</strong><br />

<strong>and</strong> USA. Isolates 9074 <strong>and</strong> 9076<br />

were collected from Richl<strong>and</strong> <strong>and</strong><br />

Gallatin counties in Montana,<br />

respectively. These two isolates<br />

were associated with mild symptom<br />

severity on detached wheat leaves<br />

<strong>and</strong> seedlings, <strong>and</strong> considered to be<br />

less aggressive isolates (Krupinsky,<br />

1997a,b). Plants were inoculated by<br />

spraying a 3-ml pycnidiospore<br />

suspension (3 x 10 6 spores/ml) per<br />

pot onto 10-day-old seedlings.<br />

Three wheat cultivars (Alba, Begra,<br />

<strong>and</strong> Liwilla), two triticales (Bogo<br />

<strong>and</strong> Pinokio) <strong>and</strong> one rye (Zduno)<br />

were used as host plants. The<br />

plants were maintained in a high<br />

humidity environment for 72 h<br />

after spraying. Fourteen days after<br />

inoculation, 10 primary leaves were<br />

assessed for percentage necrosis<br />

<strong>and</strong> scored using a 1 (resistant) to 9<br />

(susceptible) scale. The experiments<br />

were repeated three times.<br />

Determination <strong>of</strong> mating type<br />

Isolates 9074 <strong>and</strong> 9076 were<br />

mated in vitro with reference “+”<br />

<strong>and</strong> “–“ strains <strong>of</strong> the pathogen<br />

with a procedure developed in the<br />

lab (Arseniuk et al., 1997).<br />

DNA isolation <strong>and</strong><br />

characterization <strong>of</strong> ITS <strong>of</strong><br />

nuclear rDNA<br />

Growth <strong>of</strong> fungal culture in a<br />

liquid medium, isolation <strong>and</strong><br />

purification <strong>of</strong> genomic DNA<br />

mainly followed the procedures<br />

83<br />

described earlier (Ueng et al., 1992).<br />

The internal transcribed spacer<br />

(ITS) region <strong>of</strong> the nuclear rDNA<br />

repeat units was amplified by PCR<br />

following the protocol reported<br />

earlier (Ueng et al., 1998; White et<br />

al., 1990). Two highly aggressive S.<br />

nodorum isolates, 8408 <strong>and</strong> 9506-2<br />

(Krupinsky, 1997a; Krupinsky,<br />

unpublished data) were compared<br />

with the two less aggressive ones,<br />

9074 <strong>and</strong> 9076. The PCR products <strong>of</strong><br />

the ITS region were restricted with<br />

endonuclease enzymes, DraI,<br />

HaeIII, HpaII, <strong>and</strong> PvuII. The<br />

sequences <strong>of</strong> PCR-amplified ITS<br />

regions were determined using a<br />

DNA sequencer (Applied<br />

Biosystem 373A, Perkin Elmer,<br />

Foster City, CA).<br />

AFLP analysis<br />

The amplified restriction<br />

fragment polymorphism (AFLP)<br />

technique was used to explore<br />

DNA polymorphisms in four S.<br />

nodorum isolates with different<br />

degrees <strong>of</strong> aggressiveness on<br />

wheat. The AFLP Analysis<br />

System I (Life Technologies,<br />

Gaithersburg, MD) kit was used<br />

following the protocol provided by


84<br />

Session 4 — E. Arseniuk, H.S. Tsang, J.M. Krupinsky, <strong>and</strong> P.P. Ueng<br />

the manufacturer. Cluster analysis<br />

using the unweighted pair group<br />

method with an arithmetic average<br />

(UPGMA) algorithm was<br />

performed to produce a phenogram<br />

based on the b<strong>and</strong>ing patterns<br />

produced by AFLP (Rohlf, 1990).<br />

Results <strong>and</strong> Discussion<br />

In the greenhouse seedling<br />

inoculations, isolates 9074 <strong>and</strong> 9076<br />

generally showed lower symptom<br />

severity on wheats, triticales, <strong>and</strong><br />

rye (Table 1). The results agreed<br />

with the previous work on<br />

detached wheat leaves <strong>and</strong><br />

seedlings (Krupinsky, 1997a,b). It<br />

was also shown that Polish isolate<br />

Sn48-1, originally isolated from rye,<br />

caused severe symptom severity on<br />

rye (Table 1) <strong>and</strong> less symptom<br />

severity on triticale (a cross<br />

between wheat <strong>and</strong> rye).<br />

After the isolates mated,<br />

pseudothecia with mature<br />

ascospores were formed after 50-80<br />

days <strong>of</strong> incubation. Isolates 9074<br />

<strong>and</strong> 9076 were both shown to be<br />

“+” mating types. The ITS regions<br />

<strong>of</strong> rDNA in these two isolates had<br />

the same endonuclease restriction<br />

patterns <strong>and</strong> DNA sequences as the<br />

two highly aggressive isolates 8408<br />

<strong>and</strong> 9506-2, <strong>and</strong> other wheatbiotype<br />

S. nodorum isolates (Beck<br />

<strong>and</strong> Ligon, 1995; Ueng et al., 1998).<br />

This would confirm that these two<br />

less aggressive isolates are wheatbiotype<br />

S. nodorum isolates. Of a<br />

total 390 b<strong>and</strong>s generated by 48<br />

AFLP reaction combinations, only<br />

78 b<strong>and</strong>s were shared by all four<br />

isolates in this study. The genetic<br />

diversity was high in highly<br />

aggressive isolates, <strong>and</strong> between<br />

highly <strong>and</strong> less aggressive ones<br />

(Figure 1). Cluster analysis showed<br />

great similarity between the two<br />

less aggressive isolates. The<br />

presence <strong>of</strong> differential AFLP<br />

fragments in the highly aggressive<br />

isolates may be related to<br />

aggressiveness <strong>and</strong> may be useful<br />

to identify virulent elements in the<br />

future.<br />

Acknowledgments<br />

The authors thank Yan Zhao <strong>of</strong><br />

USDA-ARS, MPPL, Beltsville, MD,<br />

for his technical support <strong>and</strong><br />

Weidong Chen <strong>of</strong> the University <strong>of</strong><br />

Illinois for his comments <strong>and</strong><br />

suggestions.<br />

References<br />

Allingham, E.A., <strong>and</strong> L.F. Jackson.<br />

1981. Variation in pathogenicity,<br />

virulence, <strong>and</strong> aggressiveness <strong>of</strong><br />

<strong>Septoria</strong> nodorum in Florida.<br />

Phytopathology 71:1080-1085.<br />

Arseniuk, E., P.C. Czembor, <strong>and</strong> B.M.<br />

Cunfer. 1997. Segregation <strong>and</strong><br />

recombination <strong>of</strong> PCR-based<br />

markers in progenies <strong>of</strong> in vitro<br />

mated isolates <strong>of</strong> Phaeosphaeria<br />

nodorum (Müller) Hedjaroude.<br />

Phytopathology 87 (Supplement)<br />

S5.<br />

Beck, J.J., <strong>and</strong> J.M. Ligon. 1995.<br />

Polymerase chain reaction assays<br />

for the detection <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum <strong>and</strong> <strong>Septoria</strong> tritici in<br />

wheat. Phytopathology 85:319-324.<br />

Table 1. Comparison <strong>of</strong> <strong>Stagonospora</strong> nodorum isolates on cereal seedlings.<br />

Wheat Triticale Rye<br />

Isolates Alba Begra Liwilla Bogo Pinokio Zduno<br />

Sn26-1 4.1+1.63 6.2+1.70 1.8+0.70b 4.3+1.24 3.1+1.55 3.0+1.97f<br />

Sn48-1 2.2+1.00 4.3+1.46 2.1+0.96b 6.6+0.94 6.4+1.72 8.3+1.03<br />

9074 1.5+0.57 3.1+1.92a 1.0+0.56c 1.7+0.99d 1.8+0.73e 2.2+2.02f, g<br />

9076 2.1+0.52 2.9+0.92a 1.0+0.53c 1.4+0.57d 1.8+0.71e 2.0+1.80g<br />

The rating scale for fungal infection is from 1 (resistant) to 9 (susceptible).With T-test at the confidence<br />

level <strong>of</strong> 0.05, the pairs with the same letters are not significantly different.<br />

Krupinsky, J.M. 1997. Aggressiveness<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum isolates<br />

obtained from wheat in the<br />

northern great plains. Plant Dis.<br />

81:1027-1031.<br />

Krupinsky, J.M. 1997. Aggressiveness<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum isolates<br />

from perennial grasses on wheat.<br />

Plant Dis. 81:1032-1036.<br />

Rohlf, F.J. 1990. Numerical Taxonomy<br />

<strong>and</strong> Multivariate Analysis System.<br />

Version 1.6. Exeter, Setauket, NY.<br />

Scharen, A.L., Z. Eyal, M.D. Huffman,<br />

<strong>and</strong> J.M. Prescott. 1985. The<br />

distribution <strong>and</strong> frequency <strong>of</strong><br />

virulence genes in geographically<br />

separated populations <strong>of</strong><br />

Leptosphaeria nodorum.<br />

Phytopathology 75:1463-168.<br />

Sprague, R. 1950. <strong>Diseases</strong> <strong>of</strong> cereals<br />

<strong>and</strong> grasses in North America.<br />

Ronald Press Co., New York.<br />

Ueng, P.P., G.C. Bergstrom, R.M. Slay,<br />

E.A. Geiger, G. Shaner, <strong>and</strong> A.L.<br />

Scharen. 1992. Restriction fragment<br />

length polymorphisms in the<br />

wheat glume blotch fungus,<br />

Phaeosphaeria nodorum.<br />

Phytopathology 82:1302-1305.<br />

Ueng, P.P., K. Subramaniam, W. Chen,<br />

E. Arseniuk, L. Wang, A.M.<br />

Cheung, G.M. H<strong>of</strong>fmann, <strong>and</strong> G.C.<br />

Bergstrom. 1998. Intraspecific<br />

genetic variation <strong>of</strong> <strong>Stagonospora</strong><br />

avenae <strong>and</strong> its differentiation from<br />

S. nodorum. Mycol. Res. 102:607-<br />

614.<br />

White, T.J., T. Bruns, S. Lee, <strong>and</strong> J.<br />

Taylor. 1990. Amplification <strong>and</strong><br />

direct sequencing <strong>of</strong> fungal<br />

ribosomal RNA genes for<br />

phylogenetics. PCR Protocols.<br />

Innis, M.A., D.H. Gelf<strong>and</strong>, J.J.<br />

Sninsky, <strong>and</strong> T.J. White (eds.). pp.<br />

315-322.<br />

Dice Similarity<br />

0.5 0.6 0.7 0.8 0.9 1.0<br />

8408<br />

9506-2<br />

9074<br />

9076<br />

Figure 1. UPGMA phenogram <strong>of</strong> four<br />

<strong>Stagonospora</strong> nodorum isolates based on the<br />

Dice Similarity Coefficients (Sd ) <strong>of</strong> 390<br />

individual AFLP DNA fragments.


A Vertically Resistant Wheat Selects for Specifically<br />

Adapted Mycosphaerella graminicola Strains<br />

C. Cowger, C.C. Mundt, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

Department <strong>of</strong> Botany <strong>and</strong> Plant Pathology, Oregon State University, Corvallis, OR, USA<br />

Abstract<br />

At its commercial release in Oregon, USA, in 1992, the winter bread wheat cultivar Gene was highly resistant to<br />

Mycosphaerella graminicola. Within just four years, that resistance had substantially disintegrated. In 1997, we collected<br />

M. graminicola isolates from experimental field plots <strong>of</strong> Gene <strong>and</strong> two other cultivars in Corvallis, Oregon, <strong>and</strong> in a<br />

greenhouse experiment inoculated seedlings <strong>of</strong> the same three cultivars with those isolates. Gene seedlings were resistant to all<br />

isolates derived from the other two cultivars, but were susceptible to six <strong>of</strong> seven isolates derived from Gene. A similar<br />

experiment using a small number <strong>of</strong> 1992 isolates from the same three cultivars had previously shown no isolates virulent on<br />

Gene, in keeping with other published data on Gene’s early resistance. Commercial cultivation <strong>of</strong> Gene, which increased to<br />

about 15% <strong>of</strong> the local wheat area during the period 1992-1995, appears to have selected for strains <strong>of</strong> M. graminicola<br />

adapted to specific virulence on Gene. Cultivar specificity is a feature <strong>of</strong> the M. graminicola-wheat pathosystem.<br />

The issue <strong>of</strong> whether M.<br />

graminicola is specific in its<br />

interactions with cultivars <strong>of</strong> a<br />

single wheat species (wheat or<br />

durum) has been widely debated.<br />

Recent studies (Ahmed et al., 1995;<br />

Ahmed, et al., 1996; Kema, et al.,<br />

1996; Kema, et al., 1997)<br />

consistently indicate the existence<br />

<strong>of</strong> interactions between cultivars <strong>of</strong><br />

a single species <strong>and</strong> isolates<br />

derived from that species.<br />

However, most experiments where<br />

interaction has been measured<br />

have utilized tester cultivars <strong>and</strong><br />

isolates not originating on them.<br />

Such experiments can reveal only<br />

the potential for cultivar-specific<br />

adaptation. Here, we report the<br />

appearance <strong>of</strong> cultivar specificity<br />

when tester cultivars were<br />

challenged with isolates selected on<br />

those same cultivars in the field.<br />

Materials <strong>and</strong> Methods<br />

We conducted two experiments<br />

involving the s<strong>of</strong>t white winter<br />

wheat cultivars Gene (P.I. 560129),<br />

Madsen (P.I. 511673), <strong>and</strong> Stephens<br />

(C.I. 017596). Monopycnidial<br />

isolates <strong>of</strong> M. graminicola were<br />

obtained from Corvallis, Oregon,<br />

field plots <strong>of</strong> these cultivars in 1992<br />

<strong>and</strong> again in 1997. Greenhouse<br />

experiments were performed with<br />

these isolates in 1994 <strong>and</strong> 1998,<br />

respectively. We used two isolates<br />

derived from each cultivar in 1992,<br />

<strong>and</strong> seven or eight isolates from<br />

each cultivar in 1997. Seedlings <strong>of</strong><br />

each cultivar were grown in pots in<br />

the greenhouse, <strong>and</strong> each tester pot<br />

was inoculated at 21 days with an<br />

individual isolate in a factorial<br />

design. Both experiments were<br />

replicated four times. The percent<br />

<strong>of</strong> leaf area covered by lesions was<br />

assessed at 21 days after<br />

inoculation. Data were log e -<br />

transformed <strong>and</strong> subjected to<br />

analysis <strong>of</strong> variance.<br />

Results<br />

Of the six 1992 isolates tested,<br />

including two from Gene, none<br />

was virulent to Gene (data not<br />

shown). By contrast, the 1997<br />

isolates derived from Gene were<br />

generally virulent to Gene, while<br />

85<br />

the 1997 isolates derived from the<br />

other two cultivars were avirulent<br />

to Gene (Table 1). Analysis <strong>of</strong><br />

variance showed no significant<br />

cultivar-<strong>of</strong>-origin x tester<br />

interaction for the 1992 isolates, <strong>and</strong><br />

a significant cultivar-<strong>of</strong>-origin x<br />

tester interaction for the 1997<br />

isolates (not shown). Linear<br />

contrasts (not shown) were used to<br />

dissect the cultivar-<strong>of</strong>-origin x tester<br />

interaction, <strong>and</strong> they revealed a<br />

highly significant interactive effect<br />

<strong>of</strong> the tester Gene vs. testers<br />

Madsen <strong>and</strong> Stephens (P = 0.0068).<br />

Discussion<br />

From its release in 1992, Gene<br />

exp<strong>and</strong>ed to occupy 15% <strong>of</strong> the<br />

Willamette Valley wheat area by<br />

1995, but then declined (Figure 1)<br />

as its resistance to M. graminicola<br />

broke down (<strong>and</strong> incidence <strong>of</strong><br />

septoria nodorum blotch, to which<br />

Gene was always susceptible,<br />

increased).


86<br />

Session 4 — C. Cowger, C.C. Mundt, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

Our first greenhouse<br />

experiment used only six isolates,<br />

but it supports published field data<br />

on Gene’s resistance during the<br />

years leading up to its release in<br />

1992 (Ahmed et al., 1995; Mundt et<br />

al., 1999). Our data suggest that<br />

cultivar specificity can develop in<br />

the M. graminicola-wheat<br />

pathosystem, <strong>and</strong> that this<br />

pathogen is capable <strong>of</strong> rapidly<br />

adapting to qualitative host<br />

resistance. Past observations<br />

suggested that resistance to M.<br />

graminicola does not break down<br />

rapidly or completely (Johnson,<br />

1992; Kema et al., 1996). However,<br />

the resistance <strong>of</strong> Gene succumbed<br />

within less than five years. We<br />

believe that Gene’s resistance is<br />

probably under monogenic or<br />

oligogenic control, <strong>and</strong> is perhaps<br />

due to the Stb4 gene (Kronstad et<br />

al., 1994). Further, M. graminicola<br />

strains able to defeat Gene’s<br />

vertical resistance appear to persist<br />

in the local population despite the<br />

limited continuing commercial<br />

production <strong>of</strong> Gene.<br />

References<br />

Ahmed, H.U., Mundt, C.C., <strong>and</strong><br />

Coakley, S.M. 1995. Host-pathogen<br />

relationship <strong>of</strong> geographically<br />

diverse isolates <strong>of</strong> <strong>Septoria</strong> tritici<br />

<strong>and</strong> wheat cultivars. Plant Pathol.<br />

44:838-847.<br />

Table 1. Percent diseased leaf area <strong>of</strong> greenhouse-grown wheat seedlings inoculated with<br />

Mycosphaerella graminicola isolates collected in 1997. a<br />

Cultivar <strong>of</strong> origin Tester cultivar<br />

& isolate nbr. Gene Madsen Stephens Mean<br />

Gene<br />

1 19.9 26.4 9.4 18.6<br />

2 3.4 18.5 16.7 12.8<br />

3 13.0 29.5 23.4 21.9<br />

4 11.2 17.9 29.6 19.5<br />

5 43.6 63.4 63.4 56.8<br />

6 21.1 32.9 33.0 29.0<br />

7 37.2 61.3 56.6 51.7<br />

Meanb 21.3ac Madsen<br />

35.7a 33.1a 30.1a<br />

8 0.9 77.1 56.0 44.7<br />

9 3.0 7.4 11.5 7.3<br />

10 4.2 30.6 28.4 21.0<br />

11 10.7 38.0 31.5 26.8<br />

12 1.0 30.1 28.7 19.9<br />

13 1.3 30.1 22.6 18.0<br />

14 0.2 19.6 7.8 9.2<br />

15 0.7 15.1 13.4 9.8<br />

Mean<br />

Stephens<br />

2.8b 31.0a 25.0a 19.6b<br />

16 3.3 22.7 24.7 16.9<br />

17 1.0 13.2 7.9 7.4<br />

18 0.4 31.0 21.7 17.7<br />

19 2.5 31.4 24.9 19.6<br />

20 2.7 22.0 16.1 13.6<br />

21 0.7 15.5 14.3 10.2<br />

22 2.2 36.1 28.3 22.2<br />

Mean 1.8b 24.6a 19.7a 15.4b<br />

Gr<strong>and</strong> mean 8.4 30.5 25.9 21.6<br />

a Experiment conducted in 1998.<br />

b Means are the untransformed percent diseased leaf area over four replications. Statistical analyses<br />

c<br />

were conducted on loge-transformed data.<br />

Within a column, means followed by the same letter are not significantly different at the 95%<br />

confidence level based on Fisher’s protected least significant difference.<br />

Ahmed, H.U., Mundt, C.C., H<strong>of</strong>fer,<br />

M.E., <strong>and</strong> Coakley, S.M. 1996.<br />

Selective influence <strong>of</strong> wheat<br />

cultivars on pathogenicity <strong>of</strong><br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici).<br />

Phytopathology 86:454-458.<br />

Johnson, R. 1992. Past, present <strong>and</strong><br />

future opportunities in breeding for<br />

disease resistance, with examples<br />

from wheat. Euphytica 63:3-22.<br />

Kema, G.H.J., Sayoud, R., Annone,<br />

J.G., <strong>and</strong> van Silfhout, C.H. 1996.<br />

Genetic variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem. II. Analysis <strong>of</strong><br />

interactions between pathogen<br />

isolates <strong>and</strong> host cultivars.<br />

Phytopathology 86:213-220.<br />

Kema, G.H.J., <strong>and</strong> van Silfhout, C.H.<br />

1997. Genetic variation for<br />

virulence <strong>and</strong> resistance in the<br />

wheat-Mycosphaerella graminicola<br />

pathosystem. III. Comparative<br />

seedling <strong>and</strong> adult plant<br />

experiments. Phytopathology<br />

87:266-272.<br />

Kronstad, W.E., Kolding, M.F., Zwer,<br />

P.K., <strong>and</strong> Karow, R.S. 1994.<br />

Registration <strong>of</strong> ‘Gene’ wheat. Crop<br />

Sci. 34:538.<br />

Mundt, C.C., H<strong>of</strong>fer, M.E., Ahmed,<br />

H.U., Coakley, S.M., DiLeone, J.A.,<br />

<strong>and</strong> Cowger, C.C. 1999. Population<br />

genetics <strong>and</strong> host resistance. Pages<br />

115-130 in: <strong>Septoria</strong> on <strong>Cereals</strong>: A<br />

Study <strong>of</strong> Pathosystems. J.A. Lucas,<br />

P. Bowyer, <strong>and</strong> A.M. Anderson, eds.<br />

CAB International, Wallingford,<br />

United Kingdom.<br />

Percent <strong>of</strong> valley wheat area<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Gene<br />

Madsen<br />

Stephens<br />

0<br />

1990 91 92 93 94 95 96 97<br />

Figure 1. Percent <strong>of</strong> s<strong>of</strong>t white winter wheat<br />

area occupied by three cultivars in the<br />

Willamette Valley <strong>of</strong> Oregon during the period<br />

1990-97.


Genetic Variability in a Collection <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum Isolates from Western Australia<br />

N.E.A. Murphy, 1 R. Loughman, 2 E.S. Lagudah, 3 R. Appels, 3 <strong>and</strong> M.G.K. Jones1 1 WA State Agriculture Biotechnology Centre, Division <strong>of</strong> Science <strong>and</strong> Engineering, Murdoch University,<br />

Murdoch, Australia<br />

2 Agriculture Western Australia, Western Australia<br />

3 Plant Industries, CSIRO, Canberra, Australia<br />

Abstract<br />

<strong>Stagonospora</strong> nodorum isolates were collected from the Western Australian cereal-belt during 1993. These isolates<br />

<strong>and</strong> a subset <strong>of</strong> isolates taken from a single location were used to assay the level <strong>of</strong> variation within the pathogen<br />

population. The isolates were compared using anonymous nuclear DNA markers. Three low copy number <strong>and</strong> a single<br />

high copy number probe were used to generate restriction fragment length polymorphisms (RFLPs). The collection<br />

exhibited a high genotypic diversity for the high copy number probe. This is consistent with the high level <strong>of</strong> sexual<br />

reproduction previously found in the fungal population. Only minor differences between the total collection <strong>of</strong> isolates<br />

<strong>and</strong> the subset <strong>of</strong> isolates taken from the single location were found.<br />

<strong>Septoria</strong> nodorum blotch is one<br />

<strong>of</strong> the most important leaf diseases<br />

<strong>of</strong> wheat (Triticum aestivum L.) in the<br />

Western Australian cereal-belt<br />

(Murray <strong>and</strong> Brown, 1987). It is<br />

caused by <strong>Stagonospora</strong> nodorum<br />

(Berk.) Castellani & E. Germano<br />

(teleomorph Phaeosphaeria nodorum<br />

(E.Muller) Hedjaroude). Previous<br />

work on the variability <strong>of</strong> S. nodorum<br />

has been conducted principally on<br />

phenotypic traits such as<br />

aggressiveness (Allingham <strong>and</strong><br />

Jackson, 1981) <strong>and</strong> culture color<br />

(Scharen <strong>and</strong> Krupinsky, 1970).<br />

These studies have shown that there<br />

is considerable variation between<br />

isolates collected within relatively<br />

small distances (Allingham et al.,<br />

1981). More recently studies have<br />

assayed the genotypic variability<br />

using molecular techniques.<br />

RFLPs were used to look at two<br />

populations <strong>of</strong> S. nodorum in the USA<br />

using a hierarchical system <strong>of</strong><br />

sampling (McDonald et al., 1994).<br />

Considerable genetic variation was<br />

found to occur on a relatively small<br />

scale. The same set <strong>of</strong> eight probes<br />

were used to study a population <strong>of</strong><br />

S. nodorum isolates gathered in<br />

Switzerl<strong>and</strong> <strong>and</strong> two populations<br />

from the USA using a similar<br />

hierarchical sampling system. The<br />

Swiss population was genetically<br />

similar to the populations from the<br />

USA, providing evidence for gene<br />

flow between the populations (Keller<br />

et al., 1997a) <strong>and</strong> the gene diversity<br />

for four <strong>of</strong> the seven RFLP loci tested<br />

was not significantly different<br />

between the populations (Keller et<br />

al., 1997a).<br />

The aim <strong>of</strong> this investigation was<br />

to examine the level <strong>of</strong> genetic<br />

diversity among a collection <strong>of</strong><br />

Western Australian S. nodorum<br />

isolates using RFLPs <strong>and</strong> to compare<br />

the results obtained with similar<br />

work done in other regions <strong>of</strong> the<br />

world.<br />

Materials <strong>and</strong> Methods<br />

A total <strong>of</strong> 118 isolates from 38<br />

locations were used in the study.<br />

Thirty-nine <strong>of</strong> the isolates were<br />

obtained from an existing collection.<br />

Of these 39 historical isolates, 26<br />

87<br />

were from a single site at<br />

Badgingarra Research Station,<br />

recognized as an area where high<br />

levels <strong>of</strong> disease severity could be<br />

expected. These isolates were<br />

captured as single ascospores in<br />

spore traps during 1991. The<br />

remaining 79 isolates were collected<br />

from 26 locations throughout the<br />

cereal belt in 1993. They were<br />

isolated as ascospores from wheat<br />

stubble remaining from the 1992<br />

crop.<br />

The cultures <strong>of</strong> S. nodorum were<br />

revived on wheat meal agar plates to<br />

confirm their identity through<br />

pycnidiospore production. Mycelia<br />

was cultured in liquid broth <strong>and</strong> the<br />

DNA extracted (Lagudah et al., 1991).<br />

The genomic DNA was digested<br />

with HindIII <strong>and</strong> run on an agarose<br />

gel. A capillary alkali method was<br />

used to transfer the DNA onto the<br />

membrane. Three low copy number<br />

probes (pASN15, pASN125 <strong>and</strong><br />

pJSN73) were used to produce a<br />

multilocus haplotype, <strong>and</strong> a single<br />

high copy number probe (pSNS4)<br />

was used to generate a fingerprint for


88<br />

each isolate. Two <strong>of</strong> the probes used<br />

were from the genomic library <strong>of</strong> a<br />

Western Australian isolate <strong>of</strong> S.<br />

nodorum (pASN15 <strong>and</strong> pASN125)<br />

<strong>and</strong> the remaining two probes were a<br />

gift from Bruce McDonald <strong>of</strong> Texas<br />

A&M University (pJSN73, pSNS4).<br />

Both genomic libraries were set up<br />

by the restriction <strong>of</strong> genomic DNA<br />

from a S. nodorum isolate using PstI.<br />

Each probe was labelled with alpha<br />

32 P. Hybridization was carried out<br />

for a minimum <strong>of</strong> 12 hours. The<br />

membranes were washed <strong>and</strong> placed<br />

on x-ray film for up to 14 days.<br />

Each probe was assumed to<br />

represent a single RFLP locus <strong>and</strong><br />

each restriction fragment length<br />

variant were treated as an allele<br />

(McDonald et al., 1994). Each allele<br />

for each RFLP locus was designated<br />

an arbitrary number. The high copy<br />

number probe (pSNS4) was used to<br />

identify clones within the same<br />

multilocus haplotype (McDonald et<br />

al., 1994). The subset <strong>of</strong> the isolates<br />

from Badgingarra was treated as a<br />

separate collection to determine if<br />

the same level <strong>of</strong> genetic diversity<br />

present in the total collection was<br />

also present at a single location. The<br />

allelic frequencies, Nei’s measure <strong>of</strong><br />

gene diversity (Nei, 1973), the<br />

genotypic diversity, <strong>and</strong> the gametic<br />

disequilibrium were all calculated as<br />

described in McDonald et al. (1994).<br />

Results<br />

There were two alleles for probes<br />

pASN125 <strong>and</strong> pJSN73 <strong>and</strong> four for<br />

probe pASN15. Only four S. nodorum<br />

isolates had the same DNA<br />

fingerprint using the high copy<br />

number probe pSNS4. Nei’s measure<br />

<strong>of</strong> gene diversity was 0.23 for<br />

pJSN73-HindIII, 0.50 for pASN15-<br />

HindIII <strong>and</strong> 0.56 for pASN15-HindIII.<br />

The average over all three loci was<br />

0.43 in the clone-corrected data. For<br />

the Badgingarra subset no clones <strong>of</strong><br />

S. nodorum were identified. The<br />

allelic frequencies were very similar<br />

for the three loci compared with the<br />

total population (being 0.27, 0.49,<br />

<strong>and</strong> 0.60, respectively) <strong>and</strong> Nei’s<br />

measure <strong>of</strong> gene diversity averaged<br />

0.45 over the three loci.<br />

For the clone-corrected data <strong>of</strong><br />

the total population, 25% <strong>of</strong> the<br />

allele pairs were in significant<br />

gametic disequilibrium (Table 1).<br />

When all <strong>of</strong> the alleles were added<br />

together for each locus pair in the<br />

clone-corrected data, one <strong>of</strong> the<br />

locus pairs was in significant<br />

gametic disequilibrium (Table 1). In<br />

the Badgingarra subset, 20% <strong>of</strong> all<br />

the allelic combinations had a<br />

significant disequilibrium<br />

coefficient, <strong>and</strong> for each locus pair,<br />

one pair were in significant gametic<br />

disequilibrium (Table 1).<br />

There were 100 different<br />

fingerprint haplotypes produced<br />

from 103 isolates, giving a genotypic<br />

diversity <strong>of</strong> 92.2, 90% <strong>of</strong> its<br />

maximum value. In the subset<br />

collection from Badgingarra, the<br />

genotypic diversity was 15, 100% <strong>of</strong><br />

its maximum value.<br />

Table 1. Gametic disequilibrium for the clonecorrected<br />

data <strong>of</strong> the total population <strong>and</strong> the<br />

Badgingarra subset for each <strong>of</strong> pair <strong>of</strong> alleles<br />

between each pair <strong>of</strong> loci <strong>and</strong> for each pair <strong>of</strong> loci.<br />

Probe pASN15 pJSN73 pASN125<br />

pASN15 4/8 0/8<br />

15.7** (3) 7.15 (3)<br />

pJSN73 0/8 1,2 1/4<br />

2.05 (3) 3.70 (1)<br />

pASN125 3/8 0/4<br />

9.62 (3) 1.71 (1)<br />

1 Clone-corrected data only; values above the<br />

diagonal are for the total population <strong>and</strong> values<br />

below the diagonal are for the Badgingarra subset.<br />

2 For each set <strong>of</strong> comparisons, the first line shows the<br />

number <strong>of</strong> allele pairs that are in significant gametic<br />

disequilibrium (p


ecombination plays in the<br />

population dynamics. This level <strong>of</strong><br />

variation may have implications for<br />

the cereal plant breeding <strong>and</strong><br />

selection program if it is a reflection<br />

<strong>of</strong> the potential variability for<br />

aggressiveness in the pathogen. If<br />

the level <strong>of</strong> genotypic variation does<br />

reflect the level <strong>of</strong> variation in<br />

aggressiveness, selecting resistance<br />

using the widest possible range <strong>of</strong> S.<br />

nodorum isolates would be desirable.<br />

The most practical method to<br />

screen breeding lines for a wide<br />

selection <strong>of</strong> S. nodorum isolates is to<br />

use straw from previously infected<br />

wheat as a source <strong>of</strong> inoculum<br />

(Holmes <strong>and</strong> Colhoun, 1975).<br />

Previous work investigating the<br />

effect <strong>of</strong> the host genotype upon the<br />

selection <strong>of</strong> S. nodorum provided no<br />

evidence <strong>of</strong> specialization <strong>of</strong> the<br />

pathogen with the host cultivar<br />

(Keller et al., 1997b) according to<br />

neutral RFLP markers. It would also<br />

be unnecessary to use straw from<br />

numerous sites, as the genotypic<br />

variation encountered at a single site<br />

can account for as much as 95% <strong>of</strong><br />

the total genotypic variation within a<br />

population (Keller et al., 1997b).<br />

Using straw from a collection <strong>of</strong><br />

cultivars <strong>and</strong> from a number <strong>of</strong> sites<br />

may be unnecessary, as it would<br />

provide only a modest increase in<br />

the overall genotypic diversity.<br />

References<br />

Allingham, E.A., <strong>and</strong> L.F. Jackson. 1981.<br />

Variation in pathogenicity, virulence<br />

<strong>and</strong> aggressiveness <strong>of</strong> <strong>Septoria</strong><br />

nodorum in Florida. Phytopathology<br />

71:1080-1085.<br />

Bathgate, J.A., <strong>and</strong> R. Loughman. 1995.<br />

Ascospores as primary inoculum <strong>of</strong><br />

Phaeosphaeria spp. in Western<br />

Australia. 10th Biennial Australasian<br />

Plant Pathology Society Conference.<br />

New Zeal<strong>and</strong>, Lincoln University,<br />

28-30th August. p. 100.<br />

Holmes, S.J.I., <strong>and</strong> J. Colhoun. 1975.<br />

Straw-borne inoculum <strong>of</strong> <strong>Septoria</strong><br />

nodorum <strong>and</strong> <strong>Septoria</strong> tritici in<br />

relation to incidence <strong>of</strong> disease on<br />

wheat plants. Plant Pathology 24:63-<br />

66.<br />

Keller, S.M., J.M. McDermott, R.E.<br />

Pettway, M.S. Wolfe, <strong>and</strong> B.A.<br />

McDonald. 1997a. Gene flow <strong>and</strong><br />

sexual reproduction in the wheat<br />

glume blotch pathogen Phaesophaeria<br />

nodorum (anamorph Stagonosopora<br />

nodorum). Phytopathology 87:353-<br />

358.<br />

89<br />

Keller, S.M., M.S. Wolfe, J.M.<br />

McDermott, <strong>and</strong> B.A. McDonald.<br />

1997b. High genetic similarity<br />

among populations <strong>of</strong> Phaeosphaeria<br />

nodorum across wheat cultivars <strong>and</strong><br />

regions in Switzerl<strong>and</strong>.<br />

Phytopathology 87:1134-1139.<br />

Lagudah, E.S., R. Appels, <strong>and</strong> D.<br />

McNeil. 1991. The Nor-D3 locus <strong>of</strong><br />

Triticum tauschii: natural variation<br />

<strong>and</strong> genetic linkage to markers in<br />

chromosome 5. Genome 34:387-395.<br />

McDonald, B.A., J. Miles, L.R. Nelson,<br />

<strong>and</strong> R.E. Pettway. 1994. Genetic<br />

variability in nuclear DNA in field<br />

populations <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum. Phytopathology 84:250-<br />

255.<br />

Murray, G.M., <strong>and</strong> J.F. Brown. 1987.<br />

The incidence <strong>and</strong> relative<br />

importance <strong>of</strong> wheat diseases in<br />

Australia. Australasian Plant<br />

Pathology 16:34-37.<br />

Nei, M. 1973. Analysis <strong>of</strong> gene diversity<br />

in subdivided populations.<br />

Proceedings <strong>of</strong> the National<br />

Academy <strong>of</strong> Science, USA 70:3321-<br />

3323.<br />

Scharen, A.L., <strong>and</strong> J.M. Krupinsky.<br />

1970. Cultural <strong>and</strong> Inoculation<br />

studies <strong>of</strong> <strong>Septoria</strong> nodorum cause <strong>of</strong><br />

glume blotch <strong>of</strong> wheat.<br />

Phytopathology 60:1480-1485.


90<br />

Mating Type-Specific PCR Primers for <strong>Stagonospora</strong><br />

nodorum Field Studies<br />

Bennett, R.S., 1 S.-H. Yun, 1 T.Y. Lee, 1 B.G. Turgeon, 1 B. Cunfer, 2<br />

E. Arseniuk, 3 <strong>and</strong> G.C. Bergstrom 1 * (Poster)<br />

1 Department <strong>of</strong> Plant Pathology, Cornell University, Ithaca, NY, USA<br />

2 Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Georgia, Griffin, GA, USA<br />

3 Plant Breeding <strong>and</strong> Acclimatization Institute, Radzików, Pol<strong>and</strong><br />

* corresponding author<br />

Abstract<br />

Conserved regions <strong>of</strong> the mating type genes (MAT) in Cochliobolus heterostrophus, Mycosphaerella zeaemaydis,<br />

<strong>and</strong> Alternaria alternata were used to design primers to identify the mating type genes <strong>and</strong> thus assign<br />

mating types to <strong>Stagonospora</strong> nodorum. These primers successfully distinguished between mating types <strong>of</strong> S.<br />

nodorum isolates that had been identified through traditional crosses. The primers were used to screen a small sample <strong>of</strong><br />

S. nodorum isolates from New York, thus revealing the presence <strong>of</strong> both mating types. This efficient method <strong>of</strong><br />

identifying mating types may help determine the role <strong>of</strong> sexual recombination in the epidemiology <strong>of</strong> S. nodorum, as<br />

well as elucidate phylogenetic relationships among related species.<br />

<strong>Stagonospora</strong> nodorum (Berk.)<br />

Castellani & E.G. Germano<br />

(teleomorph = Phaeosphaeria<br />

nodorum (E. Müller) Hedjaroude) is<br />

the causal agent <strong>of</strong> stagonospora<br />

nodorum blotch, a major disease <strong>of</strong><br />

wheat around the world.<br />

Phaeosphaeria nodorum is<br />

heterothallic <strong>and</strong> thus requires<br />

strains <strong>of</strong> the two different mating<br />

types to produce ascospores<br />

(Müller, 1989). The potential<br />

significance <strong>of</strong> sexual reproduction<br />

in the epidemiology <strong>of</strong><br />

stagonospora nodorum blotch has<br />

been heightened by the high RFLP<br />

variability found among field<br />

isolates (McDonald et al., 1994).<br />

Fungi that regularly reproduce<br />

sexually have more opportunities<br />

to develop fungicide resistance <strong>and</strong><br />

to overcome cultivar resistance.<br />

Furthermore, airborne ascospores<br />

would permit dissemination over<br />

longer distances than splashdispersed<br />

conidia.<br />

MAT genes have been identified<br />

for several species <strong>of</strong><br />

Loculoascomycetes including<br />

Setosphaeria turcica, Mycosphaerella<br />

zeae-maydis, Alternaria alternata, <strong>and</strong><br />

several Cochliobolus spp. (Arie et al.,<br />

1997). Both genes, MAT-1 <strong>and</strong> MAT-<br />

2, contain conserved regions that<br />

allow the relatively rapid<br />

identification <strong>of</strong> these genes from an<br />

increasing group <strong>of</strong> ascomycetes. A<br />

high mobility group (HMG) <strong>and</strong> abox<br />

motif are the conserved regions<br />

in MAT-2 <strong>and</strong> MAT-1, respectively<br />

(Turgeon, 1998). These sequences,<br />

besides elucidating sexual<br />

mechanisms <strong>and</strong> phylogeny, may<br />

also help address epidemiological<br />

questions.<br />

We have developed primers to<br />

identify the mating types <strong>of</strong> P.<br />

nodorum from the conserved MAT<br />

regions <strong>of</strong> related fungi. Preliminary<br />

data on the mating type prevalence<br />

in a small sample <strong>of</strong> New York field<br />

isolates is also presented.<br />

Materials <strong>and</strong> Methods<br />

Fungal isolates <strong>and</strong> media<br />

Three S. nodorum isolates <strong>of</strong><br />

different mating types (two (-)<br />

(SN435PL-98, SN437GA-98) <strong>and</strong><br />

one (+) (SN436GA-98)) (Arseniuk<br />

et al., 1997a,b) were grown on V-8<br />

juice agar (200 ml V8 juice, 3 g<br />

CaCO3 , 15 g agar per 800 ml <strong>of</strong><br />

distilled water). After<br />

approximately seven days, mycelial<br />

plugs were taken <strong>and</strong> placed into<br />

yeast-malt-sucrose broth (5 g yeast<br />

extract, 5 g malt extract, 20 g<br />

sucrose per 1 liter <strong>of</strong> distilled<br />

water), <strong>and</strong> placed on a shaker (150<br />

rpm) at room temperature. Mycelia<br />

were harvested after 5-7 days by<br />

filtering through 3-ply sterile<br />

cheesecloth lining a Buchner funnel<br />

<strong>and</strong> were rinsed with distilled<br />

water. The samples were then<br />

frozen <strong>and</strong> lyophilized overnight.<br />

MAT-1 <strong>and</strong> MAT-2 C. heterostrophus<br />

laboratory strains (C5 <strong>and</strong> C4,<br />

respectively) were used as controls.


Thirty-eight arbitrarily chosen<br />

field isolates <strong>of</strong> S. nodorum collected<br />

from New York in different places<br />

<strong>and</strong> years were grown on<br />

cellophane discs overlaid on V-8<br />

juice agar. After approximately<br />

seven days, mycelia were scraped<br />

<strong>of</strong>f the cellophane <strong>and</strong> placed in<br />

microcentrifuge tubes to be<br />

lyophilized.<br />

DNA extraction<br />

Lyophilized mycelia were<br />

ground in liquid nitrogen. The field<br />

isolates were ground directly in<br />

microcentrifuge tubes with a small<br />

amount <strong>of</strong> white quartz s<strong>and</strong> (-50 +<br />

70 mesh, Sigma Chemical<br />

Company, St. Louis, MO). DNA<br />

was extracted by a miniprep<br />

procedure (Wirsel et al., 1996).<br />

Amplification <strong>of</strong> the HMG<br />

<strong>and</strong> a-boxes by PCR<br />

PCR primers (Sn-HMG1 <strong>and</strong><br />

Sn-HMG2) for the HMG box (Table<br />

1) were designed from conserved<br />

regions in the HMG box <strong>of</strong> the MAT<br />

gene <strong>of</strong> Cochliobolus heterostrophus,<br />

M. zeae-maydis, <strong>and</strong> A. alternata,<br />

whereas the primers (Sn-ab1 <strong>and</strong><br />

Sn-ab2) for the a-box were designed<br />

from the a-boxes <strong>of</strong> M. zeae-maydis<br />

<strong>and</strong> A. alternata. The primers were<br />

synthesized by the Cornell<br />

University DNA Services Facility<br />

<strong>and</strong> were dissolved (100 µM) in<br />

Table 1. PCR primers used to amplify<br />

fragments <strong>of</strong> MAT genes in <strong>Stagonospora</strong><br />

nodorum.<br />

MAT-1<br />

Sn-ab1 5’ AA(A/G)GCN(C/T)TNAA(C/<br />

T)GCNTT (C/T)GTNGG 3’<br />

Sn-ab2 5’ TC(C/T)TTNCC(A/G/T)AT(C/T)<br />

TG(A/G)TCNCG(A/G/T)AT 3’<br />

MAT-2<br />

Sn-HMG1 5’ AA(A/G)GCNCCN(AC)<br />

GNCCNATGAA 3’<br />

Sn-HMG2 5’ TT(C/T)TT(C/T)TT(C/T)T(CG)<br />

NCCNGG(C/T)TT 3’<br />

sterile distilled water <strong>and</strong> stored at<br />

–20C. Each PCR reaction mixture<br />

had approximately 20 ng <strong>of</strong><br />

genomic DNA in 50 µl reaction<br />

buffer [1X PCR Buffer (Perkin<br />

Elmer, Norwalk, CT), 0.2 mM<br />

dNTPs, 2.5 mM MgCl2, , 2 µM each<br />

primer, <strong>and</strong> 0.025 U Taq<br />

polymerase]. The samples were<br />

denatured at 95C for 2 min, <strong>and</strong><br />

then subjected to 30 cycles <strong>of</strong> 95C<br />

for 1 min, 50C for 30 sec, <strong>and</strong> 72C<br />

for 1.5 min. After extension at 72C<br />

for 10 min, the samples were kept<br />

at 4C. 5 µl aliquots <strong>of</strong> the PCR<br />

products were analyzed on a 2%<br />

agarose gel in TAE buffer.<br />

Cloning, sequencing, <strong>and</strong><br />

analysis <strong>of</strong> PCR products<br />

The MAT sequences from<br />

related fungi predicted that a PCR<br />

product <strong>of</strong> ~270 bp for the HMG<br />

box <strong>and</strong> a product <strong>of</strong> ~230 bp for<br />

the a-boxes would result if the<br />

amplifications were successful.<br />

Products <strong>of</strong> these sizes were cloned<br />

from the PCR reaction into the<br />

vector pCR2.1, using the guidelines<br />

Mating Type-Specific PCR Primers for <strong>Stagonospora</strong> nodorum Field Studies 91<br />

<strong>of</strong> the manufacturer (Invitrogen<br />

Co., San Diego, CA). DNA<br />

sequences were determined at the<br />

Cornell University DNA Services<br />

Facility. Sequences were aligned<br />

with the LaserGene program<br />

MegAlign (DNASTAR Inc.,<br />

Madison, WI), using the clustal<br />

method. A BLAST search was done<br />

using the NCBI/Genbank internet<br />

databases.<br />

Gel blot hybridization<br />

A st<strong>and</strong>ard DNA gel blot<br />

hybridization procedure was<br />

followed (Sambrook et al., 1989).<br />

Results<br />

HMG alpha<br />

The primers for the HMG box<br />

were used with genomic DNA <strong>of</strong><br />

(+) <strong>and</strong> (-) S. nodorum strains as<br />

templates. A PCR product<br />

approximately 0.3 kb, matching the<br />

~270 bp predicted, was present in<br />

the (+) (MAT-2) strain, but was not<br />

present in the (-) (MAT-1) strains<br />

(Figure 1). The primers for the abox,<br />

however, amplified the<br />

1 2 3 4 5 6 7 8 9 101112131415161718<br />

Figure 1. PCR products amplified with the HMG <strong>and</strong> a-box primers. Lanes 2-9 were amplified using<br />

HMG (MAT-2) primers, <strong>and</strong> lanes 10-17 were amplified with a-box (MAT-1) primers. Lanes 2,10 <strong>and</strong><br />

3,11: C. heterostrophus C5 (MAT-1) <strong>and</strong> C. heterostrophus C4 (MAT-2), respectively; lanes 4,12 <strong>and</strong><br />

5,13: plasmid with cloned a-box from SN435PL-98 <strong>and</strong> plasmid with cloned HMG box from SN436GA-<br />

98, respectively; lanes 6 <strong>and</strong> 14: SN436GA-98; lanes 7 <strong>and</strong> 15: SN437GA-98; lanes 8 <strong>and</strong> 16: SN005NY-<br />

85; lanes 9 <strong>and</strong> 17: SN197NY-88. A ~0.3-kb b<strong>and</strong> (arrowhead) was amplified in the MAT-2 isolates<br />

only, <strong>and</strong> a b<strong>and</strong> above 0.2-kb (asterisk) was amplified only in the MAT-1 isolates.


92<br />

Session 4 — Bennett, R.S., S.-H. Yun, T.Y. Lee, B.G. Turgeon, B. Cunfer, E. Arseniuk, <strong>and</strong> G.C. Bergstrom<br />

expected product (~230 bp) from<br />

only the (-) (MAT-1) strains. The<br />

primers were also able to<br />

distinguish between the MAT-1 <strong>and</strong><br />

MAT-2 C. heterostrophus controls.<br />

The BLAST search <strong>of</strong> the S.<br />

nodorum cloned products showed<br />

highest homology to the MAT<br />

genes <strong>of</strong> Cochliobolus spp.<br />

In gel blot hybridization, a<br />

single b<strong>and</strong> was obtained from<br />

DNA <strong>of</strong> the (+) SN436GA-98 isolate<br />

only when probed with the HMG<br />

box; no hybridization occurred on<br />

the (-) SN435PL-98 isolate.<br />

Likewise, a single b<strong>and</strong> was<br />

obtained when SN435PL-98 was<br />

probed with the a-box, <strong>and</strong> no<br />

hybridization occurred to the DNA<br />

<strong>of</strong> SN436GA-98.<br />

When these primers were used<br />

to test the 38 field isolates, 20 were<br />

shown to be <strong>of</strong> the MAT-1 mating<br />

type <strong>and</strong> 18 were <strong>of</strong> the MAT-2<br />

mating type.<br />

Discussion<br />

A PCR technique that readily<br />

identifies the mating type can be a<br />

useful tool for studying S. nodorum.<br />

Surveying for the distribution <strong>of</strong><br />

mating types in the field may result<br />

in a better underst<strong>and</strong>ing <strong>of</strong> the<br />

role <strong>of</strong> sexual reproduction in the<br />

disease cycle. The preliminary data<br />

given here tested S. nodorum<br />

isolates that were collected in New<br />

York wheat fields between 1985<br />

<strong>and</strong> 1995. These isolates are not<br />

only temporally r<strong>and</strong>om, but<br />

spatially unrelated. More<br />

informative surveys <strong>of</strong> spatially<br />

distributed isolates from single<br />

fields are being conducted.<br />

However, the preliminary data<br />

presented here do indicate that<br />

both mating types exist in New<br />

York <strong>and</strong> that sexual recombination<br />

is possible.<br />

Mating types may also serve as<br />

an additional marker for<br />

experiments requiring genetically<br />

characterized isolates. As<br />

previously established (Arie et<br />

al,.1997), we refer to (-) isolates<br />

containing the a-box as MAT-1, <strong>and</strong><br />

(+) isolates containing the HMG<br />

box as MAT-2. Finally, as outlined<br />

in a review by Turgeon (1998), MAT<br />

gene sequences may provide<br />

valuable information for<br />

phylogenetic analyses, particularly<br />

<strong>of</strong> related species, the study <strong>of</strong><br />

pathogenesis, <strong>and</strong> the underlying<br />

mechanisms <strong>of</strong> reproductive life<br />

styles in fungi.<br />

References<br />

Arie, T., S.K. Christiansen, O.C. Yoder,<br />

<strong>and</strong> B.G. Turgeon. 1997. Efficient<br />

cloning <strong>of</strong> ascomycete mating type<br />

genes by PCR amplification <strong>of</strong> the<br />

conserved MAT HMG box. Fungal<br />

Genetics <strong>and</strong> Biology 21:118-130.<br />

Arseniuk, E., B.M. Cunfer, S. Mitchell,<br />

<strong>and</strong> S. Kresovich. 1997a.<br />

Characterization <strong>of</strong> genetic<br />

similarities among isolates <strong>of</strong><br />

<strong>Stagonospora</strong> spp. <strong>and</strong> <strong>Septoria</strong> tritici<br />

by amplified fragment length<br />

polymorphism analysis.<br />

Phytopathology 87 (Supplement) S5.<br />

Arseniuk, E., P.C. Czembor, <strong>and</strong> B.M.<br />

Cunfer. 1997b. Segregation <strong>and</strong><br />

recombination <strong>of</strong> PCR-based<br />

markers inprogenies <strong>of</strong> in vitro<br />

mated isolates <strong>of</strong> Phaeosphaeria<br />

nodorum (Müller) Hedjaroude.<br />

Phytopathology 87(Supplement), S5<br />

.<br />

McDonald, B.A., J. Miles, L.R. Nelson,<br />

<strong>and</strong> R.E. Pettway. 1994. Genetic<br />

variability in nuclear DNA in field<br />

populations <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum. Phytopathology 84:250-<br />

255.<br />

Müller, E. 1989. On the taxonomic<br />

position <strong>of</strong> <strong>Septoria</strong> nodorum <strong>and</strong><br />

<strong>Septoria</strong> tritici, p. 11-12. In:<br />

Proceedings from the Third<br />

International Workshop on <strong>Septoria</strong><br />

<strong>Diseases</strong> <strong>of</strong> <strong>Cereals</strong>. Zurich,<br />

Switzerl<strong>and</strong>.<br />

Sambrook, J., E.F. Fritsch, <strong>and</strong> T.<br />

Maniatis. 1989. Molecular Cloning:<br />

A Laboratory Manual, 2 nd ed. Cold<br />

Spring Harbor Laboratory Press:<br />

Cold Spring Harbor, NY.<br />

Turgeon, B.G. 1998. Application <strong>of</strong><br />

mating type gene technology to<br />

problems in fungal biology. Annual<br />

Review <strong>of</strong> Phytopathology 36:115-<br />

137.<br />

Wirsel, S., B.G. Turgeon, <strong>and</strong> O.C.<br />

Yoder. 1996. Deletion <strong>of</strong> the<br />

Cochliobolus heterostrophus mating<br />

type (MAT) locus promotes function<br />

<strong>of</strong> MAT transgenes. Current<br />

Genetics 29:241-249.


Session 5: Epidemiology<br />

Epidemiology <strong>of</strong> Mycosphaerella graminicola <strong>and</strong><br />

Phaeosphaeria nodorum: An Overview<br />

M.W. Shaw<br />

Department <strong>of</strong> Agricultural Botany, School <strong>of</strong> Plant Sciences, The University <strong>of</strong> Reading, Reading, Engl<strong>and</strong><br />

Abstract<br />

The within-season <strong>and</strong> between-crop methods <strong>of</strong> multiplication <strong>and</strong> survival, <strong>and</strong> their environmental relations are<br />

reviewed. Mycosphaerella graminicola multiplies within a season by conidia which are primarily but not exclusively<br />

dispersed by rain. Arguments are given that the influence <strong>of</strong> ascospores within a crop will be minor, but they are the<br />

major source <strong>of</strong> movement <strong>of</strong> the pathogen into new crops. Phaeosphaeria nodorum also multiplies within a season by<br />

conidia, but has clearer associations with wet weather. The role <strong>of</strong> ascospores in movement between crops may vary<br />

geographically; seed transmission seems to be very important in some areas.<br />

In this contribution, I have tried<br />

to summarize what is understood<br />

<strong>of</strong> the epidemiology <strong>of</strong> these two<br />

diseases. My aim has been to<br />

produce a concise summary to<br />

introduce the detailed <strong>and</strong> novel<br />

contributions that follow. I have<br />

tried to give slightly more extended<br />

discussion <strong>of</strong> those areas where<br />

new ideas have arisen or our<br />

underst<strong>and</strong>ing has changed<br />

substantially in the last few years.<br />

The relevant questions for both<br />

diseases fall into two classes. First,<br />

qualitative: what conditions allow<br />

inoculum transfer, permit infection,<br />

<strong>and</strong> encourage sporulation?<br />

Second: quantitative: in a given<br />

agro-ecosystem, what factors in<br />

practice regulate pathogen<br />

population size? The two questions<br />

are related, but both need to be<br />

answered if the diseases are to be<br />

managed most effectively. The<br />

answers also depend greatly on<br />

scale: within a region <strong>and</strong> over<br />

several years, very different<br />

processes <strong>and</strong> factors may need to<br />

be considered from those operating<br />

within a crop <strong>and</strong> within a season.<br />

The paper is restricted to wheat.<br />

Mycosphaerella<br />

graminicola<br />

Within a crop<br />

As discussed later, infection <strong>of</strong> a<br />

crop is usually initiated by airborne<br />

ascospores (Shaw <strong>and</strong> Royle, 1989).<br />

The density <strong>of</strong> initial infections is<br />

such that once a few sporulating<br />

lesions per square meter exist, a<br />

polycyclic epidemic on successive<br />

leaf layers follows (Shaw <strong>and</strong><br />

Royle, 1993). Pycnidia are produced<br />

within roughly 14 to 40 days,<br />

depending on both temperature<br />

<strong>and</strong> host cultivar. Conidia may be<br />

dispersed by single rain splashes<br />

within a circle <strong>of</strong> about 1 m radius,<br />

the number dispersed decreasing<br />

exponentially with distance, with<br />

half distances <strong>of</strong> the order <strong>of</strong> 10 cm<br />

(Bannon <strong>and</strong> Cooke, 1998; Brennan<br />

et al., 1985a; Brennan et al., 1985b).<br />

Initial dispersal is to the ground or<br />

to surface water on a leaf, whence<br />

further dispersal is possible.<br />

However, spores contacting leaf<br />

surfaces are bound to the surface<br />

within a short time. The average<br />

number <strong>of</strong> splashes moving a spore<br />

during rain <strong>of</strong> given intensity <strong>and</strong><br />

duration is hard to estimate, but is<br />

93<br />

unlikely to be large, so half<br />

distances for effective horizontal<br />

dispersal will be <strong>of</strong> the order <strong>of</strong> 20-<br />

50 cm at most. Each initial infection<br />

may plausibly produce 50,000 to<br />

500,000 conidia (10-100 pycnidia <strong>of</strong><br />

ca. 5000 spores) (Eyal, 1971).<br />

Although most <strong>of</strong> these are not<br />

dispersed far, considering them as<br />

evenly dispersed over a circle <strong>of</strong> 0.5<br />

m radius gives 5-50 spores per<br />

square centimeter from initial<br />

infections spaced at about 1/m 2 . If<br />

they had 2-20% infection efficiency,<br />

the crop would be saturated with<br />

latent lesions. Fortunately, infection<br />

efficiency is usually lower than this,<br />

but if few spores are present, 1% <strong>of</strong><br />

those applied may cause infection<br />

under good infection conditions.<br />

The actual environmental<br />

conditions permitting infection are<br />

lax because the pathogen tolerates<br />

extended breaks in humidity<br />

during the infection process (Shaw,<br />

1991a; Shaw <strong>and</strong> Royle, 1993).<br />

Certainly, within two infection<br />

cycles the pathogen population in<br />

crops with moderate initial<br />

amounts <strong>of</strong> disease will be limited<br />

by the rate <strong>of</strong> growth <strong>of</strong> leaf area


94<br />

Session 5 — M.W. Shaw<br />

<strong>and</strong> the favorability <strong>of</strong> the<br />

environment rather than by<br />

shortage <strong>of</strong> inoculum. However,<br />

spread by splash is very inefficient<br />

over other than short distances, <strong>and</strong><br />

if initial infections are widely<br />

spaced, then the disease should<br />

remain patchy <strong>and</strong> therefore cause<br />

limited damage.<br />

The preceding paragraph<br />

ignored vertical spread, which has<br />

been argued to be the key factor<br />

determining severe disease on the<br />

uppermost leaf layers <strong>of</strong> a wheat<br />

crop. Vertical movement <strong>of</strong> splash<br />

droplets is limited, declining<br />

exponentially with typical halfdistances<br />

<strong>of</strong> 5 cm, but varies<br />

between storms (Bannon <strong>and</strong><br />

Cooke, 1998; Shaw, 1987; Shaw,<br />

1991b). Probably more critically, the<br />

distance between leaf layers with<br />

<strong>and</strong> without infection varies greatly<br />

according to both the architecture<br />

<strong>of</strong> the wheat cultivar <strong>and</strong> the latent<br />

period <strong>of</strong> the pathogen on the<br />

cultivar. Lovell et al. (1997) have<br />

shown how the interaction <strong>of</strong> these<br />

causes great variation in the<br />

potential for spread <strong>of</strong> pathogen to<br />

the upper part <strong>of</strong> the crop, where it<br />

becomes damaging.<br />

Once a few sporulating lesions<br />

are present on a reasonable<br />

proportion <strong>of</strong> the uppermost leaves<br />

<strong>of</strong> a wheat crop, there is potential<br />

for multiplication leading to<br />

widespread infection <strong>and</strong><br />

premature leaf death, except in very<br />

dry weather.<br />

The major advance since the last<br />

Workshop is the recognition that<br />

pseudothecia are produced<br />

regularly throughout the year<br />

under at least some conditions on<br />

at least some cultivars (Hunter et<br />

al., 1999; Kema et al., 1996). This<br />

means that sparse initial infections<br />

could spread much more effectively<br />

by windblown ascospores, <strong>and</strong> that<br />

multiplication <strong>of</strong> the disease could<br />

take place efficiently in the absence<br />

<strong>of</strong> rain, dew alone providing<br />

wetness for infection by dry<br />

dispersed spores. This is a<br />

potentially dramatic change to our<br />

view <strong>of</strong> the ecology <strong>of</strong> the organism.<br />

However, the effect on epidemics<br />

within a season is probably limited,<br />

because the pseudothecia are<br />

always produced long after<br />

pycnidia. Two extreme situations<br />

can be imagined: one in which<br />

pycnidia can infect <strong>of</strong>ten during the<br />

season, one in which they can never<br />

infect.<br />

In the first situation, even in the<br />

fastest quoted latent periods for<br />

perithecia, the effect on the<br />

epidemic rate <strong>of</strong> ascospore<br />

production is equivalent to perhaps<br />

1000-10,000 extra spores from a<br />

single infection, produced at<br />

roughly the time when the next<br />

generation <strong>of</strong> lesions from the<br />

pycnidia produced earlier are<br />

themselves producing pycnidia. If<br />

the multiplication ratio <strong>of</strong> lesions<br />

from one generation to the next is R,<br />

then the pycnidial route is<br />

producing <strong>of</strong> the order <strong>of</strong> R * 10,000<br />

conidia in the time during which<br />

the sexual route is producing 10,000<br />

spores: a ratio <strong>of</strong> 1:R. Order <strong>of</strong><br />

magnitude estimates <strong>of</strong> R can be<br />

obtained in at least two ways. First,<br />

the ratio between the number <strong>of</strong><br />

lesions a latent period after disease<br />

first entered a crop <strong>and</strong> the initial<br />

number <strong>of</strong> lesions. For the crops<br />

observed by Shaw <strong>and</strong> Royle (1989)<br />

this ratio was about 100. Second, the<br />

ratio between the initial number <strong>of</strong><br />

lesions on the uppermost leaves <strong>of</strong> a<br />

crop, <strong>and</strong> the number one latent<br />

period later (Shaw <strong>and</strong> Royle 1993):<br />

this too suggests a large number.<br />

The effect on the epidemic rate<br />

must be very small. This is borne<br />

out by detailed modelling (Eriksen<br />

et al., in prep.).<br />

In the second situation,<br />

pseudothecia are the only route by<br />

which infection can progress. In<br />

this case, the latent period is at least<br />

double the latent period for an<br />

epidemic otherwise similar but<br />

driven by pycnidia. Since the latent<br />

period is at least doubled, <strong>and</strong> any<br />

gains in infection or dispersal<br />

efficiency are <strong>of</strong>fset by the reduced<br />

number <strong>of</strong> ascospores produced,<br />

the rate <strong>of</strong> the epidemic is at least<br />

halved. A rather more qualitative<br />

prediction would be that where<br />

conditions prevent conidial<br />

dispersal, epidemics will develop<br />

very slowly, <strong>and</strong> be damaging only<br />

where initial disease is rather<br />

widespread. It will clearly be very<br />

interesting to study epidemic<br />

progress in areas where pycnidial<br />

progression is not possible, <strong>and</strong> see<br />

whether this prediction is true.<br />

In fact, even this may overstate<br />

the case, since the fungus is<br />

heterothallic (Kema et al., 1996) <strong>and</strong><br />

presumably therefore requires two<br />

infections <strong>of</strong> opposite mating type<br />

to be adjacent in order to produce<br />

ascospores. It is not known how<br />

close together infections must be to<br />

mate effectively. Metcalfe (1999)<br />

found internal hyphae spanning an<br />

average maximum <strong>of</strong> 11 mm in a<br />

susceptible wheat cultivar just<br />

before pycnidia were produced, so<br />

infections more than 20 mm apart<br />

may be isolated. In a sparse, slow<br />

epidemic this may be rare. Mating<br />

will be commonest in situations<br />

where inoculum is not limiting the<br />

pathogen population in any way,<br />

but in these cases, the extra<br />

inoculum provided by ascospores<br />

will not affect epidemic rate.


When inoculum is common low<br />

in the canopy, but transport is<br />

potentially limiting infection <strong>of</strong><br />

upper leaves, ascospores could<br />

pr<strong>of</strong>oundly alter the risk <strong>of</strong> severe<br />

infection on the uppermost leaves,<br />

because rainfall is not required for<br />

their movement to these leaves.<br />

However, recent work confirms<br />

(under UK conditions at least) the<br />

apparent association <strong>of</strong> severe M.<br />

graminicola with rain (Lovell et al.,<br />

1997); <strong>and</strong> windblown transport <strong>of</strong><br />

spores from close to the ground to<br />

the upper leaves <strong>of</strong> a dense canopy<br />

is not efficient because windspeeds<br />

are not high within the canopy.<br />

Nonetheless, the simple picture <strong>of</strong><br />

rain moving spores from the base <strong>of</strong><br />

the canopy onto the upper leaves as<br />

the only risk factor requires<br />

modification <strong>and</strong> may well be<br />

misleading under some<br />

circumstances.<br />

Interactions between genetics<br />

<strong>and</strong> ecology are <strong>of</strong> interest, <strong>and</strong><br />

could complicate the population<br />

dynamics <strong>of</strong> the pathogen<br />

considerably, at worst rendering<br />

forecasts almost impossible. The<br />

fittest isolate on a given cultivar can<br />

vary with temperature (B. al-<br />

Hamar, University <strong>of</strong> Reading,<br />

personal communication). This<br />

means that equivalent sized<br />

populations could have<br />

intrinsically different growth-rates<br />

according to the combination <strong>of</strong><br />

weather conditions which gave rise<br />

to them, so that identical<br />

environmental conditions <strong>and</strong><br />

initial population sizes would give<br />

rise to different population<br />

responses.<br />

Between crops<br />

Windblown ascospores appear<br />

to be the main mechanism by<br />

which M. graminicola moves<br />

between crops, both from one<br />

Epidemiology <strong>of</strong> Mycosphaerella graminicola <strong>and</strong> Phaeosphaeria nodorum: An Overview 95<br />

season to the next <strong>and</strong> within a<br />

season. However, there is little<br />

documentation <strong>of</strong> the relative<br />

importance <strong>of</strong> local debris <strong>and</strong><br />

widely dispersed, generalized<br />

sources <strong>of</strong> inoculum for new crops.<br />

In northern Europe, with most<br />

wheat grown in rotation <strong>and</strong> with<br />

full tillage, the generalized air<br />

spora appear to dominate sources.<br />

However, this situation must differ<br />

in different farming systems. Where<br />

wheat is a scarcer crop, the<br />

generalized concentration <strong>of</strong><br />

ascospores in the air must be lower,<br />

<strong>and</strong> where no-till <strong>and</strong> continuous<br />

cropping is common but the wheat<br />

fields are scattered, one would<br />

expect local sources to dominate.<br />

There is relatively little evidence<br />

published about this outside<br />

Europe <strong>and</strong> northern America. A<br />

curious feature <strong>of</strong> the disease in, for<br />

example, the UK seems to be that<br />

the extremely efficient control <strong>of</strong><br />

the pathogen achieved on the<br />

upper part <strong>of</strong> the plant by fungicide<br />

does not appear to have reduced<br />

the ability <strong>of</strong> the fungus to produce<br />

ascospores <strong>and</strong> infect subsequent<br />

crops. This could have two<br />

explanations: 1) in stubbles,<br />

ascospores are produced as much<br />

on the lower parts <strong>of</strong> plants, never<br />

managed by fungicide, as on the<br />

upper parts; or 2) the dominant<br />

triazole fungicides act only as<br />

fungistats <strong>and</strong> after the death <strong>of</strong><br />

leaves, fungal development <strong>and</strong> the<br />

production <strong>of</strong> pseudothecia can<br />

continue as if the treatment had<br />

never taken place. Genetic <strong>and</strong><br />

physiological evidence suggests<br />

that alternate hosts are unimportant<br />

in perpetuating the disease on<br />

wheat.<br />

As noted in a previous review<br />

(Shaw, 1999), the proportion <strong>of</strong> l<strong>and</strong><br />

devoted to wheat in a region could<br />

affect the importance <strong>of</strong> M.<br />

graminicola as a pathogen. As the<br />

proportion increases, so the<br />

distance between last year’s fields<br />

<strong>and</strong> this year’s will decline. Even<br />

by wind, the number <strong>of</strong> spores<br />

moving a given distance declines<br />

rapidly with distance, so moderate<br />

changes in average distance<br />

between fields could cause<br />

substantial changes in the density<br />

<strong>of</strong> initial infections. In turn, at low<br />

densities, this could affect the<br />

importance <strong>of</strong> the disease at the end<br />

<strong>of</strong> the cropping season. The detail<br />

depends on how cropl<strong>and</strong> is<br />

organized <strong>and</strong> how new<br />

wheatl<strong>and</strong>s tend to be disposed<br />

relative to old, as well as on the<br />

strength <strong>of</strong> linkages between initial<br />

<strong>and</strong> end-<strong>of</strong>-season disease.<br />

At the Long Ashton conference<br />

in 1997, a crucial question in<br />

underst<strong>and</strong>ing the uniformity in<br />

genetic composition <strong>of</strong> M.<br />

graminicola worldwide was whether<br />

seed-borne transport could ever<br />

occur. This apparently simple<br />

question remains open, <strong>and</strong> is far<br />

from simple, since very, very rare<br />

events could provide enough<br />

migration to leave populations<br />

genetically linked. Suppose a<br />

million seeds each from a diseased<br />

crop <strong>and</strong> a fungicide treated crop<br />

are grown, <strong>and</strong> a lesion appears on<br />

one plant from the diseased crop.<br />

Can we confidently say that lesion<br />

arose from seed transmission or<br />

could it have been attributable to a<br />

failure <strong>of</strong> isolation <strong>and</strong> a stray<br />

ascospore? Some form <strong>of</strong> gene<br />

exchange between widely<br />

dispersed populations apparently<br />

occurs, so the geneticists tell us, but<br />

it may be almost impossible to<br />

distinguish the various categories<br />

<strong>of</strong> very rare events giving rise to<br />

this.


96<br />

Session 5 — M.W. Shaw<br />

The pathogen appears to have<br />

been becoming more widespread<br />

worldwide. A most interesting<br />

series <strong>of</strong> data is provided by Gilbert<br />

(1998) showing a more or less<br />

linear increase in M. graminicola<br />

incidence between 1990 <strong>and</strong> 1995 in<br />

southern Manitoba, from a very<br />

scarce disease to the most<br />

prevalent. Coupled with the<br />

genetic evidence that populations<br />

worldwide are genetically very<br />

similar (McDonald et al., 1995) it<br />

does seem possible that a novel<br />

form <strong>of</strong> the pathogen has been<br />

spreading worldwide. If true, this<br />

would raise the interesting<br />

question as to what<br />

epidemiological characteristic<br />

confers the new form’s<br />

invasiveness.<br />

Phaeosphaeria<br />

nodorum<br />

Within crops<br />

Multiplication is more rain<br />

dependent than M. graminicola,<br />

since pycnidia are produced in<br />

response to rain as well as<br />

dispersed by it. Conidia require<br />

wet periods to infect <strong>and</strong> do not<br />

tolerate interruptions well. The<br />

pathogen also appears to multiply<br />

ineffectively in cold conditions.<br />

Correlations with weather<br />

conditions are much better (Djurle<br />

et al., 1996; Gilbert et al., 1998) <strong>and</strong><br />

more closely linked in time to<br />

serious outbreaks than for M.<br />

graminicola. In wet, warm weather,<br />

typical <strong>of</strong> summer in some regions,<br />

latent periods can be shorter than<br />

in M. graminicola. Progress up the<br />

growing plant is correspondingly<br />

faster <strong>and</strong> more regular<br />

(Mittermeier <strong>and</strong> H<strong>of</strong>fmann, 1984).<br />

Like M. graminicola spores,<br />

spores <strong>of</strong> some isolates <strong>of</strong> P.<br />

nodorum infect with a startlingly<br />

high probability when rare on a leaf<br />

surface, the infection efficiency per<br />

spore falling rapidly with total<br />

numbers attacking in a droplet<br />

(Jeger et al., 1985)(MWS,<br />

unpublished). The ecological<br />

advantage is presumably that<br />

reinfection <strong>of</strong> a new crop is very<br />

effective, after which survival <strong>of</strong> the<br />

isolate is more or less certain. The<br />

epidemiological consequence is<br />

surprisingly rapid early spread <strong>of</strong><br />

the pathogen with subsequent rapid<br />

decline in apparent epidemic rate.<br />

Between crops<br />

Phaeosphaeria nodorum can be<br />

seedborne, <strong>and</strong> before routine<br />

fungicide treatment <strong>of</strong> seed <strong>and</strong> the<br />

upper parts <strong>of</strong> the crop, substantial<br />

numbers <strong>of</strong> seeds were infected. In<br />

some regions, such as northern<br />

Europe, the disease appears to have<br />

become much less important during<br />

the 1980s <strong>and</strong> 1990s. It is tempting<br />

to ascribe this decline to the<br />

increased use <strong>of</strong> fungicides<br />

reducing seed infection both in the<br />

ear <strong>and</strong> at planting. Recent<br />

experiments in other parts <strong>of</strong> the<br />

world suggest that seedborne<br />

inoculum is crucial in many places<br />

(Milus <strong>and</strong> Chalkley, 1997).<br />

Alternate hosts seem to harbor<br />

isolates <strong>of</strong> the pathogen capable <strong>of</strong><br />

attacking wheat, but there is<br />

sufficient partial specialization that<br />

these do not act as the source <strong>of</strong><br />

most infections on wheat as well as<br />

isolates clearly partially specialized<br />

to other hosts (Krupinsky, 1997a;<br />

1997b). Crop residues have been<br />

shown to be important (Cunfer,<br />

1998) but international survey<br />

evidence does not suggest they are<br />

the dominant source <strong>of</strong> inoculum<br />

(Leath et al., 1994).<br />

Ascospores have been found,<br />

but the evidence for their ubiquity<br />

<strong>and</strong> epidemiological importance<br />

seems to vary geographically.<br />

Trapping experiments like those<br />

used to show how airborne<br />

inoculum <strong>of</strong> P. nodorum varies<br />

through the year have been<br />

published by Cunfer (1998). These<br />

showed that in Georgia, USA,<br />

inoculum persisted for long periods<br />

in stubble, detectable sporadically<br />

for at least 18 months after the crop<br />

was harvested. As little as 17 m<br />

from a field edge, inoculum was<br />

trapped only once, in March, even<br />

though both mating types <strong>of</strong> the<br />

fungus were present <strong>and</strong> could<br />

potentially form perithecia. This<br />

suggests that sexual reproduction is<br />

unusual, though possibly not<br />

absent.<br />

These data are consistent with<br />

the relatively old published data on<br />

ascospore release in France (Rapilly<br />

et al., 1973). Anecdotally, near<br />

Reading in 1994 we observed a<br />

sudden patch <strong>of</strong> the disease in May<br />

in a first wheat. Although all<br />

isolations were from a single plot,<br />

<strong>and</strong> the disease had not been<br />

previously observed in that crop, 13<br />

genotypes tested using RFLPs were<br />

<strong>of</strong> different genotypes (C.N.F.M.R.J<br />

Pijls <strong>and</strong> M.W. Shaw, unpublished).<br />

However, the most extensive<br />

study made, in Pol<strong>and</strong>,<br />

demonstrated a quite different<br />

picture, with ascospores <strong>of</strong> P.<br />

nodorum present throughout the<br />

year <strong>and</strong> at densities that appeared<br />

to be independent <strong>of</strong> distance from<br />

the nearest infected plot (Arseniuk<br />

et al., 1998). Ascospores were also<br />

found throughout the year in<br />

northern Germany, but the trap was<br />

operated over an unharvested field,<br />

so it is not possible to say whether


the spores were from local or<br />

distant sources (Mittelstadt <strong>and</strong><br />

Fehrmann, 1987).<br />

References<br />

Arseniuk, E., T. Goral, <strong>and</strong> A.L.<br />

Scharen. 1998. Seasonal patterns <strong>of</strong><br />

spore dispersal <strong>of</strong> Phaeosphaeria<br />

spp. <strong>and</strong> <strong>Stagonospora</strong> spp. Plant<br />

Disease 82:187-194.<br />

Bannon, F.J., <strong>and</strong> B.M. Cooke. 1998.<br />

Studies on dispersal <strong>of</strong> <strong>Septoria</strong><br />

tritici pycnidiospores in wheatclover<br />

intercrops. Plant Pathology<br />

47:49-56.<br />

Brennan, R.M., B.D.L. Fitt, G.S. Taylor,<br />

<strong>and</strong> J. Colhoun. 1985a. Dispersal <strong>of</strong><br />

<strong>Septoria</strong> nodorum pycnidiospores by<br />

simulated rain <strong>and</strong> wild. Journal <strong>of</strong><br />

Phytopathology 112:291-297.<br />

Brennan, R.M., B.D.L. Fitt, G.S. Taylor,<br />

<strong>and</strong> J. Colhoun. 1985b. Dispersal <strong>of</strong><br />

<strong>Septoria</strong> nodorum pycnidiospores by<br />

simulated raindrops in still air.<br />

Journal <strong>of</strong> Phytopathology 112:281-<br />

290.<br />

Cunfer, B.M. 1998. Seasonal<br />

availability <strong>of</strong> inoculum <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum in the field in<br />

the southeastern USA. Cereal<br />

Research Communications 26:259-<br />

263.<br />

Djurle, A., B. Ekbom, <strong>and</strong> J.E. Yuen.<br />

1996. The relationship <strong>of</strong> leaf<br />

wetness duration <strong>and</strong> disease<br />

progress <strong>of</strong> glume blotch, caused<br />

by <strong>Stagonospora</strong> nodorum, in winter<br />

wheat to st<strong>and</strong>ard weather data.<br />

European Journal <strong>of</strong> Plant<br />

Pathology 102:9-20.<br />

Eriksen, L., M.W. Shaw, <strong>and</strong> H.<br />

Østergård. 1999. A model <strong>of</strong> the<br />

effect <strong>of</strong> pseudothecia on epidemic<br />

progress <strong>and</strong> genetic composition<br />

in Mycosphaerella graminicola. (In<br />

preparation.)<br />

Eyal, Z. 1971. The kinetics <strong>of</strong><br />

pycnospore liberation in <strong>Septoria</strong><br />

tritici. Canadian Journal <strong>of</strong> Botany<br />

49:1095-1099.<br />

Gilbert, J., S.M. Woods, <strong>and</strong> A. Tekauz.<br />

1998. Relationship between<br />

environmental variables <strong>and</strong> the<br />

prevalence <strong>and</strong> isolation frequency<br />

<strong>of</strong> leaf-spotting pathogen in spring<br />

wheat. Canadian Journal <strong>of</strong> Plant<br />

Pathology 20:158-164.<br />

Epidemiology <strong>of</strong> Mycosphaerella graminicola <strong>and</strong> Phaeosphaeria nodorum: An Overview 97<br />

Hunter, T., R.R. Coker, <strong>and</strong> D.J. Royle.<br />

1999. The teleomorph stage,<br />

Mycosphaerella graminicola, in<br />

epidemics <strong>of</strong> septoria tritici blotch<br />

on winter wheat in the UK. Plant<br />

Pathology 48:51-57.<br />

Jeger, M.J., E. Griffiths, <strong>and</strong> D.G.<br />

Jones. 1985. The effects <strong>of</strong> postinoculation<br />

wet <strong>and</strong> dry periods,<br />

<strong>and</strong> inoculum concentration, on<br />

lesion numbers <strong>of</strong> <strong>Septoria</strong> nodorum<br />

in spring wheat seedlings. Annals<br />

<strong>of</strong> Applied Biology 106:55-63.<br />

Kema, G.H.J., E.C.P. Verstappen, M.<br />

Todorova, <strong>and</strong> C. Waalwijk. 1996.<br />

Successful crosses <strong>and</strong> molecular<br />

tetrad <strong>and</strong> progeny analyses<br />

demonstrate heterothallism in<br />

Mycosphaerella graminicola. Current<br />

Genetics 30:251-258.<br />

Krupinsky, J.M. 1997a. Aggressiveness<br />

<strong>of</strong> <strong>Stagonospora</strong> nodorum isolates<br />

from perennial grasses on wheat.<br />

Plant Disease 81:1032-1036.<br />

Krupinsky, J.M. 1997b. Stability <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum isolates from<br />

perennial grass hosts after passage<br />

through wheat. Plant Disease<br />

81:1037-1041.<br />

Leath, S., A.L. Scharen, M.E.<br />

Dietzholmes, <strong>and</strong> R.E. Lund. 1994.<br />

Factors associated with global<br />

occurrences <strong>of</strong> septoria nodorum<br />

blotch <strong>and</strong> septoria tritici blotch <strong>of</strong><br />

wheat. Plant Disease 77:1266-1270.<br />

Lovell, D.J., S.R. Parker, T. Hunter, D.J.<br />

Royle, <strong>and</strong> R.R. Coker. 1997.<br />

Influence <strong>of</strong> crop growth <strong>and</strong><br />

structure on the risk <strong>of</strong> epidemics<br />

by Mycosphaerella graminicola<br />

(<strong>Septoria</strong> tritici) in winter wheat.<br />

Plant Pathology 46:126-138.<br />

McDonald, B.A., R.E. Pettway, R.S.<br />

Chen, J.M. Boerger, <strong>and</strong> J.P.<br />

Martinez. 1995. The population<br />

genetics <strong>of</strong> <strong>Septoria</strong> tritici<br />

(teleomorph Mycosphaerella<br />

graminicola). Canadian Journal <strong>of</strong><br />

Plant Pathology 73:292-301.<br />

Metcalfe, R.J. 1999. Selection for<br />

resistance to demethylation<br />

inhibitor fungicides in<br />

Mycosphaerella graminicola on<br />

wheat. PhD, The University <strong>of</strong><br />

Reading.<br />

Milus, E.A., <strong>and</strong> D.B. Chalkley. 1997.<br />

Effect <strong>of</strong> previous crop, seedborne<br />

inoculum, <strong>and</strong> fungicides on<br />

development <strong>of</strong> <strong>Stagonospora</strong><br />

blotch. Plant Disease 81:1279-1283.<br />

Mittelstadt, A., <strong>and</strong> H. Fehrmann.<br />

1987. Zum Auftreten der<br />

Hauptfruchtform von <strong>Septoria</strong><br />

nodorum in der Bundesrepublik<br />

Deutschl<strong>and</strong>. Zeitschrift für<br />

Pflanzenkrankheiten und<br />

Pflanzenschütz 94:380-385.<br />

Mittermeier, L., <strong>and</strong> G.M. H<strong>of</strong>fmann.<br />

1984. Untersuchungen zur<br />

Populationsentwicklung von<br />

<strong>Septoria</strong> nodorum in Feldbestund<br />

von Weizen. Zeitschrift fur<br />

Pflanzenkrankheiten und<br />

Pflanzenschutz 91:629-640.<br />

Rapilly, F., B. Foucault, <strong>and</strong> J.<br />

Lacazedieux. 1973. Études sur<br />

l’inoculum de <strong>Septoria</strong> nodorum<br />

Berk. (Leptosphaeria nodorum<br />

Müller) agent de la septoriose du<br />

blé. I. Les ascospores. Annales de<br />

Phytopathologie 5:131-141.<br />

Shaw, M.W. 1987. Assessment <strong>of</strong><br />

upward movement <strong>of</strong> rain splash<br />

using a fluorescent tracer method<br />

<strong>and</strong> its application to the<br />

epidemiology <strong>of</strong> cereal pathogens.<br />

Plant Pathology 36:201-213.<br />

Shaw, M.W. 1991a. Interacting effects<br />

<strong>of</strong> interrupted humid periods <strong>and</strong><br />

light on infection <strong>of</strong> wheat leaves<br />

by <strong>Septoria</strong> tritici (Mycosphaerella<br />

graminicola). Plant Pathology<br />

40:595-607.<br />

Shaw, M.W. 1991b. Variation in the<br />

height to which tracer is moved by<br />

splash during natural summer rain<br />

in the UK. Agricultural <strong>and</strong> Forest<br />

Meteorology 55:1-14.<br />

Shaw, M.W. 1999. Population<br />

dynamics <strong>of</strong> septoria in the crop<br />

ecosystem. In <strong>Septoria</strong> on <strong>Cereals</strong>:<br />

A Study <strong>of</strong> Pathosystems (J.A.<br />

Lucas, P. Bowyer <strong>and</strong> H.A.<br />

Anderson, eds.). CABI Publishing,<br />

Wallingford, UK. pp. 82-95.<br />

Shaw, M.W., <strong>and</strong> D.J. Royle. 1989.<br />

Airborne inoculum as a major<br />

source <strong>of</strong> <strong>Septoria</strong> tritici<br />

(Mycosphaerella graminicola)<br />

infections in winter wheat crops in<br />

the UK. Plant Pathology 38:35-43.<br />

Shaw, M.W., <strong>and</strong> D.J. Royle. 1993.<br />

Factors determining the severity <strong>of</strong><br />

Mycosphaerella graminicola (<strong>Septoria</strong><br />

tritici) on winter wheat in the UK.<br />

Plant Pathology 42:882-899.


98<br />

Spore Dispersal <strong>of</strong> Leaf Blotch Pathogens <strong>of</strong> Wheat<br />

(Mycosphaerella graminicola <strong>and</strong> <strong>Septoria</strong> tritici)<br />

C.A. Cordo, 1, 3 M.R. Simón, 2 A.E. Perelló, 1, 4 <strong>and</strong> H.E. Alippi1 1 Centro de Investigaciones de Fitopatología (CIDEFI), Facultad de Ciencias Agrarias y Forestales de La Plata,<br />

La Plata, Argentina<br />

2 Laboratorio de Cerealicultura, Facultad de Ciencias Agrarias y Forestales de La Plata, La Plata, Argentina<br />

3 Comisión de Investigaciones Científicas (CIC), Provincia de Buenos Aires, Argentina<br />

4 Consejo Nacional de Investigaciones Científicas y Tecnológicas (CONICET), Provincia de Buenos Aires,<br />

Argentina<br />

Abstract<br />

The spatial <strong>and</strong> temporal patterns <strong>of</strong> discharge <strong>and</strong> dissemination <strong>of</strong> air-borne Mycosphaerella graminicola spores<br />

<strong>and</strong> <strong>of</strong> <strong>Septoria</strong> tritici spores through rain splash were studied. Spore traps were used to monitor both ascospores <strong>and</strong><br />

pycnidiospores when the wheat crop was in the vegetative <strong>and</strong> debris states. Relationships between distance from a point<br />

source <strong>and</strong> weather variables such as rainfall, relative humidity, <strong>and</strong> air temperature were analyzed. The release <strong>of</strong><br />

pycnidiospores was favored by rainfall, as was explained through the multiple regression model (18% <strong>of</strong> the variation). Air<br />

dispersal <strong>of</strong> ascospores <strong>and</strong> splash dispersal <strong>of</strong> pycnidiospores were significantly influenced by each <strong>of</strong> the weather<br />

variables. The number <strong>of</strong> air-borne ascospores increased in association with rainfall. The multiple regression model<br />

explained 59% <strong>of</strong> the variation. The correlation analysis showed significant association with temperature, humidity, <strong>and</strong><br />

rainfall; the regression coefficients <strong>of</strong> the climatic variables were significant. The effect <strong>of</strong> different distances from the<br />

inoculum source on the density <strong>of</strong> rain-splashed pycnidiospores <strong>and</strong> wind-borne ascospores was not significant.<br />

Pycnidiospores were the omnipresent inoculum in the cereal-producing area during the observed period. Thus this<br />

inoculum poses a risk to crop production <strong>and</strong> may be important to the epidemiology <strong>of</strong> septoria diseases under the climatic<br />

conditions in the wheat-producing areas <strong>of</strong> Argentina.<br />

Leaf blotch, caused by <strong>Septoria</strong><br />

tritici Rob.ex Desm. (teleomorph<br />

Mycosphaerella graminicola (Fuckel)<br />

Schroeter, in Cohn) is an important<br />

wheat (Triticum aestivum) disease<br />

that causes yield losses in different<br />

regions <strong>of</strong> the world every year<br />

(Shipton et al., 1971; Eyal, 1981;<br />

Eyal et al., 1987). Its incidence<br />

depends on cultivar susceptibility,<br />

inoculum availability, crop<br />

management practices, <strong>and</strong><br />

favorable environmental conditions<br />

(cool temperature, high humidity,<br />

<strong>and</strong> frequent rain). Climatic factors,<br />

especially precipitation, affect<br />

fungal growth <strong>and</strong> the amount <strong>and</strong><br />

timing <strong>of</strong> spore production, as well<br />

as the release, dispersal, <strong>and</strong><br />

deposition <strong>of</strong> spores. Unusually<br />

intense rain may cause the onset <strong>of</strong><br />

a S. tritici epidemic in a wheat crop.<br />

The greatest risk to a crop is related<br />

to the occurrence <strong>of</strong> conditions that<br />

favor spore dispersal during <strong>and</strong><br />

shortly after flag leaf emergence.<br />

Spore dispersal <strong>and</strong> infection at this<br />

time favors a second generation <strong>of</strong><br />

pathogens. Spore dissemination<br />

patterns <strong>of</strong> Phaeosphaeria spp. <strong>and</strong><br />

<strong>Stagonospora</strong> spp. have been<br />

described (Arseniuk <strong>and</strong> Góral,<br />

1998; Góral <strong>and</strong> Arseniuk, 1998;<br />

1991). The release <strong>of</strong> M. graminicola<br />

ascospores occurred at two peak<br />

times <strong>of</strong> the year (at crop<br />

emergence <strong>and</strong> emergence <strong>of</strong> the<br />

upper two leaves) (T. Hunter<br />

unpublished).<br />

This work aimed at producing a<br />

mathematical model <strong>of</strong><br />

pycnidiospore <strong>and</strong> ascospore<br />

dispersal by:<br />

1. determining the pattern <strong>of</strong> spore<br />

dispersal during a 6-month<br />

period;<br />

2. investigating the effect <strong>of</strong><br />

climatic variables (rainfall,<br />

temperature, humidity) on<br />

density <strong>of</strong> air-borne spores; <strong>and</strong><br />

3. doing a preliminary analysis <strong>of</strong><br />

the dispersal distance for both<br />

types <strong>of</strong> spores.<br />

Materials <strong>and</strong> Methods<br />

In the experiment station<br />

situated in Los Hornos, near La<br />

Plata, Buenos Aires Province, airborne<br />

spores were collected on 13<br />

spore traps made <strong>of</strong> PVC capsules<br />

containing slides covered with<br />

petroleum jelly. The capsules were<br />

fixed to wooden stakes 0.7 m above<br />

the soil surface. Glass tubes 0.03 m<br />

in diameter <strong>and</strong> 0.16 m long were


placed near the capsules to catch<br />

rain-splashed spores. The capsules<br />

moved with the wind to ensure that<br />

the slides caught spores in different<br />

weather conditions. The spore traps<br />

were located at different distances<br />

(11 <strong>of</strong> them at 1.5 m <strong>and</strong> 2 at 12 m)<br />

from a wheat plot 30 m long <strong>and</strong> 10<br />

m wide. Spore samplers were<br />

arranged parallel to the length <strong>of</strong><br />

the test plot.<br />

The plot was sown on 24 June<br />

1998 <strong>and</strong> inoculated twice (17 July<br />

<strong>and</strong> 12 August) to obtain high<br />

disease incidence. Plants were<br />

inoculated when the second leaf<br />

unfolded (GS12). The main shoot<br />

<strong>and</strong> second side shoot (GS 22)<br />

(Zadoks et al., 1974) were<br />

inoculated by spraying a<br />

pycnidiospore suspension (5 x 10 6<br />

spores/ml) <strong>of</strong> one S. tritici isolate.<br />

Slides were collected <strong>and</strong> examined<br />

weekly from October 1998 to March<br />

1999. The rain-splashed spores were<br />

identified by observing one drop <strong>of</strong><br />

rain water stained with aniline blue<br />

(1/100) solution under the<br />

microscope. Four spore counts (1<br />

cm 2 each) were done on every<br />

sample. Furthermore, ascospores<br />

<strong>and</strong> pycnidiospores on the entire<br />

slide area covered with petroleum<br />

jelly were counted.<br />

Different populations <strong>of</strong> spores<br />

were observed under a light<br />

microscope at 100x magnification.<br />

Identification was based on the<br />

morphology <strong>of</strong> Mycosphaerella spp.<br />

<strong>and</strong> Phaeosphaeria spp. ascospores<br />

<strong>and</strong> <strong>Septoria</strong> spp. pycnidiospores.<br />

Viability <strong>of</strong> air-borne spores was<br />

determined by recording their<br />

germination directly on the<br />

petroleum jelly on the slides.<br />

Climatic variables (rainfall,<br />

temperature, <strong>and</strong> relative humidity)<br />

were recorded at the meteorological<br />

observatory in La Plata.<br />

Spore Dispersal <strong>of</strong> Leaf Blotch Pathogens <strong>of</strong> Wheat (Mycosphaerella graminicola <strong>and</strong> <strong>Septoria</strong> tritici) 99<br />

Multiple regression analyses<br />

were used to explore the<br />

relationship between the number <strong>of</strong><br />

pycnidiospores in rain water <strong>and</strong><br />

petroleum jelly, <strong>and</strong> ascospores in<br />

petroleum jelly, <strong>and</strong> the three<br />

weather variables. An analysis <strong>of</strong><br />

variance was performed to<br />

determine significant differences<br />

between spore traps placed at<br />

different distances. Pycnidiospore<br />

<strong>and</strong> ascospore dispersion was<br />

plotted in relation to climatic<br />

variables.<br />

Results <strong>and</strong> Discussion<br />

During the observed period, the<br />

weekly mean temperature ranged<br />

between 14ºC <strong>and</strong> 25ºC, relative<br />

humidity between 64 <strong>and</strong> 84%, <strong>and</strong><br />

rainfall between 0 <strong>and</strong> 135 mm.<br />

Figure 1a <strong>and</strong> b, <strong>and</strong> Figure 2a,<br />

b, <strong>and</strong> c show the distribution <strong>of</strong><br />

pycnidospores in rain water <strong>and</strong><br />

petroleum jelly <strong>and</strong> <strong>of</strong> ascospores in<br />

no. spores/slide/week<br />

no. spores/slide/week<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

5/10/98<br />

12/10/98<br />

5/10/98<br />

12/10/98<br />

18/10/98<br />

25/10/98<br />

3/11/98<br />

8/11/98<br />

Pycnidiospores<br />

Ascospores<br />

18/10/98<br />

25/10/98<br />

3/11/98<br />

8/11/98<br />

15/11/98<br />

22/11/98<br />

15/11/98<br />

22/11/98<br />

29/11/98<br />

5/12/98<br />

29/11/98<br />

5/12/98<br />

12/12/98<br />

12/12/98<br />

rain water with weather variables<br />

throughout the 24 weeks <strong>of</strong><br />

sampling.<br />

During the 6-month period,<br />

pycnidiospores were always<br />

present in both types <strong>of</strong> sampling.<br />

Figure 1a shows an increase in rainsplashed<br />

pycnidiospores around<br />

the week <strong>of</strong> 21 December. This<br />

increase was associated with a high<br />

rainfall regimen in which intense<br />

rains lasting as long as five hours<br />

without interruption occurred.<br />

However, correlation coefficients<br />

<strong>and</strong> partial correlation coefficients<br />

with weather variables were not<br />

significant (Table 1). Application <strong>of</strong><br />

a multiple regression model did not<br />

yield significant numbers either<br />

(Table 2). The increased amount <strong>of</strong><br />

pycnidiospores was associated with<br />

the increase in rainfall between 22<br />

January <strong>and</strong> 9 February 1999. The<br />

pycnidiospore pattern showed a<br />

marked reduction from 30<br />

December to 9 February. This was<br />

Mean <strong>of</strong> pycnidiospores<br />

21/12/98<br />

30/12/98<br />

Sampling date<br />

21/12/98<br />

30/12/98<br />

6/1/99<br />

15/1/99<br />

6/1/99<br />

15/1/99<br />

22/1/99<br />

29/1/99<br />

Mean <strong>of</strong> pycnidiospores <strong>and</strong> ascospores<br />

Sampling date<br />

22/1/99<br />

29/1/99<br />

9/2/99<br />

17/2/99<br />

9/2/99<br />

17/2/99<br />

26/2/99<br />

5/3/99<br />

26/2/99<br />

5/3/99<br />

12/3/99<br />

19/3/99<br />

12/3/99<br />

19/3/99<br />

Figure 1. Mean counts <strong>of</strong> pycnidiospores in rainwater (a), <strong>and</strong> pycnidiospores <strong>and</strong> ascospores in<br />

petroleum jelly (b) settled on microscope slides.<br />

A<br />

27/3/99<br />

B<br />

27/3/99


Session 5 — C.A. Cordo, M.R. Simón, A.E. Perelló, <strong>and</strong> H.E. Alippi<br />

100<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

5/10/98<br />

5/10/98<br />

5/10/98<br />

12/10/98<br />

18/10/98<br />

12/10/98<br />

18/10/98<br />

12/10/98<br />

18/10/98<br />

25/10/98<br />

3/11/98<br />

25/10/98<br />

3/11/98<br />

25/10/98<br />

3/11/98<br />

8/11/98<br />

8/11/98<br />

8/11/98<br />

15/11/98<br />

22/11/98<br />

15/11/98<br />

22/11/98<br />

15/11/98<br />

22/11/98<br />

29/11/98<br />

5/12/98<br />

29/11/98<br />

5/12/98<br />

29/11/98<br />

5/12/98<br />

Rainfall (mm)<br />

12/12/98<br />

21/12/98<br />

30/12/98<br />

6/1/99<br />

Sampling date<br />

Temperature (°C)<br />

12/12/98<br />

21/12/98<br />

30/12/98<br />

6/1/99<br />

Sampling date<br />

Relative humidity (%)<br />

12/12/98<br />

21/12/98<br />

30/12/98<br />

6/1/99<br />

Sampling date<br />

15/1/99<br />

22/1/99<br />

15/1/99<br />

22/1/99<br />

15/1/99<br />

22/1/99<br />

29/1/99<br />

9/2/99<br />

29/1/99<br />

9/2/99<br />

29/1/99<br />

9/2/99<br />

17/2/99<br />

26/2/99<br />

17/2/99<br />

26/2/99<br />

17/2/99<br />

26/2/99<br />

5/3/99<br />

5/3/99<br />

5/3/99<br />

12/3/99<br />

19/3/99<br />

12/3/99<br />

19/3/99<br />

12/3/99<br />

19/3/99<br />

Figure 2. Weekly mean rainfall (a), weekly mean temperature (b), <strong>and</strong> weekly mean relative<br />

humidity (c).<br />

related to the end <strong>of</strong> the crop<br />

vegetative period <strong>and</strong> the advent <strong>of</strong><br />

saprophytic pathogens on the<br />

wheat debris. A new increase in<br />

pycnidiospores on 5 March was<br />

related to the first natural infection<br />

<strong>of</strong> seedlings by septoria leaf blotch.<br />

The pycnidiospore pattern in<br />

petroleum jelly was constant from<br />

the first week <strong>of</strong> observation<br />

(Figure 1b). The increased density<br />

was associated with the amount <strong>of</strong><br />

rain water. The correlation<br />

A<br />

27/3/99<br />

B<br />

27/3/99<br />

C<br />

27/3/99<br />

coefficient <strong>and</strong> the partial<br />

regression coefficient between<br />

pycnidiospores <strong>and</strong> rainfall were<br />

significant (Table 1), as was the<br />

regression coefficient for rainfall in<br />

the multiple regression model. This<br />

model explained 18% <strong>of</strong> the<br />

variance (Table 2). Temperature <strong>and</strong><br />

humidity during this period were<br />

probably not so high as to affect the<br />

number <strong>of</strong> pycnidiospores released.<br />

Similar to observations by Arseniuk<br />

<strong>and</strong> Góral (1998), we recorded high<br />

Table 1. Correlation coefficients <strong>and</strong> partial<br />

correlation coefficients between number <strong>of</strong><br />

pycnidiospores in rain water <strong>and</strong> petroleum<br />

jelly <strong>and</strong> number <strong>of</strong> ascospores <strong>of</strong> <strong>Septoria</strong><br />

tritici <strong>and</strong> Mycosphaerella graminicola with<br />

weather variables.<br />

Temp- Relative<br />

erature humidity Rainfall<br />

Pycnidiospores/<br />

rain water -0.20 ns a 0.06 ns 0.26 ns<br />

-0.19 ns b 0.04 ns 0.04 ns<br />

Pycnidiospores/<br />

p. jelly 0.12 ns 0.11 ns 0.53**<br />

0.27 ns 0.01 ns 0.46*<br />

Ascospores 0.53** 0.46* 0.57**<br />

0.62** 0.42* 0.60**<br />

a= correlation coefficient.<br />

b= partial correlation coefficient.<br />

Table 2. Multiple regression model <strong>of</strong><br />

weather variables <strong>and</strong> number <strong>of</strong><br />

pycnidiospores in rain water, petroleum jelly,<br />

<strong>and</strong> ascospores <strong>of</strong> <strong>Septoria</strong> tritici <strong>and</strong><br />

Mycosphaerella graminicola.<br />

Significance<br />

Coefficient level<br />

No. <strong>of</strong> pycnidiospores<br />

in water<br />

Constant 9.33<br />

Temperature -0.26 ns<br />

Relative humidity -0.01 ns<br />

Rainfall<br />

No. <strong>of</strong> pycnidiospores<br />

in p. jelly<br />

0.02 ns<br />

Constant 11.31 ns<br />

Temperature 0.20 ns<br />

Relative humidity 0.07 ns<br />

Rainfall<br />

No. <strong>of</strong> ascospores<br />

0.08 P


ainfall increased. Periods <strong>of</strong> low<br />

rainfall were associated with low<br />

ascospore density (30 December to<br />

15 January, 17 February, <strong>and</strong> 5<br />

March). Temperatures after week 12<br />

were in general above 20ºC,<br />

whereas in the previous period they<br />

were generally lower. It is thus<br />

likely that both the disease cycle<br />

<strong>and</strong> the increase in temperature<br />

affected ascospore release. Analysis<br />

<strong>of</strong> the partial correlation, which<br />

considers variables as being<br />

independent, showed significant<br />

association with temperature,<br />

relative humidity, <strong>and</strong> rainfall<br />

(Table 1). The multiple regression<br />

model (Table 2) explained 59% <strong>of</strong><br />

the variance, <strong>and</strong> the regression<br />

coefficients for climatic variables<br />

were significant. As Arseniuk <strong>and</strong><br />

Góral established in 1998, the<br />

abundance <strong>of</strong> ascospores reached<br />

its peak at harvest time. After<br />

harvest, plant debris decomposed<br />

<strong>and</strong> the amount <strong>of</strong> air-borne spores<br />

was reduced. Despite this<br />

reduction, ascospores were also<br />

released when there was low<br />

rainfall <strong>and</strong> adequate relative<br />

humidity. Arseniuk <strong>and</strong> Góral<br />

(1998) observed that less than 1 mm<br />

<strong>of</strong> rainfall <strong>and</strong> high relative<br />

humidity were enough to increase<br />

the number <strong>of</strong> air-borne ascospores.<br />

Spore Dispersal <strong>of</strong> Leaf Blotch Pathogens <strong>of</strong> Wheat (Mycosphaerella graminicola <strong>and</strong> <strong>Septoria</strong> tritici) 101<br />

Finally, the effect <strong>of</strong> different<br />

distance from the inoculum source<br />

(1 m <strong>and</strong> 12 m) on the density <strong>of</strong><br />

ascospores <strong>and</strong> pycnidiospores was<br />

not significant, in agreement with<br />

Góral <strong>and</strong> Arseniuk (in press 1998;<br />

Arseniuk <strong>and</strong> Góral, 1998) (Table<br />

3). The number <strong>of</strong> spores near the<br />

inoculum source <strong>and</strong> at 12 m was<br />

similar.<br />

It is concluded that air dispersal<br />

<strong>of</strong> ascospores <strong>and</strong> pycnidiospores<br />

plays a role in the epidemiology <strong>of</strong><br />

Mycosphaerella (<strong>Septoria</strong>) blotch<br />

under the climatic conditions east<br />

<strong>of</strong> Buenos Aires, Argentina.<br />

Pycnidiospores were produced in<br />

greater abundance than ascospores<br />

<strong>and</strong> they were the predominant<br />

source <strong>of</strong> inoculum. The abundance<br />

<strong>of</strong> pycnidiospores probably reflects<br />

their greater importance in the<br />

epidemiology <strong>of</strong> leaf blotch <strong>of</strong><br />

wheat.<br />

Table 3. Analysis <strong>of</strong> variance for mean values<br />

<strong>of</strong> pycnidiospores <strong>and</strong> ascospores sampled at<br />

different distances from an inoculum source.<br />

Source <strong>of</strong> variation Mean square<br />

Distance 53.7 ns<br />

Error 35.0<br />

ns= not significant according to the F test. P=0.05.<br />

Acknowledgments<br />

We are grateful to the Comisión<br />

de Investigaciones Científicas,<br />

Buenos Aires Province, <strong>and</strong> the<br />

Consejo Nacional de<br />

Investigaciones Científicas y<br />

Tecnológicas for their financial<br />

support. We also thank Ing. Agr. D.<br />

Bayo, Mrs. N. Kripelz, Mr. B. Cordo<br />

López, <strong>and</strong> J. Balonga for their<br />

technical assistance.<br />

References<br />

Arseniuk, E., <strong>and</strong> T. Góral. 1998.<br />

Seasonal patterns <strong>of</strong> spore<br />

dispersal <strong>of</strong> Phaeosphaeria spp. <strong>and</strong><br />

<strong>Stagonospora</strong> spp. Plant Dis. 82:187-<br />

194.<br />

Eyal, Z. 1981 Integrated control <strong>of</strong><br />

<strong>Septoria</strong> diseases <strong>of</strong> wheat. Plant<br />

Dis. 65:763-768.<br />

Eyal, Z., A.L. Scharen, M.J. Prescott,<br />

<strong>and</strong> M. van Ginkel. 1987. The<br />

<strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Concepts <strong>and</strong> Methods <strong>of</strong> Disease<br />

Management. Mexico, D.F.:<br />

<strong>CIMMYT</strong>.<br />

Góral, T., <strong>and</strong> E. Arseniuk. 1991.<br />

Effect <strong>of</strong> climatic conditions on<br />

liberation <strong>and</strong> dispersal <strong>of</strong> spores<br />

<strong>of</strong> Leptosphaeria spp. in the air.<br />

Phytopath. Polonica 2 (XIV):28-34.<br />

Góral, T., <strong>and</strong> E. Arseniuk. 1998. The<br />

study <strong>of</strong> relationship between<br />

distance from a point inoculum<br />

source, weather variables, <strong>and</strong> the<br />

air presence <strong>of</strong> Phaeosphaeria spp.<br />

ascospores in warmer <strong>and</strong> cooler<br />

sub-periods <strong>of</strong> a growing season.<br />

Plant Breeding <strong>and</strong> Seed Science<br />

(in press).<br />

Shipton, W.A., W.J.R. Boyd, A.A.<br />

Rossielle, <strong>and</strong> B.I. Shearer. 1971.<br />

The common <strong>Septoria</strong> diseases <strong>of</strong><br />

wheat. Bot. Rev. 27:231-262.<br />

Zadoks, J.C., Chang, T.T., <strong>and</strong><br />

Konzak, C.F. 1974. A decimal code<br />

for the growth stages <strong>of</strong> cereals.<br />

Weed Research 14: 415-421.


102<br />

Epidemiology <strong>of</strong> Seedborne <strong>Stagonospora</strong> nodorum:<br />

A Case Study on New York Winter Wheat<br />

D.A. Shah <strong>and</strong> G.C. Bergstrom<br />

Department <strong>of</strong> Plant Pathology, Cornell University, Ithaca, New York, USA<br />

Abstract<br />

<strong>Stagonospora</strong> nodorum blotch is the most important component <strong>of</strong> the foliar disease complex that attacks New York winter<br />

wheat. Ascospores <strong>of</strong> the teleomorph Phaeosphaeria nodorum are not commonly observed in New York, where wheat is<br />

generally preceded in sequence by several years <strong>of</strong> nonhost, rotational crops. Wheat seed infection by <strong>Stagonospora</strong><br />

nodorum is common, the extent <strong>and</strong> range depending mainly on rainfall during the production season. Transmission <strong>of</strong> the<br />

pathogen from infected seeds to coleoptiles can approach 100% over a wide soil temperature range; transmission to the first<br />

leaves is less than 50% <strong>and</strong> is most efficient at soil temperatures below 17 ºC. Nevertheless, under the high densities at which<br />

wheat is sown, a significant number <strong>of</strong> infected seedlings per unit area can originate from relatively low initial seed infection<br />

levels <strong>and</strong> transmission efficiencies. <strong>Stagonospora</strong> nodorum blotch epidemics arising from infected seed potentially can be<br />

managed by reducing initial seedborne inoculum <strong>and</strong> its transmission. Wheat cultivars exhibit differential responses to<br />

infection <strong>of</strong> seed by S. nodorum, <strong>and</strong> this seed resistance appears to be under genetic control separate from foliar resistance.<br />

Breeding for reduced frequencies <strong>of</strong> seed infection <strong>and</strong> transmission, along with improved disease management in seed<br />

production fields <strong>and</strong> the application <strong>of</strong> seed fungicides, may comprise an effective integrated strategy for managing<br />

stagonospora nodorum blotch in wheat production systems similar to that in New York.<br />

The Pathosystem<br />

in New York<br />

Approximately 53,000 hectares<br />

<strong>of</strong> s<strong>of</strong>t winter wheat (mainly white<br />

cultivars for pastry flour) are<br />

cultivated annually as a rotational<br />

crop on vegetable, dairy, <strong>and</strong> cash<br />

grain farms in western <strong>and</strong> central<br />

New York. <strong>Stagonospora</strong> nodorum<br />

blotch is the most prevalent <strong>and</strong><br />

severe foliar disease affecting the<br />

crop (Schilder <strong>and</strong> Bergstrom,<br />

1989). The disease is currently<br />

undermanaged. Adapted wheat<br />

cultivars are susceptible, <strong>and</strong><br />

because <strong>of</strong> low grain prices <strong>and</strong><br />

inconsistent returns on fungicide<br />

expenditure, few producers apply<br />

foliar fungicide.<br />

Several observations led to the<br />

hypothesis that infected seed is the<br />

primary inoculum source for<br />

stagonospora nodorum blotch<br />

epidemics in New York. Firstly,<br />

infected wheat debris, regarded as<br />

the primary inoculum source in<br />

continuously or frequently cropped<br />

wheat systems, is eliminated from<br />

wheat fields by three- to six-year<br />

rotations. New York wheat fields<br />

are also relatively small <strong>and</strong><br />

scattered, reducing the chances <strong>of</strong><br />

infected debris blowing from one<br />

wheat field to another. <strong>Stagonospora</strong><br />

nodorum infects several grasses, but<br />

their role in epidemic initiation on<br />

wheat is likely minimal (Krupinsky,<br />

1997), <strong>and</strong> there is some evidence<br />

<strong>of</strong> host specificity (Ueng et al., 1994;<br />

Krupinsky, 1997). The teleomorphic<br />

stage, P. nodorum, formed typically<br />

on wheat straw, has not been<br />

recovered in New York, despite<br />

concerted searches <strong>of</strong> debris <strong>and</strong><br />

aerial spore trapping from fields<br />

affected heavily by stagonospora<br />

nodorum blotch (Bergstrom,<br />

unpublished). Moreover, most<br />

straw is baled <strong>and</strong> removed from<br />

fields shortly after harvest, <strong>and</strong> the<br />

remaining stubble is sometimes<br />

plowed under in late summer<br />

before planting alfalfa or grass, or<br />

is plowed the following spring<br />

prior to planting <strong>of</strong> corn,<br />

vegetables, or soybean.<br />

About 40% <strong>of</strong> New York winter<br />

wheat is underseeded with red<br />

clover, which serves as a winter<br />

cover crop. Clover attains a height<br />

<strong>of</strong> 30 to 40 cm, enough to<br />

completely cover unplowed<br />

stubble 20 to 25 cm high, thus<br />

impeding spore dispersal.<br />

Assuming that the P. nodorum stage<br />

does occur undetected, current<br />

cultural practices would diminish<br />

substantially its role in disease<br />

epidemiology in New York wheat.<br />

Shah et al. (1995), using DNA<br />

fingerprinted isolates <strong>of</strong> the<br />

pathogen, demonstrated that<br />

seedborne S. nodorum could initiate<br />

foliar epidemics under New York<br />

field conditions <strong>and</strong> foliar disease<br />

severity during grain development<br />

may be correlated with seed<br />

infection incidence (Shah et al.,<br />

1995).


Seed Surveys<br />

Most New York wheat<br />

producers sow certified seed that<br />

has passed minimal st<strong>and</strong>ards for<br />

germination, cultivar purity,<br />

contamination by weed seed <strong>and</strong><br />

foreign materials, <strong>and</strong> infection by<br />

smuts <strong>and</strong> bunts. Infection by S.<br />

nodorum is not part <strong>of</strong> the<br />

certification st<strong>and</strong>ard. Winter<br />

wheat seedlots were surveyed from<br />

1990 to 1996 for S. nodorum by<br />

assaying seed samples on SNAW<br />

medium (Man<strong>and</strong>har <strong>and</strong> Cunfer,<br />

1991). Parts <strong>of</strong> the survey have<br />

been published (Shah <strong>and</strong><br />

Bergstrom, 1993). In some years, all<br />

sampled lots contained seed<br />

infected by S. nodorum, <strong>and</strong> no<br />

cultivar was completely resistant to<br />

seed infection. Seasons with higher<br />

rainfall during wheat elongation<br />

through grain formation stages<br />

resulted in elevated seed infection<br />

incidences compared to low rainfall<br />

seasons (Figure 1).<br />

Transmission from<br />

Infected Seed<br />

Four lots (representing the cvs.<br />

Cayuga, Geneva, <strong>and</strong> Harus),<br />

ranging in infection incidence from<br />

Seed infection incidence (%)<br />

50<br />

140<br />

40<br />

120<br />

100<br />

30<br />

20<br />

1-71<br />

2-77<br />

80<br />

60<br />

40<br />

10<br />

0-19<br />

0-42<br />

20<br />

0<br />

1990 1991 1995<br />

Year<br />

1996<br />

0<br />

Figure 1. Mean seed infection incidence by<br />

<strong>Stagonospora</strong> nodorum in New York winter<br />

wheat lots in four production seasons.<br />

Incidence <strong>of</strong> infection range is shown over<br />

each bar. Rainfall averaged over the months<br />

<strong>of</strong> April-June <strong>and</strong> the 15 major wheat<br />

growing counties in western <strong>and</strong> central<br />

New York are shown by the line graph.<br />

Rain (mm)<br />

Epidemiology <strong>of</strong> Seedborne <strong>Stagonospora</strong> nodorum: A Case Study on New York Winter Wheat 103<br />

53 to 96%, were sown in soil at five<br />

temperatures (9, 13, 17, 21, 25 ºC),<br />

at 90 % relative humidity <strong>and</strong> a 16h<br />

photoperiod. All seedlings were<br />

harvested at the second leaf<br />

emerging stage, <strong>and</strong> coleoptiles<br />

were inspected for lesions induced<br />

by S. nodorum. Segments <strong>of</strong> the first<br />

leaf were plated onto Bannon’s<br />

medium (Bannon, 1978). Plates<br />

were incubated for 14 d under near<br />

ultraviolet light at room<br />

temperature. Leaf pieces were then<br />

inspected for S. nodorum pycnidia,<br />

which confirmed infection <strong>of</strong> the<br />

leaf. Transmission frequencies to<br />

the coleoptiles <strong>and</strong> first leaves were<br />

analyzed by SAS Proc Mixed<br />

(version 6.12; SAS Institute Inc.,<br />

Cary, NC).<br />

Germination was slightly, but<br />

not significantly, lower at 9 ºC.<br />

Transmission to the coleoptiles was<br />

high at each temperature, but the<br />

effect <strong>of</strong> temperature was<br />

significant (P = 0.0001), reflecting<br />

differences in transmission at the<br />

extremes <strong>of</strong> 9 <strong>and</strong> 25 ºC. Pycnidia<br />

were found sometimes on the<br />

coleoptiles. Transmission to the<br />

first leaf was much less frequent<br />

<strong>and</strong> was temperature dependent.<br />

The average transmission<br />

frequency at 9 ºC was 0.39, but<br />

dropped to 0.03 at 25 ºC. The<br />

pathogen was not distributed<br />

evenly in leaf tissue. It was<br />

recovered at higher frequency from<br />

sections <strong>of</strong> the first leaf proximal to<br />

the stem than from sections taken<br />

from the leaf apex region.<br />

Approximately 50% <strong>of</strong> infected first<br />

leaves were asymptomatic.<br />

At a transmission frequency <strong>of</strong><br />

0.1, based conservatively on our<br />

results, a seedlot with 5% infected<br />

seed would give rise to an average<br />

<strong>of</strong> one infected seedling per m 2 in a<br />

st<strong>and</strong> <strong>of</strong> 200 seedlings per m 2 .<br />

Simulations<br />

We simulated stagonospora<br />

nodorum blotch epidemics from<br />

infected seedlings at main shoot <strong>and</strong><br />

3 tillers to early dough, a period <strong>of</strong><br />

about 60 to 70 d for New York wheat,<br />

using apparent infection rates (sensu<br />

van der Plank, 1963) derived from or<br />

reported in the literature (Jeger et al.,<br />

1983; Rambow, 1989) <strong>and</strong> the logistic<br />

model (van der Plank, 1963). Near<br />

100% incidence <strong>of</strong> infection by the<br />

early dough stage is possible from as<br />

low as one infected plant per m 2<br />

(Figure 2). Relatively low seed<br />

infection incidences <strong>and</strong> transmission<br />

rates, without any increase in disease<br />

incidence over the winter, could<br />

conceiveably result in 50% <strong>of</strong> all<br />

plants infected by flowering (Figure<br />

3). Decreasing the initial seedling<br />

disease incidence can delay the<br />

epidemic significantly (Figure 3).<br />

Differential Seed<br />

Infection Among<br />

Cultivars<br />

To examine the possibility <strong>of</strong><br />

cultivar differential response to seed<br />

infection, we assayed seed samples<br />

Disease incidence<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 10 20 30 40 50 60 70<br />

Days<br />

Figure 2. Increase in stagonospora nodorum<br />

blotch incidence with apparent infection rates<br />

<strong>of</strong> 0.13, 0.14, 0.15, 0.16, 0.17, 0.18, or 0.19, starting<br />

from an initial seedling disease incidence <strong>of</strong><br />

0.005, which corresponds to 1 infected plant<br />

per m2 at a density <strong>of</strong> 200 per m2 . A period <strong>of</strong> 70<br />

days corresponds to the interval from main<br />

shoot <strong>and</strong> 3-tiller stage to early dough stage<br />

under New York winter wheat production<br />

conditions.


Session 5 — D.A. Shah <strong>and</strong> G.C. Bergstrom<br />

104<br />

from each <strong>of</strong> four New York winter<br />

wheat trial locations in 1995 <strong>and</strong><br />

1996. Data for two 1995 sites are<br />

shown in Figure 4. Results for 1996<br />

were similar. The effect <strong>of</strong> local<br />

environment is significant; seed<br />

infection at site B is greater for all<br />

cultivars than at site A. The initial<br />

inoculum sources were the same at<br />

both sites since seed used for<br />

sowing came from the same lot for<br />

any given cultivar. Some cultivars<br />

(Delaware <strong>and</strong> Houser, for<br />

instance) had consistently less seed<br />

infection than others, even across<br />

different production environments<br />

(Figure 4).<br />

Screening <strong>of</strong> the cultivars for<br />

differential responses to seed<br />

infection by S. nodorum was<br />

continued under glasshouse<br />

conditions. Flag leaves <strong>and</strong> ears<br />

were inoculated with spore<br />

suspensions <strong>of</strong> a single isolate at<br />

10 6 spores/ml. Flag leaves were<br />

assessed for percent leaf area<br />

diseased 15 d post-inoculation, <strong>and</strong><br />

seeds were assayed on SNAW<br />

medium upon maturity. Seed<br />

infection was high, but again, some<br />

cultivars had less seed infection<br />

than others. There was a strong<br />

correlation between the field based<br />

Disease incidence<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 10 20 30 40 50 60 70<br />

Days<br />

Figure 3. Increase in stagonospora nodorum<br />

blotch incidence from initial seedling<br />

disease incidences <strong>of</strong> 0.001, 0.003, 0.005, 0.01,<br />

0.02, or 0.04, given an apparent infection rate<br />

<strong>of</strong> 0.14. The 70 day period corresponds to the<br />

interval from main shoot <strong>and</strong> 3-tiller stage to<br />

early dough stage under New York winter<br />

wheat production conditions.<br />

observations <strong>and</strong> the results <strong>of</strong> the<br />

quantitative glasshouse<br />

inoculations when cultivars were<br />

ranked according to seed infection<br />

level. There appears to be genetic<br />

mechanisms <strong>of</strong> resistance to seed<br />

infection by S. nodorum. The poor<br />

correlation between flag leaf<br />

disease severity <strong>and</strong> seed infection<br />

incidence suggests that seed <strong>and</strong><br />

foliar resistance are under separate<br />

genetic control.<br />

Potential for Integrated<br />

Management<br />

No single control tactic is likely<br />

to reduce seedborne inoculum<br />

below levels that could initiate<br />

epidemics under favorable<br />

environments. The most effective<br />

triazole seed fungicides can reduce<br />

pathogen transmission in grossly<br />

infected seedlots so that only 1 to<br />

2% <strong>of</strong> seedlings are infected, still<br />

enough for epidemics to develop<br />

(Bergstrom, unpublished). A<br />

strategy integrating breeding for<br />

reduced seed infection <strong>and</strong><br />

transmission frequencies, improved<br />

disease management in seed<br />

production fields, <strong>and</strong> the<br />

application <strong>of</strong> seed fungicides may<br />

Seed infection incidence (%)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Site A<br />

Site B<br />

Delaware<br />

Houser<br />

Annett<br />

Chelsea<br />

ACRon<br />

Pioneer 2737W<br />

NYBatavia<br />

Karena<br />

Harus<br />

Caledonia<br />

Geneva<br />

Arkona<br />

Diana<br />

Year<br />

Figure 4. Differential seed infection <strong>of</strong> winter<br />

wheat cultivars by <strong>Stagonospora</strong> nodorum at<br />

two New York sites (A <strong>and</strong> B), located 5 km<br />

apart but differing in microclimate, in 1995.<br />

be employed to manage<br />

stagonospora nodorum blotch<br />

effectively in wheat production<br />

systems similar to that in New York.<br />

References<br />

Bannon, E. 1978. A method <strong>of</strong> detecting<br />

<strong>Septoria</strong> nodorum on symptomless<br />

leaves <strong>of</strong> wheat. Ir. J. Agric. Res.<br />

17:323-325.<br />

Jeger, M.J., Gareth-Jones, D., <strong>and</strong><br />

Griffiths, E. 1983. Disease spread <strong>of</strong><br />

non-specialised fungal pathogens<br />

from inoculated point sources in<br />

intraspecific mixed st<strong>and</strong>s <strong>of</strong> cereal<br />

cultivars. Ann. Appl. Biol. 102:237-<br />

244.<br />

Krupinsky, J.M. 1997. Stability <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum isolates from<br />

perennial grass hosts after passage<br />

through wheat. Plant Dis. 81:1037-<br />

1041.<br />

Man<strong>and</strong>har, J.B., <strong>and</strong> Cunfer, B.M. 1991.<br />

An improved selective medium for<br />

the assay <strong>of</strong> <strong>Septoria</strong> nodorum from<br />

wheat seed. Phytopathology 81:771-<br />

773.<br />

Rambow, M. 1989. Befallsentwicklung<br />

von <strong>Septoria</strong> nodorum Berk. in<br />

winterweizenbeständen.<br />

Nachrichtenbl. Pflanzenschutz DDR<br />

43:209-212.<br />

Schilder, A.M.C., <strong>and</strong> Bergstrom, G.C.<br />

1989. Distribution, prevalence, <strong>and</strong><br />

severity <strong>of</strong> fungal leaf <strong>and</strong> spike<br />

diseases <strong>of</strong> winter wheat in New<br />

York in 1986 <strong>and</strong> 1987. Plant Dis.<br />

73:177-182.<br />

Shah, D., <strong>and</strong> Bergstrom, G.C. 1993.<br />

Assessment <strong>of</strong> seedborne<br />

<strong>Stagonospora</strong> nodorum in New York<br />

s<strong>of</strong>t white winter wheat. Plant Dis.<br />

77:468-471.<br />

Shah, D., Bergstrom, G.C., <strong>and</strong> Ueng,<br />

P.P. 1995. Initiation <strong>of</strong> <strong>Septoria</strong><br />

nodorum blotch epidemics in winter<br />

wheat by seedborne <strong>Stagonospora</strong><br />

nodorum. Phytopathology 85:452-<br />

457.<br />

Ueng, P.P., Cunfer, B.M., <strong>and</strong> Chen, W.<br />

1994. Identification <strong>of</strong> the wheat <strong>and</strong><br />

barley biotypes <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum using restriction fragment<br />

length polymorphisms <strong>and</strong><br />

biological characteristics.<br />

Phytopathology 84:1146.<br />

Van der Plank, J.E. 1963. Plant <strong>Diseases</strong>:<br />

Epidemics <strong>and</strong> Control. New York<br />

<strong>and</strong> London. Academic Press. 349<br />

pp.


Sessions 6A <strong>and</strong> 6B: Cultural Practices <strong>and</strong> Disease Management<br />

Influence <strong>of</strong> Cultural Practices on <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

<strong>Diseases</strong><br />

J.M. Krupinsky<br />

USDA-Agricultural Research Service, Northern Great Plains Research Lab, M<strong>and</strong>an, ND, USA<br />

Abstract<br />

The use <strong>of</strong> cultural management practices is probably one <strong>of</strong> the oldest approaches to controlling plant diseases. How<br />

cultural management practices such as crop rotation, tillage (residue level), fertilizer application, seeding operations, <strong>and</strong><br />

disease-free seed can influence the incidence <strong>and</strong> severity <strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong> diseases are reviewed. Considering that<br />

early reports on cultural practices have been reviewed previously, recent research will be emphasized. Management practices<br />

such as the use <strong>of</strong> resistant cultivars, cultivar blends, <strong>and</strong> the use <strong>of</strong> fungicides will be covered by others at the workshop.<br />

From ancient times various<br />

cultural practices have been used to<br />

reduce plant diseases <strong>and</strong> have<br />

paralleled the development <strong>of</strong><br />

agriculture (Howard, 1996).<br />

Disease severity can be modified<br />

by various cultural practices, but<br />

the magnitude <strong>and</strong> direction <strong>of</strong><br />

these effects are not always<br />

consistent (King et al., 1983;<br />

Shipton et al., 1971; Tompkins et al.,<br />

1993). In some studies, disease<br />

severity can be affected more by<br />

trial location than by the agronomic<br />

practices being tested (Tompkins et<br />

al., 1993).<br />

Considering that early reports<br />

on cultural practices have been<br />

reviewed (Eyal, 1981; King et al.,<br />

1983; Shipton et al., 1971), more<br />

recent research will be emphasized<br />

in this paper. A number <strong>of</strong> cultural<br />

practices such as crop rotation,<br />

tillage, fertilization, seeding, <strong>and</strong><br />

disease-free seed will be reviewed.<br />

Other practices, including the use<br />

<strong>of</strong> resistant cultivars, cultivar<br />

blends, <strong>and</strong> fungicides are<br />

management practices that are<br />

covered by other authors in these<br />

proceedings.<br />

Crop Rotation: A Key<br />

Factor<br />

The survival <strong>of</strong> <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> spp. on residue from a<br />

previous wheat crop is the most<br />

important means <strong>of</strong> carryover from<br />

one crop to the next. Researchers<br />

have recommended proper crop<br />

rotations in combination with other<br />

practices to reduce the severity <strong>of</strong><br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> diseases.<br />

Plowing <strong>and</strong> burning have also<br />

been used to reduce the amount <strong>of</strong><br />

residue-borne inoculum, but these<br />

practices alone may still leave<br />

sufficient inoculum for the next<br />

wheat crop (Eyal, 1981; King et al.,<br />

1983; Shipton et al., 1971). Although<br />

resistant cultivars are covered by<br />

other authors in these proceedings,<br />

the impact <strong>of</strong> resistant cultivars on<br />

the carryover <strong>of</strong> pathogens should<br />

also be considered. Murray et al.<br />

(1990) pointed out with septoria<br />

tritici blotch (STB) that resistant<br />

cultivars not only reduce crop losses<br />

but the residue can also influence<br />

the next crop. Residue from<br />

resistant crops reduces the potential<br />

disease severity the following year<br />

because less inoculum is provided<br />

to initiate the epidemic.<br />

105<br />

Early researchers, such as Rosen<br />

(1921), included proper crop<br />

rotation in their recommendations<br />

for control <strong>of</strong> stagonospora<br />

nodorum blotch (SNB). Crop<br />

rotations take advantage <strong>of</strong> the fact<br />

that plant pathogens important on<br />

one crop may not cause disease<br />

problems on another crop.<br />

Appropriate crop rotations lengthen<br />

the time between crop types so that<br />

pathogen populations have time to<br />

decline. Although pathogens may<br />

not be completely eliminated, there<br />

is a reduction in the pathogen<br />

population. By rotating among crop<br />

types, the pathogens on residue<br />

from the previous crop will not<br />

infect the crop being grown<br />

(Krupinsky et al., 1997). Crop<br />

rotation allows time for the<br />

decomposition <strong>of</strong> residue on which<br />

pathogens carry over, <strong>and</strong> natural<br />

competitive organisms reduce the<br />

pathogens on the remaining residue<br />

while unrelated crops are being<br />

grown. Crop rotation is a key factor<br />

affecting the health <strong>and</strong> productivity<br />

<strong>of</strong> future wheat crops (Cook <strong>and</strong><br />

Veseth, 1991). Crop rotation <strong>and</strong><br />

residue management are most<br />

effective with pathogens that are<br />

disseminated only short distances<br />

but not as effective for pathogens


106<br />

Session 6A / Session 6B — J.M. Krupinsky<br />

disseminated over long distances.<br />

When inoculum is produced within<br />

a field <strong>and</strong> is disseminated only<br />

short distances, as with<br />

<strong>Stagonospora</strong> nodorum in the<br />

southeastern USA, crop rotation is<br />

an important management practice<br />

(Cunfer, 1998).<br />

When evaluating methods <strong>of</strong><br />

reducing the risk <strong>of</strong> disease<br />

severity with reduced tillage,<br />

Bailey <strong>and</strong> Duczek (1996) indicated<br />

that the most effective means <strong>of</strong><br />

reducing disease is crop rotation. In<br />

Saskatchewan, Canada, Pedersen<br />

<strong>and</strong> Hughes (1992) reported that<br />

crop rotation was effective in<br />

reducing the severity <strong>of</strong> epidemics<br />

caused by a leaf spot disease<br />

complex <strong>of</strong> S. nodorum <strong>and</strong> <strong>Septoria</strong><br />

tritici. In general, <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> disease severity is<br />

greater when wheat is grown in<br />

monoculture, compared to wheat<br />

following an alternative crop. In<br />

Germany, Käesbohrer <strong>and</strong><br />

H<strong>of</strong>fmann (1989) reported that<br />

winter wheat was infected earlier<br />

<strong>and</strong> more frequently by S. nodorum<br />

in wheat after wheat, compared<br />

with wheat in a crop rotation.<br />

Sutton <strong>and</strong> Vyn (1990) found that<br />

the severity <strong>of</strong> SNB on spikes <strong>of</strong><br />

winter wheat was higher after<br />

wheat compared to wheat after<br />

soybeans.<br />

Disease severity was<br />

significantly higher under<br />

monoculture than under rotation,<br />

particularly with no tillage, with a<br />

leaf-spot disease complex<br />

composed <strong>of</strong> Pyrenophora triticirepentis,<br />

Bipolaris sorokiniana, <strong>and</strong> S.<br />

nodorum (Reis et al., 1992, 1997).<br />

With P. tritici-repentis <strong>and</strong> S. tritici<br />

as the dominant pathogens, STB<br />

was significantly higher on winter<br />

wheat when grown in monoculture<br />

(Odorfer et al., 1994). With P. triticirepentis<br />

<strong>and</strong> S. nodorum as the<br />

dominant pathogens, Fern<strong>and</strong>ez et<br />

al. (1998) observed that wheat after<br />

wheat had a higher disease severity<br />

than wheat grown after flax or<br />

lentil. This was particularly evident<br />

in years with high disease pressure<br />

but not in years with low disease<br />

pressure.<br />

In general, most carry-over<br />

inoculum <strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

diseases is minimized by crop<br />

rotations that include wheat or<br />

other cereals every third year<br />

(Wiese, 1987). Considering that<br />

plant residue decays more rapidly<br />

in warm humid conditions than in<br />

drier <strong>and</strong> cooler conditions, the<br />

amount <strong>of</strong> time necessary between<br />

wheat crops may vary depending<br />

on the environment <strong>and</strong> location.<br />

In Israel, Eyal (1981) reported that a<br />

3-5 year rotation is needed to<br />

decrease the incidence <strong>of</strong> STB.<br />

Under favorable disease conditions<br />

in Saskatchewan, Canada, a<br />

rotation <strong>of</strong> two years between<br />

spring wheat crops reduced disease<br />

severity <strong>and</strong> provided adequate<br />

control <strong>of</strong> STB <strong>and</strong> SNB (Pedersen<br />

<strong>and</strong> Hughes, 1992). Also in<br />

Saskatchewan, Fern<strong>and</strong>ez et al.<br />

(1998) recommended two years <strong>of</strong><br />

spring wheat followed by two<br />

years <strong>of</strong> a non-cereal crop, or by a<br />

non-cereal crop <strong>and</strong> summer fallow<br />

to reduce disease severity (P. triticirepentis<br />

<strong>and</strong> S. nodorum).<br />

Since Fern<strong>and</strong>ez et al. (1993)<br />

indicated that S. nodorum, along<br />

with Fusarium graminearum <strong>and</strong> B.<br />

sorokiniana, survived longer on<br />

wheat residue after summer fallow<br />

compared to wheat residue after<br />

corn or soybean crops, summer<br />

fallow may not be as effective as an<br />

alternative crop in reducing<br />

inoculum levels. In the<br />

southeastern USA, S. nodorum<br />

inoculum may be present on wheat<br />

stubble on the soil surface 22<br />

months after a wheat crop is<br />

harvested, leading to<br />

recommendations for crop rotation<br />

<strong>and</strong> tillage (Cunfer, 1998). In<br />

addition to crop rotation,<br />

intercropping can be used to<br />

reduce inoculum movement in the<br />

field. In Irel<strong>and</strong>, Bannon <strong>and</strong> Cooke<br />

(1998) reported that the dispersal <strong>of</strong><br />

pycnidiospores <strong>of</strong> S. tritici was<br />

reduced 33% horizontally <strong>and</strong> 63%<br />

vertically in a wheat-clover<br />

intercrop.<br />

Tillage<br />

Infested residue provides an<br />

important mechanism for the<br />

carryover <strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

spp. from one crop to the next.<br />

Tillage promotes residue<br />

decomposition by fracturing the<br />

residue <strong>and</strong> exposing it to residuedecomposing<br />

microorganisms.<br />

Thus, tillage operations such as<br />

plowing have been recommended<br />

as a means to reduce residue (King<br />

et al., 1983; Shipton et al., 1971).<br />

In recent years, conservation<br />

tillage practices have been<br />

promoted to reduce soil erosion<br />

<strong>and</strong> conserve soil moisture. These<br />

minimum or no tillage operations<br />

increase the quantity <strong>of</strong> crop<br />

residue on the soil surface.<br />

Increased levels <strong>of</strong> crop residues<br />

may influence the incidence <strong>and</strong><br />

severity <strong>of</strong> plant diseases,<br />

depending upon the disease <strong>and</strong><br />

region. Conservation tillage or<br />

reduced tillage practices increase,<br />

decrease, or have no effect on plant<br />

diseases (Bailey <strong>and</strong> Duczek, 1996;<br />

Sumner et al., 1981). Conservation<br />

tillage may have different effects on<br />

plant diseases depending on the<br />

soils, region, or prevailing<br />

environment, <strong>and</strong> the biology <strong>of</strong><br />

the disease organism.


Differences in weather cycles in<br />

wheat growing areas are <strong>of</strong>ten<br />

dramatic, which may account for<br />

some <strong>of</strong> the differences in the<br />

impact <strong>of</strong> diseases described in the<br />

literature. When comparing tillage<br />

effects in northeastern North<br />

Dakota, USA, Stover et al. (1996)<br />

observed that different diseases<br />

dominated each year with different<br />

weather conditions. They found<br />

that P. tritici-repentis dominated in<br />

the first two years <strong>of</strong> the study, <strong>and</strong><br />

septoria foliar diseases <strong>and</strong><br />

fusarium head blight dominated in<br />

the third year.<br />

Tillage or lack <strong>of</strong> tillage can also<br />

affect the severity <strong>of</strong> individual<br />

diseases. Sutton <strong>and</strong> Vyn (1990)<br />

reported that P. tritici-repentis <strong>and</strong><br />

S. nodorum were promoted when<br />

wheat was grown after wheat <strong>and</strong><br />

minimum or no tillage was used,<br />

whereas S. tritici was suppressed.<br />

The reverse occurred when wheat<br />

followed alternative crops in all<br />

tillage treatments or following<br />

wheat under conventional tillage.<br />

With airborne ascospores <strong>of</strong> S.<br />

tritici implicated as the source <strong>of</strong><br />

primary inoculum in Oklahoma,<br />

USA, Schuh (1990) concluded that<br />

the amount <strong>of</strong> residue after<br />

different tillage operations did not<br />

have a strong effect on disease<br />

severity.<br />

In western Canada, Bailey <strong>and</strong><br />

Duczek (1996) indicated that foliar<br />

diseases increase under reduced<br />

tillage but not always to damaging<br />

levels. With conditions for low<br />

disease development, Bailey et al.<br />

(1992) did not observe consistent<br />

tillage effects on foliar diseases. In<br />

Kentucky, USA, Ditsch <strong>and</strong> Grove<br />

(1991) reported that SNB was not<br />

affected by tillage but powdery<br />

mildew infection (Erysiphe<br />

graminis) was higher under no<br />

tillage. In North Dakota, USA, with<br />

a leaf-spot disease complex<br />

composed mainly <strong>of</strong> P. triticirepentis<br />

<strong>and</strong> S. nodorum, Krupinsky<br />

et al. (1998) found generally higher<br />

levels <strong>of</strong> necrosis <strong>and</strong> chlorosis<br />

associated with no tillage<br />

compared to minimum or<br />

conventional tillage. However, in<br />

some trials the tillage effect varied<br />

depending on the nitrogen<br />

treatment. With a leaf-spot disease<br />

complex composed <strong>of</strong> P. triticirepentis,<br />

B. sorokiniana, <strong>and</strong> S.<br />

nodorum, Reis et al. (1992, 1997)<br />

reported that disease incidence <strong>and</strong><br />

severity were higher with<br />

minimum <strong>and</strong> no tillage treatments<br />

in Brazil. In the eastern USA, both<br />

tillage <strong>and</strong> crop rotation are<br />

recommended to avoid losses to<br />

SNB (Cunfer, 1998; Milus <strong>and</strong><br />

Chalkley, 1997).<br />

Burning <strong>of</strong> Residues<br />

Although burning crop residue<br />

was recommended in the past<br />

(King et al., 1983; Shipton et al.,<br />

1971), it is no longer recommended<br />

because <strong>of</strong> environmental concerns.<br />

In addition, burning may not be<br />

hot enough to eliminate all residue,<br />

leaving sufficient infested residue<br />

to provide carryover <strong>of</strong> inoculum<br />

for another wheat crop (Eyal, 1981).<br />

In northeastern North Dakota,<br />

USA, Stover et al. (1996) compared<br />

chisel plowing (high residue) to<br />

moldboard plowing <strong>and</strong> burning<br />

followed by moldboard plowing.<br />

The effect <strong>of</strong> chisel plowing on<br />

early season foliar disease did not<br />

consistently carry over to late<br />

season severity ratings. In the first<br />

year, chisel plowing resulted in the<br />

highest late season disease severity,<br />

in the second year, there were no<br />

differences, <strong>and</strong> in the third year,<br />

the chisel plow treatment had the<br />

lowest late season disease severity.<br />

Influence <strong>of</strong> Cultural Practices on <strong>Septoria</strong>/<strong>Stagonospora</strong> <strong>Diseases</strong> 107<br />

Overall, yields were not affected by<br />

a tillage (burning)-related foliar<br />

disease effect.<br />

Plant Nutrition: Nitrogen<br />

Fertilizer<br />

As with other management<br />

practices discussed above,<br />

environmental conditions influence<br />

disease development <strong>and</strong> treatment<br />

responses, leading to variation in<br />

research results among different<br />

geographical regions. The severity <strong>of</strong><br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> diseases may<br />

increase or decrease with increasing<br />

nitrogen rates depending on the<br />

region. Tiedemann (1996) suggested<br />

that inconsistencies <strong>of</strong> <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> disease response to<br />

nitrogen fertilization may be due to<br />

in part to environmental factors such<br />

as ozone.<br />

High rates <strong>of</strong> nitrogen fertilizer<br />

have been reported to increase the<br />

severity <strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

diseases. In addition, increased<br />

nitrogen can increase the danger <strong>of</strong><br />

lodging <strong>and</strong> delayed maturity<br />

(Shipton et al., 1971). In<br />

Pennsylvania, USA, Broscious et al.<br />

(1985) reported that the severity <strong>of</strong><br />

SNB <strong>and</strong> powdery mildew increased<br />

significantly on winter wheat as the<br />

rate <strong>of</strong> spring applied nitrogen<br />

fertilizer increased. Similarly, Ditsch<br />

<strong>and</strong> Grove (1991) in Kentucky, USA,<br />

reported that SNB <strong>and</strong> powdery<br />

mildew were lowest on winter<br />

wheat at the zero nitrogen treatment<br />

but increased as nitrogen rates<br />

increased. In New York, USA, under<br />

conditions <strong>of</strong> low disease severity,<br />

Cox et al. (1989) suggested that high<br />

rates <strong>of</strong> nitrogen have the potential<br />

to increase yield <strong>and</strong> foliar disease<br />

severity on winter wheat in the<br />

northeastern USA. In Tennessee,<br />

USA, Howard et al. (1994) reported<br />

that the severity <strong>of</strong> three foliar


108<br />

Session 6A / Session 6B — J.M. Krupinsky<br />

diseases (S. tritici, S. nodorum, <strong>and</strong><br />

leaf rust [Puccinia recondita])<br />

increased on winter wheat with<br />

higher nitrogen rates, especially<br />

when fungicides were not applied.<br />

In the United Kingdom, Leitch <strong>and</strong><br />

Jenkins (1995) reported that<br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> disease<br />

(principally STB) development on<br />

winter wheat was enhanced with<br />

the application <strong>of</strong> nitrogen<br />

throughout the season. A wide<br />

range <strong>of</strong> timing <strong>and</strong> splits <strong>of</strong><br />

nitrogen application did not<br />

significantly influence the level <strong>of</strong><br />

disease severity after anthesis. Also<br />

in the United Kingdom, Jenkyn <strong>and</strong><br />

King (1988) found an increase in<br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> diseases<br />

(mostly STB) on winter wheat after<br />

fallow compared to winter wheat<br />

after ryegrass. They attributed the<br />

increase in disease severity to an<br />

increased accumulation <strong>of</strong> available<br />

nitrogen during the fallow period.<br />

There have also been reports <strong>of</strong><br />

no effect or a decrease in the<br />

severity <strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

diseases on wheat with increased<br />

nitrogen rates. In Germany,<br />

assessing disease damage by the<br />

number <strong>of</strong> pycnidia <strong>and</strong> number <strong>of</strong><br />

latent infections <strong>of</strong> S. nodorum on<br />

winter wheat, Büschbell <strong>and</strong><br />

H<strong>of</strong>fmann (1992) reported that the<br />

influence <strong>of</strong> nitrogen rates was not<br />

significant. Also in Germany,<br />

Tiedemann (1996) reported that<br />

increased nitrogen reduced the<br />

severity <strong>of</strong> SNB on spring wheat,<br />

while increasing the severity <strong>of</strong><br />

powdery mildew <strong>and</strong> leaf rust.<br />

This nitrogen effect on the disease<br />

severity <strong>of</strong> SNB was reversed at<br />

elevated ozone concentrations. In<br />

the United Kingdom, late season<br />

applications <strong>of</strong> urea solution<br />

reduced the severity <strong>of</strong> STB on the<br />

flag leaf <strong>of</strong> winter wheat (Gooding<br />

et al., 1988). Naylor <strong>and</strong> Su (1988)<br />

reported that the severity <strong>of</strong> SNB<br />

on winter wheat was not affected<br />

by increased nitrogen levels early<br />

in the season <strong>and</strong> even decreased<br />

with increasing nitrogen later in the<br />

season.<br />

In Maryl<strong>and</strong>, USA, SNB was<br />

reduced on winter wheat with a<br />

higher nitrogen fertility rate (Orth<br />

<strong>and</strong> Grybauskas, 1994). They<br />

suggested that the reduction <strong>of</strong><br />

SNB was apparently due to<br />

interference <strong>of</strong> splash dispersal <strong>of</strong><br />

spores in a denser canopy <strong>and</strong> the<br />

suppressive effect <strong>of</strong> high nitrogen<br />

fertility. In glasshouse trials, they<br />

also reported that increased<br />

nitrogen fertility decreased the<br />

severity <strong>of</strong> SNB on the same winter<br />

wheat cultivars tested in the field.<br />

In North Dakota, USA, with a leafspot<br />

disease complex composed<br />

mainly <strong>of</strong> P. tritici-repentis <strong>and</strong> S.<br />

nodorum, Krupinsky et al. (1998)<br />

reported that with low nitrogen<br />

levels disease severity was higher<br />

in no tillage compared to<br />

conventional tillage. At higher<br />

nitrogen levels, the difference in<br />

disease severity for tillage<br />

treatments was greatly reduced.<br />

In Saskatchewan, Canada, the<br />

development <strong>of</strong> <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases on winter<br />

wheat was influenced by nitrogen<br />

fertility in one trial out <strong>of</strong> nine<br />

(Tompkins et al., 1993). Greater<br />

disease severity was associated<br />

with low nitrogen fertility. They<br />

suggested that lesion development<br />

may be promoted by nitrogen<br />

deficiency or a nutrient imbalance.<br />

Also in Saskatchewan, Fern<strong>and</strong>ez<br />

et al. (1998) reported an increase in<br />

disease severity with an increase in<br />

nitrogen deficiency in dry years<br />

with a leaf-spot disease complex<br />

composed mainly <strong>of</strong> P. triticirepentis<br />

<strong>and</strong> S. nodorum. Disease<br />

severity declined with treatments<br />

receiving no phosphorus.<br />

Seeding Operations<br />

A higher disease severity <strong>of</strong> STB was<br />

associated with earlier sowings in New<br />

South Wales, Australia (Murray et al.,<br />

1990). The longer time between sowing<br />

<strong>and</strong> heading probably leads to a higher<br />

disease severity. When studying<br />

seeding rates, row spacing, <strong>and</strong> depth<br />

<strong>of</strong> seeding in Pennsylvania, USA,<br />

Broscious et al. (1985) reported that<br />

increased seeding rates increased SNB<br />

in four out <strong>of</strong> 13 trials <strong>and</strong> decreased<br />

SNB in one trial. They suggested that<br />

growers could reduce row spacing from<br />

18 cm to 13 cm to increase yields<br />

without increasing disease severity. In<br />

Saskatchewan, Tompkins et al. (1993)<br />

reported that the severity <strong>of</strong> <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases increased with a<br />

higher seeding rate. Disease severity<br />

was not influenced by row spacing.<br />

Narrow row spacing (10 cm) with<br />

increased nitrogen rates reduced SNB<br />

on winter wheat <strong>and</strong> increased yields in<br />

Maryl<strong>and</strong>, USA (Orth <strong>and</strong> Grybauskas,<br />

1994).<br />

Use <strong>of</strong> Disease-Free Seed<br />

Seed infected with S. nodorum can be<br />

a source <strong>of</strong> primary inoculum <strong>and</strong> a<br />

probable source <strong>of</strong> transmission from<br />

one wheat crop to the next. Inoculum<br />

can survive in seed for extended<br />

periods <strong>of</strong> time. <strong>Septoria</strong> tritici can be<br />

seed-borne but is not a significant<br />

source <strong>of</strong> inoculum (Eyal, 1981; King et<br />

al., 1983; Shipton et al., 1971). Although<br />

the use <strong>of</strong> disease-free seed may not be<br />

considered a cultural practice, it should<br />

be evaluated by the producer as a<br />

management practice for reducing<br />

disease severity. Cultural practices such<br />

as crop rotation may not be effective if<br />

seed infected with S. nodorum is used<br />

for planting (Luke et al., 1983). The<br />

beneficial effect <strong>of</strong> disease-free seed can<br />

be negated by sowing into residue from<br />

a previous wheat crop (Luke et al., 1983,<br />

1985; Milus <strong>and</strong> Chalkley, 1997).


In Germany, Rambow (1990)<br />

indicated that infected seed was the<br />

main source <strong>of</strong> primary inoculum. In<br />

Pol<strong>and</strong>, Arseniuk et al. (1998)<br />

reported that higher seed infestation<br />

by S. nodorum resulted in higher<br />

disease levels in the field.<br />

<strong>Stagonospora</strong> nodorum was found in<br />

98% <strong>of</strong> the wheat seed samples<br />

tested in North Carolina, USA, <strong>and</strong><br />

the fungus survived for more than<br />

two years on stored seed (Babadoost<br />

<strong>and</strong> Hebert, 1984). Using seed lots<br />

infected 1% to 40% with S. nodorum,<br />

Luke et al. (1986) reported that 10%<br />

seed infection was adequate to cause<br />

a severe epidemic in the southeastern<br />

USA. In the northeastern USA, Shah<br />

et al. (1995) found that in 1990-91, a<br />

season mildly conductive to disease<br />

development, plots sown to seed<br />

with less than 1% infection by S.<br />

nodorum developed epidemics. In<br />

1991-92, epidemics were initiated<br />

with seed infection levels as low as<br />

0.5%. They demonstrated that the<br />

same isolates in the seed used for<br />

planting were found in the seed<br />

harvested. This indicated that seed<br />

populations <strong>of</strong> S. nodorum could<br />

initiate epidemics <strong>of</strong> SNB in new<br />

locations <strong>and</strong> could provide year-toyear<br />

perpetuation <strong>of</strong> these<br />

populations. In Manitoba, Canada,<br />

plants from shriveled seed caused by<br />

S. nodorum had lower seedling vigor<br />

but with increased number <strong>of</strong> tillers<br />

on plants from the shriveled seed,<br />

the yield from plump or shriveled<br />

seed plots could not be differentiated<br />

(Gilbert et al., 1995).<br />

When disease-free seed is not<br />

available, chemical treatments <strong>and</strong><br />

fungicides can be used to reduce or<br />

eliminate <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

pathogens from seed used for<br />

planting (Eyal, 1981; King et al., 1983;<br />

Shipton et al., 1971). The use <strong>of</strong><br />

fungicides for seed treatment are<br />

covered by other authors in these<br />

proceedings.<br />

Conclusion<br />

With relatively low value crops<br />

such as small cereal grains, there is<br />

a need for cultural management<br />

practices that can reduce the<br />

impact <strong>of</strong> diseases (Howard, 1996).<br />

With the adoption <strong>of</strong> conservation<br />

tillage practices, additional longterm<br />

research is needed to<br />

minimize disease severity with<br />

integrated diverse cropping<br />

systems. Hopefully new<br />

experiments that study the<br />

interaction <strong>of</strong> cultural practices,<br />

including the development <strong>of</strong> new<br />

crop rotation sequences, will<br />

provide additional information in<br />

the future. Adapted resistant<br />

cultivars, which are an important<br />

factor in reducing disease<br />

epidemics, should be used when<br />

available. The best strategy to<br />

minimize the impact <strong>of</strong> <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases is to integrate<br />

several cultural practices into one<br />

system. The integration <strong>of</strong> several<br />

cultural practices should minimize<br />

disease impacts on the economics<br />

for the grower without increasing<br />

the cost <strong>of</strong> production.<br />

Acknowledgments<br />

I thank B. M. Cunfer <strong>and</strong> D. E.<br />

Mathre for their reviews <strong>and</strong><br />

constructive comments. Mention <strong>of</strong><br />

a trademark, proprietary product,<br />

or company by USDA personnel is<br />

intended for explicit description<br />

only <strong>and</strong> does not constitute a<br />

guarantee or warranty <strong>of</strong> the<br />

product by the USDA <strong>and</strong> does not<br />

imply its approval to the exclusion<br />

<strong>of</strong> other products that may also be<br />

suitable. USDA-ARS, Northern<br />

Plains Area, is an equal<br />

opportunity/affirmative action<br />

employer <strong>and</strong> all agency services<br />

are available without<br />

discrimination.<br />

Influence <strong>of</strong> Cultural Practices on <strong>Septoria</strong>/<strong>Stagonospora</strong> <strong>Diseases</strong> 109<br />

References<br />

Arseniuk, E., Góral, T., Sowa, W.,<br />

Czembor, H.J., Krysiak, H., <strong>and</strong><br />

Scharen, A.L. 1998. Transmission <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum <strong>and</strong> Fusarium<br />

spp. on triticale <strong>and</strong> wheat seed <strong>and</strong><br />

the effect <strong>of</strong> seedborne <strong>Stagonospora</strong><br />

nodorum on disease severity under<br />

field conditions. J. Phytopathology<br />

146:339-345.<br />

Babadoost, M. <strong>and</strong> Hebert, T.T. 1984.<br />

Incidence <strong>of</strong> <strong>Septoria</strong> nodorum in<br />

wheat seed <strong>and</strong> its effects on plant<br />

growth <strong>and</strong> grain yield. Plant Dis.<br />

68:125-129.<br />

Bailey, K.L., <strong>and</strong> Duczek, L.J. 1996.<br />

Managing cereal diseases under<br />

reduced tillage. Can. J. Plant Pathol.<br />

18:159-167.<br />

Bailey, K.L., Mortensen K., <strong>and</strong> Lafond,<br />

G.P. 1992. Effects <strong>of</strong> tillage systems<br />

<strong>and</strong> crop rotations on root <strong>and</strong> foliar<br />

diseases <strong>of</strong> wheat, flax, <strong>and</strong> field<br />

peas in Saskatchewan. Can. J. Plant<br />

Sci. 72:583-591.<br />

Bannon, F.J., <strong>and</strong> Cooke, B.M. 1998.<br />

Studies on dispersal <strong>of</strong> <strong>Septoria</strong><br />

tritici pycnidiospores in wheatclover<br />

intercrops. Plant Pathol.<br />

47:49-56.<br />

Broscious, S.C., Frank, J.A., <strong>and</strong><br />

Frederick, J.R. 1985. Influence <strong>of</strong><br />

winter wheat management practices<br />

on the severity <strong>of</strong> powdery mildew<br />

<strong>and</strong> septoria blotch in Pennsylvania.<br />

Phytopathology 75:536-542.<br />

Büschbell, T., <strong>and</strong> H<strong>of</strong>fmann, G.M.<br />

1992. [The effects <strong>of</strong> different<br />

nitrogen regimes on the<br />

epidemiological development <strong>of</strong><br />

pathogens on winter wheat <strong>and</strong><br />

their control.] Z. PflKrankh.<br />

Pflchutz 99:381-403.<br />

Cook, R.J., <strong>and</strong> Veseth, R.J. 1991. Wheat<br />

health management. APS Press, St.<br />

Paul, MN, 152 pages.<br />

Cox, W.J., Bergstrom, G.C., Reid, W.S.,<br />

Sorrells, M.E., <strong>and</strong> Otis, D.J. 1989.<br />

Fungicide <strong>and</strong> nitrogen effects on<br />

winter wheat under low foliar<br />

disease severity. Crop Sci. 29:164-<br />

170.<br />

Cunfer, B.M. 1998. Seasonal availability<br />

<strong>of</strong> inoculum <strong>of</strong> <strong>Stagonospora</strong><br />

nodorum in the field in the<br />

southeastern U.S. Cereal Res.<br />

Comm. 26:259-263.<br />

Ditsch, D.C., <strong>and</strong> Grove, J.H. 1991.<br />

Influence <strong>of</strong> tillage on plant<br />

populations, disease incidence, <strong>and</strong><br />

grain yield <strong>of</strong> two s<strong>of</strong>t red winter<br />

wheat cultivars. J. Prod. Agric.<br />

4:360-365.


110<br />

Session 6A / Session 6B — J.M. Krupinsky<br />

Eyal, Z. 1981. Integrated control <strong>of</strong><br />

<strong>Septoria</strong> diseases <strong>of</strong> wheat. Plant Dis.<br />

65:763-768.<br />

Fern<strong>and</strong>ez, M.R., Fern<strong>and</strong>es, J.M., <strong>and</strong><br />

Sutton, J.C. 1993. Effects <strong>of</strong> fallow <strong>and</strong><br />

<strong>of</strong> summer <strong>and</strong> winter crops on<br />

survival <strong>of</strong> wheat pathogens in crop<br />

residues. Plant Dis. 77:698-703.<br />

Fern<strong>and</strong>ez, M.R., Zentner, R.P.,<br />

McConkey, B.G., <strong>and</strong> Campbell, C.A.<br />

1998. Effects <strong>of</strong> crop rotations <strong>and</strong><br />

fertilizer management on leaf<br />

spotting diseases <strong>of</strong> spring wheat in<br />

southwestern Saskatchewan. Can. J.<br />

Plant Sci. 78:489-496.<br />

Gilbert, J., Tekauz, A., <strong>and</strong> Woods, S.M.<br />

1995. Effect <strong>of</strong> Phaeosphaeria nodoruminduced<br />

seed shriveling on<br />

subsequent wheat emergence <strong>and</strong><br />

plant growth. Euphytica 82:9-16.<br />

Gooding, M.J., Kettlewell, P.S., <strong>and</strong><br />

Davies, W.P. 1988. Disease<br />

suppression by late season urea<br />

sprays on winter wheat <strong>and</strong><br />

interaction with fungicide. J.<br />

Fertilizer Issues 5:19-23.<br />

Howard, D.D., Chambers, A.Y., <strong>and</strong><br />

Logan, J. 1994. Nitrogen <strong>and</strong><br />

fungicide effects on yield components<br />

<strong>and</strong> disease severity in wheat. J. Prod.<br />

Agric. 7:448-454.<br />

Howard, R.J. 1996. Cultural control <strong>of</strong><br />

plant disease: a historical perspective.<br />

Can. J. Plant Pathol. 18:145-150.<br />

Jenkyn, J.F., <strong>and</strong> King, J.E. 1988. Effects<br />

<strong>of</strong> treatments to perennial ryegrass on<br />

the development <strong>of</strong> <strong>Septoria</strong> spp. in a<br />

subsequent crop <strong>of</strong> winter wheat.<br />

Plant Pathol. 37:112-119.<br />

Käesbohrer, M., <strong>and</strong> H<strong>of</strong>fmann, G.M.,<br />

1989. [Contribution to the population<br />

dynamics <strong>of</strong> <strong>Septoria</strong> nodorum in<br />

wheat crop systems.] Z. PflKrankh.<br />

Pflchutz 96:379-392.<br />

King, J.E., Cook, R. J., <strong>and</strong> Melville, S.C.<br />

1983. A review <strong>of</strong> septoria diseases <strong>of</strong><br />

wheat <strong>and</strong> barley. Ann. Appl. Biol.<br />

103:345-373.<br />

Krupinsky, J.M., Halvorson, A.D., <strong>and</strong><br />

Black, A.L. 1998. Leaf spot diseases <strong>of</strong><br />

wheat in a conservation tillage study.<br />

Pages 322-326. In: Helminthosporium<br />

Blights <strong>of</strong> Wheat: Spot Blotch <strong>and</strong> Tan<br />

Spot. Duveiller, E., Dubin, H.J.,<br />

Reeves, J. <strong>and</strong> McNab, A. eds.<br />

Mexico, D.F.: <strong>CIMMYT</strong>. 376 pages.<br />

Krupinsky, J.M., McMullen, M., Bailey,<br />

K.L., Duczek,L.J., <strong>and</strong> Gossen, B.D.<br />

1997. <strong>Diseases</strong>. Pages 29-33. In: Zero<br />

tillage, Advancing the Art. Manitoba-<br />

North Dakota Zero-till Tillage<br />

Farmers Association. 40 pages.<br />

Leitch, M.H., <strong>and</strong> Jenkins, P.D. 1995.<br />

Influence <strong>of</strong> nitrogen on the<br />

development <strong>of</strong> <strong>Septoria</strong> epidemics<br />

in winter wheat. J. Agric. Sci. 124:361-<br />

368.<br />

Luke, H.H., Pfahler, P.L., <strong>and</strong> Barnett,<br />

R.D. 1983. Control <strong>of</strong> <strong>Septoria</strong><br />

nodorum on wheat with crop rotation<br />

<strong>and</strong> seed treatment. Plant Dis. 67:949-<br />

951.<br />

Luke, H.H., Barnett, R.D., <strong>and</strong> Pfahler,<br />

P.L. 1985. Influence <strong>of</strong> soil infestation,<br />

seed infection, <strong>and</strong> seed treatment on<br />

<strong>Septoria</strong> nodorum blotch <strong>of</strong> wheat.<br />

Plant Dis. 69:74-76.<br />

Luke, H.H., Barnett, R.D., <strong>and</strong> Pfahler,<br />

P.L. 1986. Development <strong>of</strong> <strong>Septoria</strong><br />

nodorum blotch on wheat from<br />

infected <strong>and</strong> treated seed. Plant Dis.<br />

70:252-254.<br />

Milus, E.A., <strong>and</strong> Chalkley, D.B. 1997.<br />

Effect <strong>of</strong> previous crop, seedborne<br />

inoculum, <strong>and</strong> fungicides on<br />

development <strong>of</strong> <strong>Stagonospora</strong> blotch.<br />

Plant Dis. 81:1279-1283.<br />

Murray, G.M., Martin, R.H., <strong>and</strong> Cullis,<br />

B.R. 1990. Relationship <strong>of</strong> the severity<br />

<strong>of</strong> <strong>Septoria</strong> tritici blotch <strong>of</strong> wheat to<br />

sowing time, rainfall at heading <strong>and</strong><br />

average susceptibility <strong>of</strong> wheat<br />

cultivars in the area. Aust. J. Agric.<br />

Res. 41:307-315.<br />

Naylor, R.E.L., <strong>and</strong> Su, J. 1988.<br />

Comparison <strong>of</strong> disease incidence on<br />

triticale <strong>and</strong> wheat at different<br />

nitrogen levels without fungicide<br />

treatment. Test <strong>of</strong> Agrochemicals <strong>and</strong><br />

Cultivars 9:110-111.<br />

Odorfer, A., Obst, A., <strong>and</strong> Pommer, G.<br />

1994. [The effects <strong>of</strong> different leaf<br />

crops in long lasting monoculture<br />

with winter wheat. 2 nd<br />

communication: Disease<br />

development <strong>and</strong> effects <strong>of</strong><br />

phytosanitary measures]. Agribiol.<br />

Res. 47:56-66.<br />

Orth, C.E., <strong>and</strong> Grybauskas, A.P. 1994.<br />

Development <strong>of</strong> <strong>Septoria</strong> nodorum<br />

blotch on winter wheat under two<br />

cultivation schemes in Maryl<strong>and</strong>.<br />

Plant Dis. 78.736-741.<br />

Pedersen, E.A., <strong>and</strong> Hughes, G.R. 1992.<br />

The effect <strong>of</strong> crop rotation on<br />

development <strong>of</strong> the septoria complex<br />

on spring wheat in Saskatchewan.<br />

Can. J. Plant Pathol. 14:152-158.<br />

Rambow, M. 1990. [The importance <strong>of</strong><br />

seed infestation as an infection<br />

source for <strong>Septoria</strong> nodorum Berk.]<br />

Plant Protection Bull. 44:153-155.<br />

Reis, E.M., Santos, H.P., Lhamby, J.C.B.,<br />

<strong>and</strong> Blum, M.C. 1992. Effect <strong>of</strong> soil<br />

management <strong>and</strong> crop rotation on<br />

the control <strong>of</strong> leaf blotches <strong>of</strong> wheat<br />

in Southern Brazil. Pages 217-236.<br />

In: Congresso Interamericano de<br />

Siembra Directa, 1/ Jornadas<br />

Binacionales de Cero Labranza, 2,<br />

Villa Giardino.<br />

Reis, E.M., Casa, R.T., Blum, M.M.C.,<br />

Santos, H.P., <strong>and</strong> Medeiros, C.A.<br />

1997. Effects <strong>of</strong> cultural practices on<br />

the severity <strong>of</strong> leaf blotches <strong>of</strong><br />

wheat <strong>and</strong> their relationship to the<br />

incidence <strong>of</strong> pathogenic fungi in the<br />

harvested seed. Fitopatologia<br />

Brasileira 22:407-412.<br />

Rosen, H.R. 1921. <strong>Septoria</strong> glume<br />

blotch <strong>of</strong> wheat. Univ. <strong>of</strong> Arkansas<br />

Bull. No. 175. 17 pages.<br />

Schuh, W. 1990. Influence <strong>of</strong> tillage<br />

system on disease intensity <strong>and</strong><br />

spatial pattern <strong>of</strong> <strong>Septoria</strong> leaf<br />

blotch. Phytopathology 80:1337-<br />

1340.<br />

Shah, D., Bergstrom, G.C., <strong>and</strong> Ueng,<br />

P.P. 1995. Initiation <strong>of</strong> <strong>Septoria</strong><br />

nodorum blotch epidemics in<br />

winter wheat by seedborne<br />

<strong>Stagonospora</strong> nodorum.<br />

Phytopathology 85:452-457.<br />

Shipton, W.A., Boyd, W.R.J., Rosielle,<br />

A.A., <strong>and</strong> Shearer, B.I. 1971. The<br />

common septoria diseases <strong>of</strong> wheat.<br />

Bot. Rev. 37:231-262.<br />

Stover, R.W., Francl, L.J., <strong>and</strong> Jordahl,<br />

J.G. 1996. Tillage <strong>and</strong> fungicide<br />

management <strong>of</strong> foliar diseases in a<br />

spring wheat monoculture. J. Prod.<br />

Agric. 9:261-265.<br />

Sumner, D.R., Doupnik, Jr., B., <strong>and</strong><br />

Boosalis, M.G. 1981. Effects <strong>of</strong><br />

reduced tillage <strong>and</strong> multiple<br />

cropping on plant diseases. Ann.<br />

Rev. Phytopathol. 19:167-87.<br />

Sutton, J.C., <strong>and</strong> Vyn, T.J. 1990. Crop<br />

sequences <strong>and</strong> tillage practices in<br />

relation to diseases <strong>of</strong> winter wheat<br />

in Ontario. Can. J. Plant Pathol.<br />

12:358-368.<br />

Tiedemann, A.V. 1996. Single <strong>and</strong><br />

combined effects <strong>of</strong> nitrogen<br />

fertilization <strong>and</strong> ozone on fungal<br />

leaf diseases on wheat. J. Plant Dis.<br />

<strong>and</strong> Protection 103:409-419.<br />

Tompkins, D.K., Fowler, D.B., <strong>and</strong><br />

Wright, A.T. 1993. Influence <strong>of</strong><br />

agronomic practices on canopy<br />

microclimate <strong>and</strong> septoria<br />

development in no-till winter wheat<br />

produced in the Parkl<strong>and</strong> region <strong>of</strong><br />

Saskatchewan. Can. J. Plant Sci.<br />

73:331-344.<br />

Wiese, M.V. 1987. Compendium <strong>of</strong><br />

wheat diseases. APS Press, St. Paul,<br />

Minn. 112 pages.


Disease Management Using<br />

Varietal Mixtures<br />

C.C. Mundt, C. Cowger, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

Department <strong>of</strong> Botany <strong>and</strong> Plant Pathology, Oregon State University, Corvallis, OR, USA<br />

Abstract<br />

Few data are available that evaluate the impact <strong>of</strong> variety mixtures on the <strong>Septoria</strong>/<strong>Stagonospora</strong> diseases <strong>of</strong> cereals.<br />

The splash-dispersed nature <strong>of</strong> the pathogens <strong>and</strong> the predominance <strong>of</strong> quantitative variation for host resistance <strong>and</strong><br />

pathogenicity in the <strong>Septoria</strong>/<strong>Stagonospora</strong> diseases may make interactions more complex than for the rusts <strong>and</strong><br />

mildews, where major gene interactions have been the main object <strong>of</strong> study. We have found that epidemic progression <strong>of</strong><br />

septoria tritici blotch can be substantially suppressed in mixtures <strong>of</strong> a susceptible <strong>and</strong> a moderately resistant variety,<br />

sometimes to below the level <strong>of</strong> the more resistant variety grown in pure st<strong>and</strong>. Mycosphaerella graminicola populations<br />

sampled from variety mixtures have been found to be reduced in pathogenicity in all mixtures that have been investigated<br />

thus far. Disruptive selection may be an important mechanism affecting disease progression <strong>and</strong> evolution <strong>of</strong> M.<br />

graminicola in variety mixtures. Major genes for resistance have the potential to contribute substantially to the use <strong>of</strong><br />

variety mixtures for control <strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong> diseases, both through their epidemiological impacts on disease<br />

spread, <strong>and</strong> through effects <strong>of</strong> induced resistance between avirulent <strong>and</strong> virulent genotypes <strong>of</strong> the pathogen that may occur<br />

in variety mixtures.<br />

Variety mixtures have been<br />

investigated to the greatest extent<br />

for obligate parasites <strong>of</strong> small<br />

grains (Garrett <strong>and</strong> Mundt, 1999;<br />

Wolfe, 1985), <strong>and</strong> are being utilized<br />

commercially for control <strong>of</strong> these<br />

diseases (Garrett <strong>and</strong> Mundt, 1999;<br />

Mundt, 1994). Much less<br />

information is available regarding<br />

the impact <strong>of</strong> variety mixtures on<br />

diseases caused by non-obligate<br />

pathogens.<br />

The effects <strong>of</strong> variety mixtures<br />

may differ for the <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases compared to<br />

rusts <strong>and</strong> mildews for at least two<br />

reasons. First, steep dispersal<br />

gradients caused by splash<br />

dispersal <strong>of</strong> conidia may decrease<br />

inoculum exchange among host<br />

genotypes in mixture <strong>and</strong>, hence,<br />

reduce the efficacy <strong>of</strong> mixtures for<br />

disease control (Fitt <strong>and</strong><br />

McCartney, 1986; Mundt <strong>and</strong><br />

Leonard, 1986). Second, qualitative<br />

incompatibility reactions are much<br />

less common for septoria diseases<br />

in current agricultural systems<br />

compared to the rusts <strong>and</strong><br />

mildews.<br />

Though there clearly is much<br />

more host/pathogen specificity<br />

than has been suggested in the past<br />

for Mycosphaerella graminicola<br />

(Ahmed et al., 1995; Ahmed et al.,<br />

1996; Cowger et al., this volume;<br />

Kema et al., 1986; 1987), the general<br />

lack <strong>of</strong> use <strong>of</strong> major genes for<br />

resistance to the <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases in agriculture<br />

places much greater emphasis on<br />

quantitative interactions. For M.<br />

graminicola, there clearly can be<br />

substantial quantitative variation<br />

for pathogenicity that is under the<br />

influence <strong>of</strong> host selection (Ahmed<br />

et al., 1995; 1996; Mundt et al.,<br />

1999) <strong>and</strong> may be influenced by<br />

host diversity. In addition, Jeger et<br />

al. (1981a) have derived models<br />

111<br />

which indicate, based on<br />

epidemiological considerations,<br />

that variety mixtures can impact<br />

disease severity even in the absence<br />

<strong>of</strong> host/pathogen specificity.<br />

There are a limited number <strong>of</strong><br />

studies that address the impact <strong>of</strong><br />

variety mixtures on <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> diseases in the field.<br />

Jeger et al. (1981b) found that<br />

mixing wheat varieties with<br />

apparent non-specific resistance<br />

reduced the severity <strong>of</strong><br />

stagonospora nodorum blotch by<br />

more than 50% relative to the<br />

mixture components grown<br />

separately in pure st<strong>and</strong>s, though<br />

overall severity levels were very<br />

low. Mundt et al. (1995) studied all<br />

possible equiproportional mixtures<br />

among two susceptible, one<br />

moderately resistant, <strong>and</strong> one<br />

highly resistant wheat variety.<br />

Mean disease reductions late in the


Session 6A / Session 6B — C.C. Mundt, C. Cowger, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

112<br />

epidemics were 27, 9, <strong>and</strong> 15% for<br />

the three seasons, with substantial<br />

differences among specific<br />

mixtures.<br />

The purpose <strong>of</strong> the work<br />

reported herein was to investigate<br />

in greater detail the effects <strong>of</strong> wheat<br />

variety mixtures on epidemic<br />

progression <strong>of</strong> septoria tritici blotch<br />

<strong>and</strong> to determine the influence <strong>of</strong><br />

host diversity on pathogenicity <strong>of</strong><br />

M. graminicola populations derived<br />

from field plots <strong>of</strong> pure <strong>and</strong> mixed<br />

st<strong>and</strong>s <strong>of</strong> wheat varieties in the<br />

Willamette Valley <strong>of</strong> Oregon, USA.<br />

This location is highly conducive<br />

for development <strong>of</strong> the <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> blotches <strong>of</strong> wheat, with<br />

frequent rains from November<br />

through June. Naturally occurring<br />

epidemics have varied from<br />

moderate to severe over the past<br />

decade, <strong>and</strong> commercial fields <strong>of</strong><br />

susceptible varieties are sprayed<br />

every year.<br />

In Oregon, septoria tritici blotch<br />

is currently more important than<br />

stagonospora nodorum blotch,<br />

perhaps due to competitive<br />

exclusion caused by earlier<br />

ascospore showers <strong>of</strong> M. graminicola<br />

as compared to Phaeosphaeria<br />

nodorum (DiLeone et al., 1997).<br />

<strong>Septoria</strong> tritici blotch epidemics in<br />

Oregon begin from a very dominant<br />

sexual stage, with primary<br />

infections resulting from heavy<br />

ascospore showers (Mundt et al.,<br />

1999). Ascospore showers <strong>of</strong>ten<br />

peak in November, after fall rains<br />

have begun, but continue at some<br />

level throughout the crop season<br />

(DiLeone et al., 1997), including<br />

secondary cycles <strong>of</strong> sexual<br />

recombination (Zhan et al., 1998).<br />

Materials <strong>and</strong> Methods<br />

Field plots<br />

Plots were grown at the Botany<br />

<strong>and</strong> Plant Pathology Field<br />

Laboratory near Corvallis, OR,<br />

during the 1997-98 winter wheat<br />

season. Plots were sown at a rate <strong>of</strong><br />

322 seeds/m2 on 20 October 1997.<br />

Each plot was 3.0 m (12 rows) x 5.2<br />

m, <strong>and</strong> planted within a<br />

“checkerboard” <strong>of</strong> barley (Hordeum<br />

vulgare) plots <strong>of</strong> equal size.<br />

St<strong>and</strong>ard fertilization <strong>and</strong> weed<br />

control practices were used,<br />

supplemented by h<strong>and</strong>-weeding.<br />

Epidemics were initiated from<br />

naturally occurring ascospore<br />

showers.<br />

Treatments included two<br />

susceptible (Stephens <strong>and</strong> W-301)<br />

<strong>and</strong> two moderately resistant<br />

(Cashup <strong>and</strong> Madsen) varieties,<br />

<strong>and</strong> the four possible<br />

equiproportional mixtures <strong>of</strong> a<br />

susceptible <strong>and</strong> a moderately<br />

resistant variety. Treatments were<br />

replicated four times in a<br />

r<strong>and</strong>omized complete block design.<br />

Percent <strong>of</strong> leaf area covered by<br />

lesions, on a whole-canopy basis,<br />

was recorded from early February<br />

through mid-June. Each plot<br />

received a rating that was the<br />

average <strong>of</strong> two observers.<br />

At the end <strong>of</strong> the season, leaves<br />

were sampled from each plot to<br />

obtain isolates for greenhouse tests<br />

<strong>of</strong> pathogenicity, as described<br />

below. The pathogen was sampled<br />

along a diagonal across the inner 10<br />

rows <strong>of</strong> each plot, with one<br />

sampling point per row. Two or<br />

three flag leaves were collected<br />

from each sampling point. Leaf<br />

samples were collected on 12 June<br />

1998, a time when the pathogen<br />

had been exposed to repeated<br />

generations <strong>of</strong> selection on the host<br />

varieties.<br />

Greenhouse experiment<br />

A greenhouse experiment was<br />

conducted in spring 1999 to<br />

determine pathogenicity <strong>of</strong> the<br />

field-collected isolates. Plants <strong>of</strong><br />

Stephens, W-301, Cashup, <strong>and</strong><br />

Madsen were raised in 10-cm<br />

plastic pots filled with a<br />

greenhouse mix that included a<br />

slow-release fertilizer.<br />

Approximately 15 seeds <strong>of</strong> a<br />

cultivar were sown in each pot.<br />

Before inoculation, plants were<br />

thinned to 10 per pot. Daytime<br />

greenhouse temperature was<br />

maintained from 20-25°C, <strong>and</strong><br />

sodium-halide lights were used to<br />

extend daylength to 16 h.<br />

Cultures were obtained from<br />

field-collected leaves, stored, <strong>and</strong><br />

increased by st<strong>and</strong>ard methods that<br />

have previously been described<br />

(Ahmed et al., 1996). Equal<br />

numbers <strong>of</strong> spores <strong>of</strong> the 10 isolates<br />

(one from each <strong>of</strong> the sampling<br />

points <strong>of</strong> each plot) were combined<br />

<strong>and</strong> the total suspension adjusted<br />

to 105 spores/ml.<br />

At 21 days after seeding, groups<br />

<strong>of</strong> four pots (one <strong>of</strong> each <strong>of</strong> the four<br />

varieties) were inoculated with a<br />

suspension derived from each field<br />

plot by placing them on a turntable<br />

at 16 rpm <strong>and</strong> applying 33.3 ml <strong>of</strong><br />

the suspension with a h<strong>and</strong><br />

sprayer. Surfactant (Tween) was<br />

added at the rate <strong>of</strong> one drop per 50<br />

ml <strong>of</strong> spore suspension. After<br />

inoculation, the plants were kept in


a moist chamber (wooden frame<br />

covered with a polyethylene<br />

sheet) for 72 hours. High<br />

humidity (95% or above) was<br />

maintained using a humidifier<br />

(ultrasonic, cool mist type) inside<br />

the moist chamber. The pots were<br />

subsequently returned to a<br />

greenhouse bench in a<br />

r<strong>and</strong>omized complete block<br />

design.<br />

Percent lesion area on the<br />

second leaf from the base <strong>of</strong> each<br />

<strong>of</strong> the 10 plants in each pot was<br />

estimated visually at three weeks<br />

after inoculation. Each <strong>of</strong> two<br />

observers half <strong>of</strong> the plants in<br />

each pot.<br />

% diseased leaf area<br />

% diseased leaf area<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

6-Feb<br />

6-Feb<br />

Madsen (MR)<br />

Stephens (S)<br />

Mixture<br />

27-Feb<br />

20-Mar<br />

Madsen (MR)<br />

W-301 (S)<br />

Mixture<br />

27-Feb<br />

20-Mar<br />

10-Apr<br />

10-Apr<br />

Effects <strong>of</strong> pathogen diversity<br />

on epidemic progression<br />

A field experiment that allowed<br />

us to examine the influence <strong>of</strong><br />

pathogen diversity on epidemic<br />

increase in variety mixtures was<br />

conducted in the 1994-95 season.<br />

These plots were treated in a<br />

manner highly similar to those<br />

described above, <strong>and</strong> have been<br />

discussed in more detail elsewhere<br />

(Zhan et al., 1998). Treatments were<br />

a complete factorial <strong>of</strong> three host<br />

treatments (pure st<strong>and</strong> <strong>of</strong> Madsen,<br />

pure st<strong>and</strong> <strong>of</strong> Stephens, <strong>and</strong> a 1:1<br />

mixture <strong>of</strong> Madsen:Stephens) <strong>and</strong><br />

two inoculation treatments (natural<br />

<strong>and</strong> artificial) in a completely<br />

r<strong>and</strong>om design with three<br />

replications. One set <strong>of</strong> plots was<br />

naturally inoculated, while the<br />

1-May<br />

1-May<br />

22-May<br />

22-May<br />

12-Jun<br />

12-Jun<br />

% diseased leaf area<br />

% diseased leaf area<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

6-Feb<br />

6-Feb<br />

Cashup (MR)<br />

Stephens (S)<br />

Mixture<br />

27-Feb<br />

20-Mar<br />

Cashup (MR)<br />

W-301 (S)<br />

Mixture<br />

27-Feb<br />

20-Mar<br />

Disease Management Using Varietal Mixtures 113<br />

other was sprayed in November<br />

1994 with a mixture <strong>of</strong> 10 isolates to<br />

competitively exclude the highly<br />

diverse inoculum provided by<br />

outside ascospore showers.<br />

Results<br />

The Madsen/Stephens <strong>and</strong><br />

Cashup/Stephens mixtures<br />

suppressed epidemic development<br />

below that <strong>of</strong> the more resistant<br />

component in each mixture (Figure<br />

1). In contrast, the mixtures<br />

containing the variety W-301 as the<br />

susceptible component performed<br />

less well, with the Cashup/W-301<br />

mixture showing no epidemic<br />

suppression relative to the mean <strong>of</strong><br />

the component pure st<strong>and</strong>s<br />

(Figure 1).<br />

Figure 1. Disease progress <strong>of</strong> septoria tritici blotch in naturally-occurring epidemics <strong>of</strong> pure st<strong>and</strong> varieties<br />

<strong>and</strong> 1:1 variety mixtures. MR = moderately resistant <strong>and</strong> S = susceptible.<br />

10-Apr<br />

10-Apr<br />

1-May<br />

1-May<br />

22-May<br />

22-May<br />

12-Jun<br />

12-Jun


Session 6A / Session 6B — C.C. Mundt, C. Cowger, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

114<br />

Greenhouse tests indicated that non-inoculated genotypes<br />

host diversity had a substantial comprising 35-40% <strong>of</strong> the<br />

impact on pathogenicity <strong>of</strong> M. population later in the season. In<br />

graminicola (Table 1). All populations contrast, natural inoculum has been<br />

derived from the mixtures produced estimated to result in hundreds <strong>of</strong><br />

less disease on its component unique genotypes per m<br />

varieties in the greenhouse than the<br />

mean <strong>of</strong> the populations derived<br />

from the pure st<strong>and</strong> components in<br />

the field.<br />

Molecular analyses (Zhan et al.,<br />

1998) have previously shown that the<br />

artificial inoculation used in the<br />

1994/95 season competitively<br />

excluded 97% <strong>of</strong> potential ascospore<br />

infections early in the season, though<br />

immigration <strong>and</strong> sexual<br />

recombination among inoculated<br />

<strong>and</strong>/or naturally occurring<br />

genotypes within plots resulted in<br />

2<br />

(McDonald et al., 1996). For the<br />

artificially-inoculated plots,<br />

epidemic progression for the<br />

Madsen/Stephens mixture was<br />

approximately midway between<br />

the two pure st<strong>and</strong>s throughout the<br />

epidemic (Figure 2). By contrast, in<br />

the naturally-inoculated plots,<br />

epidemic progression in the<br />

mixture began midway between<br />

the pure st<strong>and</strong>s, but became<br />

suppressed to near that <strong>of</strong> the<br />

moderately resistant variety as the<br />

season progressed (Figure 2).<br />

Table 1. Disease severity (percent <strong>of</strong> second leaf<br />

area covered by septoria tritici blotch lesions)<br />

caused by Mycosphaerella graminicola populations<br />

derived from single wheat varieties <strong>and</strong> variety<br />

mixtures in replicated field plots.<br />

Source <strong>of</strong> population<br />

Variety mixture Mixturea Pure st<strong>and</strong>sb Madsen/Stephens 9.5 15.9<br />

Madsen/W-301 9.3 11.6<br />

Cashup/Stephens 10.8 13.4<br />

Cashup/W-301 6.6 9.0<br />

a Mean disease severity for populations derived from a<br />

given mixture when tested separately on each <strong>of</strong> the<br />

component varieties. For example, populations<br />

collected from the Madsen/Stephens mixture in the<br />

field were tested separately on Madsen <strong>and</strong> Stephens<br />

in the greenhouse.<br />

b Mean disease severity for populations derived from the<br />

component pure st<strong>and</strong>s <strong>and</strong> tested on the same variety.<br />

For example, to compare with populations derived from<br />

the Madsen/Stephens mixture, populations collected<br />

from pure st<strong>and</strong>s <strong>of</strong> Madsen in the field were tested on<br />

Madsen in the greenhouse <strong>and</strong> populations derived<br />

from pure st<strong>and</strong>s <strong>of</strong> Stephens in the field were tested<br />

on Stephens in the greenhouse.<br />

percent disease<br />

percent disease<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

25-Jan<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Discussion<br />

Our results indicate that some<br />

variety mixtures suppress epidemic<br />

progression <strong>of</strong> septoria tritici blotch<br />

even in the absence <strong>of</strong> major genes for<br />

resistance. There are at least two<br />

plausible explanations for this result.<br />

First, the models <strong>of</strong> Jeger et al. (1981a)<br />

have shown that, in absence <strong>of</strong> host/<br />

pathogen specificity, mixtures can<br />

decrease, increase, or have no effect<br />

on epidemic progression, depending<br />

on the relative levels <strong>of</strong> sporulation<br />

<strong>and</strong> infection frequency in the<br />

component varieties. Thus,<br />

differences between mixtures <strong>and</strong><br />

pure st<strong>and</strong>s in our experiment might<br />

be explained by differences in<br />

resistance components between<br />

cultivars, though we have not<br />

measured these components.<br />

Artificially inoculated plots<br />

Madsen<br />

Stephens<br />

Stephens/Madsen mixture<br />

15-Mar 4-May 23-Jun<br />

Naturally inoculated plots<br />

Madsen<br />

Stephens<br />

Stephens/Madsen mixture<br />

30<br />

20<br />

10<br />

0<br />

25-Jan 15-Mar 4-May 23-Jun<br />

Figure 2. Disease progress <strong>of</strong> septoria tritici blotch on a<br />

moderately resistant variety (Madsen), a susceptible<br />

variety (Stephens) <strong>and</strong> a 1:1 mixture <strong>of</strong> the two varieties<br />

in epidemics initiated from artificial inoculation with 10<br />

isolates or from naturally occurring inoculum.


A second explanation is that<br />

disruptive selection may reduce the<br />

fitness <strong>of</strong> pathogen populations<br />

that cycle between host genotypes<br />

in mixtures. The disruptive<br />

selection hypothesis is supported<br />

by the reduced fitness <strong>of</strong> M.<br />

graminicola populations derived<br />

from the variety mixtures as<br />

compared to the pure st<strong>and</strong>s. It<br />

should be noted that the mean 27%<br />

fitness reduction <strong>of</strong> populations<br />

from mixtures that was found in<br />

our study is likely an<br />

underestimation, as disease<br />

severity in a monocyclic test<br />

accounts for only infection<br />

efficiency <strong>and</strong> lesion expansion, but<br />

not other fitness components such<br />

as sporulation rate or latent period.<br />

Additional support for the<br />

disruptive selection hypothesis is<br />

that variety mixtures suppressed<br />

epidemic development under<br />

conditions <strong>of</strong> natural inoculation,<br />

but not when artificially inoculated<br />

with a limited number <strong>of</strong> isolates.<br />

This result would be expected<br />

under the disruptive selection<br />

hypothesis, as there would be<br />

insufficient pathogen variation<br />

within the artificially inoculated<br />

plots for disruptive selection to<br />

occur. Such disruptive selection<br />

was earlier suggested to increase<br />

the efficacy <strong>of</strong> barley variety<br />

mixtures for control <strong>of</strong> powdery<br />

mildew (caused by Erysiphe<br />

graminis) (Wolfe et al., 1981) <strong>and</strong> to<br />

reduce the fitness <strong>of</strong> mildew<br />

genotypes that were able to<br />

overcome more than one major<br />

resistance gene in the mixtures<br />

(Chin <strong>and</strong> Wolfe, 1984).<br />

The levels <strong>of</strong> disease<br />

suppression contributed by our<br />

variety mixtures may not be<br />

sufficient as the sole control<br />

practice, but would certainly be<br />

useful in an integrated<br />

management program. In addition,<br />

we are currently more interested in<br />

the potential for varietal<br />

diversification to slow pathogen<br />

evolution towards increased<br />

virulence <strong>and</strong>/or aggressiveness<br />

(sensu V<strong>and</strong>erplank, 1968) than we<br />

are for the immediate effects on<br />

epidemic progression. This is<br />

crucial in the Willamette Valley <strong>of</strong><br />

Oregon, as there is evidence that<br />

even quantitative resistance to M.<br />

graminicola may erode in our<br />

environment (Mundt et al., 1999).<br />

Though not addressed in the<br />

research reported here, major genes<br />

for resistance may contribute<br />

substantially to control <strong>of</strong> septoria<br />

diseases in wheat variety mixtures.<br />

Our earlier studies showed that<br />

mixtures containing both a<br />

susceptible <strong>and</strong> a highly resistant<br />

variety provided greater disease<br />

control than other types <strong>of</strong><br />

mixtures (Mundt et al. 1995). The<br />

subsequent “breakdown” <strong>of</strong> the<br />

high-level resistance in the variety<br />

Gene (Cowger et al., this volume)<br />

might be seen to increase the<br />

overall level <strong>of</strong> disease in such<br />

mixtures. On the other h<strong>and</strong>,<br />

mixtures <strong>of</strong> a virulent <strong>and</strong> avirulent<br />

pathogen isolate have been shown<br />

to reduce pycnidial production <strong>of</strong><br />

M. graminicola very substantially<br />

(Halperin et al., 1996), which could<br />

greatly increase the value <strong>of</strong><br />

“defeated” major genes in variety<br />

mixtures.<br />

Disease Management Using Varietal Mixtures 115<br />

In summary, there appears to be<br />

some potential for variety mixtures<br />

to suppress septoria blotch<br />

epidemics <strong>and</strong> to slow pathogen<br />

evolution. Currently available data<br />

are very scant, however, <strong>and</strong><br />

substantially greater field<br />

evaluation is required to determine<br />

the short- <strong>and</strong> long-term value <strong>of</strong><br />

the diversity approach for control<br />

<strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong> diseases.<br />

References<br />

Ahmed, H.U., Mundt, C.C., <strong>and</strong><br />

Coakley, S.M. 1995. Host-pathogen<br />

relationship <strong>of</strong> geographically<br />

diverse isolates <strong>of</strong> <strong>Septoria</strong> tritici<br />

<strong>and</strong> wheat cultivars. Plant Pathol.<br />

44:838-847.<br />

Ahmed, H.U., Mundt, C.C., H<strong>of</strong>fer,<br />

M.E., <strong>and</strong> Coakley, S.M. 1996.<br />

Selective influence <strong>of</strong> wheat<br />

cultivars on pathogenicity <strong>of</strong><br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici).<br />

Phytopathology 86:454-458.<br />

Chin, K.M., <strong>and</strong> Wolfe, M.S. 1984.<br />

Selection on Erysiphe graminis in<br />

pure <strong>and</strong> mixed st<strong>and</strong>s <strong>of</strong> barley.<br />

Plant Pathol. 33:535-546.<br />

DiLeone, J.A., Karow, R.S., Coakley,<br />

S.M., <strong>and</strong> Mundt, C.C. 1997.<br />

Biology <strong>and</strong> Control <strong>of</strong> <strong>Septoria</strong><br />

<strong>Diseases</strong> <strong>of</strong> Winter Wheat in Western<br />

Oregon. Oregon State Univ. Ext.<br />

Serv. Crop Sci. Bull. 109. 17 pp.<br />

Fitt, B.D.L., <strong>and</strong> McCartney, H.A.<br />

1986. Spore dispersal in relation to<br />

epidemic models. Pages 311-345 in:<br />

Plant Disease Epidemiology, Vol. 1.<br />

K.J. Leonard <strong>and</strong> W.E. Fry, eds.<br />

McGraw-Hill, New York.<br />

Garrett, K.A., <strong>and</strong> Mundt, C.C. 1999.<br />

Epidemiology in mixed host<br />

populations (mini-review).<br />

Phytopathology 89:accepted<br />

pending revisions.<br />

Halperin, T., Schuster, S., Pnini-<br />

Cohen, S., Zilberstein, A., <strong>and</strong><br />

Eyal, Z. 1996. The suppression <strong>of</strong><br />

pycnidial production on wheat<br />

seedlings following sequential<br />

inoculation by isolates <strong>of</strong> <strong>Septoria</strong><br />

tritici. Phytopathology 86:728-732.


Session 6A / Session 6B — C.C. Mundt, C. Cowger, <strong>and</strong> M.E. H<strong>of</strong>fer<br />

116<br />

Jeger, M.J., Griffiths, E., <strong>and</strong> Jones,<br />

D.G. 1981a. Disease progress <strong>of</strong><br />

non-specialised fungal pathogens<br />

in intraspecific mixed st<strong>and</strong>s <strong>of</strong><br />

cereal cultivars. I. Models. Ann.<br />

Appl. Biol. 98:187-198.<br />

Jeger, M.J., Jones, D.G., <strong>and</strong> Griffiths,<br />

E. 1981b. Disease progress <strong>of</strong> nonspecialised<br />

fungal pathogens in<br />

intraspecific mixed st<strong>and</strong>s <strong>of</strong> cereal<br />

cultivars. II. Field experiments.<br />

Ann. Appl. Biol. 98:199-210.<br />

Kema, G.H.J., Sayoud, R., <strong>and</strong> van<br />

Silfhout, C.H. 1996. Genetic<br />

variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem. II. Analysis <strong>of</strong><br />

interactions between pathogen<br />

isolates <strong>and</strong> host cultivars.<br />

Phytopathology 86:213-220.<br />

Kema, G.H.J., <strong>and</strong> van Silfhout, C.H.<br />

1997. Genetic variation for<br />

virulence <strong>and</strong> resistance in the<br />

wheat-Mycosphaerella graminicola<br />

pathosystem. III. Comparative<br />

seedling <strong>and</strong> adult plant<br />

experiments. Phytopathology<br />

86:213-220.<br />

McDonald, B.A., Mundt, C.C., <strong>and</strong><br />

Chen, R.S. 1996. The role <strong>of</strong><br />

selection on the genetic structure<br />

<strong>of</strong> pathogen populations: Evidence<br />

from field experiments with<br />

Mycosphaerella graminicola.<br />

Euphytica 92:73-80.<br />

Mundt, C.C. 1994. Use <strong>of</strong> host genetic<br />

diversity to control cereal diseases:<br />

Implications for rice blast. Pages<br />

293-307 in: Rice Blast Disease. S.A.<br />

Leong, R.S. Zeigler, <strong>and</strong> P.S. Teng,<br />

eds. CAB International,<br />

Cambridge, <strong>and</strong> the International<br />

Rice Research Institute, Manila.<br />

Mundt, C.C., <strong>and</strong> Leonard, K.J. 1986.<br />

Analysis <strong>of</strong> factors affecting<br />

disease increase <strong>and</strong> spread in<br />

mixtures <strong>of</strong> immune <strong>and</strong><br />

susceptible plants in computersimulated<br />

epidemics.<br />

Phytopathology 76:832-840.<br />

Mundt, C.C., H<strong>of</strong>fer, M.E., Ahmed,<br />

H.U., Coakley, S.M., DiLeone, J.A.,<br />

<strong>and</strong> Cowger, C. 1999. Population<br />

genetics <strong>and</strong> host resistance. Pages<br />

115-130 in: <strong>Septoria</strong> in <strong>Cereals</strong>: a<br />

Study <strong>of</strong> Pathosystems. Lucas, J.A.,<br />

Bowyer, P., Anderson, H.M., eds.<br />

CABI Publishing, Wallingford, UK.<br />

Mundt, C.C., Brophy, L.S., <strong>and</strong><br />

Schmitt, M.E. 1995. Choosing crop<br />

cultivars <strong>and</strong> mixtures under high<br />

versus low disease pressure: A<br />

case study with wheat. Crop Prot.<br />

14:509-515.<br />

V<strong>and</strong>erplank, J.E. 1968. Disease<br />

Resistance in Plants. Academic<br />

Press, New York. 206 pp.<br />

Wolfe, M.S. 1985. The current status<br />

<strong>and</strong> prospects <strong>of</strong> multiline<br />

cultivars <strong>and</strong> variety mixtures for<br />

disease resistance. Annu. Rev.<br />

Phytopathol. 23:251-273.<br />

Wolfe, M.S., Barrett, J.A., <strong>and</strong> Jenkins,<br />

J.E.E. 1981. The use <strong>of</strong> cultivar<br />

mixtures for disease control. Pages<br />

73-80 in: J.F. Jenkyn <strong>and</strong> R.T.<br />

Plumb, eds. Strategies for the<br />

Control <strong>of</strong> Cereal Disease. Blackwell,<br />

Oxford.<br />

Zhan, J., Mundt, C.C., <strong>and</strong> McDonald,<br />

B.A. 1998. Measuring immigration<br />

<strong>and</strong> sexual reproduction in field<br />

populations <strong>of</strong> Mycosphaerella<br />

graminicola. Phytopathology<br />

88:1330-1337.


Session 6C: Breeding for Disease Resistance<br />

Breeding for Resistance to the <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

Blights <strong>of</strong> Wheat<br />

M. van Ginkel <strong>and</strong> S. Rajaram<br />

Wheat Program, <strong>CIMMYT</strong>, El Batan, Mexico<br />

Abstract<br />

Genetic resistance remains the first line <strong>of</strong> defense against the septoria foliar blights, especially in developing countries.<br />

Resistance genes with major or minor effects may be either recessive or dominant. A few genes may be enough to confer<br />

resistance, as in the case <strong>of</strong> partial resistance. Additive gene effects contribute more to resistance than dominance effects.<br />

Among normally maturing semidwarf wheats possessing resistance, those carrying Rht2 may be more resistant to <strong>Septoria</strong><br />

tritici than those with Rht1. In the case <strong>of</strong> <strong>Stagonospora</strong> nodorum, resistance <strong>of</strong> the flag leaf <strong>and</strong> <strong>of</strong> the spike may be at least<br />

partly under separate genetic control.<br />

Five confounding factors complicate selection for resistance: 1) maturity <strong>and</strong> plant height affect the expression <strong>of</strong><br />

resistance; 2) the relationship between seedling <strong>and</strong> adult plant responses is highly inconsistent; 3) the correlation between<br />

disease response <strong>and</strong> yield loss is very variable; 4) there is interaction among fungal isolates on the leaf surface; <strong>and</strong> 5) it is<br />

essential to determine the implications <strong>of</strong> the existence <strong>of</strong> races for resistance breeding.<br />

Durability <strong>of</strong> resistance across years <strong>and</strong> locations is a particular concern, <strong>and</strong> examples <strong>of</strong> both erosion <strong>of</strong> resistance <strong>and</strong><br />

stable resistance have been put forward. There is clear pro<strong>of</strong> <strong>of</strong> differential variety by isolate interactions at the seedling stage<br />

<strong>and</strong> on adult plants in the field. However, almost without exception, the size <strong>of</strong> the interaction component is about a<br />

magnitude smaller than that <strong>of</strong> the main effects due to varieties <strong>and</strong> isolates. Recent molecular experiments on the genetic<br />

structure <strong>of</strong> S. tritici populations have found a lack <strong>of</strong> adaptation to the host genotype.<br />

The methodology applied at <strong>CIMMYT</strong> for pyramiding resistance genes in wheat includes utilizing proven resistance<br />

sources as parents, selecting for resistance using a shuttle breeding program between contrasting locations, <strong>and</strong> confirming<br />

resistance through global multilocation testing. The best resulting lines are fed back into the crossing program as parents. The<br />

first genepool exploited for resistance to S. tritici by <strong>CIMMYT</strong> breeders originated in Brazil <strong>and</strong> Argentina, followed by<br />

winter wheats <strong>and</strong>, subsequently, Chinese materials. New sources <strong>of</strong> resistance, such as highly promising synthetic wheats,<br />

are currently being exploited in combination with the previous sources.<br />

Globally the septoria diseases<br />

have increased in importance over<br />

the past decades, <strong>and</strong> as Walter<br />

Spurgeon Beach wrote in 1919:<br />

“The genus is the more worthy <strong>of</strong><br />

study on account <strong>of</strong> its high<br />

economic importance.” The two<br />

main septoria blights addressed in<br />

this paper are caused by <strong>Septoria</strong><br />

tritici <strong>and</strong> <strong>Stagonospora</strong> nodorum.<br />

However, inferences can be drawn<br />

for the septoria blights in other<br />

cereals, such as <strong>Septoria</strong> passerinii on<br />

barley. Several excellent overviews<br />

have been written on the genetics <strong>of</strong><br />

resistance to the septoria diseases <strong>of</strong><br />

cereals <strong>and</strong> on how to breed for<br />

such resistance (Shipton et al., 1971;<br />

King et al., 1983; Nelson <strong>and</strong><br />

Marshall, 1990; Lucas et al., 1999;<br />

Eyal, 1999). Most <strong>of</strong> those reviews<br />

were published 10 years or more<br />

ago, but a great deal <strong>of</strong> the<br />

information they contain is still<br />

correct <strong>and</strong> relevant today.<br />

117<br />

This paper emphasizes some <strong>of</strong><br />

the more recent literature (post 1990)<br />

on the genetics <strong>and</strong> breeding <strong>of</strong><br />

resistance, <strong>and</strong> how these new<br />

insights may be applied in developing<br />

varieties that express durable<br />

resistance to the septoria blights.


118<br />

Session 6C — M. van Ginkel <strong>and</strong> S. Rajaram<br />

The Need for Resistance<br />

In most wheat production<br />

environments, although not in all,<br />

genetic resistance is the most<br />

economical approach to control<br />

fungal diseases. There are, however,<br />

two other control methods—cultural<br />

<strong>and</strong> chemical—that may be utilized.<br />

In the case <strong>of</strong> the septoria foliar<br />

blights, crop residues play a critical<br />

role in the over-seasoning <strong>of</strong> the<br />

pathogen <strong>and</strong> in supplying the<br />

primary inoculum for the<br />

subsequent cropping cycle. In recent<br />

years zero <strong>and</strong> reduced tillage<br />

practices have been gaining<br />

popularity for reasons ranging from<br />

the economics <strong>of</strong> production to<br />

protection <strong>of</strong> the environment, <strong>and</strong><br />

the trend is likely to become<br />

stronger in the near term. Hence<br />

cultural control methods involving<br />

specific residue removal seem less <strong>of</strong><br />

an option than a few decades ago.<br />

However, in the more distant future,<br />

we may once again see a<br />

contraction, particularly in the use<br />

<strong>of</strong> the zero tillage option, as the<br />

problems <strong>of</strong> excessive fungicide use<br />

<strong>and</strong> disease buildup both above <strong>and</strong><br />

below the soil become more<br />

intractable.<br />

Chemical control will always<br />

need to be available as an option. In<br />

particular when conditions are<br />

unexpectedly conducive to<br />

disease—for example, due to<br />

excessive rainfall—a wellresearched<br />

chemical option must be<br />

close at h<strong>and</strong>. Under such<br />

circumstances chemical control may<br />

<strong>of</strong>ten be justified because it can<br />

avoid production losses <strong>of</strong> great<br />

economic value.<br />

Considering the above<br />

mentioned control options,<br />

breeding for resistance must<br />

remain the first line <strong>of</strong> defense,<br />

especially in developing<br />

countries. In such countries other<br />

options, while attractive in<br />

theory, <strong>of</strong>ten cannot be<br />

implemented in a timely fashion<br />

<strong>and</strong>/or the resources needed for<br />

their development <strong>and</strong> largescale<br />

application are lacking,<br />

even in an emergency.<br />

Where resistance is not<br />

effective, tolerance can be sought<br />

(McKendry <strong>and</strong> Henke, 1994a).<br />

Despite more than 25 years <strong>of</strong><br />

interest in tolerance to the<br />

septoria blights, its mechanism(s)<br />

in general remains vague<br />

(Zuckerman et al., 1997). There is,<br />

however, little doubt that<br />

tolerance is widely available <strong>and</strong><br />

operational in modern-day<br />

germplasm.<br />

Sources <strong>of</strong> Resistance<br />

Many sources <strong>of</strong> resistance to<br />

the septoria foliar blights,<br />

including some wild relatives <strong>of</strong><br />

wheat, have been studied over<br />

the years. The interest in “alien<br />

sources” is not new; in the early<br />

part <strong>of</strong> this century, wild relatives<br />

resistant to the <strong>Septoria</strong> spp. were<br />

already being tested <strong>and</strong><br />

identified (Beach, 1919; Mackie,<br />

1929). Some <strong>of</strong> those sources were<br />

reported in the <strong>of</strong>ficial literature,<br />

while others were used in<br />

breeding programs <strong>and</strong> reported<br />

less in written form.<br />

Although grouping the<br />

sources <strong>of</strong> resistance based on<br />

origin or wheat class is to a<br />

certain extent arbitrary, commonly<br />

used classifications are: derived<br />

from South American (in particular<br />

Brazilian) sources, winter wheats,<br />

Chinese origin, <strong>and</strong> wild relatives<br />

<strong>of</strong> wheat. Some <strong>of</strong> the major<br />

sources confirmed by several<br />

authors are listed in Table 1.<br />

Table 1. Sources <strong>of</strong> resistance to <strong>Septoria</strong> tritici<br />

<strong>and</strong> <strong>Stagonospora</strong> nodorum confirmed by several<br />

authors. Generally only one key reference is given.<br />

<strong>Septoria</strong> tritici Reference<br />

Anza Wilson (1994)<br />

Bezostaya 1 Danon et al. (1982)<br />

Bobwhite Gilchrist et al. (1995)<br />

Bulgaria 88 Rillo <strong>and</strong> Caldwell (1966)<br />

Carifen Lee & Gough (1984)<br />

Colotana Danon et al. (1982)<br />

Fortaleza 1 Danon et al. (1982)<br />

Israel 493 Wilson (1979)<br />

Milan Gilchrist et al. (1995)<br />

Nabob Narvaez & Caldwell (1957)<br />

Oasis Danon et al. (1982)<br />

Seabreeze Rosielle & Brown (1979)<br />

Sheridan<br />

Synthetic wheats<br />

Danon et al. (1982)<br />

(some) May & Lagudah (1992)<br />

Tadinia Somasco et al. (1996)<br />

Tinamou Gilchrist et al. (1995)<br />

T. dicoccon Gilchrist & Skovm<strong>and</strong> (1995)<br />

T. speltoides McKendry & Henke (1994)<br />

T. tauschii Appels & Lagudah (1990); May<br />

& Lagudah (1992); McKendry<br />

& Henke (1994b)<br />

Veranopolis Rosielle & Brown (1979)<br />

Vilmorin<br />

<strong>Stagonospora</strong><br />

nodorum<br />

Gough & Smith (1985)<br />

Aegilops longissima Ecker et al. (1990a)<br />

Atlas 66 Kleijer et al. (1977)<br />

Blueboy II Nelson (1980)<br />

Cotipora Bostwick et al. (1993)<br />

Frondoso Mullaney et al. (1982)<br />

Fronthatch Mullaney et al. (1982)<br />

Oasis Nelson (1980)<br />

T. monococcum Ma & Hughes (1993)<br />

T. speltoides Ecker et al. (1990b)<br />

T. tauschii Ma & Hughes (1993)<br />

T. timopheevii Ma & Hughes (1993)<br />

Inheritance <strong>of</strong><br />

Resistance<br />

Resistance genes having major<br />

effects have been identified in<br />

wheat (Nelson <strong>and</strong> Marshall, 1990).<br />

They have been found to transmit


their resistance in either a recessive<br />

or dominant fashion (Eyal, 1999).<br />

However, in most cases, resistance<br />

appears dominant, with the F1<br />

from a cross between a resistant<br />

<strong>and</strong> a susceptible parent expressing<br />

an intermediate level <strong>of</strong> resistance<br />

that is more similar to that <strong>of</strong> the<br />

resistant parent. A few genes may<br />

be enough to confer resistance that<br />

will hold up in farmers’ fields<br />

(Dubin <strong>and</strong> Rajaram, 1996). While<br />

heritabilities tend to be only <strong>of</strong><br />

moderate magnitude, progress in<br />

breeding for resistance is evident.<br />

Although the number <strong>of</strong> genes<br />

available may be quite high,<br />

accumulating a few key ones may<br />

be sufficient to achieve resistance.<br />

As has been shown for durable<br />

resistance to leaf rust (Singh et al.,<br />

1991), several <strong>of</strong> the alleged<br />

components <strong>of</strong> partial resistance to<br />

S. tritici may also be controlled by<br />

only one or a just a few genes<br />

(Jlibene <strong>and</strong> El Bouami, 1995). It<br />

would seem that those components<br />

that are genetically different could<br />

be combined into the same genetic<br />

background by crossing. Once<br />

available, molecular markers will<br />

be <strong>of</strong> tremendous use for<br />

accumulating resistance to such<br />

environmentally sensitive diseases.<br />

The expression <strong>of</strong> S. tritici<br />

resistance was studied in T. tauschii<br />

accessions, synthetic wheats<br />

derived from them, plus<br />

derivatives <strong>of</strong> synthetic wheats<br />

crossed to common wheats. The<br />

results obtained all pointed to<br />

inheritance based on one or a few<br />

dominant genes (Appels <strong>and</strong><br />

Lagudah, 1990; May <strong>and</strong><br />

Lagudah, 1992).<br />

Chromosome 5D <strong>of</strong> T. tauschii<br />

contributed a high level <strong>of</strong> resistance<br />

to S. nodorum in a synthetic cross<br />

with a T. dicoccum line <strong>and</strong> in<br />

derived substitution lines<br />

(Nicholson et al., 1993).<br />

Chromosomes 5D, 3D, <strong>and</strong> 7D<br />

contributed resistance in decreasing<br />

order <strong>of</strong> importance. Aegilops<br />

longissima <strong>and</strong> Ae. speltoides were<br />

shown to contribute S. nodorum<br />

resistance based on two to four<br />

partially dominant to overdominant<br />

genes (Ecker et al., 1990a<br />

<strong>and</strong> b). Triticum timopheevii<br />

transmitted one S. nodorum<br />

resistance gene located on<br />

chromosome 3A to its progeny from<br />

a cross with a susceptible durum<br />

parent (Ma <strong>and</strong> Hughes, 1995).<br />

In most published accounts <strong>of</strong><br />

quantitative analyses, additive gene<br />

effects or general combining ability<br />

(GCA) effects contributed more to<br />

resistance than dominance or special<br />

combining ability (SCA) effects (van<br />

Ginkel <strong>and</strong> Scharen, 1987; Bruno<br />

<strong>and</strong> Nelson, 1990; Danon <strong>and</strong> Eyal,<br />

1990; Wilkinson et al., 1990; Jonsson,<br />

1991; Jlibene et al., 1994). However,<br />

significant non-additive effects were<br />

<strong>of</strong>ten identified in the same studies.<br />

SCA effects tend to be an order <strong>of</strong><br />

magnitude weaker than GCA<br />

effects. Indirectly Nelson <strong>and</strong> Fang<br />

(1994) confirmed these conclusions<br />

when they observed an absence <strong>of</strong><br />

heterosis for components <strong>of</strong> S.<br />

nodorum resistance, indicating a<br />

general lack <strong>of</strong> over-dominance<br />

effects that enhance resistance.<br />

Jlibene et al. (1994) found a small<br />

cytoplasmic effect in a few <strong>of</strong><br />

their crosses.<br />

Breeding for Resistance to the <strong>Septoria</strong>/<strong>Stagonospora</strong> Blights <strong>of</strong> Wheat 119<br />

Triticales were found not to be<br />

any more resistant than most<br />

wheats, <strong>and</strong> a range <strong>of</strong> reactions to<br />

both S. tritici <strong>and</strong> S. nodorum was<br />

observed (Scharen et al., 1990).<br />

Tritordeum amphiploids (doubled<br />

derivatives from crosses between<br />

Hordeum chilense <strong>and</strong> Triticum spp.)<br />

expressed the S. tritici resistance <strong>of</strong><br />

the barley parent both as seedlings<br />

in greenhouse tests <strong>and</strong> adult<br />

plants in field trials (Rubiales et<br />

al., 1992).<br />

Tolerance to S. nodorum was<br />

shown to be a mechanism that is<br />

genetically different from (partial)<br />

resistance <strong>and</strong> involves several<br />

chromosomes (Rapilly et al., 1988).<br />

Distinct characters were identified<br />

within each <strong>of</strong> these two defense<br />

mechanisms.<br />

Although the introduction <strong>of</strong><br />

alien cytoplasm from wild wheat<br />

relatives reduced partial resistance<br />

somewhat, it did increase tolerance<br />

as measured by reduced yield<br />

losses (Keane <strong>and</strong> Jones, 1990).<br />

Agronomic Traits<br />

<strong>and</strong> Resistance<br />

After a period in the 1960s <strong>and</strong><br />

1970s <strong>of</strong> seemingly increased levels<br />

<strong>of</strong> disease susceptibility in the<br />

semidwarf wheats (Santiago,<br />

1970), modern semidwarf<br />

germplasm carrying high levels <strong>of</strong><br />

resistance was subsequently<br />

identified. Many are now widely<br />

grown as commercial varieties.<br />

As in previous studies,<br />

Camacho-Casas et al. (1995)<br />

showed that it is possible to breed<br />

normally maturing semidwarf<br />

wheats if proper disease conditions


120<br />

Session 6C — M. van Ginkel <strong>and</strong> S. Rajaram<br />

are created during the selection<br />

process <strong>and</strong> attention is paid to<br />

achieving the desired maturity.<br />

Nevertheless, tall <strong>and</strong>/or late<br />

versions <strong>of</strong> similar isogenic lines<br />

tend to be less affected by the<br />

septoria/stagonospora blights than<br />

their shorter, earlier sister lines.<br />

The roles in conferring S. tritici<br />

resistance <strong>of</strong> the two most common<br />

dwarfing genes may not be the<br />

same (Baltazar et al., 1990). In the<br />

four crosses studied, the Rht2<br />

dwarfing gene tended to be<br />

associated more <strong>of</strong>ten with<br />

increased levels <strong>of</strong> resistance.<br />

Plants carrying Rht1 were more<br />

susceptible.<br />

In the case <strong>of</strong> S. nodorum,<br />

resistance <strong>of</strong> the flag leaf <strong>and</strong> <strong>of</strong> the<br />

spike may be at least partly under<br />

separate genetic control (Rosielle<br />

<strong>and</strong> Brown, 1980; Bostwick et al.,<br />

1993; Hu et al., 1996). Resistance <strong>of</strong><br />

the flag leaf was found to be coded<br />

for by genes on chromosomes 3A,<br />

4A, <strong>and</strong> 3B, while spike resistance<br />

was located on the same<br />

chromosomes <strong>and</strong> on 7A (Hu et al.,<br />

1996).<br />

Determination<br />

<strong>of</strong> Resistance<br />

One <strong>of</strong> the most confounding<br />

factors in determining resistance to<br />

the septoria blights has been <strong>and</strong><br />

continues to be the interaction<br />

between resistance <strong>and</strong> maturity<br />

<strong>and</strong> plant height, as mentioned<br />

above. Parlevliet (1990) describes<br />

how the ranking <strong>of</strong> resistance may<br />

even change dramatically, once<br />

these factors have been<br />

corrected for.<br />

Deviation from regression was<br />

proposed by van Beuningen <strong>and</strong><br />

Kohli (1990) to correct for the<br />

confounding factors <strong>of</strong> heading <strong>and</strong><br />

plant height. Loughman et al.<br />

(1994a) proposed a numerical<br />

classification approach that creates<br />

clusters <strong>of</strong> similar entries based on<br />

response patterns including disease<br />

reaction, maturity, <strong>and</strong> height. If<br />

proper checks are included, cluster<br />

analysis helps to identify resistant<br />

material.<br />

A second confounding factor in<br />

quantifying resistance is the<br />

relationship between seedling <strong>and</strong><br />

adult plant responses. Somasco et<br />

al. (1996) showed that long-term<br />

resistance to S. tritici, such as that<br />

found in Tadinia (derived from the<br />

winter wheat Tadorna), was<br />

strongly expressed in both<br />

seedlings <strong>and</strong> adult plants. Kema<br />

<strong>and</strong> van Silfhout (1997) showed<br />

that not all isolates respond<br />

similarly to seedling <strong>and</strong> adult<br />

plant infection.<br />

Several studies have found that<br />

resistance to S. nodorum is<br />

expressed differently in seedlings<br />

than in adult plants (Koric, 1988;<br />

Arseniuk et al., 1991). These studies<br />

concluded that while only<br />

preliminary seedling data should<br />

be considered, subsequent adult<br />

plant screening is imperative.<br />

In a set <strong>of</strong> bread wheats <strong>and</strong><br />

durum wheats, seedling response<br />

to a crude toxic extract <strong>of</strong> S. tritici<br />

produced a varietal ranking<br />

different from the adult plant<br />

reaction to field inoculation with a<br />

spore mixture <strong>of</strong> the same fungus<br />

(Harrabi et al., 1995).<br />

However, media containing<br />

crude extracts from grain inoculated<br />

with S. nodorum elicited a differential<br />

response that correlated well with<br />

field resistance or susceptibility <strong>of</strong><br />

the spike (Keller et al., 1994).<br />

Resistant lines showed a higher<br />

percentage <strong>of</strong> embryo-forming callus<br />

in the presence <strong>of</strong> the extract.<br />

A third major confounding factor<br />

is the correlation between disease<br />

response <strong>and</strong> yield loss. Under field<br />

conditions visually observed<br />

symptoms <strong>of</strong> S. nodorum on the flag<br />

leaf or spike were not always well<br />

correlated to yield losses (Tvaruzek<br />

<strong>and</strong> Klem, 1994). Lack <strong>of</strong><br />

consistently high correlations<br />

between S. nodorum disease scores<br />

<strong>and</strong> final yield loss motivated<br />

Walther (1990) to study a more<br />

extensive scoring method. When<br />

flag leaves <strong>of</strong> protected <strong>and</strong> nonprotected<br />

plots were scored on three<br />

dates around the middle <strong>of</strong> the<br />

epidemic, correlations rose to 0.73.<br />

<strong>Stagonospora</strong> nodorum severity<br />

scores on the flag leaf had the<br />

highest correlations with yield loss,<br />

followed by those on the flag leaf -1<br />

(Walther <strong>and</strong> Bohmer, 1992). Spike<br />

infection was poorly correlated with<br />

yield. Heritability <strong>of</strong> disease<br />

variables could be further increased<br />

if plant height <strong>and</strong> maturity were<br />

corrected for, with heritability values<br />

for weighted assessments<br />

reaching 0.89.<br />

The practice <strong>of</strong> using detached<br />

leaves to determine resistance<br />

appears to have diminished in the<br />

past decade. Most work is now<br />

carried out on seedlings <strong>and</strong> adult<br />

plants in the greenhouse or the field.


Durability <strong>of</strong> Resistance<br />

Durability <strong>of</strong> S. nodorum<br />

resistance within a crop cycle<br />

decreased with leaf age, <strong>and</strong> the<br />

older lower leaves proved more<br />

susceptible than the younger upper<br />

leaves (Jonsson, 1991). Due to this<br />

effect, late maturing germplasm is<br />

<strong>of</strong>ten wrongly considered more<br />

resistant than early maturing lines.<br />

Thus it is crucial to measure<br />

resistance at the same adult<br />

growth stage.<br />

More debated than durability<br />

within a crop cycle is the issue <strong>of</strong><br />

durability <strong>of</strong> resistance across years<br />

<strong>and</strong> locations. Johnson in his 1992<br />

general review <strong>of</strong> breeding for<br />

disease resistance highlights the<br />

difficulty <strong>of</strong> establishing the<br />

presence <strong>of</strong> differential hostpathogen<br />

interaction in the septoria<br />

foliar blights. He compares data<br />

gathered by the Eyal group (Eyal et<br />

al., 1973) with those <strong>of</strong> van Ginkel<br />

(van Ginkel <strong>and</strong> Scharen, 1985) <strong>and</strong><br />

concludes that even the<br />

interpretation <strong>of</strong> quite similar data<br />

may differ. In practice, he<br />

concludes, no unequivocal demise<br />

<strong>of</strong> varieties due to truly differential<br />

races <strong>of</strong> the septoria foliar blights<br />

has been reported.<br />

The increase <strong>of</strong> disease severity<br />

over time on the same variety has<br />

been noted in the Netherl<strong>and</strong>s,<br />

Australia, Israel (Jorgensen <strong>and</strong><br />

Smedegaard-Petersen, 1999), <strong>and</strong><br />

the USA (Mundt et al., 1999).<br />

However, it was not unequivocally<br />

shown whether this was due to an<br />

evolution <strong>of</strong> aggressiveness in the<br />

pathogen population or due to<br />

differential host-pathogen<br />

interaction based on the evolution<br />

<strong>of</strong> new virulence patterns. Mundt et<br />

al. (1999) conclude that the<br />

cultivation <strong>of</strong> susceptible hosts will<br />

result in selection for<br />

aggressiveness. In Germany several<br />

varieties were shown to retain their<br />

resistance to isolates <strong>of</strong> S. nodorum<br />

collected annually from all over the<br />

former East Germany for more than<br />

10 years, except for small variations<br />

over years due to specific weather<br />

conditions (Walther, 1993).<br />

Clear pro<strong>of</strong> <strong>of</strong> differential<br />

variety by isolate interaction at the<br />

seedling stage was presented by<br />

Kema et al. (1996). However, as in<br />

previous studies by this group <strong>and</strong><br />

others, the size <strong>of</strong> the interaction<br />

component was about a magnitude<br />

smaller than that <strong>of</strong> the main effects<br />

due to varieties <strong>and</strong> isolates. In a<br />

later monocyclic field experiment,<br />

Kema <strong>and</strong> van Silfhout (1997)<br />

applied isolates specifically selected<br />

for their virulence differences at the<br />

seedling level, to adult plant plots.<br />

Again, small but significant<br />

interactions between isolates <strong>and</strong><br />

varieties were apparent in this<br />

monocyclic situation. The<br />

significant interaction effect was<br />

largely due to two <strong>of</strong> the three<br />

isolates interacting differentially<br />

with two <strong>of</strong> the 22 varieties tested.<br />

Recent work on Kenyan isolates<br />

indicated that variety by isolate<br />

interaction in the field, while<br />

present, was small, with the<br />

interaction component accounting<br />

for only about 5% <strong>of</strong> the main<br />

effects (Arama, 1996).<br />

Breeding for Resistance to the <strong>Septoria</strong>/<strong>Stagonospora</strong> Blights <strong>of</strong> Wheat 121<br />

Recent experiments on the<br />

genetic structure <strong>of</strong> S. tritici<br />

populations in response to host<br />

differences found a total lack <strong>of</strong><br />

evidence for adaptation to the host<br />

genotype (McDonald et al., 1996).<br />

Different hosts had no differential<br />

effects on the dynamics <strong>of</strong> the<br />

pathogen population. On the other<br />

h<strong>and</strong>, Ahmed et al. (1996) found<br />

that susceptible varieties selected<br />

for more aggressive pathogen<br />

populations. However, the<br />

virulence levels <strong>of</strong> the pathogen<br />

isolates were associated with the<br />

variety <strong>of</strong> origin.<br />

Varietal mixtures <strong>of</strong> resistant<br />

<strong>and</strong> susceptible varieties have met<br />

with mixed success. Recent<br />

comparisons by Loughman et al.<br />

(1994b) <strong>of</strong> pure lines <strong>and</strong> mixtures<br />

did not find that mixtures have any<br />

advantage when it comes to<br />

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

Interaction among isolates on<br />

the leaf, which can enhance or<br />

reduce expected disease severities<br />

(Eyal, 1992; Gilchrist <strong>and</strong><br />

Velazquez, 1994), adds another<br />

component to the mix <strong>of</strong> virulence<br />

<strong>and</strong> aggressiveness.<br />

The implications for resistance<br />

breeding <strong>of</strong> the above findings on<br />

variety by isolate interactions<br />

remain to be tested further in real<br />

agricultural settings. However, it<br />

does appear that the response <strong>of</strong><br />

the septoria/stagonospora foliar<br />

blights to their hosts cannot<br />

compare to the highly volatile<br />

population dynamics we see in the<br />

wheat rusts.


122<br />

Session 6C — M. van Ginkel <strong>and</strong> S. Rajaram<br />

It is essential to determine the<br />

practical relevance <strong>of</strong> the presence <strong>of</strong><br />

races <strong>of</strong> the septoria foliar blights<br />

when applying a resistance breeding<br />

strategy. Indeed, the importance <strong>of</strong><br />

this topic cannot be overstated. New<br />

insights into the genetic structure <strong>of</strong><br />

the pathogen population <strong>and</strong> its<br />

dynamics, especially as supported<br />

by biotechnological tools, may<br />

increase our underst<strong>and</strong>ing<br />

significantly in the near term.<br />

Breeding for Resistance<br />

at <strong>CIMMYT</strong><br />

The methodology applied at<br />

<strong>CIMMYT</strong> for breeding for disease<br />

resistance in wheat has been<br />

outlined in previous publications,<br />

including the proceedings <strong>of</strong> several<br />

international septoria workshops<br />

(Mann et al. 1985; van Ginkel <strong>and</strong><br />

Rajaram, 1989; van Ginkel <strong>and</strong><br />

Rajaram, 1993; Gilchrist et al., 1995;<br />

van Ginkel <strong>and</strong> Rajaram, 1995). The<br />

methodology consists <strong>of</strong> three key<br />

components: utilizing proven<br />

resistance sources as parents,<br />

selecting for increased resistance in<br />

a shuttle breeding program, <strong>and</strong><br />

confirming resistance by<br />

multilocation testing. The best<br />

identified lines are fed back into the<br />

crossing program as parents to<br />

further accumulate resistance,<br />

bringing the process full circle.<br />

The materials that provide the<br />

best resistance are made available<br />

by <strong>CIMMYT</strong> to all collaborators<br />

around the world who request<br />

them. Lists <strong>of</strong> such materials are<br />

published periodically (Gilchrist,<br />

1994). Gilchrist et al. (this volume)<br />

have provided a new overview <strong>of</strong><br />

lines found to be consistently<br />

resistant by <strong>CIMMYT</strong> researchers.<br />

Every year <strong>CIMMYT</strong><br />

distributes to collaborators the<br />

newest wheat lines that combine S.<br />

tritici resistance with a full<br />

“agronomic package,” including<br />

adaptation to high rainfall<br />

environments, superior yield, a<br />

certain level <strong>of</strong> resistance to other<br />

prevalent diseases such as the rusts<br />

<strong>and</strong> scab, <strong>and</strong> acceptable grain<br />

quality.<br />

The parental stocks emphasized<br />

in developing these lines have a<br />

relatively long history <strong>of</strong> proven<br />

resistance in many locations<br />

around the world. The first<br />

genepool exploited for resistance to<br />

S. tritici by <strong>CIMMYT</strong> breeders was<br />

that represented by such lines as<br />

IAS20 from Brazil <strong>and</strong> Klein Atlas<br />

from Argentina. The Tinamou line<br />

is an example <strong>of</strong> a widely adapted,<br />

resistant line that was developed<br />

from those resistance sources. Then<br />

followed the winter wheats, in<br />

particular those from the former<br />

USSR, such as Aurora. Examples <strong>of</strong><br />

elite lines emanating from that<br />

work are the group <strong>of</strong> Bobwhite<br />

lines, such as the cultivar<br />

PROINTA Federal <strong>and</strong> the Milan<br />

lines, which have shown strong<br />

disease resistance throughout the<br />

world, including in the hotspots <strong>of</strong><br />

eastern Africa <strong>and</strong> the Southern<br />

Cone <strong>of</strong> South America (Kohli,<br />

1995; Dubin <strong>and</strong> Rajaram, 1996).<br />

Subsequently, Chinese materials,<br />

such as the Ning, Shanghai, <strong>and</strong><br />

Suzhoe series, plus other varieties<br />

from China, were heavily crossed<br />

by <strong>CIMMYT</strong> breeders. A key<br />

resistant line that resulted from<br />

that effort was Catbird. This<br />

history is discussed in more detail<br />

by Gilchrist et al. (this volume).<br />

New sources <strong>of</strong> resistance, such<br />

as highly promising synthetic<br />

wheats, are currently being<br />

exploited. <strong>CIMMYT</strong>’s Wide Crosses<br />

Program produces synthetic wheats<br />

by crossing <strong>CIMMYT</strong> elite durum<br />

wheats to accessions <strong>of</strong> Aegilops<br />

squarrosa (syn. Triticum tauschii <strong>and</strong><br />

Ae. tauschii). The resulting triploid<br />

seedlings are treated with colchicine<br />

to double the number <strong>of</strong><br />

chromosomes, resulting in manmade<br />

(hence “synthetic”) hexaploid<br />

bread wheats. Certain <strong>of</strong> those<br />

wheats have shown remarkable<br />

levels <strong>of</strong> S. tritici resistance, <strong>and</strong> low<br />

severities bordering on immunity<br />

have also been observed. Since<br />

synthetic wheats are relatively easy<br />

to cross with common wheats, their<br />

resistance can be readily<br />

introgressed into agronomically<br />

acceptable plant types, <strong>and</strong><br />

combined with other resistances.<br />

<strong>CIMMYT</strong>’s breeding<br />

methodology emphasizes regularly<br />

exposing segregating materials to<br />

disease epidemics. Since the<br />

breeding program also targets<br />

resistance to several other diseases<br />

besides S. tritici, straw is used as the<br />

source <strong>of</strong> primary inoculum. Straw<br />

residues may carry inoculum for tan<br />

spot, fusarium head scab, fusarium<br />

foliar blight, <strong>and</strong> different species <strong>of</strong><br />

bacteria in addition to the septorias.<br />

In special cases <strong>and</strong> in specific<br />

studies, pure liquid S. tritici spores<br />

are applied.<br />

<strong>CIMMYT</strong> applies a shuttle<br />

breeding methodology in which<br />

materials are alternately grown in a<br />

high rainfall site in the Mexican<br />

highl<strong>and</strong>s (Toluca) <strong>and</strong> an irrigated<br />

site in northwestern Mexico (Cd.<br />

Obregon). As a result, the material


targeted for S. tritici-prone regions<br />

is exposed to the disease 3-4 times<br />

during the segregating phase when<br />

grown in Toluca. During the<br />

selection process, selection<br />

intensity is increased as the<br />

segregating populations are<br />

advanced. At the end <strong>of</strong> the shuttle<br />

breeding process, all homozygous<br />

lines are exposed for two<br />

additional cycles at three sites in<br />

the high rainfall Mexican<br />

highl<strong>and</strong>s: Toluca, Patzcuaro<br />

(Gomez <strong>and</strong> Gonzalez, 1987), <strong>and</strong><br />

El Tigre.<br />

Shuttle breeding is followed by<br />

multilocation testing at key<br />

locations around the world through<br />

a network <strong>of</strong> cooperators. The<br />

global data allow truly outst<strong>and</strong>ing<br />

material (see Gilchrist et al., these<br />

proceedings) to be identified <strong>and</strong><br />

used as new parents in the ongoing<br />

process <strong>of</strong> recombining genetically<br />

different sources <strong>of</strong> resistance, <strong>and</strong><br />

they also <strong>of</strong>fer the prospect <strong>of</strong><br />

combining different resistance<br />

mechanisms.<br />

Future work will concentrate on<br />

combining accumulated resistances<br />

that are based on different genes<br />

<strong>and</strong>/or different resistance<br />

mechanisms, with superior yield,<br />

scab resistance, shattering<br />

tolerance, <strong>and</strong> industrial quality.<br />

Conclusions<br />

Several conclusions in regard to<br />

the impact <strong>of</strong> our increased<br />

underst<strong>and</strong>ing <strong>of</strong> resistance on<br />

related breeding aspects can be<br />

drawn from the combined research<br />

<strong>of</strong> the past decade:<br />

1. There are many sources <strong>of</strong> genetic<br />

resistance available that have<br />

either been shown to be different<br />

or seem to behave differently.<br />

Many <strong>of</strong> those resistances have<br />

proven to be quite stable.<br />

2. Seedling data at best are an<br />

indication <strong>of</strong> adult plant<br />

resistance, <strong>and</strong> testing adult<br />

plants in a field situation appears<br />

crucial.<br />

3. A variety is rarely replaced<br />

solely or primarily because it<br />

succumbs to one <strong>of</strong> the<br />

septoria/stagonospora blights.<br />

4. The relative role <strong>of</strong> virulence <strong>and</strong><br />

aggressiveness in the field may<br />

soon be elucidated using<br />

molecular tools in combination<br />

with the ability to cross among<br />

isolates. The outcome should<br />

have direct impact on breeding<br />

strategy.<br />

5. The diversity <strong>of</strong> isolates at the<br />

field level makes gaining a<br />

deeper underst<strong>and</strong>ing <strong>of</strong><br />

interactions among these isolates<br />

<strong>of</strong> paramount importance. The<br />

breeder needs to take this into<br />

consideration when planning to<br />

artificially inoculate, or rely on<br />

multisite testing.<br />

6. Multilocation confirmation <strong>of</strong><br />

advanced lines remains a<br />

necessity, while the debate on<br />

possible specificity continues.<br />

Breeding for Resistance to the <strong>Septoria</strong>/<strong>Stagonospora</strong> Blights <strong>of</strong> Wheat 123<br />

While several <strong>of</strong> these exciting<br />

issues remain in debate, advances<br />

in breeding for resistance to the<br />

septoria/stagonospora pathogens<br />

have not ceased. Breeders have<br />

continued to produce varieties with<br />

higher yields, better industrial<br />

quality, <strong>and</strong> improved resistance to<br />

multiple diseases.<br />

References<br />

Ahmed, H.U., C.C. Mundt, M.E.<br />

H<strong>of</strong>fer, <strong>and</strong> S.M. Coakley.1996.<br />

Selective influence <strong>of</strong> wheat<br />

cultivars on pathogenicity <strong>of</strong><br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici).<br />

Phytopathology 86:454-458.<br />

Arama, P.F. 1996. Effects <strong>of</strong> Cultivar,<br />

Isolate <strong>and</strong> Environment on<br />

Resistance <strong>of</strong> Wheat to <strong>Septoria</strong><br />

Tritici Blotch in Kenya. Ph.D.<br />

Thesis, Wageningen Agricultural<br />

University, Wageningen, The<br />

Netherl<strong>and</strong>s. 115 pp.<br />

Arseniuk, E., P.M. Fried, H. Winzeler,<br />

<strong>and</strong> H.J. Czembor. 1991.<br />

Comparison <strong>of</strong> resistance <strong>of</strong><br />

triticale, wheat <strong>and</strong> spelt to<br />

septoria nodorum blotch at the<br />

seedling <strong>and</strong> adult plant stages.<br />

Euphytica 55:43-48.<br />

Baltazar, B.M., A.L. Scharen, <strong>and</strong> W.E.<br />

Kronstad. 1990. Association<br />

between dwarfing genes ‘Rht 1 ’ <strong>and</strong><br />

‘Rht 2 ’ <strong>and</strong> resistance to <strong>Septoria</strong><br />

tritici blotch in winter wheat<br />

(Triticum aestivum L. em Thell).<br />

Theor. Appl. Genet. 79:422-426.<br />

Beach, W.S. 1919. Biologic<br />

specialization in the genus <strong>Septoria</strong>.<br />

Ann. J. Bot. 6:1-32.<br />

Bostwick, D.E., H.W. Ohm, <strong>and</strong> G.<br />

Shaner. 1993. Inheritance <strong>of</strong><br />

<strong>Septoria</strong> glume blotch resistance in<br />

wheats. Crop Sci. 33:439-443.<br />

Bruno, H.H., <strong>and</strong> L.R. Nelson. 1990.<br />

Partial resistance to septoria glume<br />

blotch analyzed in winter wheat<br />

seedlings. Crop Sci. 30:54-59.<br />

Camacho-Casas, M.A., W.E. Kronstad,<br />

<strong>and</strong> A.L. Scharen. 1995. <strong>Septoria</strong><br />

tritici resistance <strong>and</strong> associations<br />

with agronomic traits. Crop Sci.<br />

35:971-976.


124<br />

Session 6C — M. van Ginkel <strong>and</strong> S. Rajaram<br />

Danon, T., J.M. Sacks, <strong>and</strong> Z. Eyal.<br />

1982. The relationships among<br />

plant stature, maturity class <strong>and</strong><br />

susceptibility to septoria leaf<br />

blotch <strong>of</strong> wheat. Phytopathology<br />

72:1037-1042.<br />

Danon, T., <strong>and</strong> Z. Eyal. 1990.<br />

Inheritance <strong>of</strong> resistance to two<br />

<strong>Septoria</strong> tritici isolates in spring<br />

<strong>and</strong> winter wheat cultivars.<br />

Euphytica 47:203-214.<br />

Dubin, H.J., <strong>and</strong> S. Rajaram. 1996.<br />

Breeding disease-resistant wheats<br />

for tropical highl<strong>and</strong>s <strong>and</strong><br />

lowl<strong>and</strong>s. Annual Rev. <strong>of</strong><br />

Phytopath. 34:503-526.<br />

Ecker, R., A. Cahaner, <strong>and</strong> A. Dinoor.<br />

1990. The inheritance <strong>of</strong> resistance<br />

to septoria glume blotch. II. The<br />

wild wheat species Aegilops<br />

speltoides. Plant Breeding<br />

104:218-223.<br />

Ecker, R., A. Cahaner, <strong>and</strong> A. Dinoor.<br />

1990. The inheritance <strong>of</strong> resistance<br />

to septoria glume blotch. III. The<br />

wild wheat species Aegilops<br />

longissima. Plant Breeding<br />

104:218-223.<br />

Eyal, Z. 1992. The response <strong>of</strong> fieldinoculated<br />

wheat cultivars to<br />

mixture <strong>of</strong> <strong>Septoria</strong> tritici isolates.<br />

Euphytica 61:25-35.<br />

Eyal, Z. 1999. Breeding for resistance<br />

to <strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong><br />

<strong>Diseases</strong> in Wheat. In: <strong>Septoria</strong> in<br />

<strong>Cereals</strong>: a Study <strong>of</strong> Pathosystems.<br />

Lucas, J.A., Bowyer, P., Anderson,<br />

H.M. (eds.). CABI Publishing,<br />

Wallingford, UK. pp. 332-344.<br />

Gilchrist, L. 1994. New <strong>Septoria</strong> tritici<br />

resistance sources in <strong>CIMMYT</strong><br />

germplasm <strong>and</strong> its incorporation<br />

in the <strong>Septoria</strong> Monitoring<br />

Nursery. In: Proceedings <strong>of</strong> the 4 th<br />

International Workshop on:<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>. E. Arseniuk, T.<br />

Goral, <strong>and</strong> P. Czembor (eds.).<br />

IHAR, Radzikow, Pol<strong>and</strong>.<br />

Hodowla Roslin Aklimatyzacja I<br />

Nasiennictwo (Special edition)<br />

38(3-4):187-190.<br />

Gilchrist, L., <strong>and</strong> B. Skovm<strong>and</strong>. 1995.<br />

Evaluation <strong>of</strong> emmer wheat<br />

(Triticum dicoccon) for resistance to<br />

<strong>Septoria</strong> tritici. In: Proceedings <strong>of</strong> a<br />

<strong>Septoria</strong> tritici Workshop. L.<br />

Gilchrist, M. van Ginkel, A.<br />

McNab, <strong>and</strong> G.H.J. Kema (eds.).<br />

Mexico, D.F.: <strong>CIMMYT</strong>.<br />

pp. 130-134.<br />

Gilchrist, L., M. van Ginkel, A.<br />

McNab, <strong>and</strong> G.H.J. Kema (eds.).<br />

1995. Proceedings <strong>of</strong> a <strong>Septoria</strong><br />

tritici Workshop. Mexico, D.F.:<br />

<strong>CIMMYT</strong>. 157 pp.<br />

Gilchrist, L., <strong>and</strong> C. Velazquez. 1994.<br />

Interaction to <strong>Septoria</strong> tritici isolatewheat<br />

as adult plant under field<br />

conditions. In: Proceedings <strong>of</strong> the<br />

4 th International Workshop on:<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>. E. Arseniuk, T.<br />

Goral, <strong>and</strong> P. Czembor (eds.).<br />

IHAR, Radzikow, Pol<strong>and</strong>. Hodowla<br />

Roslin Aklimatyzacja I<br />

Nasiennictwo (Special edition)<br />

38(3-4):111-114.<br />

Gomez, B.L., <strong>and</strong> R.M. Gonzalez I.<br />

1987. Mejoramiento genetico de<br />

trigos harineros para resistencia a<br />

<strong>Septoria</strong> tritici en el area de<br />

temporal humedo en Mexico. In:<br />

Proceedings <strong>of</strong> the Regional<br />

<strong>Septoria</strong> Conference, <strong>CIMMYT</strong>,<br />

Mexico. pp 42-57.<br />

Gough, F.J., <strong>and</strong> E.L. Smith. 1985. A<br />

genetic analysis <strong>of</strong> Triticum aestivum<br />

‘Vilmorin’ resistance to speckled<br />

leaf blotch <strong>and</strong> pyrenophora tan<br />

spot. In: <strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>: Proc.<br />

<strong>of</strong> the Workshop. A.L. Scharen<br />

(ed.). Bozeman, MT. 2-4 August,<br />

1983. USDA-ARS 12. p. 36.<br />

Harrabi, M., M. Cherif, H. Amara, Z.<br />

Ennaiffer, <strong>and</strong> A. Daaloul. 1995. In<br />

vitro selection for resistance to<br />

<strong>Septoria</strong> tritici in wheat. In:<br />

Proceedings <strong>of</strong> a <strong>Septoria</strong> tritici<br />

Workshop. L. Gilchrist, M. van<br />

Ginkel, A. McNab, <strong>and</strong> G.H.J.<br />

Kema (eds.). Mexico, D.F.:<br />

<strong>CIMMYT</strong>. pp. 109-116.<br />

Hu Xueyi, D. Bostwick, H. Sharma, H.<br />

Ohm, <strong>and</strong> G. Shaner. 1996.<br />

Chromosome <strong>and</strong> chromosomal<br />

arm locations <strong>of</strong> genes for<br />

resistance to septoria glume blotch<br />

in wheat cultivar Cotipora.<br />

Euphytica 91:251-257.<br />

Jlibene, M., <strong>and</strong> F. El Bouami. 1995.<br />

Inheritance <strong>of</strong> partial resistance to<br />

<strong>Septoria</strong> tritici in hexaploid wheat<br />

(Triticum aestivum). In: Proceedings<br />

<strong>of</strong> a <strong>Septoria</strong> tritici Workshop. L.<br />

Gilchrist, M. van Ginkel, A. McNab,<br />

<strong>and</strong> G.H.J. Kema (eds.). Mexico,<br />

D.F.: <strong>CIMMYT</strong>. pp 117-125.<br />

Jlibene M., J.P. Gustafson, <strong>and</strong> S.<br />

Rajaram. 1994. Inheritance <strong>of</strong><br />

resistance to Mycosphaerella<br />

graminicola in hexaploid wheat.<br />

Plant Breeding 112:301-310.<br />

Johnson, R. 1992. Past, present <strong>and</strong><br />

future opportunities in breeding<br />

for disease resistance, with<br />

examples from wheat. Euphytica<br />

63:3-22.<br />

Jonsson, J.O. 1991. Wheat breeding<br />

against facultative pathogens.<br />

Sveriges utsadesforenings Tidskrift<br />

101:89-93.<br />

Jorgensen, H.J.L., <strong>and</strong> V.<br />

Smedegaaard-Peterson. Hostpathogen<br />

interactions in the<br />

<strong>Septoria</strong>-disease complex. In:<br />

<strong>Septoria</strong> in <strong>Cereals</strong>: a Study <strong>of</strong><br />

Pathosystems. Lucas, J.A., Bowyer,<br />

P., Anderson, H.M. (eds.). CABI<br />

Publishing, Wallingford, UK. pp.<br />

131-160.<br />

Keane, E.M., <strong>and</strong> P.W. Jones. Effects <strong>of</strong><br />

alien cytoplasm substitution on the<br />

response <strong>of</strong> wheat cultivars to<br />

<strong>Septoria</strong> nodorum. Ann. Appl. Biol.<br />

117:299-312.<br />

Keller, B., H. Winzeler, M. Winzeler,<br />

<strong>and</strong> P.M. Fried. 1994. Differential<br />

sensitivity <strong>of</strong> wheat embryos<br />

against extracts containing toxins<br />

<strong>of</strong> <strong>Septoria</strong> nodorum: First steps<br />

towards in vitro selection. J.<br />

Phytopathology 141:233-240.<br />

Kema, G. H. J, J. G., Annone, R.<br />

Sayoud, R., C. H. Van Silfhout, M.<br />

Van Ginkel, <strong>and</strong> J. De Bree. 1996.<br />

Genetic variation for virulence <strong>and</strong><br />

resistance in the Wheat-<br />

Mycosphaerella graminicola<br />

pathosystem. I. Interactions<br />

between pathogen isolates <strong>and</strong> host<br />

cultivars. Phytopathology<br />

86:200-212.<br />

Kema, G.H.J., <strong>and</strong> C.H. van Silfhout.<br />

1997. Genetic variation for<br />

virulence <strong>and</strong> resistance in the<br />

wheat-Mycosphaerella graminicola<br />

pathosystem III. Comparative<br />

seedling <strong>and</strong> adult plant<br />

experiments. Phytopathology<br />

87:266-272.<br />

King, J.E., R.J. Cook, <strong>and</strong> S.C. Melville.<br />

1983. A review <strong>of</strong> septoria diseases<br />

in wheat <strong>and</strong> barley. Ann. Appli.<br />

Biol. 103:345-373.<br />

Kleijer, G., A. Bronniman, <strong>and</strong> A.<br />

Fossati. 1977. Chromosomal<br />

location <strong>of</strong> a dominant gene for<br />

resistance at the seedling stage to<br />

<strong>Septoria</strong> nodorum Berk. in the wheat<br />

variety Atlas 66. Z.<br />

Pflanzenzuchtung 78:170-173.


Kohli, M.M. 1995. Resistance to<br />

septoria tritici blotch in Southern<br />

Cone germplasm. In: Proceedings<br />

<strong>of</strong> a <strong>Septoria</strong> tritici Workshop. L.<br />

Gilchrist, M. van Ginkel, A.<br />

McNab, <strong>and</strong> G.H.J. Kema (eds.).<br />

Mexico, D.F.: <strong>CIMMYT</strong>. pp. 62-72.<br />

Koric, B. 1988. Seedling <strong>and</strong> adult<br />

stage screening for <strong>Septoria</strong><br />

nodorum resistance in wheat.<br />

Rachis 7(1,2): 31-32.<br />

Lee, T.S., <strong>and</strong> F.J. Gough. 1984.<br />

Inheritance <strong>of</strong> <strong>Septoria</strong> leaf blotch<br />

(S. tritici) <strong>and</strong> Pyrenophora tan<br />

spot (P. tritici repentis) resistance in<br />

Triticum aestivum cv. Carifen 12.<br />

Plant Dis. 68:848-851.<br />

Loughman, R., R.E. Wilson, K.E.<br />

Basford, R.F. Gilmour, <strong>and</strong> I.H.<br />

DeLacey. 1994a. Numerical<br />

classification <strong>of</strong> cultivar interaction<br />

<strong>of</strong> septoria tritici blotch with<br />

maturity <strong>and</strong> height. In:<br />

Proceedings <strong>of</strong> the 4 th International<br />

Workshop on: <strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>.<br />

E. Arseniuk, T. Goral, <strong>and</strong> P.<br />

Czembor (eds.). IHAR, Radzikow,<br />

Pol<strong>and</strong>. Hodowla Roslin<br />

Aklimatyzacja I Nasiennictwo<br />

(Special edition) 38(3-4):127-132.<br />

Loughman, R., R.E. Wilson, <strong>and</strong> G.J.<br />

Thomas. 1994b. Effect <strong>of</strong> variety<br />

mixtures with complementary<br />

partial septoria resistance on<br />

disease <strong>and</strong> yield <strong>of</strong> wheat. In:<br />

Proceedings <strong>of</strong> the 4 th International<br />

Workshop on: <strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>.<br />

E. Arseniuk, T. Goral, <strong>and</strong> P.<br />

Czembor (eds.). IHAR, Radzikow,<br />

Pol<strong>and</strong>. Hodowla Roslin<br />

Aklimatyzacja I Nasiennictwo<br />

(Special edition) 38(3-4):203-208.<br />

Lucas, J.A., P. Bowden, <strong>and</strong> H.M.<br />

Anderson (eds). 1999. <strong>Septoria</strong> on<br />

<strong>Cereals</strong>: A Study <strong>of</strong> Pathosystems.<br />

CABI. 353 pp.<br />

Ma, H., <strong>and</strong> G.R. Hughes. 1993.<br />

Resistance to septoria nodorum<br />

blotch in several Triticum species.<br />

Euphytica 70:151-157.<br />

Ma, H., <strong>and</strong> G.R. Hughes. 1995.<br />

Genetic control <strong>and</strong> chromosomal<br />

location <strong>of</strong> Triticum timopheevii<br />

derived resistance to septoria<br />

nodorum blotch in durum wheat.<br />

Genome 38:332-338.<br />

Mackie, W.W. 1929. Resistance to<br />

<strong>Septoria</strong> tritici in wheat.<br />

Phytopathology 19:1139-1140.<br />

Mann, C.E., S. Rajaram, <strong>and</strong> R.L.<br />

Villareal. 1985 progress in<br />

breeding for septoria tritici<br />

resistance in semidwarf spring<br />

wheat at <strong>CIMMYT</strong>. In: <strong>Septoria</strong><br />

<strong>of</strong> <strong>Cereals</strong>: Proc. <strong>of</strong> the<br />

Workshop. A.L. Scharen (ed.).<br />

Bozeman, MT. 2-4 August, 1983.<br />

USDA-ARS ARS-12. pp 22-26.<br />

May, C.E., <strong>and</strong> E.S. Lagudah. 1992.<br />

Inheritance in hexaploid wheat<br />

<strong>of</strong> septoria tritici blotch<br />

resistance <strong>and</strong> other<br />

characteristics derived from<br />

Triticum tauschii. Aust. J. Agric.<br />

Res. 43:433-442.<br />

McDonald, B.A., C.C. Mundt, <strong>and</strong><br />

C. Ruey-Shyang. 1996. The role<br />

<strong>of</strong> selection on the genetic<br />

structure <strong>of</strong> pathogen<br />

populations: Evidence from field<br />

experiments with Mycosphaerella<br />

graminicola on wheat. Euphytica<br />

92:73-80.<br />

McKendry, A.L., <strong>and</strong> G.E. Henke.<br />

1994a. Tolerance to septoria<br />

tritici blotch in s<strong>of</strong>t red winter<br />

wheat. Cer. Res. Comm. 22(4):<br />

353-359).<br />

McKendry, A.L., <strong>and</strong> G.E. Henke.<br />

1994b. Evaluation <strong>of</strong> wheat wild<br />

relatives for resistance to<br />

septoria tritici blotch. Crop Sci.<br />

34:1080-1084.<br />

Mullaney, E.J., J.M. Martin, <strong>and</strong> A.L.<br />

Scharen. 1982. Generation mean<br />

analysis to identify <strong>and</strong> partition<br />

the components <strong>of</strong> genetic<br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Euphytica 31:539-545.<br />

Mundt, C.C., M.E. H<strong>of</strong>fer, H.U.<br />

Ahmed, S.M. Coakley, J.A.<br />

DiLeone, <strong>and</strong> C. Cowger. 1999.<br />

Population genetics <strong>and</strong> host<br />

resistance. In: <strong>Septoria</strong> in<br />

<strong>Cereals</strong>: a Study <strong>of</strong><br />

Pathosystems. Lucas, J.A.,<br />

Bowyer, P., Anderson, H.M.<br />

(eds.). CABI Publishing,<br />

Wallingford, UK. pp. 115-130.<br />

Narvaez, I., <strong>and</strong> R.M. Caldwell.<br />

1957. Inheritance <strong>of</strong> resistance to<br />

leaf blotch <strong>of</strong> wheat caused by<br />

<strong>Septoria</strong> tritici. Phytopathology<br />

47:529-530.<br />

Breeding for Resistance to the <strong>Septoria</strong>/<strong>Stagonospora</strong> Blights <strong>of</strong> Wheat 125<br />

Nelson, L.R. 1980. Inheritance <strong>of</strong><br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Crop Sci. 20:447-449.<br />

Nelson, L.R., <strong>and</strong> Xiaobing Fang.<br />

1994. Effect <strong>of</strong> heterosis on <strong>Septoria</strong><br />

nodorum disease level <strong>and</strong> plant<br />

yield in wheat. In: Proceedings <strong>of</strong><br />

the 4 th International Workshop on:<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>. E. Arseniuk, T.<br />

Goral, <strong>and</strong> P. Czembor (eds.).<br />

IHAR, Radzikow, Pol<strong>and</strong>.<br />

Hodowla Roslin Aklimatyzacja I<br />

Nasiennictwo (Special edition)<br />

38(3-4):213-218.<br />

Nelson, L.R., <strong>and</strong> D. Marshall, 1990.<br />

Breeding wheat for resistance to<br />

<strong>Septoria</strong> nodorum <strong>and</strong> <strong>Septoria</strong><br />

tritici. Advances in Agronomy, Vol.<br />

44.<br />

Nicholson, P., H.N. Rezanoor, <strong>and</strong> A.J.<br />

Worl<strong>and</strong>. 1993. Chromosomal<br />

location <strong>of</strong> resistance to <strong>Septoria</strong><br />

nodorum in a synthetic hexaploid<br />

wheat determined by the study <strong>of</strong><br />

chromosomal substitution lines in<br />

‘Chinese Spring’ wheat. Plant<br />

Breeding 110:177-184.<br />

Parlevliet, J.E. 1987. Breeding for<br />

durable resistance to pathogens.<br />

In: Proceedings <strong>of</strong> the 6 th Regional<br />

Wheat Workshop for Eastern,<br />

Central <strong>and</strong> Southern Africa <strong>and</strong><br />

the Indian Ocean. Addis Ababa,<br />

Ethiopia, October 2-6, 1989. D.G.<br />

Tanner, M. van Ginkel, <strong>and</strong> W.<br />

Mwangi (eds.). <strong>CIMMYT</strong>. pp. 14-<br />

27.<br />

Rapilly, F., P. Auriau, H. Richard, <strong>and</strong><br />

C. Depatureaux. 1988. Monosomic<br />

analysis <strong>of</strong> partial resistance <strong>and</strong><br />

tolerance <strong>of</strong> wheat to <strong>Septoria</strong><br />

nodorum. Agronomie 8(9):801-809.<br />

Rillo, A.O., <strong>and</strong> R.M. Caldwell. 1966.<br />

Inheritance <strong>of</strong> resistance to <strong>Septoria</strong><br />

tritici in Triticum aestivum subsp.<br />

vulgare, Bulgaria 88. (Abstr).<br />

Phytopathology 56:897.<br />

Rosielle, A.A., <strong>and</strong> A.G.P. Brown.<br />

1979. Inheritance, heritability <strong>and</strong><br />

breeding behaviour <strong>of</strong> three<br />

sources <strong>of</strong> resistance to <strong>Septoria</strong><br />

tritici in wheat. Euphytica 28:385-<br />

392.<br />

Rosielle, A.A., <strong>and</strong> A.G.P. Brown.<br />

1980. Selection for resistance to<br />

<strong>Septoria</strong> nodorum in wheat.<br />

Euphytica 29:337-346.


126<br />

Session 6C — M. van Ginkel <strong>and</strong> S. Rajaram<br />

Rubiales, D., J. Ballesteros, <strong>and</strong> A.<br />

Martin. 1992. Resistance to <strong>Septoria</strong><br />

tritici in Hordeum chilense x Triticum<br />

spp. amphiploids. Plant Breeding<br />

109:281-286.<br />

Santiago, J.C. 1970. Resultado das<br />

observacoes effectuadas em<br />

Marrocos na Tunisia respeifantes<br />

as doencas e pragas doe cereais,<br />

principalmente em recalcao aos<br />

trigos Mexicanos em 1969 a<br />

expensas das Missoes Americanas<br />

de auxilio technico a Marrocos e<br />

Tunisia. 12. Junho de 1970. Elvas.<br />

Portugal, estacao de<br />

Melhoramento.<br />

Scharen, A.L., E. Arseniuk, W. Sowa,<br />

J. Zimmy, <strong>and</strong> W. Podyma. 1990.<br />

Seedling resistance <strong>of</strong> triticale <strong>and</strong><br />

Triticum spp. germplasm to<br />

<strong>Septoria</strong> nodorum <strong>and</strong> S. tritici. In:<br />

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

International Triticale Symposium,<br />

1-5 October, 1990, Passo Fundo,<br />

Rio Gr<strong>and</strong>e do Sul, Brazil. pp 256-<br />

259.<br />

Shipton, W.A., W.R.J. Boyd, A.A.<br />

Rosielle, <strong>and</strong> B.I. Shearer. 1971. The<br />

common septoria diseases <strong>of</strong><br />

wheat. Bot. Rev. 37:231-262.<br />

Singh, R.P., T.S. Payne, <strong>and</strong> S.<br />

Rajaram. 1991. Characterization <strong>of</strong><br />

variability <strong>and</strong> relationships<br />

among components <strong>of</strong> partial<br />

resistance to leaf rust in <strong>CIMMYT</strong><br />

bread wheats. Theor. Appl. Genet.<br />

82:674-680.<br />

Somasco, O.A., C.O. Qualset, <strong>and</strong><br />

D.G. Gilchrist. 1996. Single-gene<br />

resistance to <strong>Septoria</strong> tritici blotch<br />

in the spring wheat cultivar<br />

‘Tadinia’. Plant Breeding 115:261-<br />

267.<br />

Tvaruzek, L., <strong>and</strong> K. Klem. 1994.<br />

Varieties <strong>and</strong> lines <strong>of</strong> winter wheat<br />

with stable tolerance <strong>and</strong> low yield<br />

loss to <strong>Septoria</strong> nodorum (Berk). Cer.<br />

Res. Comm. 22(4): 369-374.<br />

Van Beuningen, L.T., <strong>and</strong> M.M. Kohli.<br />

1990. Deviation from the<br />

regression <strong>of</strong> infection on heading<br />

<strong>and</strong> height as a measure <strong>of</strong><br />

resistance to septoria tritici blotch<br />

on wheat. Plant Disease 74:488-<br />

493.<br />

Van Ginkel, M., <strong>and</strong> A.L. Scharen.<br />

1987. Generation mean analysis<br />

<strong>and</strong> heritabilities <strong>of</strong> resistance to<br />

<strong>Septoria</strong> tritici in durum wheat.<br />

Phytopathology 77:1629-1633.<br />

Van Ginkel, M., <strong>and</strong> S. Rajaram. 1989.<br />

Breeding for global resistance to<br />

<strong>Septoria</strong> tritici in wheat. In: <strong>Septoria</strong><br />

<strong>of</strong> <strong>Cereals</strong>. P.M. Fried (ed.). July 4-<br />

7, Swiss Federal Research Station<br />

for Agronomy, Zurich,<br />

Switzerl<strong>and</strong>. pp 174-176.<br />

Van Ginkel, M., <strong>and</strong> S. Rajaram. 1993.<br />

Breeding for Durable Resistance to<br />

<strong>Diseases</strong> in Wheat: An<br />

International Perspective. In:<br />

Durability <strong>of</strong> Disease Resistance.<br />

Th. Jacobs <strong>and</strong> J.E. Parlevliet (eds.).<br />

Kluwer Academic Publishers, The<br />

Netherl<strong>and</strong>s. pp. 259-272.<br />

Van Ginkel, M., <strong>and</strong> S. Rajaram. 1995.<br />

Breeding for Resistance to <strong>Septoria</strong><br />

tritici at <strong>CIMMYT</strong>. In: Proceedings<br />

<strong>of</strong> a <strong>Septoria</strong> tritici Workshop. L.<br />

Gilchrist, M. van Ginkel, A.<br />

McNab, <strong>and</strong> G.H.J. Kema (eds.).<br />

Mexico, D.F.: <strong>CIMMYT</strong>. pp. 55-61.<br />

Walther, H. 1990. An improved<br />

assessment procedure for breeding<br />

for resistance to <strong>Septoria</strong> nodorum<br />

in wheat. Plant Breeding 105:<br />

53-61.<br />

Walther, H. Durability <strong>and</strong> stability <strong>of</strong><br />

resistance <strong>of</strong> wheat to <strong>Septoria</strong><br />

nodorum (glume blotch) as<br />

assessed by means <strong>of</strong> disease<br />

progress on flag leaves. In:<br />

Durability <strong>of</strong> Disease Resistance.<br />

Th. Jacobs <strong>and</strong> J.E. Parlevliet (eds.).<br />

Kluwer Academic Publishers, The<br />

Netherl<strong>and</strong>s. p. 354.<br />

Walther, H., <strong>and</strong> M. Bohmer. 1992.<br />

Improved quantitative-genetic<br />

selection in breeding for resistance<br />

to <strong>Septoria</strong> nodorum (Berk.) in<br />

wheat. J. <strong>of</strong> Plant Dis. <strong>and</strong> Prot.<br />

99:371-380.<br />

Wilkinson, C.A., J.P. Murphy, <strong>and</strong><br />

R.C. Rufty. 1990. Diallel analysis <strong>of</strong><br />

components <strong>of</strong> partial resistance to<br />

<strong>Septoria</strong> nodorum in wheat. Plant<br />

Dis. 74:47-50.<br />

Wilson, R.E. 1979. Resistance to<br />

<strong>Septoria</strong> tritici in two wheat<br />

cultivars, determined by<br />

independent, single dominant<br />

genes. Aust. Plant Pathol. 8:16-18.<br />

Wilson, R.E. 1994. Progress toward<br />

breeding for resistance to the two<br />

septoria diseases <strong>of</strong> wheat in<br />

Australia. In: Proceedings <strong>of</strong> the<br />

4 th International Workshop on:<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>. E. Arseniuk, T.<br />

Goral, <strong>and</strong> P. Czembor (eds.).<br />

IHAR, Radzikow, Pol<strong>and</strong>.<br />

Hodowla Roslin Aklimatyzacja I<br />

Nasiennictwo (Special edition)<br />

38(3-4):149-152.<br />

Zuckerman, E., A. Eshel, <strong>and</strong> Z Eyal.<br />

1997. Physiological aspects related<br />

to tolerance <strong>of</strong> spring wheat<br />

cultivars to septoria tritici blotch.<br />

Phytopathology 87:60-65.


Breeding for Resistance to <strong>Septoria</strong> <strong>and</strong><br />

<strong>Stagonospora</strong> Blotches in Winter Wheat in<br />

the United States<br />

G. Shaner<br />

Department <strong>of</strong> Botany <strong>and</strong> Plant Pathology, Purdue University,<br />

West Lafayette, IN, USA<br />

This report will concentrate on<br />

resistance breeding efforts in the<br />

winter wheat region <strong>of</strong> the United<br />

States, where leaf blotch, caused by<br />

either <strong>Septoria</strong> tritici or <strong>Stagonospora</strong><br />

nodorum, has been a chronic<br />

disease. Leaf blotch has been a<br />

recognized problem in the eastern<br />

s<strong>of</strong>t winter wheat region <strong>of</strong> the US<br />

for more than 50 years. Historically,<br />

<strong>Stagonospora</strong> nodorum was the major<br />

pathogen in the southeastern US<br />

<strong>and</strong> <strong>Septoria</strong> tritici was the major<br />

pathogen in the north-central<br />

regions. Since the mid 1980s,<br />

however, S. nodorum has been at<br />

least as damaging in the northcentral<br />

region as S. tritici.<br />

The Purdue University-USDA<br />

small grain improvement program<br />

provides an example <strong>of</strong> efforts to<br />

manage leaf blotch by use <strong>of</strong><br />

genetic resistance. During the mid<br />

1950s, wheat breeders began to<br />

incorporate resistance to S. tritici<br />

into adapted s<strong>of</strong>t red winter wheat<br />

cultivars. At that time, S. tritici was<br />

clearly the dominant leaf blotch<br />

pathogen. The primary sources <strong>of</strong><br />

resistance were wheat cultivars<br />

from South America, e.g. Bulgaria<br />

88, Sao Sepe, <strong>and</strong> Sudeste.<br />

Eventually, most effort<br />

concentrated on Bulgaria 88 as the<br />

source <strong>of</strong> resistance. Cultivars<br />

Oasis <strong>and</strong> Sullivan, released in 1973<br />

<strong>and</strong> 1977, respectively, carried this<br />

resistance (Patterson et al., 1975;<br />

1979). When Oasis or Sullivan are<br />

inoculated at the adult plant stage<br />

in the greenhouse with spores <strong>of</strong> S.<br />

tritici, they are highly resistant.<br />

Infection results in small chlorotic<br />

flecks, with little or no chlorosis<br />

<strong>and</strong> no formation <strong>of</strong> pycnidia.<br />

Resistance in these cultivars<br />

segregates as a single, dominant<br />

Mendelian factor when they are<br />

crossed to a susceptible cultivar.<br />

Although Oasis derived its<br />

resistance to S. tritici mainly from<br />

Bulgaria 88, it evidently carried<br />

additional factors for resistance to<br />

S. nodorum because it exhibited a<br />

degree <strong>of</strong> resistance <strong>and</strong> performed<br />

well in areas <strong>of</strong> the southeastern US<br />

where this species was the<br />

dominant leaf spotting pathogen.<br />

This additional resistance may<br />

have resulted from the fact that<br />

selection for resistance in Indiana<br />

was entirely in the field under<br />

conditions <strong>of</strong> natural infection, <strong>and</strong><br />

there may have been more S.<br />

nodorum present than was<br />

suspected.<br />

Since the latter half <strong>of</strong> the 1980s,<br />

S. nodorum has emerged as the<br />

dominant leaf spotting pathogen in<br />

Indiana (Shaner <strong>and</strong> Buechley,<br />

1995). This shift in pathogen<br />

populations has evidently occurred<br />

throughout the north central region<br />

127<br />

<strong>of</strong> the US, <strong>and</strong> leaf blotch must now<br />

be regarded as a complex involving<br />

both S. tritici <strong>and</strong> S. nodorum. In<br />

more southern regions, S. nodorum<br />

continues to be the major leaf<br />

spotting pathogen.<br />

Management <strong>of</strong> this disease<br />

complex by genetic resistance<br />

requires resistance to both<br />

pathogens. Despite at least 50 years<br />

<strong>of</strong> effort in breeding for resistance<br />

to leaf blotch, progress has been<br />

modest. Evidence for this<br />

conclusion comes from descriptions<br />

<strong>of</strong> cultivars developed in this<br />

region <strong>and</strong> from direct observation<br />

<strong>of</strong> many cultivars in field trials.<br />

An estimation <strong>of</strong> the degree <strong>of</strong><br />

resistance available in currently<br />

available cultivars <strong>of</strong> s<strong>of</strong>t red<br />

winter wheat can be obtained from<br />

data collected in statewide wheat<br />

performance trials conducted each<br />

year in Indiana. For many years,<br />

entries in this trial have been<br />

evaluated for disease. Leaf blotch is<br />

encountered every year to some<br />

degree in Indiana wheat fields, but<br />

at varying severity. For the period<br />

1973–91, leaf blotch symptoms<br />

reached the flag leaf <strong>of</strong> susceptible<br />

cultivars within 26 days <strong>of</strong> heading<br />

in 10 <strong>of</strong> the 19 years (Shaner <strong>and</strong><br />

Buechley, 1995).


Session 6C — G. Shaner<br />

128<br />

An examination <strong>of</strong> leaf blotch<br />

severity data from 1986 through<br />

1996, collected from two locations<br />

in Indiana (Tippecanoe <strong>and</strong><br />

Daviess Counties), can be used to<br />

estimate progress in development<br />

<strong>of</strong> cultivars with resistance to this<br />

disease complex. The number <strong>and</strong><br />

identity <strong>of</strong> entries in these trials<br />

varied from year to year. Typically,<br />

a cultivar would be included for<br />

several consecutive years, but not<br />

for the full period covered by these<br />

trials. Each year, the trial included<br />

most cultivars <strong>of</strong> s<strong>of</strong>t red winter<br />

wheat <strong>of</strong>fered for sale in Indiana<br />

<strong>and</strong> adjacent states. The cultivars<br />

Arthur <strong>and</strong> Caldwell were<br />

included in every trial as long-term<br />

checks. Severity <strong>of</strong> leaf blotch in<br />

these trials was estimated using a 0<br />

to 9.5 scale, modified from one<br />

developed at <strong>CIMMYT</strong> (Shaner <strong>and</strong><br />

Buechley, 1995). This scale is<br />

reproduced here as Table 1 for<br />

convenient reference. The ratings <strong>of</strong><br />

leaf blotch severity discussed here<br />

were made when wheat was in the<br />

early to mid dough stage <strong>of</strong><br />

development.<br />

Table 1. Leaf blotch severity scale for wheat.<br />

There are several ways<br />

that the question <strong>of</strong><br />

improvement in resistance to<br />

leaf blotch can be<br />

investigated using data from<br />

these trials. If the general<br />

level <strong>of</strong> resistance among<br />

cultivars increased over the<br />

14-year period, then the trial<br />

mean leaf blotch severity,<br />

based on all cultivars,<br />

should decline relative to the<br />

severities for the check<br />

cultivars. Among the 16<br />

trials, trial means ranged<br />

from 5.4 to 9.2, reflecting<br />

variation among years in the<br />

suitability <strong>of</strong> weather for leaf<br />

blotch development. This<br />

variation in trial means was<br />

closely followed by the<br />

values for the two long-term<br />

check cultivars, Arthur <strong>and</strong><br />

Caldwell. The trial mean<br />

was always less than the<br />

severity for Caldwell <strong>and</strong><br />

usually less than the severity<br />

for Arthur, but there was no<br />

evidence for a greater<br />

disparity between the trial<br />

mean <strong>and</strong> the checks over<br />

Severity<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

Mean<br />

Arthur<br />

Caldwell<br />

86 PAF<br />

87 PAF<br />

89 PAF<br />

90 PAF<br />

90 DAV<br />

91 PAF<br />

9 DAV<br />

93 PAF<br />

93 DAV<br />

94 PAF<br />

94 DAV<br />

96 PAF<br />

96 DAV<br />

97 DAV<br />

98 DAV<br />

99 PAF<br />

Trial<br />

Figure 1. Trial means <strong>and</strong> means for check cultivars<br />

Arthur <strong>and</strong> Caldwell for leaf blotch ratings, 1986<br />

through 1999, at two locations in Indiana. PAF =<br />

Purdue Agronomy Farm (Tippecanoe County); DAV =<br />

Daviess County.<br />

time (Figure 1). The conclusion<br />

from this is that the newer cultivars<br />

are not more resistant collectively<br />

than those included in trials during<br />

earlier years, as represented by the<br />

two check cultivars.<br />

If only a few cultivars in a trial<br />

had substantially improved<br />

resistance, then the mean <strong>of</strong> all<br />

entries in that trial might not reflect<br />

this improvement, but the st<strong>and</strong>ard<br />

deviation <strong>of</strong> cultivar means should<br />

be greater than in trials lacking<br />

exceptionally resistant cultivars. If<br />

there has been progress in<br />

Range in percent severity on indicated leaf<br />

Scale value Flag Flag-1 Flag-2 Flag-3 Mean severityx 1 0-5 0.1<br />

2 5-20 2.9<br />

3 20-40 8.1<br />

4 1-10 40-70 15.6<br />

5 0-1 10-25 70-90 25.5<br />

6 1-10 25-75 90-100 37.8<br />

7 10-50 75-100 100 52.3<br />

8 1-20 50-90 100 100 69.3<br />

9 20-90 90-100 100 100 88.5<br />

9.5 90-100 100 100 100 99.1<br />

x Mean severity is the average for the four leaves, based on midpoint values for each range. Mean severity (P) can be<br />

calculated from the scale value (S) according to: P = -0.38253 – 0.69435 S + 1.17499 S2 (R2 = 0.999).


development <strong>of</strong> cultivars resistant to<br />

leaf blotch, then the variance <strong>of</strong><br />

cultivar means should be seen to<br />

increase over the years. Plotting trial<br />

means <strong>and</strong> st<strong>and</strong>ard deviations over<br />

time revealed no tendency for the<br />

st<strong>and</strong>ard deviation to increase with<br />

time or to vary with the trial mean<br />

(Figure 2).<br />

Examination <strong>of</strong> data from<br />

individual trials may reveal to what<br />

extent resistance is expressed under<br />

conditions <strong>of</strong> severe, moderate, or<br />

mild disease pressure. Results from<br />

three trials are presented here for<br />

illustration. The trial with the<br />

greatest severity <strong>of</strong> leaf blotch was<br />

in Daviess County during 1990. The<br />

trial mean severity was 9.2. The<br />

cultivars with the least disease in<br />

this trial had a rating <strong>of</strong> 8. Thus,<br />

Severity<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Breeding for Resistance to <strong>Septoria</strong> <strong>and</strong> <strong>Stagonospora</strong> Blotches in Winter Wheat in the United States 129<br />

Mean<br />

St dev<br />

86 PAF<br />

87 PAF<br />

89 PAF<br />

90 PAF<br />

90 DAV<br />

91 PAF<br />

9 DAV<br />

93 PAF<br />

93 DAV<br />

94 PAF<br />

94 DAV<br />

96 PAF<br />

96 DAV<br />

97 DAV<br />

98 DAV<br />

99 PAF<br />

Trial<br />

Figure 2. Mean <strong>and</strong> st<strong>and</strong>ard deviation <strong>of</strong> leaf blotch<br />

severity on wheat in cultivar trials from 1986 through<br />

1999 at two locations in Indiana.<br />

8<br />

;; yy<br />

6<br />

;; yy ;; yy<br />

4<br />

;; yy ;; yy ;; yy ;; yy ;; yy ;; yy ;; yy<br />

2<br />

;; yy<br />

0<br />

;; yy ;; yy ;; yy ;; yy ;; yy ;; yy<br />

4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5<br />

Severity class<br />

Figure 4. Frequency distribution <strong>of</strong> mean leaf blotch<br />

severities on wheat cultivars evaluated during 1993<br />

in Tippecanoe County, Indiana.<br />

Number <strong>of</strong> entries<br />

under conditions <strong>of</strong> great disease<br />

pressure, symptoms progressed to<br />

the flag leaves <strong>of</strong> all cultivars<br />

(Figure 3). The entire range in<br />

resistance was confined to the<br />

degree <strong>of</strong> flag leaf area that was<br />

blighted. All lower leaves were<br />

completed destroyed by leaf blotch.<br />

Mean leaf blotch severity was<br />

very low in the trial conducted in<br />

Tippecanoe County during 1993<br />

(Figure 4). The range among<br />

cultivar severity means was<br />

somewhat greater than in the<br />

previous example, because severity<br />

values were not truncated by the<br />

maximum possible value. The most<br />

severely affected cultivars showed<br />

only a few lesions on the flag leaf.<br />

On most <strong>of</strong> the cultivars, leaf blotch<br />

was at most only 10% on leaf F-1.<br />

Number <strong>of</strong> entries<br />

Number <strong>of</strong> entries<br />

15<br />

The trial conducted during 1997<br />

in Daviess County represents<br />

moderately severe disease pressure.<br />

The trial mean was 7.8 (Figure 5).<br />

Blotch symptoms reached the flag<br />

leaf on fewer than half <strong>of</strong> the<br />

cultivars in this trial, but the<br />

cultivars with the least amount <strong>of</strong><br />

disease had symptoms on leaf F-1.<br />

Regardless <strong>of</strong> overall disease<br />

pressure, the range in leaf blotch<br />

severity among cultivars within a<br />

trial was not great.<br />

Analysis <strong>of</strong> these field trial data<br />

provides no evidence for substantial<br />

increases in resistance in a region <strong>of</strong><br />

the United States where leaf blotch<br />

is a chronic <strong>and</strong> <strong>of</strong>ten severe disease<br />

<strong>of</strong> wheat. At best, a resistant cultivar<br />

will hold symptom development<br />

back by a few days on the flag <strong>and</strong><br />

flag-1 leaves.<br />

10<br />

5<br />

0<br />

4 4.5 5 5.5<br />

;; yy ; y ;<br />

;; yy;<br />

y; 6 6.5 7 7.5 8 8.5 9 9.5<br />

Severity class<br />

; ;; yy<br />

;; yy<br />

;; yy<br />

Figure 3. Frequency distribution <strong>of</strong> mean leaf blotch<br />

severities on wheat cultivars evaluated during 1990<br />

in Daviess County, Indiana.<br />

20 ;; yy<br />

15<br />

10 ;; yy ;; yy<br />

5 ;; yy ;; yy ;; yy<br />

0<br />

;; yy ;; yy ;; yy ;; yy<br />

4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5<br />

Severity class;;<br />

Figure 5. Frequency distribution <strong>of</strong> mean leaf blotch<br />

severities on wheat cultivars evaluated during 1997<br />

in Daviess County, Indiana.<br />

y<br />

y<br />

y


Session 6C — G. Shaner<br />

130<br />

Another approach to<br />

determining progress in breeding<br />

winter wheat for resistance to leaf<br />

blotch is to examine cultivar<br />

registrations in the journal Crop<br />

Science over a period <strong>of</strong> years.<br />

Registrations for 131 wheat<br />

cultivars were published in<br />

volumes 28-39, <strong>of</strong> which 96 were<br />

for winter wheat. Of these winter<br />

wheat cultivars, 25 were reported<br />

to have some degree <strong>of</strong> resistance<br />

to S. tritici or S. nodorum <strong>and</strong> 2 were<br />

reported to be susceptible. For the<br />

other 69 cultivars there was no<br />

mention <strong>of</strong> reaction to either <strong>of</strong><br />

these pathogens. The summary for<br />

winter wheat cultivars includes<br />

both hard <strong>and</strong> s<strong>of</strong>t, <strong>and</strong> red <strong>and</strong><br />

white classes, covering the eastern<br />

portion <strong>of</strong> the US, the Great Plains,<br />

<strong>and</strong> the Pacific Coast states.<br />

Because leaf blotch has historically<br />

been a greater problem in the<br />

eastern region <strong>of</strong> the US, data were<br />

summarized for this region. Of 39<br />

wheat cultivars developed in the<br />

eastern s<strong>of</strong>t wheat region <strong>of</strong> the US,<br />

15 were described as having some<br />

resistance, none were described as<br />

susceptible, <strong>and</strong> there was no<br />

mention <strong>of</strong> reaction for 24 cultivars.<br />

The absence <strong>of</strong> any comment about<br />

reaction to leaf blotch is interesting,<br />

because for many other wheat<br />

pathogens <strong>and</strong> pests (e.g. rusts,<br />

powdery mildew, viruses, Hessian<br />

fly), the registration articles<br />

commonly documented<br />

susceptibility as well as resistance.<br />

Of the 15 cultivars described as<br />

being resistant, 12 were described<br />

as having resistance to S. nodorum,<br />

7 <strong>of</strong> which were described as<br />

having resistance to S. tritici as<br />

well. The other three cultivars were<br />

described as being resistant only to<br />

S. tritici.<br />

An association has long been<br />

noted between tall stature, late<br />

maturity, <strong>and</strong> resistance to leaf<br />

blotch caused by either S. tritici or<br />

S. nodorum (Camacho-Casas et al.,<br />

1995; Scott et al., 1982). Late<br />

maturity <strong>and</strong> tall stature per se may<br />

confer a significant degree <strong>of</strong> leaf<br />

blotch in such cultivars. Where leaf<br />

blotch is the greatest threat in<br />

North America, the eastern s<strong>of</strong>t<br />

wheat region, breeders have<br />

emphasized development <strong>of</strong> short<br />

(90 to 100 cm), early maturing<br />

cultivars, as a way to reduce risk<br />

from rust infection, to permit grain<br />

filling before the excessive hot <strong>and</strong><br />

humid conditions <strong>of</strong> summer, <strong>and</strong><br />

to permit double-cropping with<br />

soybeans. The desired st<strong>and</strong>ards <strong>of</strong><br />

early maturity <strong>and</strong> short stature<br />

may be an important reason for the<br />

lack <strong>of</strong> progress in achieving<br />

adequate levels <strong>of</strong> resistance to the<br />

leaf blotch pathogens.<br />

Hectares <strong>of</strong> s<strong>of</strong>t wheat<br />

production in the eastern US have<br />

been declining for many years.<br />

Poor yields <strong>and</strong> poor grain quality<br />

because <strong>of</strong> disease have been a<br />

major reason for the ab<strong>and</strong>onment<br />

<strong>of</strong> wheat production by many<br />

farmers. Unless greater degrees <strong>of</strong><br />

resistance can be bred into highyielding,<br />

early-maturing, shortstatured<br />

wheat cultivars, this<br />

downward trend in production will<br />

likely continue.<br />

References<br />

Camacho-Casas, M.A., W.E.<br />

Kronstad, <strong>and</strong> A.L. Scharen. 1995.<br />

<strong>Septoria</strong> tritici resistance <strong>and</strong><br />

associations with agronomic traits<br />

in a wheat cross. Crop Sci. 35:971-<br />

976.<br />

Patterson, F.L., Roberts, J.J., Finney,<br />

R.E., Shaner, G.E., Gallun, R.L.,<br />

<strong>and</strong> Ohm, H.W. 1975. Registration<br />

<strong>of</strong> Oasis wheat. Crop Sci. 15:736-<br />

737.<br />

Patterson, F.L., Shaner, G.E., Huber,<br />

D.M., Ohm, H.W., Finney, R.E.,<br />

Gallun, R.L., <strong>and</strong> Roberts, J.J. 1979.<br />

Registration <strong>of</strong> Sullivan wheat<br />

Crop Sci. 19:297.<br />

Scott, P.R., Benedikz, P.M., <strong>and</strong> Cox,<br />

C.J. 1982. A genetic study <strong>of</strong> the<br />

relationship between height, time<br />

<strong>of</strong> ear emergence <strong>and</strong> resistance to<br />

<strong>Septoria</strong> nodorum in wheat. Plant<br />

Pathology 31:45-60.<br />

Shaner, G., <strong>and</strong> Buechley, G. 1995.<br />

Epidemiology <strong>of</strong> leaf blotch <strong>of</strong> s<strong>of</strong>t<br />

red winter wheat caused by<br />

<strong>Septoria</strong> tritici <strong>and</strong> <strong>Stagonospora</strong><br />

nodorum. Plant Disease 79:928-938.


<strong>Septoria</strong> tritici Resistance <strong>of</strong> Wheat<br />

Cultivars at Different Growth Stages<br />

M. Díaz de Ackermann, 1 M.M. Kohli, 2 <strong>and</strong> V. Ibañez1 1 INIA La Estanzuela, Colonia, Uruguay<br />

2 <strong>CIMMYT</strong> Regional Wheat Program, Uruguay<br />

Abstract<br />

Nine wheat cultivars planted in the greenhouse were inoculated with a mixture <strong>of</strong> three isolates <strong>of</strong> <strong>Septoria</strong> tritici at<br />

four different growth stages. The first evaluation, done six weeks after inoculation, showed that the facultative wheats at<br />

growth stage (GS) 39 (flag leaf ligule just visible) <strong>and</strong> spring wheats at GS 65 (flowering halfway complete) reached the<br />

highest levels <strong>of</strong> infection, with two exceptions, cvs E. Pelón 90 <strong>and</strong> P. Oasis. In the second evaluation a week later, the<br />

highest level <strong>of</strong> infection was reached when facultative wheats were inoculated at GS 32-33 (stem elongation-first node<br />

detectable) <strong>and</strong> spring wheats at GS 39 (flag leaf ligule just visible). In both evaluations, the effects <strong>of</strong> cultivars, seeding<br />

date/GS at inoculation, <strong>and</strong> the interaction between cultivars <strong>and</strong> GS at inoculation were significant. <strong>Septoria</strong> tritici<br />

susceptibility increased with the age <strong>of</strong> the plants. The increase in the level <strong>of</strong> infection was significantly lower in cv P.<br />

Superior (Bobwhite), indicating a certain level <strong>of</strong> adult plant resistance.<br />

Wheat is an important crop in the<br />

western region <strong>of</strong> Uruguay, where it<br />

is grown for grain <strong>and</strong> forage<br />

(double purpose) in the mixed<br />

farming system, or solely for grain<br />

production. Among diseases present<br />

in the country, <strong>Septoria</strong> tritici leaf<br />

blotch is one <strong>of</strong> the most important,<br />

<strong>and</strong> it can be particularly severe<br />

during the wheat-growing season in<br />

this temperate high rainfall<br />

environment. The wide range <strong>of</strong><br />

planting dates from May (or even<br />

earlier for double purpose crops) to<br />

August determines the growth stage<br />

(GS) at the time <strong>of</strong> primary infection,<br />

later in the fall or in early spring.<br />

Facultative wheats, sown in the fall,<br />

are infected very early (seedling<br />

stage) or at the beginning <strong>of</strong> spring<br />

(boot stage), depending on the year.<br />

Spring wheats, sown later than the<br />

facultative types, are exposed mainly<br />

to spring infections.<br />

Reduced disease severity<br />

associated with tall plant stature <strong>and</strong><br />

late maturity has been observed<br />

(Danon et al., 1982; Eyal <strong>and</strong> Talpaz,<br />

1990; van Beuningen <strong>and</strong> Kohli,<br />

1990; Tavella, 1978). However, the<br />

relationship between the resistance<br />

expressed at the seedling stage <strong>and</strong><br />

the adult plant stage, or vice versa,<br />

is not fully understood. Kema <strong>and</strong><br />

van Silfhout (1995) found<br />

significant correlations for one <strong>of</strong><br />

three isolates inoculated on 22<br />

wheat cultivars at the seedling <strong>and</strong><br />

adult plant stages. On the other<br />

h<strong>and</strong>, Arama, Parlevliet, <strong>and</strong> van<br />

Silfhout (1994) described three<br />

types <strong>of</strong> resistance: resistance only<br />

in seedling (seedling resistance),<br />

resistance only in adult plant (adult<br />

plant resistance), <strong>and</strong> resistance in<br />

both seedling <strong>and</strong> adult plant<br />

(overall resistance).<br />

The objective <strong>of</strong> this study was<br />

to determine the level <strong>of</strong> resistance<br />

in nine wheat cultivars,<br />

representing different reactions to S.<br />

tritici in the field, at different<br />

growth stages <strong>and</strong> under controlled<br />

conditions.<br />

Materials <strong>and</strong> Methods<br />

131<br />

Nine wheat cultivars with<br />

different growth habits <strong>and</strong> field<br />

reactions to <strong>Septoria</strong> tritici were<br />

planted in three pots each, five<br />

plants per pot, on four seeding<br />

dates: 28 April . , 16 May . , 31 May . ,<br />

<strong>and</strong> 27 June 1995. The pots<br />

measured 29 <strong>and</strong> 20 cm in their<br />

upper <strong>and</strong> lower diameters,<br />

respectively, <strong>and</strong> 20 cm in height.<br />

The cultivars’ growth habit,<br />

progenitors, <strong>and</strong> field reaction to S.<br />

tritici are presented in Table 1. The<br />

concentration <strong>of</strong> S. tritici spore<br />

suspension was adjusted to 106 conidia per ml, with the aid <strong>of</strong> a<br />

hemacytometer. The cultivars were<br />

inoculated with a mixture <strong>of</strong> three<br />

isolates (26S, 4407, <strong>and</strong> E. Federal)<br />

on 17 July, at the growth stages<br />

(Zadoks et al., 1974) indicated in<br />

Table 2. The selected isolates were<br />

characterized by Diaz de<br />

Ackermann et al. (1994). The plants<br />

were kept in a humid chamber until<br />

20 July. Due to the delayed


Session 6C — M. Díaz de Ackermann, M.M. Kohli, <strong>and</strong> V. Ibañez<br />

132<br />

appearance <strong>of</strong> symptoms, two<br />

evaluations were done, the first<br />

on 28 August . <strong>and</strong> the second on<br />

5 September.<br />

The infection was scored as<br />

percent <strong>of</strong> leaf area affected by<br />

the disease, on two tillers per<br />

plant <strong>and</strong> five leaves per tiller.<br />

Transformed data (Loge +0.5) for<br />

the flag leaf, F-1, <strong>and</strong> F-2, as<br />

well as the average <strong>of</strong> the three<br />

leaves, were analyzed using<br />

SAS GLM procedure (Version<br />

6.12; SAS Institute, Cary, NC).<br />

Infection on the lower two<br />

leaves was not included in this<br />

analysis due to many missing<br />

values.<br />

Results <strong>and</strong><br />

Discussion<br />

The analysis <strong>of</strong> variance<br />

(ANOVA) for each leaf <strong>and</strong> the<br />

average <strong>of</strong> the three leaves<br />

showed significant differences<br />

between cultivars <strong>and</strong> seeding<br />

dates, as well as significant<br />

interaction between seeding dates<br />

<strong>and</strong> cultivars in both evaluations. As<br />

the results were similar for<br />

individual leaves <strong>and</strong> the average <strong>of</strong><br />

the three top leaves, only data for<br />

the latter are presented (Table 3).<br />

Differences among cultivars<br />

were found on the first two seeding<br />

dates (28/04 <strong>and</strong> 16/05) during the<br />

first evaluation, <strong>and</strong> on the<br />

intermediate dates (16/05 <strong>and</strong> 31/<br />

05) during the second evaluation.<br />

All the cultivars except B.<br />

Charrúa showed different infection<br />

levels on different seeding dates at<br />

first evaluation. This interaction is<br />

explained by the cultivars’ different<br />

Table 1. Cross, growth habit <strong>and</strong> field reaction to <strong>Septoria</strong> tritici <strong>of</strong> nine wheat cultivars.<br />

Cultivar Cross Growth habit Reaction<br />

E. Cardenal Veery#3 Spring S1 INIA Mirlo Car 853/Coc//Vee#5/3/Ures Spring R<br />

E Pelón 90 Kavkaz/Torim73 Spring R-MR<br />

ProINTA Superior Bobwhite Spring MR<br />

ProINTA Oasis Oasis/Torim Spring S<br />

E. Federal E.Hornero/CNT8 Facultative MR<br />

E Halcón Buck6/MR74507 Facultative MR<br />

Buck Charrúa RAP/RE//IRAP/3/LOV/4/RAP/RE//IRAP Facultative R<br />

LE 2196 E. Jilguero/ND526 Facultative R<br />

1 S: susceptible, R: resistant, MR: moderately resistant.<br />

Table 2. Growth stages <strong>of</strong> spring <strong>and</strong> facultative wheat varieties at time <strong>of</strong> inoculation.<br />

Seeding date 28/04 16/05 31/05 27/06<br />

Spring wheats 65 39 30-31 22-23<br />

Facultative wheats 39 32-33 30 24-25<br />

Table 3. Average level <strong>of</strong> infection <strong>of</strong> top three leaves, comparison among date <strong>of</strong> seeding (small<br />

letter) <strong>and</strong> cultivars (capital letter). Evaluation August 28th <strong>and</strong> September 5th .<br />

Sowing date 28/04 16/05 31/05 27/06<br />

GS Facultative 39 32/33 30 24/25<br />

GS Spring 65 39 30/31 22/23<br />

Cultivar 1st . eval 1st . eval 2nd . eval 1st . eval 2nd . eval 1st . eval 2nd . eval<br />

E.Cardenal 11.65 aA 3.81 bA 46.24 aA 0.16 c 3.57 bA 0.00 c 0.45 c<br />

I. Mirlo 1.98 aCD 2.17 aAB 23.93 aB 0.00 b 1.96 bB 0.00 b 0.66 c<br />

E. Pelón 90 1.32 aDE 2.20 aAB 16.86 aC 0.02 b 1.76 bB 0.00 b 0.19 c<br />

P. Superior 2.92 aC 1.50 bB 8.14 aD 0.00 c 1.86 bB 0.00 c 0.11 c<br />

P. Oasis 1.79 bCD 4.25 aA 27.22 aB 0.26 c 2.18 bB 0.02 c 0.44 c<br />

E. Federal 7.64 aB 3.17 bA 19.15 aC 0.00 c 0.50 bD 0.00 c 0.43 b<br />

E. Halcón 1.25 aE 0.23 bC 3.61 aE 0.03 b 1.13 bC 0.00 b 0.17 c<br />

B. Charrúa 0.14 aE 0.03 aC 2.91 aE 0.00 a 0.56 bCD 0.00 a 0.44 b<br />

LE 2196 1.66 aDE 0.22 bC 6.65 aD 0.03 b 1.23 bB 0.00 b 0.33 c<br />

Average 3.37 1.95 17.19 0.06 1.64 0.00 0.36<br />

behavior on the different seeding<br />

dates/GS at inoculation. Early<br />

seedling infection demonstrated<br />

higher sensitivity for detecting<br />

differences among cultivars.<br />

During this evaluation, the<br />

facultative wheat varieties were<br />

found to have higher <strong>Septoria</strong><br />

susceptibility at GS 39 (flag leaf<br />

ligule just visible) <strong>and</strong> spring<br />

varieties at GS 65 (flowering<br />

halfway complete). These were<br />

followed by GS 32/33 (stem<br />

elongation-first node detectable)<br />

<strong>and</strong> GS 39 for facultative <strong>and</strong><br />

spring varieties, respectively.<br />

Inoculations at GS 22 to GS 31<br />

did not produce adequate<br />

infection levels.<br />

At the second evaluation on 5<br />

September, the group seeded on the<br />

first date was already dry. For the<br />

remaining three sowing dates, the<br />

highest infection level was shown by<br />

facultative wheat varieties inoculated<br />

at GS 32-33 <strong>and</strong> by spring wheat<br />

varieties inoculated at GS 39.<br />

Comparison <strong>of</strong> the cultivars’ field<br />

reactions to S. tritici with infection<br />

under controlled conditions showed<br />

differences in I. Mirlo, P. Superior, E.<br />

Halcón, <strong>and</strong> P. Oasis. These cultivars<br />

are considered resistant (R),<br />

moderately resistant (MR), moderately<br />

resistant (MR), <strong>and</strong> susceptible (S),<br />

respectively, under field conditions;<br />

under controlled conditions they were


MS, R, R, <strong>and</strong> MS, respectively. The<br />

interaction between cultivar <strong>and</strong><br />

growth stage at inoculation<br />

indicated the lack <strong>of</strong> precision that<br />

may occur while evaluating<br />

resistance <strong>of</strong> a variety.<br />

In the first evaluation, with<br />

relatively low levels <strong>of</strong> infection,<br />

cultivars such as P. Oasis showed<br />

significantly lower infection levels<br />

at GS 65 than at GS 39. E. Pelón 90<br />

showed similar behavior without<br />

being significantly different at<br />

these two growth stages. However,<br />

in contrast to P. Oasis <strong>and</strong> E. Pelón<br />

90, P. Superior presented a<br />

significantly higher infection level<br />

at GS 65 than at GS 39 (Table 3;<br />

Figure 1). Although the infection<br />

level <strong>of</strong> P. Superior at the initial<br />

stages was almost equal to that <strong>of</strong><br />

% <strong>of</strong> infection<br />

0<br />

Sowing date 27/06 31/05 16/05 28/04<br />

ZGS 22-23 30-31 39 65<br />

Figure 1. Level <strong>of</strong> infection, average <strong>of</strong> three<br />

% <strong>of</strong> infection<br />

12<br />

9<br />

6<br />

3<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Leaves, August 28th.<br />

Cardenal<br />

Oasis<br />

Mirlo<br />

Pelón<br />

Superior<br />

Leaves, September 5th.<br />

Charrúa<br />

Halcón<br />

LE2196<br />

Federal<br />

Superior<br />

Pelón<br />

Mirlo<br />

Oasis<br />

Cardenal<br />

0<br />

Sowing date 27/06 31/05 16/05<br />

S-ZGS 22-23 30-31 39<br />

F-ZGS 24-25 30 32-33<br />

Figure 2. Level <strong>of</strong> infection, average <strong>of</strong> three<br />

susceptible varieties P. Oasis <strong>and</strong> E.<br />

Cardenal at the second evaluation, P.<br />

Superior showed the slowest<br />

increase over time to become the<br />

most resistant genotype at the adult<br />

plant stage. This suggests that P.<br />

Superior (Bobwhite) may have a<br />

certain degree <strong>of</strong> adult plant<br />

resistance.<br />

In general, spring habit, high<br />

yielding, early maturing semidwarf<br />

wheats showed higher levels <strong>of</strong><br />

<strong>Septoria</strong> infection than the facultative<br />

wheats (tall <strong>and</strong> late maturing). The<br />

highest infection level observed in a<br />

facultative wheat (E. Federal) is<br />

equivalent to the level demonstrated<br />

by a moderately resistant spring<br />

variety such as E. Pelón 90 (Figure 2).<br />

As per the definition put forward by<br />

Arama, Parleviet, <strong>and</strong> van Silfhout<br />

(1994), Buck Charrúa seems to<br />

possess overall resistance. A high<br />

degree <strong>of</strong> resistance was observed at<br />

the juvenile stage in all varieties.<br />

Only P. Superior demonstrated<br />

adult-plant resistance. Further<br />

studies are required to evaluate the<br />

level <strong>of</strong> resistance in stages earlier<br />

than GS 22-25.<br />

Conclusions<br />

The S. tritici susceptibility <strong>of</strong><br />

different wheat varieties seems to<br />

increase with the age <strong>of</strong> the plants.<br />

However, the rate <strong>of</strong> this increase<br />

(the slope <strong>of</strong> the curve) is different<br />

for each cultivar. As a result<br />

interactions were observed between<br />

the cultivars’ susceptibility (infection<br />

level) <strong>and</strong> growth stage at the time <strong>of</strong><br />

infection (inoculation). This suggests<br />

that several evaluations may be<br />

required to determine a cultivar’s<br />

potential as a parent in a crop<br />

improvement program.<br />

<strong>Septoria</strong> tritici Resistance <strong>of</strong> Wheat Cultivars at Different Growth Stages 133<br />

References<br />

Arama, P.F., Parlevliet, J.E., <strong>and</strong> van<br />

Silfhout, C.H.1994. Effect <strong>of</strong><br />

plant height <strong>and</strong> days to heading<br />

on the expression <strong>of</strong> resistance in<br />

Triticum aestivum to <strong>Septoria</strong> tritici<br />

in Kenya. In: Proceedings <strong>of</strong> the<br />

4 th . International Workshop on:<br />

<strong>Septoria</strong> <strong>of</strong> cereals. E. Arseniuk,<br />

T. Goral, <strong>and</strong> P. Czembor (eds.).<br />

July 4-7, 1994. IHAR Radzikow,<br />

Pol<strong>and</strong>. pp. 153-157.<br />

Danon, T., Sacks, J.M., <strong>and</strong> Eyal, Z.<br />

1982. The relationship among<br />

plant stature, maturity class, <strong>and</strong><br />

susceptibility to septoria leaf<br />

blotch <strong>of</strong> wheat. Phytopathology<br />

72:1037-1042.<br />

Díaz de Ackermann, M., Stewart, S.,<br />

Ibañez,V., Capdeville, F., <strong>and</strong><br />

Stoll, M. 1994. Pathogenic<br />

variability <strong>of</strong> <strong>Septoria</strong> tritici in<br />

isolates from South America. In:<br />

Proceedings <strong>of</strong> the 4 th<br />

International Workshop on:<br />

<strong>Septoria</strong> <strong>of</strong> cereals. July 4-7, 1994.<br />

Ihar Radzikow, Pol<strong>and</strong>.<br />

pp. 335-338.<br />

Eyal, Z., <strong>and</strong> Talpaz, H. 1990. The<br />

combined effect <strong>of</strong> plant stature<br />

<strong>and</strong> maturity on the response <strong>of</strong><br />

wheat <strong>and</strong> triticale accessions to<br />

<strong>Septoria</strong> tritici. Euphytica<br />

46:133-141.<br />

Kema, G.H.J., <strong>and</strong> van Silfhout,<br />

C.H. 1995. Comparative<br />

virulence analysis <strong>of</strong> <strong>Septoria</strong><br />

tritici to seedling <strong>and</strong> adult plant<br />

resistance. In: Breeding for<br />

disease resistance with emphasis<br />

on durability. Regional<br />

Workshop for Eastern, Central<br />

<strong>and</strong> Southern Africa. D.I. Danial<br />

(ed.). Njoro, Kenya, October 4-7,<br />

1994. p. 221.<br />

Tavella, C.M. 1978. Date <strong>of</strong> heading<br />

<strong>and</strong> plant height <strong>of</strong> wheat<br />

cultivars, as related to septoria<br />

leaf blotch damage. Euphytica<br />

27:577-580.<br />

Van Beuningen, L.T., <strong>and</strong> Kohli,<br />

M.M. 1990. Deviation from the<br />

regression <strong>of</strong> infection on<br />

heading <strong>and</strong> height as a measure<br />

<strong>of</strong> resistance to septoria tritici<br />

blotch in wheat. Plant Disease<br />

74:488-493.<br />

Zadoks, J.C., Chang, T.T., <strong>and</strong><br />

Konzak, C.F. 1974. A decimal<br />

code for the growth stages <strong>of</strong><br />

cereals. Weed Res. 14:415-4214.


134<br />

<strong>Septoria</strong> tritici Resistance Sources <strong>and</strong> Breeding<br />

Progress at <strong>CIMMYT</strong>, 1970-99<br />

L. Gilchrist, 1 B. Gomez, 2 R. Gonzalez, 2 S. Fuentes, 3 A. Mujeeb-Kazi, 1 W. Pfeiffer, 1 S. Rajaram, 1 R. Rodriguez, 3 B.<br />

Skovm<strong>and</strong>, 1 M. van Ginkel, 1 <strong>and</strong> C. Velazquez1 (Field presentation)<br />

1 <strong>CIMMYT</strong> Wheat Program, El Batan, Mexico<br />

2 National Livestock <strong>and</strong> Agricultural Research Institute (INIFAP), Mexico<br />

3 Retired <strong>CIMMYT</strong> Staff<br />

Abstract<br />

An overview is provided <strong>of</strong> the sources <strong>of</strong> <strong>Septoria</strong> tritici resistance used in the <strong>CIMMYT</strong> bread wheat program, <strong>and</strong> <strong>of</strong><br />

progress in breeding for resistance for wheat mega-environment 2 (high rainfall) in 1970-99. Several phases can be<br />

distinguished: use <strong>of</strong> Russian winter wheat, lines from the Southern Cone <strong>of</strong> South America (Brazil, Chile <strong>and</strong> Argentina)<br />

<strong>and</strong> USA as resistance sources; introduction <strong>of</strong> disease resistance from Brazilian sources into semi-dwarf <strong>and</strong> early<br />

backgrounds; introduction <strong>of</strong> genes from triticale <strong>and</strong> Triticum diccocon into common wheat; introgression <strong>of</strong> resistance<br />

from synthetic hexaploids <strong>and</strong> derivatives into susceptible bread wheat stocks; pyramiding <strong>of</strong> Chinese resistance sources with<br />

the sources mentioned above, <strong>and</strong> combining them into adapted agronomic types. Some preliminary work on resistant durum<br />

wheat sources, including some <strong>of</strong> the best ones from Tunisia, is presented.<br />

Bread Wheat Program<br />

The semidwarf wheats<br />

developed in Mexico in the 1960s<br />

were widely distributed. The<br />

modern wheat varieties that<br />

replaced local wheats were<br />

successful because <strong>of</strong> their wide<br />

adaptation, high yield potential,<br />

<strong>and</strong> disease resistance (mainly to<br />

stem <strong>and</strong> leaf rust) for favorable,<br />

irrigated production conditions<br />

(Rajaram et al., 1994).<br />

The emphasis <strong>of</strong> <strong>CIMMYT</strong>’s<br />

bread wheat breeding program on<br />

disease resistance was exp<strong>and</strong>ed in<br />

the 1970s to include resistance to<br />

Puccinia striiformis <strong>and</strong> <strong>Septoria</strong><br />

tritici for the high rainfall wheat<br />

growing areas <strong>of</strong> the world (megaenvironment<br />

2). These are<br />

temperate environments with an<br />

average precipitation <strong>of</strong> >500 mm.<br />

The search for S. tritici resistance,<br />

the accumulation <strong>of</strong> resistance<br />

genes, <strong>and</strong> the incorporation <strong>of</strong><br />

resistance into high yielding<br />

advanced lines have been very<br />

successful in the <strong>CIMMYT</strong> bread<br />

wheat program.<br />

Three main groups <strong>of</strong> sources<br />

were used to introduce resistance:<br />

a) Russian winter wheat, b) lines<br />

from the wheat growing areas <strong>of</strong><br />

Brazil, Chile, <strong>and</strong> Argentina<br />

(Southern Cone <strong>of</strong> South America),<br />

<strong>and</strong> c) to a lesser extent, lines from<br />

the USA (Mann et al., 1985)<br />

(Table 1). The resistance <strong>of</strong> these<br />

lines was identified <strong>and</strong>/or<br />

confirmed in different countries<br />

through the International <strong>Septoria</strong><br />

Nursery (ISEPTON), beginning<br />

in 1971.<br />

The breeding program focused<br />

on combining the semidwarf plant<br />

type with high yield potential <strong>and</strong><br />

resistance to leaf <strong>and</strong> stripe rust<br />

plus septoria leaf blotch.<br />

Germplasm was screened <strong>and</strong><br />

selected for septoria leaf blotch at<br />

two locations in Mexico: Patzcuaro<br />

<strong>and</strong> Toluca. Patzcuaro is located at<br />

2,200 m altitude in the western<br />

Mexican highl<strong>and</strong>s (Michoacan<br />

state). It proved suitable for testing<br />

advanced lines, since epidemics <strong>of</strong><br />

S. tritici occur naturally every year.<br />

In some years we can also select for<br />

resistance to Pyrenophora triticirepentis<br />

<strong>and</strong> <strong>Stagonospora</strong> nodorum.<br />

The Atizapan experiment station in<br />

Toluca (Mexico state) at 2,600 m<br />

altitude has an intense rainy season<br />

during the summer months (800-<br />

900 mm). Artificial epidemics are<br />

created by inoculating with<br />

infected straw or a spore<br />

suspension involving a mixture <strong>of</strong><br />

isolates.<br />

Numerous lines with acceptable<br />

resistance levels were derived from<br />

crosses using the above mentioned<br />

sources <strong>and</strong> subsequent testing in<br />

Toluca <strong>and</strong> Patzcuaro (Table 2).


Another set <strong>of</strong> resistant lines<br />

was created by the wheat<br />

germplasm enhancement<br />

section. The S. tritici resistance <strong>of</strong><br />

lines from Brazil (IAS 20, Pelotas<br />

Arthur, Colotana, IAS 58, <strong>and</strong><br />

Maringa), which are tall <strong>and</strong><br />

late-maturing with weak straw,<br />

was incorporated into earlymaturing,<br />

high yielding<br />

semidwarf lines. Thus earlymaturing,<br />

short, high yielding<br />

lines were developed that had a<br />

level <strong>of</strong> resistance similar to that<br />

<strong>of</strong> the original resistant parent<br />

(Table 3).<br />

Bread wheat lines were also<br />

crossed with triticale, <strong>and</strong> the<br />

resulting lines showed high<br />

levels <strong>of</strong> S. tritici resistance<br />

(Table 4). Subsequently, some <strong>of</strong><br />

the lines listed in Table 3 were<br />

combined with triticale-derived<br />

lines (Table 5) in an attempt to<br />

pyramid several resistance<br />

genes.<br />

<strong>CIMMYT</strong>’s wheat wide<br />

crosses program has generated a<br />

wide range <strong>of</strong> resistant<br />

germplasm from D genome<br />

synthetics <strong>and</strong> their synthetic<br />

derivatives (Mujeeb-Kazi et al.,<br />

1996; 1998; 1999). These<br />

materials express high levels <strong>of</strong><br />

resistance to several leaf<br />

pathogens, including Bipolaris<br />

sorokiniana, Pyrenophora triticirepentis,<br />

<strong>and</strong> S. tritici. Their<br />

resistance may be to one or a<br />

combination <strong>of</strong> these pathogens.<br />

Table 6 shows a group <strong>of</strong><br />

representative synthetic wheats<br />

with high levels <strong>of</strong> resistance to<br />

S. tritici.<br />

<strong>Septoria</strong> tritici Resistance Sources <strong>and</strong> Breeding Progress at <strong>CIMMYT</strong>, 1970-99 135<br />

Table 1. Main sources <strong>of</strong> <strong>Septoria</strong> tritici resistance used by the <strong>CIMMYT</strong> wheat breeding<br />

program in the 1970s.<br />

Ex-USSR Southern Cone <strong>of</strong> SA USA spring Other<br />

winter wheat spring wheat wheat sources<br />

Aurora IAS 55 Chris Enkoy<br />

Kavkaz IAS 58 Era Salomoni Seafoam<br />

Bezostaja 1 PF 70254 Frontana-Kenya Kvz/K4500 L.6A.6<br />

Maringa x Newthatch<br />

Carazinho<br />

IAS 63<br />

Lagoa Vermelha<br />

Tizano Pinto Precoz<br />

IAS 62<br />

CTN 7<br />

CTN 8<br />

Gaboto<br />

IAS 20<br />

Table 2. <strong>CIMMYT</strong>-derived lines with acceptable <strong>Septoria</strong> tritici resistance from the ex-USSR,<br />

USA, France, Brazil, Chile, <strong>and</strong> Romania.<br />

Advanced line Pedigree cross Number S. tritici score a Origin<br />

Veery Kvz/Buho//Kal/Bb CM 33027 76-86 Ex-USSR<br />

Bobwhite Au//Kal/Bb/3/Wop CM 33203 32-76 Ex-USSR<br />

Musala Lee/Kvz/3/Cc//Ron/Cha CM16780 Ex-USSR<br />

Sunbird Gll/Cuc//Kvz/Sx CM34630 42-77 Ex-USSR<br />

Milan VS73.600/Mirlo/3/ CM75113 21-74 France/Ex<br />

Bow//Ye/Trf USSR<br />

Attila ND/VG9144//Kal/Bb/ CM85836 73-76 USA/Ex-<br />

3/Yaco/4/Veery USSR<br />

Bagula Ore F1 158/Fdl// CM59123 53-75 USA/Romania<br />

Mfn/2*Tiba63/3/Coc<br />

/4/Bb/GLl//Carp/3/Pvn<br />

Prinia Prl/Vee#6/Myna/Vul CM90722 74-76 Ex-USSR<br />

Chilero 4777*2//Fkn/Gb54/3/ CM66684 74-76 Ex-USSR<br />

Vee#5/4/Buc/Pvn<br />

Corydon Car 853/Coc/Vee/3/ CM81074 41-75 Chile/ Brazil/<br />

E7408/Pam//Hork/PF73226 Ex - USSR<br />

Tinamu IAS58/4/Kal/Bb// CM81812 32-75 Brazil/Ex-<br />

Cjm71/Ald/5/Au// USSR<br />

Kal/Bb/3/Wop<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

Table 3. Short, early-maturing, <strong>Septoria</strong> tritici resistant lines derived from tall, late-maturing,<br />

resistant Brazilian varieties, with a resistance similar to that <strong>of</strong> the original resistant parents.<br />

Lines Cross number S. tritici score a<br />

IAS 20/H567.71//4*IAS 20 CMH 79.243 35-52<br />

IAS 20/H567.71//IAS 20/3/IAS 58 CMH 78.390 31-42<br />

IAS 20/H567.71//IAS20/3/3*MRNG CMH 78.443 32-63<br />

CLT/H471.71A//4*CLT CMH 83.2277 32-53<br />

H.567.71/3*P.Ar CMH 77.308 32-74<br />

IAS 20 – 11-62<br />

IAS 58 – 31-54<br />

Maringa – 31-54<br />

Colotona – 11-51<br />

Pelotas Arthur – 21-71<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.


Session 6C — L. Gilchrist et al.<br />

136<br />

Table 4. <strong>Septoria</strong> tritici resistant lines derived from bread wheat x triticale crosses.<br />

Lines Cross number S. tritici score a<br />

INIA 66/RYE*2//ARM/3/H277.69 CMH 72.A576 21-42**<br />

M2A/CML//2*NYUBAY CMH 80.A1267 22-52<br />

M2A/CML//Nyubay/3/ CMH72A576/MRG CMH 82.961 32-43<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

Table 5. <strong>Septoria</strong> tritici resistant lines combining resistance from Brazilian sources <strong>and</strong> bread<br />

wheat x triticale derived lines.<br />

Lines Cross number S. tritici score a<br />

H.567.71/3*P. Ar/MRNG//IAS20/H567.71/<br />

IAS20/3/3*MRNG/79.243/4/M2A/CML//<br />

NYUBAY/3/INIA66/Rye*2//ARM/3/H277.69<br />

CMH 85.3215 31-42<br />

EAGLE/H567.71//4*EAGLE/3/2*IAS20<br />

/H567.71//4*IAS20<br />

CMH 79.243 11-43<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

Table 6. Synthetic hexaploid wheats (Triticum turgidum x Aegilops tauschii; 2n=6x=42) with high<br />

levels <strong>of</strong> resistance to <strong>Septoria</strong> tritici.<br />

Synthetic hexaploids Cross number S. tritici score a<br />

Aco 89/Ae. tauschii (309) b CIGM90.595 11-11<br />

Croc1/Ae. tauschii (879) CIGM89.479 11-21<br />

Doy1/Ae. tauschii (372) CIGM93.229 11-21<br />

Sty-US/Celta//Pals/3/SRN-5/4/Ae. tauschii (502) CIGM93.261 11-11<br />

Altar 84/Ae. tauschii (502) CIGM93.395 11-11<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

b Accession number in the wide cross working collection in <strong>CIMMYT</strong>’s wheat genebank.<br />

Table 7. Synthetic hexaploids crossed with susceptible <strong>and</strong> moderately susceptible bread<br />

wheats to develop advanced derivatives resistant to <strong>Septoria</strong> tritici.<br />

Advanced derivative Cross number S. tritici score a<br />

Altar 84// Aegilops tauschii (191) b CIGM92.337 21-52<br />

//Yaco/3/2*Bau<br />

Croc 1/ Aegilops tauschii (205)//Kauz CIGM90.248 11-31<br />

Croc 1/ Aegilops tauschii (205)/5/Br12*3/4/ CIGM90.252 11-32<br />

Ias 55*4…..<br />

Altar 84/Aegilops tauschii (219)//Opata CIGM90.429 11-32<br />

Altar 84/Aegilops tauschii (219)//2*Seri CMSS92YO1855M 21-32<br />

Altar 84/Aegilops tauschii (224)//2*Yaco CIGM91.191 11-32<br />

Bcn/3/Fgo/USA2111//Aegilops tauschii (658) CASS94Y00146S 11-22<br />

Opata/6/68111/Rgb-V//Ward/3/Fgo/4/ CASS94Y00247S 11-22<br />

Rabi/5/Aegilops tauschii (878)<br />

Filin/4/Snip/Yan79//Dack/Teal/3/ CASS94Y00072S 32-32<br />

Aegilops tauschii (633)<br />

Altar 84/Aegilops tauschii (191)// Opata/3/ CIGM93.566 11-22<br />

Altar 84/Aegilops tauschii (224)//Yaco(1)…<br />

Mayoor//TkSN1081/Aegilops tauschii (222) CASS94Y00009S 11-22<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

b Accession number in the wide cross working collection in <strong>CIMMYT</strong>’s wheat genebank.<br />

The high levels <strong>of</strong> resistance<br />

present in the synthetic hexaploids<br />

were incorporated into susceptible<br />

bread wheat advanced lines <strong>and</strong> in<br />

some cases were used to reinforce<br />

the intermediate susceptibility<br />

present in some lines (Table 7). To<br />

pyramid the resistance genes,<br />

resistant alien germplasm<br />

possessing tertiary gene pool<br />

diversity was crossed with resistant<br />

Chinese germplasm (Table 8).<br />

Another effort to diversify the<br />

sources <strong>of</strong> resistance focused on the<br />

Triticum dicoccon collection in the<br />

wheat genebank (Gilchrist <strong>and</strong><br />

Skovm<strong>and</strong>, 1995). Early tall<br />

accessions with resistance to P.<br />

striiformis <strong>and</strong> S. tritici, <strong>and</strong> weak<br />

straw were identified from<br />

genebank collections. One <strong>of</strong> these<br />

accessions was selected <strong>and</strong><br />

crossed to a very susceptible<br />

selection <strong>of</strong> the line Kauz. The<br />

resulting lines combine high levels<br />

<strong>of</strong> resistance to P. striiformis <strong>and</strong> S.<br />

tritici with good agronomic type<br />

(Table 9).<br />

In a second effort <strong>of</strong> the bread<br />

wheat breeding program (1986-87)<br />

to increase variability <strong>of</strong> the<br />

resistance base, a group <strong>of</strong><br />

advanced bread wheat lines<br />

consisting <strong>of</strong> resistance sources in<br />

an adapted background (Table 10)<br />

was crossed with a group <strong>of</strong><br />

Chinese lines having acceptable<br />

levels <strong>of</strong> resistance (Sumai #3,<br />

Suzhoe #8, Suzhoe #6 Ningmai #4,<br />

Yangmai #4, Ning 8401, YMI #6,<br />

Shangai #5, Shangai #, Nanjing<br />

7840, Wuhan 2, <strong>and</strong> Wuhan 3).<br />

Once the resistance <strong>of</strong> the<br />

Chinese materials had been<br />

introduced, a new effort was


initiated to combine high yield<br />

potential <strong>and</strong> pyramid resistance<br />

genes (Table 11).<br />

Shuttle breeding between<br />

different sites is very important<br />

for selection <strong>and</strong> helps to increase<br />

effective resistance across<br />

different locations where S. tritici<br />

is a problem. A large number <strong>of</strong><br />

advanced lines showing very high<br />

levels <strong>of</strong> resistance have been<br />

selected through shuttle breeding.<br />

New sources have also been<br />

selected with this methodology. A<br />

representative example <strong>of</strong> new<br />

sources coming from the Southern<br />

Cone <strong>of</strong> South America that were<br />

identified in the region through<br />

the LACOS (Advanced Lines for<br />

the Southern Cone) nursery is<br />

given in Table 12.<br />

Durum Wheat Program<br />

In the countries <strong>of</strong> West Asia<br />

<strong>and</strong> North Africa (WANA), severe<br />

epidemics <strong>of</strong> septoria leaf blotch<br />

have occurred, especially in<br />

Morocco, Tunisia, <strong>and</strong> Turkey<br />

(Saari, 1974). Isolates <strong>of</strong> the<br />

pathogen have been identified in<br />

different regions, along with their<br />

specificity for durum <strong>and</strong> bread<br />

wheat. Breeding efforts have<br />

increased resistance levels, but<br />

selection for resistance is<br />

complicated by different disease<br />

reactions at different locations<br />

<strong>and</strong> in different years. This may<br />

be due to different environmental<br />

conditions or to the variability <strong>of</strong><br />

S. tritici populations (Arama et al.,<br />

1988); van Silfhout et al. (1989)<br />

have reported specificity in some<br />

durum <strong>and</strong> bread wheat isolates.<br />

Complementary information was<br />

<strong>Septoria</strong> tritici Resistance Sources <strong>and</strong> Breeding Progress at <strong>CIMMYT</strong>, 1970-99 137<br />

Table 8. Advanced wheat lines resistant to <strong>Septoria</strong> tritici <strong>and</strong> derived from crosses <strong>of</strong> perennial<br />

Triticeae species <strong>and</strong> Chinese germplasm.<br />

Lines Crosses S. tritici score a<br />

CS/Thinopyrum curvifolium//Glenn.81/3/Ald/Pvn CIGM84295 31-43<br />

Cno79/4/CS/Thinopyrum curvifolium// CIGM88829 32-62<br />

Glenn.81/3/Ald/Pvn<br />

Cno 79/4/CS/Thinopyrum curvifolium//Gen/3/ CIGM8876 32-42<br />

Ald/Pvn/4/CS/Leymus racemosus//2*CS/3/Cno 79<br />

Inia66/Thinopyrum distichum//Inia66/3//4/Gen/ CIGM88734 31-62<br />

CS/Thinopyrum curvifolium//Glenn.81/3/Ald/Pvn<br />

CS/Thinopyrum curvifolium//Glenn 81/3/Ald/Pvn/ CIGM87.116 51-72<br />

4/Ningmai#4/Oleson//Ald/Yangmai#4<br />

CS/Thinopyrum curvifolium/Glenn 81/3/ CIGM87.123 31-42<br />

Ald/Pvn/4/Suzhoe#8<br />

CS/Thinopyrum curvifolium//Glenn 81/3/ CIGM87.1109 41-53<br />

Ald/Pvn/4/Ningmai 8401<br />

Chir3/5/CS/Thinopyrum curvifolium//Glenn 81 CIGM93.612 21-22<br />

/3/Ald/Pvn/4/Cs/Leymus racemosus//2*CS/3/Cno79<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

Table 9. Triticum dicoccon/Kauz advanced lines with resistance to <strong>Septoria</strong> tritici.<br />

Advanced lines <strong>and</strong> pedigree Cross number S. tritici score a<br />

T. dicoccon PI 1254156/2*Kauz GRSS93SH18-0M-12SH-2M-1Y 11-51<br />

T. dicoccon PI 1254156/2*Kauz GRSS93SH18-0M-13SH-3M-1Y 11-21<br />

T. dicoccon PI 1254156/2*Kauz GRSS93SH18-0M-13SH-3M-1Y 11-21<br />

T. dicoccon PI 1254156/2*Kauz GRSS93SH27-OM-10SH-3M-1Y 11-72<br />

T. dicoccon PI 1254156/2*Kauz GRSS93SH29-0M-5SH-3M-1Y 32-73<br />

Kauz (susceptible parent) 89-99<br />

Bobwhite (moderately resistance check) 41-46<br />

T. dicoccon PI 1254156 (resistant parent) 11-71<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

Table 10. Advanced lines consisting <strong>of</strong> sources <strong>of</strong> <strong>Septoria</strong> tritici resistance in an adapted<br />

background, crossed with a group <strong>of</strong> resistant Chinese lines.<br />

Advanced line Cross S. tritici score a<br />

Shangai#5/Bow CM91100 32-72<br />

Ald/Pvn//YM#6 CM91065 21-72<br />

Suzhoe#6//Ald/Pvn CM91128 32-71<br />

YM#6/Thb*2 CMH87.2698 42-72<br />

Ald/Pvn//Ning#7840 CMH91325 11-47<br />

Sha#5/Weaver CM95103 11-32<br />

Sha#8/Gen CM95124 11-32<br />

Milan/Sha#7 CM97550 52-63<br />

Sabuf CM95073 53-32<br />

Catbird CM91045 42-76<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.


Session 6C — L. Gilchrist et al.<br />

138<br />

Table 11. Bread Wheat Program advanced lines with high yield potential containing different<br />

combinations <strong>of</strong> <strong>Septoria</strong> tritici resistance sources in their pedigrees.<br />

Advanced lines Cross S. tritici score a<br />

Trap#1/Bow CM 84548 21-35<br />

PF70354/Bow CM 67910 21-54<br />

Gov/Az//Mus/3/Dodo/4/Bow CM 79515 31-64<br />

Ald/Pvn//YMI#6/3/Ald/Pvn CM 96723 11-41<br />

Ias20*3/H567.71//Sara CM 81021 11-31<br />

Ning8675/Catbird CMSS92Y00639S 32-52<br />

Thb/CEP7780//Suz#9/Weaver/3/Ng8675 CMSS92Y02302T 52-63<br />

Cbrd/5/Cs/Thinopyrum curvifolium//<br />

Glen/3/Gen/4/L2266/1406.101//Buc/<br />

3/Vpm/Mos83.11.4.8//Nac<br />

CMBW91MO3723 11-21<br />

Sha3/Seri//Kauz/3/Sha4/Chil CMBW91M03723 11-21<br />

Bau/Milan CM103873 11-21<br />

Lfn/II58.57//PRL/3/Hahn CM77224 31-42<br />

Sha #3/Seri//Psn/Bow CMBW90M2470 31-42<br />

Suz #6/Weaver//Tui CMBW90M2474 11-21<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

Table 12. Bread wheat germplasm resistant to <strong>Septoria</strong> tritici that was selected through shuttle<br />

breeding <strong>and</strong> collaborative nurseries (<strong>CIMMYT</strong>/national agricultural research programs).<br />

S. tritici<br />

Lines <strong>and</strong> crosses score a Origin<br />

T00011/T00007<br />

A8972-1T-2N-1B-2T-2T-0T<br />

53 Argentina<br />

L.A.CIAT(Santa Cruz) 63 Bolivia<br />

Talhuen INIA 31 Chile<br />

Br14/CEP847<br />

B31615-0A-0Z-1A-15A-0A<br />

21 Brazil<br />

Iapi/Fink<br />

CP3409-1E-0Y-0E-18Y-0E<br />

31 Paraguay<br />

E Fed/F.5.83.7792 (Bajas) 42 Uruguay<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

found by Eyal et al. (1985), whose<br />

studies revealed that isolates from<br />

Syria <strong>and</strong> Tunisia were more<br />

virulent on tetraploid than on<br />

hexaploid varieties.<br />

Based on this information, elite<br />

durum wheat lines were selected in<br />

Toluca <strong>and</strong> Patzcuaro, as was<br />

described earlier for bread wheat. A<br />

representative group <strong>of</strong> these<br />

selections is listed in Table 13. This<br />

germplasm is under observation in<br />

the WANA region, but the dry<br />

conditions that have prevailed in<br />

recent years have not permitted<br />

accurate information to be obtained<br />

on the resistance <strong>of</strong> germplasm preselected<br />

in Mexico. However,<br />

information from Tunisia (Table 14)<br />

has contributed greatly to<br />

enhancing the effective resistance<br />

in durum wheat germplasm, <strong>and</strong><br />

the disease data from that location<br />

have been used to design crosses.<br />

Conclusion<br />

In conclusion, there are<br />

numerous sources <strong>of</strong> resistance to<br />

S. tritici. For the most part, the<br />

genetic constitution <strong>of</strong> these<br />

sources remains unclear, but the<br />

information contained in this paper<br />

suggests that they are genetically<br />

different. Resistance sources in<br />

<strong>CIMMYT</strong>’s genebank collections<br />

are available on request.<br />

References<br />

Arama, P.F., van Silfhout, C.H., <strong>and</strong><br />

Kema, G.H.J. 1988. Report on the<br />

Cooperative research project<br />

between IPO, Tel Aviv University<br />

<strong>and</strong> <strong>CIMMYT</strong>. 31 pp.<br />

Eyal, Z., Scharen, A.L., H<strong>of</strong>fman,<br />

M.D., <strong>and</strong> J.M. Prescott. 1985.<br />

Global insights into virulence<br />

frequencies <strong>of</strong> Mycosphaerella<br />

graminicola. Phytopathology<br />

75(12):1456-1462.<br />

Eyal, Z., A.L. Sharen, J.M. Prescott,<br />

<strong>and</strong> M. van Ginkel. 1987. The<br />

<strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Concepts <strong>and</strong> Methods <strong>of</strong> Disease<br />

Management. Mexico, D.F.:<br />

<strong>CIMMYT</strong>. 46 pp.<br />

Gilchrist, L., <strong>and</strong> Skovm<strong>and</strong>, B. 1995.<br />

Evaluation <strong>of</strong> emmer wheat<br />

(Triticum dicoccon) for resistance to<br />

<strong>Septoria</strong> tritici . In: Gilchrist, L., van<br />

Ginkel, M., McNab, A., <strong>and</strong> Kema,<br />

G.H.J., eds. Proceedings <strong>of</strong> a<br />

<strong>Septoria</strong> tritici workshop. Mexico,<br />

D.F.: <strong>CIMMYT</strong>. Pp 130-134.<br />

Mann, C.E., Rajaram, S., <strong>and</strong> R.L.<br />

Villarreal. 1985. Progress in<br />

breeding for <strong>Septoria</strong> tritici<br />

resistance in semidwarf spring<br />

wheat at <strong>CIMMYT</strong>. In: Sharen, A.L.<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>: Proceedings <strong>of</strong><br />

the workshop, held August 2-4,<br />

1983, at Montana State University,<br />

Bozeman, Montana. Pp. 22-26.


Mujeeb-Kazi, A., V. Rosas, <strong>and</strong> S.<br />

Roldan. 1996. Conservation <strong>of</strong> the<br />

genetic variation <strong>of</strong> Triticum<br />

tauschii (Coss) Schmalh. (Aegilops<br />

squarrosa auct. Non L.) in synthetic<br />

hexaploid wheats (T. turgidum L.s.<br />

lat X T. tauschii; 2n=6x=42,<br />

AABBDD) <strong>and</strong> its utilization for<br />

wheat improvement. Genetic<br />

Resources <strong>and</strong> Crop Evolution<br />

43:129-134.<br />

Mujeeb-Kazi, A., Gilchrist, L.,<br />

Villareal, R.L., <strong>and</strong> Delgado, R.<br />

1998. D-genome synthetic<br />

hexaploids: Production <strong>and</strong><br />

utilization in wheat improvement<br />

In: Triticeae III, Jaradat, A.A.(ed.),<br />

ICARDA, Aleppo. Pp. 369-374.<br />

Mujeeb-Kazi, A., Gilchrist, L.,<br />

Villareal, R.L., <strong>and</strong> Delgado, R.<br />

1999. Registration <strong>of</strong> ten wheat<br />

germplasm lines resistant to<br />

<strong>Septoria</strong> tritici leaf blotch. Crop<br />

Science (in press).<br />

Rajaram, S., van Ginkel, M., <strong>and</strong><br />

Fischer, R.A. 1994. <strong>CIMMYT</strong>’s<br />

wheat breeding megaenvironments<br />

(ME). In:<br />

Proceedings <strong>of</strong> the 8 th International<br />

Wheat Genetics Symposium, July<br />

19-24, Beijing, China. Pp 1101-1106.<br />

Saari, E.E. 1974. Results <strong>of</strong> studies on<br />

<strong>Septoria</strong> in the near East <strong>and</strong><br />

Africa. Proc. <strong>of</strong> Fourth FAO<br />

Rockefeller Foundation Wheat<br />

Seminar. Tehran, Iran, May 21-June<br />

2, 1973. Pp. 275-283.<br />

Saari, E.E., <strong>and</strong> Prescott, J.M. 1975. A<br />

scale for appraising the foliar<br />

intensity <strong>of</strong> wheat diseases. Plant.<br />

Dis. Rep. 59:377-380.<br />

van Silfhout, C.H., Arama, P.F., <strong>and</strong><br />

Kema, G.H.J. 1989. International<br />

survey <strong>of</strong> factors <strong>of</strong> virulence <strong>of</strong><br />

<strong>Septoria</strong> tritici. In: Fried, M.P., ed.,<br />

Proceedings, <strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>.<br />

Zurich, Switzerl<strong>and</strong>, July 4-7, 1989.<br />

<strong>Septoria</strong> tritici Resistance Sources <strong>and</strong> Breeding Progress at <strong>CIMMYT</strong>, 1970-99 139<br />

Table 13. Elite durum wheat lines with resistance to <strong>Septoria</strong> tritici selected at Toluca <strong>and</strong> Patzcuaro.<br />

Lines Cross S. tritici score a<br />

Aaz77 3/Olus CD 94270 21-22<br />

Ajaia 12/F3Local<br />

(Sel.Ethio.135.85)//Plata 13<br />

CD 98331 21-32<br />

Eco/CMH76A.722//Yav/3/<br />

Altar 84/4/Ajaia 2/5/Kjove 1<br />

CD 91B2636 32-42<br />

Gs/Cra//Sba81/3/Ho/4/Mex 1/<br />

5/ Memo/6/2*Altar 84<br />

CDSS 92B1193 32-33<br />

Himan9/ Bejah 6 CD91B2096 21-21<br />

Kucuk CD 91B2620 T-11<br />

Lhnke//Gs/Str/3/Altar 84/4/Focha 1 CDSS92B1076 21-22<br />

Liro 2 CD 93352 31-31<br />

Lotus 1/Espe CD97015 11-21<br />

Lymno 8 CD92314 21-42<br />

Nus/Silver//CMH82A.1062/<br />

Rissa/3/Chen/Altar84/4/Don<br />

CD91B2664 21-32<br />

Patka 3 CD 78995 11-21<br />

Plata1 Smn//Plata 9 CD97899 T- 11<br />

Plata10/6/Mque/4/Usda573//Qfn/<br />

Aa 7/3/Alba-D/5/Avohui/7/ Plata 1<br />

CD98581 32-43<br />

Plata 6/ Green17 CD96789 11-21<br />

Plata7/Fillo 9//Sbak CD99545 31-32<br />

Plata 8/4/Garza/Afn//Cra/<br />

3/Gta/5/Rascon<br />

CD91B2061 32-42<br />

Rascon 21/Long CDWS91M377 32-42<br />

Rascon 37/Green 2 CD91B1975 31-32<br />

Rascon 37/Tarro 2//Rascon 37 CDSS92B1022 31-32<br />

Rascon 39/Tilo 1 CDSS92B61 21-22<br />

Sn Turrk Mi 83-84 503/<br />

Lotus 4//Lotus 5<br />

CD97775 T - 11<br />

Sooty 9/Rascon 37 CD91B1938 31-32<br />

Sora/2*Plata 12 CD92011 T - 11<br />

Topdy 18/Focha 1//Altar 84 CDSS92B1034 22-32<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.<br />

Table 14. Durum wheat germplasm with resistance to <strong>Septoria</strong> tritici from Tunisia.<br />

Lines Cross S. tritici score a<br />

Karim (Check S ) CM9799 98-97<br />

Khiar (check S) CD57.005 97-97<br />

BD 2337 – 53-42<br />

BD 2338 – 64<br />

BD 2339 – 64<br />

Src2/Src1 ICD88.66 75-74<br />

Gdfl/ T. dicoccoides-<br />

SY20013/Bcr<br />

– 64-53<br />

Bcr/Guerou 1 ICD87.0572 75-76<br />

Zeina 2 ICD88.1233 44-32<br />

Zeina 4 ICD88.1233 64-32<br />

Aus 1/5/Cndo/4/Bry*2/<br />

Tace//II27655/3/Tme/Zb/2*W<br />

ICD88.1120 53-76<br />

Srn 3/Ajaia 15//Don 87 CD96855 73-42<br />

Porron 1 CD85328 54<br />

Poho/Ajaia CD99818 64<br />

Cali/ship 2// Fillo 7 CD98278 42<br />

Plata 6/Green 17 CD96789 52<br />

CMH82A.1062/3/Ggovz394//Sba81/<br />

Plc/4/Aaz 1/Crex/5/Hui//Cit71/CII<br />

CD97395 43<br />

a Double digit scale (Saari <strong>and</strong> Prescott, 1975; Eyal et al., 1987) indicating the range <strong>of</strong> damage during two<br />

screening cycles in Toluca.


140<br />

Selecting Wheat for Resistance to<br />

<strong>Septoria</strong>/<strong>Stagonospora</strong> in Patzcuaro, Michoacan, Mexico<br />

R.M. Gonzalez I., 1 S. Rajaram, 2 <strong>and</strong> M. van Ginkel2 1 Campo Morelia, National Livestock <strong>and</strong> Agricultural Research Institute, Mexico<br />

2 <strong>CIMMYT</strong> Bread Wheat Program<br />

Abstract<br />

The research reported in this paper was conducted in the humid, temperate area <strong>of</strong> Patzcuaro in the state <strong>of</strong> Michoacan, in<br />

Mexico, where there is a high natural incidence <strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong> spp. Nine wheat genotypes that had previously been<br />

selected for resistance to <strong>Septoria</strong>/<strong>Stagonospora</strong> spp. were included in the study. They were compared to two check varieties, one<br />

tolerant <strong>and</strong> one susceptible. The materials were tested with <strong>and</strong> without chemical protection. Three years’ data were analyzed <strong>and</strong><br />

relative yield losses <strong>of</strong> 10-32% were found. Two different responses were observed among the outst<strong>and</strong>ing genotypes. On the one<br />

h<strong>and</strong>, several lines expressed little reduction in yield when challenged by the septoria pathogens. Depending on their yield potential,<br />

final yield could be moderate to quite good. These could be classified as resistant or tolerant to <strong>Septoria</strong>/<strong>Stagonospora</strong> spp. Some <strong>of</strong><br />

these same materials did show a considerable level <strong>of</strong> severity; their response would thus be better classified as tolerance rather than<br />

resistance. On the other h<strong>and</strong>, certain materials with high yield potential did lose yield following the attack by the septoria foliar<br />

blights, but retained sufficient expression <strong>of</strong> yield potential to compete well with the local check variety. These should not be classified<br />

as resistant nor tolerant, but may well be desirable from a production st<strong>and</strong>point. The line that best combined these traits is IAS20/<br />

H567.71/5*IAS 20. It is being considered for release under the name Patzcuaro. Severity <strong>of</strong> foliar disease on the flag leaf proved to be<br />

well correlated with yield loss.<br />

Over the past 15 years, Mexico’s<br />

National Livestock <strong>and</strong><br />

Agricultural Research Institute<br />

(INIFAP) <strong>and</strong> <strong>CIMMYT</strong> have<br />

worked together on breeding bread<br />

wheats for resistance to <strong>Septoria</strong><br />

tritici <strong>and</strong> <strong>Stagonospora</strong> nodorum in<br />

the Patzcuaro, Michoacan, region <strong>of</strong><br />

Mexico. Natural conditions in this<br />

area favor the yearly development<br />

<strong>of</strong> these two <strong>Septoria</strong> species<br />

(Gomez <strong>and</strong> Gonzalez, 1987). The<br />

area enjoys a temperate climate,<br />

with mean temperatures <strong>of</strong> 8-21ºC,<br />

> 800 mm annual rainfall, <strong>and</strong> ><br />

85% relative humidity. The isolates<br />

<strong>of</strong> <strong>Septoria</strong>/<strong>Stagonospora</strong> spp. that<br />

thrive under these conditions are<br />

very aggressive, which makes this<br />

a stress environment for wheat.<br />

According to Eyal et al. (1985), the<br />

<strong>Septoria</strong> isolates present in the<br />

Patzcuaro area are among the most<br />

virulent in the world.<br />

Materials <strong>and</strong> Methods<br />

In this study, 11 advanced<br />

wheat lines from <strong>CIMMYT</strong><br />

nurseries (Table 1) targeted for high<br />

rainfall production environments<br />

(e.g., HRWSN <strong>and</strong> ASWSN) were<br />

evaluated with <strong>and</strong> without<br />

chemical protection during 1994-96.<br />

The variety Curinda M-87,<br />

which has functioned as reference<br />

in our work for the past 10 years,<br />

was used as the resistant/tolerant<br />

check variety in the trials (Table 1).<br />

The variety Batan was the<br />

susceptible check. Although both S.<br />

tritici <strong>and</strong> S. nodorum were present,<br />

the former was found in a higher<br />

proportion in most years.<br />

Fungicides Terbuconazole <strong>and</strong><br />

Tecto were used for chemical<br />

control. Terbuconazole was applied<br />

every 10 days starting from<br />

tillering to grain milk stage at a<br />

dosage <strong>of</strong> 0.5 l/ha. Tecto was<br />

applied in the same dosage every<br />

eight days, starting at the end <strong>of</strong><br />

the booting stage.<br />

Days to flowering, percentage<br />

flag leaf area affected by <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> spp. (% flag leaf<br />

severity, FLS) at grain milk stage,<br />

plant height, grain yield, <strong>and</strong> test<br />

weight were measured. Yield losses<br />

<strong>and</strong> correlations were calculated for<br />

individual years.<br />

Results <strong>and</strong> Discussion<br />

Several lines expressed higher<br />

levels <strong>of</strong> resistance than the<br />

resistant check variety Curinda<br />

(29% FLS) (Table 1). Severities as<br />

low as 1-3% were noted in<br />

individual years. Across the three


Selecting Wheat for Resistance to <strong>Septoria</strong>/<strong>Stagonospora</strong> in Patzcuaro, Michoacan, Mexico 141<br />

Table 1. Response <strong>of</strong> 11 wheat lines to natural infection by <strong>Septoria</strong>/<strong>Stagonospora</strong>, in regard to flag leaf severity<br />

(FLS in %), yield, percentage loss, in field plots either protected or unprotected with fungicide, in Pátzcuaro,<br />

Michoacan. México, during 1994, 1995, <strong>and</strong> 1996.<br />

Yield (kg/ha) %<br />

Entry Cross/Selection history CC * % ** UP % Yield loss % Flag leaf<br />

1 CURINDA<br />

Tolerant check<br />

2875 e *** 100 2042 d 100 29 28<br />

2 BATAN<br />

Susceptible check<br />

3521 b 122 2271 c 111 36 23<br />

3 IAS20/H567.71/5* IAS20<br />

CM78A-544-7B-1Y-1B-1Y-2B-1Y-0B<br />

3826 b 133 3694 a 181 4 1<br />

4 CHIL//ALD/PVN<br />

CM92801-65Y-0M-0Y-4M-0RES<br />

2820 e 98 1590 e 78 44 36<br />

5 THB/CNT 7<br />

CM 830057-0Z-0A-1A-1A-1A-0Y<br />

2632 f 92 2243 c 110 15 10<br />

6 LIRA/TAN//SPB<br />

CM96824-U-0Y-0H-0SY-4M-0RES<br />

2792 f 97 1847 d 90 34 18<br />

7 ALD/PVN//YMI 6<br />

CM 91065-2M-0M-0Y-0M-2Y-0B-5PZ<br />

3681 b 128 2681 b 131 27 3<br />

8 BAU/MILAN<br />

CM103873-2M-030Y-020Y-010M-4Y-0M<br />

3312 c 115 1986 d 97 40 15<br />

9 ISEPTON89-82E/BR6//PF83144<br />

F32938-2M-0AL-0AL-0AL-4Y-0M<br />

3049 d 106 3014 b 148 1 1<br />

10 BOW/MII//RES/NAC/3/PFAU<br />

CM1033644-2FC-0M-1FC-0C<br />

4432 a 154 2688 d 132 39 8<br />

11 CAR853/COC//VEE/3/E7408/PAM//HORK/PF73226<br />

CM80174-32Y-04M-0Y-3M-1M-0M<br />

3854 b 134 2292 d 112 41 24<br />

LSD 498 545<br />

Mean 3344 2396<br />

C.V. 10 16<br />

Yield (kg/ha) %<br />

Entry Cross/Selection history CC * % ** UP % Yield loss % Flag leaf<br />

1 CURINDA<br />

Tolerant check<br />

4153 c 100 3125 a 100 25 18<br />

2 BATAN<br />

Susceptible check<br />

3492 e 84 2285 c 73 35 53<br />

3 IAS20/H567.71/5* IAS20<br />

CM78A-544-7B-1Y-1B-1Y-2B-1Y-0B<br />

4521 b 108 3375 a 108 35 26<br />

4 CHIL//ALD/PVN<br />

CM92801-65Y-0M-0Y-4M-0RES<br />

4007 d 96 2722 b 87 32 45<br />

5 THB/CNT 7<br />

CM 830057-0Z-0A-1A-1A-1A-0Y<br />

4035 d 97 3153 a 101 22 21<br />

6 LIRA/TAN//SPB<br />

CM96824-U-0Y-0H-0SY-4M-0RES<br />

3854 d 92 2986 a 96 22 30<br />

7 ALD/PVN//YMI 6<br />

CM 91065-2M-0M-0Y-0M-2Y-0B-5PZ<br />

4722 b 114 2875 a 92 39 16<br />

8 BAU/MILAN<br />

CM103873-2M-030Y-020Y-010M-4Y-0M<br />

3952 d 95 3333 a 102 20 13<br />

9 ISEPTON89-82E/BR6//PF83144<br />

F32938-2M-0AL-0AL-0AL-4Y-0M<br />

4007 d 97 3368 a 108 17 5<br />

10 BOW/MII//RES/NAC/3/PFAU<br />

CM1033644-2FC-0M-1FC-0C<br />

5292 a 127 3236 a 104 39 13<br />

11 CAR853/COC//VEE/3/E7408/PAM//HORK/PF73226<br />

CM80174-32Y-04M-0Y-3M-1M-0M<br />

4549 b 110 3222 a 103 29 21<br />

LSD 484 592<br />

Mean 4235 3045<br />

C.V. 8 13<br />

* (CC) Chemical control <strong>and</strong> (UP) Unprotected.<br />

** Percentage relative to the tolerant check, Curinda M87.<br />

*** Values followed by the same letter are not significantly different at the 0.05 probability level.<br />

1994<br />

1995


Session 6C — R.M. Gonzalez I., S. Rajaram, <strong>and</strong> M. van Ginkel<br />

142<br />

Table 1. Continued...<br />

Yield (kg/ha) %<br />

Entry Cross/Selection history CC * % ** UP % Yield loss % Flag leaf<br />

1 CURINDA<br />

Tolerant check<br />

2896 c 100 2826 c 100 3 40<br />

2 BATAN<br />

Suscepible check<br />

2038 e 70 1637 f 57 19 61<br />

3 IAS20/H567.71/5* IAS20<br />

CM78A-544-7B-1Y-1B-1Y-2B-1Y-0B<br />

3359 b 116 3173 a 112 6 27<br />

4 CHIL//ALD/PVN<br />

CM92801-65Y-0M-0Y-4M-0RES<br />

2644 d 91 2289 e 81 13 75<br />

5 THB/CNT 7<br />

CM 830057-0Z-0A-1A-1A-1A-0Y<br />

3206 b 111 2766 c 98 14 31<br />

6 LIRA/TAN//SPB<br />

CM96824-U-0Y-0H-0SY-4M-0RES<br />

2710 d 94 2655 c 94 2 62<br />

7 ALD/PVN//YMI 6<br />

CM 91065-2M-0M-0Y-0M-2Y-0B-5PZ<br />

3961 a 137 3291 a 116 17 28<br />

8 BAU/MILAN<br />

CM103873-2M-030Y-020Y-010M-4Y-0M<br />

3028 c 105 2326 d 82 23 10<br />

9 ISEPTON89-82E/BR6//PF83144<br />

F32938-2M-0AL-0AL-0AL-4Y-0M<br />

3146 b 109 2847 b 101 10 22<br />

10 BOW/MII//RES/NAC/3/PFAU<br />

CM1033644-2FC-0M-1FC-0C<br />

3456 b 119 3185 a 113 26 38<br />

11 CAR853/COC//VEE/3/E7408/PAM//HORK/PF73226<br />

CM80174-32Y-04M-0Y-3M-1M-0M<br />

3156 b 110 2852 b 101 28 33<br />

LSD 412 345<br />

Mean 3094 2713<br />

C.V. 9 11<br />

* (CC) Chemical control <strong>and</strong> (UP) Unprotected.<br />

** Percentage relative to the tolerant check, Curinda M87.<br />

*** Values followed by the same letter are not significantly different at the 0.05 probability level.<br />

years Curinda, when unprotected,<br />

yielded on average 2688 kg/ha,<br />

expressing a 19% loss (Table 2).<br />

Several lines were also superior to<br />

the check in yield, <strong>and</strong> suffered less<br />

yield loss.<br />

In this study percent flag leaf<br />

area affected correlated negatively<br />

with yield loss (-0.71 to -0.83;<br />

Table 3). This negative correlation<br />

appears strong <strong>and</strong> reliable over<br />

years.<br />

Although grain yield <strong>of</strong> most<br />

lines tested varied over years<br />

depending on climatic <strong>and</strong><br />

production conditions (Figure 1), in<br />

general their performance relative<br />

to the resistant check Curinda<br />

showed impressive consistency<br />

over time. In contrast, the<br />

susceptible check Batan yielded<br />

less every year as a percentage <strong>of</strong><br />

the check (Tables 1 <strong>and</strong> 2; Figure 1),<br />

<strong>and</strong> seemed to be losing its<br />

resistance progressively.<br />

The plots protected with<br />

fungicide provide a partial<br />

expression <strong>of</strong> each line’s yield<br />

potential in the absence <strong>of</strong> disease<br />

stress. However, other abiotic<br />

stresses, such as soil acidity <strong>and</strong><br />

nutrient imbalances related to<br />

leaching following excessive rain,<br />

also reduce yield in this high<br />

rainfall environment. The<br />

difference between the protected<br />

<strong>and</strong> unprotected plots provides a<br />

measure <strong>of</strong> yield loss due to<br />

disease. Mean yield losses were in<br />

the 10-32% range. The line<br />

ISEPTON89-82E/BR6//PF83144<br />

experienced the lowest loss (10%),<br />

while the susceptible check Batan<br />

suffered the highest (32%).<br />

1996<br />

Comparing protected <strong>and</strong><br />

unprotected plots provides a measure<br />

<strong>of</strong> the contribution <strong>of</strong> resistance or<br />

tolerance to yield. If the yield<br />

difference is small (i.e. the loss is<br />

small), the genotype tested<br />

apparently has a certain level <strong>of</strong><br />

resistance/tolerance that enables it to<br />

express most, if not all, <strong>of</strong> its yield<br />

potential. However, useful<br />

information can also be gained from<br />

direct comparison <strong>of</strong> lines in a<br />

disease-prone setting, in companion<br />

plots without the use <strong>of</strong> fungicides, in<br />

particular when screening germplasm<br />

to identify potential varieties. This is<br />

less expensive than also growing<br />

protected plots, <strong>and</strong> hence allows<br />

larger numbers <strong>of</strong> entries to be tested.<br />

To be acceptable to farmers, a<br />

variety must yield the same or better<br />

than the commercial check variety.<br />

C<strong>and</strong>idate varieties can be acceptable


<strong>and</strong> appear well adapted to the<br />

point <strong>of</strong> being releasable to farmers<br />

for two reasons, or a combination <strong>of</strong><br />

both. On the one h<strong>and</strong>, a line may<br />

express true resistance or tolerance.<br />

In that case challenge by the<br />

pathogen does not result in great<br />

reductions in yield (Tables 1 <strong>and</strong> 2).<br />

Yield potential is roughly<br />

maintained due to resistance, with<br />

only minimum disease severity, or<br />

due to tolerance when considerable<br />

disease severity may be noted but<br />

losses are little. The line<br />

ISEPTON89-82E/BR6//PF83144 is a<br />

good example <strong>of</strong> a resistant<br />

Selecting Wheat for Resistance to <strong>Septoria</strong>/<strong>Stagonospora</strong> in Patzcuaro, Michoacan, Mexico 143<br />

genotype (9% FLS) with minimum<br />

yield loss (10%) <strong>and</strong> a high mean<br />

yield <strong>of</strong> 3034 kg/ha, 13% more<br />

than the tolerant check, Curinda,<br />

in the presence <strong>of</strong> disease. Most<br />

lines suffered yield losses which<br />

percentage-wise were roughly<br />

equivalent to about two-thirds <strong>of</strong><br />

their percentage-wise flag leaf<br />

severity scores. Good examples <strong>of</strong><br />

very sensitive lines are ALD/<br />

PVN//YMI 6 <strong>and</strong> BAU/MILAN,<br />

whose mean % yield losses (27%-<br />

28%) were almost twice that <strong>of</strong><br />

their FLS scores (13%-16%).<br />

Table 2. Mean response <strong>of</strong> 11 wheat lines to natural infection by <strong>Septoria</strong>/<strong>Stagonospora</strong>, in regard to flag leaf<br />

severity, yield, percentage loss, in field plots either protected or unprotected with fungicide, in Patzcuaro,<br />

Michoacan, Mexico, over 1994, 1995, <strong>and</strong> 1996.<br />

Mean values<br />

% Overall<br />

Yield mean yield<br />

Entry Cross/Selection history CC * % ** UP % loss % Flag leaf kg/ha %<br />

1 CURINDA<br />

Tolerant check<br />

3376 c *** 100 2688 d 100 19 29 3005 c 100<br />

2 BATAN<br />

Susceptible check<br />

3016 e 89 2003 h 75 32 46 2470 f 82<br />

3 IAS20/H567.71/5* IAS20<br />

CM78A-544-7B-1Y-1B-1Y-2B-1Y-0B<br />

4094 b 121 3380 a 126 13 18 3709 a 123<br />

4 CHIL//ALD/PVN<br />

CM92801-65Y-0M-0Y-4M-0RES<br />

3062 d 91 2216 g 82 30 52 2606 d 87<br />

5 THB/CNT 7<br />

CM 830057-0Z-0A-1A-1A-1A-0Y<br />

3298 c 98 2727 c 101 17 21 2991 c 99<br />

6 LIRA/TAN//SPB<br />

CM96824-U-0Y-0H-0SY-4M-0RES<br />

3098 d 91 2519 e 94 20 37 2786 d 93<br />

7 ALD/PVN//YMI 6<br />

CM 91065-2M-0M-0Y-0M-2Y-0B-5PZ<br />

4160 a 123 2998 b 112 28 16 3534 a 118<br />

8 BAU/MILAN<br />

CM103873-2M-030Y-020Y-010M-4Y-0M<br />

3449 c 102 2473 f 92 27 13 2924 d 97<br />

9 ISEPTON89-82E/BR6//PF83144<br />

F32938-2M-0AL-0AL-0AL-4Y-0M<br />

3486 c 103 3034 b 113 10 9 3242 b 108<br />

10 BOW/MII//RES/NAC/3/PFAU<br />

CM1033644-2FC-0M-1FC-0C<br />

4441 a 132 3058 b 114 31 20 3696 a 123<br />

11 CAR853/COC//VEE/3/E7408/PAM//HORK/PF73226<br />

CM80174-32Y-04M-0Y-3M-1M-0M<br />

3870 b 115 2798 b 104 28 26 3292 b 109<br />

LSD 346 312 314<br />

Mean 3577 2717 3114<br />

C.V. 12 15 18<br />

Chemical control (CC) 3577 a<br />

Unprotected (UP) 2717 b<br />

1994 3344 b ns<br />

1995 4234 a ns<br />

1996 3153 c ns<br />

LSD 181<br />

* (CC) Chemical control <strong>and</strong> (UP) Unprotected.<br />

** Percentage relative to the tolerant check, Curinda M87.<br />

*** Values followed by the same letter are not significantly different at the 0.05 probability level.<br />

On the other h<strong>and</strong>, a line may<br />

have a very high intrinsic yield<br />

potential in the absence <strong>of</strong> biotic<br />

stresses, such as attacks by <strong>Septoria</strong>/<br />

<strong>Stagonospora</strong> spp. Even if<br />

susceptible to the fungus, such<br />

lines may express a residual yield<br />

that competes well with that <strong>of</strong> the<br />

check variety (<strong>and</strong> may even<br />

outyield it in certain cases). These<br />

lines are not resistant—on the<br />

contrary, they may be quite<br />

susceptible. Their virtue lies in<br />

their high initial yield potential,<br />

which allows a considerable loss<br />

due to disease, <strong>and</strong> still produces


Session 6C — R.M. Gonzalez I., S. Rajaram, <strong>and</strong> M. van Ginkel<br />

144<br />

an impressive final yield when<br />

challenged by the fungus. In this<br />

study BOW/MII//RES/NAC/3/<br />

PFAU is an example <strong>of</strong> a line with<br />

high yield potential, which despite<br />

having 20% <strong>of</strong> its flag leaf area<br />

damaged by <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

spp., resulting in a 31% mean yield<br />

loss, still yielded 14% more than<br />

the commercial check.<br />

The most desirable genotype is<br />

one that combines high yield<br />

potential with resistance/tolerance<br />

to the fungus; this results in<br />

relatively low losses in the presence<br />

;<br />

200<br />

;;;<br />

% yield <strong>of</strong> check Curinda<br />

150<br />

100<br />

CC 1994<br />

CC 1995<br />

CC 1996<br />

UP 1994<br />

UP 1995<br />

UP 1996<br />

<strong>of</strong> disease, <strong>and</strong> outst<strong>and</strong>ing yields<br />

in its absence. The line IAS20/<br />

H567.71/5*IAS 20 combines these<br />

traits; only 18% <strong>of</strong> its foliage was<br />

affected, which translated into a<br />

minimal yield loss <strong>of</strong> 13% <strong>and</strong><br />

resulted in the trial’s top mean<br />

yield <strong>of</strong> 3380 kg/ha, 26% over the<br />

commercial check, Curinda, in the<br />

presence <strong>of</strong> disease (Table 2 <strong>and</strong><br />

Figure 1). It produced a very<br />

respectable yield <strong>of</strong> 4094 kg/ha, the<br />

third highest <strong>of</strong> the group, in the<br />

absence <strong>of</strong> disease. This line is now<br />

being considered for release under<br />

the name Patzcuaro.<br />

References<br />

Table 3. Correlation coefficients <strong>of</strong> flag leaf severity <strong>and</strong> yield with various other characters, Patzcuaro,<br />

Michoacan, Mexico, during 1994, 1995, <strong>and</strong> 1996.<br />

Eyal, Z., A.L. Scharen, M.D. Huffman,<br />

<strong>and</strong> J.M. Prescott. 1985. Global<br />

insights into virulence frequencies<br />

<strong>of</strong> Mycosphaerella graminicola.<br />

Phytopathology 75:1456-1462.<br />

Gomez, B.L., <strong>and</strong> R.M. Gonzalez I.<br />

1987. Mejoramiento genético de<br />

trigos harineros para resistencia a<br />

<strong>Septoria</strong> tritici en el área de<br />

temporal húmedo en Mexico. In<br />

Conferencia regional sobre la<br />

septoriosis del trigo, M.M. Kohli<br />

<strong>and</strong> L.T. van Beuningen, eds.<br />

Mexico, D.F.: <strong>CIMMYT</strong>. pp 42-57.<br />

Variables Variables 1994 1995 1996<br />

Flag leaf severity due to infection Yield -0.81 -0.83 -0.71<br />

by <strong>Septoria</strong>/<strong>Stagonospora</strong> Test weight -0.72 -0.34 -0.44<br />

Plant height -0.48 -0.55 -0.62<br />

Yield loss (%) 0.71 0.33 0.71<br />

Days to flowering -0.42 -0.71 -0.84<br />

Yield Test weight 0.72 0.85 0.76<br />

Plant height 0.62 0.44 0.55<br />

Yield loss (%) -0.75 -0.49 -0.86<br />

Days to flowering 0.46 0.86 0.79<br />

;<br />

; ; ;<br />

; ; ; ; ;<br />

;<br />

; ; ; ; ;; ; ; ;<br />

; ;; ; ; ; ; ;; ; ; ;<br />

; ; ; ; ;; ; ; ;<br />

; ;; ; ;; ; ; ; ; ; ; ; ; ;<br />

; ;;<br />

; ;; ;; ; ;<br />

; ;; ;; ; ; ;<br />

; ; ; ; ; ; ;<br />

;<br />

; ; ; ; ; ; ;<br />

; ;;;<br />

; ; ; ; ; ;<br />

;;; ; ;<br />

; ; ;<br />

;;<br />

;; ; ; ;<br />

; ;; ; ; ; ;<br />

; ;; ; ; ; ;<br />

; ;; ; ; ; ;<br />

50 Curinda Batan 3 4 5 6 7 8 9 10 11<br />

Figure 1. Yield as percentage <strong>of</strong> the tolerant check Curinda, in chemically controlled plots (CC) <strong>and</strong><br />

unprotected plots (UP) <strong>of</strong> nine wheat lines, during 1994, 1995 <strong>and</strong> 1996. Patzcuaro, Michoacan, Mexico.


Varieties <strong>and</strong> Advanced Lines Resistant to <strong>Septoria</strong><br />

<strong>Diseases</strong> <strong>of</strong> Wheat in Western Australia<br />

R. Loughman, 1 R.E. Wilson, 1 I.M. Goss, 1 D.T. Foster, 1 <strong>and</strong> N.E.A. Murphy2 1 Agriculture Western Australia, Bentley Delivery Centre, Western Australia<br />

2 WA State Agriculture Biotechnology Centre, Division <strong>of</strong> Science <strong>and</strong> Engineering, Murdoch University,<br />

Murdoch, Australia<br />

Abstract<br />

The development <strong>of</strong> bread wheats in Western Australia has resulted in the release <strong>of</strong> several varieties with improved<br />

response to the commonly occurring leaf spot diseases caused by Phaeosphaeria nodorum, Mycosphaerella graminicola,<br />

<strong>and</strong> Pyrenophora tritici-repentis. In a yield trial evaluating resistance to M. graminicola, susceptible varieties suffered 20-<br />

50% yield loss, while yield loss <strong>of</strong> resistant crossbreds <strong>and</strong> varieties was 0-15%. Similar responses were observed in most lines<br />

for grain weight when assessed in M. graminicola-infected row plots. In P. nodorum-infected row plots, susceptible<br />

varieties suffered 30-50% grain weight loss, while resistant lines suffered 0-30% loss in grain weight depending on year<br />

<strong>of</strong> testing.<br />

In Western Australia resistance<br />

to Phaeosphaeria nodorum is sought<br />

in combination with Mycosphaerella<br />

graminicola <strong>and</strong> Pyrenophora triticirepentis<br />

resistance, as these<br />

pathogens frequently occur in<br />

combination (Loughman et al.,<br />

1994). Improved resistance to these<br />

diseases is sought by continually<br />

crossing the most resistant, best<br />

adapted lines with resistance<br />

donors. A selection <strong>of</strong> varieties,<br />

advanced lines, <strong>and</strong> potential<br />

resistance sources were assessed for<br />

resistance in replicated field<br />

experiments using susceptible<br />

adapted wheats with comparable<br />

maturities as benchmarks.<br />

Materials <strong>and</strong> Methods<br />

Lines were evaluated for<br />

response to infection with M.<br />

graminicola or P. nodorum in<br />

separate experiments in 1995-96<br />

using split plot designs <strong>of</strong> three<br />

replicates. At Mt. Barker lines were<br />

sown as 5-row, 5-m subplots. Main<br />

plots were established in early<br />

spring as either fungicide<br />

protected (three applications <strong>of</strong><br />

125 g tebuconazole/ha) or<br />

inoculated twice with aqueous<br />

pycnidiospore suspensions <strong>of</strong> the<br />

pathogen. Growth stage (Zadoks<br />

et al., 1974) <strong>and</strong> % infection <strong>of</strong> the<br />

flag-3 leaves were assessed in late<br />

spring. Yields were assessed. At<br />

South Perth lines were sown as<br />

40-cm, 2-row subplots with 18-cm<br />

row spacings, each subplot<br />

separated from its neighbor by a<br />

single row <strong>of</strong> barley. Main plots<br />

were established with fungicide<br />

or inoculation as above. Separate<br />

experiments were established for<br />

M. graminicola <strong>and</strong> P. nodorum.<br />

Heading date (as day <strong>of</strong> the year)<br />

<strong>and</strong> disease severity as percent<br />

infection <strong>of</strong> the flag or flag <strong>and</strong><br />

second top leaves were recorded<br />

in spring. Grain weights were<br />

assessed from a r<strong>and</strong>om<br />

subsample <strong>of</strong> harvested grain.<br />

Relative grain weights <strong>of</strong> infected<br />

plots were calculated as a<br />

percentage <strong>of</strong> fungicide<br />

protected plots.<br />

Results <strong>and</strong><br />

Discussion<br />

Resistance to M.<br />

graminicola<br />

Cascades (Tadorna.Inia /<br />

3*Aroona), Tammin (Bod/Eradu<br />

sib.xbvt//Atlas66.2Madden),<br />

<strong>and</strong> Nyabing (Ar-<br />

3Ag3(WT329)/<br />

(IW753)Jab.Ag4C) developed<br />

low to moderate infection <strong>and</strong><br />

yielded 5.4, 5.2, <strong>and</strong> 4.1 t/ha,<br />

respectively in the presence <strong>of</strong><br />

disease. These yields were not<br />

significantly different from<br />

fungicide sprayed plots (yield<br />

diseased was 89-99% relative to<br />

the fungicide protected plots). In<br />

contrast, two susceptible<br />

varieties <strong>of</strong> similar maturity,<br />

Gamenya <strong>and</strong> Datatine,<br />

developed high infection scores<br />

<strong>and</strong> yielded 1.5 <strong>and</strong> 3.2 t/ha,<br />

respectively when diseased.<br />

These yields were significantly<br />

different from fungicide sprayed<br />

plots (yield diseased was 48-72%<br />

relative to fungicide protected).<br />

145


Session 6C — R. Loughman, R.E. Wilson, I.M. Goss, D.T. Foster, <strong>and</strong> N.E.A. Murphy<br />

146<br />

Table 1. Maturity (growth stage), percent leaf disease (Dis), yield, hundred-grain weight (HGW), <strong>and</strong> relative<br />

yield (RY) <strong>and</strong> relative grain weight (RGW) <strong>of</strong> wheat varieties infected with Mycosphaerella graminicola in a<br />

yield trial at Mt. Barker <strong>and</strong> small row plots at South Perth, 1995. Comparisons for percent flag leaf disease<br />

should be made between varieties <strong>of</strong> similar maturity.<br />

Mt. Barker South Perth<br />

Growth Dis Yield(t/ha) RY Dis HGW (g) RGW<br />

stage Flag-3 Fung Inoc % Flag Fung Inoc %<br />

Resistant <strong>and</strong> moderately resistant lines<br />

77Z893 44 2 5.4 5.3 98 86 3.6 2.5 70<br />

Nyabing 42 30 4.1 4.1 99 38 4.2 3.4 81<br />

Cascades 41 23 6.0 5.4 89 86 4.2 2.5 57<br />

Tammin 41 3 5.8 5.2 90 23 4.2 3.1 76<br />

483W3 40 0 5.1 5.2 101 21 3.2 3.4 105<br />

486W189 39 0 4.6 4.3 92 32 3.9 4.2 106<br />

78Z976 39 5 5.7 5.0 89 40 3.1 2.7 87<br />

Calingiri 38 1 5.9 5.0 85 43 4.0 3.7 94<br />

Resistant <strong>and</strong> moderately resistant control varieties<br />

Millewa 43 9 4.5 3.9 86 78 3.3 2.2 69<br />

Aroona 40 13 3.9 4.4 114 99 3.6 2.5 72<br />

Corrigin 40 2 5.7 5.2 91 70 3.1 2.5 80<br />

Susceptible control varieties<br />

Kulin 55 85 4.0 2.2 55 99 2.7 1.6 63<br />

Tincurrin 42 90 4.5 3.6 80 99 3.6 2.2 68<br />

Datatine 41 50 4.5 3.2 72 73 3.4 2.5 72<br />

Gamenya 41 88 3.2 1.5 48 99 3.6 1.9 59<br />

Stiletto 38 23 4.5 3.2 70 98 4.1 2.8 69<br />

lsd 5% 2.5 23 1.0 32 0.7 23<br />

cv % 4 41 13 44 13 19<br />

Table 2. Maturity (heading day <strong>of</strong> year, Hd), percent leaf disease (Dis), hundred-grain weight (HGW), <strong>and</strong><br />

relative grain weight (RGW) <strong>of</strong> wheat varieties infected with Phaeosphaeria nodorum in short row plot<br />

experiments in South Perth, 1995-96 (- denotes testing discontinued). Comparisons for percent flag leaf<br />

disease should be made between varieties <strong>of</strong> similar maturity.<br />

1995 1996<br />

Hd Dis HGW (g) RGW Dis HGW(g) RGW<br />

d.o.y.<br />

Moderately resistant lines<br />

Flag Fung Inoc % F,F-1 Fung Inoc %<br />

Brookton 225 64 3.6 3.7 101 52 3.5 2.8 80<br />

W48795 228 46 3.6 3.6 101 4 3.8 2.9 76<br />

81W1137 230 43 3.8 3.6 97 - - - -<br />

Tammin 231 32 3.9 3.9 103 8 4.3 3.0 71<br />

W48433 232 33 3.2 3.1 98 30 3.8 3.0 81<br />

Resistant <strong>and</strong> moderately resistant controls<br />

ALMRES-83-18 225 72 4.3 3.6 83 56 3.9 2.8 74<br />

CNT2 236 21 4.6 4.2 93 45 4.1 3.9 103<br />

Aus20917 236 40 3.1 2.5 81 27 2.8 2.4 87<br />

Qualset601-68<br />

Intermediate lines<br />

238 18 3.6 3.0 85 20 4.0 2.9 73<br />

W486110 220 50 4.9 4.2 87 39 4.0 3.3 83<br />

Westonia 220 83 4.1 3.3 80 76 3.2 2.2 69<br />

Cascades 226 83 4.1 3.2 80 83 3.7 2.0 54<br />

Carnamah 226 68 3.4 2.9 86 32 4.1 2.1 51<br />

Cunderdin 228 50 3.7 3.0 82 41 2.9 2.4 81<br />

Susceptible control varieties<br />

Kulin 221 99 3.4 2.2 66 91 3.2 1.5 46<br />

Aroona 225 87 4.6 2.2 49 92 3.2 1.6 50<br />

Millewa 225 96 2.6 2.0 77 97 3.3 1.3 44<br />

Gamenya 228 77 3.4 2.1 64 51 3.2 2.0 63<br />

Cranbrook 236 71 3.1 2.5 80 48 3.4 2.0 61<br />

Spear 242 59 4.2 2.6 63 1 4.1 3.0 75<br />

lsd 5% 23 0.6 24 18 0.8 28<br />

cv % 26 10 19 23 16 26


Calingiri (Chino/Kulin//<br />

Reeves) also exhibited resistance<br />

<strong>and</strong> while some <strong>of</strong> this low disease<br />

expression may have been due to<br />

later maturity, the performance was<br />

measurably superior to the similar<br />

maturing but susceptible variety<br />

Stiletto. Crossbreds 483W3, 78Z976<br />

<strong>and</strong> 77Z893 (Corrigin sib) also<br />

expressed high levels <strong>of</strong> resistance,<br />

as indicated by low disease scores,<br />

the non-significant yield losses, <strong>and</strong><br />

the resulting high relative yields<br />

(yield diseased was 89-101%<br />

relative to fungicide protected)<br />

(Table 1).<br />

Resistance to P. nodorum<br />

Five moderately resistant lines<br />

achieved high grain weights in<br />

diseased plots relative to fungicide<br />

protected plots. Variety Brookton<br />

(Torres/Cranbrook//76W596/<br />

Cranbrook) developed significantly<br />

less disease <strong>and</strong> no significant<br />

grain weight reduction when<br />

compared with susceptible<br />

varieties <strong>of</strong> equivalent maturity<br />

such as Aroona or Millewa.<br />

W48795, Tammin, <strong>and</strong> sister line<br />

81W1137 had similar maturity to<br />

susceptible variety Gamenya.<br />

These lines developed significantly<br />

less disease than Gamenya <strong>and</strong> had<br />

smaller grain weight reductions in<br />

the presence <strong>of</strong> disease. Line<br />

W48433, derived from CNT2, was<br />

similar to Tammin <strong>and</strong> developed<br />

low disease in 1995 <strong>and</strong> high<br />

relative grain weights in 1995 <strong>and</strong><br />

1996. This line was comparable<br />

with the resistant parent CNT2 for<br />

disease severity <strong>and</strong> relative grain<br />

weight in 1995, though relative<br />

grain weight appeared less robust<br />

than that <strong>of</strong> CNT2 in 1996 (Table 2).<br />

Varieties <strong>and</strong> Advanced Lines Resistant to <strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat in Western Australia 147<br />

Westonia (SpicaTimg.Tosc/<br />

Cr:J2.Bob), Cunderdin (Cranbrook<br />

sib/Sunfield sib), Carnamah<br />

(Bolsena-1ch(RAC529)/77W660),<br />

<strong>and</strong> Cascades gave responses that<br />

were intermediate between the<br />

resistant <strong>and</strong> susceptible lines.<br />

Although hundred grain weight<br />

reductions in the presence <strong>of</strong><br />

disease were generally significant,<br />

the grain weight reductions were<br />

less than for susceptible lines.<br />

Although excellent levels <strong>of</strong> M.<br />

graminicola resistance from diverse<br />

sources have been observed in lines<br />

such as 483W3 <strong>and</strong> 78Z976,<br />

moderate resistance derived from<br />

WW15 via Aroona <strong>and</strong> Condor<br />

(Wilson, 1994) has had the greatest<br />

impact in the development <strong>of</strong><br />

commercial varieties. While this<br />

moderate resistance can be more<br />

difficult to identify in intensive<br />

disease nurseries, its effect has been<br />

observed <strong>of</strong>ten in practice <strong>and</strong><br />

previously reported (Loughman<br />

<strong>and</strong> Thomas, 1992).<br />

Significant progress has been<br />

made in developing moderate<br />

resistance to P. nodorum.<br />

Commercial varieties currently<br />

exhibit improved responses<br />

compared with their susceptible<br />

predecessors, <strong>and</strong> some varieties<br />

<strong>and</strong> advanced lines perform at<br />

similar levels to resistant control<br />

varieties. In some cases moderate<br />

resistance to different leaf spot<br />

diseases has been combined.<br />

Cascades combines moderate<br />

resistance to both M. graminicola<br />

<strong>and</strong> P. tritici-repentis (Loughman et<br />

al., 1998). Brookton, Cunderdin,<br />

<strong>and</strong> Westonia have moderate<br />

resistance to P. tritici-repentis<br />

(Loughman et al., 1998), in<br />

combination with partial resistance<br />

to P. nodorum. Because P. nodorum,<br />

M. graminicola, <strong>and</strong> P. tritici-repentis<br />

frequently occur as disease<br />

complexes in Western Australia,<br />

combinations <strong>of</strong> resistance are<br />

important for effective disease<br />

management.<br />

References<br />

Loughman, R., <strong>and</strong> Thomas, G.J. 1992.<br />

Fungicides <strong>and</strong> cultivar control <strong>of</strong><br />

<strong>Septoria</strong> diseases <strong>of</strong> wheat. Crop<br />

Protection 11: 349-54.<br />

Loughman, R., Wilson, R.E., <strong>and</strong><br />

Thomas, G.J. 1994. Influence <strong>of</strong><br />

disease complexes involving<br />

Leptosphaeria (<strong>Septoria</strong>) nodorum on<br />

detection <strong>of</strong> resistance to three leaf<br />

spot diseases in wheat. Euphytica<br />

72: 31-42.<br />

Loughman, R., Wilson, R.E., Roake,<br />

J.E., Platz, G.J., Rees, R.G., <strong>and</strong><br />

Ellison, F.W. 1998. Crop<br />

management <strong>and</strong> breeding for<br />

control <strong>of</strong> Pyrenophora triticirepentis<br />

causing yellow spot <strong>of</strong><br />

wheat in Australia. In Duveiller, E.,<br />

H.J. Dubin, J. Reeves, <strong>and</strong> A.<br />

McNab (eds.). Helminthosporium<br />

Blights <strong>of</strong> Wheat: Spot Blotch <strong>and</strong><br />

Tan Spot. Mexico, D.F.: <strong>CIMMYT</strong>.<br />

Wilson, R.E. 1994. Progress toward<br />

breeding for resisance to the two<br />

septoria disease <strong>of</strong> wheat in<br />

Australia. Pp. 149-152 In: E.<br />

Arseniuk (ed.). <strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong>,<br />

Proc. Fourth International<br />

Workshop on <strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong><br />

<strong>Cereals</strong>. IHAR, Radzikow, Pol<strong>and</strong>.<br />

Zadoks, J.C., Chang, T.T., <strong>and</strong><br />

Konzak, C.F. 1974. A decimal code<br />

for the growth stages <strong>of</strong> cereals.<br />

Weed Research 14: 415-21.


148<br />

Field Resistance <strong>of</strong> Wheat to <strong>Septoria</strong> Tritici Leaf Blotch,<br />

<strong>and</strong> Interactions with Mycosphaerella graminicola<br />

Isolates<br />

J.K.M. Brown, 1 G.H.J. Kema, 2 H.-R. Forrer, 3 E.C.P. Verstappen, 2 L.S. Arraiano, 1 P.A. Brading, 1 E.M. Foster, 1 A.<br />

Hecker, 3 <strong>and</strong> E. Jenny3 1 John Innes Centre, Norwich, UK<br />

2 DLO-Research Institute for Plant Protection, Wageningen, The Netherl<strong>and</strong>s<br />

3 Swiss Federal Research Station for Agroecology <strong>and</strong> Agriculture, Zürich, Switzerl<strong>and</strong><br />

Abstract<br />

The resistance <strong>of</strong> 71 varieties <strong>of</strong> bread wheat to six isolates <strong>of</strong> Mycosphaerella graminicola was studied in field trials in<br />

the Netherl<strong>and</strong>s, Switzerl<strong>and</strong>, <strong>and</strong> the UK, carried out over three years. There was a wide range <strong>of</strong> <strong>Septoria</strong> tritici infection.<br />

Some varieties had especially good resistance, including lines from Europe (especially Switzerl<strong>and</strong>) <strong>and</strong> Latin America. Many<br />

interactions between varieties <strong>and</strong> isolates were detected. In particular, 27 varieties <strong>of</strong> diverse origins were specifically<br />

resistant to the isolate IPO323. Variety-by-isolate interactions were stable over years <strong>and</strong> locations. The existence <strong>of</strong> these<br />

interactions <strong>and</strong> the fact that they are stable over environments implies that certain widely used resistances to <strong>Septoria</strong> tritici<br />

might break down through the evolution <strong>of</strong> specific virulence in the fungus. Breeders should take these interactions into<br />

account in their efforts to develop varieties with durable resistance.<br />

<strong>Septoria</strong> tritici leaf blotch is<br />

now the most economically<br />

important foliar disease <strong>of</strong> wheat in<br />

Europe, <strong>and</strong> controlling it with<br />

fungicides costs several hundred<br />

million dollars a year. Resistance to<br />

septoria tritici leaf blotch is<br />

therefore a major target for most<br />

European wheat breeders.<br />

However, if resistant varieties are<br />

to be economically worthwhile,<br />

resistance must be both effective<br />

<strong>and</strong> durable.<br />

Strong, specific interactions<br />

between wheat varieties <strong>and</strong><br />

isolates <strong>of</strong> the pathogen,<br />

Mycosphaerella graminicola, have<br />

been found in both seedlings <strong>and</strong><br />

adult plants (Kema <strong>and</strong> van<br />

Silfhout, 1997), while the resistance<br />

<strong>of</strong> at least one variety, Gene, has<br />

“broken down” through the<br />

evolution <strong>of</strong> a virulent pathogen<br />

population (Mundt et al., 1999).<br />

This means that breeders need to<br />

know not only which varieties may<br />

be good sources <strong>of</strong> resistance in<br />

breeding programs, but also<br />

whether or not resistance genes in<br />

these varieties are at risk from<br />

virulent isolates <strong>of</strong> M. graminicola.<br />

This paper reports a series <strong>of</strong><br />

field trials carried out in Engl<strong>and</strong>,<br />

The Netherl<strong>and</strong>s, <strong>and</strong> Switzerl<strong>and</strong><br />

between 1994 <strong>and</strong> 1997. The aims<br />

were to investigate resistance to<br />

septoria tritici leaf blotch in a<br />

representative set <strong>of</strong> European<br />

varieties, to evaluate potential new<br />

sources <strong>of</strong> resistance, to study the<br />

responses <strong>of</strong> varieties to different<br />

isolates <strong>of</strong> M. graminicola <strong>and</strong> to test<br />

the stability <strong>of</strong> variety-by-isolate<br />

interactions over a range <strong>of</strong><br />

environments.<br />

Materials <strong>and</strong> Methods<br />

A total <strong>of</strong> 71 wheat varieties <strong>of</strong><br />

diverse origins were included in<br />

the field trials. The majority were<br />

cultivars or breeding lines<br />

developed by European breeders.<br />

Most <strong>of</strong> these were winter wheats.<br />

Several varieties that are potential<br />

sources <strong>of</strong> septoria tritici leaf blotch<br />

resistance were also included, as<br />

were a number <strong>of</strong> varieties that are<br />

parents <strong>of</strong> precise genetic stocks<br />

held by the John Innes Centre.<br />

Trials were inoculated with six<br />

monospore isolates <strong>of</strong> M.<br />

graminicola from the IPO-DLO<br />

collection. Six trials were<br />

conducted, three in the<br />

Netherl<strong>and</strong>s, one in Switzerl<strong>and</strong>,<br />

<strong>and</strong> two in Engl<strong>and</strong>. Each trial was<br />

sown in a split-plot design with<br />

two replicate blocks. Within each<br />

block, there were five or six main


plots, each inoculated with one<br />

isolate, such that isolates were<br />

r<strong>and</strong>omized within blocks, while<br />

varieties were r<strong>and</strong>omized within<br />

main plots. The percentage <strong>of</strong> the<br />

leaf area that was necrotic or<br />

covered by lesions bearing<br />

pycnidia was estimated visually.<br />

Results<br />

Levels <strong>of</strong> necrosis <strong>and</strong> pycnidia<br />

were highly correlated, so further<br />

analysis was based on pycnidia<br />

scores. There was highly significant<br />

variation between varieties in the<br />

level <strong>of</strong> disease <strong>and</strong> highly<br />

significant interactions between<br />

varieties <strong>and</strong> isolates. However,<br />

there was no significant variation<br />

in variety-by-isolate interactions<br />

among the six trials. This implies<br />

that variety-by-isolate interactions<br />

were stable over the range <strong>of</strong><br />

environments used in this series <strong>of</strong><br />

trials.<br />

There was a wide range <strong>of</strong><br />

resistance among the varieties<br />

grown in the trials. The existence <strong>of</strong><br />

strong variety-by-isolate<br />

interactions <strong>and</strong> the small number<br />

<strong>of</strong> isolates used means that one<br />

cannot be certain which varieties<br />

have good, overall resistance to the<br />

European population <strong>of</strong> M.<br />

graminicola. However, some lines<br />

were identified that should be<br />

tested further, preferably with more<br />

isolates. The most resistant was the<br />

Brazilian cultivar Veranopolis, a<br />

well-known source <strong>of</strong> resistance.<br />

Four <strong>of</strong> the top ten varieties were<br />

from the FAP breeding program in<br />

Switzerl<strong>and</strong>, including the popular<br />

cultivar Arina. Other varieties<br />

which had very good resistance to<br />

Field Resistance <strong>of</strong> Wheat to <strong>Septoria</strong> Tritici Leaf Blotch, <strong>and</strong> Interactions with Mycosphaerella graminicola Isolates 149<br />

the isolates used in these trials<br />

included cultivars <strong>and</strong> breeding<br />

lines from several European<br />

countries, as well as another<br />

well-known source <strong>of</strong> resistance,<br />

Kavkaz-K4500 l.6.a.4, <strong>and</strong><br />

Frontana, one <strong>of</strong> the parents <strong>of</strong><br />

Veranopolis.<br />

Variety-by-isolate interactions<br />

were most pronounced in tests<br />

with the isolate IPO323. A total <strong>of</strong><br />

27 cultivars <strong>and</strong> breeding lines,<br />

from several European countries,<br />

China, Israel, <strong>and</strong> the USA,<br />

showed specific resistance to this<br />

isolate. One or more varieties<br />

showed specific resistance to four<br />

<strong>of</strong> the other five isolates, while<br />

specific susceptibility to four<br />

isolates was also detected.<br />

Discussion<br />

There is good resistance to<br />

septoria tritici in a wide range <strong>of</strong><br />

European wheat germplasm<br />

from several countries. Plant<br />

breeders may be able to exploit<br />

this resistance by recombining<br />

quantitative resistance from<br />

different sources. There may be<br />

some additional value in<br />

introducing resistance from<br />

sources outside Europe. The<br />

striking success <strong>of</strong> the Swiss<br />

breeding program may be<br />

attributed to a long tradition <strong>of</strong><br />

selection for foliar disease<br />

resistance, using both artificial<br />

<strong>and</strong> natural inoculation.<br />

Furthermore, trials are conducted<br />

without fungicides in several<br />

locations with diverse climatic<br />

conditions.<br />

The existence <strong>of</strong> strong<br />

variety-by-isolate interactions, as<br />

well as the fact that they are<br />

stable over environments, means<br />

that breeders must take account<br />

<strong>of</strong> the possibility that certain<br />

resistances may not be durable.<br />

This is particularly the case with<br />

resistance to IPO323, although it<br />

is not known if this resistance is<br />

controlled by the same genes in<br />

different varieties. One strategy<br />

for breeding might be to use<br />

several specific resistance genes<br />

in a breeding program <strong>and</strong> aim to<br />

select varieties with combinations<br />

<strong>of</strong> resistances. An alternative<br />

would be to avoid selecting for<br />

specific resistances altogether,<br />

aiming instead for a high level <strong>of</strong><br />

quantitative resistance.<br />

Acknowledgments<br />

This research was supported<br />

in part by the EU Framework 4<br />

Biotechnology program, the UK<br />

Ministry <strong>of</strong> Agriculture, Fisheries<br />

<strong>and</strong> Food, <strong>and</strong> Praxis XXI<br />

(Portugal).<br />

References<br />

Kema, G.H.J., <strong>and</strong> C.H. van<br />

Silfhout. 1997. Genetic variation<br />

for virulence <strong>and</strong> resistance in<br />

the wheat-Mycosphaerella<br />

graminicola pathosystem III.<br />

Comparative seedling <strong>and</strong> adult<br />

plant experiments.<br />

Phytopathology 87:266-272.<br />

Mundt, C.C., M.E. H<strong>of</strong>fer, H.U.<br />

Ahmed, S.M. Coakley, J.A.<br />

DiLeone, <strong>and</strong> C. Cowger. 1999.<br />

Population genetics <strong>and</strong> host<br />

resistance. In: <strong>Septoria</strong> on<br />

<strong>Cereals</strong>: a Study <strong>of</strong><br />

Pathosystems. J.A. Lucas, P.<br />

Bowyer, <strong>and</strong> H.A. Anderson<br />

(eds.). CAB Publishing,<br />

Wallingford, UK. pp. 115-130.


150<br />

Using Precise Genetic Stocks to Investigate the Control <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum Resistance in Wheat<br />

C.M. Ellerbrook, 1 V. Korzun, 2 <strong>and</strong> A.J. Worl<strong>and</strong> 1 (Poster)<br />

1 John Innes Centre, Colney, Norwich, UK (Poster)<br />

2 IPK, Gatersleben, Germany<br />

Abstract<br />

A substitution series <strong>of</strong> ‘Synthetic 6x’ into ‘Chinese Spring’ had previously been studied to determine which <strong>of</strong> the 21<br />

chromosomes <strong>of</strong> ‘Synthetic 6x’ conferred resistance to <strong>Stagonospora</strong> nodorum; from this single chromosome recombinant<br />

lines (SCRs) have been developed for substitution lines that improve the resistance <strong>of</strong> ‘Chinese Spring’ to the pathogen. The<br />

SCRs were screened for their response to infection. Marker assisted screening was also employed on the SCR populations;<br />

micro-satellite, RFLP <strong>and</strong> iso-electric focusing, coupled with conventional QTL analysis <strong>and</strong> morphological character scoring<br />

(vernalization response). As these markers had been previously mapped, it would be possible to pinpoint the resistance gene(s)<br />

<strong>and</strong> find a linked marker. Chromosome 5D had been shown to be the most effective against <strong>Stagonospora</strong> nodorum <strong>and</strong> is<br />

also known to carry the isozyme genes, Ibf-1 <strong>and</strong> Mdh-3, a range <strong>of</strong> micro-satellite markers, a sparse selection <strong>of</strong> RFLP<br />

markers, <strong>and</strong> a vernalization response gene (Vrn 3). Screening revealed that the resistance gene (Srb3) was located on the<br />

long arm between Ibf-1 <strong>and</strong> the RFLP probe Xpsr 912, but being 15.6 <strong>and</strong> 12.0 cM, respectively, away from Srb3, they could<br />

not be used as markers for resistance. SCRs have been developed for the remaining chromosomes, <strong>and</strong> screening continues to<br />

precisely map the resistance gene(s).<br />

<strong>Stagonospora</strong> nodorum (<strong>Septoria</strong><br />

nodorum Berk.) is the causal agent<br />

<strong>of</strong> leaf <strong>and</strong> glume blotch in wheat<br />

with infection reducing yield by up<br />

to 40%. Conventional fungicide<br />

control has significant<br />

environmental <strong>and</strong> economic<br />

consequence, thus alternative<br />

methods <strong>of</strong> control are being<br />

sought. Several Aegilops species<br />

including Ae. squarrosa exhibit<br />

partial resistance to the pathogen.<br />

This resistance is expressed in a<br />

synthetic amphiploid ‘Synthetic 6x’<br />

derived from a cross between Ae.<br />

squarrosa <strong>and</strong> Triticum dicoccum<br />

(McFadden <strong>and</strong> Sears, 1946; Sears,<br />

1976). A substitution series <strong>of</strong><br />

‘Synthetic 6x’ chromosomes into<br />

‘Chinese Spring’ was developed<br />

<strong>and</strong> is maintained at the John Innes<br />

Centre. This series was utilized in<br />

disease tests to determine the<br />

chromosomal location <strong>of</strong> the<br />

resistance gene(s) (Nicholson et al.,<br />

1993) as ‘Chinese Spring’ was<br />

shown to be susceptible to<br />

<strong>Stagonospora</strong> nodorum (Scott <strong>and</strong><br />

Benedikz, 1977).<br />

From the D genome Ae.<br />

squarrosa parent, chromosome 5D<br />

conferred a high level <strong>of</strong> resistance,<br />

as did 3D <strong>and</strong> 7D. Resistance was<br />

most pronounced in 5D where the<br />

percentage leaf infection was near<br />

to that <strong>of</strong> the amphiploid. Three<br />

chromosomes from the T. dicoccum<br />

parent (AB genome) 2A, 3B, <strong>and</strong> 5A<br />

conferred resistance but to a lesser<br />

extent than that promoted by the D<br />

genome chromosomes. SCRs were<br />

then developed for these<br />

chromosomes. This paper reports<br />

the results <strong>of</strong> experiments to<br />

localize the resistance genes <strong>and</strong><br />

assign markers linked to the genes.<br />

Materials <strong>and</strong> Methods<br />

Single chromosome<br />

recombinant lines were developed<br />

by crossing the individual<br />

chromosome substitution lines,<br />

previously identified as conferring<br />

resistance, to ‘Chinese Spring’<br />

euploid (or ditelocentric), <strong>and</strong> then<br />

backcrossing the F 1 with the<br />

corresponding monosomic from<br />

the ‘Chinese Spring’ monosomic<br />

series (Sears, 1954). Monosomic<br />

plants were then selected in the<br />

backcrossed F 1 progeny <strong>and</strong> selfed<br />

for disomic extraction (Figure 1).<br />

Infection response was scored<br />

on juvenile plants. Seeds had been<br />

placed on moist filter paper in petri<br />

dishes <strong>and</strong> incubated at 25 o C for 2<br />

days (with 2 days cold shock at 4 o C<br />

to promote germination if<br />

necessary) <strong>and</strong> then planted<br />

individually in pots <strong>of</strong> John Innes<br />

no. 2 compost. In all experiments<br />

the plantlets were grown fully<br />

r<strong>and</strong>omized over 10 replications<br />

with an additional uninfected<br />

replicate as a control.


The seedlings were inoculated<br />

when the second leaves were fully<br />

exp<strong>and</strong>ed with a 1 x 10 6 spore/ml<br />

suspension <strong>of</strong> an aggressive isolate<br />

(S353/88). The plantlets were then<br />

placed in propagators <strong>and</strong><br />

transferred to growth chambers for<br />

72 h at 15 o C in complete darkness<br />

to establish infection. Once this had<br />

been completed, the plants were<br />

moved to a containment glasshouse<br />

<strong>and</strong> removed from their<br />

propagators. Disease was scored at<br />

10 <strong>and</strong> 17 days post-inoculation<br />

<strong>and</strong> expressed as the percentage <strong>of</strong><br />

leaf area lesioned (at the intervals 1,<br />

5, 10, 25, 40, 60, 75, 90, <strong>and</strong> 100%).<br />

The scores were transformed using<br />

the angular transformation <strong>and</strong> an<br />

analysis <strong>of</strong> variance carried out<br />

using the GENSTAT 5 statistical<br />

package. The significance <strong>of</strong><br />

Chinese Spring ditelocentric for the long arm<br />

(2n = 42 tt)<br />

Background<br />

Chinese Spring monosomic (2n = 41)<br />

Selfing <strong>of</strong> each monosomic plant<br />

Using Precise Genetic Stocks to Investigate the Control <strong>of</strong> <strong>Stagonospora</strong> nodorum Resistance in Wheat 151<br />

differences in the mean scores<br />

relative to ‘Chinese Spring’ <strong>and</strong><br />

‘Synthetic 6x’ was estimated using<br />

a two-tailed t-test.<br />

Morphological characters were<br />

scored on field plots, with three<br />

replicates <strong>of</strong> 11 plants in<br />

r<strong>and</strong>omized blocks to enable QTL<br />

analysis <strong>and</strong> screening for Vrn3.<br />

Biochemical analysis involved<br />

the isoelectric focusing technique,<br />

using seed embryo or endosperm<br />

on flat bed electrophoresis<br />

apparatus, showing separation <strong>of</strong><br />

the proteins at their relevant<br />

isoelectric points (pI) (Liu <strong>and</strong><br />

Gale, 1989).<br />

X<br />

X<br />

Cs/Syn 5D (2n = 42)<br />

Selection <strong>of</strong> monosomic plants (2n = 41 ) or monotelosomics (2n = 41 t)<br />

Selection <strong>of</strong> disomic (2n = 42) single chromosome recombinant lines (SCRs)<br />

Figure 1. Development <strong>of</strong> single chromosome recombinant lines (SCRs) for the long arm <strong>of</strong> chromosome<br />

5D (18).<br />

F1<br />

Molecular marker techniques<br />

required extraction <strong>of</strong> DNA (phenol<br />

chlor<strong>of</strong>orm method) <strong>and</strong> in the case<br />

<strong>of</strong> RFLPs digesting with enzymes<br />

that have a 4 base-pair recognition<br />

sequence (Bam HI, Eco RI) <strong>and</strong><br />

then hybridizing with P 32 labeled<br />

CTP (Southern, 1975). The microsatellite<br />

technique was PCR-based<br />

<strong>and</strong> used micro-satellite probes<br />

developed at IPK-Gatersleben. Both<br />

techniques clearly revealed the SCR<br />

progeny as either ‘Chinese Spring’<br />

or ‘Synthetic 6x’ type.<br />

Marker results were input into<br />

Mapmaker (Lincoln et al., 1992) to<br />

compare the segregation <strong>of</strong><br />

markers in the progeny, <strong>and</strong> a map<br />

<strong>of</strong> resistance gene(s) location<br />

5D


Session 6C — C.M. Ellerbrook, V. Korzun, <strong>and</strong> A.J. Worl<strong>and</strong><br />

152<br />

related to the markers analyzed<br />

was compiled. QTLs were also<br />

assigned to markers on the<br />

chromosome using the Students ttest.<br />

To date only the 5D SCR<br />

population has been analyzed for<br />

disease resistance <strong>and</strong> associated<br />

markers.<br />

6.8 cM<br />

8.2 cM<br />

2.1 cM<br />

3.2 cM<br />

12.0 cM<br />

12.0 cM<br />

15.6 cM<br />

7.8 cM<br />

7.4 cM<br />

16.1 cM<br />

7.4 cM<br />

10.3 cM<br />

Results <strong>and</strong> Discussion<br />

Following analysis <strong>of</strong> the 5D<br />

population (85 lines), it was found<br />

that S. nodorum resistance was<br />

under the control <strong>of</strong> a single gene<br />

Srb3, which was located between<br />

Ibf-1 <strong>and</strong> Xpsr 912, with a QTL for<br />

plant height (Qph) being linked to<br />

Ibf-1 (Figure 2). However, these<br />

genes are too remote from Srb3 to<br />

Mdh 3<br />

Wms 205<br />

Dms 3<br />

Xpsr 628<br />

Xpsr 906<br />

Xpsr 574<br />

Xpsr 912<br />

Srb 3<br />

Ibf-1<br />

Wms 182<br />

Wms 174<br />

Wms 292<br />

Xpsr 819<br />

Vrn 3<br />

Qsy (QTL for plant yield)<br />

Qey (QTL for ear yield)<br />

Qph (QTL for plant height)<br />

Qph (QTL for plant height)<br />

be useful as flanking markers.<br />

Work is continuing to locate more<br />

markers on chromosome 5D, as the<br />

region in which Srb3 is located<br />

currently has a low marker<br />

concentration. RFLP screening is<br />

also underway with recently<br />

developed markers <strong>and</strong> with the<br />

collaboration at IPK-Germany; any<br />

new 5D micro-satellites produced<br />

will also be screened. It is hoped<br />

Qsn (QTL for spikelet number)<br />

Qph (QTL for plant height)<br />

Figure 2. Relative position <strong>of</strong> the Stagnospora nodorum resistance gene Srb3 <strong>and</strong> previously mapped markers.


that this continuing work will help<br />

to complete the map more<br />

successfully.<br />

Mapping, field trials, <strong>and</strong><br />

disease screening are now in<br />

progress on the SCR populations<br />

for chromosomes 2A, 3D, <strong>and</strong> 7D.<br />

Our results have shown that S.<br />

nodorum resistance in wheat is<br />

inherited in a complex manner<br />

involving several genes. By<br />

studying them in single<br />

chromosome recombinant lines, we<br />

should be more able to fully<br />

underst<strong>and</strong> their heritability as<br />

well as highlighting linked<br />

markers.<br />

Using Precise Genetic Stocks to Investigate the Control <strong>of</strong> <strong>Stagonospora</strong> nodorum Resistance in Wheat 153<br />

References<br />

Lincoln, S.E., Daly, M., <strong>and</strong> L<strong>and</strong>er,<br />

E.S. 1992. Constructing genetic<br />

maps with MAPMAKER/EXP 3.0.<br />

Whitehead Institute Technical<br />

Report (Second edition).<br />

Lui, C.J., <strong>and</strong> Gale, M.D. 1989. Ibf-1, a<br />

highly variable marker system in<br />

Triticeae. TAG 77:233-240.<br />

McFadden, E.S., <strong>and</strong> Sears, E.R. 1946.<br />

The origin <strong>of</strong> Triticum spelta <strong>and</strong> its<br />

free-threshing hexaploid relatives.<br />

J. Heredity 37:81-89 <strong>and</strong> 107-116.<br />

Nicholson, P., Rezanoor, H.N., <strong>and</strong><br />

Worl<strong>and</strong>, A.J. 1993. Chromosomal<br />

location <strong>of</strong> resistance to <strong>Septoria</strong><br />

nodorum in a synthetic hexaploid<br />

wheat determined by the study <strong>of</strong><br />

chromosomal substitution lines in<br />

‘Chinese Spring’ wheat. Plant<br />

Breeding 110:177-184.<br />

Scott, P.R., <strong>and</strong> Benedikz, P.W. 1977.<br />

<strong>Septoria</strong> Plant Breeding Institute<br />

Annual Report 128-129.<br />

Sears, E.R. 1954. The aneuploids <strong>of</strong><br />

common wheat. Missouri<br />

Agricultural Experimental Station<br />

Research Bulletin 572, pp. 1-58.<br />

Sears, E.R. 1976. A synthetic<br />

hexaploid wheat with fragile<br />

rachis. Wheat Information Service<br />

41-42, 31-32.<br />

Southern, E.M. 1975. Detection <strong>of</strong><br />

specific sequences among DNA<br />

fragments separated by gel<br />

electrophoresis. J. Mol. Biol.<br />

98:503-517.


154<br />

Evaluating Triticum durum x Triticum tauschii Germplasm<br />

for Resistance to <strong>Stagonospora</strong> nodorum<br />

L.R. Nelson 1 <strong>and</strong> M.E. Sorrells 2 (Poster)<br />

1 Agricultural Research & Extension Center, Texas A&M University, Overton, TX, USA<br />

2 Dept. <strong>of</strong> Plant Breeding, Cornell University, Ithaca, NY, USA<br />

Abstract<br />

This study was conducted to determine resistance <strong>of</strong> 50 <strong>CIMMYT</strong> Triticum durum x Triticum tauschii populations to<br />

septoria glume blotch. Seedlings were inoculated at the 2-leaf stage, <strong>and</strong> data were recorded on disease severity <strong>and</strong> latent<br />

period. Several lines from this population demonstrated good resistance compared to check entries <strong>and</strong> are being used as<br />

sources <strong>of</strong> resistance to septoria glume blotch in the wheat breeding program.<br />

<strong>Septoria</strong> glume blotch caused<br />

by <strong>Stagonospora</strong> nodorum is a<br />

significant disease <strong>of</strong> wheat in the<br />

southern <strong>and</strong> eastern USA.<br />

Immunity <strong>of</strong> wheat to S. nodorum<br />

has not been found (Dantuma,<br />

1955), but variation in resistance is<br />

apparent. New sources <strong>of</strong> genetic<br />

resistance are needed in order to<br />

reduce yield losses cause by this<br />

fungal disease. In our research we<br />

have studied components <strong>of</strong> partial<br />

resistance (Nelson <strong>and</strong> Marshall,<br />

1990). A relatively simple test to<br />

evaluate resistance is based on<br />

measuring incubation period <strong>and</strong><br />

disease severity at a different<br />

length(s) <strong>of</strong> time after inoculation.<br />

In this study, we utilized these two<br />

components to evaluate resistance<br />

<strong>of</strong> wheat germplasm to septoria<br />

glume blotch. The objective <strong>of</strong> this<br />

study was to screen a collection <strong>of</strong><br />

50 exotic <strong>CIMMYT</strong> germplasm<br />

accessions for resistance to septoria<br />

glume blotch, <strong>and</strong> compare any<br />

resistance found with that <strong>of</strong><br />

moderately resistant checks to<br />

determine the potential <strong>of</strong> exotic<br />

germplasm.<br />

Materials <strong>and</strong> Methods<br />

Seed <strong>of</strong> 50 <strong>CIMMYT</strong><br />

populations <strong>and</strong> a susceptible<br />

(TAM 107) <strong>and</strong> a resistant (TX76-<br />

40-2) check (Nelson et al., 1994)<br />

were planted in rows in soil in flats.<br />

The <strong>CIMMYT</strong> germplasm consisted<br />

<strong>of</strong> 50 amphiploids from crosses<br />

between Triticum durum <strong>and</strong><br />

Triticum tauschii obtained from Dr.<br />

A. Mujeeb Kazi, <strong>CIMMYT</strong>, via Dr.<br />

Mark Sorrells at Cornell University.<br />

The wheat was germinated <strong>and</strong><br />

grown in the greenhouse until the<br />

2-leaf stage. At that time spores <strong>of</strong><br />

S. nodorum were sprayed on the<br />

wheat, which was then placed in a<br />

humidity chamber. The plants were<br />

removed from the humidity<br />

chamber after 50 hours. During this<br />

50 hour period, a cool mist blower<br />

operated for 15 minutes every 2<br />

hours during the day, <strong>and</strong> for 15<br />

minutes every 4 hours during the<br />

night, which kept the leaves moist,<br />

but did not cause run-<strong>of</strong>f. After the<br />

50-hour period, the flats were<br />

removed from the chamber <strong>and</strong><br />

allowed to continue growing in the<br />

greenhouse. Five days after being<br />

removed from the humidity<br />

chamber, some plants began to<br />

show leaf necrosis. Disease ratings<br />

were recorded for each row, or<br />

entry, on a 0 to 9 disease rating (0 =<br />

no symptoms; 9 = severe necrosis)<br />

on day 5, 8, <strong>and</strong> 11, to obtain an<br />

estimate <strong>of</strong> incubation period <strong>and</strong><br />

disease severity. A mean rating<br />

over the above three days was also<br />

calculated.<br />

Results<br />

Data are presented in Table 1<br />

for each day’s rating <strong>and</strong> the mean<br />

disease rating. The susceptible <strong>and</strong><br />

resistant checks provided a<br />

benchmark or st<strong>and</strong>ard upon<br />

which to compare these lines. Any<br />

lines that have disease ratings<br />

similar or less than the resistant<br />

check (TX76-40-2) may be useful in<br />

a wheat-breeding program <strong>and</strong><br />

could provide new resistance. Any<br />

lines similar to the susceptible<br />

check TAM 107 will not be useful<br />

for increasing resistance to septoria<br />

glume blotch. TAM 107 was<br />

already highly diseased by day 5,<br />

indicating a very short incubation<br />

period. Some lines had low disease<br />

ratings on day 5, but by day 11, had<br />

high disease ratings, similar to<br />

TAM 107. This indicates a longer<br />

incubation period, <strong>and</strong> this<br />

resistance may be useful in a<br />

breeding program to delay the<br />

buildup <strong>of</strong> the disease.


Table 1. Disease rating <strong>of</strong> <strong>CIMMYT</strong> durum x T.<br />

tauschii wheat lines for septoria glume blotch on<br />

a 0 to 9 scale where 0 = no disease <strong>and</strong> 9 = very<br />

high disease rating.<br />

Rating Rating Rating Mean<br />

Entry number day 5 day 8 day 11 rating<br />

TAM 107 Susc. Ck 7 7 7 7<br />

Syn-1 Altar 21 3 7 7 6<br />

Syn-2 Altar 219 1 4 6 4<br />

Syn-3 Altar 219 4 6 6 5<br />

Syn-4 Altar 219 7 6 6 6<br />

Syn-5 Altar 224 7 7 7 7<br />

Tx76-40-2 Res. Ck 0 1 1 1<br />

Syn-6 Altar 224 4 3 4 4<br />

Syn-7 Altar 224 3 3 6 4<br />

Syn-8 Altar 221 3 4 6 4<br />

Syn-9 Altar 223 1 2 3 2<br />

Syn-10 Altar 223 2 3 4 3<br />

Syn-11 Altar 192 1 3 4 3<br />

Syn-12 Altar 198 1 3 2 2<br />

Syn-13 Altar 198 0 0 1 1<br />

Syn-14 Altar 198 1 2 2 2<br />

Syn-15 Altar 211 1 3 2 2<br />

Syn-16 Chen 205 7 6 7 7<br />

Syn-17 Chen 205 8 8 8 8<br />

Syn-18 Chen 205 6 7 7 7<br />

Syn-19 Chen 205 5 5 5 5<br />

Syn-20 Chen 205 3 3 5 4<br />

Syn-21 Chen 205 2 3 5 3<br />

Syn-22 Chen 205 5 5 7 6<br />

Syn-23 Chen 205 4 4 6 5<br />

Syn-24 Chen 205 3 3 3 3<br />

Syn-25 Chen 205 5 5 5 5<br />

Syn-26 Chen 215 3 3 3 3<br />

Syn-27 Chen 215 0 2 2 1<br />

Syn-28 Chen 224 7 6 6 6<br />

Syn-29 Chen 224 7 7 4 6<br />

Syn-30 Chen 224 7 7 4 6<br />

Syn-31 Chen 224 4 5 5 5<br />

Syn-32 Chen 224 3 3 4 3<br />

Syn-33 Chen 224 5 5 5 5<br />

Syn-34 Chen 224 3 5 5 4<br />

Syn-35 Chen 224 6 6 6 6<br />

Syn-36 Chen 224 6 6 4 5<br />

Syn-37 Chen 224 7 7 7 7<br />

Syn-38 Chen 224 5 6 6 6<br />

Syn-39 Cndo/R143//<br />

Ent “S”/Mex 214 7 6 6 6<br />

Syn-40 “ “ 5 6 5 5<br />

Syn 41 “ “ 3 5 5 4<br />

Syn-43 Cndo/R143//<br />

Ent”S”/Mex 221 5 5 5 5<br />

Syn-44 Cndo/R143//<br />

Ent”S”/Mex 221 5 4 4 4<br />

Syn-45 “ “ 4 7 7 6<br />

Syn-45 “ “ 3 4 5 4<br />

Syn-46 “ “ 6 7 6 6<br />

Syn-45 Laru’s 309 3 5 5 4<br />

Syn-48 5 6 6 6<br />

Syn-49 “ “ 8 7 8 8<br />

Syn-50 “ “ 2 3 3 3<br />

Evaluating Triticum durum x Triticum tauschii Germplasm for Resistance to <strong>Stagonospora</strong> nodorum 155<br />

Highly resistant lines were<br />

Syn 9, Syn 12, Syn 13, Syn 14,<br />

Syn 15, <strong>and</strong> Syn 27. Their<br />

resistance was similar to TX76-<br />

40-2. Moderately resistant lines<br />

were Syn 10, Syn 11, Syn 32, <strong>and</strong><br />

Syn 50. Several other lines also<br />

demonstrated a moderate<br />

degree <strong>of</strong> partial resistance.<br />

Lines Syn 1 through 15 may<br />

have a similar genetic<br />

background for the A <strong>and</strong> B<br />

genomes because they all have<br />

Altar as the durum parent. Of<br />

the resistant lines, the T. tauschii<br />

parents were 223 for Syn 9 <strong>and</strong><br />

10, 192 for Syn 11, 198 for Syn<br />

12-14 <strong>and</strong> 211 for Syn 15. ‘Chen’<br />

<strong>and</strong> ‘Laru’ “S” were the durum<br />

parents for Syn 27 <strong>and</strong> 50 <strong>and</strong><br />

the T. tauschii lines were 215 <strong>and</strong><br />

309, respectively. It appears that<br />

at least some <strong>of</strong> the resistance in<br />

this germplasm originated in the<br />

T. tauschii lines used to make the<br />

amphiploids.<br />

Discussion<br />

It appears that we have<br />

several promising new sources<br />

<strong>of</strong> resistance in a wheat<br />

background that can be crossed<br />

with our adapted winter wheat.<br />

We are presently increasing seed<br />

<strong>of</strong> these 50 lines in our<br />

greenhouse. We have crossed<br />

several <strong>of</strong> the best lines for<br />

resistance to septoria glume blotch<br />

with both hard <strong>and</strong> s<strong>of</strong>t red winter<br />

wheat in our crossing program<br />

during 1998/99. Since this<br />

germplasm contains sources <strong>of</strong><br />

resistance to septoria glume blotch<br />

which are likely not present in our<br />

s<strong>of</strong>t red winter wheat, there may be<br />

an opportunity to pyramid genes<br />

<strong>and</strong> make significant improvement<br />

in resistance to this disease.<br />

Acknowledgment<br />

Appreciation is expressed to<br />

<strong>CIMMYT</strong> for developing <strong>and</strong><br />

sharing the germplasm.<br />

References<br />

Dantuma, G. 1955. The heavy attack<br />

<strong>of</strong> diseases during the ripening <strong>of</strong><br />

wheat in 1954. Euphytica 5:94-95.<br />

Nelson, L.R., R.D. Barnett, D.<br />

Marshall, C.A. Erickson, M.E.<br />

McDaniel, W.D. Worrall, N.A.<br />

Tuleen, <strong>and</strong> M.D. Lazar. 1994.<br />

Registration <strong>of</strong> TX76-40-2 wheat<br />

germplasm. Crop Sci. 34:1137.<br />

Nelson, L.R., <strong>and</strong> D. Marshall. 1990.<br />

Breeding wheat for resistance to<br />

<strong>Septoria</strong> nodorum <strong>and</strong> <strong>Septoria</strong><br />

tritici. Advances in Agronomy<br />

44:257-277.


156<br />

Sources <strong>of</strong> Resistance to <strong>Septoria</strong> passerinii in Hordeum<br />

vulgare <strong>and</strong> H. vulgare subsp. spontaneum<br />

H. Toubia-Rahme <strong>and</strong> B.J. Steffenson (Poster)<br />

North Dakota State University, Department <strong>of</strong> Plant Pathology, Fargo, ND, USA<br />

Abstract<br />

<strong>Septoria</strong> speckled leaf blotch (SSLB), incited by <strong>Septoria</strong> passerinii <strong>and</strong> <strong>Stagonospora</strong> avenae f. sp. triticea, has<br />

become one <strong>of</strong> the most serious diseases <strong>of</strong> barley in the Upper Midwest region <strong>of</strong> the USA. In barley SSLB can cause<br />

significant losses in both the yield <strong>and</strong> quality. The major malting <strong>and</strong> feed barley cultivars in the Upper Midwest are very<br />

susceptible to SSLB. Resistance breeding is the most effective strategy for controlling this disease. A diverse group <strong>of</strong> barley<br />

germplasm including Hordeum vulgare subsp. spontaneum accessions, advanced midwestern breeding lines, <strong>and</strong><br />

commercial cultivars was evaluated at the seedling stage in the greenhouse for reaction to S. passerinii. Of 200 accessions<br />

tested, 79 were found resistant. Most <strong>of</strong> H. vulgare subsp. spontaneum accessions (17 <strong>of</strong> 24) were resistant to S. passerinii.<br />

These accessions all originated from the Middle East, except one that was from Tibet. Most <strong>of</strong> the advanced midwestern<br />

breeding lines found resistant to this pathogen have Gloria”S”/Copal”S” (an ICARDA/<strong>CIMMYT</strong> barley line) in their<br />

pedigree, which is believed to be a source <strong>of</strong> resistance to S. passerinii. From this study, it is evident that many barley<br />

accessions possess resistance to S. passerinii. Additional evaluations will be made on this germplasm to S. avenae f. sp.<br />

triticea to identify accessions that possess effective resistance to both SSLB pathogens.<br />

<strong>Septoria</strong> speckled leaf blotch<br />

(SSLB), caused by <strong>Septoria</strong> passerinii<br />

<strong>and</strong> <strong>Stagonospora</strong> avenae f. sp.<br />

triticea, has become an increasingly<br />

important disease <strong>of</strong> barley<br />

(Hordeum vulgare L.) over the past<br />

decade in the Upper Midwest<br />

region <strong>of</strong> the USA. In surveys <strong>of</strong><br />

North Dakota barley fields in 1998,<br />

S. passerinii <strong>and</strong> S. avenae f. sp.<br />

triticea were recovered from 45%<br />

<strong>and</strong> 37% <strong>of</strong> leaves exhibiting leaf<br />

spot symptoms, respectively<br />

(Krupinsky <strong>and</strong> Steffenson, this<br />

volume). Yield losses <strong>of</strong> 20% have<br />

been recorded in barley due to S.<br />

passerinii infection (Green <strong>and</strong><br />

Bendelow, 1961). In recent trials<br />

conducted in North Dakota, yield<br />

losses <strong>of</strong> 23% to 38% were<br />

observed in Robust barley infected<br />

with SSLB (J. Lukach <strong>and</strong> B.<br />

Steffenson, unpublished data). In<br />

addition to reducing yield, SSLB<br />

also reduces kernel plumpness <strong>and</strong><br />

malt extract, which are important<br />

malt quality characters (Green <strong>and</strong><br />

Bendelow, 1961).<br />

Host plant resistance provides<br />

the most practical <strong>and</strong><br />

environmentally safe method <strong>of</strong><br />

disease control. Unfortunately, all<br />

major malting <strong>and</strong> feed barley<br />

cultivars in the Upper Midwest<br />

region are susceptible to SSLB. The<br />

objectives <strong>of</strong> this study were to 1)<br />

identify sources <strong>of</strong> resistance to S.<br />

passerinii in commercial barley<br />

cultivars, agronomically advanced<br />

midwestern breeding lines, <strong>and</strong><br />

Hordeum vulgare subsp. spontaneum<br />

accessions; <strong>and</strong> 2) evaluate<br />

previously reported sources <strong>of</strong> S.<br />

passerinii resistance to a<br />

midwestern isolate <strong>of</strong> this<br />

pathogen.<br />

Materials <strong>and</strong> Methods<br />

In total, 200 barley entries<br />

including 24 H. vulgare subsp.<br />

spontaneum accessions were<br />

evaluated at the seedling stage in<br />

the greenhouse. Five seeds <strong>of</strong> each<br />

entry were planted in pots (10 x 10<br />

cm) filled with a potting mix<br />

consisting <strong>of</strong> peat moss (75%) <strong>and</strong><br />

perlite (25%). Slow-release (14-14-<br />

14, N-P-K, 2 g/pot) <strong>and</strong> watersoluble<br />

(15-0-15, N-P-K, 2 g/pot)<br />

fertilizers were added at the time <strong>of</strong><br />

planting. All seedlings were grown<br />

in the greenhouse at 20±3ºC with a<br />

14-h photoperiod. The fungal<br />

isolate (ND97-15) used in this<br />

study was obtained from naturally<br />

infected barley leaves collected<br />

from a commercial field in<br />

Bottineau county, North Dakota, in<br />

1997. For inoculum production, the<br />

isolate was grown on yeast malt<br />

agar (YMA) (Eyal et al., 1987) in<br />

plastic Petri dishes at 21ºC with a<br />

12-h photoperiod (cool-white<br />

fluorescent tubes). When pycnidia<br />

developed <strong>and</strong> sporulated, mass<br />

spore transfers were made by<br />

removing with a sterile needle<br />

cirrhi from pycnidia <strong>and</strong><br />

transferring them to other YMA<br />

plates. After 4-5 days <strong>of</strong> incubation<br />

under the same conditions,<br />

pycnidia were harvested by<br />

flooding the surface <strong>of</strong> the plates


with double-distilled water <strong>and</strong><br />

scraping the agar surface with a<br />

rubber spatula. This suspension <strong>of</strong><br />

pycnidia was blended for 30 s in a<br />

blender to release pycnidiospores<br />

<strong>and</strong> then filtered through four<br />

layers <strong>of</strong> cheesecloth. Tween-20<br />

(polyoxyethylene-20-sorbitan<br />

monolaurate) was added to the<br />

pycnidiospore suspension at a rate<br />

<strong>of</strong> 100 µl/l to facilitate the uniform<br />

distribution <strong>and</strong> adsorption <strong>of</strong><br />

inoculum onto the leaf surfaces.<br />

Barley seedlings were<br />

inoculated with the pycnidiospore<br />

suspension (5 x 10 5<br />

pycnidiospores/ml) at the two-leaf<br />

stage (10-12 days old). Inoculated<br />

seedlings were incubated at 21ºC/<br />

dark <strong>and</strong> 25ºC/light for 72 h in<br />

mist chambers, where the relative<br />

humidity was maintained near<br />

100%. The first 40 h <strong>of</strong> incubation<br />

was in darkness, followed by a<br />

photoperiod <strong>of</strong> 5 h for next two<br />

days. Plants were allowed to dry<br />

<strong>of</strong>f slowly before being transferred<br />

to the greenhouse under the same<br />

conditions previously described.<br />

The reaction <strong>of</strong> the entries to S.<br />

passerinii was assessed on the<br />

second leaves <strong>of</strong> seedlings 17 days<br />

after inoculation using a 0-5 rating<br />

scale where 0, 1, <strong>and</strong> 2 are<br />

indicative <strong>of</strong> resistance <strong>and</strong> 3, 4,<br />

<strong>and</strong> 5 <strong>of</strong> susceptibility. Resistant<br />

(cv. Atlas) <strong>and</strong> susceptible (cv.<br />

Betzes) checks were included in the<br />

experiment.<br />

Results <strong>and</strong> Discussion<br />

Hordeum vulgare <strong>and</strong> H. vulgare<br />

subsp. spontaneum accessions were<br />

classified as susceptible or resistant<br />

based on their reaction to S.<br />

passerinii at the seedling stage.<br />

Sources <strong>of</strong> Resistance to <strong>Septoria</strong> passerinii in Hordeum vulgare <strong>and</strong> H. vulgare subsp. spontaneum 157<br />

Marked differences were observed<br />

in the reaction <strong>of</strong> barley accessions<br />

to S. passerinii infection. In total, 79<br />

lines were found resistant to S.<br />

passerinii. Of the 24 H. vulgare<br />

subsp. spontaneum accessions<br />

tested, 17 were resistant. These<br />

accessions all originated from the<br />

Middle East, except one, which was<br />

from Tibet. Similar results were<br />

obtained by Metcalfe et al. (1977)<br />

who found that all H. vulgare<br />

subsp. spontaneum accessions<br />

collected in the Middle East were<br />

resistant to S. passerinii.<br />

All <strong>of</strong> the major 6-rowed<br />

malting (Foster, St<strong>and</strong>er, <strong>and</strong><br />

Robust), <strong>and</strong> 2-rowed feed<br />

(Bowman, Conlon, <strong>and</strong> Logan)<br />

cultivars grown in the Upper<br />

Midwest region were susceptible.<br />

Of 120 advanced midwestern 6<strong>and</strong><br />

2-rowed breeding lines, 41<br />

were resistant. Most <strong>of</strong> these<br />

breeding lines have Gloria”S”/<br />

Copal”S” (an ICARDA/<strong>CIMMYT</strong><br />

barley line) in their pedigree. This<br />

line exhibited S. passerinii resistance<br />

under field conditions in North<br />

Dakota <strong>and</strong> is presumed to be the<br />

source <strong>of</strong> resistance to S. passerinii<br />

in these breeding lines (J.<br />

Franckowiak, personal<br />

communication).<br />

Nine accessions (AC Hamilton,<br />

Atlas, Bolron, CIho 4439, CIho<br />

4780, CIho 10644, Feebar, Nomini,<br />

<strong>and</strong> Starling) previously reported<br />

to have resistance to S. passerinii<br />

were also resistant to the North<br />

Dakota isolate (ND97-15) used in<br />

this study. Other barley accessions<br />

resistant to this isolate were: Atlas<br />

54, Baronesse, Belford, CIho 4428,<br />

CIho 4940, Flynn 1, Glacier, <strong>and</strong><br />

Vaughn. Only a few studies have<br />

been advanced on the genetics <strong>of</strong><br />

resistance in barley to S. passerinii.<br />

Peterson (1956) indicated that the<br />

variety Atlas possesses dominant<br />

resistance genes to S. passerinii.<br />

Buchannon (1961) found two<br />

recessive genes conferring<br />

resistance to S. passerinii in the<br />

cultivar Feebar, <strong>and</strong> Rasmusson<br />

<strong>and</strong> Rogers (1963) reported two<br />

different dominant resistant genes,<br />

Sep2 <strong>and</strong> Sep3, in the accessions<br />

CIho 4780 <strong>and</strong> CIho 10644,<br />

respectively. Metcalfe et al. (1970)<br />

found that a single dominant gene<br />

governs resistance to S. passerinii in<br />

CIho 4439. It is evident that many<br />

barley accessions possess resistance<br />

to this pathogen; however, none<br />

has been exploited in midwest<br />

barley breeding programs, as all <strong>of</strong><br />

the major malting <strong>and</strong> feed<br />

cultivars are susceptible to S.<br />

passerinii.<br />

Segregating populations<br />

derived from some <strong>of</strong> the described<br />

sources <strong>of</strong> resistance <strong>and</strong><br />

susceptible commercial cultivars<br />

are under evaluation in our<br />

laboratory to study the genetics <strong>of</strong><br />

resistance in barley to S. passerinii<br />

<strong>and</strong> to identify molecular markers<br />

linked to S. passerinii resistance<br />

gene(s). All accessions found<br />

resistant to S. passerinii in this study<br />

will also be evaluated to S. avenae f.<br />

sp. triticea. This test will enable us<br />

to identify sources <strong>of</strong> resistance to<br />

both S. passerinii <strong>and</strong> S. avenae f. sp.<br />

triticea <strong>and</strong> to determine whether<br />

resistance to both pathogens is<br />

governed by the same or by<br />

different gene(s). The development<br />

<strong>of</strong> barley cultivars with resistance<br />

to both S. passerinii <strong>and</strong> S. avenae f.<br />

sp. triticea is necessary, as both<br />

pathogens are common in the<br />

Upper Midwest production area.


Session 6C — H. Toubia-Rahme <strong>and</strong> B.J. Steffenson<br />

158<br />

References<br />

Buchannon, K.W. 1961. Inheritance <strong>of</strong><br />

reaction to <strong>Septoria</strong> passerinii Sacc.<br />

<strong>and</strong> Pyrenophora teres (Died.)<br />

Drechsl., <strong>and</strong> <strong>of</strong> row number, in<br />

barley. Ph.D. diss. University <strong>of</strong><br />

Saskatchewan, Saskatoon. 40 pp.<br />

Eyal, Z., Scharen, A.L., Prescott, J.M.,<br />

<strong>and</strong> M. van Ginkel. 1987. The<br />

<strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Concepts <strong>and</strong> Methods <strong>of</strong> Disease<br />

Management. Mexico, D.F.:<br />

<strong>CIMMYT</strong>. 46 pp.<br />

Green, G.J., <strong>and</strong> V.M. Bendelow. 1961.<br />

Effect <strong>of</strong> speckled leaf blotch,<br />

<strong>Septoria</strong> passerinii Sacc., on the<br />

yield <strong>and</strong> malting quality <strong>of</strong> barley.<br />

Can. J. Plant Sci 41:431-435.<br />

Metcalfe, D.R., Buchannon, K.W.,<br />

McDonald, W.C., <strong>and</strong> E. Reinbergs.<br />

1970. Relationships between the<br />

‘Jet’ <strong>and</strong> ‘Milton’ genes for<br />

resistance to loose smut <strong>and</strong> genes<br />

for resistance to other barley<br />

diseases. Can. J. Plant Sci 50:423-<br />

427.<br />

Metcalfe, D.R., Chiko, A.W., Martens,<br />

J.W., <strong>and</strong> A. Tekauz. 1977. Reaction<br />

<strong>of</strong> barleys from the Middle East to<br />

Canadian pathogens. Can. J. Plant<br />

Sci 57: 995-999.<br />

Peterson, R.F. 1956. Progress report,<br />

Cereal Breeding Laboratory,<br />

Winnipeg, Manitoba, 1949-54. 36<br />

pp.<br />

Rasmusson, D.C., <strong>and</strong> W.E. Rogers.<br />

1963. Inheritance <strong>of</strong> resistance to<br />

<strong>Septoria</strong> in barley. Crop Sci 3:161-<br />

162.


Partial Resistance to <strong>Stagonospora</strong> nodorum in Wheat 159<br />

S<strong>of</strong>t Red Winter Wheat with Resistance to <strong>Stagonospora</strong><br />

nodorum <strong>and</strong> Other Foliar Pathogens<br />

B.M. Cunfer1 <strong>and</strong> J.W. Johnson2 (Poster)<br />

1 Department <strong>of</strong> Plant Pathology, <strong>and</strong> 2 Department <strong>of</strong> Crops <strong>and</strong> Soils, Griffin Campus, University <strong>of</strong> Georgia,<br />

Griffin, GA, USA<br />

S<strong>of</strong>t red winter wheat<br />

germplasm adapted to the<br />

southeastern USA that has a high<br />

level <strong>of</strong> resistance to <strong>Stagonospora</strong><br />

nodorum has been difficult to<br />

identify. Some partially resistant<br />

cultivars with long-lasting<br />

resistance have been developed,<br />

but these are replaced frequently<br />

due to loss <strong>of</strong> resistance to<br />

powdery mildew (Blumeria<br />

graminis), leaf rust (Puccinia<br />

recondita), <strong>and</strong> Hessian fly<br />

(Mayetiola destructor). Four lines,<br />

GA 84202, GA 85240, GA 85410AB,<br />

<strong>and</strong> GA 861460, with good<br />

agronomic traits have been selected<br />

<strong>and</strong> included among elite lines in<br />

the Georgia wheat breeding<br />

program in the past eight years.<br />

These lines have resistance to S.<br />

nodorum which is equal or better<br />

than that <strong>of</strong> older cultivars such as<br />

Oasis. They are also resistant to<br />

current populations <strong>of</strong> leaf rust<br />

present in the southeastern USA,<br />

most races <strong>of</strong> powdery mildew, <strong>and</strong><br />

Hessian fly.<br />

The lines selected were<br />

evaluated in the greenhouse <strong>and</strong><br />

field over a six-year period against<br />

100 or more advanced <strong>and</strong> elite<br />

lines <strong>and</strong> st<strong>and</strong>ard check cultivars<br />

each year. Seedling plants were<br />

inoculated in the greenhouse each<br />

year. Adult plants were inoculated<br />

in the field with S. nodorum <strong>and</strong><br />

exposed to natural infection. Data<br />

were collected in field trials at<br />

Griffin <strong>and</strong> Plains, GA, under<br />

moderate to severe disease<br />

pressure from powdery mildew,<br />

leaf rust, <strong>and</strong> leaf <strong>and</strong> glume<br />

blotch. Results from the<br />

greenhouse <strong>and</strong> field were<br />

generally in good agreement.<br />

Seeds <strong>of</strong> each line have been<br />

deposited in the USDA Small<br />

Grains Collection, Aberdeen, ID.<br />

These lines may be useful in other<br />

regions where they are adapted.<br />

For example, in addition to being<br />

agronomically adapted, 85410AB<br />

<strong>and</strong> 84202 were found to be<br />

resistant to stripe rust (P. striiformis)<br />

<strong>and</strong> leaf rust in a field trial at<br />

Colonia, Uruguay, in 1995. These<br />

lines are facultative wheats<br />

adapted to winter wheat culture in<br />

regions with a mild to moderate<br />

winter climate. They have been<br />

used as parents in several<br />

advanced <strong>and</strong> elite breeding lines<br />

currently being evaluated in the<br />

Georgia breeding program.


160<br />

Partial Resistance to <strong>Stagonospora</strong> nodorum in Wheat<br />

C.G. Du, 1 L.R. Nelson, 2 <strong>and</strong> M.E. McDaniel 3 (Poster)<br />

1 Dept. <strong>of</strong> Computer Science, Texas A&M Univ., College Station, TX. USA<br />

2 Texas A&M Univ. Agri. Res. & Ext. Center, Overton, TX, USA<br />

3 Soil <strong>and</strong> Crop Sciences Dept., Texas A&M Univ., College Station, TX, USA<br />

Abstract<br />

Seedling plants <strong>of</strong> parents, F 1 crosses, <strong>and</strong> F 2 populations were inoculated at the two-leaf stage with spores <strong>of</strong><br />

<strong>Stagonospora</strong> nodorum in a humidity chamber to determine incubation period (IP), latent period (LP) <strong>and</strong> necrosis<br />

percentage (NP). Incubation period, LP, <strong>and</strong> NP exhibited polygenic inheritance <strong>and</strong> were controlled by 2-3, 3, <strong>and</strong> 1-4 genes,<br />

respectively. Each <strong>of</strong> the components <strong>of</strong> partial resistance showed moderate to high heritabilities.<br />

Breeding wheat for resistance to<br />

<strong>Stagonospora</strong> nodorum, common<br />

name septoria glume blotch (SGB),<br />

has been difficult because no genes<br />

for immunity exist. There are<br />

numerous sources <strong>of</strong> genetic<br />

resistance; however, most have<br />

been labeled as partial resistance<br />

<strong>and</strong> in some manner delay the<br />

growth <strong>of</strong> the pathogen. In this<br />

study our objectives were to<br />

conduct a quantitative genetic<br />

analysis <strong>of</strong> three components <strong>of</strong><br />

resistance to SGB in wheat. Second,<br />

to compare 10 wheat parents, 21 F 1 ,<br />

<strong>and</strong> 21 F 2 crosses for foliar disease<br />

reaction induced by inoculation,<br />

<strong>and</strong> third, to investigate the<br />

heritability estimates <strong>and</strong> gene<br />

numbers governing SGB resistance.<br />

Materials <strong>and</strong> Methods<br />

Six s<strong>of</strong>t winter wheat <strong>and</strong> four<br />

hard wheat genotypes that varied<br />

greatly in resistance to SGB were<br />

used in two incomplete diallel<br />

crosses. S<strong>of</strong>t wheat lines were<br />

TX92D7374, TX82-11, L890682,<br />

18NT, ‘Coker 9803’, <strong>and</strong> ‘Coker<br />

9543’. Hard wheat lines were<br />

TX91V3308, TX84V344, ‘TAM 300’,<br />

<strong>and</strong> SWM14240. SWM14240 is a<br />

wheat line selected in the<br />

northwestern US from <strong>CIMMYT</strong><br />

germplasm <strong>and</strong> therefore is likely<br />

to have different genes for SGB<br />

resistance. Seeds were germinated<br />

on moist filter paper for two days<br />

<strong>and</strong> then transplanted into pots<br />

containing a soil/peat moss<br />

mixture. Two plants in each pot<br />

were treated as two samples <strong>of</strong> the<br />

particular hybrid or parent. Plants<br />

were inoculated <strong>and</strong> placed in a<br />

humidity chamber for 50 h, as<br />

previously described (Nelson,<br />

1980). Hayman’s approach (1954)<br />

was used to calculate components<br />

<strong>of</strong> genetic variations.<br />

Results<br />

Estimation <strong>of</strong> components<br />

<strong>of</strong> variation<br />

Hard winter wheat: Incubation<br />

period. Parents with the highest<br />

frequency <strong>of</strong> dominant alleles have<br />

the smallest Vr (variance) <strong>and</strong> Wr<br />

(parent-<strong>of</strong>fspring covariance); thus<br />

their relative position along the Vr/<br />

Wr regression line reflects the<br />

frequency <strong>of</strong> dominant alleles.<br />

Sorting the parents used in these<br />

crosses in order <strong>of</strong> decreasing<br />

dominance gave the follow<br />

sequence: TX91V3308, TAM 300,<br />

TX84V344, <strong>and</strong> SWM14240.<br />

SWM14240 had least dominant<br />

effect (Figure 1). The fixable<br />

variation D <strong>and</strong> the dominance<br />

component H were calculated<br />

(Table 1). Narrow sense heritability<br />

was 66% <strong>and</strong> 61% in F 1 in F 2<br />

generations, respectively. Gene<br />

number governing the incubation<br />

period was 2-3 (Table 1).<br />

Latent period. The analysis <strong>of</strong> Wr,<br />

Vr, <strong>and</strong> Wr/Vr graphical statistics<br />

in both F 1 <strong>and</strong> F 2 provided detail<br />

information on the interrelation<br />

between the parents. The order <strong>of</strong><br />

decreasing dominance was<br />

TX84V344, TAM 300, TX91V3308,<br />

<strong>and</strong> SWM14240 for F 1 (Figure 1);<br />

TAM 300, TX91V3308, TX84V344,<br />

<strong>and</strong> SWM14240 for F 2 . Average<br />

degree <strong>of</strong> dominance was partial<br />

dominance because <strong>of</strong> the positive<br />

intercepts. SWM14240 had the least<br />

dominant effect. TAM 300 had the<br />

2.0<br />

(F1 LP) 4<br />

4<br />

(F2 LP)<br />

Wr<br />

2<br />

1<br />

4<br />

(F1 LP)<br />

0.5<br />

0<br />

3<br />

2<br />

1<br />

3<br />

1<br />

3<br />

1<br />

2<br />

3<br />

2<br />

4<br />

(F1 LP)<br />

0.5 Vr 2.0<br />

Figure 1. Parent-<strong>of</strong>fspring covariance (Wr)/<br />

variance (Vr) for SNB incubation period <strong>and</strong><br />

latent period <strong>of</strong> F1 <strong>and</strong> F2 hard winter wheat<br />

crosses.<br />

1 = TX91V3308 2 = TX84V344<br />

3 = TAM 300 4 = SWM14240


largest dominant effect. Narrow<br />

sense heritability was 64% <strong>and</strong> 68%<br />

in F 1 <strong>and</strong> F 2 generations,<br />

respectively (Table 1). There were<br />

about three genes governing the<br />

resistance <strong>of</strong> latent period. The<br />

additive effect D <strong>and</strong> dominance<br />

component H1 <strong>and</strong> H2 were also<br />

calculated. Additive D was 3.74 for<br />

the F 1 <strong>and</strong> 3.75 for the F 2<br />

generation.<br />

S<strong>of</strong>t winter wheat: Incubation<br />

period. The order <strong>of</strong> decreasing<br />

dominance was TX82-11,<br />

TX92D7374, 18NT, Coker 9803,<br />

Coker 9543, <strong>and</strong> L890682 for F 1 ,<br />

<strong>and</strong> TX82-11, 18NT, Coker 9803,<br />

Coker 9543, L890682, <strong>and</strong><br />

TX92D7374 for F 2 generations.<br />

Mean square <strong>of</strong> Wr-Vr was not<br />

significant. Dominance seems to<br />

account for the major proportion <strong>of</strong><br />

the non-additive variation. The<br />

additive variation D <strong>and</strong> the<br />

dominant component H were<br />

estimated in Table 2. The<br />

dominance ratio H1/D, an estimate<br />

<strong>of</strong> the average level <strong>of</strong> dominance,<br />

was 0.7 <strong>and</strong> 0.6 for F 1 <strong>and</strong> F 2<br />

generations, respectively. This<br />

indicates that resistance for<br />

incubation period was partially<br />

dominant. There were 2-3 genes<br />

controlling IP in the s<strong>of</strong>t wheat.<br />

Percent necrosis. The order <strong>of</strong><br />

decreasing dominance was TX82-<br />

11, 18NT, Coker 9803, TX92D7374,<br />

Coker 9543, <strong>and</strong> L890682.<br />

Dominance <strong>and</strong> non-allelic<br />

interaction was important in<br />

necrosis. H/D>1 indicates that<br />

necrosis had over-dominance<br />

effects (Table 2). Heritability<br />

estimated for necrosis was similar<br />

by different methods <strong>of</strong> calculation.<br />

Narrow sense heritability for<br />

necrosis was relatively low. Only<br />

one gene controlled % necrosis in<br />

s<strong>of</strong>t winter wheat. It was different<br />

from hard winter wheat <strong>and</strong> may<br />

have resulted from the genotype X<br />

environment interaction-over<br />

dominance effects.<br />

Discussion<br />

Nelson <strong>and</strong> Marshall (1990)<br />

stated that resistance that reduced<br />

the infection rate had polygenic<br />

inheritance. Jeger (1980) indicated<br />

that resistance to SGB may involve<br />

four independent polygenes. The<br />

results <strong>of</strong> this study agree with<br />

these previous studies. IP, LP <strong>and</strong><br />

PN had polygenic inheritance.<br />

Incubation period, LP, <strong>and</strong> NP were<br />

controlled by 2 to 3, 3, <strong>and</strong> 1 to 4<br />

genes in the hard <strong>and</strong> s<strong>of</strong>t wheat<br />

studies, respectively. Previous<br />

studies (Nelson 1980; Mullaney et<br />

al., 1982; Scott et al., 1982) also<br />

provided evidence for polygenic<br />

control <strong>of</strong> host reaction to S.<br />

nodorum both at the seedling plant<br />

stage <strong>and</strong> the mature plant stage.<br />

The most accepted hypothesis is<br />

that these polygenes are<br />

Table 1. Genetic variation <strong>of</strong> components <strong>of</strong> partial resistance for hard winter wheat crosses.<br />

Incubation period Latent period Necrosis<br />

Component F1 F2 F1 F2 F1 F2<br />

D 2.51 2.38 3.74 3.75 365.76 360.37<br />

H1 2.03 2.51 2.16 2.34 459.48 441.15<br />

H2 0.88 1.34 1.33 0.52 265.07 59.52<br />

F 2.40 1.87 3.10 3.24 475.17 350.18<br />

E 0.32 0.51 0.76 0.70 37.12 42.52<br />

H narrow 66% 61% 64% 68% 30% 77%<br />

Number <strong>of</strong> genes 3 2 3 3 4 3<br />

Partial Resistance to <strong>Stagonospora</strong> nodorum in Wheat 161<br />

independently inherited. Our<br />

research may indicate some <strong>of</strong> the<br />

polygenes may have multi-effects.<br />

For example, one <strong>of</strong> the latent<br />

period resistant genes may also be<br />

a resistant gene in incubation<br />

period polygenes.<br />

Analysis <strong>of</strong> components <strong>of</strong><br />

variation. The H1/D value <strong>of</strong> F 1<br />

hard wheat, <strong>and</strong> F 1 <strong>and</strong> F 2 s<strong>of</strong>t<br />

wheat were less than one with the<br />

exception <strong>of</strong> F 2 hard winter wheat<br />

(1.1). This means that additive gene<br />

action was <strong>of</strong> greater importance in<br />

the components in this study.<br />

Therefore, incubation period could<br />

be a very useful selection criterion<br />

for SGB resistant breeding. Both F 1<br />

<strong>and</strong> F 2 hard winter wheat crosses<br />

had larger additive value D than<br />

dominant value H1. It is likely that<br />

the non-additive genetic effects<br />

were due predominantly to<br />

contribution from additive x<br />

additive epistatic gene action rather<br />

than dominance.<br />

Additive gene action <strong>and</strong><br />

additive x additive epistatic gene<br />

action are most easily utilized <strong>and</strong><br />

exploited in homozygous<br />

genotypes. Latent period would<br />

also be a good selection criterion<br />

for SGB resistant breeding. H1/D<br />

values for hard <strong>and</strong> s<strong>of</strong>t wheat<br />

crosses were greater than 1. Over<br />

dominance may have been present<br />

for necrosis. Nelson (1980),<br />

Table 2. Genetic variation <strong>of</strong> components <strong>of</strong><br />

partial resistance for s<strong>of</strong>t winter wheat crosses.<br />

Component Incubation period Necrosis<br />

F1 F2 F1<br />

D 1.19 1.60 3.66<br />

H 1 0.83 0.92 148.27<br />

H 2 0.50 1.44 123.54<br />

F 0.55 1.24 -4.00<br />

E 0.91 0.72 26.99<br />

H narrow 32% 39% 22%<br />

Number <strong>of</strong> genes 2 3 1


Session 6C — C.G. Du, L.R. Nelson, <strong>and</strong> M.E. McDaniel<br />

162<br />

Wilkinson et al. (1990), <strong>and</strong><br />

Bostwick et al. (1993) had similar<br />

results. It is reasonable for the<br />

components <strong>of</strong> resistance to show<br />

dominant gene action in the F 1<br />

generation because <strong>of</strong> gene<br />

interaction. From the breeder’s<br />

viewpoint, necrosis should not be<br />

an early generation selection<br />

criterion for SGB resistant breeding.<br />

Heritability. Estimates <strong>of</strong><br />

heritability in all crosses studied<br />

were moderate to high. Relatively<br />

high heritability for spike <strong>and</strong> flag<br />

leaf reaction to SGB have been<br />

estimated in previous studies<br />

(Rosielle <strong>and</strong> Brown, 1980; Fried<br />

<strong>and</strong> Meister, 1987; Bostwick et al.<br />

1993). Therefore, early generation<br />

selection for SGB should be<br />

effective.<br />

References<br />

Bostwick, D.E., H.W. Ohm, <strong>and</strong> G.<br />

Shaner. 1993. Inheritance <strong>of</strong><br />

septoria glume blotch resistance in<br />

wheat. Crop Sci. 33:493-443.<br />

Fried, P.M., <strong>and</strong> E. Meister. 1987.<br />

Inheritance <strong>of</strong> leaf <strong>and</strong> head<br />

resistance <strong>of</strong> winter wheat to<br />

<strong>Septoria</strong> nodorum in a diallel cross.<br />

Genetics 77:1371-1375.<br />

Hayman, B.I. 1954. The theory <strong>and</strong><br />

analysis <strong>of</strong> diallel crosses. Genetics<br />

39:789-809.<br />

Jeger, M.J. 1980. Ultivariate models <strong>of</strong><br />

the components <strong>of</strong> partial<br />

resistance. Protection Ecology<br />

2:265-269.<br />

Mullaney, E.J., J.M. Martin, <strong>and</strong> A.L.<br />

Scharen. 1982. Generation mean<br />

analysis to identify <strong>and</strong> partition<br />

the components <strong>of</strong> genetic<br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Euphytica 31:539-545.<br />

Nelson, L.R. 1980. Inheritance <strong>of</strong><br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Crop Sci. 20:447-449.<br />

Nelson, L.R., <strong>and</strong> D. Marshall. 1990.<br />

Breeding wheat for resistance to<br />

<strong>Septoria</strong> nodorum <strong>and</strong> S. tritici.<br />

Advances in Agronomy 44:257-<br />

277.<br />

Rosielle, A.A., <strong>and</strong> A.G.P. Brown.<br />

1980. Selection for resistance to<br />

<strong>Septoria</strong> nodorum Berk. in wheat.<br />

Euphytica 29:337-346.<br />

Scott, P.R., P.W. Benedikz, <strong>and</strong> C.J.<br />

Cox. 1982. A genetic study on the<br />

relationship between height, time<br />

<strong>of</strong> ear emergence, <strong>and</strong> resistance to<br />

<strong>Septoria</strong> nodorum in wheat. Plant<br />

Pathol. 31:45-60.<br />

Wilkinson, C.A., J.P. Murphy, <strong>and</strong><br />

R.R. Rufty. 1990. Diallel analysis <strong>of</strong><br />

components <strong>of</strong> partial resistance to<br />

<strong>Septoria</strong> nodorum in wheat. Plant<br />

Dis. 74:47-50.


Comparison <strong>of</strong> Methods <strong>of</strong> Screening for <strong>Stagonospora</strong><br />

nodorum Resistance in Winter Wheat<br />

D.E. Fraser, 1 J.P. Murphy, 1 <strong>and</strong> S. Leath 2 (Poster)<br />

1 Department <strong>of</strong> Plant Pathology, North Carolina State Univ., Raleigh, NC, USA<br />

2 Department <strong>of</strong> Plant Pathology, USDA-ARS, NCSU, Raleigh, NC, USA<br />

Abstract<br />

Isolates <strong>of</strong> <strong>Stagonospora</strong> nodorum that varied in levels <strong>of</strong> aggressiveness were used in both controlled environment <strong>and</strong><br />

field tests to determine whether isolate aggressiveness enhanced selection <strong>of</strong> resistant wheat genotypes. Two segregating wheat<br />

populations that differed in mean levels <strong>of</strong> resistance to S. nodorum were developed. Components <strong>of</strong> resistance measured in<br />

controlled environments indicated significant differences among isolate treatments. Measurement <strong>of</strong> disease severity in field<br />

studies also indicated significant differences among the isolate treatments. For the breeding population (r<strong>and</strong>om selections <strong>of</strong><br />

genotypes in F 3 , F 4 , F 5 , <strong>and</strong> F 6 generations), incubation period measured in both juvenile <strong>and</strong> adult plant tests were most<br />

highly correlated with resistance measured in the field. For the biparental population (progeny <strong>of</strong> a resistant by susceptible<br />

cross), incubation period <strong>and</strong> percent leaf area diseased on the flag leaf for adult plant tests were most highly correlated with<br />

resistance measured in the field. Rankings <strong>of</strong> genotypes based on their resistance in both field <strong>and</strong> controlled environment<br />

tests indicated that five <strong>of</strong> the ten most resistant genotypes measured under controlled conditions were also identified as<br />

resistant in field tests. Similarly, four <strong>of</strong> the ten most susceptible genotypes measured under controlled conditions were also<br />

identified as susceptible in field tests. Combinations <strong>of</strong> adult <strong>and</strong> juvenile screening tests may allow for the identification <strong>of</strong> up<br />

to eight <strong>of</strong> ten genotypes resistant in controlled environments that are also resistant in field tests.<br />

Glume blotch, caused by the<br />

ascomycete fungus <strong>Stagonospora</strong><br />

nodorum (teleomorph Leptosphaeria<br />

nodorum) is an economically<br />

important disease <strong>of</strong> wheat.<br />

Resistance to S. nodorum is<br />

controlled by multiple genes (Jeger<br />

et al., 1983; Mullaney et al., 1982;<br />

Nelson <strong>and</strong> Gates, 1982; Wilkinson<br />

et al., 1990) <strong>and</strong> expression <strong>of</strong><br />

resistance is strongly affected by<br />

the environment.<br />

Previous studies attempting to<br />

correlate components <strong>of</strong> host<br />

resistance measured in controlled<br />

environments with host resistance<br />

under field conditions have<br />

reported variable results<br />

(Wilkinson et al., 1990; Rufty et al.,<br />

1981). A better underst<strong>and</strong>ing <strong>of</strong><br />

the relationship between isolate<br />

aggressiveness <strong>and</strong> host response<br />

could provide more effective<br />

selection methods for resistance to<br />

S. nodorum. The objective <strong>of</strong> this<br />

study was 1) to use isolates with<br />

different levels <strong>of</strong> aggressiveness to<br />

measure differences in components<br />

<strong>of</strong> resistance among genotypes in<br />

two segregating wheat<br />

populations, under both field <strong>and</strong><br />

controlled environments <strong>and</strong> 2) to<br />

determine which components are<br />

the most effective in selection <strong>of</strong><br />

resistant genotypes.<br />

Materials <strong>and</strong> Methods<br />

Two wheat populations were<br />

used in all studies. Population 1<br />

(biparental) consisted <strong>of</strong> the F3:5<br />

progeny <strong>of</strong> NKC 8427 (resistant) by<br />

Caldwell (susceptible) cross.<br />

Population 2 (breeding) was<br />

composed <strong>of</strong> r<strong>and</strong>om F3, F4, F5,<br />

<strong>and</strong> F6 lines from a southern US<br />

s<strong>of</strong>t red winter wheat nursery in<br />

North Carolina. A total <strong>of</strong> 50 lines<br />

from each population were used in<br />

all studies. A collection <strong>of</strong> 60 NC<br />

field isolates was screened in a<br />

seedling assay <strong>and</strong> the most<br />

163<br />

aggressive isolate <strong>and</strong> least<br />

aggressive isolate in the collection<br />

were selected for further study. The<br />

same isolate treatments were used<br />

in all subsequent studies <strong>and</strong><br />

consisted <strong>of</strong> the most <strong>and</strong> least<br />

aggressive isolates (A <strong>and</strong> D,<br />

respectively) applied singly, in<br />

combination (AD), <strong>and</strong> a noninoculated<br />

control.<br />

Field tests<br />

Hill-plots were planted on a 0.3m<br />

square grid in three replications<br />

at Laurel Springs, NC, in 1997 <strong>and</strong><br />

at both Laurel Springs <strong>and</strong> Kinston,<br />

NC, in 1998. Each population <strong>of</strong> 50<br />

genotypes was inoculated with a<br />

spore suspension <strong>of</strong> a different<br />

isolate treatment at either Feekes<br />

growth stage (GS) 10.1 in 1997 <strong>and</strong><br />

Feekes GS 9 in 1998. Percent leaf<br />

area diseased (% LAD) was<br />

recorded at four bi-weekly<br />

intervals after inoculation using the<br />

Saari-Prescott scale.


Session 6C — D.E. Fraser, J.P. Murphy, <strong>and</strong> S. Leath<br />

164<br />

Greenhouse tests<br />

Seeds <strong>of</strong> each genotype were<br />

over-planted <strong>and</strong> thinned to two<br />

plants per pot <strong>and</strong> allowed to grow<br />

in alternating 12-hour light <strong>and</strong><br />

dark periods at 16 to 26 o C.<br />

Experimental design was a<br />

r<strong>and</strong>omized complete block with<br />

four replications in time.<br />

At GS 10 the penultimate leaf <strong>of</strong><br />

one plant/pot was marked <strong>and</strong> a 2<br />

ml droplet <strong>of</strong> a single isolate spore<br />

suspension st<strong>and</strong>ardized to 1 x 10 6<br />

spores ml -1 applied to the leaf.<br />

Presence/absence <strong>of</strong> a lesion at 5<br />

<strong>and</strong> 10 days was recorded.<br />

Following the last rating on the<br />

penultimate leaf <strong>of</strong> one plant, the<br />

flag leaf <strong>of</strong> the other plant/pot was<br />

marked. Spore suspensions <strong>of</strong> each<br />

isolate treatment were applied to 50<br />

genotypes per population using<br />

h<strong>and</strong> atomizers until run-<strong>of</strong>f<br />

occurred. Glass slides placed in the<br />

plant canopy during inoculation<br />

indicated that 50-75 spores were<br />

deposited per mm -2 . Incubation<br />

period <strong>and</strong> % LAD on each plant<br />

were recorded.<br />

Growth chamber<br />

Seeds <strong>of</strong> each genotype were<br />

planted <strong>and</strong> grown for 21 days in<br />

12 h light <strong>and</strong> darkness periods in<br />

day/night temperatures <strong>of</strong> 22/<br />

18 o C. At the 2-leaf stage, spore<br />

suspensions <strong>of</strong> the same four<br />

isolate treatments as described<br />

previously were applied to 50<br />

genotypes per population using<br />

h<strong>and</strong> atomizers until run-<strong>of</strong>f<br />

occurred, with approximately 65<br />

spores deposited per mm -2 <strong>of</strong> leaf<br />

tissue. Incubation period <strong>and</strong> %<br />

LAD on each plant were recorded.<br />

Experimental design was a<br />

r<strong>and</strong>omized complete block with<br />

four replications.<br />

Detached leaf test–adult <strong>and</strong><br />

juvenile<br />

Seeds <strong>of</strong> each genotype were<br />

planted <strong>and</strong> grown in the<br />

greenhouse for three weeks<br />

(juvenile) <strong>and</strong> ten weeks (adult),<br />

respectively. For the juvenile test,<br />

fully exp<strong>and</strong>ed second leaves were<br />

cut from each genotype in both<br />

populations. For the adult test,<br />

plants were grown to Feeke’s GS<br />

10.1, <strong>and</strong> the flag leaf was excised<br />

from each plant. Each single leaf<br />

was cut into four equal sized pieces<br />

<strong>and</strong> placed on 150 ppm<br />

benzimidazole agar in a box<br />

divided in 4-cm grid squares. Each<br />

grid square represented each <strong>of</strong> the<br />

isolate treatments on a single<br />

genotype. A 2 ml droplet <strong>of</strong> a single<br />

isolate spore suspension was<br />

placed in the center <strong>of</strong> each leaf<br />

piece. Incubation period, lesion<br />

expansion rate, <strong>and</strong> final lesion size<br />

was recorded. Both the adult <strong>and</strong><br />

juvenile tests were replicated four<br />

times.<br />

Results <strong>and</strong> Discussion<br />

The level <strong>of</strong> resistance in the<br />

breeding population was generally<br />

higher than the biparental<br />

population in all tests. The AUDPC<br />

values indicated significant<br />

differences among isolate<br />

treatments in the field at Laurel<br />

Springs in 1997 <strong>and</strong> at both<br />

locations in 1998 (Table 1).<br />

Genotype by environment<br />

interaction was significant.<br />

Isolate treatments in<br />

greenhouse evaluations were<br />

significantly different for most <strong>of</strong><br />

the components <strong>of</strong> resistance<br />

measured (Tables 2 <strong>and</strong> 3). Results<br />

<strong>of</strong> inoculations with the most <strong>and</strong><br />

least aggressive isolates in<br />

combination (AD) were similar to<br />

inoculations with the most<br />

aggressive isolate alone <strong>and</strong> both <strong>of</strong><br />

these treatments resulted in higher<br />

numbers <strong>of</strong> lesions, more rapid<br />

rates <strong>of</strong> lesion expansion <strong>and</strong><br />

higher % LAD than inoculation the<br />

least aggressive isolate. Shorter<br />

incubation periods were not<br />

consistently associated with the<br />

most aggressive isolate or the most<br />

<strong>and</strong> least aggressive isolate<br />

together. Components <strong>of</strong> resistance<br />

measured in controlled<br />

environments were highly<br />

correlated with each other (Table 4).<br />

Few components <strong>of</strong> resistance<br />

measured in controlled<br />

environments correlated with field<br />

measured disease severity. A<br />

comparison <strong>of</strong> the controlled<br />

environment studies with field data<br />

from Kinston in 1998 indicated that<br />

adult plant incubation period <strong>and</strong><br />

% LAD on the flag leaf for adult<br />

plant spray inoculation were highly<br />

correlated (P>0.01) with field<br />

Table 1. Area under the disease progress curve values for disease severities measured on hill<br />

plots inoculated with isolates <strong>of</strong> <strong>Stagonospora</strong> nodorum.<br />

Laurel Springs 1997 Laurel Springs 1998 Kinston 1998<br />

Isolate treatments Breeding Biparental Breeding Biparental Breeding<br />

A (most) a 2724b 1669b 2196a 1678a 1587a<br />

D (least) 2783a 1667b 1972c 1557c 1561c<br />

AD (both) 2770a 1806a 1733d 1608b 1628a<br />

NI (control) 2694c 1829a 2087b 1367d 1452d<br />

a ‘Most’ <strong>and</strong> ‘least’ aggressive refer to isolate reaction measured in the isolate screening assay done on<br />

seedlings in controlled conditions.


disease severity in the biparental<br />

population. Adult <strong>and</strong> juvenile<br />

plant incubation period <strong>and</strong> lesion<br />

expansion rate measured in the<br />

detached leaf test were highly<br />

correlated with field disease<br />

severity measured at Kinston in<br />

1998 in the breeding population.<br />

Similar results were obtained<br />

for disease measured at Laurel<br />

Springs in 1997. Components <strong>of</strong><br />

resistance were not highly<br />

correlated with field disease<br />

measured at Laurel Springs in 1998.<br />

Comparison <strong>of</strong> Methods <strong>of</strong> Screening for <strong>Stagonospora</strong> nodorum Resistance in Winter Wheat 165<br />

Resistant genotypes<br />

A comparison <strong>of</strong> genotype<br />

rankings measured in the field with<br />

those measured in controlled<br />

environments indicate that for the<br />

biparental population at Kinston in<br />

1998, five <strong>of</strong> the ten most resistant<br />

genotypes in the field were also<br />

identified by the measurements <strong>of</strong><br />

the incubation period <strong>and</strong> % LAD<br />

on the flag leaf in the adult plant<br />

tests. Similarly, rankings based on<br />

% LAD measured in the adult plant<br />

spray test ranked five <strong>of</strong> the same<br />

ten most resistant genotypes<br />

identified by the field disease<br />

severity values at Laurel Springs in<br />

1998. The numbers were slightly<br />

lower for the breeding population.<br />

At Kinston in 1998, four <strong>of</strong> the ten<br />

most resistant genotypes in the<br />

field were also identified as<br />

resistant by measurements <strong>of</strong> %<br />

LAD for adult plant spray<br />

inoculation. Combining adult plant<br />

greenhouse tests <strong>and</strong> juvenile<br />

detached leaf tests could increase<br />

this number to seven or eight <strong>of</strong> the<br />

top ten resistant genotypes<br />

identified in controlled<br />

Table 2. Components <strong>of</strong> resistance to <strong>Stagonospora</strong> nodorum measured on F 3:5 lines <strong>of</strong> NKC 8433 x Caldwell cross in controlled environments.<br />

Dlt a Juvenile Dlt Adult Phytotron(Juvenile) Greenhouse<br />

Isolate Incubation # <strong>of</strong> Incubation Lesion Incubation # <strong>of</strong> Droplet Spray<br />

treatments b (days) lesions (days) expansion ( days) lesions incubation %LAD<br />

A (most) 7.9 a 2.4 b 12.7 c 0.14 b 12.9 a 2.4 bc 3.8 a 5.5 ab<br />

D (least) 8.1 b 2.3 b 10.4 b 0.13 a 13.1 a 1.7 a 4.3 a 3.9 a<br />

AD (both) 7.6 a 2.7 c 9.4 a 0.20 c 12.9 a 1.8 b 5.4 b 4.8 a<br />

NI(control) 20.0 c 0.0 a - - 17.1 b 1.2 a 20.0 c -<br />

a Detached leaf test.<br />

‘b ‘Most’ <strong>and</strong> ‘least’ aggressive refer to isolate reaction measured in the isolate screening assay done on seedlings in controlled conditions.<br />

Table 3. Components <strong>of</strong> resistance to <strong>Stagonospora</strong> nodorum measured on r<strong>and</strong>om F 3 , F 4 , F 5 <strong>and</strong> F 6 lines from a southern US winter wheat nursery<br />

in controlled environments.<br />

Dlt a Juvenile Dlt Adult Phytotron(Juvenile) Greenhouse<br />

Isolate Incubation # <strong>of</strong> Incubation Lesion Incubation # <strong>of</strong> Droplet Spray<br />

treatments b (days) lesions (days) expansion ( days) lesions incubation %LAD<br />

A (most) 8.4 a 2.5 b 12.7 b 0.11 a 13.5 a 2.1 c 5.1 b 3.7 a<br />

D (least) 9.0 b 2.4 b 10.8 a 0.12 b 12.8 a 1.6 b 3.4 a 2.8 a<br />

AD (both) 8.3 a 2.4 b 11.0 a 0.19 c 13.7 a 2.2 c 4.9 b 4.1 ab<br />

NI (control) 20.0 c 0.0 a - - 17.5 b 0.9 a 20.0 c -<br />

a Detached leaf test.<br />

b ‘ Most’ <strong>and</strong> ‘least’ aggressive refer to isolate reaction measured in the isolate screening assay done on seedlings in controlled conditions.<br />

Table 4. Correlations among resistance components measured in controlled environments <strong>and</strong> in field tests.<br />

Treatment A (most) a Treatment AD(most/least) Treatment D(least)<br />

Resistance component Biparental Breeding Biparental Breeding Biparental Breeding<br />

Adult dlt – incubation 0.23 * 0.30 * NS NS NS 0.25 *<br />

Adult dlt- lesion expansion NS -0.34 ** NS NS 0.2 0* 0.20 *<br />

Juvenile dlt – incubation NS -0.40 ** NS NS NS NS<br />

Juvenile dlt – lesion size NS NS 0.20 * 0.20* 0.27 * NS<br />

Adult spray test - %LAD NS NS 0.23 * NS 0.33** 0.14<br />

a ‘Most’ <strong>and</strong> ‘least’ aggressive refer to isolate reaction measured in the isolate screening assay done on seedlings in controlled conditions.


Session 6C — C.G. Du, L.R. Nelson, <strong>and</strong> M.E. McDaniel<br />

166<br />

environment tests also being<br />

identified as resistant under field<br />

conditions.<br />

Susceptible genotypes<br />

Of the ten most susceptible<br />

genotypes in each population, four<br />

were also identified as susceptible<br />

by measurement <strong>of</strong> juvenile plant<br />

incubation period. Combining<br />

results from both the adult plant<br />

<strong>and</strong> juvenile plant detached leaf<br />

tests resulted in as high as seven <strong>of</strong><br />

the ten most resistant genotypes in<br />

the field also being identified as<br />

resistant under controlled<br />

conditions.<br />

In general, evaluation <strong>of</strong><br />

components <strong>of</strong> resistance may<br />

provide a more efficient selection<br />

method by allowing for the<br />

removal <strong>of</strong> extremely susceptible<br />

types <strong>and</strong> identification <strong>of</strong> resistant<br />

types before large-scale field<br />

selection. Based on our results <strong>and</strong><br />

practicability to the plant breeder, a<br />

combination <strong>of</strong> juvenile detached<br />

leaf tests <strong>and</strong> adult plant<br />

inoculations (droplet or spray) in<br />

the greenhouse could aid in the<br />

selection <strong>of</strong> parental materials with<br />

resistance to S. nodorum. These<br />

methods could also be used for<br />

screening the progeny <strong>of</strong><br />

segregating populations. Highly<br />

aggressive isolates or combinations<br />

<strong>of</strong> isolates with different levels <strong>of</strong><br />

aggressiveness help to maximize<br />

differences among genotypes in<br />

both field <strong>and</strong> controlled<br />

environments<br />

Epidemiology<br />

In addition to measuring<br />

components <strong>of</strong> resistance to S.<br />

nodorum in the field <strong>and</strong> in<br />

controlled conditions, molecular<br />

analyses <strong>of</strong> the development <strong>of</strong> the<br />

epidemics in the field were also<br />

studied using these same isolate<br />

treatments <strong>and</strong> populations. The<br />

information provided by the<br />

molecular analyses may explain<br />

differences among treatments in<br />

the field <strong>and</strong> help elucidate the role<br />

<strong>of</strong> the natural inoculum in<br />

epidemic development.<br />

References<br />

Jeger, M.J., D.G. Jones, <strong>and</strong> H.<br />

Griffiths. 1983. Components <strong>of</strong><br />

partial resistance <strong>of</strong> wheat<br />

seedlings to <strong>Septoria</strong> nodorum.<br />

Euphytica 32:575-584.<br />

Mullaney, E.J., J.M. Martin, <strong>and</strong> A.L.<br />

Scharen. 1982. Generation mean<br />

analysis to identify <strong>and</strong> partition<br />

the components <strong>of</strong> genetic<br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Euphytica 31:539-545.<br />

Nelson, L.R., <strong>and</strong> C.E. Gates. 1982.<br />

Genetics <strong>of</strong> host plant resistance <strong>of</strong><br />

wheat to <strong>Septoria</strong> nodorum. Crop<br />

Sci. 22:771-773.<br />

Rufty, R.C., T.T. Hebert, <strong>and</strong> C.F.<br />

Murphy. 1981. Evaluation <strong>of</strong><br />

resistance to <strong>Septoria</strong> nodorum in<br />

wheat. Plant Dis. 65:406-409.<br />

Wilkinson, C.A., J.P. Murphy, <strong>and</strong><br />

R.C. Rufty. 1990. Diallel analysis <strong>of</strong><br />

components <strong>of</strong> partial resistance to<br />

<strong>Septoria</strong> nodorum in wheat. Plant<br />

Dis. 74:47-50.


Response <strong>of</strong> Winter Wheat Genotypes to Artificial<br />

Inoculation with Several <strong>Septoria</strong> tritici Populations<br />

M. Mincu (Poster)<br />

Research Institute for <strong>Cereals</strong> <strong>and</strong> Industrial Crops, Fundulea, Romania<br />

Abstract<br />

A study was conducted to identify genotypic differences in the response <strong>of</strong> winter wheat to artificial inoculation with<br />

several <strong>Septoria</strong> tritici populations. Twenty-two genotypes were grown in two locations <strong>and</strong> artificially inoculated with four<br />

<strong>Septoria</strong> tritici populations. Significant effects <strong>of</strong> host genotype, pathogen population, <strong>and</strong> host-pathogen interaction were<br />

found. Most released cultivars were susceptible. Several entries, including some Aegilops spp. derivatives, were resistant to<br />

all <strong>Septoria</strong> populations in both locations. <strong>Septoria</strong> tritici populations differed both in aggressiveness <strong>and</strong> virulence. A<br />

strong host-pathogen interaction was found in several cultivars with medium average resistance.<br />

<strong>Septoria</strong> tritici is a major foliar<br />

disease <strong>of</strong> wheat in Romania. Yield<br />

losses caused by septoria leaf<br />

blotch can reach 25% in years that<br />

favor the disease. Most currently<br />

grown wheat cultivars are more or<br />

less susceptible to <strong>Septoria</strong> tritici.<br />

Therefore, resistance to this<br />

pathogen is a high-priority<br />

breeding goal.<br />

Material <strong>and</strong> Methods<br />

Twenty-two winter wheat<br />

genotypes were grown on 2-m+<br />

plots using three replications in<br />

each <strong>of</strong> two locations (Fundulea in<br />

the south <strong>and</strong> Brasov in the central<br />

part, near the mountains). Artificial<br />

inoculation was conducted by<br />

spraying the plants at heading with<br />

a spore suspension diluted to 10<br />

spores/ml, from four <strong>Septoria</strong> tritici<br />

populations, collected from several<br />

regions <strong>of</strong> Romania. After spraying,<br />

the plants were covered with<br />

plastic sacks to maintain the<br />

humidity necessary for infection.<br />

The intensity <strong>of</strong> the attack was read<br />

30 days after inoculation, using the<br />

0 to 9 scale developed by Saari <strong>and</strong><br />

Prescott at <strong>CIMMYT</strong> (1995). For<br />

statistical analysis, the arc sin<br />

transformation was used.<br />

Significance <strong>of</strong> differences among<br />

genotypes was determined using<br />

the Duncan test.<br />

Results <strong>and</strong> Discussion<br />

The analysis <strong>of</strong> variance<br />

(ANOVA) shows there are<br />

significant effects <strong>of</strong> wheat<br />

genotypes, pathogen populations,<br />

<strong>and</strong> host x pathogen interaction<br />

(Table 1). Genotype classifications<br />

for each pathogen population in<br />

the two locations are presented in<br />

Tables 2 <strong>and</strong> 3. Most released<br />

cultivars were susceptible to all<br />

populations <strong>and</strong> in both locations.<br />

On the other h<strong>and</strong>, several entries<br />

showed low intensities <strong>of</strong> attack<br />

with all populations <strong>and</strong> in both<br />

locations (e.g. F91552G8-01, Admis,<br />

ZE16547, KS93U134, G1662-51,<br />

G557-5, G557-6). The last four are<br />

derivatives from crosses with<br />

Aegilops spp. Finally, some entries<br />

(such as Turda 81) showed a very<br />

variable response to different<br />

<strong>Septoria</strong> populations.<br />

167<br />

The differences between host<br />

genotypes are more pronounced<br />

than between pathogen<br />

populations. The pathogen<br />

populations produced different<br />

attack intensities in the two<br />

locations. In Fundulea, all <strong>Septoria</strong><br />

populations had a higher<br />

pathogenicity than in Brasov. In<br />

Brasov, the pathogenicity <strong>of</strong> all<br />

populations was approximately<br />

equal, but in Fundulea Population<br />

4 was less pathogenic, compared<br />

with the other populations. The<br />

host-pathogen interactions in the<br />

two locations are exemplified in<br />

Figures 1-4. A strong host-pathogen<br />

interaction was noticed for several<br />

cultivars in both locations.<br />

Reference<br />

Saari, E.E., <strong>and</strong> Prescott, M. 1975. A<br />

scale for appraising the foliar<br />

intensity <strong>of</strong> wheat diseases. Plant<br />

Dis. Rep. 59:377-380.<br />

Table 1. Analysis <strong>of</strong> variance (MS) for the<br />

intensity <strong>of</strong> <strong>Septoria</strong> tritici attack on the leaf in<br />

Fundulea <strong>and</strong> Brasov.<br />

Source <strong>of</strong> variation Fundulea Brasov<br />

Genotypes (G) 1479.276*** 696.308***<br />

Error (a) 81.600 5.582<br />

Populations (P) 3041.026*** 688.315***<br />

G x P 339.740*** 69.820***<br />

Error (b) 30.020 3.300


Session 6C — M. Mincu<br />

168<br />

Arcsin (sqrt (intensity %))<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

ZE-16547<br />

G 1662-51<br />

G 557-5<br />

KS93U134<br />

20<br />

P4 P2 P1 P3<br />

<strong>Septoria</strong> tritici populations<br />

Figure 1. Interaction between genotypes <strong>and</strong><br />

population (intensity <strong>of</strong> the attack on the leaf),<br />

Fundulea - sources <strong>of</strong> resistance.<br />

Arcsin (sqrt (intensity %))<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

Admis<br />

Dropia<br />

Lovrin 34<br />

Turda 81<br />

Apullum<br />

20<br />

P4 P2 P1 P3<br />

<strong>Septoria</strong> tritici populations<br />

Figure 3. Interaction between genotypes <strong>and</strong><br />

population (intensity <strong>of</strong> the attack on the leaf),<br />

Fundulea - cultivars.<br />

Arcsin (sqrt (intensity %))<br />

Arcsin (sqrt (intensity %))<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

P1 P4 P3 P2<br />

<strong>Septoria</strong> tritici populations<br />

Figure 2. Interaction between genotypes <strong>and</strong><br />

population (intensity <strong>of</strong> the attack on the leaf),<br />

Brasov - sources <strong>of</strong> resistance.<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

ZE-16547<br />

G 1662-51<br />

G 557-5<br />

KS93U134<br />

Admis<br />

Dropia<br />

Lovrin 34<br />

Turda 81<br />

Apullum<br />

20<br />

P1 P4 P3 P2<br />

<strong>Septoria</strong> tritici populations<br />

Figure 4. Interaction between genotypes <strong>and</strong><br />

population (intensity <strong>of</strong> the attack on the leaf),<br />

Brasov - cultivars.<br />

Table 2. The reaction <strong>of</strong> genotypes to <strong>Septoria</strong> tritici infection–intensity <strong>of</strong> attack on the leaf in Fundulea.<br />

Genotypes Population 1 Population 2 Population 3 Population 4 Mean<br />

ZE16547 26.6 a † 26.6 a 26.6 a 26.6 a 26.6<br />

G557-5 26.6 a 26.6 a 26.6 a 26.6 a 26.6<br />

G557-6 26.6 a 26.6 a 26.6 a 26.6 a 26.6<br />

KS93U134 26.6 a 26.6 a 26.6 a 26.6 a 26.6<br />

91552G8-01 28.8 a 28.8 a 30.8 a 26.6 a 28.7<br />

G1662-51 26.6 a 26.6 a 37.2 b 26.6 a 29.2<br />

649T2-1 28.8 a 33.0 ab 39.2 bc 26.6 a 31.9<br />

Lovrin 34 51.2 bc 43.1 cde 38.9 bc 34.9 ab 32.3<br />

Rapid 45.0 b 28.8 a 31.0 ab 26.6 a 32.8<br />

Admis 26.6 a 49.2 ef 43.3 cd 26.6 a 36.4<br />

577U1-106 46.9 b 48.9 ef 50.8 d 26.6 a 36.6<br />

Dropia 44.9 b 36.9 bc 42.7 c 33.0 a 39.4<br />

93396G1-1 26.6 a 52.8 ef 59.0 ef 26.6 a 41.2<br />

92033G2-1 55.0 bcd 28.8 a 51.2 de 33.2 ab 42.0<br />

91375GA-13 57.0 cd 31.0 b 63.9 fg 26.6 a 44.6<br />

93009G11-12 59.0 de 35.2 abc 61.2 ef 26.6 a 45.5<br />

Fundulea – 4 61.2 def 46.9 de 55.0 e 35.2 ab 49.6<br />

93122G6-2 26.6 a 61.2 gh 63.5 f 48.9 d 50.0<br />

Apullum 68.9 f 30.8 ab 59.7 ef 43.0 bc 50.6<br />

508U3-201 26.6 a 56.6 fg 72.8 g 46.9 cd 50.7<br />

Turda 81 83.9 g 43.0 cd 83.9 h 37.2 b 62.0<br />

Cadril 249 66.2 de 63.9 h 72.3 g 48.9 d 62.8<br />

† Means in columns followed by the same letter are not significantly different by Duncan’s test at the 0.05% level.


Response <strong>of</strong> Winter Wheat Genotypes to Artificial Inoculation with Several <strong>Septoria</strong> tritici Populations 169<br />

Table 3. The reaction <strong>of</strong> genotypes to <strong>Septoria</strong> tritici infection–intensity <strong>of</strong> attack on the leaf in Brasov.<br />

Genotypes Population 1 Population 2 Population 3 Population 4 Mean<br />

ZE16547 26.6 a † 26.6 a 26.6 a 26.6 a 26.6<br />

G557-6 26.6 a 26.6 a 26.6 a 26.6 a 26.6<br />

KS93U134 26.6 a 26.6 a 26.6 a 26.6 a 26.6<br />

91552G8-01 26.6 a 26.6 a 26.6 a 26.6 a 26.6<br />

Admis 26.6 a 26.6 a 26.6 a 26.6 a 26.6<br />

649T2-1 26.6 a 28.8 a 26.6 a 26.6 a 27.1<br />

G557-5 26.6 a 26.6 a 26.6 a 28.8 a 27.1<br />

Cadril 249 26.6 a 31.0 b 26.6 a 26.6 a 27.7<br />

G1662-51 26.6 a 33.2 b 26.6 a 26.6 a 28.2<br />

Rapid 26.6 a 28.8 a 37.2 c 28.8 a 30.3<br />

93396G1-1 26.6 a 37.2 c 33.2 b 26.6 a 30.9<br />

93009G11-12 26.6 a 39.2 c 33.2 b 26.6 a 31.4<br />

508U3-201 26.6 a 37.2 c 35.2 b 26.6 a 31.4<br />

Apullum 26.6 a 33.2 b 37.2 c 31.0 b 32.0<br />

Turda 81 33.2 b 26.6 a 43.1 d 31.0 b 33.5<br />

91375GA-13 31.0 b 45.0 d 35.2 b 26.6 a 34.5<br />

93122G6-2 26.6 a 45.0 d 37.2 c 31.0 b 35.0<br />

577U1-106 33.2 b 39.2 c 37.2 c 33.2 c 35.7<br />

92033G2-1 37.2 d 48.9 e 45.0 d 37.2 d 42.1<br />

Dropia 33.2 b 56.8 g 43.1 d 50.8 e 46.0<br />

Fundulea – 4 50.8 e 52.8 f 52.8 e 39.2 d 48.90<br />

Lovrin – 34 35.2 c 52.8 f 68.9 f 50.8 e 51.92<br />

† Means in columns followed by the same letter are not significantly different by Duncan’s test at the 0.05% level.


170<br />

Comparison <strong>of</strong> Greenhouse <strong>and</strong> Field Levels <strong>of</strong> Resistance<br />

to <strong>Stagonospora</strong> nodorum<br />

S.L. Walker, 1,2 S. Leath, 1,3 <strong>and</strong> J.P. Murphy 2 (Poster)<br />

1 Department <strong>of</strong> Plant Pathology, North Carolina State University, Raleigh, USA<br />

2 Department <strong>of</strong> Crop Science, North Carolina State University, Raleigh, USA<br />

3 USDA-ARS, North Carolina State University, Raleigh, USA<br />

Abstract<br />

A biparental population <strong>of</strong> 150 recombinant wheat inbred (RI) lines segregating for reaction to <strong>Stagonospora</strong> nodorum<br />

was developed <strong>and</strong> tested in the greenhouse <strong>and</strong> field for levels <strong>of</strong> resistance. The F 3:4 lines were tested in the greenhouse for<br />

incubation period <strong>and</strong> percent leaf infection 5 <strong>and</strong> 10 days after initial symptoms <strong>of</strong> disease. Plants were tested at both the<br />

juvenile <strong>and</strong> adult stages. Adult plants were also tested for percent disease on the spike. The F 3:5 lines were tested in the field<br />

over four locations for level <strong>of</strong> adult leaf <strong>and</strong> head resistance. Heritability <strong>of</strong> leaf resistance to S. nodorum was calculated to<br />

be 0.56 on an entry means basis <strong>and</strong> 0.17 on a per plot basis. Heritability <strong>of</strong> resistance <strong>of</strong> the spike was calculated to be 0.57 on<br />

an entry means basis <strong>and</strong> 0.20 on a per plot basis. Both heritability estimates are based on field data. Correlations between<br />

field data <strong>and</strong> greenhouse data were generally weak, notably between field data <strong>and</strong> greenhouse performance <strong>of</strong> juvenile plants.<br />

However, resistance <strong>of</strong> the spike in the field <strong>and</strong> the greenhouse were highly correlated (r = 0.74).<br />

Resistance to <strong>Stagonospora</strong><br />

nodorum is an economical way to<br />

control glume blotch as well as a<br />

component <strong>of</strong> integrated<br />

management <strong>of</strong> the disease.<br />

Resistance to S. nodorum is complex<br />

<strong>and</strong> polygenic, particularly in adult<br />

plants. There is some evidence that<br />

resistance in the spike <strong>and</strong> at the<br />

juvenile growth stage are under<br />

different genetic control than adult<br />

leaf resistance (Eyal et al., 1987). By<br />

underst<strong>and</strong>ing the relationships<br />

between juvenile, adult leaf, <strong>and</strong><br />

resistance <strong>of</strong> the spike as well as<br />

field <strong>and</strong> greenhouse resistance,<br />

one could improve progress in<br />

selecting the most resistant lines.<br />

Also, an underst<strong>and</strong>ing <strong>of</strong> the<br />

various components <strong>of</strong> resistance<br />

<strong>and</strong> the epidemiology <strong>of</strong> glume<br />

blotch in a given area would allow<br />

deploying the most useful<br />

component <strong>of</strong> resistance in that<br />

environment.<br />

Materials <strong>and</strong> Methods<br />

A cross <strong>of</strong> the s<strong>of</strong>t red winter<br />

wheat cultivars Caldwell= (CItr<br />

17897) ( mod. susceptible) x Coker<br />

8427 (9766) (PI601429) (mod.<br />

resistant) was made, <strong>and</strong> progeny<br />

were advanced in bulk to the F 3<br />

generation. Individual spikes were<br />

taken from the F 3 generation <strong>and</strong><br />

used to form 150 F 3:4 derived lines.<br />

The F 3:4 lines were tested in the<br />

greenhouse for incubation period<br />

<strong>and</strong> percent leaf area diseased at<br />

both the juvenile <strong>and</strong> adult plant<br />

growth stages. Percent diseased<br />

tissue on the spike was measured<br />

at the adult growth stage.<br />

A test consisted <strong>of</strong> a single plant<br />

from each F 3:4 line using a<br />

completely r<strong>and</strong>omized design. A<br />

total <strong>of</strong> four tests were replicated in<br />

time during the winter <strong>of</strong> 1997-98.<br />

Plants were inoculated with a<br />

mixture <strong>of</strong> three S. nodorum isolates<br />

at a concentration <strong>of</strong> 1.0 x 10 6<br />

conidia/ml plus 0.01% Tween 80.<br />

Inoculations on juvenile plants<br />

were performed at the three-leaf<br />

stage, while adult inoculations<br />

occurred at anthesis. After<br />

inoculation, plants were kept in<br />

high humidity for 72 hours. Plants<br />

were rated each day for incubation<br />

period, <strong>and</strong> then at five <strong>and</strong> ten<br />

days after the incubation period<br />

was determined for each individual<br />

line. Leaf <strong>and</strong> spike ratings in the<br />

greenhouse were recorded as a<br />

percentage <strong>of</strong> the total leaf or head<br />

tissue displaying symptoms. Spike<br />

ratings underwent square root<br />

transformation prior to statistical<br />

analysis.<br />

Field testing was performed<br />

during the 1998-99 growing season<br />

at three locations: Johnston Co.,<br />

Lenoir Co., <strong>and</strong> Ashe Co., North<br />

Carolina, USA. F 3:5 lines were<br />

planted as hillplots, consisting <strong>of</strong> 25<br />

seeds planted together as a group.<br />

Hillplots were planted on a grid,<br />

with 0.3 m separating each plot.<br />

Three repetitions <strong>of</strong> each line were<br />

planted in a r<strong>and</strong>omized complete


lock design within each location.<br />

Dried wheat straw cut the previous<br />

season from a field heavily infected<br />

with S. nodorum was distributed at<br />

each location to induce disease.<br />

Overhead irrigation was used at<br />

the Lenoir <strong>and</strong> Ashe County<br />

locations, but was unavailable at<br />

the Johnston County location.<br />

Heritability estimates were<br />

calculated as described by Fehr<br />

(1991).<br />

Each hillplot was rated for level<br />

<strong>of</strong> disease on the leaf <strong>and</strong> on the<br />

spike. Leaf disease ratings were<br />

done using the double digit scale<br />

(Eyal et al., 1987), indicating both<br />

the highest point <strong>of</strong> disease in a<br />

hillplot <strong>and</strong> the percentage <strong>of</strong><br />

disease at that point. Disease on the<br />

spike was recorded as the<br />

percentage <strong>of</strong> diseased tissue <strong>and</strong><br />

underwent square root<br />

transformation prior to statistical<br />

analysis. Statistical analysis was<br />

performed using SAS release 6.12.<br />

Correlation analysis used the<br />

Spearman ranked test.<br />

Results <strong>and</strong> Discussion<br />

A range <strong>of</strong> values for each<br />

measured trait was produced<br />

among lines in both the greenhouse<br />

<strong>and</strong> in the field (Table 1). The<br />

Comparison <strong>of</strong> Greenhouse <strong>and</strong> Field Levels <strong>of</strong> Resistance to <strong>Stagonospora</strong> nodorum 171<br />

analysis <strong>of</strong> variance demonstrated<br />

significant differences among lines<br />

(p < 0.001) for each trait in both the<br />

greenhouse <strong>and</strong> the field (data not<br />

shown). Using field data,<br />

heritability <strong>of</strong> adult leaf resistance<br />

was calculated <strong>and</strong> found to have a<br />

value <strong>of</strong> 0.56 on an entry means<br />

basis <strong>and</strong> a value <strong>of</strong> 0.17 on a per<br />

hillplot basis. Heritability <strong>of</strong><br />

resistance <strong>of</strong> the spike was 0.57 on<br />

an entry means basis <strong>and</strong> 0.20 on a<br />

per plot basis. These data indicate<br />

the genetic component <strong>of</strong> resistance<br />

in this population has a large effect<br />

on disease expression. Ranking <strong>of</strong><br />

mean resistance scores among lines<br />

was consistent over three locations.<br />

Correlation analysis<br />

demonstrated a negative<br />

correlation between adult<br />

incubation period <strong>and</strong> leaf<br />

resistance in the greenhouse <strong>and</strong><br />

field, as well as a small positive<br />

correlation between adult<br />

greenhouse leaf resistance <strong>and</strong> field<br />

leaf resistance (Table 2). The largest<br />

<strong>and</strong> most significant correlation<br />

was between greenhouse <strong>and</strong> field<br />

resistance <strong>of</strong> the spike (0.74). This<br />

indicates screening for resistance <strong>of</strong><br />

the spike in the greenhouse may<br />

provide an alternative to field<br />

testing. In the greenhouse, juvenile<br />

plant reactions demonstrated little<br />

Table 1. Means, st<strong>and</strong>ard deviations, <strong>and</strong> ranges <strong>of</strong> <strong>Stagonospora</strong> nodorum traits measured in the greenhouse <strong>and</strong> field.<br />

correlation with adult plant<br />

resistance traits either in the<br />

greenhouse or in the field,<br />

indicating a possible separate<br />

genetic mechanism for juvenile<br />

resistance. Resistance <strong>of</strong> the spike<br />

<strong>and</strong> leaf resistance in the field<br />

produced a correlation value <strong>of</strong><br />

0.55, indicating some relationship<br />

between the traits, but a large<br />

degree <strong>of</strong> variation between the<br />

traits remained unexplained.<br />

We are currently using AFLP<br />

<strong>and</strong> RAPD markers in an attempt<br />

to match polymorphisms at the<br />

molecular level with traits<br />

measured in the field <strong>and</strong><br />

greenhouse to gain some insight<br />

into the genetics <strong>of</strong> resistance to<br />

this disease.<br />

References<br />

Eyal, Z., A.L. Scharen, J.M. Prescott,<br />

<strong>and</strong> M. van Ginkel. 1987. The<br />

<strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Concepts <strong>and</strong> Methods <strong>of</strong> Disease<br />

Management. Mexico, D.F.:<br />

<strong>CIMMYT</strong>. pp. 33-35.<br />

Fehr, W.R. 1991. Principles <strong>of</strong> Cultivar<br />

Development. Vol 1. MacMillan<br />

Publishing. USA. p. 99.<br />

Trait Growth Stage <strong>and</strong> Environment Mean Std. Dev. Min. Max.<br />

Incubation period Juvenile, greenhouse 5.8 days 0.94 days 4.0 days 10.0 days<br />

Leaf area disease<br />

after 5 days Juvenile, greenhouse 5.5% 4.7% 0.5% 30.0%<br />

Leaf area disease<br />

after 10 days Juvenile, greenhouse 10.2% 7.8% 0.4% 42.5%<br />

Incubation period Adult, greenhouse 5.4 days 0.86 days 3.5 days 9.5 days<br />

Leaf area disease<br />

after 5 days Adult, greenhouse 8.4% 7.2% 1.0% 48.8%<br />

Leaf area disease<br />

after 10 days Adult, greenhouse 27.8% 19.0% 3.0% 90.0%<br />

Head area disease Adult, greenhouse 3.2% 0.8% 0% 23.0%<br />

Leaf rating Adult, field 72.1% 7.7% 48.0% 88.8%<br />

Head rating Adult, field 25.0% 1.0% 0% 100%


Session 6C — S.L. Walker, S. Leath, <strong>and</strong> J.P. Murphy<br />

172<br />

Table 2. Spearman correlation coefficients between <strong>Stagonospora</strong> nodorum traits measured in the greenhouse <strong>and</strong> field.<br />

Adult a Adult b Adult c Adult d Juv. e Juv. f Juv. g Field h Field i<br />

incub. day 5 day 10 head incub. day 5 day 10 leaf head<br />

Adult incub. 1.0 -0.40*** -0.36*** -0.11 0.04 -0.12 -0.16* -0.26** -0.13<br />

Adult day5 -0.40*** 1.0 0.75*** 0.12 0.00 0.16* 0.20** 0.18* 0.21**<br />

Adult day10 -0.36*** 0.75*** 1.0 0.13 -0.03 0.14 0.18* 0.23** 0.19*<br />

Adult head -0.11 0.12 0.13 1.0 0.11 0.03 0.06 0.48*** 0.74***<br />

Juv. incub. 0.04 0.00 -0.03 0.11 1.0 -0.05 -0.15 -0.08 0.00<br />

Juv. day5 -0.12 0.16* 0.14 0.03 -0.05 1.0 0.69*** 0.10 0.08<br />

Juv. day10 -0.16* 0.20* 0.18* 0.06 -0.15 0.69*** 1.0 0.09 0.08<br />

Field leaf -0.26** 0.18* 0.23** 0.48*** -0.08 0.10 0.09 1.0 0.55***<br />

Field head -0.13 0.21** 0.19* 0.74*** 0.00 0.08 0.08 0.55*** 1.0<br />

*, **, *** p < 0.05, 0.01, 0.0001, respectively.<br />

a. Incubation period <strong>of</strong> adult plants in greenhouse test.<br />

b. Percent leaf area disease <strong>of</strong> adult plants five days after first sign <strong>of</strong> disease in greenhouse test.<br />

c. Percent leaf area disease <strong>of</strong> adult plants ten days after first sign <strong>of</strong> disease in greenhouse test.<br />

d. Percent area head disease <strong>of</strong> adult plants in greenhouse test.<br />

e. Incubation period <strong>of</strong> juvenile plants in greenhouse test.<br />

f. Percent leaf area disease <strong>of</strong> juvenile plants five days after first sign <strong>of</strong> disease in greenhouse test.<br />

g. Percent leaf area disease <strong>of</strong> juvenile plants five days after first sign <strong>of</strong> disease in greenhouse test.<br />

h. Leaf disease rating in field tests.<br />

i. Head disease rating in field tests.


Session 6D: Chemical Control<br />

Adjusting Thresholds for <strong>Septoria</strong> Control in Winter Wheat<br />

Using Strobilurins<br />

L.N. Jørgensen, 1 K.E. Henriksen, 1 <strong>and</strong> G.C. Nielsen2 1 Department <strong>of</strong> Crop Protection, Danish Institute <strong>of</strong> Agricultural Sciences, Slagelse, DK<br />

2 The Danish Agricultural Advisory Centre, Skejby, DK<br />

Abstract<br />

In semi-field trials in spring wheat, azoxystrobin has shown a longer residual <strong>and</strong> preventive effect on <strong>Stagonospora</strong><br />

nodorum than the triazole propiconazole. Three weeks after application, more than 90% control was still obtained with one<br />

quarter <strong>of</strong> the recommended rate <strong>of</strong> azoxystrobin. Propiconazole gave in comparison only 50% control. In another semi-field<br />

trial using <strong>Septoria</strong> tritici, azoxystrobin <strong>and</strong> propiconazole had a similar curative effect, both being significantly better than<br />

chlorothalonil. Different dosages (12-100% <strong>of</strong> normal rate <strong>of</strong> azoxystrobin) were tested in field trials. When taking into<br />

account the different timing, the most pr<strong>of</strong>itable dose has varied between 50 <strong>and</strong> 100% <strong>of</strong> normal rate, depending on growth<br />

stage <strong>and</strong> disease pressure. If optimal timing is used, 25-50% <strong>of</strong> normal rate has generally been sufficient. In 1998 in field<br />

trials azoxystrobin <strong>and</strong> the co-formulation propiconazole + fenpropimorph or tebuconazole provided similar control <strong>of</strong> S.<br />

tritici when applied after 4 or 8 days <strong>of</strong> precipitation, respectively. The test model for azoxystrobin recommended between 36<br />

<strong>and</strong> 45% <strong>of</strong> the normal rates. Azoxystrobin gave an increase in yield <strong>of</strong> 100-600 kg ha -1 above the co-formulation or<br />

tebuconazole. So far azoxystrobin has shown pr<strong>of</strong>itable yields in all trials carried out in Denmark since 1994. However, it is<br />

still not known how many days <strong>of</strong> precipitation are required to make spraying with azoxystrobin pr<strong>of</strong>itable.<br />

Based on historical trial data, it<br />

has been shown that in Denmark an<br />

economically important attack <strong>of</strong><br />

<strong>Septoria</strong> tritici or <strong>Stagonospora</strong><br />

nodorum requires more than 7-8<br />

days <strong>of</strong> precipitation (>1 mm rain)<br />

between GSZ 32 <strong>and</strong> 30 days<br />

thereafter (Hansen et al., 1994). This<br />

model has been incorporated into<br />

the decision support system PC-<br />

Plant Protection <strong>and</strong> is widely used<br />

by Danish farmers as a guide for<br />

septoria control (Secher et al., 1995).<br />

Under validation the model has<br />

given correct answers in 75% <strong>of</strong> the<br />

events studied (Jørgensen et al.,<br />

1999). The experience using<br />

triazoles for control <strong>of</strong> septoria<br />

diseases in Denmark has been that<br />

their application is justified in<br />

approximately half <strong>of</strong> the years.<br />

Adjusting the model for<br />

strobilurins requires knowledge <strong>of</strong><br />

their preventive <strong>and</strong> curative effect,<br />

the residual effect, the dose<br />

response at different timing, <strong>and</strong> the<br />

yield responses measured in relation<br />

to fungicides. Several <strong>of</strong> these<br />

aspects are under investigation<br />

using azoxystrobin as representative<br />

<strong>of</strong> the strobilurins.<br />

Materials <strong>and</strong> Methods<br />

In outdoor semi-field trials,<br />

spring wheat (cv. Dragon) was<br />

grown in 8-liter pots using 20 seeds<br />

per pot <strong>and</strong> 4 replicates. The plants<br />

were exposed to normal weather<br />

conditions (May, June, <strong>and</strong> July).<br />

Two types <strong>of</strong> trials were carried out<br />

using either artificial inoculation <strong>of</strong><br />

S. nodorum or S. tritici at growth<br />

stage 39-45. The plants were<br />

inoculated by spraying each pot<br />

with 25 ml <strong>of</strong> a suspension<br />

containing 2.5 x 10 6 spores ml -1 <strong>of</strong><br />

either S. nodorum or S. tritici. After<br />

inoculation the pots were covered<br />

with polyethylene for two days to<br />

provide high humidity conditions<br />

173<br />

for infection. Plants were treated<br />

with fungicide either after<br />

inoculation to investigate the<br />

curative effect, or before inoculation<br />

to examine the preventive <strong>and</strong><br />

residual effect. Azoxystrobin,<br />

propiconazole, <strong>and</strong> chlorothalonil<br />

were tested at 100-25% <strong>of</strong> the<br />

normal rate using a laboratory pot<br />

sprayer with flat-fan nozzles (Hardi<br />

4110-14) <strong>and</strong> a water volume <strong>of</strong> 167 l<br />

ha -1 . Percent total coverage <strong>of</strong> the<br />

plants by disease as well as percent<br />

severity on individual leaves were<br />

assessed.<br />

Dose response field trials were<br />

carried out by the Danish Institute<br />

<strong>of</strong> Agricultural Sciences (DIAS), <strong>and</strong><br />

the trials testing the septoria model<br />

using either 4 or 8 days <strong>of</strong><br />

precipitation were conducted by<br />

DIAS <strong>and</strong> the Danish Agricultural<br />

Advisory Centre (DAAC). The<br />

design <strong>of</strong> the trials was in the first<br />

case split plot <strong>and</strong> in the second


Session 6D — L.N. Jørgensen, K.E. Henriksen, <strong>and</strong> G.C. Nielsen<br />

174<br />

systematic complete block design<br />

with 5 replicates <strong>and</strong> a plot size <strong>of</strong><br />

20-32 m 2 . The fungicides were<br />

applied with knapsack sprayers at<br />

low pressure (2-3 bar) using flat<br />

fan nozzles in a volume <strong>of</strong> 200-300<br />

l ha -1 . The following products<br />

were tested: Tilt top (125 g<br />

propiconazole + 375 g<br />

fenpropimorph per liter), Folicur<br />

(250 g/l tebuconazole per liter),<br />

<strong>and</strong> Amistar (250 g azoxystrobin<br />

per liter). <strong>Septoria</strong> attacks were<br />

observed in all trials, with S. tritici<br />

being dominant, but mixed<br />

infections with S. nodorum were<br />

also seen. Plots were harvested<br />

with a plot combine harvester,<br />

<strong>and</strong> grain yield was corrected to<br />

15% moisture content. When<br />

calculating net yields (pr<strong>of</strong>it), the<br />

prices used were: Tilt top 375<br />

% control<br />

10<br />

8<br />

6<br />

4<br />

2<br />

- 1 1 1<br />

Days before <strong>and</strong> after application<br />

Figure 1. Curative control <strong>of</strong> <strong>Septoria</strong> tritici using 1 / 2<br />

rate <strong>of</strong> three different fungicides at different timing.<br />

12% attack in untreated.<br />

% control<br />

50<br />

40<br />

30<br />

20<br />

10<br />

vs. 65vs.55-<br />

Chlorothalonil<br />

vs.39- DKkr per liter <strong>and</strong> Folicur 400<br />

DKkr per liter, Amistar 583 dkr per<br />

liter, <strong>and</strong> 60 dkr per application;<br />

grain price, 750 dkr per ton. Danish<br />

fungicide prices include a 33% tax.<br />

Results <strong>and</strong> Discussion<br />

Semi-field trials have shown<br />

that azoxystrobin has a curative<br />

effect on S. tritici similar to<br />

propiconazole <strong>and</strong> much better<br />

than chlorothalonil (Figure 1).<br />

Compared to very effective<br />

triazoles like epoxiconazole, the<br />

effect <strong>of</strong> azoxystrobin is, however,<br />

lower (data not shown).<br />

Azoxystrobin showed a very<br />

effective, long preventive effect<br />

against S. nodorum (Figure 2),<br />

lasting for three weeks or more. In<br />

comparison propiconazole +<br />

Propiconazole<br />

Azoxystrobin<br />

vs. 33-<br />

0<br />

100 75 50 25 12.5<br />

Percent <strong>of</strong> normal dose rate<br />

Figure 3. Percent attack <strong>of</strong> <strong>Septoria</strong> tritici at GS 75<br />

after application <strong>of</strong> different dosages<br />

<strong>of</strong> azoxystroin at different growth stages. 2 trials, 1998.<br />

% control<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

fenpropimorph had only a residual<br />

preventive effect for 10 days. In<br />

field trials the early season effect<br />

(GSZ 31) <strong>of</strong> azoxystrobin has been<br />

found to be inferior to the effect <strong>of</strong><br />

the co-formulation propiconazole +<br />

fenpropimorph, something which<br />

has not been observed following<br />

applications carried out around<br />

heading (Jørgensen <strong>and</strong> Nielsen,<br />

1998).<br />

Field trials using four different<br />

dosages <strong>of</strong> azoxystrobin, along<br />

with several other trial series<br />

(Jørgensen <strong>and</strong> Nielsen, 1998), have<br />

shown that azoxystrobin <strong>and</strong> other<br />

strobilurins may be used at a<br />

reduced dose (Figures 3 <strong>and</strong> 4).<br />

Different dosages, 12-100% <strong>of</strong> the<br />

normal azoxystrobin rate, have<br />

been tested in trials. When taking<br />

Propiconazole<br />

Azoxystrobin<br />

1 6 12 18 24<br />

Spraying day after inoculation<br />

Figure 2. Control <strong>of</strong> <strong>Stagonospora</strong> nodorum<br />

preventatively using 1 / 2 rate <strong>of</strong> two fungicides. 35%<br />

attack in untreated.<br />

t/ha<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

0<br />

vs.39- vs.65- vs.33- vs.55- 100 75 50 25 12.5<br />

Percent <strong>of</strong> normal dose rate<br />

Figure 4. Net yield in winter wheat after one<br />

application af azoxystrobin for control <strong>of</strong> <strong>Septoria</strong><br />

tritici using different dosages <strong>and</strong> timing. 2 trials, 1998.


into account different timing, the<br />

most pr<strong>of</strong>itable dose under severe<br />

disease pressure in 1998 varied<br />

between 50 <strong>and</strong> 100% <strong>of</strong> the normal<br />

rate, depending on growth stage<br />

<strong>and</strong> disease pressure. Using<br />

optimal timing, 50% was sufficient.<br />

When summarizing results from all<br />

trials with azoxystrobin, the most<br />

pr<strong>of</strong>itable treatment is most likely<br />

applied between GSZ 39 <strong>and</strong> 55,<br />

using 25 to 50% <strong>of</strong> the normal dose.<br />

In three seasons, different<br />

thresholds have been investigated<br />

as background for recommending<br />

spraying azoxystrobin for septoria<br />

control. In 1998 the difference<br />

between using a threshold <strong>of</strong> 4 or 8<br />

days <strong>of</strong> precipitation above 1 mm<br />

was minor (Table 1). Similar results<br />

were found in two trials in 1997<br />

(data not shown). In two trials in<br />

1996, four days was the optimal<br />

threshold (Jørgensen et al., 1999).<br />

Mixing azoxystrobin with<br />

Adjusting Thresholds for <strong>Septoria</strong> Control in Winter Wheat Using Strobilurins 175<br />

Table 1. Results from trials using 4 <strong>and</strong> 8 days <strong>of</strong> precipitation as a threshold model for control <strong>of</strong> septoria diseases in winter wheat.<br />

Treatment 9 trials 1999 No. <strong>of</strong> appl. TFI a % septoria GS 69-71 Yield <strong>and</strong> yield inc. (t/ha) Net yield (t/ha)<br />

Untreated 0 0 21 6.67<br />

Tilt top 1.0 GSZ 45 1 1 8 1.05 0.47<br />

PC-P Tilt top 8 days 0.9 0.36 12 0.80 0.55<br />

PC-P Tilt top 4 days 1.4 0.65 7 0.91 0.47<br />

Amistar 1.0 GSZ 45 1 1 7 1.67 0.81<br />

PC-P Amistar 8 days 0.9 0.36 12 1.20 0.85<br />

PC-P Amistar 4 days 1.3 0.43 7 1.19 0.75<br />

PC-P Amistar+Folicur 8 days 0.9 0.44 12 1.26 0.90<br />

PC-P Amistar+Folicur 4 days 1.3 0.52 8 1.18 0.74<br />

LSD95 % septoria<br />

0.35<br />

Treatment 3 trials 1998 No. <strong>of</strong> appl. TFIa Flag leaf 2nd leaf Yield <strong>and</strong> yield inc. (t/ha) Net yield (t/ha)<br />

Untreated 0 0 33.2 80.6 7.40 -<br />

Folicur 1.0 GSZ 45 1 1 2.0 32.5 1.70 1.09<br />

PC-P Folicur 8 days 1 0.41 4.1 43.6 1.16 0.86<br />

PC-P Folicur 4 days 2 0.84 5.6 28.7 1.76 1.15<br />

Amistar 1.0 GSZ 45 1 1 2.8 42.3 2.00 1.14<br />

PC-P Amistar 8 days 1 0.41 5.5 55.7 1.79 1.39<br />

PC-P Amistar 4 days 1.3 0.42 9.2 44.7 1.89 1.46<br />

PC-P Amistar+Folicur 8 days 1 0.5 3.6 43.8 1.88 1.47<br />

PC-P Amistar+Folicur 4 days 1.3 0.5 10.9 35.8 1.91 1.45<br />

LSD95 6.8 10.9 5.3<br />

a TFI = Treatment frequency index.<br />

tebuconazole has given results<br />

similar to azoxystrobin applied alone<br />

at both 4 <strong>and</strong> 8 days. These results<br />

indicate that the curative effect <strong>of</strong><br />

azoxystrobin against S. tritici under<br />

Danish conditions are similar or<br />

better than the triazoles<br />

propiconazole <strong>and</strong> tebuconazole<br />

used on the Danish market.<br />

Based on collected information,<br />

so far the test model for strobilurins<br />

recommends spraying after four<br />

days <strong>of</strong> precipitation starting at GSZ<br />

32. If spraying is carried out before<br />

GSZ 51, 10 days’ protection can be<br />

expected before days <strong>of</strong> precipitation<br />

have to be counted again. If applying<br />

after GSZ 51, 20 days’ effect can be<br />

expected from azoxystrobin, which<br />

in practice means that no further<br />

application is needed. The<br />

recommended rates <strong>of</strong> azoxystrobin<br />

in the model vary between 35 <strong>and</strong><br />

45% <strong>of</strong> the normal rate.<br />

References<br />

Hansen, J.G., Secher, B.J., Jørgensen,<br />

L.N., <strong>and</strong> Welling, B. 1994.<br />

Threshold for control <strong>of</strong> <strong>Septoria</strong><br />

spp. in winter wheat based on<br />

precipitation <strong>and</strong> growth stage.<br />

Plant Pathology 43:183-189.<br />

Jørgensen, L.N., <strong>and</strong> Nielsen, G.C.<br />

1998. Reduced dosages <strong>of</strong><br />

strobilurins for disease<br />

management in winter wheat.<br />

Brighton Crop Protection<br />

Conference. Pest <strong>and</strong> <strong>Diseases</strong> 993-<br />

998.<br />

Jørgensen, L.N.; B.J.M. Secher, <strong>and</strong> H.<br />

Hossy. 1999. Decision support<br />

systems featuring <strong>Septoria</strong><br />

Management. In: <strong>Septoria</strong> on<br />

<strong>Cereals</strong>: a Study <strong>of</strong> Pathosystems.<br />

J.A. Lucas, P. Bowyer, <strong>and</strong> H. M.<br />

Anderson (eds.). CABI Publishing,<br />

Wallingford, UK. pp. 251-262.<br />

Secher, B.J., Jørgensen, L.N., Murali,<br />

N.S., <strong>and</strong> Boll, P.S. 1995. Field<br />

validation <strong>of</strong> a decision support<br />

system for control <strong>of</strong> pests <strong>and</strong><br />

diseases in cereals in Denmark.<br />

Pesticide Science 45:195-199.


176


Concluding Remarks<br />

The <strong>Septoria</strong>/<strong>Stagonospora</strong> Blotch <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Past, Present, <strong>and</strong> Future<br />

Z. Eyal (paper presented by A.L. Scharen)<br />

Department <strong>of</strong> Plant Sciences <strong>and</strong> the Institute for Cereal Crops Improvement (ICCI), Tel Aviv University, Tel Aviv,<br />

Israel<br />

Abstract<br />

<strong>Septoria</strong> tritici <strong>and</strong> stagonospora nodorum blotch <strong>of</strong> wheat are regarded as major diseases because <strong>of</strong> their impact on crop<br />

management <strong>and</strong> wheat production. Both pathogens mainly affect the crop’s grain filling processes. Early research dealt<br />

mostly with genetic, cultural, <strong>and</strong> chemical control measures. Chemical control remains one <strong>of</strong> the major means <strong>of</strong> protecting<br />

wheat production, mostly through the use <strong>of</strong> new families <strong>of</strong> systemic fungicides. Emphasis has been given to epidemiological<br />

studies, in many cases associated with chemical control; however, in recent years epidemiological studies have exp<strong>and</strong>ed to<br />

include the effect <strong>of</strong> primary inoculum initiated from dispersal <strong>of</strong> ascospores <strong>of</strong> both pathogens. That shift in focus has<br />

introduced issues such as disease cycle, mating types, <strong>and</strong> their effect—together with pathogenicity—on each pathogen’s<br />

population structure. Recent findings attributed population structure to the pathogens’ sexual rather than asexual stage. The<br />

association between virulence spectrum <strong>and</strong> pathogenicity <strong>and</strong> the contribution <strong>of</strong> the sexual stage remains to be investigated.<br />

The impact <strong>of</strong> such studies on population structure may dictate adapting the proper breeding strategies. A population that is<br />

ever changing due to recombination <strong>and</strong> gene flow may influence the use <strong>of</strong> sources for specific resistance. As for the<br />

introduction <strong>of</strong> DNA technology into <strong>Septoria</strong>/<strong>Stagonospora</strong>: wheat so far has been able to identify population genetic<br />

trends in both pathogens, but they have yet to be linked to virulence patterns. The use <strong>of</strong> genetic transformation in both<br />

pathogens makes it possible to follow post-penetration processes by following reporter genes. This may reveal that<br />

synchronous events associated with the disease cycle, such as the building <strong>of</strong> fungal biomes <strong>and</strong> the initiation <strong>of</strong> picnidia<br />

formation, are all under the control <strong>of</strong> host-parasite interaction. Mutations in virulence genes facilitate studying their<br />

function <strong>and</strong> genome mapping. Concurrently resistance genes can be identified, studied, <strong>and</strong> mapped. This is an exciting era<br />

for <strong>Septoria</strong> tritici/<strong>Stagonospora</strong> nodorum x wheat interaction.<br />

The septoria/stagonospora<br />

blotch diseases <strong>of</strong> wheat are incited<br />

by <strong>Septoria</strong> tritici Roberge in<br />

Desmaz. (teleomorph:<br />

Mycosphaerella graminicola (Fückel)<br />

J. Schrot. in Cohn) <strong>and</strong> by<br />

<strong>Stagonospora</strong> nodorum (Berk) E.<br />

Castellani <strong>and</strong> E. G. Germano<br />

(teleomorph: Phaeosphaeria nodorum<br />

(E. Müller) Hedjaroude),<br />

respectively. The two pathogens<br />

cause major foliar diseases <strong>of</strong><br />

wheat, inflicting considerable yield<br />

losses in many countries<br />

worldwide (King et al., 1983; Pnini-<br />

Cohen et al., 1998). It is <strong>of</strong> interest<br />

to note that these two pathogens<br />

are rare in many rice-wheat<br />

management systems in southeast<br />

Asia or others in Africa.<br />

The increased economic<br />

importance <strong>of</strong> these pathogens,<br />

especially <strong>of</strong> S. tritici, can be<br />

attributed to the cultivation <strong>of</strong><br />

susceptible cultivars with a<br />

concomitant enhancement <strong>of</strong><br />

resistance to other foliar pathogens<br />

(e.g., rusts, powdery mildew),<br />

predominance <strong>of</strong> wheat in crop<br />

management systems usually<br />

characterized by poor management<br />

<strong>of</strong> crop residues, increased nitrogen<br />

fertilization, high summer rainfall,<br />

earlier sowing, <strong>and</strong> resistance to<br />

MBC fungicides (Eyal, 1999).<br />

A surge in scientific research on<br />

these two pathogens occurred in<br />

the early 1980s, when a wide range<br />

<strong>of</strong> scientific data were published in<br />

177<br />

control-related disciplines (mostly<br />

chemical control <strong>and</strong><br />

epidemiology) <strong>and</strong> less on<br />

biological <strong>and</strong> genetic aspects<br />

associated with the pathogens<br />

(Eyal, 1999). Breeding for disease<br />

resistance lagged due to scarcity <strong>of</strong><br />

resistant germplasm <strong>and</strong> poor<br />

underst<strong>and</strong>ing <strong>of</strong> host-pathogen<br />

relationships <strong>and</strong> <strong>of</strong> how to<br />

manipulate resistance sources in<br />

breeding schemes.<br />

The shift in importance from<br />

stagonospora nodorum blotch to<br />

septoria tritici blotch (Pnini-Cohen<br />

et al., 1998) that occurred in certain<br />

European countries was explained<br />

by the biological differences in their<br />

disease cycles (shorter for S. tritici)


Concluding Remarks — Z. Eyal<br />

178<br />

<strong>and</strong> possibly due to differential<br />

sensitivity <strong>of</strong> the two pathogens to<br />

commercially used fungicides<br />

(Shaw, 1999). These explanations—<br />

which may reflect the situation in<br />

certain Western European<br />

countries—may not hold true in<br />

other wheat management systems.<br />

Still, the verification <strong>of</strong> the<br />

mechanism(s) affecting such shifts<br />

is extremely important, since it may<br />

dictate adoption <strong>of</strong> different control<br />

strategies.<br />

The sources <strong>of</strong> primary<br />

inoculum are rather variable:<br />

airborne ascospores,<br />

pycnidiospores from plant refuse,<br />

wild grasses, <strong>and</strong>, possibly, infested<br />

(infected?) seeds (Brokenshire,<br />

1975). Several wild grass species<br />

can serve as sources <strong>of</strong> inoculum to<br />

cultivated wheat, <strong>and</strong> some <strong>of</strong> them<br />

can contribute to speciation in<br />

pathogen populations. Artificial<br />

inoculation <strong>of</strong> wheat seedlings by<br />

isolates secured from grasses may<br />

not provide information on their<br />

contribution under natural<br />

infection. The possible contribution<br />

<strong>of</strong> infected seeds to the onset <strong>of</strong><br />

disease was confirmed for S.<br />

nodorum, but is not well established<br />

for S. tritici. The current world<br />

trade <strong>of</strong> grain <strong>and</strong>, for that matter,<br />

<strong>of</strong> infested/infected grain provides<br />

ample opportunities for the<br />

dissemination <strong>of</strong> inoculum<br />

worldwide if such a distribution<br />

system is operative.<br />

Global virulence surveys<br />

conducted by Eyal et al. (1985) <strong>and</strong><br />

Kema et al. (1996) have shown<br />

considerable diversity at the<br />

species level (Triticum spp.) <strong>and</strong><br />

differential response on the selected<br />

“differential” set <strong>of</strong> wheat cultivars.<br />

McDonald et al. (1999) stated that<br />

“since it has been shown that M.<br />

graminicola can infect seed<br />

(Brokenshire,1975), we consider it<br />

likely that this is also the<br />

mechanism for long distance gene<br />

flow in M. graminicola.” The<br />

authors quoted King et al. (1983)<br />

that in the case <strong>of</strong> S. nodorum, the<br />

most likely mechanism for<br />

intercontinental dispersal is<br />

infected seed. It should be noted<br />

that Brokenshire (1975) stated that<br />

“infection <strong>of</strong> seedlings from<br />

infected untreated seed samples<br />

has proved unsuccessful with<br />

Triticum dicoccum.” The report on<br />

seed infection <strong>of</strong> S. tritici by<br />

Brokenshire (1975) was not further<br />

verified <strong>and</strong> thus introduced<br />

ambiguity in the underst<strong>and</strong>ing <strong>of</strong><br />

the possible dissemination <strong>of</strong> this<br />

pathogen by seeds. The<br />

epidemiological implications <strong>of</strong><br />

naturally-infected grasses for shortscale<br />

dissemination <strong>and</strong> seeds<br />

(especially for S. tritici) for longdistance<br />

transport, warrant<br />

detailed investigations.<br />

The supposition that<br />

population genetics <strong>of</strong> M.<br />

graminicola using single-locus<br />

probes to measure gene diversity<br />

for individual RFLP loci ,<br />

population subdivision, <strong>and</strong><br />

genetic similarity among<br />

populations (McDonald et al., 1999)<br />

can be superimposed on virulence<br />

patterns was not supported in<br />

studies on populations from<br />

California <strong>and</strong> Oregon (Mundt et<br />

al., 1999). Isolates <strong>of</strong> S. tritici from<br />

these states differed significantly in<br />

their pathogenicity (Ahmed et al.,<br />

1996), yet they manifested<br />

similarity in allele frequencies as<br />

measured by RFLPs (Chen <strong>and</strong><br />

McDonald, 1996). The finding that<br />

the genetic diversity measured<br />

among 122 isolates collected in an<br />

S. tritici nursery at Patzcuaro,<br />

Mexico, was among the lowest,<br />

does not imply that the pathogen<br />

has a narrow virulence pattern. The<br />

Patzcuaro S. tritici population was<br />

reported to possess virulence to<br />

some <strong>of</strong> the most resistant wheat<br />

cultivars (e.g., Bobwhite”S” <strong>and</strong><br />

Kavkaz/K4500 L.6.A.4) (McDonald<br />

et al., 1999).<br />

Based on genetic population<br />

studies, McDonald et al. (1999)<br />

suggest that “if <strong>CIMMYT</strong> continues<br />

to use Patzcuaro as a field site to<br />

screen for resistance to M.<br />

graminicola <strong>and</strong> S. nodorum, it may<br />

want to consider introducing more<br />

diverse fungal populations from<br />

other parts <strong>of</strong> Mexico into this<br />

disease nursery.” It should be<br />

mentioned that some <strong>of</strong> the testing<br />

sites in the Patzcuaro area were<br />

artificially inoculated in the early<br />

1980s with a mixture <strong>of</strong> Mexican S.<br />

tritici isolates to ensure selection<br />

pressure on breeding materials<br />

(Santiago Fuentes, personal<br />

communication). The suggestion<br />

made by McDonald et al. (1999)<br />

implies that genetic diversity<br />

measured by r<strong>and</strong>om RFLP alleles<br />

can be used as an estimator <strong>of</strong><br />

virulence pattern. It is therefore<br />

proposed that unless genetic<br />

diversity measured by r<strong>and</strong>om<br />

alleles is not highly correlated with<br />

diversity in virulence patterns, the<br />

former should not be used as an<br />

estimator <strong>of</strong> virulence in disease<br />

resistance breeding schemes. The<br />

issue <strong>of</strong> selecting relevant fungal<br />

isolates in screening for disease<br />

resistance will be further discussed.<br />

The quantitative estimation <strong>of</strong><br />

host response percent (pycinidia<br />

<strong>and</strong>/or necrosis) for both S. tritici<br />

<strong>and</strong> S. nodorum <strong>and</strong> thereafter<br />

analyses <strong>of</strong> the interaction term <strong>of</strong><br />

cultivar x isolates by statistical<br />

means (ANOVA) have been used to<br />

estimate genetic variation in these<br />

pathosystems (Eyal <strong>and</strong> Levy,<br />

1987). Such a statistical measure


equires a predetermined set <strong>of</strong><br />

“differentiating cultivars,” which<br />

was not agreed to by <strong>Septoria</strong> tritici/<br />

<strong>Stagonospora</strong> nodorum workers,<br />

suitable methodology (sampling,<br />

culturing <strong>of</strong> pathogen, inoculum<br />

age <strong>and</strong> dosage, incubation<br />

categories, disease assessment,<br />

etc.), <strong>and</strong> controlled environmental<br />

conditions to ensure reliability <strong>of</strong><br />

results. The lack <strong>of</strong> a uniform<br />

methological approach <strong>and</strong>, in<br />

particular, <strong>of</strong> a common <strong>and</strong><br />

agreeable set <strong>of</strong> differentials<br />

introduces difficulties in comparing<br />

<strong>and</strong> drawing conclusions on a<br />

wider basis.<br />

The suggested speciation <strong>of</strong> M.<br />

graminicola on durum (Triticum<br />

durum) <strong>and</strong> common wheat (T.<br />

aestivum) may require the inclusion<br />

<strong>of</strong> differentiating cultivars <strong>of</strong> both<br />

species when they are grown<br />

together or in studies where both<br />

are tested (Kema et al., 1996;<br />

Saadaoui, 1987). Isolates from one<br />

species may not be pathogenic to<br />

the other Triticum species; still, they<br />

may show a significant interaction<br />

term when a proper set <strong>of</strong> cultivars<br />

<strong>of</strong> a species is being inoculated<br />

with isolates from the same species.<br />

Cross infection <strong>of</strong> the two species<br />

was reported for isolates secured<br />

from either Triticum species (Eyal,<br />

1999; Kema et al., 1996). The<br />

specific pathogenicity towards T.<br />

durum may have bearing on the<br />

structure <strong>of</strong> S. tritici populations<br />

(Saadaoui, 1987). Alternate<br />

cropping <strong>of</strong> bread wheat cultivars<br />

in a durum wheat management<br />

system may express a lower than<br />

expected infection level on the<br />

former than under uninterrupted<br />

bread wheat management.<br />

Today it is generally accepted<br />

that the specificity in virulence is<br />

low in S. nodorum, but can be<br />

The <strong>Septoria</strong>/<strong>Stagonospora</strong> Blotch <strong>Diseases</strong> <strong>of</strong> Wheat: Past, Present, <strong>and</strong> Future 179<br />

detected in S. tritici populations<br />

provided proper differentiating<br />

germplasm is used (Eyal, 1995).<br />

The magnitude <strong>of</strong> specificity in the<br />

S. tritici - wheat pathosystem <strong>and</strong><br />

its implications for breeding for<br />

disease resistance requires<br />

elaboration. Few dominant single<br />

genes conferring resistance were<br />

identified (e.g. Stb 1 - Bulgaria 88,<br />

Stb 2 - Veranopolis; Stb 4 - Tadinia)<br />

(Somasco et al., 1996). There are<br />

several reports that these resistant<br />

sources are not providing the<br />

claimed protection when moved<br />

across S. tritici populations with<br />

wide virulence patterns (Ballantyne<br />

<strong>and</strong> Thomson, 1995; Eyal, 1999).<br />

It is proposed that specificity<br />

<strong>and</strong> inheritance studies should be<br />

conducted with germplasm that is<br />

agronomically relevant to breeding,<br />

preferably exhibiting resistance to a<br />

wide virulence pattern (such as<br />

“Bobwhite“S”, IAS20-IASSUL, or<br />

other Frontana derivatives,<br />

Kavkaz/K4500 L.6.A.4, <strong>and</strong> other<br />

sources identified through<br />

multilocation testing) (Eyal et al.,<br />

1987). Special attention should be<br />

given to resistant wheat accessions<br />

developed in wide-cross programs.<br />

Some <strong>of</strong> these accessions may be<br />

susceptible to other foliar diseases<br />

that may need attention when<br />

incorporating <strong>Septoria</strong>/<strong>Stagonospora</strong><br />

resistance. Germplasm used in<br />

virulence studies is usually derived<br />

from disease nurseries naturally or<br />

artificially infected with the<br />

pathogen <strong>of</strong> interest. Only in a few<br />

cases has the virulence spectrum <strong>of</strong><br />

these populations been categorized.<br />

It is therefore likely that<br />

identified resistant germplasm may<br />

have only limited use either as a<br />

breeding source or as<br />

“differentials.” Resistant<br />

germplasm that has withstood<br />

prolonged testing to variable<br />

pathogen populations can be<br />

considered a potential source for<br />

breeding for resistance. Special<br />

attention should be given to avoid<br />

selecting tall stature <strong>and</strong> late<br />

maturing germplasm that may<br />

introduce such non-genetic factors<br />

into germplasm evaluation, or<br />

germplasm with poor agronomic<br />

characteristics.<br />

Isolates possessing virulence to<br />

important resistance sources can<br />

become the “core virulence<br />

spectrum” for screening<br />

germplasm. There is no criterion as<br />

yet for selecting “relevant” isolates<br />

in artificially inoculated breeding<br />

trials. The criteria dictating the<br />

choice <strong>of</strong> isolates for genetic studies<br />

<strong>of</strong> virulence (Kema et al., 1999) may<br />

use considerations other than<br />

breeding. The choice by Kema et al.<br />

(1999) <strong>of</strong> S. tritici isolates such as<br />

IPO323 (avirulent on Veranopolis,<br />

Kavkaz, <strong>and</strong> Shafir) <strong>and</strong> IP094269,<br />

which is virulent on these cultivars,<br />

in studying the genetics <strong>of</strong><br />

avirulence in this pathogen merits<br />

adoption by other <strong>Septoria</strong> tritici /<br />

<strong>Stagonospora</strong> nodorum investigators.<br />

The correlation between<br />

seedling <strong>and</strong> adult host response<br />

has been substantiated by several<br />

investigators (Eyal, 1999; Kema <strong>and</strong><br />

van Silfhout, 1997). Screening for<br />

resistance at the seedling stage<br />

does not provide an integral view<br />

<strong>of</strong> the tested germplasm. Seedling<br />

tests can serve as a supplementary<br />

measure <strong>and</strong> are an excellent tool<br />

for detailed, controlled studies on a<br />

multitude <strong>of</strong> biological issues<br />

associated with host-pathogen<br />

interactions. The seedling test as a<br />

screening measure is hampered by<br />

not knowing whether the used<br />

isolates are relevant to the<br />

virulence spectrum <strong>of</strong> the


Concluding Remarks — Z. Eyal<br />

180<br />

populations to which the test<br />

germplasm will be subjected under<br />

field conditions.<br />

Cross testing <strong>of</strong> germplasm to<br />

both seedling <strong>and</strong> field conditions<br />

is usually performed with<br />

predetermined isolate(s) that can<br />

serve as a “basic virulence set”<br />

(Eyal, 1999). Its composition may<br />

change by introducing isolates<br />

found to be virulent on specific<br />

resistant germplasm, or omitting<br />

others. The set should include<br />

isolate(s) whose virulence on<br />

specific accessions has been<br />

repeatedly confirmed. The<br />

multilocation <strong>Septoria</strong> Monitoring<br />

Nursery initiated by <strong>CIMMYT</strong> can<br />

provide information as to the<br />

magnitude <strong>of</strong> protection <strong>of</strong> some <strong>of</strong><br />

the identified sources <strong>and</strong><br />

contribute information on<br />

worldwide virulence patterns.<br />

The contribution <strong>of</strong> the sexual<br />

stage to the virulence spectrum<br />

needs special attention. Evidence<br />

from studies on population<br />

genetics emphasizes the major<br />

influence <strong>of</strong> the sexual stage on the<br />

genetic diversity <strong>of</strong> the population<br />

(Chen <strong>and</strong> McDonald, 1986; Eyal<br />

<strong>and</strong> Levy, 1987; Eyal et al., 1985;<br />

Jlibene et al., 1994; McDonald et al.,<br />

1999; Zhan et al., 1998). McDonald<br />

et al. (1999) stated that asexual<br />

reproduction may have an<br />

important impact over a limited<br />

area (e.g. few square meters),<br />

whereas sexual reproduction has<br />

much greater consequences for<br />

population genetics.<br />

The limited scope <strong>of</strong> clonality in<br />

M. graminicola <strong>and</strong> P. nodorum<br />

populations strengthens the need<br />

for comparative studies on<br />

virulence. The limitation <strong>of</strong> such<br />

studies resides with selected<br />

“wheat differentials.” This<br />

difficulty can be partly overcome<br />

by tracing specific virulences<br />

within the pathogen population on<br />

certain resistance accessions (e.g.<br />

Kavkaz/K4500 L.6.A.4). Low<br />

frequency <strong>of</strong> virulence to Kavkaz/<br />

K4500 L.6.A.4 was reported for the<br />

first time in Israel (Ezrati et al.,<br />

1998), interestingly, in Nahal Oz,<br />

where McDonald et al. (1999)<br />

detected the greatest genetic<br />

diversity using r<strong>and</strong>om RFLP<br />

alleles. Low frequency <strong>of</strong> virulence<br />

on Kavkaz/K4500 L.6.A.4 was<br />

reported in the global virulence<br />

surveys conducted by Eyal et al.<br />

(1995) <strong>and</strong> Kema et al. (1996). This<br />

germplasm was not used in<br />

international <strong>and</strong> Israeli breeding<br />

programs in the past <strong>and</strong> therefore<br />

no selective advantage can explain<br />

its detection. It is expected that the<br />

presence <strong>of</strong> virulence on this<br />

accession at Nahal Oz can be<br />

accounted for by the fact that the<br />

sexual stage is operative there, as<br />

implied by McDonald et al. (1999).<br />

Chen <strong>and</strong> McDonald (1996)<br />

hypothesized that M. graminicola<br />

isolates can be produced via<br />

genetic recombination with the<br />

same combination <strong>of</strong> virulence<br />

genes that may not have the same<br />

recent ancestors, thus contributing<br />

to multiple clonal lineages in the<br />

population. The distribution <strong>of</strong> this<br />

virulence in space (locations) <strong>and</strong><br />

time (years) is under study.<br />

The expression <strong>of</strong> virulence on<br />

certain wheat germplasm was<br />

reported to be altered upon<br />

inoculation with mixtures <strong>of</strong> S.<br />

tritici isolates (Ezrati et al., 1998;<br />

Zelikovitch <strong>and</strong> Eyal, 1991). When<br />

a resistant cultivar (e.g., Seri 82) is<br />

inoculated with a mixture <strong>of</strong><br />

avirulent (ISR398) <strong>and</strong> virulent<br />

(ISR8036) isolates, the level <strong>of</strong><br />

pycnidia produced on the seedlings<br />

<strong>and</strong> adult plants in the field<br />

resembles that produced by the<br />

avirulent isolate inoculated singly<br />

(Ezrati et al., 1998). The virulent<br />

isolate ISR8036 predominated in<br />

the population <strong>of</strong> pycnidia on both<br />

seedlings <strong>and</strong> adult plants <strong>of</strong> the<br />

susceptible cultivar Shafir<br />

inoculated with a 1:1 mixture <strong>of</strong> the<br />

two isolates. It was suggested that<br />

isolate ISR8036, which induced the<br />

same level <strong>of</strong> pycnidia on Shafir as<br />

ISR398, was more aggressive than<br />

the latter, giving it a competitive<br />

advantage in the population.<br />

The suppression phenomenon<br />

was also operative when an array<br />

<strong>of</strong> resistant cultivars (e.g.<br />

Bobwhite”S”, IAS 20-IASSUL, <strong>and</strong><br />

Kavkaz/K4500 L.6.A.4) were<br />

inoculated with appropriate<br />

mixtures <strong>of</strong> avirulent <strong>and</strong> virulent<br />

isolates (Ezrati et al., 1998). These<br />

findings are indicative <strong>of</strong> the<br />

commonality <strong>of</strong> the phenomenon<br />

where resistance induced by an<br />

avirulent S. tritici isolate can<br />

provide tissue protection against<br />

the virulent isolate. The level <strong>of</strong><br />

protection, the conditions under<br />

which it is expressed, <strong>and</strong> the<br />

mechanism(s) associated with the<br />

induction needs further<br />

investigation.<br />

It is possible that under field<br />

conditions, induced resistance may<br />

be operative under low to<br />

moderate S. tritici epidemics,<br />

provided a certain level <strong>of</strong> genetic<br />

resistance is present in the infected<br />

wheat cultivar. The reported<br />

differences in aggressiveness<br />

among isolates (Ahmed et al., 1996;<br />

Ezrati et al., 1998; Mundt et al.,<br />

1999) suggest that the structure <strong>of</strong><br />

an asexual population<br />

progressively developed on wheat<br />

plants can be strongly affected by<br />

both virulence <strong>and</strong> aggressiveness.<br />

If the sexual stage is the sole


provider <strong>of</strong> primary inoculum from<br />

distant sources (Shaw <strong>and</strong> Royle,<br />

1989) or from within the crop<br />

(Gilchrist <strong>and</strong> Velazquez, 1994), the<br />

structure <strong>of</strong> the population will<br />

probably change annually.<br />

The vertical progression within<br />

the crop <strong>of</strong> pseudothecia from<br />

lower to upper leaves (Gilchrist<br />

<strong>and</strong> Velazquez, 1994) may give an<br />

advantage to “local” populations,<br />

though it does not exclude the<br />

possibility that the primary<br />

inoculum may come from a<br />

distance <strong>and</strong> then perpetuate itself<br />

within the crop. Hunter et al. (1999)<br />

stressed that the continuous<br />

development <strong>of</strong> a functional sexual<br />

stage suggests that there is genetic<br />

exchange throughout the growing<br />

season that can respond over time<br />

to selection pressure exerted by<br />

resistant germplasm. The role <strong>of</strong><br />

the asexual stage in epidemics <strong>and</strong><br />

in formulating the virulence<br />

structure <strong>of</strong> the population under<br />

such conditions is not clear.<br />

The genetic variation for<br />

virulence <strong>and</strong> resistance in the<br />

wheat-S. tritici pathosystem, in<br />

addition to having implications for<br />

breeding for disease resistance,<br />

provides an opportunity to exp<strong>and</strong><br />

our knowledge on the biology <strong>and</strong><br />

genetics <strong>of</strong> the interaction. The<br />

recognized specificity in the<br />

relationship linked with the<br />

implementation <strong>of</strong> new DNA<br />

technology allows for a more<br />

thorough underst<strong>and</strong>ing <strong>of</strong> the<br />

pathogen, the host, <strong>and</strong> the<br />

interaction. This makes it possible<br />

to adopt methodologies from other<br />

host-pathogen systems <strong>and</strong> apply<br />

them to the economically important<br />

wheat-S. tritici/S. nodorum<br />

pathosystems.<br />

The <strong>Septoria</strong>/<strong>Stagonospora</strong> Blotch <strong>Diseases</strong> <strong>of</strong> Wheat: Past, Present, <strong>and</strong> Future 181<br />

Issues associated with the<br />

ability to infect wheat <strong>and</strong><br />

processes related to the infectioncycle<br />

(biochemical, involvement <strong>of</strong><br />

toxins, tissue necrosis, induction <strong>of</strong><br />

pycnidia formation, <strong>and</strong><br />

suppression <strong>of</strong> production),<br />

virulence (genotypic diversity <strong>and</strong><br />

population structure), pathogen<br />

migration, the potential <strong>of</strong> genetic<br />

recombination on pathogen<br />

structure <strong>and</strong> then on disease<br />

management, <strong>and</strong> genetic host<br />

resistance (specific, non-specific)<br />

are currently being investigated<br />

with the aid <strong>of</strong> molecular tools<br />

(Caten, 1999; Eyal, 1999; Kema et<br />

al., 1999; Baker et al., 1997; Eyal et<br />

al., 1985). Genetic diversity <strong>and</strong><br />

population structure are being<br />

revealed with the aid <strong>of</strong> RFLP,<br />

RAPD probes, <strong>and</strong> DNA<br />

fingerprinting (McDonald et al.,<br />

1999).<br />

The issue <strong>of</strong> mating types <strong>and</strong><br />

possible sequencing <strong>of</strong> genes<br />

associated with these types, <strong>and</strong><br />

mapping <strong>of</strong> the M. graminicola <strong>and</strong><br />

P. nodorum genomes are being<br />

investigated by Caten (1999) <strong>and</strong><br />

Kema et al. (1999). Genetically<br />

transformed S. tritici isolates <strong>and</strong><br />

tagged mutants with altered<br />

virulence are being investigated by<br />

Pnini-Cohen et al. (1998). The use<br />

<strong>of</strong> reporter gene(s) (e.g. GUS, GFP)<br />

in genetically transformed S. tritici<br />

<strong>and</strong> S. nodorum isolates <strong>and</strong><br />

immunological assays can greatly<br />

contribute to the underst<strong>and</strong>ing <strong>of</strong><br />

in planta qualitative <strong>and</strong><br />

quantitative post-inoculation<br />

events. The interrelations between<br />

isolates in a population <strong>and</strong> the<br />

effect <strong>of</strong> induced resistance <strong>and</strong><br />

competition on a pathogen<br />

population are elucidated with the<br />

aid <strong>of</strong> isolate-specific PCR primers<br />

(Ezrati et al., 1998). The elucidation<br />

<strong>of</strong> loci associated with host<br />

resistance employ QTL analyses,<br />

<strong>and</strong> genomic analysis may some<br />

day contribute to the identification<br />

<strong>of</strong> sequences linked to resistance.<br />

It is likely that genes associated<br />

with resistance to S. tritici <strong>and</strong> S.<br />

nodorum share products with<br />

structural similarities to other plant<br />

defense systems (Baker et al., 1997).<br />

The structural features shared by<br />

several R (resistance) gene products<br />

are a leucine-rich repeat (LRR)<br />

motif or a serine-threonine kinase<br />

domain. Some <strong>of</strong> these genes<br />

encode cytoplasmic receptor-like<br />

proteins that contain an LRR<br />

domain <strong>and</strong> a nucleotide binding<br />

site (NBS). It is therefore possible<br />

that these gene products are<br />

operative in wheat <strong>and</strong> can be<br />

tagged. The identification <strong>of</strong><br />

specific interactions between<br />

avirulence genes in S. tritici<br />

(AVRmg) <strong>and</strong> genes for resistance<br />

in the wheat plant may pave the<br />

way for the “genetic dissection <strong>of</strong> R<br />

gene-mediated induction <strong>of</strong><br />

hypersensitive (?)” (or suppression<br />

<strong>of</strong> symptoms in isolate mixtures ?)<br />

host defense. This insight can be<br />

used to genetically engineer wheat<br />

cultivars resistant to a broad<br />

spectrum <strong>of</strong> pathogens. It will<br />

require a better underst<strong>and</strong>ing <strong>of</strong><br />

avirulence, specific resistance, <strong>and</strong><br />

host-pathogen interactions <strong>and</strong><br />

their products in those<br />

pathosystems, prior to the<br />

formulation <strong>of</strong> a protection<br />

strategy. As a consequence, the<br />

economic <strong>and</strong> ecological (chemical<br />

protection) impact on wheat<br />

production will be reduced.


Concluding Remarks — Z. Eyal<br />

182<br />

References<br />

Ahmed, H.U., Mundt, C.C., H<strong>of</strong>fer,<br />

M.E., <strong>and</strong> Coakley, S.M. 1996.<br />

Selective influence <strong>of</strong> wheat<br />

cultivars on pathogenicity <strong>of</strong><br />

Mycosphaerella graminicola<br />

(anamorph <strong>Septoria</strong> tritici).<br />

Phytopathology 86:454-458.<br />

Baker, B., Zambryski, P., Staskawicz,<br />

B., <strong>and</strong> Dinesh-Kumar, S.P. 1997.<br />

Signaling in plant-microbe<br />

interactions. Science 276:726-733.<br />

Ballantyne, B., <strong>and</strong> Thomson, F. 1995.<br />

Pathogenic variation in Australian<br />

isolates <strong>of</strong> Mycosphaerella<br />

graminicola. Australian J. <strong>of</strong> Agric.<br />

Res. 46:921-934.<br />

Brokenshire, T. 1975. Wheat seed<br />

infection by <strong>Septoria</strong> tritici.<br />

Transactions <strong>of</strong> the British<br />

Mycological Society 64:331-334.<br />

Caten, C.E. 1999. Molecular genetics<br />

<strong>of</strong> <strong>Stagonospora</strong> <strong>and</strong> <strong>Septoria</strong>. Pages<br />

26-43 in: <strong>Septoria</strong> on <strong>Cereals</strong>: A<br />

Study <strong>of</strong> Pathosystems. J.A. Lucas,<br />

P. Bowyer, <strong>and</strong> H.M. Anderson,<br />

eds. CAB International,<br />

Wallingford, UK.<br />

Chen, R.S., <strong>and</strong> McDonald, B.A. 1996.<br />

Sexual reproduction plays a major<br />

role in the genetic structure <strong>of</strong><br />

populations <strong>of</strong> the fungus<br />

Mycosphaerella graminicola.<br />

Genetics 142:1119-1127.<br />

Eyal, Z. 1995. Virulence in <strong>Septoria</strong><br />

tritici, the causal agent <strong>of</strong> septoria<br />

tritici blotch <strong>of</strong> wheat. Pages 27-33<br />

in: Proceedings <strong>of</strong> a <strong>Septoria</strong> tritici<br />

workshop. L. Gilchrist, M. van<br />

Ginkel, A McNab, <strong>and</strong> G.H.J.<br />

Kema, eds. Mexico, D.F.: <strong>CIMMYT</strong>.<br />

Eyal, Z. 1999. <strong>Septoria</strong> <strong>and</strong><br />

stagonospora diseases <strong>of</strong> cereals: A<br />

comparative prespectives. Pages 1-<br />

25 in: <strong>Septoria</strong> on <strong>Cereals</strong>: A Study<br />

<strong>of</strong> Pathosystems. J.A. Lucas, P.<br />

Bowyer, <strong>and</strong> H.M. Anderson, eds.<br />

CAB International, Wallingford,<br />

UK.<br />

Eyal, Z. 1999. Breeding for disease<br />

resistance to septoria <strong>and</strong><br />

stagonospora diseases <strong>of</strong> wheat.<br />

Pages 332-344 in: <strong>Septoria</strong> on<br />

<strong>Cereals</strong>: A Study <strong>of</strong> Pathosystems.<br />

J.A. Lucas, P. Bowyer, <strong>and</strong> H.M.<br />

Anderson, eds. CAB International,<br />

Wallingford, UK.<br />

Eyal, Z., <strong>and</strong> Levy, E. 1987. Variations<br />

in pathogenicity patterns <strong>of</strong><br />

Mycosphaerella graminicola within<br />

Triticum spp. in Israel. Euphytica<br />

36:237-250.<br />

Eyal, Z., Scharen, A.L., Huffman, M.D.,<br />

<strong>and</strong> Prescott, J.M. 1985. Global<br />

insights into virulence frequencies <strong>of</strong><br />

Mycosphaerella graminicola.<br />

Phytopathology 75:1456-1462.<br />

Eyal, Z., Scharen, A.L., Prescott, J.M.,<br />

<strong>and</strong> van Ginkel, M. 1987. The<br />

<strong>Septoria</strong> <strong>Diseases</strong> <strong>of</strong> Wheat:<br />

Concepts <strong>and</strong> Methods <strong>of</strong> Disease<br />

Management. <strong>CIMMYT</strong>, Mexico,<br />

D.F. 51 pp.<br />

Ezrati, S., Schuster, S., Eshel, A., <strong>and</strong><br />

Eyal, Z. 1998. The interrelations<br />

between isolates <strong>of</strong> <strong>Septoria</strong> tritici<br />

varying in virulence.<br />

Phytopathology 88<br />

(Supplement):S27.<br />

Gilchrist, L., <strong>and</strong> Velazquez, C. 1994.<br />

Interaction to <strong>Septoria</strong> tritici isolate<br />

<strong>of</strong> wheat on adult plant under field<br />

conditions. Pages 111-114 in:<br />

Proceedings <strong>of</strong> the 4 th International<br />

<strong>Septoria</strong> <strong>of</strong> <strong>Cereals</strong> Workshop. E.<br />

Arseniuk, T. Goral, <strong>and</strong> P. Czembor,<br />

eds.IHAR, Radzikow, Pol<strong>and</strong>.<br />

Hunter, T., Coker, R.R., <strong>and</strong> Royle, D.J.<br />

1999. The teleomorph stage,<br />

Mycosphaerella graminicola, in<br />

epidemics <strong>of</strong> septoria tritici blotch<br />

on winter wheat in the UK. Plant<br />

Pathology 48:51-57.<br />

Jlibene, M., Gustafson, J.P., <strong>and</strong><br />

Rajaram, S. 1994. Inheritance <strong>of</strong><br />

resistance to Mycosphaerella<br />

graminicola in hexaploid wheat.<br />

Plant Breeding 112:301-310.<br />

Kema, G.H.J., <strong>and</strong> van Silfhout, C.H.<br />

1997. Genetic variation for virulence<br />

<strong>and</strong> resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem III. Comparative<br />

seedling <strong>and</strong> adult plant<br />

experiments. Phytopathology<br />

87:266-272.<br />

Kema, G.H.J., Sayoud, R., Annone, J.G.,<br />

<strong>and</strong> van Silfhout, C.H. 1996. Genetic<br />

variation for virulence <strong>and</strong><br />

resistance in the wheat-<br />

Mycosphaerella graminicola<br />

pathosystem II. Analysis <strong>of</strong><br />

interactions between pathogen<br />

isolates <strong>and</strong> host cultivars.<br />

Phytopathology 86:213-220.<br />

Kema, G.H.J., Verstappen, E.C.P.,<br />

Waalwijk, C., Bonants, P.J.M., de<br />

Koning, J.R.A., Hagenaar-de Weerdt,<br />

M., Hamza, S., Koeken, J.G.P., <strong>and</strong><br />

van der Lee, T.A.J. 1999. Genetics <strong>of</strong><br />

biological <strong>and</strong> molecular markers in<br />

Mycosphaerella graminicola, the cause<br />

<strong>of</strong> septoria tritici leaf blotch <strong>of</strong><br />

wheat. Pages 161-180 in: <strong>Septoria</strong> on<br />

<strong>Cereals</strong>: A Study <strong>of</strong> Pathosystems.<br />

J.A. Lucas, P. Bowyer, <strong>and</strong> H.M.<br />

Anderson, eds. CAB International,<br />

Wallingford, UK.<br />

King, J.E., Cook, R.J., <strong>and</strong> Melville, S.C.<br />

1983. A review <strong>of</strong> septoria diseases<br />

<strong>of</strong> wheat <strong>and</strong> barley. Annals <strong>of</strong><br />

Applied Biology 103:345-373.<br />

McDonald, B.A., Zhan, J., Yarden, O.,<br />

Hogan, K., Garton, J., <strong>and</strong> Pettway,<br />

R.E. 1999. The population genetics<br />

<strong>of</strong> Mycosphaerella graminicola <strong>and</strong><br />

<strong>Stagonospora</strong> nodorum. Pages 44-69<br />

in: <strong>Septoria</strong> on <strong>Cereals</strong>: A Study <strong>of</strong><br />

Pathosystems. J.A. Lucas, P. Bowyer,<br />

<strong>and</strong> H.M. Anderson, eds. CAB<br />

International, Wallingford, UK.<br />

Mundt, C.C., H<strong>of</strong>fer, M.E., Ahmed,<br />

H.U., Coakley, S.M., DiLeone, J.A.,<br />

<strong>and</strong> Cowger, C. 1999. Pages 115-130<br />

in: <strong>Septoria</strong> on <strong>Cereals</strong>: A Study <strong>of</strong><br />

Pathosystems. J.A. Lucas, P. Bowyer,<br />

<strong>and</strong> H.M. Anderson, eds. CAB<br />

International, Wallingford, UK.<br />

Pnini-Cohen, S., Zilberstein, A., <strong>and</strong><br />

Eyal, Z. 1998. Molecular tools for<br />

studying <strong>Septoria</strong> tritici virulence. In:<br />

Proceedings 7 th International<br />

Congress Plant Pathology. 9-16<br />

August, 1998, Edinburgh, Scotl<strong>and</strong>,<br />

UK.<br />

Polley, R.W., <strong>and</strong> Thomas, M.R. 1991.<br />

Surveys <strong>of</strong> diseases <strong>of</strong> winter wheat<br />

in Engl<strong>and</strong> <strong>and</strong> Wales, 1976-1988.<br />

Annals <strong>of</strong> Applied Biology 119:1-20.<br />

Saadaoui, E.M. 1987. Physiologic<br />

specialization <strong>of</strong> <strong>Septoria</strong> tritici in<br />

Morocco. Plant Disease 71:153-155.<br />

Shaw, M.W. 1999. Population dynamics<br />

<strong>of</strong> <strong>Septoria</strong> in the crop ecosystem.<br />

Pages 82-95 in: <strong>Septoria</strong> on <strong>Cereals</strong>:<br />

A Study <strong>of</strong> Pathosystems. J.A.<br />

Lucas, P. Bowyer, <strong>and</strong> H.M.<br />

Anderson, eds. CAB International,<br />

Wallingford, UK.<br />

Shaw, M.W., <strong>and</strong> Royle, D.J. 1989.<br />

Airborne inoculum as a major<br />

source <strong>of</strong> <strong>Septoria</strong> tritici<br />

(Mycosphaerella graminicola)<br />

infections in winter wheat crops in<br />

the UK. Plant Pathology 38:35-43.<br />

Somasco, O.A., Qualset, C.O., <strong>and</strong><br />

Gilchrist, D.G. 1996. Single-gene<br />

resistance to septoria tritici blotch in<br />

the spring wheat cultivar Tadinia.<br />

Plant Breeding 115:261-267.<br />

Zelikovitch, N., <strong>and</strong> Eyal, Z. 1991.<br />

Reduction in pycnidial coverage<br />

after inoculation <strong>of</strong> wheat with<br />

mixtures <strong>of</strong> isolates <strong>of</strong> <strong>Septoria</strong> tritici.<br />

Plant Disease 75:907-910.<br />

Zhan, J., Mundt, C.C., <strong>and</strong> McDonald,<br />

B.A. 1998. Measuring immigration<br />

<strong>and</strong> sexual reproduction in field<br />

populations <strong>of</strong> Mycosphaerella<br />

graminicola. Phytopathology<br />

88:1330-1337.


List <strong>of</strong> Participants<br />

Mr. Shaukat Ali<br />

Graduate Student<br />

North Dakota State University<br />

320 Walster Hall<br />

Fargo ND 58105<br />

Phone: (701) 231-7855<br />

Email: sali@prairie.NoDak.edu<br />

Ms. Lia Arraiano<br />

Phd Student<br />

John Innes Centre<br />

Colney Lane<br />

Norwich Research Park<br />

Norwich Norfolk NR4 7UH<br />

United Kingdom<br />

Phone: +44 (1603) 452571 ext. 2618<br />

Fax: +44 (1603) 502241<br />

Email: lia.arraiano-e-castroalves@bbsrc.ac.uk<br />

Pr<strong>of</strong>essor Edward Arseniuk<br />

Plant Breeding <strong>and</strong> Acclimatization<br />

Institute<br />

Radzikow<br />

05-870 Blonie<br />

Pol<strong>and</strong><br />

Phone: (48-22) 725 4536; (48-22) 349 470<br />

[home]<br />

Fax: (48-22) 725 4714<br />

Email: e.arseniuk@ihar.edu.pl<br />

Dr. Gary C. Bergstrom<br />

Pr<strong>of</strong>essor<br />

Department <strong>of</strong> Plant Pathology<br />

Cornell University<br />

334 Plant Science Building<br />

Ithaca NY 14853-4203<br />

Phone: +1(607) 255-7849<br />

Fax: +1(607) 255-4471<br />

Email: gcb3@cornell.edu<br />

Dr. Penny Brading<br />

Post-Doctoral Scientist<br />

<strong>Cereals</strong> Research Dept.<br />

John Innes Centre<br />

Colney Lane<br />

Norwich Research Park<br />

Norwich Norfolk NR4 7UH<br />

United Kingdom<br />

Phone: +44 (1603) 452571 ext. 2618<br />

Fax: +44 (1603) 502241<br />

Email: penelope.brading@bbsrc.ac.uk<br />

Dr. James K.M. Brown<br />

Pathologist & Geneticist<br />

<strong>Cereals</strong> Research Department<br />

John Innes Centre<br />

Colney Lane<br />

Norwich Norfolk NR4 7UK<br />

United Kingdom<br />

Phone: +44 (1603) 452571<br />

Fax: +44 (1603) 456844<br />

Email: james.brown@bbsrc.ac.uk<br />

Dr. Cristina C. A. Cordo<br />

Facultad de Agronomia<br />

Catedra de Fitopatologia La Plata<br />

Universidad Nacional de la Plata<br />

Calle 60 y 119; c.c. 31<br />

La Plata 1900<br />

Argentina<br />

Phone: +54 0221 4 83 18 31<br />

Fax: +54 0221 4 25 23 46<br />

Email: criscordo@infovia.com.ar<br />

Ms. Christina Cowger<br />

Graduate Research Assistant<br />

Botany Dept.<br />

Oregon State University<br />

2082 Cordley Hall<br />

Corvallis OR 97331-2902<br />

Phone: +1 (541) 737-4408<br />

Fax: +1 (541) 737-3573<br />

Email: cowgerc@bcc.orst.edu<br />

Dr. Barry M. Cunfer<br />

Pr<strong>of</strong>essor<br />

Dept. Plant Pathology<br />

University <strong>of</strong> Georgia<br />

1109 Experiment St.<br />

Griffin GA 30223-1797<br />

United States<br />

Phone: +1(770) 412 4012<br />

Fax: +1(770) 228 7305<br />

Email: bcunfer@gaes.griffin.peachnet.edu<br />

Dr. Pawel Czembor<br />

Research Assistant<br />

Plant Breeding & Acclimatization Inst.<br />

Radzikow<br />

05-870 Blonie<br />

Pol<strong>and</strong><br />

Phone: +48 (22) 7253611<br />

Fax: +48 (22) 7254714<br />

Email: p.czembor@ihar.edu.pl<br />

Pr<strong>of</strong>. Amos Dinoor<br />

Pr<strong>of</strong>essor <strong>of</strong> Plant Pathology<br />

Faculty <strong>of</strong> Agriculture, Food <strong>and</strong><br />

Environmental Quality Sciences<br />

The Hebrew University <strong>of</strong> Jerusalem<br />

P.O. Box 12<br />

Rehovot 76100<br />

Israel<br />

Phone: + 972 (8) 9481358<br />

Fax: +972 (8) 9466794<br />

Email: dinoor@agri.huji.ac.il<br />

Dr. Annika Djurle<br />

Plant Pathology 1<br />

Department <strong>of</strong> Ecology <strong>and</strong> Crop<br />

Production Sciences<br />

SLU<br />

P.O. Box 7043<br />

SE-75007 UPPSALA<br />

Sweden<br />

Phone: +46 (18) 67 16 02<br />

Fax: +46 (18) 67 28 90<br />

Email: annikad@pinus.slu.se<br />

Mr. Keith E. Duncan<br />

Biologist<br />

DuPont Agricultural Biotechnology<br />

E. I. DuPont de Nemours Co., Inc.<br />

E402 Experimental Station<br />

Wilmington , DE 19880-0402<br />

United States<br />

Phone: 302 695 4298; 302 695<br />

3891(laboratory)<br />

Fax: 302 695 4509<br />

Email: keith.e.duncan@usa.dupont.com<br />

Ms. Clare Marie Ellerbrook<br />

Research Scientist<br />

John Innes Centre<br />

Colney Lane<br />

Norwich UK NR4 7UH<br />

United Kingdom<br />

Phone: +44 (1603) 452571<br />

Fax: +44 (1603) 456844<br />

Email: clare.ellerbrook@bbsrc.ac.uk<br />

Ms. Smadar Ezrati<br />

Ph. D. Student<br />

Department <strong>of</strong> Botany<br />

Tel-Aviv University<br />

Tel Aviv 69978<br />

Israel<br />

Phone: +972 (3) 640 9766<br />

Fax: +972 (3) 640 9380<br />

Email: ezrati@post.tau.ac.il<br />

Dr. Lucy Gilchrist<br />

Pathologist<br />

Wheat Program<br />

<strong>CIMMYT</strong><br />

Lisboa 27, Col. Juárez<br />

Delegación Cuauhtemoc<br />

Apdo. Postal 6-641<br />

06600 Mexico D.F.<br />

Mexico<br />

Phone: +52 5804 2004<br />

Fax: +52 5804 7558/9<br />

Email: l.gilchrist@cgiar.org<br />

183<br />

Ing. Rebeca Margarita Gonzalez Iñiguez<br />

Investigador de Trigo y Triticale<br />

Instituto Nacional de Investigaciones<br />

Forestales y Agropecuarias<br />

Silvestre Guerrero 449, Colonia 5 de<br />

Diciembre<br />

58280 Morelia Mich.<br />

Mexico<br />

Phone: (43) 15 9489 y 15 9021<br />

Fax: (43) 151091<br />

Dr. Stephen B. Goodwin<br />

Plant Pathologist<br />

Department <strong>of</strong> Botany <strong>and</strong> Plant Pathology<br />

USDA/ARS Purdue University<br />

1155 Lily Hall<br />

West Lafayette IN 47907-1155<br />

United States<br />

Phone: +1(765) 494-4635<br />

Fax: +1(765) 494-0363<br />

Email: goodwin@btny.purdue.edu


184<br />

Dr. Patrice Halama<br />

Pr<strong>of</strong>essor<br />

Institut Superieur d’Agriculture<br />

41 rue du port<br />

59046 Lille cedex<br />

France<br />

Phone: +33 (03) 28 38 48 48<br />

Fax: +33 (03) 28 38 48 47<br />

Email: p.halama@isa.fupl.asso.fr<br />

Dr. Sonia Hamza<br />

Associate Pr<strong>of</strong>essor<br />

Laboratoire de Genetique<br />

INAT<br />

43 Av. Chrales Nicolle<br />

1082 El Mahrajene Tunis<br />

Tunisia<br />

Phone: +216 (1) 840 270<br />

Fax: +216 (1) 799 391<br />

Email: hamza.sonia@inat.agrinet.tn<br />

Dr. Karen K. Hanson<br />

Cereal Pathology Specialist<br />

Plant Pathology<br />

Zeneca Agrochemicals<br />

Jealott’s Hill Research Station<br />

RG42 6ET Bracknell Berishire<br />

Ukraine<br />

Phone: +44 (01344) 414488<br />

Fax: +44 (01344) 414502<br />

Email:<br />

Karen.Hanson@AGUK.ZECECA.com<br />

Pr<strong>of</strong>. Mouncef Harrabi<br />

Pr<strong>of</strong>essor <strong>and</strong> Director General<br />

Crop Science Department<br />

INAT<br />

43 Ave. Charles Nicolle<br />

1082 Tunis<br />

Tunisia<br />

Phone: +216 (1) 840270<br />

Fax: +216 (1) 799391<br />

Email: harrabi.moncef@inat.agrinet.tn<br />

Dr. Pavel Horcicka<br />

Head<br />

Wheat Breeding Department<br />

SELGEN<br />

Plant Breeding St. Stupice<br />

25084 Sibrina<br />

Czech Republic<br />

Phone: +42 (2) 81972462<br />

Fax: +42 (2) 81970465<br />

Email: horcicka@zero.cz<br />

Dr. Jerry W. Johnson<br />

Wheat Breeder<br />

Crop <strong>and</strong> Soil Sciences Division<br />

University <strong>of</strong> Georgia<br />

Georgia Station<br />

Griffin GA 30223<br />

United States<br />

Phone: +1 (770) 228 7321<br />

Fax: +1 (770) 229 3215<br />

Email: jjohnso@gaes.griffin.peachnet.edu<br />

Ms. Lise Nistrup Jorgensen<br />

Senior Scientist<br />

Research Centre Flakkebjerg<br />

Danish Institute <strong>of</strong> Agricultural Sciences<br />

Slagelse<br />

DK-4200 Flakkebjerg<br />

Denmark<br />

Phone: +45 (58) 113300<br />

Fax: +45 (58) 113301<br />

Email: LiseN.Jorgensen@agrsci.dk<br />

Dr. Ute Kastirr<br />

Scientific Collatorator<br />

Federal Centre for Breeding Research on<br />

Cultivated Plants<br />

Institute for Resistance Research <strong>and</strong><br />

Pathogendiagnostics<br />

Theodor-Roemer-Weg 4<br />

D-06449 Aschersleben<br />

Germany<br />

Phone: +49 (0) 3473 879197<br />

Fax: +49 (0) 3473 879200<br />

Email: u.kastirr@bromo.qlb.bafz.de<br />

Dr. Gert H.J. Kema<br />

Senior Scientist<br />

IPO-DLO<br />

P.O. Box 9060<br />

6700 GW Wageningen<br />

Netherl<strong>and</strong>s<br />

Phone: +31 317 476149<br />

Fax: +31 317 410113<br />

Email: G.H.J.Kema@IPO.DLO.NL<br />

Mr. Awgechew Kidane<br />

Quarantine Pathologist<br />

Ethiopian Agricultural Research<br />

Organization,<br />

Holetta ARC<br />

P.O. Box 2003<br />

Addis Ababa<br />

Ethiopia<br />

Phone: +251<br />

Fax: +251<br />

Email: <strong>CIMMYT</strong>-Ethiopia@cgiar.org<br />

Dr. Theodore J. Kisha<br />

Research Agronomist<br />

Agronomy Department<br />

Purdue University<br />

1150 Lilly Hall<br />

West Lafayette IN 47907<br />

United States<br />

Phone: +1 (765) 496 1917<br />

Fax: +1 (765) 496 2926<br />

Email: tkisha@purdue.edu<br />

Dr. Holger Klink<br />

Plant Pathologist at the University <strong>of</strong> Kiel<br />

Inst. for Phytopathlogie<br />

University <strong>of</strong> Kiel<br />

Hermann-Rodewald-Str.9<br />

D-24118 KIEL<br />

Germany<br />

Phone: +49 431 880 2994<br />

Fax: +49 461 880 1583<br />

Email:<br />

Barbara.Muth@ageurope.zeneca.com;<br />

Evelyn.Badeck@ageurope.zeneca.com<br />

Notes: HK@phytomet.uni.kiel.de<br />

Dr. Manfred Konradt<br />

Technical Manager<br />

ZENECA Agrochemicals<br />

Cmil-von-Behringstr.2<br />

D-60439 Frankfurt/Main<br />

Germany<br />

Phone: +49 069 58 01 414<br />

Fax: +49 069 5801 672<br />

Email: Barbara.Muth@ageurope.zeneca.com;<br />

Evelyn.Badeck@ageurope.zeneca.com<br />

Notes:<br />

Manfred.Konradt@geurope.zeneca.com<br />

Dr. Joseph M. Krupinsky<br />

Research Plant Pathologist<br />

Northern Great Plains Research Lab<br />

USDA-ARS<br />

P.O. Box 459<br />

M<strong>and</strong>an ND 58554-0459<br />

United States<br />

Phone: +1(701) 667-3011<br />

Fax: +1(701) 667-3054<br />

Email: Dvorakl@m<strong>and</strong>an.ars.usda.gov;<br />

krupinsj@m<strong>and</strong>an.ars.usda.gov<br />

Dr. Steven Leath<br />

Research Plant Pathologist<br />

Department <strong>of</strong> Plant Pathology<br />

USDA-ARS<br />

Box 7616, NCSU<br />

Raleigh NC 27695<br />

United States<br />

Phone: 919-515 6819<br />

Fax: 919-515 7716<br />

Email: Judith_Sulentic@ncsu.edu<br />

Notes: steven_leath@ncsu.edu<br />

Dr. Robert Loughman<br />

Senior Plant Pathologist<br />

Plant Protection Branch<br />

Agriculture Western Australia<br />

Plant Research <strong>and</strong> Development Services<br />

Locked Bag No. 4<br />

Bentley Delivery Centre W.A. 6983<br />

Australia<br />

Phone: +61 (618) 93683691<br />

Fax: +61 (618) 93672625<br />

Email: jtoms@agric.wa.gov.au;<br />

rloughman@agric.wa.gov.au<br />

Pr<strong>of</strong>. Dr. Bruce McDonald<br />

Group Leader<br />

Institute <strong>of</strong> Plant Sciences/Phytopathology<br />

Federal Insitute <strong>of</strong> Technology<br />

ETH-Zentrum, LFW<br />

Universitaetstr 2/LFW-B16<br />

CH-8092 Zuerich<br />

Switzerl<strong>and</strong><br />

Phone: +41 (1) 632 3847<br />

Fax: +41 (1) 632 15 72<br />

Email: Bruce.McDonald@ipw.agrl.ethz.ch


Dr. Ehud Meidan<br />

Wheat Breeder<br />

HAZERA Quality Seeds<br />

Mivhor<br />

M.P. Lachish-Dargm 75354<br />

Israel<br />

Phone: +972 (7) 6878155<br />

Fax: +972 (7) 6814057<br />

Email: Udi_Meidan@hazera.com<br />

Dr. Eugene Milus<br />

Associated Pr<strong>of</strong>essor<br />

Department <strong>of</strong> Plant Pathology<br />

University <strong>of</strong> Arkansas<br />

217 Plant Science Bldg.<br />

Fayetteville AR 72701<br />

United States<br />

Phone: +1(501) 575-2676<br />

Fax: +1(501) 575-7601<br />

Email: gmilus@comp.uark.edu<br />

Miss Mihaela V. Mincu<br />

Junior Research worker<br />

Research Institute for <strong>Cereals</strong> <strong>and</strong><br />

Industrial Crops<br />

FUNDULEA<br />

CP 22-171<br />

Bucuresti<br />

Romania<br />

Phone: +40 3154040<br />

Fax: +40 3110722<br />

Email: fundulea@cons.incerc.ro<br />

Mrs. Rose John Mongi<br />

Wheat Breeder<br />

MARTI-Uyole<br />

<strong>CIMMYT</strong><br />

P.O. Box 400<br />

Mbeya<br />

Tanzania<br />

Phone: + 255 (1) 614-645<br />

Fax: +255<br />

Email: c/o <strong>CIMMYT</strong>-Ethiopia@cgiar.org<br />

Dr. Chris C. Mundt<br />

Pr<strong>of</strong>essor<br />

Dept. <strong>of</strong> Botany <strong>and</strong> Plant Pathology<br />

Oregon State University<br />

2082 Cordley Hall<br />

Corvallis OR 97331-2902<br />

United States<br />

Phone: +1 541 737 5256<br />

Fax: +1 541 737 3573<br />

Email: mundtc@bcc.orst.edu<br />

Dr. Noel E.A. Murphy<br />

Research Fellow<br />

Murdoch University<br />

DSE South St<br />

Murdoch 6150<br />

Australia<br />

Phone: +61 (8) 9360 6097<br />

Fax: +61 (8) 9360 6303<br />

Email: nmurphy@central.murdoch.edu.au<br />

Dr. Lloyd R. Nelson<br />

Wheat Breeder <strong>and</strong> Pr<strong>of</strong>essor<br />

Texas A&M Research <strong>and</strong> Extension Center<br />

P.O. Box 200<br />

Overton TX 75684-0200<br />

United States<br />

Phone: +1(903) 834 6191<br />

Fax: +1(903)-834 7140<br />

Email: Ir-nelson@tamu.edu<br />

Ms. Guita Cordsen Nielsen<br />

Senior Adviser<br />

The National Department <strong>of</strong> Plant<br />

Production<br />

The Danish Agricultural Advisory Centre<br />

Udkaersvej 15<br />

DK-8200 Skejby Aarhus<br />

Denmark<br />

Phone: +45 (87) 405439; +45 (87) 405000;<br />

+45 20282695 mobile<br />

Fax: +45 (87) 405010<br />

Email: gcn@lr.dk<br />

Notes: Damgaard, Vroldvej 168; DK-8660<br />

Skaderborg; Phone: +45 (86) 57 98 00<br />

Dr. Thomas S. Payne<br />

Wheat Breeder/Pathologist<br />

Wheat Program<br />

<strong>CIMMYT</strong>-East Africa<br />

P.O. Box 5689<br />

Addis Ababa<br />

Ethiopia<br />

Phone: +251 (1) 615-127<br />

Fax: +251 (1) 614-645<br />

Email: t.payne@cgiar.org<br />

Telex: 21207 ILCA ET<br />

Dr. Sanjaya Rajaram<br />

Director<br />

Wheat Program<br />

<strong>CIMMYT</strong><br />

Lisboa 27, Col. Juárez<br />

Delegación Cuauhtemoc<br />

Apdo. Postal 6-641<br />

06600 Mexico D.F.<br />

Mexico<br />

Phone: +52 5804 2004<br />

Fax: +52 5804 7558/9<br />

Email: s.rajaram@cgiar.org<br />

Dr. Albert L. Scharen<br />

Pr<strong>of</strong>essor Emeritus<br />

Department <strong>of</strong> Plant Sciences<br />

Montana State University<br />

Bozeman MT 59717<br />

United States<br />

Phone: +1(406) 994-5162<br />

Fax: +1 (406) 994-1848<br />

Email: uplas@montana.campuscwix.net<br />

Mrs. Silvia Schuster<br />

Research Assitant<br />

Plant Sciences Department<br />

Tel-Aviv University<br />

Tel Aviv<br />

Israel<br />

Phone: +972 (3) 640 9766<br />

Fax: +972 (3) 640 9380<br />

Email: schus@post.tau.ac.il<br />

Dr. Gregory Shaner<br />

Pr<strong>of</strong>essor<br />

Botany <strong>and</strong> Plant Pathology<br />

Purdue University<br />

1155 Lilly Hall <strong>of</strong> Life Sciences<br />

West Lafayette IN 47907-1155<br />

United States<br />

Phone: +1(765) 494 4651<br />

Fax: +1 (765) 494 0363<br />

Email: shaner@btny.purdue.edu<br />

Dr. Michael W. Shaw<br />

Reader, Sch. <strong>of</strong> Plant Sciences<br />

The University <strong>of</strong> Reading<br />

Whiteknights<br />

Reading, RG6 6AS UK<br />

United Kingdom<br />

Phone: +44 118 931 8091<br />

Fax: +44 118 931 6577<br />

Email: M.W.Shaw@reading.ac.uk<br />

Dr. Ravi P. Singh<br />

Rust Genetist<br />

Wheat Program<br />

<strong>CIMMYT</strong><br />

Lisboa 27, Col. Juárez<br />

Delegación Cuauhtemoc<br />

Apdo. Postal 6-641<br />

06600 Mexico D.F.<br />

Mexico<br />

Phone: +52 5804 2004<br />

Fax: +52 5804 7558/9<br />

Email: r.singh@cgiar.org<br />

Dr. Bent Skovm<strong>and</strong><br />

Head, Wheat Genetic Research<br />

Wheat Program<br />

International Maize <strong>and</strong> Wheat<br />

Improvement Center<br />

Lisboa 27, Col. Juárez<br />

Delegación Cuauhtemoc<br />

Apdo. Postal 6-641<br />

06600 México D.F.<br />

Mexico<br />

Phone: +52 5804 2004 ext. 2226<br />

Fax: +52 5804 7558/9<br />

Email: b.skovm<strong>and</strong>@cgiar.org<br />

Dr. Brian J. Steffenson<br />

Associate Pr<strong>of</strong>essor<br />

Department <strong>of</strong> Plant Pathology<br />

North Dakota State University<br />

P.O. Box 5012<br />

Fargo ND 58105-5012<br />

United States<br />

Phone: +1(701)231-7078<br />

Fax: +1(701)231-7851<br />

Email: bsteffen@badl<strong>and</strong>s.nodak.edu<br />

Dr. Enrique Torres<br />

Calle 72 A No. 16-15 (301)<br />

Bogota<br />

Colombia<br />

Phone: +57-1-2359861, dentro Colombia:<br />

91-2359861<br />

Fax: +57-1-2359861<br />

Email: etorres@hotmail.com<br />

185


186<br />

Dr. Hala Toubia-Rahme<br />

Research Associate<br />

Dept. <strong>of</strong> Plant Pathology<br />

North Dakota State University<br />

P.O. Box 5012<br />

Fargo ND 58105-5012<br />

United States<br />

Phone: +1 (701) 231 7018<br />

Fax: +1 (701) 231 7851<br />

Email:<br />

Hala_Toubia_rahme@ndsu.nodak.edu<br />

Dr. Ludvik Tvaruzek<br />

Research Worker<br />

Agricultural Research Insitute Kromeriz<br />

Havlickova 2787<br />

767 01 Kromeriz<br />

Czech Republic<br />

Phone: +42 634317138<br />

Fax: +42 63422725<br />

Email: tvaruzek@vukrom.cz<br />

Dr. Peter Ueng<br />

Scientist<br />

Plant Molecular Biology Laboratory<br />

USDA-ARS<br />

MPPL, BARC-West, Bldg. 011A<br />

Beltsville MD 20705<br />

United States<br />

Phone: +1 (301) 504 6308<br />

Fax: +1 (301) 504 5449<br />

Email: pueng@asrr.arsusda.gov<br />

Dr. Maarten van Ginkel<br />

Head, Bread Wheat Program<br />

Wheat Program<br />

<strong>CIMMYT</strong><br />

Lisboa 27, Col. Juárez<br />

Delegación Cuauhtemoc<br />

Apdo. Postal 6-641<br />

06600 México D.F.<br />

Mexico<br />

Phone: +52 5804 2004<br />

Fax: +52 5804 7558/9<br />

Email: m.van-ginkel@cgiar.org; http://<br />

www.cimmyt.mx<br />

Ms. Carmen Velazquez<br />

Wheat Program<br />

<strong>CIMMYT</strong><br />

Lisboa 27, Col. Juarez<br />

Delegacion Cuauhtemoc<br />

Apdo. Postal 6-641<br />

Mexico D.F.<br />

Mexico<br />

Phone: +52 5804 2004<br />

Fax: +52 5804 7558/9<br />

Email: c.velazquez@cgiar.org<br />

Pr<strong>of</strong>. Dr. Joseph Alex<strong>and</strong>er Verreet<br />

Pr<strong>of</strong>essor<br />

Inst. for Phytopathlogie<br />

University <strong>of</strong> Kiel<br />

Hermann-Rodewald-Str.9<br />

D-24118 KIEL<br />

Germany<br />

Phone: +49 431 880 2996<br />

Fax: +49 461 880 1583<br />

Email:<br />

Barbara.Muth@ageurope.zeneca.com;<br />

Evelyn.Badeck@ageurope.zeneca.com<br />

Notes: JAV@phytomet.uni.kiel.de<br />

Mr. Robin Wilson<br />

Senior Wheat Breeder<br />

Crop Industries<br />

Agriculture Western Australia<br />

Locked Bag No. 4<br />

Bentley Delivery Centre W.A. 6983<br />

Australia<br />

Phone: +61 (618) 9368 3691<br />

Fax: +61 (618) 9367 2625<br />

Email: rwilson@agric.wa.gov.au<br />

Dr. Bruno Zwatz<br />

Head <strong>of</strong> the Institute<br />

Federal Office <strong>and</strong> Research Centre for<br />

Agriculture<br />

Institute <strong>of</strong> Phytomedicine<br />

Spargelfeldstrasse 191<br />

Vienna Wien A-1226<br />

Austria<br />

Phone: +43(1)73216-5500<br />

Fax: +43(1)73216-5194<br />

Email: bzwatz@bfl.gv.at;<br />

margarethe.zaufal@relay.bfl.at


ISBN: 970-648-035-8<br />

International Maize <strong>and</strong> Wheat Improvement Center<br />

Centro Internacional de Mejoramiento de Maíz y Trigo<br />

Lisboa 27, Apartado postal 6-641 México, D.F., México

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!