28.03.2013 Views

sdu faculty of forestry journal special edition 2009 - Orman Fakültesi

sdu faculty of forestry journal special edition 2009 - Orman Fakültesi

sdu faculty of forestry journal special edition 2009 - Orman Fakültesi

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

SDU<br />

FACULTY<br />

OF<br />

FORESTRY<br />

JOURNAL<br />

SPECIAL EDITION<br />

<strong>2009</strong><br />

PROCEEDINGS OF<br />

THE CONFERENCE OF<br />

IUFRO WORKING<br />

PARTY 7.02.02<br />

EĞİRDİR, TURKEY<br />

11-16 MAY <strong>2009</strong><br />

Guest Editor<br />

H. Tuğba<br />

DOĞMUŞ-LEHTIJÄRVI


SDU Faculty <strong>of</strong> Forestry Journal


SDU<br />

FACULTY OF FORESTRY JOURNAL<br />

Serial: A, Number: Special Edition, Year: <strong>2009</strong>, ISSN: 1302-7085<br />

Indexed in<br />

CAB<br />

TÜBİTAK<br />

PUBLICATION COMMITTEE<br />

Editor<br />

Asst. Pr<strong>of</strong>. Dr. Nevzat GÜRLEVİK<br />

Associate Editors<br />

Asst. Pr<strong>of</strong>. Dr. H. Oğuz ÇOBAN<br />

Asst. Pr<strong>of</strong>. Dr. Mehmet TOPAY<br />

Asst. Pr<strong>of</strong>. Dr. Abdullah SÜTÇÜ<br />

Res. Asst. Alper BABALIK<br />

Res. Asst. Yılmaz ÇATAL<br />

Expert Süleyman UYSAL<br />

Guest Editor<br />

H. Tuğba DOĞMUŞ-LEHTIJÄRVI<br />

COVER DESIGN<br />

SDU Press and Public Relations<br />

PRESS<br />

Fakülte Kitabevi-ISPARTA<br />

SDU Faculty <strong>of</strong> Forestry Journal is a refereed <strong>journal</strong> and published twice a year.<br />

Responsibility for the published papers concern to the Authors<br />

<strong>2009</strong> – SDU FFJ<br />

CORRESPONDENCE<br />

SDÜ <strong>Orman</strong> <strong>Fakültesi</strong>, 32260, ISPARTA<br />

Phone: + 90 246 2113198 Fax: +90 246 2371810<br />

e-mail: dergi@orman.<strong>sdu</strong>.edu.tr<br />

http://ormanweb.<strong>sdu</strong>.edu.tr/dergi


SDU Faculty <strong>of</strong> Forestry Journal


SDU FACULTY OF FORESTRY JOURNAL<br />

SERIAL: A, NUMBER: SPECIAL ISSUE, YEAR: <strong>2009</strong>, ISSN: 1302-7085<br />

Asko T. LEHTIJÄRVI<br />

Ayşe Gülden ADAY<br />

Organization Committee<br />

V<br />

SDU Faculty <strong>of</strong> Forestry<br />

SDU Faculty <strong>of</strong> Forestry<br />

Canpolat KAYA SDU Faculty <strong>of</strong> Forestry<br />

Denizhan ULUSAN SDU Faculty <strong>of</strong> Forestry<br />

Funda OSKAY<br />

H. Tuğba DOĞMUŞ-LEHTIJÄRVI<br />

Mertcan KARADENIZ<br />

Musa GENÇ<br />

Mustafa AVCI<br />

SDU Faculty <strong>of</strong> Forestry<br />

SDU Faculty <strong>of</strong> Forestry<br />

SDU Faculty <strong>of</strong> Forestry<br />

SDU Faculty <strong>of</strong> Forestry<br />

SDU Faculty <strong>of</strong> Forestry<br />

Süleyman UYSAL SDU Faculty <strong>of</strong> Forestry<br />

Special <strong>edition</strong> <strong>of</strong> the <strong>journal</strong> is financially supported by<br />

The Scientific and Technological Research Council <strong>of</strong> Turkey (TÜBİTAK)<br />

& Süleyman Demirel University


SDU FACULTY OF FORESTRY JOURNAL<br />

SERIAL: A, NUMBER: SPECIAL ISSUE, YEAR: <strong>2009</strong>, ISSN: 1302-7085<br />

VI


Preface<br />

Full papers<br />

SDU FACULTY OF FORESTRY JOURNAL<br />

SERIAL: A, NUMBER: SPECIAL ISSUE, YEAR: <strong>2009</strong>, ISSN: 1302-7085<br />

Foliage diseasas <strong>of</strong> conifers<br />

Needle diseases<br />

CONTENTS<br />

DOTHRISTROMA AND LECANOSTICTA NEEDLE BLIGHT IN THE CR<br />

Libor JANSKOVSKY, Miroslava BEDNAROVA, Milon DVORAK,<br />

Dagmar PALOVCIKOVA and Michal TOMSOVSKY..........................................................7-14<br />

TWO NEW SPECIES OF Lophodermium COLONISE SCOTS PINES NEEDLES IN<br />

SCOTLAND<br />

Sabrina N. A. REIGNOUX, Richard A. ENNOS, Sarah GREEN .........................................15-23<br />

RED BAND NEEDLE BLIGHT IN FINLAND, SYMPTOMS AND DISTRIBUTION<br />

Martti VUORINEN ..........................................................................................................24-26<br />

Scleroderris canker<br />

THE OCCURRENCE OF MICROCONIDIA ON Gremmeniella abietina (LAGERB.)<br />

MORELET<br />

Antti UOTILA ...................................................................................................................29-32<br />

CENTRAL NEWFOUNDLAND: ESCAPE FROM QUARANTINE<br />

Gary R. WARREN and Gaston LAFLAMME .....................................................................33-38<br />

Shoot blights<br />

CONE DAMAGES BY Diplodia pinea AND SEED BORING INSECTS ON Pinus pinea<br />

L. (ITALIAN STONE PINE) IN CENTRAL ITALY<br />

Paolo CAPRETTI, Matteo FEDUCCI, Martina CAMBI, Alessia PEPORI,<br />

Daniele BENASSAI ...........................................................................................................41-47<br />

SUSCEPTIBILITY OF DIFFERENT CONIFEROUS SEEDLINGS INOCULATED<br />

WITH Diplodia pinea<br />

H. Tuğba DOĞMUŞ-LEHTIJÄRVI, Asko LEHTIJÄRVI, Gürsel KARACA,<br />

A. Gülden ADAY and Funda OSKAY ...............................................................................48-56<br />

SITE AND STAND CHARACTERISTICS OF A Pinus brutia STAND INFECTED<br />

WITH Diplodia pinea IN TURKEY<br />

Nevzat GÜRLEVIK, H. Tuğba DOĞMUŞ-LEHTIJÄRVI, Asko LEHTIJÄRVI,<br />

A. Gülden ADAY...............................................................................................................57-64<br />

THE EFFECTS OF Sirococcus SHOOT BLIGHT AND VITALITY FERTILIZATION<br />

ON GROWTH OF MATURE NORWAY SPRUCE<br />

Markus HUBER, Erhard HALMSCHLAGER and Hubert STERBA....................................65-70<br />

INTERACTION BETWEEN Diplodia pinea, D. scrobiculata AND SEVERAL FUNGAL<br />

ENDOPHYTES IN RED AND JACK PINE SEEDLINGS<br />

Oscar SANTAMARÍA, Denise R. SMITH, Glen R. STANOSZ ..........................................71-84<br />

ADELGID GALLS ON SPRUCE AS A RESERVOIR INOCULUM SOURCE FOR<br />

THE SHOOT BLIGHT PATHOGEN Diplodia pinea<br />

Glen R. STANOSZ, Denise R. SMITH -, and S. ZHOU.......................................................85-92<br />

VII


SDU FACULTY OF FORESTRY JOURNAL<br />

SERIAL: A, NUMBER: SPECIAL ISSUE, YEAR: <strong>2009</strong>, ISSN: 1302-7085<br />

Dieback and canker diseases<br />

Dieback dieases<br />

THE CURRENT SITUATION OF ASH DIEBACK CAUSED BY Chalara fraxinea IN<br />

AUSTRIA<br />

Thomas KIRISITS, Michaela MATLAKOVA, Susanne MOTTINGER-KROUPA,<br />

Thomas L. CECH, Erhard HALMSCHLAGER................................................................. 97-119<br />

DIEBACK ON Fraxinus ornus IN KONYA REGION<br />

Asko LEHTIJÄRVI, H. Tuğba DOĞMUŞ-LEHTIJÄRVI, Mertcan KARADENİZ,<br />

Mustafa UYGUN........................................................................................................... 120-123<br />

ASH DIEBACK IN THE CZECH REPUBLIC<br />

Petr STASTNY, Dagmar PALOVCIKOVA and Libor JANKOVSKY............................. 124-128<br />

Canker dieases<br />

HORSE CHESTNUT BLEEDING CANKER – BAGGING THE BUG!<br />

Sarah GREEN, Bridget LAUE, Grace MACASKILL, Heather STEELE.......................... 131-135<br />

AN OVERVIEW OF POTENTIAL INFECTION COURTS FOR Neonectria fuckeliana,<br />

THE CAUSAL AGENT OF NECTRIA FLUTE CANKER IN Pinus radiata IN NEW<br />

ZEALAND<br />

Anna J..M. HOPKINS, PATRICIA E. CRANE and Margaret A. DICK ........................... 136-140<br />

PRELIMINARY RESULTS OF MYCOFLORA ASSOCIATED WITH CANKERS ON<br />

Cupressus sempervirens var. horizontalis (Mill.) GORDON IN TURKEY<br />

Asko LEHTIJÄRVI, H. Tuğba DOĞMUŞ- LEHTIJÄRVI, Funda OSKAY,<br />

A. Gülden ADAY .......................................................................................................... 141-149<br />

SOME MORPHOLOGICAL ASPECTS OF EUTYPELLA CANKER OF MAPLE<br />

(Eutypella parasitica)<br />

Nikica OGRIS, Barbara PIŠKUR, and Dušan JURC........................................................ 150-161<br />

OCCURRENCE OF Pseudomonas syringae ON POPLAR DAMAGED BY NECROSIS<br />

AND CANKER<br />

Irmtraut ZASPEL and Volker SCHNECK....................................................................... 162-167<br />

Rust dieases<br />

SEASONAL FRUITING AND SPORULATION OF THEKOPSORA AND<br />

CHRYSOMYXA CONE RUSTS IN NORWAY SPRUCE CONES AND ALTERNATE<br />

HOSTS IN FINLAND<br />

Juha KAITERA, Eila TILLMAN-SUTELA and A. KAUPPI........................................... 171-176<br />

PRELIMINARY STUDIES ON GENETIC VARIATION IN Gymnosporangium fuscum<br />

IN THE LAKES DISTRICT OF TURKEY DETECTED WITH M13 MINISATELLITE<br />

MARKER<br />

Asko LEHTIJÄRVI, H. Tuğba DOĞMUŞ-LEHTIJÄRVI, A. Gülden ADAY,<br />

Funda OSKAY .............................................................................................................. 177-181<br />

FACTORS FAVOURING BROOM RUST INFECTION IN ADVANCE PLANTINGS<br />

OF Abies alba IN SW-GERMANY<br />

Tilo PODNER, Berthold METZLER .............................................................................. 182-186<br />

VIII


SDU FACULTY OF FORESTRY JOURNAL<br />

SERIAL: A, NUMBER: SPECIAL ISSUE, YEAR: <strong>2009</strong>, ISSN: 1302-7085<br />

Foliage diseases <strong>of</strong> hardwood<br />

NON-NATIVE HOSTS AND CONTROL OF Rhytisma acerinum CAUSING TAR SPOT<br />

OF MAPLE<br />

Tom HSIANG, Tian LYNN, Coralie SOPHER............................................................... 189-193<br />

BIOLOGICAL CONTROL TRIALS OF BEECH BARK DISEASE UNDER<br />

LABORATORY CONDITIONS<br />

Gaston LAFLAMME, Simon BOUDREAULT, Robert LAVALLÉE, Martine BLAIS,<br />

Jean Yves BLANCHETTE ............................................................................................. 194-199<br />

PATHOGENICITY OF Fusarium circinatum NIREMBERG & O’DONNELL ON<br />

SEEDS AND SEEDLINGS OF RADIATA PINE<br />

Pablo MARTÍNEZ-ÁLVAREZ, Juan BLANCO, Milagros DE VALLEJO,<br />

Fernando M. ALVES-SANTOS, Julio Javier DIEZ ........................................................ 200-205<br />

POWDERY MILDEW ON WOODY PLANTS IN THE CZECH REPUBLIC<br />

Dagmar PALOVČÍKOVÁ, Hana DANČÁKOVÁ, Hana MATOUŠKOVÁ,<br />

Jindřiška JUNÁŠKOVÁ and Libor JANKOVSKÝ.......................................................... 206-215<br />

Abiotic diseases and other diseases<br />

URBAN TREE HEALTH OF 49 GREEN SPACES IN MADRID (SPAIN)<br />

Eva ALFONSO CORZO, María Jesús GARCIA and J. A SAIZ DE OMEÑACA............. 219-232<br />

CHARACTERISATION OF CZECH Ophiostoma novo-ulmi ISOLATES<br />

Milon DVORAK, Libor JANKOVSKY, J .KRAJNAKOVA ........................................... 233-237<br />

HAIL DAMAGE OF FOREST TREES IN WESTERN CANADA<br />

Yasuyuki HIRATSUKA................................................................................................. 238-240<br />

Extended abstracts<br />

THREATENING TREE DISEASE IN EAST AFRICA<br />

Pia BARKLUND, Jane NJUGUNA, Abdella GURE, Philip NYEKO, Katarina IHRMARK<br />

and Jan STENLID . ........................................................................................................ 243-244<br />

PREMATURE DEFOLIATION OF Cedrus libani IN SOUTH- WESTERN TURKEY<br />

Asko LEHTIJARVI, H. Tuğba DOĞMUŞ-LEHTIJÄRVI....................................................... 245<br />

CONTRIBUTIONS TO THE PHYLOGENY OF EUROPEAN Porodaedalea species<br />

(BASIDIOMYCETES, HYMENOCHAETALES)<br />

Michal TOMSOVSKY, Libor JANKOVSKY.................................................................. 246-251<br />

Abstracts<br />

EXAMINING THE GEOGRAPHIC DISTRIBUTION OF Diplodia pinea AND D.<br />

scrobiculata: A CASE STUDY FROM MINNESOTA, USA<br />

J. S. ALBERS, Denise R. SMITH, Glen R. STANOSZ........................................................... 255<br />

FOREST INVASIVE ALIEN FUNGAL SPECIES PRESENT IN LIVE PLANT<br />

MATERIAL<br />

Jean A. BERUBE.................................................................................................................. 256<br />

NEW ADVANCES IN THE STUDY OF THE TAXONOMY OF THE EUROPEAN<br />

RACE OF Gremmeniella abietina<br />

Leticia BOTELLA, Julio Javier DIEZ and Jarkko HANTULA................................................ 257<br />

IX


SDU FACULTY OF FORESTRY JOURNAL<br />

SERIAL: A, NUMBER: SPECIAL ISSUE, YEAR: <strong>2009</strong>, ISSN: 1302-7085<br />

STUDIES ON THE SIGNIFICANCE, CAUSAL AGENTS AND CONTROL<br />

METHODS OF DAMPING- OFF DISEASE IN FOREST NURSERIES OF AEGEAN<br />

AND LAKES DISTRICT<br />

H. Tuğba DOĞMUŞ LEHTIJÄRVI and Gülay TURHAN..................................................... 258<br />

A FOLIAR DISEASE OF Celtis glabrata IN THE LAKES REGION<br />

Gürsel KARACA, H. Tuğba DOĞMUŞ LEHTIJÄRVI –, Hüseyin FAKIR ............................. 259<br />

DETERMINATION OF MACROMYCETES IN THE REGION OF KOCAELI<br />

Ayhan KARAKAYA ............................................................................................................ 260<br />

ARE SUBPOPULATIONS OF Heterobasidion parviporum DIFFERENTIATED BY<br />

LOCAL CLIMATE?<br />

Michael M. MÜLLER, Nicola LA PORTA, Jaana EKOJÄRVI, Jarkko HANTULA and Kari<br />

KORHONEN........................................................................................................................ 261<br />

ATTEMPTS TO NATURALLY REGENRATE RED PINE CAN BE THREATENED<br />

BY DIPLODIA SHOOT BLIGHT DAMAGE TO UNDERSTORY SEEDLINGS<br />

B.W. OBLINGER, Denise R. SMITH and Glen R. STANOSZ............................................... 262<br />

SOME FUNGAL SPECIES ON Pinus pinaster Ait. AND Pinus radiata D. Don<br />

PLANTATIONS IN MARMARA REGION OF TURKEY<br />

Fazıl SELEK......................................................................................................................... 263<br />

RESPONSE OF Alnus tenuifolia TO INOCULATION WITH Valsa melanodiscus.<br />

Glen R. STANOSZ, L. M. TRUMMER, J. K. ROHRS-RICHEY, - G.C. ADAMS and<br />

J. T. WORRALL................................................................................................................... 264<br />

GREMMENIELLA INFECTIONS ON SEEDLINGS AFTER REPLANTING<br />

SEVERELY INFECTED PINE FOREST<br />

Elna.STENSTRÖM........................................................................................................ 265-266<br />

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

X


Preface<br />

The meeting <strong>of</strong> IUFRO Working Party (WP) 7.02.02 “Foliage, Shoot<br />

and Stem Diseases <strong>of</strong> Forest Trees” was held in Eğirdir, Isparta,<br />

Turkey. Local organizers were Dr. H. Tuğba Doğmuş-Lehtijärvi and<br />

her colleagues from the Faculty <strong>of</strong> Forestry, Süleyman Demirel<br />

University. In the opening ceremony, vice rectors <strong>of</strong> the university,<br />

Dr. Vecihi Kırdemir and Dr. Mehmet Ali Koyuncu and also dean <strong>of</strong><br />

the Forestry Faculty, Dr. Musa Genç and the Coordinator <strong>of</strong> the<br />

IUFRO Forest Health Division, Dr. Gaston Laflamme welcomed the<br />

participants <strong>of</strong> the WP. We thank the attendees who presented oral<br />

and poster presentations, local organisers for their great effort,<br />

Süleyman Demirel University for their kind support and The<br />

Scientific and Technological Research Council <strong>of</strong> Turkey (TUBİTAK)<br />

for financial support.<br />

Forest pathology has developed a lot during the past 40 years. The<br />

most remarkable progress is the possibility to use DNA methods in<br />

studying genetics <strong>of</strong> pathogens and host plants. DNA methods are<br />

helping us also when we try to find out the origin <strong>of</strong> new diseases,<br />

alien or original? What is the next step? Also the development <strong>of</strong><br />

information technology has affected a lot <strong>of</strong> our daily working<br />

compared to time before computers and internet. The first step in<br />

forest pathology is to describe the pathogen and disease symptoms.<br />

The pathogenicity should be tested according to Koch’s postulates.<br />

Then the ecology <strong>of</strong> the disease should be examined experimentally<br />

including the interactions between the pathogen, host, other<br />

microbes and environment. DNA methods can be used in studying<br />

the variation <strong>of</strong> pathogens, taxonomy, endophytes and<br />

pathogenesis.<br />

The first meeting <strong>of</strong> the IUFRO Working party 7.02.02 was<br />

arranged in 1973 in Minneapolis, USA. Now Turkey was the 10th<br />

country to host the meeting so far. In the first meetings one <strong>of</strong> the<br />

main subjects was Scleroderris canker which was an important<br />

disease in Europe, Northern America and Japan. Surprisingly, the<br />

subjects <strong>of</strong> the presentations in this meeting were scattered,<br />

including some other important biotic and abiotic diseases <strong>of</strong> forest<br />

trees.<br />

XI


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

When new disease appears we should start from the beginning.<br />

Thanks to new methods or tools we can get the same knowledge<br />

much faster than in the 1900´s. It seems that new diseases appear<br />

continuously, f. ex. Phytophthora alni and Chalara fraxinea in<br />

Europe. The human interest to grow exotic tree species and even the<br />

commercial international seedling trade keeps the forest<br />

pathologists in work in future too.<br />

The climate change means also challenges to forest pathologists.<br />

The warming climate is a fact based on the laws <strong>of</strong> physics. The<br />

concentration <strong>of</strong> carbon dioxide and other greenhouse gases are<br />

rising fast in atmosphere. Near the arid areas the warming means<br />

the drought problems and in the north it means the new pathogens<br />

from milder climates together with faster forest growth. The climate<br />

change can disturb the evolutionary balance between the plants<br />

and the pathogens. But the climate change highlights also the<br />

importance <strong>of</strong> forests. At the same time, forests produce renewable<br />

materials and energy and bind the carbon dioxide from the air. A<br />

healthy forest is important in controlling the warming. This is a<br />

fact which is very important to get to a common knowledge <strong>of</strong> people<br />

and this is our task.<br />

Nowadays we can exchange information quickly by internet. Why<br />

to fly to another side <strong>of</strong> the earth to meet colleagues? I think the<br />

human being needs human contacts and conversation. This is an<br />

excellent opportunity to discuss and to start co-operation. The<br />

internet has not yet changed human genes. We had once more a<br />

very successful meeting in Eğirdir with very sincerely and relaxed<br />

atmosphere and found the opportunity to share the experiences and<br />

create the new ideas coming from basically all age groups, from<br />

very young to seniors, and we had close to 50 participants from 15<br />

countries with 29 oral presentations and 15 posters. This proceeding<br />

includes full papers as well as extended and short abstracts and we<br />

would like to mention that the responsibility for the published<br />

papers lies with the authors.<br />

Finally, we would like to see you in the next meeting to be held in<br />

Spain, 2011 organised by new deputy Julio Javier Diez Casero!<br />

Antti Uotila H. Tuğba Doğmuş-Lehtijärvi<br />

Coordinator Local organizer & Deputy<br />

IUFRO WP 7.02.02<br />

XII


Full Papers<br />

1


Foliage Diseases <strong>of</strong> Conifers<br />

3


Needle Diseases<br />

5


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 7-14<br />

DOTHISTROMA AND LECANOSTICTA NEEDLE BLIGHT IN THE CR<br />

Libor JANKOVSKÝ 1* , Miroslava BEDNÁŘOVÁ 1 , Miloň DVOŘÁK 1 ,<br />

Dagmar PALOVČÍKOVÁ 1 , Michal TOMŠOVSKÝ 1<br />

1 Dpt. <strong>of</strong> Forest Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood Technology,<br />

Mendel University <strong>of</strong> Agriculture and Forestry, Zemědělská 3, 613 00 Brno, Czech Republic<br />

ABSTRACT<br />

*jankov@mendelu.cz<br />

Dothistroma needle blight is widespread in the Czech Republic now, although<br />

the first finding was noted in 2000. To date, it has been identified on 21 species <strong>of</strong><br />

Pines, 4 species <strong>of</strong> Spruces and also on Douglas fir in the CR. Records on Scots<br />

pine were exceptionally rare in the CR and also in Europe up to spring 2008.<br />

Brown spot needle blight caused by Lecanosticta acicola was for the first time<br />

reported in the Czech Republic on June 2007, actually is known from 2 localities<br />

on Pinus rotundata. Lecanosticta acicola coincides in observed localities with<br />

Dothistroma needle blight on Scots pine Pinus sylvestris, bog pine Pinus rotundata<br />

and their hybrid P. digenea, however no finding <strong>of</strong> both diseases on the same tree<br />

was observed.<br />

Key words: Dothistroma septosporum, Lecanosticta acicola, needlecast, alien<br />

species<br />

1. INTRODUCTION<br />

Dothistroma needle blight Mycosphaerella pini E. Rostrup, resp. its anamorph<br />

Dothistroma septosporum is known from most <strong>of</strong> European countries, eg. France,<br />

Italy, Portugal, Spain, Georgia (Ivory, 1994), UK (Murray and Batko, 1962),<br />

Croatia (Novak-Agbaba et al., 1997; EPPO, 2005), Montenegro and Serbia<br />

(Karadzic, 1989, 2004), Romania (Gremmen, 1968) etc. From Central Europe it<br />

was reported from Austria (Petrak, 1961), Slovenia (Macek, 1975), Germany<br />

(Butin, 1983; Richter, 1983) and Poland (Kowalski and Jankowiak, 1998) where it<br />

7


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

was found in May 1990, Slovakia (Kunca and F<strong>of</strong>fová, 2000), Hungary (Koltay,<br />

1997) and Czech Republic (Jankovský et al., 2000, 2004). Recent findings are from<br />

Netherlands (EPPO, 2007) and Belgium (EPPO, 2008a). Up to 2008 there was no<br />

report on Dothistroma from Scandinavia, actually there are reports from Estonia<br />

(Hanso and Drenkhan 2008), Finland (EPPO, 2008b), Sweden (Stenlid, oral<br />

communication; DNA isolated from needle, no symptoms); it is also reported from<br />

Lithuania (Fig 1). European strains belong mostly to Dothistroma septosporum,<br />

however Barnes et al. (2004, 2007) recorded also D. pini from samples from<br />

Ukraine.<br />

Fig. 1 Distribution <strong>of</strong> Dothistroma needle blight in Europe. Years mean first published<br />

report, in parentheses are years <strong>of</strong> findings, if differ from year <strong>of</strong> publishing.<br />

Brown spot needle blight Mycosphaerella dearnessii M. E. Barr, resp.<br />

anamorphic stage Lecanosticta acicola (Thüm.) Syd. is in Europe reported (Fig. 2)<br />

from Austria (Petrak, 1961; Brandstetter and Cech, 1999, 2003; Kirisits and Cech,<br />

2006), France (Chandalier et al., 1993), Italy (Porta and Capretti, 2000), Germany<br />

(Butin and Richter, 1983; Pehl, 1995), Switzerland (Holdenrieder and Sieber,<br />

1995), Bulgaria and formerly Yugoslavia (Holdenrieder and Sieber, 1995), Serbia<br />

(Milanovic and Karadzic, oral communication), in 1979 it was reported from<br />

Croatia (Novak-Agbaba and Halambek, 1997; EPPO, 2007). Some new records<br />

origin from Estonia (Cech, 2008, oral communication) and Slovenia (Jurc, 2008).<br />

8


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Fig. 2. Distribution <strong>of</strong> Lecanosticta acicola in Europe. Years mean first report, in<br />

parentheses are years <strong>of</strong> findings, if differ from year <strong>of</strong> publishing.<br />

2. MATERIAL AND METHODS<br />

Records from the Czech Republic are based on monitoring carried out in 2000 –<br />

2008. Pine and also Spruce and Douglas fir needle samples were examined, mainly<br />

from regions <strong>of</strong> Southern and Central Moravia, Silesia, Eastern and Central<br />

Bohemia.<br />

The presence <strong>of</strong> the pathogen was always investigated according to<br />

characteristic symptoms such as red bands, dying tips <strong>of</strong> needles or the occurrence<br />

<strong>of</strong> subepidermal sporocarps, acervuli. Exact identification was proved on the basis<br />

<strong>of</strong> microscopic analyses <strong>of</strong> conidia.<br />

Isolation <strong>of</strong> culture was made on 3% MEA containing malt extract 30 g/l,<br />

pepton 5 g/l, agar 15 g/l, without addition <strong>of</strong> any antibiotics. Pieces <strong>of</strong> needles with<br />

acervuli 3 - 5 mm long were on surface sterilized by sodium hypochlorite 7%,<br />

subsequently by ethanol 96% and washed by sterilized water and put on malt<br />

extract agar. After 3 weeks <strong>of</strong> incubation, when new conidia occurred on fruiting<br />

bodies, conidia were inoculated into new medium.<br />

Herbarium specimens are deposited at Herbarium <strong>of</strong> Faculty <strong>of</strong> Forestry and<br />

Wood Technology (BRNL).<br />

9


3. RESULTS AND DISCUSSION<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Dothistroma needle blight is widespread across the CR now, although the first<br />

finding was noted in 2000 (eg. Jankovský et al., 2000, 2004; Bednářová et al.,<br />

2007). More than 80 host species <strong>of</strong> Dothistroma needle blight are mentioned from<br />

all continents (Bednářová et al., 2006). To date, it has been identified on 21 species<br />

<strong>of</strong> pine, 4 species <strong>of</strong> spruce and also on Douglas fir in the CR: Pinus aristata<br />

Engelm., P. attenuata Lemon, Pinus banksiana Lamb., Pinus cembra L. var.<br />

sibirica (Du Tour) G. Don, Pinus contorta Douglas ex Loudon, Pinus x digenea<br />

Beck (=P. rotundata x P. sylvestris), Pinus heldreichii H. Christ, Pinus heldreichii<br />

H. Christ var. leucodermis (Antoine) Markgraf ex Fitschen, syn. Pinus leucodermis<br />

Ant., Pinus jeffreyi Grev. et Balf, Pinus mugo Turra, Pinus nigra Arnold, Pinus<br />

ponderosa Douglas ex Lawson, Pinus pungens Lambert, Pinus rigida Miller, Pinus<br />

rotundata Link = Pinus mugo nothosubsp. rotundata (Link) Janchen & Neumayer,<br />

Pinus strobus L. var. sibirica, Pinus sylvestris L., Pinus tabuliformis Hort. ex<br />

Carrière, Pinus taeda L., Pinus thunbergii Parlatore, syn. Pinus thunbergiana<br />

Franco, Pinus wallichiana A. B. Jackson, Picea abies L. Karst., Picea pungens<br />

Engelm., Picea omorika (Pančić) Purkyně, Picea schrenkiana Fisch. & C. A. Mey,<br />

Pseudotsuga menziesii. Austrian pine Pinus nigra Arnold, mountain pine Pinus<br />

mugo Turra, Pinus ponderosa Douglas ex Lawson, Pinus jeffreyi Grev. are the<br />

most frequent and most susceptible hosts. As for species <strong>of</strong> other genera Picea<br />

pungens Engelm., Picea abies L. Karst., Picea omorika Purkyně and Picea<br />

schrenkiana Fisch. & C. A. Mey were noted as hosts. Dothistroma septosporum<br />

was also isolated from needles <strong>of</strong> Pseudotsuga menziesii. Symptoms on needles <strong>of</strong><br />

Douglas fir were not so clear, acervuli were observed exceptionally.<br />

Records on Scots pine were exceptionally rare in the CR and also in Europe up<br />

to spring 2008. Risk <strong>of</strong> Dothistroma needle blight for Scots pine in Europe is noted<br />

eg. by Lang and Karadzic (1987). According to Gadgil (1984), Pinus sylvestris is<br />

highly susceptible. Contrary, according to data from Great Britain, Peterson (1982),<br />

mentions that the attack occurs very rarely. However several hectares <strong>of</strong> Scots pine<br />

plantations, infested by Dothistroma septosporum, about 10 years old, were<br />

registered in Southern Bohemia, in Forest district Nové Hrady, Třeboň area, in<br />

March, 2008. In <strong>2009</strong> progress <strong>of</strong> infection in the same plot contrary precedent<br />

year. Large infestations were registered also on Scots pine in peat bog nature<br />

reserve Soběslavská blata in Southern Bohemia. Infected trees were origin from<br />

natural regeneration. Surrounding commercial forest was not affected. Dothistroma<br />

septosporum outbreak on Scots pines was observed simultaneously in large areas<br />

across the Central Finland in spring 2008 (EPPO, 2008). Dothistroma seems to be<br />

threat for Scots pine, including natural stands in Europe, although it was mostly<br />

reported from plantations <strong>of</strong> introduced species eg. Pinus nigra, P. ponderosa, P.<br />

contorta etc.<br />

Brown spot needle blight caused by Lecanosticta acicola was for the first time<br />

reported in the Czech Republic on June 10, 2007 (Jankovský et al. 2008). The first<br />

record is from the peat bog National Nature Reserve Červená Blata in South<br />

10


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Bohemia, close to town Třeboň; coord. N 48°51'37.06", E 14°48'44.09". The<br />

disease was observed on 10 – 40 years old trees <strong>of</strong> Pinus rotundata, (Jankovský et<br />

al., 2008). The new record in the CR is from the same host species, 10-60 years old<br />

in National Nature Reserve Borkovická blata, near town Soběslav (coord. N<br />

49°14'16.3" E 14°37'54.2") on August 7, 2008 (Jankovský et al., in press). Both<br />

places are very strictly protected areas. Typical symptoms (according to EPPO,<br />

2005) were observed in current year needles declined from tips in middle <strong>of</strong> July<br />

and in August. Brown spots with apparent yellow separation were present on green<br />

needles as well. Visible yellow belts were present between dead tissues <strong>of</strong> killed<br />

tips and green tissues. Studied conidia were subhyalinne, even dark olive green,<br />

surface <strong>of</strong> conidia echinulate to verrucose or tuberculate, straight to curved, with<br />

one to five septae, fusiform to cylindrical, size 3 - 5 μm × 21 - 44 μm. On 3% MEA<br />

medium, the fungus produced grayish green olive to olive black stromatic colonies,<br />

producing slime with conidias<br />

Lecanosticta acicola coincides in localities with causal agent <strong>of</strong> Dothistroma<br />

needle blight Dothistroma septosporum on Scots pine Pinus sylvestris, bog pine<br />

Pinus rotundata and their hybrid P.×digenea. Nevertheless, the threat <strong>of</strong> the<br />

disease spreading to Scots pines frequently planted in the region remains unclear<br />

yet. While bog pines inside the nature reserves display remarkable needle<br />

defoliation, Scots pines in surrounding managed stands are without visible<br />

symptoms <strong>of</strong> infection by Lecanosticta acicola.<br />

4. CONCLUSIONS<br />

With respect to actual epidemic situation in some countries, it is necessary to<br />

discuss the role <strong>of</strong> climatic factors in Europe and trade with plant material as main<br />

risk factors for spreading <strong>of</strong> both diseases. Dothistroma needle blight and brown<br />

spot needle blight are relatively quickly spreading needle casts in Central and also<br />

in Northern Europe. Within past 15 years, Dothistroma was reported from many<br />

new areas. Spreading <strong>of</strong> these diseases should be considered as result <strong>of</strong> climatic<br />

extremes and also one <strong>of</strong> exhibitions <strong>of</strong> climatic changes (eg. Woods et al., 2005).<br />

However reasons <strong>of</strong> spreading are still not sure. We cannot exclude also some other<br />

factors as trade with plant material. Reasons <strong>of</strong> occurrence <strong>of</strong> disease <strong>of</strong> strictly<br />

protected areas without any human interventions for many decades are not sure.<br />

Control <strong>of</strong> these needle casts is problematic due distribution in large infested areas.<br />

Dothistroma needle blight is established in most areas across Europe now and<br />

seems to be problem in Central and North Europe now due to very quick spreading<br />

and adaptation for climatic and natural conditions in new areas.<br />

5. ACKNOWLEDGEMENTS<br />

The results presented here were obtained from projects supported by NAZV<br />

QH81039.<br />

11


6. REFERENCES<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Barnes, I., Wingfield, M.J., Groenewald, M., Kirisits, T., Crous, P.W., Wingfield, B.D., 2007.<br />

Exposing the Enigma <strong>of</strong> Dothistroma Needle Blight Using Molecular Markers – a<br />

Progress Report. Acta Silv. Lign. Hung., Spec. Edition: 239-240.<br />

Barnes, I., Crous, P.W., Wingfield, M.J., Wingfield, B.D., 2004. Multigene phylogenies reveal that<br />

red band needle blight <strong>of</strong> Pinus is caused by two distinct species <strong>of</strong> Dothistroma, D.<br />

septosporum and D. pini. Stud. Mycol. 50, 551-565.<br />

Bednářová, M., Palovčíková, D., Jankovský, L., 2006. The host spektrum <strong>of</strong> Dothistroma needle<br />

blight Mycosphaerella pini E. Rostrup – new hosts <strong>of</strong> Dothistroma needle blight observed<br />

in the Czech Republic. Journal <strong>of</strong> Forest Science 52(1), 30–36.<br />

Bednářová, M., Bodejčková, I., Palovčíková, D., Jankovský, L., 2007.<br />

The contemporary situation <strong>of</strong> Dothistroma needle blight outbreak in the Czech Republic.<br />

Acta Silv. Lign. Hung. Spec. Edition, 17-22.<br />

Brandstetter M., Cech T. 2003. Lecanosticta - Kiefernnadelbäume (Mycosphaerella dearnessii Barr)<br />

in Niederösterreich = Brown Spot Needle Blight <strong>of</strong> Pines (Mycosphaerella dearnessii<br />

Barr) in Lower Austria. Centralblatt für das gesamte Forstwesen, Wien, 120 (3/4), 163-<br />

175.<br />

Brandstetter, M., Cech, T., 1999. Neue Nadelkrankheiten Kiefer. Osterreichishe Forstzeitung, 99 (3):<br />

p. 35 - 36.<br />

Butin H., Richter J., 1983 Dothistroma - Nadelbräune: Eine neue Kiefernkrankheit in der<br />

Bundesepublik Deutschland. Nachrichtenblatt Deutscher Pflanzenschutzdienst<br />

(Braunschweig), 35: p. 129 - 131.<br />

Doroguine, G., 1911. Une maladie cryptogamique du Pin. Bulletin Trimestriel de la Société<br />

Mycologique de France 27 (1): 105 – 106.<br />

EPPO, 1997. Fungal diseases <strong>of</strong> forest trees in the coastal region <strong>of</strong> Croatia. EPPO Reporting Service.<br />

1997/149.<br />

EPPO, 2005a. EPPO Standards. Diagnostic PM 7/46. OEPP/EPPO, EPPO Bulletin, 35: p. 271 – 273.<br />

EPPO, 2005b. New data on quarantine pests and pests <strong>of</strong> the EPPO Alert List. EPPO Reporting<br />

Service. 2005/33.<br />

EPPO, 2007. First report <strong>of</strong> Mycosphaerella pini in the Netherlands. EPPO Reporting Service.<br />

2008/55.<br />

EPPO, 2008a. First report <strong>of</strong> Mycosphaerella pini in Belgium. EPPO Reporting Service. 2007/211.<br />

EPPO, 2008b. First report <strong>of</strong> Mycosphaerella pini in Finland. EPPO Reporting Service. 2008/185.<br />

Gadgil, P.D., 1984. Dothistroma needle blight. Forest Pathology in New Zealand, No. 5. New<br />

Zealand Forest Service, Roturua, New Zealand.<br />

Gremmen, J., 1968. The presence <strong>of</strong> Scirrhia pini Funk et Parker in Romania (Conidial stage:<br />

Dothistroma pini Hulb.). Bulletin Trimestiel de la Société Mycologique de France, 84:<br />

489–492.<br />

Hanso, M., Drenkhan, R., 2008. First observations <strong>of</strong> Mycosphaerella pini in Estonia. New disease<br />

reports 17.<br />

Holdenrieder, O., Sieber, T.N., 1995. First report <strong>of</strong> Mycosphaerella dearnessii in Switzerland. Forest<br />

Pathology, 25(5), 293 – 295.<br />

Hulbary, R.L., 1941. A needle blight <strong>of</strong> Austrian pines: III. <strong>of</strong> the Illinois Natural History Survey<br />

Bulletin 21, 231 – 236.<br />

12


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Chandelier, P., Lafaurie, C., Maugard, F., 1994. Découverte en France de Mycosphaerella dearnessii<br />

sur Pinus attenuata x radiata. C.R. Acad. Agric. Fr., 80, 103 - 108.<br />

Ivory, M.H., 1994. Records <strong>of</strong> foliage pathogens <strong>of</strong> Pinus species in tropical countries. Plant<br />

Pathology 43, 511–518.<br />

Jankovský, L., Šindelková, M., Palovčíková, D., 2000. Karanténní sypavky Mycosphaerella pini a M.<br />

dearnessii. [The quarantine needlecasts Mycosphaerella pini and M. dearnessii] Lesnická<br />

práce 79, 370 – 372.<br />

Jankovský, L., Bednářová, M., Palovčíková, D., 2004. Dothistroma needle blight Mycosphaerella pini<br />

E. Rostrup, a new quarantine pathogen <strong>of</strong> pines in the CR. Journal <strong>of</strong> Forest Science 50,<br />

319-326.<br />

Jankovský, L., Palovčíková, D., Tomšovský, M., 2008. Brown spot needle blight associated with<br />

Mycosphaerella dearnessii occurs on Pinus rotundata in the Czech Republic. New<br />

Disease Reports, 18. .<br />

http://www.bspp.org.uk/ndr/volume18.asp.<br />

Karadzic, D.M., 1989. Scirrhia pini Funk et Parker. Life cycle <strong>of</strong> the fungus in plantations <strong>of</strong> Pinus<br />

nigra Arn. in Serbia. Eur. J. For. Path. 19 (4), 231-236.<br />

Karadzic, D. M., 2004. The distribution, hosts, epidemiology, impact and control <strong>of</strong> fungus<br />

Mycosphaerella pini E. Rostrup apud Munk. in Serbia. Glasnik Šumarskog fakulteta,<br />

Beograd 90, 7 – 35.<br />

Kirisits, T., Cech, T., 2006. Entwickelt sich die Dothistroma-Nadelbräune zu einem<br />

Forstschutzproblem in Österreich? Forstschutz aktuell, 36, 20 – 26.<br />

Kowalski, T., Jankowiak, R., 2008. First record <strong>of</strong> Dothistroma septospora (Dorog.) Morelet in<br />

Poland: a contribution to the symptomology and epidemiology. Phytopatologia Polonica<br />

16, 16-29.<br />

Koltay, A., 1997. New pathogens in Hungarian black pine stands. Novenyvedelem 33 (7), 339 – 341.<br />

Kunca, A., F<strong>of</strong>fova, E., 2000. Ohrozenie porastov borovice čiernej fytokaranténnym patogénom<br />

Dothistroma septospora (Dorog.) Morelet [Threat <strong>of</strong> stands <strong>of</strong> Austrian Pines by<br />

Dothistroma septospora (Dorog.) Morelet]. In: Varinsky J. (ed.): Aktuálne problémy v<br />

ochrane lesa 2000. Proceeding Zvolen, Slovak National Forest Centrum Zvolen, 136–<br />

139.<br />

La Porta, N., Capretti, P., 2000. Mycosphaerella dearnessii, a needle-cast pathogen on mountain pine<br />

(Pinus mugo) in Italy. Plant Disease, 84(8): 922.<br />

Lang, K. J., Karadzic, D., 1987. Dothistroma pini - eine Gefahr für Pinus sylvestris?-<br />

Forstwissenschaftliches - Zentralblatt 106 (1), 45-50.<br />

Macek, J., 1975. Scirrhia pini, the pathogen <strong>of</strong> a new disease <strong>of</strong> Pine in Slovenia. Gozdarski Vestnik,<br />

33, 9 – 11.<br />

Morelet, M., 1968. De Aliquibus in Mycologia Novitatibus (3 e note). Bull. Soc. Sci. Nat. Archeo.<br />

Toulon Var. 177, 9.<br />

Murray, J. S., Batko, S., 1962. Dothistroma pini Hulbary: A new disease on pine in Britain. Forestry<br />

3, 57–65.<br />

Novak-Agbaba, S., Halambek, M., 1997. The most important plant diseases on forest trees in the<br />

coastal region <strong>of</strong> Croatia. In: Proceedings <strong>of</strong> the 10th Congress <strong>of</strong> the Mediterranean<br />

Phytopathological Union, 1997-06-01/05, Montpellier (FR), 67-73.<br />

Pehl, L., 1995. Lecanosticta-Nadelbräune. Eine neue Kiefernkrankheit in der Bundesrepublik<br />

Deutschland. Nachrichtenbl. Deut. Pflanzenschutzd., 47, 305 - 309.<br />

13


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Peterson, G.W., 1982. Dothistroma needle blight <strong>of</strong> pines. Forest Insect and Disease Leaflet 143.<br />

U.S. Department <strong>of</strong> Agriculture Forest Services, Washington DC.<br />

Petrak, F., 1961. Die Lecanosticta Krankheit der Föhren in Osterreich [The<br />

Lecanosticta disease <strong>of</strong>. pines in Austria]. Sydowia 15, 252–256.<br />

Saccardo, P. A., 1920. Mycetes Boreali-Americani. Nuovo Giornale Botanico Italiano 27, 75–88.<br />

Woods, A., Coates, D. and Hamann, A., 2005. Is an Unprecedented Dothistroma Needle Blight<br />

Epidemic Related to Climate Change? BioScience 55, 761-769.<br />

14


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 15-23<br />

TWO NEW SPECIES OF Lophodermium COLONISE SCOTS PINES<br />

NEEDLES IN SCOTLAND<br />

ABSTRACT<br />

Sabrina. N. A. REIGNOUX 1 *, Elizabeth A.E. AITKEN, Sarah GREEN,<br />

Richard A. ENNOS<br />

1 Forest Research, Northern Research Station, Roslin, Midlothian, Scotland EH25 9SY<br />

*s.n.a.reignoux@sms.ed.ac.uk<br />

Previous work has indicated that Scots pine needles are colonized by three species <strong>of</strong><br />

Lophodermium, <strong>of</strong> which two are the endophytes L. pinastri and L. conigenum and the third<br />

is the pathogen L. seditiosum. Recent work on a DNA based Lophodermium phylogeny<br />

found huge variation among L. pinastri isolates which was interpreted as the presence <strong>of</strong><br />

two subspecies within this taxon. In this study we use a combination <strong>of</strong> sequence data,<br />

molecular markers and culture morphology to demonstrate the existence <strong>of</strong> three distinct<br />

taxa within the entity that was previously classified as L. pinastri. These three taxa co-occur<br />

within the native pine woods <strong>of</strong> Scotland<br />

Keywords: Ascomycete, Lophodermium pinastri, Pinus sylvestris, Phylogeny<br />

1. INTRODUCTION<br />

There is increasing evidence that the diverse endophytic communities within the<br />

leaves and needles <strong>of</strong> trees confer protection against pathogens (Petrini, 1991;<br />

Arnold et al., 2003). A situation in which protection by endophytes could be<br />

commercially important is found in the ascomycete genus Lophodermium Chev<br />

which is ubiquitous in pines and is distributed worldwide. Many Lophodermium<br />

species live asymptomatically as endophytes inside the needles <strong>of</strong> pines for at least<br />

part <strong>of</strong> their life cycle (Minter, 1981a; Diwani & Millar, 1987; Wilson, 1995) and<br />

could potentially help to protect against needle cast diseases (Minter, 1981b).<br />

However in order to determine the importance <strong>of</strong> endophytic Lophodermium<br />

species in protecting pines from needle diseases, it is essential to clarify their<br />

taxonomy so that they can be readily identified and their ecology and behaviour<br />

can be studied.<br />

Lophodermium Chev. includes 145 species, mostly from pine hosts, and there is<br />

some degree <strong>of</strong> host specificity (Kirk et al., 2008; Ortiz-Garcia et al., 2003).<br />

Morphological characteristics <strong>of</strong> this genus include a single longitudinal slit<br />

opening <strong>of</strong> the apothecia, and the fusiform shape <strong>of</strong> the ascospores (Darker 1967).<br />

As it is known today, the Lophodermium species complex on P. sylvestris in<br />

Scotland includes two endophytes and one pathogen. The two endophytes differ in<br />

15


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

their ecology. L. pinastri ascocarps are found on naturally shed needles, while L.<br />

conigenum fruits on prematurely killed needles (Minter and Millar, 1980). The<br />

pathogen L. seditiosum causes needlecast disease which is particularly a problem<br />

on young P. sylvestris (Diwani and Millar, 1987).<br />

Preliminary studies <strong>of</strong> Lophodermium isolates made from naturally shed needles in<br />

Scottish native pinewoods showed striking variation among cultures grown on malt<br />

agar. PCR-RFLP and sequence analysis <strong>of</strong> the ITS region in these isolates also<br />

showed greater variability than had hitherto been reported for L. pinastri (Johnston<br />

et al., 2003). In this manuscript we explore the hypothesis that the current taxon L.<br />

pinastri includes more than one species. The hypothesis is tested using data from a<br />

DNA based phylogeny and analysis <strong>of</strong> DNA markers and cultural morphology.<br />

2. MATERIAL AND METHODS<br />

2.1. Fungal Isolates and DNA extraction<br />

L. pinastri was isolated onto malt agar from naturally shed, surface sterilized<br />

needles collected in the Scottish native pine woods at Glen Affric (NH278278) and<br />

Amat (phylogenetic and population genetic analysis) and in Glen Affric, Abernethy<br />

(NJ015155) and Loch Maree (NG995654) for the analysis <strong>of</strong> culture morphology<br />

(Table 1). DNA was extracted from pure cultures grown in liquid Malt Extract<br />

medium using the Plant DNAeasy Qiagen kit.<br />

Table 1: Numbers <strong>of</strong> isolates <strong>of</strong> each clade and population included in the culture<br />

morphology test<br />

Population Clade Ia Clade Ib CladeII<br />

Loch Maree 10 20 3<br />

Glen Affric 15 20 18<br />

Abernethy 20 20 2<br />

Total 45 60 23<br />

2.2. Sequencing and Sequencing Analysis<br />

ITS regions <strong>of</strong> the Ribosomal DNA and partial ACTIN was amplified using<br />

primers ITS1 / ITS4 (White et al., 1990) and ACT-512F / ACT-783R (Carbone and<br />

Kohn 1999) respectively. Sequencing <strong>of</strong> both strands was conducted using the<br />

BigDye® Terminator v3.1 Cycle Sequencing Kit (Applied Biosystem, Foster city,<br />

USA). Sequences were aligned with Clustal X2 (Thompson et al., 1997; Larkin et<br />

al., 2007).<br />

2.3. Phylogenetic Analysis<br />

Phylogenetic analysis was conducted using parsimony in PAUP* 4.0 (Sw<strong>of</strong>ford,<br />

2003) and Baysian inference in MrBayes-3.1.2 (Huelsenbeck and Ronquist, 2001;<br />

16


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Ronquist and Huelsenbeck, 2003). Additional ITS sequences derived from<br />

Genbank were included in a broader ITS phylogeny to look at the relationship<br />

between L. pinastri isolates from Scotland and from other parts <strong>of</strong> the globe.<br />

2.4. Amplified Fragment Length Polymorphism and Corresponding Analysis<br />

AFLP (Amplified Fragment Length Polymorphism) markers were obtained using<br />

the protocol <strong>of</strong> Vos et al. (1995). Eight primer combinations were chosen which<br />

gave 549 markers across all L. pinastri isolates (Table 2). Bands were scored as<br />

present or absent, a distance matrix was calculated using the Jaccard coefficient,<br />

and this was used to conduct a principal coordinate analysis in PAST.<br />

Table 2: Numbers <strong>of</strong> AFLP markers found in each putative Lophodermium species<br />

Primer Combination Total Clade Ia Clade Ib Clade II<br />

AAC/CAA 74 27 23 24<br />

AAC/CG 86 31 33 33<br />

AAC/CT 98 24 39 48<br />

AAC/CC 91 26 33 46<br />

AAC/CAGA 53 26 7 28<br />

AAC/CTA 58 18 10 34<br />

AAC/CCG 60 20 13 32<br />

AAC/CAG 29 12 5 12<br />

2.5. Culture Morphology<br />

Isolates collected as described above from Glen Affric, Loch Maree and Abernethy<br />

(Table 1) were identified to clade on the basis <strong>of</strong> their ITS sequence. Isolates were<br />

then chosen at random within clades and populations. Inoculum measuring 5mm<br />

diameter was cut from cultures and transferred onto a 2% malt agar Petri dish.<br />

Radial growth was measured once a week. The experiment followed a randomised<br />

block design with four blocks and included two replicates <strong>of</strong> each isolate. Results<br />

were analysed using R 2.8.0 (R Development Core Team, 2008)<br />

3. RESULTS<br />

3.1. Phylogenetic and Genetic Marker Analysis<br />

Phylogentic analysis showed a total <strong>of</strong> five monophyletic clades with strong bootstrap<br />

or prior probability support within the Lophodermium complex on Pinus sylvestris.<br />

Isolates derived from shed needles in Scottish pinewoods, and previously classified as<br />

L. pinastri, fell into three <strong>of</strong> these clades named Ia, Ib and II. These clades are<br />

consistent when using both Parsimony and Baysian inference <strong>of</strong> phylogeny on<br />

combined data from the ITS and Actin loci (Fig. 1 and 2). Isolates from these three<br />

clades also form three distinct groupings in the Principal Coordinate analysis based on<br />

genetic distances obtained by scoring AFLP markers.<br />

17


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 1: Phylogeny <strong>of</strong> Lophodermium species isolated from Pinus sylvestris needles from<br />

Scotland by Parsimony criterion phylogeny <strong>of</strong> the ITS. Bootstrap values are annotated on<br />

the branch located before the corresponding node<br />

18


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Figure 2: Phylogeny <strong>of</strong> Lophodermium species isolated from Pinus sylvestris needles from<br />

Scotland by Baysian inference <strong>of</strong> phylogeny <strong>of</strong> combined ITS and Actin<br />

3.2. Culture Morphology<br />

Growth rate differs significantly among the three clades identified in the<br />

phylogenetic and genetic marker analyses (P< 0.001 ANOVA, Fig. 3). Clade Ib<br />

shows the slowest and clade Ia the fastest growth.<br />

19


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 3: Box plot representing growth rate difference between each phylogenetic clades<br />

Figure 4: Eight weeks old colonies <strong>of</strong> two individuals per clade. From left to right: Clade<br />

Ia, Clade II and clade Ib.<br />

20


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

3.3. Relationships among Lophodermium Species<br />

Clade Ia includes Genbank sequences <strong>of</strong> L. pinastri from Europe on Pinus<br />

sylvestris and from Canada on P. strobus (Fig. 5). There are no Genbank<br />

sequences <strong>of</strong> clade Ib except for those derived from the native pinewood at Glen<br />

Affric (Scotland). Clade II includes Genbank sequences <strong>of</strong> L. pinastri from P.<br />

ponderosa in North America and from P. pinaster in New-Zealand. It also<br />

includes a different species <strong>of</strong> Lophodermium, L. kumaunicum, described from<br />

the Himalayas.<br />

Figure 5: Phylogenetic relationship between species <strong>of</strong> Lophdermium Chev.<br />

21


4. DISCUSSION<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Multigene genealogy and genetic marker analysis coupled with culture morphology<br />

data demonstrate that Pinus sylvestris in Scotland is colonized by five species <strong>of</strong><br />

Lophodermium, four <strong>of</strong> which are endophytes. Isolates previously classified in a<br />

single taxon, L. pinastri, fall into three distinct species. At least one <strong>of</strong> these<br />

species, corresponding to clade II, is distributed worldwide. This clade includes<br />

Genebank entries given the name L. pinastri and more recently L. kumaunicum.<br />

Clade II <strong>of</strong> L. pinastri was previously classified as a subspecies in the phylogeny <strong>of</strong><br />

Lophodermium published by Ortiz-Garcia et al. (2003).<br />

The present study has uncovered a greater diversity <strong>of</strong> endophytes than was<br />

previously known. Clarification <strong>of</strong> their taxonomy, and the ability to recognize the<br />

Lophodermium taxa on Scots pine in Scotland will allow us to compare the genetic<br />

diversity, gene flow and mating system <strong>of</strong> each species. Ultimately this will help us<br />

to understand the differences that exist and the interactions that occur between<br />

closely related pathogens and endophytes within Lophodermium.<br />

5. ACKNOWLEDGMENT<br />

We would like to thank BBSRC and Forest Research for funding this project,<br />

Pr<strong>of</strong>essor Mark Blaxter, Dr Graham Stone and Dr Martin Jones for their<br />

contribution in the phylogenetic analysis.<br />

6. REFERENCES<br />

Arnold, A. E., Mejia, L. C., Kyllo, D., Rojas, E. I., Maynard, Z., Robbins, N., et al., 2003. Fungal<br />

endophytes limit pathogen damage in a tropical tree. Proceedings <strong>of</strong> the National<br />

Academy <strong>of</strong> Science <strong>of</strong> the United States <strong>of</strong> America, 100, 15649-15654.<br />

Carbone, I., Kohn, L.M., 1999. A Method for Designing Primer Sets for Speciation Studies in<br />

Filamentous Ascomycetes. Mycologia, 91, 553-556.<br />

Darker, G. D., 1967. A revision <strong>of</strong> the genera <strong>of</strong> the hypodermataceae, 45, 1399-1444.<br />

Diwani, S.A., & Millar, C.S., 1987. Pathogenicity <strong>of</strong> three Lophodermium species on Pinus sylvestris<br />

L. European Journal <strong>of</strong> Forest Pathology, 17, 53-58.<br />

Huelsenbeck, J.P., & Ronquist, F., 2001. MRBAYES: Bayesian inference <strong>of</strong> phylogenetic trees.<br />

Bioinformatics, 17, 754-755.<br />

Johnston, P.R., Park, D., Dick, M.A., Ortiz-Garcia, S., & Gernandt, D.S., 2003. Identifying pineinhabiting<br />

Lophodermium species using PCR-RFLP. New Zealand Journal <strong>of</strong> Forestry<br />

Science, 33, 10-24.<br />

Kirk, P.M., Cannon, F.P., Minter, W.D., & Stalpers, A.J., 2008. Dictionary <strong>of</strong> the Fungi 10th Edition<br />

(10th ed.). CABI Publishing.<br />

Larkin, M.A., Blackshields, G., Brown, N. P., Chenna, R., McGettigan, P. A., McWilliam, H., et al.,<br />

2007. Clustal W and Clustal X version 2.0. Bioinformatics, 23, 2947.<br />

Minter, D.W., 1981a. Lophodermium on pines. Commonwealth Mycological Institute.<br />

22


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Minter, D.W., 1981b. Possible biological control <strong>of</strong> Lophodermium seditiosum. In C. S. Millar (Ed.),<br />

Current Research on Conifer Needle Diseases, pp. 67-74. Aberdeen University Press,<br />

Aberdeen.<br />

Minter, D.W. & Millar, C.S., 1980. Ecology and biology <strong>of</strong> three Lophodermium species on<br />

secondary needles <strong>of</strong> Pinus sylvestris. European Journal <strong>of</strong> Forest Pathology, 10, 169-<br />

181.<br />

Ortiz-Garcia, S., Gernandt, D.S., Stone, J.K., Johnston, P.R., Chapela, I.H., Salas-Lizana, R., et al.,<br />

2003. Phylogenetics <strong>of</strong> Lophodermium from pine. Mycologia, 95 846-859.<br />

Petrini, O., 1991. Fungal endophytes <strong>of</strong> tree leaves. Microbial ecology <strong>of</strong> leaves, 179–197.<br />

R Development Core Team., 2008. R: A language and environment for statistical computing. R<br />

Foundation for Statistical Computing Vienna, Austria. Retrieved from http://www.Rproject.org.<br />

Ronquist, F., & Huelsenbeck, J.P., 2003. MrBayes 3: Bayesian phylogenetic inference under mixed<br />

models. Bioinformatics, 19, 1572–1574.<br />

Thompson, J.D., Gibson, T.J., Plewniak, F., Jeanmougin, F., & Higgins, D.G., 1997. The Clustal X<br />

windows interface: flexible strategies for multiple sequence alignment aided by quality<br />

analysis tools. Nucleic Acids Research, 24, 4876-4882.<br />

Sw<strong>of</strong>ford, D.L., 2003. PAUP–Phylogenetic Analysis Using Parsimony. Ver. 4.0 [Computer S<strong>of</strong>tware<br />

and Manual]. Sinauer Associates, Massachusetts, Sunderland.<br />

Vos, P., Hogers, R., Bleeker, M., Reijans, M., Van De Lee, T., Hornes, M., et al., 1995. AFLP: a new<br />

technique for DNA fingerprinting. Nucleic acids research, 23, 4407-4414.<br />

White, T.J., Bruns, T., Lee, S., & Taylor, J., 1990. Amplification and Direct Sequencing <strong>of</strong> Fungal<br />

Ribosomal RNA Genes for Phylogenetics. In M. A. Innis, D. H. Gelfand, & J. J. Sninsky<br />

(Eds.), PCR Protocols: A Guide to Methods and Applications, Academic Press, San<br />

Diego, pp. 315-325.<br />

Wilson, D., 1995. Endophyte: The Evolution <strong>of</strong> a Term, and Clarification <strong>of</strong> Its Use and Definition.<br />

Oikos, 73, 274-276.<br />

23


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 24-26<br />

RED BAND NEEDLE BLIGHT IN FINLAND, SYMPTOMS AND<br />

DISTRIBUTION<br />

Martti VUORINEN 1*<br />

1 Finnish Forest Research Institute, Suonenjok Research Unit FI-91500, Juntintie, Finland<br />

ABSTRACT<br />

*Martti.vuorinen@metla.fi<br />

Red band needle blight is cused by Mycosphaerella pini Rostrup (1957) also<br />

known as Scirrhia pini A. Funk & A.K. Parker and conidial state is Dothistroma<br />

septosporum (Dorog.) M Morelet. First time it has described as Cytosporina<br />

septospora Dorog. 1911 and later the asexual state has called Septoria septospora<br />

(Dorog.) Arx, Dothistroma pini Hulbary 1941, Dothistroma pini var. keniense<br />

M.H. Ivory 1967, Dothistroma pini var. lineare Thyr & C.G. Shaw 1964,<br />

Dothistroma septospora, Dothistroma septosporum (G. Doroguine) M. Morelet,<br />

Dothistroma septosporum var. keniense (M.H. Ivory) B. Sutton 1980, Dothistroma<br />

septosporum var. lineare (Thyr & D.E. Shaw) B. Sutton 1980, Dothistroma<br />

septosporum var. septosporum (Dorog.) M. Morelet 1968, Eruptio pini (Rostr.)<br />

M.E.Barr 1996. The disease is called red band needle cast or dothistroma needle<br />

cast or dothistroma needle blight or pine needle blight. In Finnish it is called<br />

punavyökariste. Reproduction happens mostly asexual through conidia during the<br />

growing season in moist conditions, in Finland from May to October. Needles <strong>of</strong><br />

all ages are infected by D. septospora. If infection is strong, needles can fall down<br />

during the same growing season, but more <strong>of</strong>ten they can stay on branches to the<br />

next season.<br />

First symptoms <strong>of</strong> red band needle cast are yellow spots on needles which turn<br />

later on brown. Red-brown coloration, commonly associated to the disease is<br />

caused by dothistromin which is a potent and broad–spectrum toxin and is<br />

responsible for the characteristic necrotic lesions and red bands on needles.<br />

Mycosphaerella pini is believed to be native to the cloud forest <strong>of</strong> Central America,<br />

but the first description has made in Russia (Dorogin, 1911). Red band needle<br />

blight has now a days world wide distribution and is most serious disease in Pinus<br />

radiata plantations in Southern Hemisphere, East Africa, New Zealand and Chile<br />

and Pinus contorta in Northern Hemisphere in British Columbia, in Canada is<br />

particularly susceptible.<br />

There are altogether 80 host species: 60 Pinus species and t.e. Picea and<br />

Pseudotsuga.<br />

24


1. INTRODUCTION<br />

Occurrence in Finland<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Red band needle blight is most common in dense 5-15 years stands and infects<br />

needles in branches mostly in 0.5m-2 m high. It can also find in the same young<br />

pine stands, where is or has been Pine needle cast epidemic caused by<br />

Lophodermium seditiosum. Jankowsky (2008) has also observed infections in 60-<br />

80 years stands. In Pinus sylvestris records has made in Hungary, Poland, Zhech<br />

Republic and Estonia, but records were rare until spring 2008 (Jankowsky, 2008)<br />

when has happened rapid outbreak throughout Europe.<br />

Limiting factors to occurrence?<br />

In Finland red band needle blight is recorded first time in autumn 2007 and in<br />

spring 2008 pycnidias and germinated conidia was found and first tree isolates<br />

have done from Hartola, Kangasniemi and Suonenjoki and ITS sequenced. The<br />

sequences are identical with each others to a number <strong>of</strong> M. pini sequences <strong>of</strong> the<br />

GenBank database.<br />

Cold tolerance test has been made in test chambers to the needles, where was<br />

able to find fresh pycnidias: one day interval +5 o C/-5 o C, continuous -10 o C, one<br />

day interval +5 o C /-20 o C, continuous -70 o C in one week testing time and control<br />

samples were stored in cold room in +4 o C. No clear differences could be<br />

observed in the germination <strong>of</strong> conidia between temperatures, which mean that<br />

winter temperature is not a limiting factor. Humid climate favors the outbreak <strong>of</strong><br />

red band needle cast, but dry periods in growing season are a stress to pathogen and<br />

the developing <strong>of</strong> the disease can discontinue.<br />

25


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Observations <strong>of</strong> the distribution <strong>of</strong> Red band needle blight in Finland 2008 by<br />

occasionally selected areas. Observations were not made systematically in whole<br />

country, e.g. south western Finland was outside observation area.<br />

These observations show areas where red band needle cast was noticed in dense<br />

pine stand. Most <strong>of</strong> the observations in 2008 have made in southern and central<br />

Finland, but some observations have made in northern Finland, too. That means,<br />

that red band needle cast has spread out to some extend in whole country.<br />

CONCLUSIONS OF THE DISEASE OUTBREAK ACCORDING TO<br />

OBSERVATIONS 2008<br />

Red band needle blight has been prevalent since 1960s in pines planted as<br />

exotics in plantation forests, particularly in Southern hemisphere, but caused<br />

normally not serious damage to native pine stands. During the last few years the<br />

incidence <strong>of</strong> red band needle blight has increased dramatically in the Northern<br />

Hemisphere and the epidemic in British Columbia, Canada, has caused extensive<br />

mortality <strong>of</strong> pines in their native ranges and the disease has been correlated with<br />

climate change ( Woods et. al., 2005). Climate change could have a positive<br />

impact to the occurrence <strong>of</strong> Dothistroma pini if weather is humid and rainy during<br />

growing season. Drought during growing season is a stress to host and to pathogen<br />

too and the outbreak can stop.<br />

REFERENCES<br />

Jankovský, L., Tomešová, V., Palovčíková, D., Dvořák, M., Bednářová, M., 2008. Dothistroma<br />

needle blight -- Dothistroma septospora epidemic on Scots pine?. In Methodology <strong>of</strong><br />

Forest Insect and Disease Survey in Central Europe. Abstrakt book. IUFRO Working<br />

Party 7. 03. 10. 1. vyd. Tatranské lomnica: IUFRO Working Party 7.03.10, 2008.<br />

Müller, M.M., Hantula, J. and Vuorinen M., <strong>2009</strong>. First Observations <strong>of</strong> Mycosphaerella pini on<br />

Scots Pine in Finland. Plant Disease ,March <strong>2009</strong>, Volume 93, Number 3<br />

Page 322<br />

Woods, A.K., Coates, D. and Hamann, A., 2005. Is an Unprecedented Dothistroma Needle Blight<br />

Epidemic Related to Climate Change? BioScience, Vol. 55, No. 9, pp. 761-769, 09/2005<br />

26


Scleroderris Canker<br />

27


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

28


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 29-32<br />

THE OCCURRENCE OF MICROCONIDIA ON Gremmeniella abietina<br />

(LAGERB.) MORELET<br />

ABSTRACT<br />

Antti UOTILA 1*<br />

1 Hyytiälä Forestry Field Station, University <strong>of</strong> Helsinki, Finland<br />

*antti.uotila@helsinki.fi<br />

The Gremmeniella abietina microconidia were observed during 1981-2002 in<br />

numerous finnish specimen and pure cultures <strong>of</strong> the fungus. Microconidia existed in<br />

type A and B on Gremmeniella abietina. Microconidia were found in pycnidia and<br />

apothecia. Grey microconidial mass was produced also in pure cultures. Size <strong>of</strong><br />

microconidium was 3-6 µm X 1 µm. The germination was not observed. The attempts<br />

to produce apothecia with the help <strong>of</strong> microconidia in lab failed.<br />

1. INTRODUCTION<br />

Finn and Helka Roll-Hansen first reported microconidia in 1973. They found<br />

microconidia in pycnidia in Norwegian samples <strong>of</strong> G. abietina. The colour <strong>of</strong><br />

microconidial mass was grey instead <strong>of</strong> pink as conidial mass. Bergdahl and<br />

Tsajkowski (1982) reported microconidia in North America. According to them<br />

microconidia rarely germinate. Microconidia are also described in taxonomical<br />

work <strong>of</strong> Petrini et al., 1989, when microconidia were found on different races or<br />

types <strong>of</strong> G. abietina. My purpose was to check the occurrence <strong>of</strong> microconidia<br />

in two types <strong>of</strong> G. abietina existing in Finland. In addition apothecia were tried<br />

to produce by transferring microconidia.<br />

2. MATERIAL AND METHODS<br />

These results based on notes which I have done on different experiments and<br />

cultures <strong>of</strong> numerous G. abietina isolates during 1981-2002. Hundreds <strong>of</strong><br />

microscope slides were examined. They were done from natural pycnidia or<br />

apothecia and also from pure cultures in laboratory. The types <strong>of</strong> G. abietina<br />

were determined by conidial septa and disease symptoms. Later the types were<br />

confirmed with fatty acid (Müller & Uotila, 1997) or RAMS method (Hantula<br />

&Müller, 1997; Uotila et al., 2000). After determining mating alleles <strong>of</strong> G.<br />

abietina (Uotila, 1992) the pairing experiment with microconidia was done. G.<br />

abietina has two mating alleles, mat 1 and mat 2. The microconidia from the<br />

monospore culture were transferred to culture having different mating allele.<br />

The cultures were grown in erlenmeyer bottles on barley groats +pine needles<br />

medium.<br />

29


3. RESULTS AND DISCUSSION<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

The size <strong>of</strong> microconidia is 3-6 µm X 1 µm (Fig. 1 and 2). Microconidium<br />

contain one nucleus. Microconidia were found regularly in types A and B <strong>of</strong><br />

Gremmeniella. Microconidia are produced also in pure cultures <strong>of</strong><br />

monoascospore isolates. The color <strong>of</strong> microconidial mass in pure culture or in<br />

pycnidia is grey instead <strong>of</strong> pink color <strong>of</strong> macroconidial mass. I have not seen<br />

germinating microconidia. The microconidia were not germinating in<br />

monospore cultures.<br />

Figure 1. Microconidia formation in pure culture <strong>of</strong> Gremmeniella abietina.<br />

Stained with anilin blue.<br />

The microconidia are formed from mycelia (Fig. 1 and 3.), conidia (Fig. 2.)<br />

or ascospores. When the contents <strong>of</strong> pycnidium with microconidia and<br />

macroconidia is spread on the agar the microconidia adhered close to<br />

macroconidia (Fig. 2.). The change <strong>of</strong> nuclei is possible in these conditions. The<br />

role <strong>of</strong> microconidia is propably to transfer the nuclei before meiosis. This<br />

sounds reasonable, but why there are microconidia also in apothecia? I tried to<br />

produce apothecia in lab with help <strong>of</strong> microconidia. G. abietina is heterothallic<br />

with two mating alleles. So the microconidia were added in cultures with<br />

different mating allele, but apothecia did not appear in these cultures. In nature<br />

we can produce apothecia by inoculating compatible mycelia close to each other<br />

in the same seedling.<br />

30


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Figure 2. The macroconidium seems to produce microconidia.<br />

Figure 3. Germinated macroconidia and the mycelia forming microconidia.<br />

31


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

It is still needed more research to describe the exact process <strong>of</strong> pairing and so<br />

to understand the role <strong>of</strong> microconidia for the fungus. This knowledge do not<br />

<strong>of</strong>fer direct possibilities to control the disease, but it is important to understand<br />

the enemy.<br />

4. REFERENCES<br />

Hantula, J. & Müller, M., 1997. Variation within Gremmeniella abietina in Finland and other<br />

countries as determined by Random Amplified Microsatellites (RAMS). Mycol. Res.<br />

101: 169-175.<br />

Müller, M. & Uotila,A., 1997. The diversity <strong>of</strong> Gremmeniella abietina var. abietina FAST<br />

pr<strong>of</strong>iles. Mycological Research 101 (5): 557-564.<br />

Petrini, O., Petrini, L. E., Laflamme, G., Ouellette, G. B., 1989. Taxonomic position <strong>of</strong><br />

Gremmeniella abietina and related species: a reappraisal. Can. J. Bot. 67:2805-2814.<br />

Roll-Hansen, F. & Roll-Hansen, H., 1973. Microconidia formed by Scleroderris lagerbergii. 2nd<br />

International congress <strong>of</strong> Plant Pathology. University <strong>of</strong> Minneapolis, MN.<br />

Uotila, A., 1983. Physiological and morphological variation among finnish Gremmeniella<br />

abietina isolates. Commun. Inst. For. Fenn. 119, 12 p.<br />

Uotila, A., 1992. Mating system and apothecia production in Gremmeniella abietina. Eur. J. For.<br />

Path. 22: 410-417.<br />

Uotila, A., Hantula, J., Väätänen, A-K. & Hamelin, R., 2000. Hybridization between two biotypes<br />

<strong>of</strong> Gremmeniella abietina var. abietina in artificial pairings. Eur. J. For. Path. 30:211-<br />

219.<br />

Zajkowski, S. J. & Bergdahl, D. R., 1982. Observations on a microconidiospore stage associated<br />

with Gremmeniella abietina. Phytopathology 72: 267.<br />

32


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 33-38<br />

CENTRAL NEWFOUNDLAND: ESCAPE from QUARANTINE<br />

Gary R. WARREN 1* and Gaston LAFLAMME 2 .<br />

1 Natural Resources Canada, C.F.S. - Canadian Wood Fibre Centre, Corner Brook, NL Canada<br />

A2H 6J3<br />

2 Natural Resources Canada, C.F.S. - Laurentian Forestry Centre, Quebec, QC. Canada G1V 4C7<br />

ABSTRACT<br />

* Gary.Warren@NRCan-RNCan.gc.ca<br />

Scleroderris canker, European race, was first detected on Austrian pine in St. John’s,<br />

Newfoundland in 1979. To prevent spread <strong>of</strong> this exotic disease, a quarantine zone was<br />

established in 1980 to all areas north <strong>of</strong> the Witless Bay Line. Later, red pine mortality near<br />

Torbay (1981), Upper Island Cove and along Salmonier Line (1996) resulted in extending<br />

the quarantine zone in 1998 to all areas east <strong>of</strong> Route #202 at the isthmus <strong>of</strong> the Avalon<br />

Peninsula. Infection on these pines was tracked back to planting stock produced at the Back<br />

River Nursery on Salmonier Line. These seedlings were planted on the Avalon and<br />

Bonavista Peninsulas from1937 to1952. Until 2007, the slow rate <strong>of</strong> spread and natural<br />

quarantine boundary limited this disease for over 60 years to the Avalon Peninsula. In<br />

2007, the European race <strong>of</strong> Scleroderris canker was detected in an isolated red pine<br />

plantation in central Newfoundland at Berry Hill Pond, 400km outside <strong>of</strong> the quarantine<br />

zone. Field observations showed that conducive conditions for the pathogen were always<br />

present in the area, explaining rapid development <strong>of</strong> the epidemic compared to slow<br />

progression in plantations on the Avalon Peninsula. Failure to publicize and enforce the<br />

quarantine and apply preventative control measures has now resulted in threats to native red<br />

pine stands and plantations established throughout central Nfld. Pruning red pines in that<br />

region will prevent any new outbreak. We cannot rely on quarantine measures alone to<br />

prevent spread <strong>of</strong> this disease.<br />

Keywords: Scleroderris canker, Gremmeniella abietina, Pinus resinosa, outbreak,<br />

quarantine.<br />

1. INTRODUCTION<br />

Scleroderris canker cause by Gremmeniella abietina (Lagerb.) Morelet is a serious<br />

disease <strong>of</strong> hard pines, causing shoot blight, branch dieback, stem cankers and tree<br />

mortality. Two races <strong>of</strong> the disease affect pines in North America, the native North<br />

American (NA) race and the introduced European (EU) race (Dorworth et al., 1977)).<br />

The NA race causes infection on lower branches in the snow, inducing dieback and<br />

mortality in pine seedlings or pines less than 2 m in height. The NA race has never<br />

been found in Newfoundland. The EU race, introduced in North America, is a very<br />

serious disease; it is not restricted at the snow level; the whole crown <strong>of</strong> large trees can<br />

be affected. Red pine (Pinus resinosa Ait.) is very susceptible to this disease (Skilling,<br />

1975); Scots (Pinus sylvestris L.) and Austrian pines (Pinus nigra Arnold) are<br />

moderately affected (Bernhold et al., <strong>2009</strong>) while jack pine (P. banksiana Lamb.) is the<br />

most resistant (Laflamme and Blais, 2000).<br />

33


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

2. FIRST REPORTS OF THE DISEASE IN NEWFOUNDLAND<br />

Scleroderris canker, EU race, was first detected on Austrian pine in St. John’s,<br />

Newfoundland, in 1979 (Singh et al., 1980). To prevent spread <strong>of</strong> this exotic disease, a<br />

quarantine zone was established in 1980 to all areas north <strong>of</strong> the Witless Bay Line<br />

(Figure 1). No movement <strong>of</strong> conifer stock from the area north <strong>of</strong> Witless Bay Line was<br />

allowed out <strong>of</strong> this zone. An information pamphlet was produced and distributed. It<br />

showed photos <strong>of</strong> symptoms to help identification <strong>of</strong> the disease as well as a map <strong>of</strong> the<br />

hazard and quarantine zone for Scleroderris canker. Within this quarantine zone, severe<br />

infection followed by tree mortality <strong>of</strong> red pine was detected in 1981; it was found in<br />

the Torbay plantation located 15 km north <strong>of</strong> St. John's (Figure 1).<br />

3. BACK RIVER NURSERY<br />

In mid 1980's, Scots pine dieback and mortality caused by Scleroderris canker<br />

EU race was observed along Salmonier Line and at a site called the old “Back<br />

River Nursery”. This tree nursery has an important role in understanding the<br />

introduction <strong>of</strong> the disease on the Island, as we will see later. Back River Nursery<br />

was outside the quarantine zone (Figure 2).<br />

In early 1990's Scleroderris canker infection and limited mortality was observed<br />

in a number <strong>of</strong> mixed pine plantations along the southern shore <strong>of</strong> Conception Bay,<br />

again outside the quarantine zone. Planting stock for all affected plantations came<br />

from the Back River Nursery.<br />

In 1996, Scleroderris canker was observed in a red pine plantation at Upper<br />

Island Cove and the whole plantation was completely destroyed by the disease.<br />

Figure 1: Quarantine zone north <strong>of</strong> Witless Bay Line established in 1980 to prevent<br />

spread <strong>of</strong> Scleroderris canker outside this area.<br />

34


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Figure 2: In1998, the quarantine zone for Scleroderris canker, European race, was<br />

extended to all areas east <strong>of</strong> Route #202, a natural barrier for the Avalon Peninsula.<br />

Following this extension <strong>of</strong> the disease, the quarantine zone was extended in<br />

1998 to all areas east <strong>of</strong> Route #202, a natural barrier for the Avalon Peninsula<br />

(Figure 2). Unfortunately, no public information or pamphlet was given or<br />

produced<br />

After these new findings, the senior author did a search <strong>of</strong> historical data on the<br />

Back River Nursery, to have information on the provenance <strong>of</strong> seed or seedlings in<br />

this nursery and the location <strong>of</strong> plantations established with seedlings produced in<br />

this nursery. A report on Reforestation in Newfoundland, from 1937 to 1952,<br />

summarized the establishment and activities <strong>of</strong> the Back River Nursery (Doyle,<br />

1967). It gives details <strong>of</strong> planting stocks produced, and identified 16 plantations<br />

established on the Avalon, Burin and Bonavista Peninsulas between 1938 and<br />

1951. This report indicated also that all stocks at the Back River Nursery were<br />

produced from seeds: white spruce (Picea glauca (Moench) Voss), white pine<br />

(Pinus strobus L.) and balsam fir (Abies balsamea (L.) P.Mill.) were <strong>of</strong> local origin<br />

while Norway spruce (Picea abies (L.) Karst.), red pine, Scots pine and jack pine<br />

came from Ontario, Canada. It is important to note that Scleroderris canker cannot<br />

be transmitted by seed. Another report provided a detailed historical account <strong>of</strong> the<br />

Back River Nursery (Baker and Miller-Pitt, 1998). In 1939 the Newfoundland<br />

Forestry Division received a gift <strong>of</strong> 30,000 seedlings <strong>of</strong> red, white, jack and Scots<br />

pines from the Province <strong>of</strong> Ontario, Canada.<br />

With this information on hand, a survey <strong>of</strong> localized plantations was done in 1998.<br />

The 16 plantations established with Back River Nursery stock, revealed that most<br />

plantations were in very poor condition, but Scleroderris canker EU race was present in<br />

a number <strong>of</strong> them. The majority <strong>of</strong> red pine had died but the disease was still present<br />

and viable in the surviving Scots and jack pines which show resistance to the disease<br />

35


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

(Bernhold et al., <strong>2009</strong>; Laflamme and Blais, 2000). Even if these pines are surviving<br />

with the disease for several years, the host persists as an endemic carrier <strong>of</strong> the<br />

pathogen. Infected plantations outside the Avalon Peninsula, at Sunnyside and on the<br />

Bonavista Peninsula, were sanitized with expectations <strong>of</strong> limiting the disease to the<br />

Avalon Peninsula, and keeping the 1998 quarantine zone in place. The quarantine and<br />

natural barrier at the isthmus <strong>of</strong> the Avalon Peninsula was successful in limiting the<br />

spread <strong>of</strong> Scleroderris canker EU race for over 60 years (Figure 2).<br />

4. OUTBREAK IN SOUTH-CENTRAL NEWFOUNDLAND<br />

In 2007, severe red pine mortality was reported in an 18-year-old red pine<br />

plantation at Berry Hill Pond, 90 km down the Bay D'Espoir highway from the<br />

Trans Canada Highway, approximately 400 km outside the quarantine zone.<br />

(Figure 3). G. abietina was isolated and molecular tests have proven the disease to<br />

be Scleroderris canker EU race. Damage assessment done in<br />

Figure 3: Outbreak in Central Newfoundland, 400 km from the quarantine zone for<br />

Scleroderris canker, European race on the Avalon Peninsula.<br />

2007 and 2008 revealed the disease had been present for 7-8 yr, building up<br />

inoculum by infecting lower branches in a center <strong>of</strong> infection. Then, optimum<br />

conditions conducive for spore release, dispersal and infection occurred 2-3 years prior<br />

to 2007 resulting in killing pines in the center <strong>of</strong> infection by shoot infection <strong>of</strong> the<br />

whole crown. This high spore load in the top <strong>of</strong> crown exposed to wind, rain and snow<br />

helped the disease to spread on top <strong>of</strong> surrounded healthy pines. Spores from these tops<br />

36


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

were rain splashed and disseminated on lower shoots, killing the residual trees the<br />

following year.<br />

Planting stock for this plantation was produced in the Wooddale provincial tree<br />

nursery in central Newfoundland. This plantation and several others in the forest<br />

management district were established in 1989 using the same red pine planting<br />

stock. An inspection in the fall <strong>of</strong> 2007 <strong>of</strong> the other plantations showed no other<br />

signs <strong>of</strong> infection, ruling out infected planting stock as the source <strong>of</strong> the disease.<br />

Geographic isolation <strong>of</strong> this plantation, surrounded primarily by bog and scrub<br />

forest with no pine content, has all but ruled out natural spread <strong>of</strong> the disease from<br />

the Avalon Peninsula. Forestry personnel from the Bay D'Espoir <strong>of</strong>fice commented<br />

that locations along the road through the plantation were common camp sites for<br />

moose hunters from the Avalon Peninsula in the late 1980's early 1990's. Two tall<br />

communication towers along the road to the plantation were easy landmarks for<br />

hunters to locate the campsites. It is suspected that hunting groups from the Avalon<br />

Peninsula brought their own kindling and firewood with them, which would have<br />

been readily available from recently killed red pine in the infected Conception Bay<br />

plantations.<br />

5. DISCUSSION<br />

After 15 years <strong>of</strong> observations on the development <strong>of</strong> a Scleroderris canker<br />

epidemic, EU race, in 50 red pine plantations located in Quebec, we can sum up the<br />

results into three steps:<br />

1- There is a build up <strong>of</strong> inoculum in a centre <strong>of</strong> infection.<br />

2- Followed by a spread <strong>of</strong> infection on lower branches in the snow over a large<br />

area.<br />

3- Finally, under conducive climatic conditions, conidia <strong>of</strong> G. abietina produced<br />

on lower branches will spread the disease higher in the crown <strong>of</strong> trees.<br />

In south-central Newfoundland, we observed a different pattern. The epidemic<br />

started in a centre <strong>of</strong> infection and killed the trees in that centre relatively rapidly.<br />

The large number <strong>of</strong> infected shoots in the crown <strong>of</strong> dead and dying trees produced<br />

a large amount <strong>of</strong> inoculum in the upper part <strong>of</strong> trees. The strong wind prevailing in<br />

that region, with rain and snow had spread the disease to the top <strong>of</strong> surrounding<br />

trees, causing their death in less than 2 to 3 years. The disease became so<br />

widespread that nothing could be done to save the plantation. Because <strong>of</strong> the<br />

climatic conditions favourable to the disease in that region, the other red pine<br />

plantations should be pruned to prevent any build up <strong>of</strong> inoculum in the eventuality<br />

<strong>of</strong> an introduction <strong>of</strong> G. abietina.<br />

From the historical reports, seedlings <strong>of</strong> pine were imported into the Island <strong>of</strong><br />

Newfoundland from Ontario, Canada. It is difficult to conclude that the<br />

introduction <strong>of</strong> the disease came with this nursery stock: the European race was not<br />

37


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

found in Ontario until 1985 and the disease had been present in Back River<br />

Nursery several years before. So the origin <strong>of</strong> the introduction <strong>of</strong> the disease on the<br />

island is still unknown. Based on preliminary molecular study, the Newfoundland<br />

introduction would be different than the one on continental North America;<br />

moreover, the pathogen population in Nfld. shows more relationship with the<br />

pathogen population from Europe (Hamelin et al., 1998). The introduction could<br />

have come from Austrian pine seedlings imported from Europe; this pine species<br />

has been planted in St. John’s area for many years. This tree species was never<br />

imported for reforestation; it was imported as ornamental probably not too long<br />

after the establishment <strong>of</strong> the Back River Nursery, but this hypothesis remains to be<br />

proven.<br />

The breach <strong>of</strong> the quarantine zone now places the natural red pine stands and<br />

numerous plantations in central Newfoundland region at a serious risk <strong>of</strong><br />

extinction. A communication plan to the public should be undertaken through<br />

information, pamphlets, display notices, roadside signs and kiosk.. It is necessary<br />

to maintain the Scleroderris canker quarantine on the Avalon Peninsula. The<br />

isthmus <strong>of</strong> the Avalon Peninsula is a natural barrier with limited alternate access,<br />

where quarantine notices, signage and material deposition depots can be setup.<br />

6. LITERATURE CITED<br />

Baker, M., Miller-Pitt. J., 1998. By wise and prudent measures: The development <strong>of</strong> <strong>forestry</strong> on<br />

Salmonier Line. Newfoundland & Labrador, Department Forest Resources & Lands<br />

Publication. 65 p.<br />

Bernhold, A., Hansson, P., Rioux, D., Simard, M., Laflamme, G., <strong>2009</strong>. Resistance to Gremmeniella<br />

abietina (EU race, large tree type) in introduced Pinus contorta and native Pinus<br />

sylvestris in Sweden. Canadian Journal <strong>of</strong> Forest Research 39, 89-96.<br />

Doyle, J.A., 1967. Report <strong>of</strong> reforestation in Newfoundland 1937-1952. Nfld. Dept. Forestry, File<br />

Report. 18 p. (Nfld. Dept. For. Library File #002509), Corner Brook, Newfoundland.<br />

Dorworth, C.E., Krywienczyk, J., Skilling, D.D., 1977. New York isolates <strong>of</strong> Gremmeniella abietina<br />

(Scleroderris lagerbergii) identical to immunogenic reaction to European isolates. Plant<br />

Disease Reporter 61, 887-890.<br />

Hamelin, C.R., Lecours, N., Laflamme, G., 1998. Molecular evidence <strong>of</strong> distinct introduction <strong>of</strong> the<br />

European race <strong>of</strong> Gremmeniella abietina into North America. Phytopathology 88, 582-<br />

588.<br />

Laflamme, G., Blais, R., 2000. Resistance <strong>of</strong> Pinus banksiana to the European race <strong>of</strong> Gremmeniella<br />

abietina. Phytoprotection 81, 49-55.<br />

Myren, D.T., Davis, C.D., 1986. European race <strong>of</strong> Scleroderris canker found in Ontario. Plant<br />

Disease 70, 475.<br />

Singh, P., Dorworth, C.E., Skilling, D.D., 1980. Gremmeniella abietina in Newfoundland. Plant<br />

Disease 64, 1117-1118.<br />

Skilling, D.D., 1975. The development <strong>of</strong> a more virulent strain <strong>of</strong> Scleroderris lagerbergii in New<br />

York State. European Journal Forest Pathology 7, 297-302.<br />

38


Shoot Blights<br />

39


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

40


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 41-47<br />

CONE DAMAGES BY Diplodia pinea AND SEED BORING INSECTS ON<br />

Pinus pinea L. (ITALIAN STONE PINE) IN CENTRAL ITALY<br />

Matteo FEDUCCI 1 , Alessia PEPORI 1 , Daniele BENASSAI 2 , Martina CAMBI 1 ,<br />

Paolo CAPRETTI 1*<br />

1 DiBA – Department <strong>of</strong> Agricultural Biotechnology Sect. <strong>of</strong> Plant Pathology. Piazzale delle<br />

Cascine, 28 – 50144 – FIRENZE - ITALY<br />

2 DiBA – Department <strong>of</strong> Agricultural Biotechnology, Sect. <strong>of</strong> Entomology Via Maragliano, 77 –<br />

50144 – FIRENZE - ITALY (at the time <strong>of</strong> research)<br />

ABSTRACT<br />

* paolo.capretti@unifi.it<br />

In Italy Pinus pinea L. (Italian Stone Pine) is cultivated for several purposes but mainly<br />

for dune protection, landscape conservation, tourism, and seeds production. This economic<br />

activity is annually negatively affected by the occurrence <strong>of</strong> fungi, insects, and abiotic<br />

agents but also by a new alien pest (Leptoglossus occidentalis Heidemann) recently<br />

introduced from North America. In order to clarify the main reasons <strong>of</strong> seed losses,<br />

monitoring investigation have been organized in different pinewood located along the<br />

Tyrrhenian cost in Tuscany. During the period 2006 – 2008 cones, having different ages<br />

(from 1 to 3 years old) were analyzed. The main agents responsible <strong>of</strong> biotic damage were<br />

identified and their incidence on losses was ranked.<br />

The study showed that: - about 64% <strong>of</strong> 1-yr old and 36% <strong>of</strong> 2-yrs old cones were<br />

affected by D. pinea; - more than 80% <strong>of</strong> old cones collected on the ground in pinewoods<br />

were colonized by the fungus; - up to 57% <strong>of</strong> old infected cones produced available<br />

inoculum, able to germinate in 6h at 25°C. The most consistent damages on the harvested<br />

mature cones were caused by D. pinea, followed by small mammals and insects including<br />

L. occidentalis, although differences among forest occurred. In infected cones by the fungus<br />

seed losses were up to 64%.<br />

Keywords: Pinus, edible seeds, insects, diseases, fungi<br />

1. INTRODUCTION<br />

Pinus pinea L., the Italian Stone Pine with its umbrella-shape crown is one <strong>of</strong><br />

the most characteristic and attractive Mediterranean trees. It is usually grown in<br />

pure pinewood plantation along the coasts, or mixed with other Mediterranean<br />

species in temperate forests. Often it is also cultivated as an ornamental tree, in<br />

gardens and public parks areas <strong>of</strong> Southern Europe (Bernetti, 1995).<br />

P. pinea is also well known for its edible seeds (pinoli) about two cm long that<br />

quite <strong>of</strong>ten constitute the main income for pinewoods owners (Bernetti, 1995;<br />

Ducci et al., 2001; Saporito, 2002). Seeds are produced by pines after a long<br />

process <strong>of</strong> maturation. The cones ripen after three years expanding to oval-shaped,<br />

300 gr. weigh. During that long period they are exposed to several biotic threats<br />

41


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

due primarily to insects Ernobius impressithorax, (Coleoptera,Anobiidae)<br />

(Dioryctria mendacella Stgr. (Lepidoptera, Pyralidae), Pissodes validirostris Gyll.<br />

(Coleoptera, curculionide), small mammals (squirrels and dormouse)<br />

(CanakcioglU, 1969; Roques, 1983; El Hassani and Messaoudi, 1986; Innocenti<br />

and Tiberi, 2002) and fungi, mainly Diplodia pinea (Petri, 1917; Petri and Adani,<br />

1916; Verona, 1950; Maresi et al., 2002). More recently a new alien insect,<br />

Leptoglossus occidentalis Heidemann (Heteroptera Coreidae), was accidentally<br />

introduced in Italy (Bernardinelli and Zandigiacomo, 2001) and contribute to<br />

enhance the amount <strong>of</strong> seed damages.<br />

Since the 2006-2007 the amount <strong>of</strong> seed production dropped down causing<br />

serious economic concern for the market and several survey activities were<br />

organized in order to evaluate the main causes <strong>of</strong> seed losses.<br />

2. MATERIALS AND METHODS<br />

The studies on the reduction <strong>of</strong> seed production from P. pinea trees were carried<br />

out in Tuscany, Central Italy during the period 2006 – 2008 in different pine<br />

forests. Two areas were located in the North, close to Pisa, in the Regional Park<br />

“Parco Migliarino S. Rossore Massaciuccoli” MSRM-1 (Migliarino) and MSRM-2<br />

(Tirrenia), while the third was situated in the South <strong>of</strong> region, close to Grosseto, in<br />

the “Parco Naturale della Maremma” PNM (Alberese). All the mentioned areas are<br />

particularly large pinewood forests from where every year cones are usually<br />

harvested and seeds are send to the market.<br />

The study was organized following different aspects:<br />

2.1. - Surveying on the Occurrence <strong>of</strong> D. pinea on cones <strong>of</strong> different age.<br />

Pinewood areas MSRM-1 and MSRM-2.<br />

a) Occurrence <strong>of</strong> D. pinea on immature 1- 2 yr-old cones. During the winter<br />

2006 five trees were felled and 150 green cones (n.100 1yr-old and n.50 2yr<br />

old samples) were randomly collected from the crown. Cones were then put<br />

in moist chambers up to 15 days and regularly checked to detect the<br />

occurrence <strong>of</strong> D. pinea (Feducci, 2007).<br />

b) Occurrence <strong>of</strong> D. pinea on mature and old cones lying on the ground.<br />

Pinewood area MSRM-1. About 600 cones were collected under the pine<br />

crown in order to detect the occurrence <strong>of</strong> the fungus. Cones were ranked<br />

according to the age (1, 2 >2 yrs) and percentage <strong>of</strong> scales cover by pycnidia<br />

(< 15% ; 15 – 50 ; > 50%) (Pepori, 2006).<br />

2.2. Detecting the inoculum availability from mature cones on the ground.<br />

Pinewood area MSRM-1. A sample <strong>of</strong> 135 cones was collected from 5 different<br />

particles under the crown <strong>of</strong> 27 pines. Cones were processed according to Munck<br />

and Stanosz (<strong>2009</strong>). The method was modified and adjusted to the size <strong>of</strong> cones<br />

(350 ml <strong>of</strong> water was used in rinsing the cones). A conidia suspension was<br />

42


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

obtained for each cone and used for germinability test on 2% Malt-agar in Petri<br />

dishes at 25°C for 6 hours (Cambi, 2008).<br />

2.3. Assessing the sanitary conditions <strong>of</strong> harvested mature cones (3 yrs-old).<br />

Pinewood areas MSRM-1, MSRM-2, and PNM. In order to evaluate the occurrence<br />

<strong>of</strong> damage on cones ready to the extraction process, about 3000 cone samples were<br />

collected (1000 from each previously mentioned area). Symptoms <strong>of</strong> damages, if<br />

present, were visually detected and classified according their main causes: insects<br />

(Dioryctria sp., Ernobius impressithorax, Pissodes validirostris), L. occidentalis,<br />

small mammals (Dormouse, Glis glis; Squirrel, Sciurus vulgaris), Diplodia pinea<br />

and abiotics. Total observed: 3045 cones (Migliarino 1045, Tirrenia 1000, Alberese<br />

1000).<br />

2.4. Test seed viability. Pinewood areas MSRM-1, MSRM-2, and PNM. From<br />

a sample <strong>of</strong> 100 cones, 80 affected by D. pinea and 20 apparently healthy as<br />

control (see 2.3) all seeds were extracted and than checked for their viability<br />

conditions, first visually and later by using the Tetrazolium Test (International<br />

Seed Testing Association., 2003; 2007; Feducci, 2007).<br />

3. RESULTS<br />

3.1. Occurrence <strong>of</strong> D. pinea on cones <strong>of</strong> different age.<br />

a) Immature 1-2 yr-old cones. D. pinea is a fungal parasite quite common in<br />

young pine cones. After evaluation about 64% <strong>of</strong> 1-yr old and 36% <strong>of</strong> 2-yrs old<br />

cones were affected by the fungus. Percentages <strong>of</strong> apparently healthy cones<br />

increase according to their age (Table 1).<br />

Table 1. Occurrence <strong>of</strong> Diplodia pinea on Pinus pinea cones <strong>of</strong> different age.<br />

Cone conditions Apparently<br />

healthy<br />

1 year-old cones 2 years- old cones<br />

Affected by<br />

D. pinea<br />

43<br />

Apparently<br />

Healthy<br />

Affected by<br />

D. pinea<br />

data in % 36.6±13.9 63.4±13.9 66.7±4.3 33.3±4.5<br />

b) Mature and old cones lying on the ground. Observations on cones on the<br />

ground under the crown <strong>of</strong> pines show that the number <strong>of</strong> old ones was quite<br />

relevant. The percentage <strong>of</strong> cones having more than 50% <strong>of</strong> scales covered by D.<br />

pinea was also consistent (Table 2).<br />

Table 2. Occurrence <strong>of</strong> pycnidia <strong>of</strong> Diplodia pinea on Pinus pinea cones.<br />

Cone<br />

Age <strong>of</strong> cones Occurrence <strong>of</strong> D. pinea on cones*<br />

conditions 1 yr 2 yrs >2 yrs 0 1 2 3<br />

Data in %<br />

±st<br />

6,7±0,8 12,1±5,6 81,2±4,0 5,9±2,1 7,4±1,1 15,0±0,2 71,3±3,<br />

2


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

* Occurrence <strong>of</strong> Diplodia pinea. 0= none or few pycnidia; 1 = pycnidia detected<br />

on < 15% <strong>of</strong> cone scales ; 2 = 15 – 50 % ; 3 = > 50%.<br />

3.2. Detecting the inoculum availability from mature cones on the ground.<br />

A large amount <strong>of</strong> cones colonized by D. pinea, and showing pycnidia (Table 3),<br />

were still actively releasing conidia. Among them 55.6% <strong>of</strong> 1-yr old cones and<br />

63.9% <strong>of</strong> >3-yrs cones were still producing inoculum. Percentage <strong>of</strong> conidia<br />

germination after 6 h at 25°C was 47.4%.<br />

Table 3. Inoculum availability <strong>of</strong> Diplodia pinea (pycnidia) on Pinus pinea cones.<br />

Ages <strong>of</strong> cones<br />

Percentage <strong>of</strong> cone scales showing D. pinea pycnidia<br />

50<br />

1-yr 15.6±0,8 66.7±5,6 17.8±4,0<br />

2-yrs 0.0 42.2±5,6 57.8±4,0<br />

>3-yrs 35.6±0,8 53.3±5,6 11.1±4,0<br />

3.3. Sanitary conditions <strong>of</strong> harvested mature cones (3 yrs-old).<br />

Data obtained from cones ready for seed extraction showed that damages<br />

ranged from 28-55% according to the different forest areas considered. The most<br />

consistent damages were caused by D. pinea (average 16%) followed by small<br />

mammals and insects (about 6%) and L. occidentalis (4%) (Table 4). Differences<br />

values among forests were quite pronounced particularly for the fungal parasite<br />

(Figure 1).<br />

Table 4. Main damages observed on 3-yrs old Pinus pinea cones.<br />

Sampling<br />

area Healthy<br />

Cone conditions Causes <strong>of</strong> damage<br />

(n.1900)<br />

Damaged<br />

(n.1145)<br />

Insects*<br />

44<br />

L.<br />

occid.<br />

Small<br />

mam.**<br />

D.<br />

pinea<br />

Other<br />

causes<br />

MSRM-1 71.1 28.9 6.3 2.5 2.5 14.5 3.1<br />

MSRM-2 44.8 55.2 4.5 8.0 1.3 32.0 9.4<br />

PNM 70.9 29.1 7.5 3.2 15.6 2.4 0.4<br />

Average±sd 62.4±15.1 37.6±15.1 6.1±1.5 4.5±3.0 6.4±7.9 16.3±14.9 4.3±4.6<br />

*Insects: Dioryctria sp., Ernobius impressithorax, Pissodes validirostris.<br />

** Small mammals: Dormouse (Glis glis), Squirrel (Sciurus vulgaris).<br />

Total cones: 3045 (Migliarino 1045, Tirrenia 1000, Alberese 1000)


35,0<br />

30,0<br />

25,0<br />

20,0<br />

15,0<br />

10,0<br />

5,0<br />

0,0<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Insects* L.<br />

occidentalis<br />

Small<br />

mammals<br />

45<br />

MSRM-1<br />

MSRM-2<br />

PNM<br />

D. pinea Other<br />

causes<br />

Figure 1. Main damages observed on 3-yrs old Pinus pinea cones according different<br />

pine forests. (MSRM-1 “Parco Migliarino S. Rossore Massaciuccoli” MSRM-2, Tirrenia;<br />

PNM “Parco Naturale della Maremma”<br />

3.4 Test seed viability. Quality <strong>of</strong> seeds resulted quite variable after visual<br />

assessment and TZ test. Seeds discarded were about 35% from apparently healthy<br />

cones and raised up to 55% and 64% from partially injured and injured cones<br />

respectively (Table 6).<br />

Table 6. Percentage <strong>of</strong> seed losses in healthy and injured cones by D. pinea.<br />

Percentage<br />

<strong>of</strong> seed losses<br />

after/<br />

Visual<br />

assessment<br />

Seeds from<br />

healthy cones<br />

(n.503)<br />

Seeds from<br />

partially Injured cones<br />

(n.1056)<br />

Seeds from<br />

Injured cones<br />

(n.1221)<br />

TOT<br />

(n.2780)<br />

34,4 50,5 60,9 52,1<br />

TZ test 0,9 4,6 3,8 3,4<br />

T. losses 35,3 55,1 64,6 55,5<br />

4. DISCUSSION<br />

Due to the losses <strong>of</strong> seed production several survey activities have been<br />

promoted by Italian local regional governments. In this study, conduced during the<br />

period 2006-2008, mainly supported by the Tuscany region, the surveys showed<br />

that seed losses can be attributed to different causes but mainly to D. pinea, the<br />

fungal parasite which is able to colonize a large amount <strong>of</strong> cones since the first


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

year after their development (up to 64% <strong>of</strong> 1-yr losses). Later on the fungus can<br />

survive for long time on old cones on the ground under the canopy (up to 80% <strong>of</strong><br />

old cones sampled). Large part <strong>of</strong> these cones, confirming Munck and Stanosz<br />

(<strong>2009</strong>), produced available inoculum, able to germinate in 6 h at 25°C and able to<br />

infect the new cones generation.<br />

Ranking the most common types <strong>of</strong> damage on the harvested mature cones, we<br />

found that large part <strong>of</strong> losses can be attributed to D. pinea, followed by small<br />

mammals and insects, including L. occidentalis. However the occurrence <strong>of</strong> the<br />

fungus was clearly related with site conditions and was more pronounced in humid<br />

areas whilst was lower in dry forests where insects and small mammals prevailed.<br />

On the basis <strong>of</strong> this data seed losses related to L. occidentalis were not so<br />

evident. Probably the role <strong>of</strong> the insect is more dangerous during the early stages <strong>of</strong><br />

cone growth and also it remain difficult to be detected on mature cones.<br />

5. ACKNOWLEDGEMENTS<br />

Authors are grateful to the Regional Agricultural Agency <strong>of</strong> Tuscany, (ARSIA)<br />

and its Monitoring service (META) for the financial support during the<br />

investigation. Pr<strong>of</strong>. Riziero Tiberi, Forest Entomologists from DIBA, is also<br />

acknowledged for useful suggestions and help for insects identification.<br />

6. REFERENCES<br />

Bernardinelli, I., Zandigiacomo, P., 2001. Leptoglossus occidentalis Heidemann (Heteroptera,<br />

Coreidae): a conifer seed bug recently found in Northern Italy. - In: Knizek M., Forster<br />

B., Grodzki W. (eds.), Methodology <strong>of</strong> forest insect and diseases survey in Central<br />

Europe. Proceedings <strong>of</strong> the 4th international Workshop <strong>of</strong> the IUFRO WP 7.03.10,<br />

September 17-20, 2001, Praha (Czech Republic). J. For. Sci., 47, Special Issue 2, 56-58.<br />

Bernetti, G., 1995. Selvicoltura <strong>special</strong>e, UTET, Torino.<br />

Cambi, M., 2008. Conservazione dell'inoculo di Diplodia pinea su pigne nel parco di Migliarino - S.<br />

Rossore - Massaciuccoli. Tesi di Laurea Scienze Forestali ed ambientali. Anno<br />

accademico 2006–07. Facoltà di Agraria, Università degli Studi di Firenze.<br />

Çanakçıoğlu, H., 1969. Insect damage on cones and seeds <strong>of</strong> forest trees in Turkey, Istanbul<br />

University <strong>Orman</strong> Fakultesi Dergisi. 19A(2), 82-88.<br />

Ducci, F., Maltoni, A., Tani, A., 2001. La raccolta del seme di specie forestali, Monti e Boschi 70, 57-<br />

62.<br />

El Hassani, A., Messaoudi, J., 1986. Italian stone pine (Pinus pinea L.): the cone and seed pest <strong>of</strong><br />

conifers and their distribution in Marocco. Proceedings <strong>of</strong> the 2nd Conference <strong>of</strong> the cone<br />

and seed insect working party S207-01. September 3-5, 1986. Briancon, 05100, France.<br />

pp 5-14.<br />

Feducci, M., 2007. Indagini sulle cause dei danni alla fruttificazione del pino domestico in Italia. Tesi<br />

di Laurea Magistrale. In Gestione dei sistemi forestali. Anno accademico 2005– 06.<br />

Facoltà di Agraria, Università degli Studi di Firenze.<br />

Innocenti, M., Tiberi, R., 2002. Cone and seed pests <strong>of</strong> Pinus pinea L. in Central Italy, Redia, 85, 21-<br />

28.<br />

46


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

International Seed Testing Association., 2003. Working Sheet on Tetrazolium Testing. Vol 1. 1st<br />

Edition. CH-Switzerland.<br />

International Seed Testing Association., 2007. Biochemical Test for Viability : The Topography<br />

Tetrazolium Test International Rules for Seed Testing 2007. ISTA. CH-Switzerland.<br />

Maresi, G., Ambrosi, P., Battisti, A., Capretti, P., Danti, R., Feci, E., Minerbi, S., Tegli, S., 2002. Pine<br />

dieback by Sphaeropsis sapinea in Northern and Central Italy, Proceedings IUFRO WP<br />

7.02.02, 17-21 June, 2001, Hyytiälä, Finland, Finnish Forest Research Institute, Research<br />

Papers 829, pp. 60-67.<br />

Munck I.A., Smith D.R., Stanosz G.R., 2007. Quantification <strong>of</strong> conidia <strong>of</strong> Diplodia sp. Extracted<br />

from Red and Jack pine. Acta Silvatic. Lign. Hung., Spec. Edition, p. 280.<br />

Munck, I.A., Stanosz, G.R., <strong>2009</strong>. Longevity <strong>of</strong> inoculum production by Diplodia pinea on red pine<br />

cones. Forest pathology, doi: 10.1111/j.1439-0329.<strong>2009</strong>.00607.x<br />

Pepori, A.L., 2006. Danni da Sphaeropsis sapinea alla fruttificazione del pino domestico nel litorale<br />

tirrenico. Tesi di Laurea Scienze Forestali ed ambientali. Anno accademico 2005– 06.<br />

Facoltà di Agraria, Università degli Studi di Firenze.<br />

Petri, L., 1917. Contributo allo studio delle condizioni di recettività del Pinus pinea per la<br />

Sphaeropsis necatrix, Ann. R. Ist. Sup. Forestale, II.<br />

Petri, L., Adani A., 1916. Ricerche sopra una Malattia dei coni del "Pinus pinea", Estratto dagli<br />

Annali della R. Accademia di Agricoltura di Torino, volume LIX, 1916.<br />

Roques, A., 1983. Les insectes ravageurs des cones et graines de conifères en France, Instutut<br />

National de la Recherche Agronomique, Paris.<br />

Salvadori, C., 2005. II cimicione americano delle conifere, Terra Trentina, Novembre: 31-33.<br />

Saporito, L., 2002. Produzione di pinoli in popolamenti di Pinus pinea L. della Sicilia orientale,<br />

Sherwood, 78, 35-39.<br />

Verona, O., 1950. Note sopra una malattia degli strobili di “Pinus pinea” prodotti da “Sphaeropsis<br />

necatrix”, Università di Pisa, Istituto di Patologia Vegetale e Microbiologia agraria, Pisa.<br />

47


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 48-56<br />

SUSCEPTIBILITY OF DIFFERENT CONIFEROUS SEEDLINGS<br />

INOCULATED WITH Diplodia pinea<br />

H. Tuğba Doğmuş-Lehtijärvi 1* , Asko Lehtijärvi 1 , Gürsel Karaca 2 ,<br />

A. Gülden Aday 1 , Funda Oskay 1<br />

1 Süleyman Demirel University, Faculty <strong>of</strong> Forestry, Dept. <strong>of</strong> Botany, Isparta-Turkey<br />

2 Süleyman Demirel University, Faculty <strong>of</strong> Agriculture, Dept. <strong>of</strong> Plant Protection, Isparta-Turkey<br />

ABSTRACT<br />

* tugba@orman.<strong>sdu</strong>.edu.tr<br />

In this study, virulence <strong>of</strong> 5 Diplodia pinea isolates on Pinus nigra, Pinus brutia, Pinus<br />

sylvestris, Cedrus libani, Abies nordmanniana ssp. bornmülleriana and Juniperus excelsa<br />

seedlings was investigated. Needle fascicles on terminal and three lateral shoots <strong>of</strong> each<br />

seedling were removed to create small wounds. Agar plugs with D. pinea mycelia were<br />

placed in the wounds and secured in position with laboratory film. Five seedlings were used<br />

for each species-isolate combination and a randomized complete block design was used in<br />

the trial. Nine weeks later dead shoots, lesion length and fungal growth were recorded.<br />

Dead shoots occurred in almost all isolate–host combinations: the only exceptions were two<br />

isolates on J. excelsa. Within host variation in dead shoot rates among the isolates was low.<br />

However, there was high variation in the mean dead shoot rate among the hosts, with the<br />

highest rate (98.0%) on C. libani, and the lowest (4.0%) on J. excelsa. On C. libani all<br />

isolates caused remarkable lesion, while on the other host species some <strong>of</strong> the isolates<br />

caused lesion. The average lengths <strong>of</strong> lesion were 15.72 mm on C. libani, 8.12 mm on P.<br />

nigra, 2.4 on P. sylvestris, 1.2 on P. brutia and 0.08 mm on A. nordmanniana ssp.<br />

bornmülleriana and J. excelsa. Similarly, average linear extension <strong>of</strong> D. pinea in the shoot<br />

tissues was high on C. libani and Pinus spp. and low on J. excelsa. The results indicate high<br />

virulence <strong>of</strong> D. pinea on C. libani and P. nigra.<br />

Keywords: Pathogenicity, Pinus nigra, Pinus brutia, Pinus sylvestris, Cedrus libani,<br />

Abies nordmanniana ssp. bornmülleriana, Juniperus excelsa<br />

1- INTRODUCTION<br />

Diplodia pinea (Desm.) Kickx. (=Sphaeropsis sapinea (Fr.) Dyko Sutton) has<br />

a worldwide distribution and causes the disease known by the common name <strong>of</strong><br />

Diplodia tip blight <strong>of</strong> pine (Stanosz et al., 1996). The fungus can affect the trees<br />

from early to older ages causing shoot blight, twig blight, dead top, sap stain, root<br />

48


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

disease and cankers on stems and branches (Brookhouser and Peterson, 1971;<br />

Peterson, 1977) and can have devastating effects on various conifers when it is<br />

associated with stress-inducing factors (Sinclair et al., 1987; Swart et al., 1987;<br />

Chou and Mackenzie, 1988; Nicholls and Ostry 1990; Stanosz et al., 2001).<br />

Although diseases caused by the fungus are usually on trees under stress,<br />

healthy trees, e<strong>special</strong>ly nursery seedlings can also be infected (Palmer and<br />

Nichols, 1985). Pinus species are the most common hosts <strong>of</strong> the fungus, but Abies<br />

Mill., Chamaecyparis Spach., Cupressus L., Larix Mill., Picea A. Dietr.,<br />

Pseudotsuga Carriere and Thuja L. species are among hosts <strong>of</strong> the pathogen<br />

(Punithalingam and Waterston, 1970; Swart and Wingfield, 1991). Previous studies<br />

reported significant interactions among host species, environmental factors and the<br />

virulence <strong>of</strong> the isolates (Swart and Wingfield, 1991; Blodgett and Stanosz, 1997;<br />

Adams et al., 2002; Blodgett and Bonello, 2003).<br />

Two different S. sapinea morphotypes have been described and are referred to<br />

as A and B (Palmer et al., 1987; Smith and Stanosz, 1995; de Wet et al., 2000;<br />

Burguess and Wingfield, 2001). While D. pinea is accepted name for morphotype<br />

A, morphotype B has been recognized as Diplodia scrobiculata J. de Wet, B.<br />

Slippers & M. J. Wingfield (de Wet et al., 2003). It is reported that they differ in<br />

morphology, cultural characteristics and aggressiveness against host plants (Smith<br />

and Stanosz, 1995; Blodgett and Stanosz, 1997; Blodgett and Bonello, 2003).<br />

Eventhough D. pinea was recorded sixteen years ago for the first time on Pinus<br />

pinea and Pinus pinaster (Unligil and Ertas, 1993), very little is known about the<br />

distribution and the damage <strong>of</strong> the disease by far in Turkey (Doğmuş- Lehtijärvi,<br />

et al. 2007). Studies carried out in the western part <strong>of</strong> Taurus Mountains located in<br />

the Mediterranean part <strong>of</strong> Turkey, in Isparta province showed that D. pinea was the<br />

main agent <strong>of</strong> the shoot blight <strong>of</strong> Calabrian pines (Pinus brutia Ten.) (Doğmuş-<br />

Lehtijärvi et al., 2007). The aim <strong>of</strong> this study was to determine the pathogenicity <strong>of</strong><br />

D. pinea isolates obtained from P. brutia showing shoot blight symptoms, on 6<br />

conifer species, under field conditions.<br />

2- MATERIALS AND METHODS<br />

Five-year-old potted Pinus nigra Arnold, Pinus brutia Ten, Pinus sylvestris L.,<br />

Cedrus libani A. Rich., Abies nordmanniana ssp. bornmülleriana Mattfelt<br />

seedlings obtained from Eskişehir and Balıkesir forest nurseries and 3- year-old<br />

Juniperus excelsa Bieb seedlings obtained from Eğirdir forest nursery were used as<br />

the plant materials. Fungal materials were the five D. pinea isolates obtained from<br />

the Calabrian pine (P. brutia) shoots showing blight symptoms in Aşağı-Gökdere<br />

location <strong>of</strong> Isparta province, Turkey. The isolates were confirmed as D. pinea by<br />

Glen Stanosz, (Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Wisconsin-Madison,<br />

personal communication) using species-specific primers (Smith and Stanosz 2006).<br />

49


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Pathogenicity trial was carried out in the campus area <strong>of</strong> Süleyman Demirel<br />

University, during October-November, 2007. The D. pinea isolates were grown on<br />

PDA (Merck) at 24ºC in the dark, for one week. Four wounds <strong>of</strong> 2 x 2 mm were<br />

made approximately 2 cm below the shoot apex, on one terminal and three lateral<br />

shoots <strong>of</strong> each seedling, removing a needle fascicle by a scalpel and agar plugs<br />

with D. pinea mycelia cut from the actively growing culture were placed myceliaside-down<br />

on the wounds (Blodget and Stanosz, 1997). Wounds were then<br />

wrapped with parafilm. Non-colonized agar plugs were placed on control<br />

seedlings. Five seedlings were used for each species-isolate combination and a<br />

randomized complete block design was used in the trial. Seedlings were incubated<br />

under field conditions for 9 weeks. Average temperature in October was 14.4°C<br />

(maximum up to 28.2 °C, minimum 0.8 °C) and in November 7.4 °C (maximum up<br />

to 22.9 °C, minimum -10°C). Average relative humidity was 58 % in October and<br />

76 % in November. The seedlings were regularly irrigated and controlled for<br />

characteristic disease symptoms. Dead shoots were recorded.<br />

At the end <strong>of</strong> the inoculation period, inoculated terminal and lateral shoots were<br />

cut 15 cm below the inoculation point and brought to the laboratory. Based on the<br />

color changes on the shoots and needles, lesion sizes were measured. Then the<br />

needles on the shoots were removed and the shoots were cut into 1 cm long<br />

segments, from the inoculation point to the cut end <strong>of</strong> the shoots. The segments<br />

were surface sterilized by keeping them 10 seconds in 96 % ethyl alcohol and 4<br />

minutes in 1 % NaOCl and dried between sterile paper towels. They were then<br />

placed on petri plates with 2.5 % bacto agar and 0.05 % tannic acid, 5 segments per<br />

plate, in a clockwise serial order.<br />

The plates were incubated at room temperature for 4 weeks and examined under<br />

stereomicroscope for the presence <strong>of</strong> D. pinea colonies growing from the segments.<br />

Dead shoot rate, lesion length and fungal growth data obtained for each tree<br />

species-isolate combination were statistically analyzed by using SPSS program.<br />

3- RESULTS<br />

Susceptibility <strong>of</strong> the tree species to D. pinea and the virulence <strong>of</strong> the D. pinea<br />

isolates among those host species were different from each other. Dead shoots were<br />

observed in all isolate-host combinations, except for P047 and P097 isolates on J.<br />

excelsa. Considering the dead shoot rates, within host variation <strong>of</strong> the isolates was<br />

low, while host species had high variation. The highest rate was on C. libani<br />

(98.0%), followed by P. nigra (51.0%) and P. brutia (31.9%). Dead shoot rates<br />

were low on Abies (16.0%), P. sylvestris (12.8%) and J. excelsa (4.0%) (Figure 1).<br />

50


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Figure 1. Dead shoot rates <strong>of</strong> five tree species inoculated with five D. pinea isolates.<br />

On C. libani all isolates caused remarkable lesion, while on the other host<br />

species only 1-3 <strong>of</strong> the isolates caused lesion. The length <strong>of</strong> lesion ranged from<br />

12.4 to 21.4 mm on C. libani and from 9.0 to 18.6 mm on P. nigra, while on the<br />

other hosts the length did not exceed 6.0 mm (Figure 2).<br />

There was a significant correlation ( r= 0.53, p< 0.01) between length <strong>of</strong><br />

lesion and linear fungal extension on all host-isolate combinations. Average<br />

length <strong>of</strong> the lesion was greater than average fungal extension in all host species<br />

including all isolates <strong>of</strong> the fungus ( Table 2, Figure 3). The average fungal<br />

extension was found 27.6 mm on P. nigra, 24.0mm on P. brutia, 18.4mm on C.<br />

libani, 16.4 mm on P. sylvestris and 8.4 mm on A. nordmanniana<br />

bornmülleriana. Similar to length <strong>of</strong> lesion , average fungal extension was the<br />

lowest on J. excelsa (1.6 mm). Control seedlings did not show any symptoms <strong>of</strong><br />

disease.<br />

51


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 2. Length <strong>of</strong> lesions <strong>of</strong> five tree species inoculated with five D. pinea isolates.<br />

Figure 3. Average fungal extension and length <strong>of</strong> lesion caused by D. pinea.<br />

52


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Table 1. Linear extension <strong>of</strong> D. pinea in the shoots <strong>of</strong> five tree species inoculated with<br />

different isolates (mm).<br />

Tree<br />

species<br />

53<br />

Isolates<br />

T008 T029 P047 T081 P097 Control Averages<br />

A.n.bor. 6.0 B ab 8.0 B a 10.0 BC a 10.0ABC a 8.0 B a 0.0 Ab 8.4AB<br />

C. libani 12.0 AB b 34.0 A a 14.0 B b 20.0AB ab 12.0 B b 0.0 Ac 18.4 BC<br />

J. excel. 2.0 B a 2.0 B a 2.0 C a 2.0 C a 0.0 C a 0.0 Aa 1.6 A<br />

P. brutia 20.0 A ab 38.0 A a 20.0 B ab 24.0 A b 18.0AB ab 0.0 Ac 24.0 BC<br />

P. nigra 22.0 A abc 22.0ABabc 54.0 A a 6.0 BC bc 34.0 A ab 0.0 Ac 27.6 C<br />

P.syl. 12.0AB ab 8.0 B bc 28.0 AB a 24.0 A ab 10.0 B abc 0.0 Ac 16.4 ABC<br />

Means in the same column followed by the same uppercase letter and means in the same row<br />

shown by the same lowercase letter were not significantly different from each other according to<br />

Duncan’s Multiple range test (P=0.05)<br />

4- DISCUSSION<br />

Inoculations with the five D. pinea isolates resulted in necrotic needles and dead<br />

shoot tips on the tested coniferous tree species, showing high virulence on C. libani<br />

and P. nigra and low on A. nordmanniana ssp. bornmülleriana and J. excelsa.<br />

Even though most <strong>of</strong> the coniferous tree species have been tested for their<br />

sensitivity to S. sapinea sensu lato (s.l.) (Chou, 1976; Brookhouser, 1971; Blodgett<br />

and Stanosz, 1997; Flowers, 2001; Blodgett and Bonello, 2003; Luchi et al., 2007;<br />

Munoz et al., 2007) this is the first study mentioning the high susceptibility <strong>of</strong> C.<br />

libani to D. pinea isolates.<br />

The time <strong>of</strong> the inoculation significantly affects the relative aggressiveness <strong>of</strong><br />

the fungus. (Literatür). The results <strong>of</strong> the experiment indicate that D. pinea were<br />

pathogenic on tested conifer tree species, e<strong>special</strong>ly on C. libani and Pinus spp.<br />

However, inoculations with most <strong>of</strong> the fungal isolates resulted in little or no<br />

measurable symptoms on shoots <strong>of</strong> J. excelsa and A. nordmanniana in October and<br />

November when the experiment was carried out. To confirm these results it may be<br />

necessary to repeat the experiment during spring when the conditions are much<br />

more favorable for the natural spread <strong>of</strong> the fungus.<br />

Blodgett and Stanosz (1997) indicated that the wound inoculation technique<br />

provides a reproducible method for comparing the aggressiveness <strong>of</strong> the isolates<br />

belonging to different species <strong>of</strong> the fungus but, they did not found this method<br />

reliable for comparing the aggressiveness <strong>of</strong> the isolates within the species. In the<br />

present experiment, the same technique was used to find out the susceptibility <strong>of</strong><br />

the host species to the disease, and similarly, there were no significant differences<br />

between the five isolates on any <strong>of</strong> the tree species. Blodgett and Stanosz (1997)<br />

conducted the inoculation experiments also with conidial suspensions to be able to


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

understand the penetration capacity <strong>of</strong> the fungus to the healthy tissue. Also in the<br />

present study more realistic results <strong>of</strong> the virulence <strong>of</strong> the isolates could have been<br />

obtained by spraying the seedlings with conidial suspensions.<br />

Virulence <strong>of</strong> S. sapinea s.l. isolates on pine varies depending on several factors,<br />

such as ecological and climatic conditions, season, species <strong>of</strong> the fungus and where<br />

the experiment takes place (Swart and Wingfield, 1991; Blodgett and Stanosz,<br />

1997). It is also known that S. sapinea s.l. is able to live without typical symptoms<br />

in its latent phase (Stanosz et al., 1997; Flowers et al., 2001; 2003; Stanosz et al.,<br />

2005; Maresi et al., 2007). In that sense, it is important to have knowledge about<br />

the latent infections in nurseries and forest stands. D. scrobiculata was not found in<br />

the area where the isolates used in the present experiment were collected. As the<br />

isolates <strong>of</strong> D. pinea were directly obtained from pycnidia located on diseased<br />

shoots, instead <strong>of</strong> isolating from host tissues, only the frequently fruiting species<br />

could be found. Since the asymptomatic persistence <strong>of</strong> the fungal species in the<br />

host tissues is important for their wide distribution, a study focusing in latent<br />

infections is needed to shed light on the Diplodia-species composition on the local<br />

tree species and their infection potential.<br />

5-ACKNOWLEDGEMENTS<br />

This research was founded by Süleyman Demirel University Scientific Projects<br />

Unit (1325-06) and Republic <strong>of</strong> Turkey Prime Ministry State Planning<br />

Organisation (DPT 2003- K- 121020 /7). We are also very grateful to Glen Stanosz<br />

and Denise Smith for their help in molecular identification <strong>of</strong> isolates <strong>of</strong> S. sapinea<br />

sensu lato, Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Wisconsin-Madison.<br />

REFERENCES<br />

Adams, G. C., Wu, N. –T., Eisenberg, B. E., 2002. Virulence and double-stranded RNA in<br />

Sphaeropsis sapinea. Forest Pathology 32 (2002) 309–329.<br />

Blodgett, J.T., Stanosz, G.R., 1997. Sphaeropsis sapinea morphotypes differ in aggressiveness, but<br />

both infect nonwounded red or jack pines. Plant Disease 81, 143-147.<br />

Blodgett, J.T., Bonello, P., 2003. The aggressiveness <strong>of</strong> Sphaeropsis sapinea on Austrian pine varies<br />

with isolate group and site <strong>of</strong> infection. Forest Pathology 33, 15-19.<br />

Burguess, T., Wingfield, M. J., 2001. Simple sequence repeat markers distinguish among<br />

morphotypes <strong>of</strong> Sphaeropsis sapinea. Applied and Environmental Microbiology 67, 354-<br />

362.<br />

Brookhouser, L. W. Peterson, G. 1971. Infection <strong>of</strong> Austrian, Scots, and panderosa pines by Diplodia<br />

pinea. Phytopathology 61, 409–414.<br />

de Wet, J., Burgess, T., Slippers, B., Preisig, O., Wingfield, B.D., Winfield, M.J., 2000.<br />

Characterization <strong>of</strong> Sphaeropsis sapinea isolates from South Africa, Mexico and<br />

Indonesia. Plant disease 84, 151-156.<br />

54


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

de Wet, J., Burgess, T., Slippers, B., Preisig, O., Wingfield, B. D., Winfield, M.J., 2003. Multiple<br />

gene genealogies and microsatellite markers reflect relationships between morphotypes <strong>of</strong><br />

Sphaeropsis sapinea and distinguish a new species <strong>of</strong> Diplodia. Mycological Research<br />

107, 557-566.<br />

Chou, C. K. S., Mackenzie, M., 1988. Effect <strong>of</strong> prunning intensity and season on Diplodia pinea<br />

infection <strong>of</strong> Pinus radiata stem through prunning wounds. Eur. J: For. Path., 18, 437-444.<br />

Doğmuş Lehtijärvi H. T., Lehtijärvi, A., Karaca, G. and Aday, A. G., 2007. Sphaeropsis sapinea<br />

Dyko& Sutton associated with shoot blight on Brutian pine in soutwestern Turkey.<br />

IUFRO Meeting- May- 2007, Hungary. Acta Silv. Lign. Hung., Spec. Edition, 95-99.<br />

Flowers, J., Hartman, J., Vaillancourt, L., 2003. Detection <strong>of</strong> latent Sphaeropsis sapinea infection in<br />

Austrian pine tissues using Nested-Polymerase Chain Reaction. Phytopathology 93, 1471-<br />

1477.<br />

Flowers, J., Nuckles, E., Hartman, J., Vaillancourt, L., 2001. Latent infection <strong>of</strong> Austrian and Scots<br />

pine tissues by Sphaeropsis sapinea. Plant Disease 85, 1107-1112.<br />

Luchi, N., Capretti, P., Maresi, G., Feducci, M., 2007. Detection <strong>of</strong> Diplodia pinea in asymptomatic<br />

pine shooths. Acta Silv. Lign. Hung. Spec. Edition, 111-114.<br />

Nicholls, T. H., Ostry, M. E. 1990. Sphaeropsis sapinea cankers on stressed red and jack pines in<br />

Minnesota and Wisconsin. Plant Dis. 74 (1): 54-56.<br />

Maresi, G., Luchi, N., Pinzani, P., Pazzagli, M., Capretti, P., 2007. Detection <strong>of</strong> Diplodia pinea in<br />

asymptomatic pine shoots and its relation to the Normalized Insolation index. Forest<br />

Pathology 37, 272-280.<br />

Muñoz, Z., Moret, A., Garcés, S., 2008. The use <strong>of</strong> Verticillum dahliae and Diplodia scrobiculata to<br />

induce resistance in Pinus halepensis against Diplodia pinea infection. European Journal<br />

<strong>of</strong> Plant Pathology 120, 331-337.<br />

Palmer, M. A. Nicholls, T.H., 1985. Shoot blight and collar rot <strong>of</strong> Pinus resinosa caused by<br />

Sphaeropsis sapinea in forest tree nurseries. Plant Dis. 69, 739- 740.<br />

Palmer, M.A., Stewart, E.L., Wingfield, M.J., 1987. Variation among isolates <strong>of</strong> Sphaeropsis sapinea<br />

in the North Central United States. Phytopathology 77, 944- 948.<br />

Peterson, G., 1977. Infection, epidemiology, and control <strong>of</strong> Diplodia blight <strong>of</strong> Austrian, ponderosa,<br />

and Scots pines. Phytopathology 67: 511–514.<br />

Punithalingam, E., Waterston, J. M.1970. Diplodia pinea. CMI Descriptions <strong>of</strong> Plant Pathogenic<br />

Fungi and Bacteria. No. 273. Commonw. Mycol. Inst./Assoc. Appl. Biol.,Kew, Surrey,<br />

Eng.<br />

Smith, D.R., Stanosz, G.R., 1995. Confirmation <strong>of</strong> two distinct populations <strong>of</strong> Sphaeropsis sapinea in<br />

the north central United States using RAPDs. Phytopathology 85, 699-704.<br />

Smith, D.R., Stanosz, G.R., 2006. A species-specific PCR assay for detection <strong>of</strong> Diplodia pinea and<br />

D. scrobiculata in dead red and jack pines with collar rot symptoms. Plant Disease 90,<br />

307-313.<br />

Stanosz, G. R. Cummings, C.J., 1996. Association <strong>of</strong> mortality <strong>of</strong> recently planted seedlings and<br />

established saplings in red pine plantations with Sphaeropsis collar rot. Plant Disease 80,<br />

750–753.<br />

55


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Stanosz, G.R., Smith, D.R., Albers, J.S., 2005. Surveys for asymptomatic persistence <strong>of</strong> Sphaeropsis<br />

sapinea on or in stems <strong>of</strong> red pine seedlings from seven Great Lakes region nurseries.<br />

Forest Pathology 35, 233-244.<br />

Stanosz, G.R., Blodgett, J.T., Smith, D.R., Kruger, E.L. 2001. Water stress and Sphaeropsis sapinea<br />

as a latent pathogen <strong>of</strong> red pine seedlings. New Phythologist 149: 531-538.<br />

Stanosz, G.R., Smith, D.R., Guthmiller, M.A., Stanosz, J.C., 1997. Persistence <strong>of</strong> Sphaeropsis<br />

sapinea on or in asymptomatic shoots <strong>of</strong> red and jack pines. Mycologia 89, 525-530.<br />

Swart, W. J., Wingfield, M. J., 1991. Biology and control <strong>of</strong> Sphaeropsis sapinea on Pinus species in<br />

South Africa. Plant Dis. 75 (8): 761-766.<br />

Swart, W. J. Wingfield, M. J., Knox, P. S., 1987. Factors associated with Sphaeropsis sapinea<br />

infection <strong>of</strong> pine trees in South Africa. Phytophylactica 19: 505-510.<br />

Unligil, H., Ertaş, A.,1993. Damage caused by Sphaeropsis sapinea to pine trees near İstanbul [<br />

İstanbul yakınlarındaki çam ağaçlarında Sphaeropsis sapinea (Fr.) Dyko & Sutton mantar<br />

hastalığı]. İstanbul Universitesi <strong>Orman</strong> <strong>Fakültesi</strong> Dergisi Seri A 43 (1): 131-137. (in<br />

Turkish).<br />

56


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 57-64<br />

SITE AND STAND CHARACTERISTICS OF A Pinus brutia STAND<br />

INFECTED WITH Diplodia pinea IN TURKEY<br />

Nevzat GÜRLEVIK 1* , H.T. Doğmuş LEHTIJÄRVI 1 , Asko LEHTIJÄRVI 1 ,<br />

ABSTRACT<br />

A. Gülden ADAY 1<br />

SDU Faculty <strong>of</strong> Forestry, 32260 Isparta, TURKEY<br />

*gurlevik@orman.<strong>sdu</strong>.edu.tr<br />

Shoot blight can be a very costly disease for coniferous forests as it results in defoliation<br />

and reduced growth. In this study, some <strong>of</strong> the site and stand characteristics <strong>of</strong> a Pinus<br />

brutia Ten. stand infected with Diplodia pinea (Desm.) Kickx has been investigated.<br />

Blight occurrence was investigated for two different years, and also some soil and foliar<br />

nutrients were determined. Our results showed that at the first sapling time in 2006, 18 out<br />

<strong>of</strong> 90 shoot samples had infection, while in 2008 this number fell down to 5 out <strong>of</strong> 90<br />

samples. Results for soil analysis generally indicated somewhat poor site conditions. On<br />

average, soil had sandy clay loam texture, reaction was 7.72 and organic matter content was<br />

moderate with 3.84 percent. However, soil had a coarse fraction up to 62 % indicating<br />

relatively poor nutrition. Similarly, foliar nutrient levels also showed poor site quality.<br />

Overall, needle N concentration was lower than 1 %, indicating N deficiency on these sites.<br />

These poor site conditions and extreme droughts may have played a role in development <strong>of</strong><br />

the shoot blight in these stands.<br />

Keywords: Diplodia pinea, Pinus brutia, drought, nutrients<br />

1. INTRODUCTION<br />

Pinus brutia Ten. is the most commonly distributed coniferous forest tree<br />

species in Turkey, covering approximately 5.4 million hectares. It can be found at<br />

elevations between sea level and 1500 m. Pinus brutia grows best at elevations <strong>of</strong><br />

600-800m, below which very high temperature and demand for evapotranspiration,<br />

and above which cold weather restrict its distribution and growth (Boydak, 2000)<br />

Diplodia pinea (Desm.) Kickx is one <strong>of</strong> the most common disease agents in<br />

coniferous forests. It causes necrosis on needles, and therefore defoliation and<br />

death <strong>of</strong> shoots. In turn, it results in reduced growth and substantial economical<br />

losses (Stanosz et al., 2004).<br />

Several studies indicate higher risk <strong>of</strong> disease if plants are stressed. E<strong>special</strong>ly,<br />

water stress seems to increase the occurrence <strong>of</strong> this disease. Paoletti et al. (2001)<br />

showed that water stressed Pinus halepensis Mill. seedlings had longer cankers <strong>of</strong><br />

Sphaeropsis sapinea sensu lato 5 months after inoculation. Blodgett et al. (1997a)<br />

have also shown that 3-year-old red pine (Pinus resinosa Sol. ex Aiton) seedlings<br />

57


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

had higher S. sapinea infection when the seedlings were subjected to water stress.<br />

Similarly, in a 9–year–old red pine plantation in Wisconsin, trees with lower predawn<br />

water potentials had more severe development <strong>of</strong> disease (Blodgett et al., 1997b). Also,<br />

nutritional imbalances and alterations <strong>of</strong> nutrients can affect development <strong>of</strong> disease<br />

(Stanosz et al., 2004). Blodgett et al. (2005) showed that fertilized red pine trees had<br />

higher foliar N and greater lesion size.<br />

In 2004 and 2005, Diplodia had devastating effect on P. brutia stands on lower<br />

ranges <strong>of</strong> its distribution area (Lehtijärvi et al., 2007). It caused defoliation <strong>of</strong> the pines<br />

and reduced growth. In this study, the objectives were to determine the occurrence <strong>of</strong><br />

disease on pines and also to determine the site and stand characteristics <strong>of</strong> this infected<br />

site.<br />

2. MATERIAL AND METHOD<br />

The study site is located in Asagi Gokdere district <strong>of</strong> Isparta, along the Isparta-<br />

Antalya highway (37 o 33" N, 30 o 45" E). The altitude is 350 m above sea level and the<br />

distance to the Mediterranean coast is 75 km. The site was planted in mid 1980’s, and<br />

it became somewhat dense over time with natural seeding and lack <strong>of</strong> tending<br />

operations. Number <strong>of</strong> trees per hectare was 2100 on average and it reached as high as<br />

2775 on one site. Mean annual precipitation is 1052 mm in Antalya weather station,<br />

and there is almost no rainfall between June and September during the growing season.<br />

Monthly average temperatures range from 0 o C to 23.4 o C, but maximum daily<br />

temperatures can be as high as 41 o C during the summer, indicating very high demand<br />

for evapotranspiration.<br />

For the study, four different treatments with three replications were applied to the<br />

stands to control the disease. These treatments included control, pruning, thinning and<br />

pruning + thinning. Each plot was 20 x 20 m in size. Remains <strong>of</strong> pruning and thinning<br />

operations were removed from the plots afterwards.<br />

We conducted some measurements to determine the disease rate and site<br />

parameters. These measurements included fungal survey, soil analysis, foliar analysis<br />

and water potential.<br />

2.1. Fungal survey<br />

10 trees were samples from each plot for fungal survey in two different years (2006<br />

and 2008). One dead branch from each tree was cut and the shoots were investigated<br />

under the stereomicroscope for the presence <strong>of</strong> fungal structures. Samples were also<br />

isolated to identify the fungal species based on conidial and pycnidial morphology.<br />

Number <strong>of</strong> shoot samples with D. pinea was recorded.<br />

2.2. Soil sampling<br />

Soil samples were taken from 5 points in each plot at 0-20 cm depth. Coarse<br />

material larger than approximately 3 cm in diameter was not included in soil<br />

58


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

samples. Samples were mixed for each plot, sieved to pass 2 mm after air drying,<br />

and physical and chemical properties were determined in laboratory.<br />

2.3. Needle samples<br />

Prior to sampling, every tree was labelled and their health condition was<br />

recorded. Five healthy and 5 unhealthy trees were chosen for foliar sampling in<br />

each plot (Figure 1). 40 fascicles were taken from each tree with a shotgun.<br />

Samples were dried at 70°C for 24 hours. The dry weights were measured and the<br />

needles ground, then chemical analysis was performed for nutrients.<br />

2.4. Plant water potential<br />

We assumed that water potential <strong>of</strong> the trees may have played a role in<br />

development <strong>of</strong> this fungal disease. Five trees were chosen in each plot for water<br />

potential measurements. From each tree, representative needle samples were taken<br />

from the bottom 1/3 <strong>of</strong> the canopy, and needle water potentials were measured with<br />

a pressure bomb. However, the data was not conclusive, so the initial results were<br />

not included here.<br />

3. RESULTS<br />

3.1. Fungal survey<br />

Figure 1. Healthy (a) and unhealthy (b) pine needles.<br />

Since start <strong>of</strong> the fungal disease in 2004, the rate <strong>of</strong> the disease has decreased<br />

over time. In 2006, the ratio <strong>of</strong> samples with Diplodia was 20 %, while in 2008 the<br />

ratio fell down to 5.5 %. As <strong>of</strong> today, the stand looks quite healthy.<br />

3.2. Soil samples<br />

On average, soil texture was sandy clay loam (Table 1). Soil reaction was relatively<br />

neutral. Organic matter and nutrient concentrations were moderate. However, soil<br />

59


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

organic matter and nutrient contents were low considering that the coarse fraction (up<br />

to 3 cm) was as high as 62 %, and fraction larger than 3 cm in size estimated to be<br />

more than 50 % in some plots.<br />

3.3. Needle samples<br />

Similarly, foliar nutrient levels also showed poor site quality (Table 2). Overall,<br />

needle N concentration was lower than 1 per cent. Usually, less than 1.2 % N indicates<br />

some deficiency and N concentration is usually between 1.4-1.6 % in first quality sites.<br />

There was no significant difference between healthy and unhealthy needles in terms <strong>of</strong><br />

nutrient concentrations (Figure 2), however fascicle weight <strong>of</strong> health needles were<br />

significantly greater (Figure 3).<br />

Table 1. Soil characteristics <strong>of</strong> the infected stands.<br />

Parameters Mean Std. Error<br />

Sand (%) 50.11 1.958<br />

Silt (%) 27.32 1.297<br />

Clay (%) 22.57 1.570<br />

Coarse faction (%) 34.63 5.664<br />

EC 188.53 8.909<br />

pH 7.72 0.044<br />

Lime (%) 16.08 1.899<br />

Organic matter (%) 3.84 0.116<br />

N (ppm) 1360.33 71.736<br />

P (ppm) 16.38 1.425<br />

K (ppm) 256.17 14.053<br />

Ca (ppm) 8130.92 361.213<br />

Mg (ppm) 248.08 13.118<br />

Na (ppm) 16.47 0.925<br />

Table 2. Foliar characteristics <strong>of</strong> the infected stands.<br />

Parameters Mean Std. Error<br />

N (%) 0.97 0.024<br />

P (%) 0.13 0.003<br />

K (%) 0.33 0.010<br />

Ca (%) 0.83 0.030<br />

Mg (%) 0.21 0.007<br />

Fe (ppm) 37.24 1.069<br />

Cu (ppm) 3.13 0.197<br />

Mn (ppm) 32.60 3.640<br />

Zn (ppm) 15.00 0.414<br />

B (ppm) 19.07 1.418<br />

Fascicle weight (g) 6.53 0.204<br />

60


N (%)<br />

P (%)<br />

K (%)<br />

1,40<br />

1,20<br />

1,00<br />

0,80<br />

0,60<br />

0,18<br />

0,14<br />

0,10<br />

0,06<br />

0,60<br />

0,40<br />

0,20<br />

0,00<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

h uh h uh h uh h uh<br />

K T P TP<br />

h uh h uh h uh h uh<br />

K T P TP<br />

h uh h uh h uh h uh<br />

K T P TP<br />

Figure 2. Nitrogen, phosphorus and potassium concentrations <strong>of</strong> healthy (h) and<br />

unhealthy (uh) needles in control (C), thinning (T), pruning (P) and thinning plus pruning<br />

(TP) treatments.<br />

61


100 fascicle weight (g)<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 3. Fascicle weights <strong>of</strong> healthy (h) and unhealthy (uh) pine needles in control (C),<br />

thinning (T), pruning (P) and thinning plus pruning (TP) treatments.<br />

4. CONCLUSION<br />

h uh h uh h uh h uh<br />

K T P TP<br />

Diplodia had some very serious effects on pine stands <strong>of</strong> Asagi Gokdere region.<br />

The symptoms diminished over the past four years since its first occurrence in this<br />

site. The site characteristics indicated poor site with low nutrients and rocky soil.<br />

Nutrient level in the needles also showed very low levels <strong>of</strong> N, indicating nitrogen<br />

deficiency. In fact, N levels in needles were almost one half <strong>of</strong> the one in a healthy<br />

needle. In addition, these stands have received no silvicultural treatments in recent<br />

years and the number <strong>of</strong> stems in these stands was almost twice as much as it<br />

should be.<br />

Although the site is located in a region which receives approximately 1000 mm<br />

rainfall, most <strong>of</strong> it is received in winter, and summer droughts are very common.<br />

Temperatures reaching above 40°C during the summer months intensify the effects<br />

<strong>of</strong> drought, creating a very high demand for evapotranspiration.<br />

These poor site conditions and extreme droughts may have played a role in<br />

development <strong>of</strong> the shoot blight in these stands. Apparently, the site was both<br />

nutrient and water limited, e<strong>special</strong>ly during the growing season. This limitation<br />

was most apparent when the canopy <strong>of</strong> the stand was investigated. In most areas <strong>of</strong><br />

the site, the individual trees had very thin and sparse canopy due to lack <strong>of</strong> growing<br />

space and site resources to develop healthy crown (Figure 4). Timely and<br />

appropriate silvicultural treatments may reduce the effects <strong>of</strong> poor soil conditions<br />

and droughts, leading to stronger and more resistant trees beforehand, and better<br />

recovery after an infection.<br />

62


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Figure 4. Defoliated branches and remaining thin canopy <strong>of</strong> Pinus brutia infected by<br />

Diplodia pine.<br />

5. ACKNOWLEDGEMENTS<br />

This work was supported by Süleyman Demirel University BAP Project (1325-<br />

M-06).<br />

6. LITERATURE CITED<br />

Blodgett J. T., Kruger E. L., Stanosz G. R., 1997a. Effects <strong>of</strong> Moderate Water Stress on Disease<br />

Development by Sphaeropsis sapinea on Red Pine. Phytopathology, 87 (4): 422-428<br />

Blodgett J. T., Kruger E. L., Stanosz G. R., 1997b. Sphaeropsis sapinea and Water Stres in a Red<br />

Pine Plantation in Central Wisconsin. Phytopathology, 87 (4): 429-434<br />

Blodgett J.T., Herms D.A. Bonello P., 2005. Effects <strong>of</strong> fertilization on the red pine defense chemistry<br />

and resistance to Sphaeropsis sapinea. Forest Ecology and Management, 2008: 373-382.<br />

Boydak, M., 2006. Biology and Silviculture <strong>of</strong> Turkish Red Pine (Pinus brutia Ten.). Lazer Ofset<br />

press, Ankara, 19-41 s<br />

63


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

GDF, 2008. Forest atlas. General Directorate <strong>of</strong> Forestry, Ankara<br />

Lehtijärvi H.T.D., Lehtijärvi A., Karaca G., Aday, A.G., 2007. Sphaeropsis sapinea Dyko & Sutton<br />

Associated with Shoot Blight on Pinus brutia Ten. in Southwestern Turkey. Acta Silv.<br />

Lign. Hung., Spec. Edition (2007) 95-99.<br />

Paoletti, E., Danti R., Strati S., 2001. Pre- and post-inoculation water stress affects Sphaeropsis<br />

sapinea canker length in Pinus halepensis seedlings. For. Path. 31: 209-218<br />

Stanosz, G. R., Trobaugh, J., Guthmiller, M. A., Stanosz, J. C., 2004. Sphaeropsis shoot blight and<br />

altered nutrition in red pine plantations treated with paper mill waste sludge. For. Path.<br />

34: 245–253<br />

64


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 65-70<br />

THE EFFECTS OF SIROCOCCUS SHOOT BLIGHT AND VITALITY<br />

FERTILIZATION ON GROWTH OF MATURE NORWAY SPRUCE<br />

Markus HUBER 1 , Erhard HALMSCHLAGER 2* , Hubert STERBA 1<br />

Department <strong>of</strong> Forest and Soil Science, University <strong>of</strong> Natural Resources and Applied Life Sciences,<br />

Vienna (BOKU),<br />

1 Institute <strong>of</strong> Forest Growth and Yield Research, Peter Jordanstraße 82,<br />

A-1190 Vienna, Austria, 2 Institute <strong>of</strong> Forest Entomology, Forest Pathology and Forest Protection<br />

(IFFF), Hasenauerstraße 38, A-1190 Vienna, Austria<br />

ABSTRACT<br />

* erhard.halmschlager@boku.ac.at<br />

The impact <strong>of</strong> Sirococcus shoot blight and vitality fertilization on the growth <strong>of</strong> mature<br />

Norway spruce (Picea abies (L.) Karst.) was studied in a single tree fertilization<br />

experiment, established in autumn 2000. A total <strong>of</strong> 144 sample trees were selected among<br />

the dominant and co-dominant trees <strong>of</strong> the 90-year-old Norway spruce stand. Half <strong>of</strong> the<br />

trees exhibited severe symptoms <strong>of</strong> Sirococcus shoot blight whereas the other half were<br />

apparently healthy and vigorous. A randomised block design with the factors “slope<br />

section” (lower slope versus upper slope) and “Sirococcus shoot blight” (severely affected<br />

versus healthy trees) was used. Within these blocks sample trees were randomly assigned to<br />

one <strong>of</strong> the three treatments (dolomitic liming, application <strong>of</strong> gypsum and kieserite,<br />

unfertilized control). Due to tree mortality caused by bark beetle infestation the<br />

experimental design became unbalanced and therefore final analyses were performed with<br />

the volume growth data <strong>of</strong> 125 sample trees only. The effects <strong>of</strong> Sirococcus shoot blight<br />

and fertilization treatments on current annual volume increment were investigated by<br />

analysis <strong>of</strong> covariance, using the average volume increment <strong>of</strong> the period 1977–1980 as a<br />

covariate attribute (assuming that tree growth was not yet affected by Sirococcus shoot<br />

blight during this period). Indeed results indicated that Sirococcus shoot blight started in<br />

1981 in the experimental stand and trees with shoot blight symptoms had a significantly<br />

lower increment over the whole period 1981–2006. Sirococcus induced increment reduction<br />

<strong>of</strong> the nonfertilized trees continuously increased from 7,5 ± 2,9% in 1981 to 37 ± 3,8% by<br />

the year 2000. A significant positive effect <strong>of</strong> vitality fertilization was only achieved with<br />

the gypsum and kieserite variant from 2002 to 2006. The highest surplus increment was<br />

found in 2004 with 31,6 ± 15,2%, calculated as average over the diseased and healthy<br />

group. However, a mitigation <strong>of</strong> increment loss caused by Sirococcus shoot blight was<br />

statistically significant only for the year 2003.<br />

Keywords: Sirococcus conigenus, Norway spruce, increment reduction, vitality<br />

fertilization<br />

65


1. INTRODUCTION<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Since the early 1980s shoot blight by Sirococcus conigenus (DC.) P. Cannon &<br />

Minter has caused severe damage on Norway spruce (Picea abies [L.] Karst.) in<br />

some parts <strong>of</strong> Austria (Fig. 1). Previous studies revealed insufficient Mg and Ca<br />

supply, enhanced N/Mg and N/Ca-ratios in the needles <strong>of</strong> severely diseased trees<br />

(Anglberger et al., 2003) and demonstrated that application <strong>of</strong> appropriate<br />

fertilizers mitigated disease severity and promoted tree recovery (Halmschlager et<br />

al., 2007). The present study was carried out in the same individual-treefertilization<br />

experiment and aimed to investigate the effect <strong>of</strong> Sirococcus shoot<br />

blight on growth <strong>of</strong> mature Norway spruce (for further details see: Huber et al.,<br />

<strong>2009</strong>).<br />

Figure 1. Severely affected mature Norway spruce.<br />

66


2. MATERIAL & METHODS<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The experiment was established in a mature Norway spruce stand in Upper<br />

Austria. Two fertilizer treatments and an unfertilized control variant were applied<br />

on a total <strong>of</strong> 144 dominant or co-dominant trees in a randomized block design<br />

(Fig. 2) in April 2001. Half <strong>of</strong> the trees exhibited symptoms <strong>of</strong> Sirococcus shoot<br />

blight (“diseased”), whereas the other trees were “healthy”. The average basal area<br />

increment <strong>of</strong> the two groups diverged after the year 1980, therefore analysis <strong>of</strong><br />

covariance was applied for testing the effects <strong>of</strong> fertilization and Sirococcus<br />

symptoms on volume increment, using the average <strong>of</strong> the current annual increment<br />

in the period 1977 to 1980 as the covariate.<br />

Figure 2. Position <strong>of</strong> the sample trees within the experimental stand. The tree position is<br />

marked with different symbols for the variants. Trees from the slope position “upper slope”<br />

are found northern to the dotted line. (Figure reprinted from Huber et al. <strong>2009</strong> with<br />

permission from Elsevier.)<br />

67


3. RESULTS & DISCUSSION<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Tree ring analyses indicated that Sirococcus shoot blight started in 1981 in the<br />

experimental stand and trees with shoot blight symptoms had a significantly lower<br />

increment over the whole period 1981–2006 (data not shown). Sirococcus induced<br />

increment reduction <strong>of</strong> the nonfertilized trees continuously increased from 7,5 ±<br />

2,9% in 1981 to 37 ± 3,8% by the year 2000 (Fig. 3). Results therefore clearly<br />

revealed a volume-growth decreasing effect <strong>of</strong> Sirococcus shoot blight in mature<br />

Norway spruce, which has been been observed only on young Norway spruce trees<br />

yet (Halmschlager et al., 2000).<br />

Figure 3. Current annual increment <strong>of</strong> the unfertilized “diseased” and “healthy” sample<br />

trees, adjusted for the same average current annual increment <strong>of</strong> 21.0 dm 3 in the period<br />

1977 to 1980, and the difference between the current annual increment <strong>of</strong> “diseased” and<br />

“healthy” trees in percent <strong>of</strong> the current annual increment <strong>of</strong> “healthy” trees. The error bars<br />

indicate the standard error <strong>of</strong> the adjusted means. (Figure reprinted from Huber et al. <strong>2009</strong><br />

with permission from Elsevier.)<br />

68


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

69<br />

Figure 4. The development <strong>of</strong> the adjusted average increment per tree by fertilization treatment for the “diseased” and “healthy” trees.<br />

The error bars indicate the standard errors <strong>of</strong> the means. The means in the subfigure on the right were calculated as average over the<br />

“diseased” and “healthy” group. (Figure reprinted from Huber et al. <strong>2009</strong> with permission from Elsevier.)


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

The effect <strong>of</strong> fertilization on tree volume growth was negative in the first year<br />

after fertilization in both treatments. A positive effect <strong>of</strong> vitality fertilization was<br />

achieved for the first time 2 years after fertilization with the gypsum and kieserite<br />

variant (Fig. 4). This significant surplus in increment was found from 2002 to 2005<br />

and culminated in 2004 with 32 ± 15%, calculated as average over the diseased and<br />

healthy group. The rapid response <strong>of</strong> the fertilized trees in this treatment can be<br />

explained by the quick plant availability <strong>of</strong> Ca and Mg from these water soluble<br />

fertilizers. In the dolomitic lime variant a positive effect occurred at first in 2004<br />

on the healthy trees and in 2006 also on the diseased trees.<br />

A significant interaction between Sirococcus symptoms and fertilization with<br />

kieserite + gypsum was found in 2003, where the increment increasing effect by<br />

fertilization was higher for the “diseased” trees. Therefore a possible mitigation <strong>of</strong><br />

Sirococcus-induced increment losses by fertilization can be suggested.<br />

4. REFERENCES<br />

Anglberger, H., Sieghardt, M., Katzensteiner, K., Halmschlager, E., 2003. Needle nutrient status <strong>of</strong><br />

Sirococcus shoot blight-diseased and healthy Norway spruces. Forest Pathology 33, 21-<br />

29.<br />

Halmschlager, E., Anglberger, H., Katzensteiner, K., Sterba, H., 2007. The effect <strong>of</strong> fertilisation on<br />

the severity <strong>of</strong> Sirococcus shoot blight in a mature Norway spruce (Picea abies [L.]<br />

Karst.) stand. Acta Silvatica & Lignaria Hungarica, Special Edition 2007,101-11.<br />

Halmschlager, E., Gabler, A., Andrae, F., 2000. The impact <strong>of</strong> Sirococcus shoot blight on radial and<br />

height growth <strong>of</strong> Norway spruce (P. abies) in young plantations. Forest Pathology 30,<br />

127–33.<br />

Huber, M., Halmschlager, E., Sterba, H., <strong>2009</strong>. The impact <strong>of</strong> forest fertilization on growth <strong>of</strong> mature<br />

Norway spruce affected by Sirococcus shoot blight. Forest Ecology and Management<br />

257, 1489-95.<br />

Acknowledgements: We gratefully acknowledge the support <strong>of</strong> the Austrian<br />

Federal Forests (Österreichische Bundesforste AG), district management Traun- &<br />

Innviertel in supplying financial support, the plot and local resources for this<br />

investigation.<br />

70


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 71-84<br />

INTERACTION BETWEEN Diplodia pinea, D. scrobiculata AND<br />

SEVERAL FUNGAL ENDOPHYTES IN RED AND JACK PINE<br />

SEEDLINGS<br />

Oscar SANTAMARÍA 1 *; Denise R. SMITH 2 and Glen R. STANOSZ 2<br />

1 Dpt. Ingeniería del Medio Agronómico y Forestal, Universidad de Extremadura. Ctra. de Cáceres<br />

s/n, 06071 Badajoz, SPAIN.<br />

2 Dpt. <strong>of</strong> Plant Pathology, University <strong>of</strong> Wisconsin-Madison, 1630 Linden Dr., Madison 53706, WI,<br />

USA.<br />

ABSTRACT<br />

* osantama@unex.es<br />

Sphaeropsis sapinea sensu lato is a conifer fungal pathogen that causes mainly shoot<br />

blight and stem cankers. Recently, the former S. sapinea has been divided in two new<br />

species, D. pinea and D. scrobiculata. The aim <strong>of</strong> the study was to evaluate the interaction<br />

between those two fungal pathogens and among them and several fungal endophytes<br />

isolated from healthy shoots <strong>of</strong> Pinus resinosa and P. banksiana adult trees. Interaction was<br />

evaluated by means <strong>of</strong> co-inoculations in Red and Jack pine seedling under greenhouse<br />

conditions. Symptom severity (distance below the inoculation site at which necrotic needles<br />

were observed) was recorded after four weeks <strong>of</strong> incubation and used as response variable.<br />

The results showed D. pinea to be much more aggressive on both hosts than D.<br />

scrobiculata. When both pathogens where inoculated in a single plant, the symptom<br />

development was mainly due to D. pinea. Furthermore, D. scrobiculata showed antagonism<br />

with D. pinea, since when both pathogens co-occurred in a single seedling, symptom<br />

severity caused by D. pinea was lower than that caused when D. pinea acted alone. The<br />

results also suggested that two <strong>of</strong> the endophytes, Trichoderma atroviride and Rosellinia<br />

subiculata, were able to inhibit the pathogen spreading and therefore they could be<br />

considered biocontrol agents against D. pinea. Further studies would be needed to confirm<br />

biocontrol.<br />

Keywords: Sphaeropsis sapinea, biocontrol, Pinus sp, inoculation.<br />

1. INTRODUCTION<br />

Sphaeropsis sapinea sensu lato is a conifer fungal pathogen <strong>of</strong> worldwide<br />

distribution which has caused significant economic damage in nurseries,<br />

plantations, and natural stands (Gibson, 1979; Nicholls and Ostry, 1990; Davison et<br />

al., 1991). The parasite can infect most parts <strong>of</strong> host plants, causing a broad range<br />

<strong>of</strong> disease symptoms: shoot blight, stem cankers, branch dieback, dead tops, death,<br />

and blue staining <strong>of</strong> cut wood (Nicholls and Ostry, 1990; Stanosz and Cummings-<br />

Carlson, 1996).<br />

71


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Two groups, A and B, were distinguished within S. sapinea sensu lato (Palmer<br />

et al., 1987) based on several characteristics such as morphology, growth, virulence<br />

and molecular markers (Palmer et al., 1987; Smith and Stanosz, 1995; de Wet et<br />

al., 2000; Burguess and Wingfield, 2001). Recently, the B group has been<br />

recognized as a distinct species and assigned the name Diplodia scrobiculata J. de<br />

Wet, B. Slippers & M. J. Wingfield (de Wet et al., 2003). At the same time, the A<br />

group was named as Diplodia pinea (Desmaz.) J. Kickx fil. (syn. Sphaeropsis<br />

sapinea (Fr.:Fr.) Dyko & Sutton in Sutton) (de Wet et al., 2003). In terms <strong>of</strong><br />

pathogenicity, D. pinea has been shown to be more aggressive than D. scrobiculata<br />

(Blodgett and Stanosz, 1997; Blodgett and Bonello, 2003).<br />

The reported geographic and host ranges <strong>of</strong> D. pinea and D. scrobiculata are<br />

broad and overlapping. Not only do distributions and hosts overlap, but D. pinea<br />

and D. scrobiculata also are known to occur together. In this sense, both species<br />

have been isolated from individual red pine (Pinus resinosa Aiton.) plantations<br />

(Palmer, 1991; Stanosz et al., 2005) or even from a single tree (Morelet and<br />

Chandelier, 1993) or a single sample (Smith and Stanosz, 2006). Those<br />

experiments stated the potential for intimate association between D. pinea and D.<br />

scrobiculata within host tissues, but relatively little is known about their local cooccurrence<br />

and the implications <strong>of</strong> this fact for the disease development, and<br />

subsequent survival.<br />

Complete eradication <strong>of</strong> the pathogen is difficult due to latent infection <strong>of</strong><br />

symptomless tissues <strong>of</strong> apparently healthy trees (Flowers et al., 2001; Stanosz et<br />

al., 2005; Maresi et al., 2007; Stanosz et al., 2007); however, proper control<br />

measures could reduce the spread and virulence <strong>of</strong> the disease. Among those<br />

measures, the use <strong>of</strong> biological control is <strong>of</strong> increasing interest since it provides an<br />

effective and environmentally safer alternative to chemical application. Several<br />

microorganisms, mainly fungi, have been observed to cause ‘systemic induced<br />

resistance’ in the host after their inoculation into the plant, which may prompt a<br />

lower host susceptibility to later infections with D. pinea (Luchi et al., 2005;<br />

Blodgett et al., 2007; Muñoz et al., 2008). On the other hand, several endophytes,<br />

which have been shown to produce secondary metabolites, some <strong>of</strong> them with<br />

antifungal properties (Tan and Zou, 2001; Schulz et al., 2002), have shown<br />

antagonism with several pathogens and they have been assessed as biological<br />

control agents (Mehrotra et al., 1988; Holdenrieder and Greig, 1998; Roy et al.,<br />

2001). However in the literature, few studies dealing with the use <strong>of</strong> endophytes as<br />

biological control agents against D. pinea can be found.<br />

Therefore, the main aim <strong>of</strong> the study was to analyse the effect <strong>of</strong> D.<br />

scrobiculata and several fungal endophytes, isolated from healthy shoots <strong>of</strong> Pinus<br />

resinosa and P. banksiana Lamb. (jack pine) adult trees, on the symptom severity<br />

72


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

caused by D. pinea by means <strong>of</strong> co-inoculations on P. resinosa and P. banksiana<br />

seedlings, with the goal <strong>of</strong> evaluating their potential role as biological control<br />

agents <strong>of</strong> Diplodia pinea.<br />

2. MATERIALS AND METHODS<br />

2.1. Plant material<br />

Dormant, 1-year-old jack pine and 2-year-old red pine nursery seedlings were<br />

lifted at the beginning <strong>of</strong> winter and transplanted into Deepot cones (conical tubes,<br />

6.4 cm wide x 25.4 cm deep; Stuewe & Sons Inc., Corvallis, OR) in a soil mix (1:1<br />

vol/vol) <strong>of</strong> Plainfield sand (containing 89% sand and 7% silt) from a 24-year-old<br />

red pine plantation in central Wisconsin and Fafard growing mix no. 2 (Conrad<br />

Fafard Inc., Inkerman, New Brunswick, Canada). Red pine seedlings had a mean<br />

stem height <strong>of</strong> 16.9 cm 0.3 standard error (SE) and jack pine had a mean stem<br />

height <strong>of</strong> 26.8 cm 0.6 SE at the time transplanted. Seedlings were placed in a<br />

greenhouse supplemented with artificial light (maximum recorded ambient<br />

greenhouse photon flux density was 1,560 µmol·s -1 ·m -2 ; supplemental photon flux<br />

density averaged 132 µmol·s -1 ·m -2 ) to provide a 16-h photoperiod. The seedlings<br />

were watered to field capacity every 2 to 3 days. The climatic conditions during the<br />

experiments in the greenhouse were: average temperature (T) 30.6º C 0.7 SE,<br />

average relative humidity (RH) 30.0% 1.5 SE.<br />

2.2. Fungal material<br />

Diplodia mycelia inoculum was produced for two monoconidial isolates <strong>of</strong> each<br />

species, D. pinea and D. scrobiculata, collected from various pine species and<br />

locations in the north central United States (Table 1). Thirty four endophytes were<br />

isolated from vigorous external twigs located at 1.5 m above ground in the canopy<br />

<strong>of</strong> 20–30 year old dominant red and jack pine trees. For isolations, needle-free<br />

shoots were surface sterilized by immersion for 30 s in 95% ethanol, and then two<br />

immersions for 2 min each in 1.05% NaClO plus two drops Tween-80 per litre.<br />

This procedure is similar to those routinely used for detection <strong>of</strong> endophytic fungi<br />

(Bills, 1996). Then, for each shoot, four 5-mm long plant segments including bark<br />

were cut and plated into a Petri dish containing PDA (Potato Dextrose Agar, 39 g·l -<br />

1 , Difco Laboratories, Detroit, MI) medium and incubated at room temperature for<br />

1 week before examination. Outgrowing colonies were transferred to PDA and<br />

incubated at 24º C. Four endophytes (Table 1) among the total were selected to<br />

perform the experiments based on a preliminary study (data not shown) where the<br />

antagonism between all <strong>of</strong> them and the<br />

73


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Table 1: Origin <strong>of</strong> Sphaeropsis sapinea sensu lato and endophyte isolates used in the<br />

experiments.<br />

Isolate Isolate no. a Species Pine host Geographic origin<br />

P1 411 D. pinea b Pinus resinosa Clearwater Co., MN<br />

P2 04-126 D. pinea b Pinus banksiana Wood Co., WI<br />

S1 124 D. scrobiculata b Pinus banksiana Jackson Co., WI<br />

S2 462 D. scrobiculta b Pinus resinosa Clearwater Co., MN<br />

E1 08-08 Not determined Pinus banksiana Jackson Co., WI<br />

E2 08-09 Not determined Pinus resinosa Jackson Co., WI<br />

E3 08-10<br />

Trichoderma atroviride<br />

Karst. c<br />

E4 08-13 Rosellinia subiculata<br />

(Schwein.) Sacc. c<br />

74<br />

Pinus resinosa Jackson Co., WI<br />

Pinus banksiana Jackson Co., WI<br />

a<br />

Culture collection numbers <strong>of</strong> M. A. PALMER (3-digit number) or G. R. STANOSZ<br />

(04-xxx).<br />

b<br />

The isolates were previously characterized according to the specific primers given by<br />

SMITH and STANOSZ (2006).<br />

c<br />

The most homologous species given in the GeneBank after introducing the ITS<br />

sequence.<br />

Diplodia isolates was evaluated in vitro. Identifications <strong>of</strong> these four endophytes<br />

were carried out by obtaining the ITS region sequence and later comparison with<br />

the GeneBank. The fungal DNA was extracted directly from mycelia previously<br />

grown in PDB (Potato Dextrose Broth, 39 g·l -1 , Difco Laboratories, Detroit, MI) by<br />

using the slightly modified method <strong>of</strong> Gilbertson et al. (1991) outlined in Smith<br />

and Stanosz (1995). The extracted DNA was amplified using ITS1 and ITS4<br />

primers, and the entire region containing both internal transcribed spacers (ITS)<br />

was sequenced by the University <strong>of</strong> Wisconsin Biotechnology Center on an<br />

Applied Biosystems 3730XL DNA sequencer.<br />

2.3. Interaction between D. pinea and D. scrobiculta experiment<br />

The elongating, asymptomatic terminal shoot <strong>of</strong> each seedling was inoculated in<br />

the greenhouse approximately 11 weeks after transplanting in the case <strong>of</strong> red pine,<br />

and 6 weeks in jack pine. On each shoot, two wounds (in the opposite sides <strong>of</strong> the<br />

shoot) were made by removing a needle fascicle per side (by a sterile scalpel cut<br />

flush to the stem) approximately 2 cm below the shoot apex. Inoculations were<br />

made by placing fungus-side-down a 4-mm-diameter plug, cut from the margin <strong>of</strong><br />

an actively growing culture on PDA (Potato Dextrose Agar, 39 g·l -1 , Difco<br />

Laboratories, Detroit, MI), on each <strong>of</strong> the two wounds. In the controls, seedlings<br />

were inoculated with sterile PDA in both wounds. Another five nonwounded<br />

seedlings for each host were incorporated in the experiments as additional controls<br />

but those were not included in the statistical analyses. Parafilm (American National<br />

Can Co., Chicago, IL) was wrapped around the shoots for 7 days. Five seedlings<br />

per treatment combination (isolate and fungal species; Table 1) were used in each<br />

<strong>of</strong> two trials separated by two weeks. Thus, for each tree species were inoculated


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

five seedlings per each one <strong>of</strong> the following treatments: 1) P1+S1, 2) P1+S2, 3)<br />

P2+S1, 4) P2+S2, 5) P1+P1, 6) P2+P2, 7) S1+S1, 8) S2+S2, and 9) Agar A+A.<br />

Then, a total <strong>of</strong> 45 seedlings per trial and tree species were inoculated. All<br />

treatments were assigned randomly. Four weeks after inoculation, the distances<br />

below the inoculation site at which necrotic needles and cankers were present were<br />

measured (symptom severity). Most <strong>of</strong> the seedlings were carried to the lab and, by<br />

means <strong>of</strong> a molecular analysis with specific primers (Smith and Stanosz, 2006), the<br />

presence <strong>of</strong> D. pinea or/and D. scrobiculata was verified in order to assure they<br />

were the causal agents <strong>of</strong> the necroses.<br />

2.4. Interaction between D. pinea and several endophytes experiment<br />

This experiment was carried out on 1-year-old jack pine seedlings. Shoots were<br />

inoculated as explained above, although D. pinea was inoculated 2 cm below the shoot<br />

apex and the endophyte 5 cm below the shoot apex. The endophytes were inoculated<br />

one week before the Diplodia inoculation. Two types <strong>of</strong> controls were used: seedlings<br />

inoculated with sterile PDA instead <strong>of</strong> the endophyte into the lowest wound (D+PDA)<br />

and seedlings nonwounded in the lower part (D). Other nonwounded seedlings were<br />

incorporated in the experiments as additional controls but those were not included in<br />

the statistical analyses. Five seedlings per treatment combination (D. pinea isolate and<br />

endophyte species) were used in each <strong>of</strong> two trials separated by two weeks. Thus, a<br />

total <strong>of</strong> 60 seedlings were double-inoculated per trial and considered in the analysis.<br />

All treatments were assigned randomly.<br />

Two and four weeks after inoculation, the distances below the site inoculated with<br />

D. pinea isolates at which necrotic needles and cankers were present (symptom<br />

severity) were measured. At the end <strong>of</strong> the experiment, 40% <strong>of</strong> the seedlings were cut<br />

and carried to the lab where, after needles were removed, cross sections, 1 cm long,<br />

centered at 0, 1.5, and 3 cm from the site inoculated with Diplodia, were aseptically cut<br />

and surface disinfected as described above. Each <strong>of</strong> the three cross sections was<br />

transferred to a PDA plate and incubated for 1 week at 25ºC and light. The presence <strong>of</strong><br />

Diplodia isolates and endophyte species in the plates was recorded.<br />

2.5. Statistical analysis<br />

Symptom severity was analyzed by two-factor analysis <strong>of</strong> variance with<br />

interaction. Factors used as main effects were treatment and trial. Fisher´s least<br />

significant difference (LSD) test was used for multiple comparison among<br />

treatments by means <strong>of</strong> the General Linear Model Procedure <strong>of</strong> SAS (Statistical<br />

Analysis S<strong>of</strong>tware v. 9.1.3, SAS Institute Inc., Cary, NC) when significant<br />

differences were found in the ANOVA table. The ln (x+1) transformation <strong>of</strong> the<br />

response variable was used to stabilize the residual variance, although backtransformed<br />

values are presented in tables and figures for clarification.<br />

Assumptions <strong>of</strong> normality and homoscedasticity were assured by Kolmogorov-<br />

Smirnov and Levene´s tests respectively. In the endophyte experiment, since<br />

symptom severity was recorded at two and four weeks after inoculation, a repeated<br />

measurements analysis was also applied by means <strong>of</strong> Repeated Procedure <strong>of</strong> SAS,<br />

to test the effect <strong>of</strong> time on the symptom severity.<br />

75


3. RESULTS<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

3.1. Interaction between D. pinea and D. scrobiculta experiment<br />

Inoculated seedlings <strong>of</strong> both tree species produced symptoms similar to those<br />

reported for seedlings in field and nursery studies, including necrotic needles, stem<br />

cankers, and crooked and dead shoot tips. Symptoms were observed in 100% <strong>of</strong> the<br />

seedlings inoculated with D. pinea, in 60% <strong>of</strong> the seedlings inoculated with D.<br />

scrobiculata and in 95% <strong>of</strong> the seedlings inoculated with both pathogens. No<br />

symptoms developed on wounded or nonwounded controls <strong>of</strong> any <strong>of</strong> the two hosts.<br />

In jack pine, the two-factor analyses <strong>of</strong> variance <strong>of</strong> the symptom severity (distance<br />

below the inoculation site at which necrotic needles were present) indicated a<br />

significant effect <strong>of</strong> the treatment (F = 50.75, p < 0.01), but not <strong>of</strong> the trial (F =<br />

1.92, p = 0.170). Thus, the multiple comparison <strong>of</strong> the treatments was performed<br />

with the pooled data <strong>of</strong> the two trials.<br />

In jack pine, symptom severity was greater on seedlings inoculated with isolates<br />

<strong>of</strong> D. pinea than on seedlings inoculated with isolates <strong>of</strong> D. scrobiculata (Figure 1).<br />

Symptom severity on seedlings inoculated with the isolate S2 <strong>of</strong> D. scrobiculata<br />

was not significantly different from that on control seedlings. It was observed that<br />

symptom severity also differed between the two isolates <strong>of</strong> D. pinea. On seedlings<br />

inoculated with the most aggressive isolate (P1), there were not statistical<br />

differences in the symptom severity between seedlings inoculated only with P1 and<br />

those inoculated with P1 in combination with either isolate <strong>of</strong> D. scrobiculata<br />

(Figure 1). However, on seedlings inoculated with the less aggressive isolate <strong>of</strong> D.<br />

pinea (P2), the co-occurrence <strong>of</strong> both fungal species in the seedling significantly<br />

reduced symptom severity in comparison to that caused by P2 when it was<br />

inoculated alone (Figure 1).<br />

Figure 1: Mean symptom severity (distance below the inoculation site at which necrotic<br />

needles were present) for each treatment on jack pine seedlings. Averages with the same<br />

letter are not significantly different according to Fisher´s least significant difference (LSD)<br />

test at a significant level <strong>of</strong> 0.05. Vertical bars indicate Standard Error.<br />

76


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

In the case <strong>of</strong> red pine seedlings, two-way ANOVA <strong>of</strong> the symptom severity<br />

showed a significant effect <strong>of</strong> the treatment (F = 15.89, p < 0.01), the trial (F =<br />

16.29, p < 0.01), but not <strong>of</strong> their interaction (F = 0.34, p = 0.88). Therefore,<br />

multiple comparison was performed separately for each trial (Figure 2). As in the<br />

case <strong>of</strong> jack pine, isolates ranked in the same order in relation to their<br />

aggressiveness, although in general terms, symptom severity was lower in red pine<br />

than in jack pine. In red pine, it was also observed that the presence <strong>of</strong> either isolate<br />

<strong>of</strong> D. scrobiculata in the seedling reduced the symptom severity caused by either<br />

isolate <strong>of</strong> D. pinea (except in trial 1 for P2 isolate), although in red pine, the<br />

differences between treatments were not so evident (Figure 2). Molecular analyses<br />

indicated that symptom severity was caused mainly by D. pinea. D. scrobiculata<br />

was only detected in the proximities <strong>of</strong> the inoculation site.<br />

Figure 2: Mean symptom severity (distance below the inoculation each trial on red<br />

pine seedlings. Averages with the same letter are not significantly different according to<br />

Fisher´s least significant difference (LSD) test at a significant level <strong>of</strong> 0.05. Vertical<br />

bars indicate Standard Error.<br />

3.2. Interaction between D. pinea and several endophytes experiment<br />

Jack pine seedlings inoculated with D. pinea isolates also produced symptoms<br />

similar to those reported above. No symptoms developed on nonwounded controls.<br />

After 2 weeks, the two-factor analyses <strong>of</strong> variance <strong>of</strong> the ‘necrosis length’ (distance<br />

below the inoculation site at which necrotic needles were present) indicated a<br />

significant effect <strong>of</strong> the ‘treatment’ (F = 4.24, p < 0.01), the trial (F = 6.53, p =<br />

0.012), but not <strong>of</strong> the interaction (F = 1.41, p = 0.18). The two-factor ANOVA <strong>of</strong><br />

the ‘necrosis length’ caused by Diplodia isolates after 4 weeks showed that only<br />

the ‘treatment’ effect was significant (F = 2.91, p < 0.01). Only the ‘Time’ effect<br />

77


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

influenced the ‘treatment’ effect (interaction time*treatment: F = 3.94, p < 0.01).<br />

LSD test showed the inoculation with endophytes 08-10 and 08-13 to produce a<br />

significantly lower symptom severity caused by the second isolate (p2) <strong>of</strong> D. pinea<br />

after 2- and 4-weeks <strong>of</strong> incubation but only when compared with the control<br />

seedlings non-wounded in the lower part (Fig. 3). However, those fungi did not<br />

cause reduced symptom severity when compared with those caused by Diplodia on<br />

the controls wounded in the lower part and inoculated with PDA instead <strong>of</strong><br />

endophyte mycelia (Controls PDA). For the most aggressive isolate <strong>of</strong> D. pinea<br />

(the p1 isolate), no endophyte was able to reduce significantly the symptom<br />

severity caused by Diplodia (Fig. 3). Re-isolation percentages <strong>of</strong> each fungus can<br />

be observed in Table 2.<br />

Figure 3: Mean necrosis length caused by the D. pinea for each endophyte treatment<br />

after 2 (figure a) and 4 weeks (figure b) <strong>of</strong> incubation. Averages with the same letter are<br />

not significantly different according to Fisher´s least significant difference (LSD) test at<br />

a significant level <strong>of</strong> 0.05. Vertical bars indicate Standard Error. ‘Control’ was not<br />

wounded 5 cm below the apex; and ‘Agar’ was a control wounded and inoculated with<br />

agar 5 cm below the shoot apex.<br />

78


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Table 2: Re-isolation percentage <strong>of</strong> each fungal species from the seedlings where they had<br />

been inoculated.<br />

Isolate<br />

79<br />

Cross section a<br />

Initial Mid Final<br />

P1 100% 85% 55%<br />

P2 100% 60% 40%<br />

08-08 0% 0% 0%<br />

08-09 0% 0% 0%<br />

08-10 0% 25% 81.3%<br />

08-13 0% 0% 6.3%<br />

a .- Initial: 1-cm cross section including the Diplodia inoculation site;<br />

Mid: 1-cm cross section between both inoculation sites; Final: 1-cm cross<br />

section including the endophyte inoculation site.<br />

4. DISCUSSION<br />

Inoculations with D. pinea isolates resulted in greater severity <strong>of</strong> symptoms<br />

than inoculations with D. scrobiculata isolates on seedlings <strong>of</strong> red and jack pine.<br />

This result agrees well with those stated in previous studies (Blodgett and Stanosz,<br />

1997; 1999) which used a similar inoculation technique on the same or different<br />

host seedlings. However, in Blodgett and Stanosz (1997) D. scrobiculata was more<br />

aggressive in jack pine and D. pinea in red pine. In our study, although results are<br />

not statistically comparable since they were obtained in separated experiments,<br />

jack pine appeared to be the most susceptible host for both fungal species.<br />

The molecular analysis confirmed that, when both pathogens were inoculated in<br />

the same seedling, D. pinea was mainly responsible for the necrosis length<br />

observed in the seedlings. This fact was consistent with the greater aggressiveness<br />

shown by D. pinea. In general terms, when both species were inoculated in the<br />

same seedling, symptom severity observed was lower than that recorded when D.<br />

pinea was inoculated alone. This fact suggests a potential antagonism between both<br />

pathogens when they co-occurred in the same plant tissue. This may have very<br />

important implications, because it is already known D. pinea and D. scrobiculata<br />

occur together in nature. In this sense, both species have been isolated from<br />

individual red pine plantations (Palmer, 1991; Stanosz et al., 2005) or even from a<br />

single tree (Morelet and Chandelier, 1993) or a single sample (Smith and Stanosz,<br />

2006). Therefore it would be very interesting to develop further studies to<br />

investigate whether where both pathogens co-occur in those plantations, the disease<br />

incidence caused by Diplodia is lower than in plantations where just one pathogen<br />

is present.


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

However, this result should be taken cautiously because the reduction <strong>of</strong> the<br />

symptom severity caused by D. pinea when D. scrobiculata was also present, was<br />

not always statistically significant. One possible explanation for this lack <strong>of</strong><br />

significance could be related to the great variability recorded between the<br />

repetitions. This great variation could be a consequence <strong>of</strong> the natural presence <strong>of</strong><br />

both pathogens, in a latent state, in the plant material used in the experiments.<br />

Latency has already been described in previous studies as a very common state for<br />

Sphaeropsis sapinea sensu lato (Stanosz et al., 1997; Flowers et al., 2001; 2003;<br />

Stanosz et al., 2005; Maresi et al., 2007), which was isolated from asymptomatic<br />

plant tissues in a percentage range <strong>of</strong> 20-85%. Even in the present study, D. pinea<br />

and D. scrobiculata occurred asymptomatically in several <strong>of</strong> the non inoculated<br />

seedlings (data not shown). Despite this natural presence <strong>of</strong> the pathogen in the<br />

seedlings <strong>of</strong> our experiments, control seedlings did not become diseased and did not<br />

show any symptom <strong>of</strong> the disease. This result was predictable since it has been<br />

proposed by several authors that activation from this latency to a pathogenic state may<br />

take place under different conditions <strong>of</strong> stress (Stanosz et al., 1997; Smith et al., 2002);<br />

and seedlings used in the present study were grown in optimal conditions. However in<br />

the case <strong>of</strong> inoculated seedlings, the fungal spreading and colonization may cause<br />

stress to the plant that could activate the latent infections <strong>of</strong> the pathogen, if present. In<br />

such cases, the symptom severity recorded could be overestimated.<br />

In general terms, when two <strong>of</strong> the endophytes, 08-10 and 08-13, were also<br />

inoculated in the seedling, symptom severity casused by the D. pinea isolates was<br />

lower than that recorded when D. pinea was inoculated alone. This fact suggests a<br />

potential antagonism between those two endophytes and D. pinea when they cooccurred<br />

in the same plant tissue. However, that reduction <strong>of</strong> the symptom severity<br />

caused by D. pinea was only statistically significant on the less aggressive isolate<br />

<strong>of</strong> D. pinea (p2). Endophytes have been demonstrated to be suitable candidates as<br />

biological control agents, as they seem to be part <strong>of</strong> the defence system <strong>of</strong> trees<br />

(Barklund and Unestam, 1988; Ranta et al., 1995) and have already been shown to<br />

be antagonistic to many fungal pathogens, including Diplodia spp. (Holdenrieder<br />

and Greig, 1998; Roy et al., 2001; Campanile et al., 2007).<br />

Among the endophytes used in the present study, isolate 08-10 was identified as<br />

Trichoderma atroviride Karst. Since the potential <strong>of</strong> this genus as a biocontrol<br />

agent <strong>of</strong> plant pathogens was first recognized in the early 1930s (Weindling, 1932),<br />

continuous studies have demonstrated the effectiveness <strong>of</strong> this species in the<br />

biological control against a number <strong>of</strong> plant pathogenic fungi, including Diplodia<br />

spp., on several forest hosts (Knudsen et al., 1991; Mousseaux et al., 1998;<br />

Campanile et al., 2007; Schubert et al. 2008). T. atroviride is a mycoparasite<br />

which, once it recognizes and attacks the fungal host, uses it’s nutrients killing the<br />

host before or just after invasion (Chet et al., 1998). The endophyte 08-13, which<br />

was also able to reduce the symptom severity caused by Diplodia on pine<br />

seedlings, was identified as Rosellinia subiculata (Schwein.) Sacc. This fungal<br />

endophyte has been shown to produce sordarin, which is an antibiotic with<br />

antifungal properties against a number <strong>of</strong> plant pathogenic fungi (Bills et al., 2002).<br />

80


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

This fact is in agreement with a preliminary assay carried out in the present study<br />

for the endophyte selection, where R. subiculata was observed to produce a reddish<br />

compound on PDA culture media that inhibited strongly Diplodia growth on<br />

contact with mycelium.<br />

Among the three types <strong>of</strong> interaction between antagonistic organisms proposed<br />

by Adams (1990), competition, antibiosis and hyperparasitism, the former could be<br />

the most important mechanism acting for D. scrobiculata, although antibiosis could<br />

be also acting since S. sapinea sensu lato has been shown to produce metabolites<br />

with antifungal activity (Cabras et al., 2006). For T. atroviride hyperparasitism<br />

may be the main process involved in the reduction <strong>of</strong> symptom severity caused by<br />

D. pinea in most <strong>of</strong> the cases. This is corroborated with the fact that in the 92.8%<br />

<strong>of</strong> the seedlings where T. atroviride was re-isolated, D. pinea isolates were not able<br />

to cause necroses farther than the site inoculated with the endophyte. In the case <strong>of</strong><br />

R. subiculata it seems clear that the type <strong>of</strong> interaction involved in the antagonism<br />

would be antibiosis as expressed above.<br />

Systemic induced resistance (SIR) is a host defense mechanism which has also<br />

been proposed by several authors (Blodgett et al., 2007; Muñoz et al., 2008) to<br />

explain why inoculations with D. scrobiculata and other fungal endophytes<br />

reduced the symptom severity on seedlings inoculated later with D. pinea. In such<br />

studies, those fungi were inoculated several weeks before inoculating with D.<br />

pinea, because plants need some time to produce the defensive action. Therefore,<br />

the reduction <strong>of</strong> symptom severity caused by D. pinea when D. scrobiculata is also<br />

inoculated into the same seedling at the same time, as it is our case, could not be<br />

explained by this mechanism <strong>of</strong> SIR. In the case <strong>of</strong> the endophytes, however, those<br />

were inoculated one week before the D. pinea inoculation. Therefore, SIR could be<br />

implicated in the reduction <strong>of</strong> the symptom severity caused by D. pinea when<br />

endophytes had been previously inoculated into the plant. This is consistent with<br />

the fact that in the present study the reduction in the symptom severity caused by<br />

D. pinea when endophytes were also inoculated into the seedlings, was only<br />

significant when compared with the controls non-wounded in the lower part, but it<br />

was not with those wounded and inoculated with PDA plugs. This fact may<br />

indicate that wounding could activate that defensive mechanism. This mechanism<br />

has been already shown to occur after wounding without any later fungal<br />

inoculation in several important components in pine resistance such as resin flow<br />

(Luchi et al., 2005). Nevertheless, this is in disagreement with the results obtained<br />

by other authors who indicated that wounding alone does not induce SIR in the<br />

case <strong>of</strong> Pinus nigra Arn. (Blodgett et al., 2007) or in P. radiata D. Don. (Bonello et<br />

al., 2001).<br />

In conclusion, the results presented here suggested that, when both D. pinea and<br />

D. scrobiculata co-occurred in a single plant, the symptom development was<br />

mainly due to D. pinea. Furthermore, D. scrobiculata and two <strong>of</strong> the endophytes<br />

tested (the isolate 08-10 <strong>of</strong> Trichoderma atroviride and the isolate 08-13 <strong>of</strong><br />

Rosellinia subiculata) showed some antagonism to D. pinea, since, when both<br />

81


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

fungi co-occurred in a single seedling, symptom severity caused by D. pinea was<br />

lower than that caused when D. pinea was acting alone (at least this happened with<br />

the less aggressive isolates <strong>of</strong> D. pinea).<br />

Acknowledgements<br />

This research was partially supported by a grant provided by the Ministry <strong>of</strong><br />

Culture and Science <strong>of</strong> Spain (Program: José Castillejo) and by the Department <strong>of</strong><br />

Plant Pathology <strong>of</strong> the University <strong>of</strong> Wisconsin-Madison. We are also very grateful<br />

to the Wisconsin Department <strong>of</strong> Natural Resources for seedlings, and JoAnne<br />

Stanosz and Lea Gardiner for technical assistance. The attendance <strong>of</strong> Oscar<br />

Santamaría to the meeting was co-funded by the regional government <strong>of</strong><br />

Extremadura (Consejería de Economía, Comercio e Innovación de la Junta de<br />

Extremadura) and by the European Regional Development Fund (ERDF).<br />

5. REFERENCES<br />

Adams, P.B., 1990. The potential <strong>of</strong> mycoparasites for biological control <strong>of</strong> plant diseases. Annual<br />

Review <strong>of</strong> Phytopathology 28, 59-72.<br />

Barklund, P., Unestam, T., 1988. Infection experiments with Gremmeniella abietina on seedlings <strong>of</strong><br />

Norway spruce and Scots pine. European Journal <strong>of</strong> Forest Pathology 18, 409-420.<br />

Bills, G.F., 1996. Isolation and analysis <strong>of</strong> endophytic fungal communities from woody plants. In:<br />

Redlin, S.C., Carris, L.M. (Eds.) Endophytic fungi in grasses and woody plants:<br />

systematics, ecology and evolution. APS Press, Saint Paul, Minnesota. pp. 31-65.<br />

Bills, G.F., Dombrowski, A.W., Horn, W.S., Jansson, R.K., Rattray, M., Schmatz, D., Schwartz, R.E.,<br />

2002. Rosellinia subiculata ATCC 74386 and fungus ATCC 74387 for producing sordarin<br />

compounds for fungi control. Merck & Co., Inc. Rahway, NJ (USA). Patent no 6436395.<br />

http://www.freepatentsonline.com/6436395.html<br />

Blodgett, J.T., Bonello, P., 2003. The aggressiveness <strong>of</strong> Sphaeropsis sapinea on Austrian pine varies<br />

with isolate group and site <strong>of</strong> infection. Forest Pathology 33, 15-19.<br />

Blodgett, J.T., Stanosz, G.R., 1997. Sphaeropsis sapinea morphotypes differ in aggressiveness, but<br />

both infect nonwounded red or jack pines. Plant Disease 81, 143-147.<br />

Blodgett, J.T., Stanosz, G. R, 1999. Differences in aggressiveness <strong>of</strong> Sphaeropsis sapinea RAPD<br />

marker group isolates on several conifers. Plant Disease 83, 853-856.<br />

Blodgett, J.T., Eyles, A., Bonello, P., 2007. Organ-dependent induction <strong>of</strong> systemic resistance and<br />

systemic susceptibility in Pinus nigra inoculated with Sphaeropsis sapinea and Diplodia<br />

scrobiculata. Tree Physiology 27, 511-517.<br />

Bonello, P., Gordon, T.R., Storer, A.J., 2001. Systemic induced resistance in Monterey pine. Forest<br />

Pathology 31, 99-106.<br />

Burguess, T., Wingfield, M. J., 2001. Simple sequence repeat markers distinguish among<br />

morphotypes <strong>of</strong> Sphaeropsis sapinea. Applied and Environmental Microbiology 67, 354-<br />

362.<br />

Cabras, A., Mannoni, M.A., Serra, S., Andolfi, A., Fiore, M., Evidente. A., 2006. Occurence,<br />

isolation and biological activity <strong>of</strong> phytotoxic metabolites produced in vitro by<br />

Sphaeropsis sapinea, pathogenic fungus <strong>of</strong> Pinus radiata. European Journal <strong>of</strong> Plant<br />

Pathology 115, 187-193.<br />

82


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Campanile, G., Ruscelli, A., Luisi, N., 2007. Antagonistic activity <strong>of</strong> endophytic fungi towards<br />

Diplodia corticola assessed by in vitro and in planta tests. European Journal <strong>of</strong> Plant<br />

Pathology 117, 237-246.<br />

Chet, I., Benhamou, N., Haran, S., 1998. Mycoparasitism and lytic enzymes. In: Harman, G.E.,<br />

Kubicek, C.P. (Eds.). Trichoderma and Gliocladium, vol. 2. Taylor & Francis, London,<br />

pp. 153–171.<br />

Davison, E.M., Tay, F.C.S., Peroni, D., 1991. Sphaeropsis sapinea on pines in Western Australia.<br />

Australasian Plant Pathology 20, 31.<br />

de Wet, J., Burgess, T., Slippers, B., Preisig, O., Wingfield, B.D., Winfield, M.J., et al., 2000.<br />

Characterization <strong>of</strong> Sphaeropsis sapinea isolates from South Africa, Mexico and<br />

Indonesia. Plant disease 84, 151-156.<br />

de Wet, J., Burgess, T., Slippers, B., Preisig, O., Wingfield, B. D., Winfield, M.J., et al., 2003.<br />

Multiple gene genealogies and microsatellite markers reflect relationships between<br />

morphotypes <strong>of</strong> Sphaeropsis sapinea and distinguish a new species <strong>of</strong> Diplodia.<br />

Mycological Research 107, 557-566.<br />

Flowers, J., Hartman, J., Vaillancourt, L., 2003. Detection <strong>of</strong> latent Sphaeropsis sapinea infection in<br />

Austrian pine tissues using Nested-Polymerase Chain Reaction. Phytopathology 93, 1471-<br />

1477.<br />

Flowers, J., Nuckles, E., Hartman, J., Vaillancourt, L., 2001. Latent infection <strong>of</strong> Austrian and Scots<br />

pine tissues by Sphaeropsis sapinea. Plant Disease 85, 1107-1112.<br />

Gibson, I.A.S., 1979. Diseases <strong>of</strong> Forest Trees Widely Planted as Exotics in the Tropic and Southern<br />

Hemisphere. Part II. The Genus Pinus. Commonwealth Mycological Institute, Kew, Eng.<br />

Gilbertson, R.L., Rojas, M.R., Russell, D.R., Maxwell, D.P., 1991. Use <strong>of</strong> the asymmetric polymerase<br />

chain reaction and DNA sequencing to determine genetic variability <strong>of</strong> bean golden<br />

mosaic geminivirus in the Dominican Republic. Journal <strong>of</strong> General Virology 72, 2843-<br />

2848.<br />

Holdenrieder, O., Greig, B.J., 1998. Biological methods <strong>of</strong> control. In: Woodward, S., Stenlid, J.,<br />

KarjalaineN, R., Hüttermann, A.; Heterobasidium annosum: Biology, Ecology, Impact<br />

and Control. Oxon, UK. CAB International, pp. 235-258.<br />

Knudsen, G.R., Eschen, D.J., Dandurand, L.M., Bin, L., 1991. Potential for biocontrol <strong>of</strong> Sclerotinia<br />

sclerotiorum through colonization <strong>of</strong> sclerotia by Trichoderma harzianum. Plant Disease<br />

75, 466-470.<br />

Luchi, N., Ma, R., Capretti, P., Bonello, P., 2005. Systemic induction <strong>of</strong> traumatic resin ducts and<br />

resin flow in Austrian pine by wounding and inoculation with Sphaeropsis sapinea and<br />

Diplodia scrobiculata. Planta 221, 75-84.<br />

Maresi, G., Luchi, N., Pinzani, P., Pazzagli, M., Capretti, P., 2007. Detection <strong>of</strong> Diplodia pinea in<br />

asymptomatic pine shoots and its relation to the Normalized Insolation index. Forest<br />

Pathology 37, 272-280.<br />

Mehrotra, R., Aneja, K., Gupta, A., Aggerwal, A., 1988. Fungi-agents <strong>of</strong> biological control. In:<br />

Mukerji, K. G.; Garg, K. L. Biocontrol <strong>of</strong> plant diseases. Vol. 1. CRC Press, Boca Raton,<br />

Fla., pp. 37-52.<br />

Morelet, P.M., Chandelier, P., 1993. Sur un cas de variabilité chez Sphaeropsis sapinea. European<br />

Journal <strong>of</strong> Forest Pathology 23, 317-320.<br />

Mousseaux, M.R., Dumroese, R.K., James, R.L., Wenny, D.L., Knudsen, G.R., 1998: Efficacy <strong>of</strong><br />

Trichoderma harzianum as a biological control <strong>of</strong> Fusarium oxysporum in containergrown<br />

Douglas-fir seedlings. New Forests 15, 11-21.<br />

83


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Muñoz, Z., Moret, A., Garcés, S., 2008. The use <strong>of</strong> Verticillum dahliae and Diplodia scrobiculata to<br />

induce resistance in Pinus halepensis against Diplodia pinea infection. European Journal<br />

<strong>of</strong> Plant Pathology 120, 331-337.<br />

Nicholls, T.H., Ostry, M.E., 1990. Sphaeropsis sapinea cankers on stressed red and jack pines in<br />

Minnesota and Wisconsin. Plant Dis. 74, 54-56.<br />

Palmer, M.A., 1991. Isolate types <strong>of</strong> Sphaeropsis sapinea associated with main stem cankers and topkill<br />

<strong>of</strong> Pinus resinosa in Minnesota and Wisconsin. Plant Disease 75, 507-510.<br />

Palmer, M.A., Stewart, E.L., Wingfield, M.J., 1987. Variation among isolates <strong>of</strong> Sphaeropsis sapinea<br />

in the North Central United States. Phytopathology 77, 944-948.<br />

Ranta, H., Neuvonen, S., Ylimartimo, A., 1995. Interactions <strong>of</strong> Gremmeniella abietina and<br />

endophytic fungi in shoots <strong>of</strong> Scots pine trees treated with simulated acid rain. Journal <strong>of</strong><br />

Applied Ecology 32, 67-75.<br />

Roy, G., Bussières, G., Laflamme, G., Dessureault, M., 2001. In vitro inhibition <strong>of</strong> Heterobasidion<br />

annosum by Phaeotheca dimorphospora. Forest Pathology 31, 395-404.<br />

Schubert, M., Fink, S., Schwarze, F.W.M.R., 2008. Evaluation <strong>of</strong> Trichoderma spp. as a biocontrol<br />

agent against wood decay fungi in urban trees. Biological Control 45, 111–123.<br />

Schulz, B., Boyle, C., Draeger, S., Römmert, A. –K.; Krohn, K., 2002. Endophytic fungi: a source <strong>of</strong><br />

novel biologically active secondary metabolites. Mycological Research 106, 996-1004.<br />

Smith, D.R.; Stanosz, G.R., 1995. Confirmation <strong>of</strong> two distinct populations <strong>of</strong> Sphaeropsis sapinea in<br />

the north central United States using RAPDs. Phytopathology 85, 699-704.<br />

Smith, D.R., Stanosz, G.R., 2006. A species-specific PCR assay for detection <strong>of</strong> Diplodia pinea and<br />

D. scrobiculata in dead red and jack pines with collar rot symptoms. Plant Disease 90,<br />

307-313.<br />

Smith, H., Wingfield, M.J., Coutinho, T.A., 2002. The role <strong>of</strong> latent Sphaeropsis sapinea infections in<br />

post-hail associated dieback <strong>of</strong> Pinus patula. Forest Ecology and Management 164, 177-<br />

184.<br />

Stanosz, G.R., Cummings-Carlson, J., 1996. Association <strong>of</strong> mortality <strong>of</strong> recently planted seedlings<br />

and established saplings in red pine plantations with Sphaeropsis collar rot. Plant Dis. 80,<br />

750-753.<br />

Stanosz, G.R., Smith, D.R., Albers, J.S., 2005. Surveys for asymptomatic persistence <strong>of</strong> Sphaeropsis<br />

sapinea on or in stems <strong>of</strong> red pine seedlings from seven Great Lakes region nurseries.<br />

Forest Pathology 35, 233-244.<br />

Stanosz, G.R., Smith, D.R., Leisso, R., 2007. Diplodia shoot blight and asymptomatic persistence <strong>of</strong><br />

Diplodia pinea on or in stems <strong>of</strong> jack pine nursery seedlings. Forest Pathology 37, 145-<br />

154.<br />

Stanosz, G.R., Smith, D.R., Guthmiller, M.A., Stanosz, J.C., 1997. Persistence <strong>of</strong> Sphaeropsis<br />

sapinea on or in asymptomatic shoots <strong>of</strong> red and jack pines. Mycologia 89, 525-530.<br />

Tan, R.X., Zou, W.X., 2001. Endophytes: a rich source <strong>of</strong> functional metabolitos. Natural Products<br />

Rep. 18, 448-459.<br />

Weindling, R., 1932. Trichoderma lignorum as a parasite <strong>of</strong> other soil fungi. Phytopathology 22, 837–<br />

845.<br />

84


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 85-92<br />

ADELGID GALLS ON SPRUCE AS A RESERVOIR INOCULUM<br />

SOURCE FOR THE SHOOT BLIGHT PATHOGEN Diplodia pinea<br />

Glen R. STANOSZ 1*, Denise R. SMITH 1 , S. ZHOU 2<br />

1 Department <strong>of</strong> Plant Pathology and 2 Genomics Center <strong>of</strong> Wisconsin, University <strong>of</strong> Wisconsin-<br />

Madison, WI, 53706, USA<br />

ABSTRACT<br />

* grs@plantpath.wisc.edu<br />

Diplodia pinea is a shoot blight and canker pathogen <strong>of</strong> many conifers and sporulates<br />

on killed needles, stems, and mature, opened seed cones. Although Colorado blue spruce<br />

(Picea pungens) is an atypical host, pycnidia <strong>of</strong> this fungus were observed on galls induced<br />

by the Cooley spruce gall adelgid (Adelges cooleyi). The elongate, cone-like galls that<br />

form on ends <strong>of</strong> shoots as a result <strong>of</strong> feeding by nymphs <strong>of</strong> this insect normally do not<br />

seriously impact tree health, but they can be considered unsightly. A survey was conducted<br />

to determine the incidence and abundance <strong>of</strong> inoculum <strong>of</strong> D. pinea that could be obtained<br />

from these galls on an otherwise healthy, ornamental Colorado blue spruce in Madison, WI,<br />

USA. Ten arbitrarily selected galls were collected from one branch at each <strong>of</strong> four<br />

directions in the top, middle, and bottom thirds <strong>of</strong> the tree crown (120 galls total). A<br />

washing and filtration technique was used to determine presence and estimate the numbers<br />

<strong>of</strong> conidia extracted from these galls, and species-specific PCR primers were used to<br />

confirm the identity <strong>of</strong> the pathogen. Conidia were obtained from each gall, but the number<br />

<strong>of</strong> spores varied greatly from gall to gall. Some galls yielded few spores, a result that<br />

suggests these conidia may have been produced elsewhere. In contrast, many thousands <strong>of</strong><br />

spores were obtained from galls on which pycnidia were abundant. Thus, in the absence <strong>of</strong><br />

usual host trees, insect-altered organs <strong>of</strong> an atypical host can be an alternative substrate for<br />

this pathogen and a reservoir inoculum source <strong>of</strong> D. pinea.<br />

Keywords: spruce, Picea, gall, Adelges, Diplodia, inoculum<br />

1. INTRODUCTION<br />

Elongate, cone-like galls form on ends <strong>of</strong> shoots <strong>of</strong> several Picea A. Dietr.<br />

species as a result <strong>of</strong> feeding by nymphs <strong>of</strong> the Cooley spruce gall adelgid (Adelges<br />

cooleyi Gillette) (Cumming, 1959; USDA Forest Service, 1985). This insect is<br />

native to North America and transcontinental in distribution. Primary hosts include<br />

Colorado blue spruce (P. pungens Engelm.), Englemann spruce (P. englemannii<br />

Parry ex Englem.), Sitka spruce (P. sitchensis (Bong.) Carrière), and white spruce<br />

(P. glauca (Moench) Voss. Douglas-fir (Pseudotsuga menziesii (Mirb.) Franco) is<br />

an alternate host. Similar galls are induced by the eastern spruce gall adelgid<br />

(Adelges abietis L.), which was introduced from Europe and is now found in<br />

85


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

eastern North America on native spruce species and on Norway spruce (Picea<br />

abies (L.) H. Karst). Although sometimes numerous, these galls normally do not<br />

seriously impact spruce health. They are, however, considered unsightly on trees<br />

grown as ornamentals or as Christmas trees.<br />

Although a single host species can support multiple generations <strong>of</strong> the Cooley<br />

spruce gall adelgid, its complex life cycle normally involves alternation between<br />

spruce and Douglas-fir hosts. The life history <strong>of</strong> this insect has been described in<br />

detail by Cumming (1959) and summarized by the USDA Forest Service (1985).<br />

Immature females overwinter as nymphs on the most recent year’s twigs <strong>of</strong> spruce.<br />

After emergence and maturation in spring, the “stem-mother” deposits up to 350<br />

eggs covered by a mass <strong>of</strong> white, waxy secretion. Eggs hatch in 1 to 2 weeks, and<br />

nymphs feed at bases <strong>of</strong> new needles on the elongating shoots. Young, green to<br />

purple, fleshy galls develop rapidly, elongating (up to 75 mm) and expanding in<br />

girth (up to 18 mm) to enclose the feeding nymphs. By late summer galls open to<br />

allow nymphs to move to needles where they molt into winged adults that fly to<br />

Douglas-fir. On Douglas-fir the adelgid is able to reproduce parthenogenically and<br />

sexually, overwinter as nymphs, and develop a winged stage that returns to spruce.<br />

Although feeding occurs on Douglas-fir, galls similar to those on spruce are not<br />

produced on this host.<br />

Diplodia pinea (Desmaz.) J. Kickx fil. (syn. Sphaeropsis sapinea) is a widely<br />

distributed, asexual fungal pathogen <strong>of</strong> conifers in native forests and where<br />

planted as exotics (Punithalingam and Waterston, 1970). Reports <strong>of</strong> severe<br />

damage caused by D. pinea most frequently involve pines (Pinus species),<br />

e<strong>special</strong>ly two- and three-needled pines <strong>of</strong> the subgenus Diploxylon. However, the<br />

fungus occasionally has been reported from spruce hosts or substrates (Farr et al.,<br />

1989; Punithalingham and Waterston, 1970). Rain-splashed conidia <strong>of</strong> D. pinea<br />

can be dispersed throughout the growing season (Palmer et al., 1988) and<br />

germinate quickly followed by penetration directly (Brookhouser and Peterson,<br />

1971; Chou, 1978) or through wounds. Disease may develop rapidly, or D. pinea<br />

may persist on or in asymptomatic hosts (Stanosz et al., 2005) with subsequent<br />

proliferation to cause disease under conditions that induce host stress (Stanosz et<br />

al., 2001). Damage includes seed rot and seedling collar rot, shoot blight, branch<br />

and bole cankers, crown wilt, and blue stain <strong>of</strong> sapwood (Chou, 1976; Chou, 1987;<br />

Palmer, 1991; Rees and Webber, 1988; Stanosz and Cummings Carlson, 1996).<br />

Diplodia pinea is frequently found sporulating on needles and stems it has killed,<br />

and also on mature, opened seed cones (Waterman, 1943; Peterson, 1977).<br />

Close examination <strong>of</strong> galls <strong>of</strong> the Cooley spruce gall adelgid from an<br />

ornamental Colorado blue spruce revealed presence <strong>of</strong> pycnidia on the galls (and<br />

attached needles) (Figure 1) that yielded D. pinea conidia. The objective <strong>of</strong> this<br />

study was to examine the frequency <strong>of</strong> occurrence and amount <strong>of</strong> potential<br />

inoculum <strong>of</strong> this pathogen from galls induced by the Cooley spruce gall adelgid.<br />

Procedures used were modified from those developed by Munck and Stanosz<br />

(<strong>2009</strong>) for water extraction <strong>of</strong> conidia from seed cones <strong>of</strong> pines.<br />

86


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Figure 1: Abundant pycnidia <strong>of</strong> Diplodia pinea on a gall induced by the Cooley spruce<br />

gall adelgid.<br />

2. MATERIAL AND METHODS<br />

Galls were collected in summer from a single mature Colorado blue spruce tree<br />

(approximately 12 m tall) growing as an ornamental on the University <strong>of</strong><br />

Wisconsin-Madison campus (43.07 N, 89.40 W). Shoots on this tree did not<br />

exhibit symptoms attributable to D. pinea. Two ornamental Austrian pines (Pinus<br />

nigra Arnold) that were damaged by shoot blight and bore D. pinea pycnidia on<br />

shoots and cones were located within approximately 20 m. The galls that were<br />

collected had matured to release nymphs the previous year and were dead. Ten<br />

arbitrarily selected galls were collected from one branch at each <strong>of</strong> the four<br />

cardinal directions in the top, middle, and bottom thirds <strong>of</strong> the tree crown (120<br />

galls total). Galls were bagged separately and frozen until extraction <strong>of</strong> conidia.<br />

Galls were processed individually using procedures similar to those described<br />

for pine cones by Munck and Stanosz (<strong>2009</strong>). Each gall was placed in a 100 ml<br />

plastic beaker containing 50 ml <strong>of</strong> distilled water with 2 drops <strong>of</strong> Tween 80 l -1<br />

(Fisher Scientific Company, Fair Lawn, NJ) and washed for 3 h on a rotary shaker<br />

at 110 rpm. The resulting suspension was then filtered through 0.8 µm pore size<br />

filters printed with a grid to delineate 3 mm x 3 mm squares (Cat. No.:<br />

AAWG047SP, Millipore Corporation, Billerica, MA). The cup with the gall then<br />

was rinsed with an additional 50 ml <strong>of</strong> distilled water with Tween and this liquid<br />

87


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

also was filtered. The filter for each gall was examined with the aid <strong>of</strong> a microscope<br />

for presence or absence <strong>of</strong> conidia recognized as those <strong>of</strong> D. pinea based on<br />

morphological characteristics. For three randomly selected galls from each branch,<br />

conidia were counted in five compound microscope fields randomly located within the<br />

filtered area at magnifications <strong>of</strong> 40 to 200x. Lower magnification was used for filters<br />

with relatively few conidia and higher magnification (i.e., smaller fields) was used for<br />

filters with many conidia. The number <strong>of</strong> conidia in five fields was multiplied by<br />

respective factors to adjust for total filtered area. Galls were then oven dried and<br />

weighed to also allow expression <strong>of</strong> conidial numbers on the basis <strong>of</strong> oven dry weight<br />

(odw).<br />

To confirm pathogen identity, 12 filters (corresponding to four galls from each third<br />

<strong>of</strong> the tree crown) on which numerous conidia had been deposited were selected. A<br />

piece <strong>of</strong> the filter approximately 5 mm x 5 mm was excised and placed in a<br />

microcentrifuge tube. This was ground and then DNA was directly extracted using<br />

methods described by Smith and Stanosz (1995). DNA was amplified using speciesspecific<br />

primers developed by Smith and Stanosz (2006) that allow differentiation <strong>of</strong><br />

D. pinea from the similar conifer pathogen D. scrobiculata.<br />

3. RESULTS AND DISCUSSION<br />

Every gall yielded conidia morphologically consistent with those <strong>of</strong> D. pinea,<br />

although the estimated numbers <strong>of</strong> conidia obtained varied widely. The range per gall<br />

was from 176 to 1,099,695 (mean = 249,599, standard error = 57,756). The range per<br />

gram odw was from 169 to 3,447,320 (mean = 462,691, standard error = 120,975).<br />

The numbers <strong>of</strong> galls categorized according to the estimated numbers <strong>of</strong> conidia<br />

extracted from each (for the 36 for which conidia were quantified) are displayed in<br />

Figure 2. For the majority <strong>of</strong> galls, this number was >10 4 conidia on both per gall and<br />

per gram odw bases.<br />

Figure 2: Numbers <strong>of</strong> Cooley spruce gall adelgid galls categorized by the<br />

estimated number <strong>of</strong> conidia extracted per gall.<br />

88


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The identity <strong>of</strong> the pathogen was confirmed as D. pinea. Eleven <strong>of</strong> the 12 filters<br />

that were tested were positive for this fungus. The test <strong>of</strong> one filter produced no<br />

result. No filters tested positive for D. scrobiculata.<br />

The estimated numbers <strong>of</strong> conidia extracted from galls (when expressed on the<br />

per gram odw basis) <strong>of</strong>ten were within the range or even greater than numbers<br />

reported by Munck and Stanosz (<strong>2009</strong>) for pine cones. In that study, cones from<br />

which D. pinea (or less frequently D. scrobiculata) were cultured were collected<br />

from the crowns <strong>of</strong> red pines (P. resinosa Aiton) and jack pines (P. banksiana<br />

Lamb.) in which typical shoot blight symptoms were not apparent. The greatest<br />

mean number <strong>of</strong> conidia extracted per gram odw for any <strong>of</strong> these locations was<br />

23,181, for cones from red pine crowns. The large numbers <strong>of</strong> conidia obtained<br />

from Cooley spruce gall adelgid galls indicate that, at least in this case, inoculum<br />

production by D. pinea is not limited by its exploitation <strong>of</strong> an atypical host as<br />

substrate.<br />

Previous researchers have noted a diversity <strong>of</strong> relationships between D. pinea<br />

and insects or host material altered by insects. Epidemics characterized by severe<br />

damage to Scots pine (P. sylvestris L.) in Ontario was associated with injuries<br />

resulting from feeding <strong>of</strong> the pine spittle-bug (Aphrophora parallela Say) (Haddow<br />

and Newman, 1942). In contrast, Feci et al. (2003) found D. pinea only<br />

infrequently on red pine shoots damaged by insects, primarily the red pine shoot<br />

moth Dioryctria resinosella Mutuura. Other studies provide evidence that the cone<br />

bug Gastrodes grossipes De Geer has a role in movement <strong>of</strong> this fungus to cones <strong>of</strong><br />

Austrian pine (Feci et al., 2002) and support the conclusion that the bark beetle Ips<br />

pini (Say) may vector D. pinea (Whitehill et al., 2007).<br />

Researchers also have noted a previous relationship between an apparent<br />

disease and adelgid galls on spruce. Audley and Skelly (1994) noted the<br />

occurrence <strong>of</strong> necrotic twigs <strong>of</strong> red spruce (Picea rubens Sargent) that bore galls <strong>of</strong><br />

the eastern spruce gall adelgid. A Phomopis Harter species was isolated from 14 <strong>of</strong><br />

33 dying, adelgid-galled twigs. This fungus was used to inoculate seedlings, and<br />

produced cankers in 29% <strong>of</strong> the attempts.<br />

Documentaton <strong>of</strong> the sporulation <strong>of</strong> D. pinea on insect altered, necrotic spruce<br />

tissues does not clarify the potential confusion about the ability <strong>of</strong> this fungus to<br />

infect and kill normal spruce shoots under natural conditions. Although spruces<br />

are included on lists <strong>of</strong> trees on which D. pinea has been reported (Farr et al., 1989;<br />

Punithalingam and Waterston, 1970) such sources do not provide details necessary<br />

to know if the fungus was causing disease or merely was present saprophytically.<br />

Interpretation <strong>of</strong> reports <strong>of</strong> D. pinea from spruces is further complicated by current<br />

knowledge that past references to D. pinea sensu lato may have referred to either <strong>of</strong><br />

two species (i.e., D. pinea or the similar fungus D. scrobiculata that was described<br />

by deWet et al., 2003). For example, Myren (1991) attributed killing <strong>of</strong> stems and<br />

branches <strong>of</strong> stressed black spruce (P. mariana (Mill.) B. S. P.) in an Ontario seed<br />

orchard to D. pinea (as S. sapinea). However, isolates later collected from black<br />

spruces at that seed orchard and each <strong>of</strong> four other Ontario seed orchards all were<br />

89


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

proven to be D. scrobiculata (Hausner et al., 1999). Blodgett and Stanosz (1999)<br />

did demonstrate the ability <strong>of</strong> well-characterized isolates <strong>of</strong> D. pinea and D.<br />

scrobiculata to kill shoots <strong>of</strong> small, potted Colorado blue spruce seedlings<br />

following wounding and inoculation in a greenhouse experiment. But the relative<br />

lack <strong>of</strong> reports <strong>of</strong> damage to spruces, even when grown in the same locations where<br />

more susceptible species are attacked (e.g., red pine in the Great Lakes region <strong>of</strong><br />

the USA) suggests that spruce species are infrequent hosts, or perhaps not <strong>of</strong>ten<br />

severely damaged by D. pinea. Unambiguous identification <strong>of</strong> the Diplodia<br />

species associated with disease symptoms <strong>of</strong> spruces should allow a better<br />

understanding <strong>of</strong> the relationships between these fungi and Picea species.<br />

Regardless <strong>of</strong> possible ambiguity in the status <strong>of</strong> D. pinea as a potential<br />

pathogen <strong>of</strong> spruce, galls <strong>of</strong> the Cooley spruce gall adelgid serve as a substrate.<br />

The production <strong>of</strong> a very large number <strong>of</strong> conidia on any individual gall is<br />

magnified by the potential for a single spruce crown to bear many hundreds <strong>of</strong><br />

galls. This utilization <strong>of</strong> galls by D. pinea to produce very large numbers <strong>of</strong><br />

conidia confirms that, in the absence <strong>of</strong> normal disease development, altered<br />

tissues on an atypical host tree species can be a potentially significant reservoir<br />

inoculum source for this pathogen.<br />

4. ACKNOWLEDGMENTS<br />

The authors thank Erin Brooks for collecting the galls used in this study, and<br />

with Mary Lynn Stanosz, for assistance examining galls in the laboratory.<br />

5. REFERENCES<br />

Audley, D. E., Skelly, J. M., 1994. A Phomopsis species associated with nonlethal adelgid galls on<br />

upper crown branchlets <strong>of</strong> red spruce in West Virginia. Plant Disease 78, 569-571.<br />

Blodgett, J. T., Stanosz, G. R., 1999. Differences in aggressiveness <strong>of</strong> Sphaeropsis sapinea RAPD<br />

marker group isolates on several conifers. Plant Disease 83, 853-856.<br />

Brookhouser, L. W., Peterson, G. W., 1971. Infection <strong>of</strong> Austrian, Scots, and ponderosa pines by<br />

Diplodia pinea. Phytopathology 61, 409-414.<br />

Chou, C. K. S., 1976. A shoot dieback in Pinus radiata caused by Diplodia pinea. I. Symptoms,<br />

disease development, and isolation <strong>of</strong> the pathogen. New Zealand Journal <strong>of</strong> Forest<br />

Science 6, 72-79.<br />

Chou, C. K. S., 1978. Penetration <strong>of</strong> young stems <strong>of</strong> Pinus radiata by Diplodia pinea. Physiological<br />

Plant Pathology 12, 189-192.<br />

Chou, C. K. S., 1987. Crown wilt <strong>of</strong> Pinus radiata associated with Diplodia pinea infection <strong>of</strong><br />

woody stems. European Journal <strong>of</strong> Forest Pathology 17, 398-411.<br />

Cumming, M. E. P., 1959. The biology <strong>of</strong> Adelges cooleyi (Gill.) (Homoptera: Phylloxeridae).<br />

Canadian Entomologist 91, 601-617.<br />

90


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

de Wet, J. T., Burgess, T., Slippers, B., Preisig, O., Wingfield, B. D., Wingfield, M. J., 2003.<br />

Multiple gene genealogies and microsatellite markers reflect relationships between<br />

morphotypes <strong>of</strong> Sphaeropsis sapinea and distinguish a new species <strong>of</strong> Diplodia.<br />

Mycological Research 107, 557-566.<br />

Farr, D., Bills, G., Chamuris, G., Rossman, A., 1989. Fungi on Plants and Plant Products in the<br />

United States. APS Press, St. Paul. 1252 pp.<br />

Feci, E., Battisti, A., Capretti, P., Tegli, S., 2002. An association between the fungus Sphaeropsis<br />

sapinea and the cone bug Gastrodes grossipes in cones <strong>of</strong> Pinus nigra in Italy. Forest<br />

Pathology 32, 241-247.<br />

Feci, E., Smith, D., Stanosz, G. R., 2003. Association <strong>of</strong> Sphaeropsis sapinea with insect-damaged<br />

red pine shoots and cones. Forest Pathology 33, 7-13.<br />

Haddow, W. R., and Newman, F. S., 1942. A disease <strong>of</strong> the Scots pine (Pinus sylvestris L.) caused<br />

by the fungus Diplodia pinea Kickx associated with the pine spittle-bug (Aphrophora<br />

parallela Say.). Transactions <strong>of</strong> the Royal Canadian Institute 24 (Part 1), 1-18.<br />

Hausner, G., Hopkin, A. A., Davis, C. N., Reid, J., 1999. Variation in culture and rDNA among<br />

isolates <strong>of</strong> Sphaeropsis sapinea from Ontario and Manitoba. Canadian Journal <strong>of</strong> Plant<br />

Pathology 21, 256-264.<br />

Munck, I., and Stanosz, G., <strong>2009</strong>. Quantification <strong>of</strong> conidia <strong>of</strong> Diplodia spp. extracted from red and<br />

jack pine cones. Plant Disease 93, 81-86.<br />

Myren, D. T., 1991. Sphaeropsis sapinea on black spruce in Ontario. Plant Disease 75, 664.<br />

Palmer, M. A., McRoberts, R. E., Nicholls, T. H., 1988. Sources <strong>of</strong> inoculum <strong>of</strong> Sphaeropsis sapinea<br />

in forest tree nurseries. Phytopathology 78, 831-835.<br />

Palmer, M. A., 1991. Isolate types <strong>of</strong> Sphaeropsis sapinea associated with main stem cankers and<br />

top-kill <strong>of</strong> Pinus resinosa in Minnesota and Wisconsin. Plant Disease 75, 507-510.<br />

Peterson, G. W., 1977. Infection, epidemiology, and control <strong>of</strong> Diplodia blight <strong>of</strong> Austrian,<br />

ponderosa, and Scots pines. Phytopathology 67, 511-514.<br />

Punithalingam, E., Waterston, J. M., 1970: Diplodia pinea. No. 273 in: Descriptions <strong>of</strong> Pathogenic<br />

Fungi and Bacteria, Commonwealth Mycological Institute, Kew, Surrey, England.<br />

Rees, A. A., Webber, J. F., 1988. Pathogenicity <strong>of</strong> Sphaeropsis sapinea to seed, seedlings, and<br />

saplings <strong>of</strong> some central American pines. Transactions <strong>of</strong> the British Mycological Society<br />

91, 273-277.<br />

Smith, D. R., Stanosz, G. R., 1995. Confirmation <strong>of</strong> two distinct populations <strong>of</strong> Sphaeropsis sapinea<br />

in the north central United States using RAPDs. Phytopathology 85, 699-704.<br />

Smith, D. R., Stanosz, G. R., 2006. A species-specific PCR assay for detection <strong>of</strong> Diplodia pinea and<br />

D. scrobiculata in dead red and jack pines with collar rot symptoms. Plant Disease 90,<br />

307-313.<br />

Stanosz, G. R., Cummings Carlson, J., 1996. Association <strong>of</strong> mortality <strong>of</strong> recently planted seedlings<br />

and established saplings in red pine plantations with Sphaeropsis collar rot. Plant Disease<br />

80, 750-753.<br />

Stanosz, G. R., Blodgett , J. T., Smith, D. R., Kruger, E. L., 2001. Water stress and Sphaeropsis<br />

sapinea as a latent pathogen <strong>of</strong> red pine seedlings. New Phytologist 149, 531-538.<br />

Stanosz, G. R., Smith, D. R., Albers, J. S., 2005. Surveys for asymptomatic persistence <strong>of</strong><br />

Sphaeropsis sapinea on or in stems <strong>of</strong> red pine seedlings from seven Great Lakes region<br />

nurseries. Forest Pathology 35, 233-244.<br />

91


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

USDA Forest Service, 1985. Insects <strong>of</strong> Eastern Forests. Misc. Publ. 1426. Washington, D.C. 608<br />

pp.<br />

Waterman, A. M., 1943. Diplodia pinea, the cause <strong>of</strong> a disease <strong>of</strong> hard pines. Phytopathology 33,<br />

1018-1031.<br />

Whitehill, J. G. A., Lehman, J. S., and Bonello, P., 2007. Ips pini (Curculionidae: Scolytinae) is a<br />

vector <strong>of</strong> the fungal pathogen, Sphaeropsis sapinea (Coelomycetes), to Austrian pines,<br />

Pinus nigra (Pinaceae). Environmental Entomology 36, 114-120.<br />

92


Dieback And Canker Diseases<br />

93


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

94


Dieback Diseases<br />

95


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

96


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 97-119<br />

THE CURRENT SITUATION OF ASH DIEBACK CAUSED BY<br />

Chalara fraxinea IN AUSTRIA<br />

Thomas Kirisits 1* , Michaela Matlakova 1 , Susanne Mottinger-Kroupa 1 ,<br />

Thomas L. Cech 2 , Erhard Halmschlager 1<br />

1* Institute <strong>of</strong> Forest Entomology, Forest Pathology and Forest Protection (IFFF), Department <strong>of</strong><br />

Forest and Soil Science, University <strong>of</strong> Natural Resources and Applied Life Sciences, Vienna<br />

(BOKU)Hasenauerstraße 38, A-1190 Vienna, Austria; 2 Federal Research and Training Centre for<br />

Forests, Natural Hazards and Landscape (BFW), Department <strong>of</strong> Forest Protection, Seckendorff-<br />

Gudent-Weg 8, A-1131 Vienna, Austria<br />

ABSTRACT<br />

*thomas.kirisits@boku.ac.at<br />

In many parts <strong>of</strong> Europe common ash, Fraxinus excelsior, is presently affected by a<br />

serious dieback <strong>of</strong> shoots, twigs and branches, causing decline and mortality <strong>of</strong> trees <strong>of</strong> all<br />

age classes. Initially thought to be primarily incited by abiotic damaging factors,<br />

accumulating evidence suggests that ash dieback is a new infectious disease caused by the<br />

hyphomycete Chalara fraxinea and its teleomorphic state, Hymenoscyphus albidus. In<br />

Austria, ash dieback was first observed in 2005 and in 2008 it occurred in all Austrian<br />

provinces. In heavily affected forests mortality is common amongst saplings and young<br />

trees. Moreover, in some areas dying <strong>of</strong> mature trees has started to occur. Chalara fraxinea<br />

was for the first time recorded in Austria in June 2007. Subsequent surveys have shown that<br />

the pathogen is widespread in the country. Until June <strong>2009</strong> it was isolated from<br />

symptomatic ash trees at 82 localities in eigth out <strong>of</strong> the nine Austrian provinces. Apart<br />

from F. excelsior, C. fraxinea was isolated from narrow-leaved ash, F. angustifolia subsp.<br />

danubialisand from weeping ash, F. excelsior ‘Pendula’. Chalara fraxinea was consistently<br />

isolated at high frequencies from ash shoots, twigs and stems showing early symptoms <strong>of</strong><br />

disease. In inoculation experiments using potted F excelsior and F. angustifolia seedlings,<br />

Kochs postulates were fulfilled for C. fraxinea, clearly suggesting that this fungus is the<br />

primary causal agent <strong>of</strong> ash dieback. It also displayed pathogenicity to Fraxinus ornus<br />

seedlings. Besides reviewing the situation <strong>of</strong> ash dieback in Austria and summarizing the<br />

results <strong>of</strong> some <strong>of</strong> our research since 2007, we generally review the knowledge on this<br />

emerging disease in Europe, describe its symptoms, present a hypothetic disease cycle for<br />

ash dieback and discuss options for disease management.<br />

Keywords: Hymenoscyphus albidus, Fraxinus excelsior, Fraxinus angustifolia subsp.<br />

danubialis, emerging disease, fungal disease, new forest health problem<br />

97


1. INTRODUCTION<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Since the early 1990s common ash, Fraxinus excelsior, has been affected by an<br />

unprecedented, serious dieback <strong>of</strong> shoots, twigs and branches, causing decline and<br />

mortality <strong>of</strong> trees <strong>of</strong> all age classes. Ash dieback was first noticed around 1992 in<br />

Poland (Kowalski and Holdenrieder, 2008), where it has been causing serious<br />

damage (Przybył, 2002; Kowalski, 2006; Kowalski and Holdenrieder, 2008). It was<br />

subsequently recorded in many other European countries and threatens common<br />

ash in large parts <strong>of</strong> its distribution range. By <strong>2009</strong> ash dieback has been reported<br />

to occur in Lithuania (Lygis et al., 2005), Latvia (T. Gaitnieks, personal<br />

communication), Estonia (M. Hanso and R. Drenkhan, personal communication),<br />

area Kaliningrad (Russia, R. Vasaitis, personal communication), Danmark<br />

(Thomsen et al., 2007; Skovsgaard et al., <strong>2009</strong>), Sweden (Bakys et al., <strong>2009</strong>a;<br />

<strong>2009</strong>b), Norway (Talgø et al., <strong>2009</strong>), Finland (J. Hantula, personal<br />

communication), Germany (Schumacher et al., 2007), Slovakia (A. Kunca,<br />

personal communication), Czech Republic (Jankovský et al., 2008), Austria (Cech,<br />

2006b; Halmschlager and Kirisits, 2008; Kirisits et al., 2008a; 2008b; <strong>2009</strong>),<br />

Hungary (Szabó, 2008), Slovenia (Ogris et al., <strong>2009</strong>), Rumania (D. Chira, personal<br />

communication), Switzerland (Engesser et al., <strong>2009</strong>) and eastern France<br />

(Chandelier et al., <strong>2009</strong>; Ioos et al., <strong>2009</strong>) It can be expected that this emerging<br />

forest health problem will in the future also occur in other parts <strong>of</strong> Europe, where it<br />

has thus-far not been found.<br />

Initially, ash dieback was suspected to be primarily incited by abiotic damaging<br />

factors (frost, drought and abrupt changes <strong>of</strong> periods with warm and cold weather<br />

conditions), with secondary, weakly virulent fungal pathogens and endophytes<br />

contributing to the syndrome (Przybył, 2002; Pukacki and Przybyl, 2005; Cech,<br />

2006b; Cech et al., 2007). This view changed with the discovery and description <strong>of</strong><br />

the anamorphic fungus Chalara fraxinea that was in Poland frequently isolated<br />

from shoots in the early phase <strong>of</strong> the pathological process (Kowalski, 2006). In the<br />

meanwhile, C. fraxinea has been detected in many <strong>of</strong> the above mentioned<br />

countries and accumulating evidence suggests that it is the cause <strong>of</strong> ash dieback<br />

(Schumacher et al., 2007; Halmschlager and Kirisits, 2008; Szabó, 2008; Kowalski<br />

and Holdenrieder, <strong>2009</strong>a; Talgø et al., <strong>2009</strong>; Bakys et al., <strong>2009</strong>a; <strong>2009</strong>b; Engesser<br />

et al., <strong>2009</strong>; Ogris et al., <strong>2009</strong>; Kirisits et al. <strong>2009</strong>; Chandelier et al., <strong>2009</strong>; Ioos et<br />

al., <strong>2009</strong>).<br />

Because <strong>of</strong> the sudden appearance, the rapid spread and the high intensity <strong>of</strong> ash<br />

dieback, C. fraxinea was thought by some forest pathologists to be an alien<br />

invasive organism (Halmschlager and Kirisits, 2008; Kirisits and Halmschlager,<br />

2008; Kowalski and Holdenrieder, 2008; Ogris et al., <strong>2009</strong>). However,<br />

Hymenoscyphus albidus, a discomycete native to Europe has recently been<br />

identified as the teleomorph <strong>of</strong> C. fraxinea (Kowalski and Holdenrieder, <strong>2009</strong>b).<br />

This ascomycete fungus has been known since 1850 as harmless decomposer <strong>of</strong><br />

leaf rachises (referred to as leaf petioles by Kowalski and Holdenrieder [<strong>2009</strong>b] but<br />

we think that ‘leaf rachis’ is the more appropriate botanical term) <strong>of</strong> common ash<br />

98


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

and it is unknown, why this fungus suddenly causes a serious, emerging disease on<br />

F. excelsior (Kowalski and Holdenrieder, <strong>2009</strong>b). Kowalski and Holdenrieder<br />

(<strong>2009</strong>b) proposed that the fungus they assigned to the morphospecies H. albidus<br />

may have undergone genetic change by mutation or hybridization with an unknown<br />

introduced species. Another possibility is that the teleomorph <strong>of</strong> C. fraxinea is not<br />

the ‘original’ H. albidus, but an exotic invasive species, that is morphologically<br />

virtually identical to H. albidus. Finally, the fungus may show unprecedented<br />

aggressiveness towards ash because <strong>of</strong> environmental factors or/and weather<br />

extremes, that could either have predisposed the host trees to fungal attack or<br />

provided ideal conditions for fungal infections (Kowalski and Holdenrieder,<br />

<strong>2009</strong>b). These three theories are the conceptual basis for future research regarding<br />

the question what triggered the epidemic <strong>of</strong> H. albidus and its anamorphic state C.<br />

fraxinea.<br />

Ash dieback is presently amongst the most important forest health problems in<br />

Austria. Here we review the situation <strong>of</strong> this emerging disease in this Central<br />

European country. We also summarize some research that has been conducted<br />

since 2007, present a hypotheticAL disease cycle for ash dieback and discuss<br />

options for disease management.<br />

2. ASH SPECIES IN AUSTRIA<br />

Three ash species are native in Austria including common ash, Fraxinus<br />

excelsior, also known as European ash, narrow-leaved ash, F. angustifolia subsp.<br />

danubialis and flowering ash, F. ornus (Adler et al., 1994). While European ash is<br />

widespread on appropriate sites in many parts <strong>of</strong> the country (Schadauer, 1994), the<br />

two other species are at the edge <strong>of</strong> their distribution ranges in Austria and are thus<br />

rare (Adler et al., 1994; Zukrigl, 1997).<br />

With a share <strong>of</strong> 2.5% (based on the number <strong>of</strong> trees) and 1.8% (based on the<br />

growing stock) ash species are the third most frequent group <strong>of</strong> hardwood species<br />

in managed forests in Austria (Table 1). All three species are included in the<br />

relative proportions reported in Table 1, but the vast majority <strong>of</strong> the share <strong>of</strong><br />

Fraxinus spp. refers to F. excelsior. Based on the growing stock only European<br />

beech (Fagus sylvatica) and oak species (mainly Quercus petraea and Quercus<br />

robur) occur more frequently than Fraxinus spp. and based on the number <strong>of</strong> trees<br />

F. sylvatica dominates amongst hardwood species, while European hornbeam<br />

(Carpinus betulus) is slightly more frequent than ash. The share <strong>of</strong> ash in the nine<br />

Austrian provinces is shown in Table 1. Ash is particularly abundant in the<br />

provinces Upper Austria, Lower Austria, Vorarlberg and Vienna and occurs least<br />

frequently in the province Tyrol that is to a large extent located in areas <strong>of</strong> the Alps<br />

that are inappropriate for the growth <strong>of</strong> F. excelsior.<br />

99


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Table 1: Share (%) <strong>of</strong> ash, based on number <strong>of</strong> trees and based on growing stock in<br />

managed forests in the nine Austrian provinces and in entire Austria. The data include all<br />

three ash species native in Austria, but the vast majority is F. excelsior, while the<br />

proportions <strong>of</strong> F. angustifolia and F. ornus are negligible. Source: Austrian Forest<br />

Inventory 2000-2002, Federal Research and Training Centre for Forests, Natural Hazards<br />

and Landscape (BFW), Department <strong>of</strong> Forest Inventory<br />

(http://bfw.ac.at/rz/bfwcms.web?dok=35).<br />

Austrian province Based on number <strong>of</strong> trees Based on growing stock<br />

Burgenland 2.1 1.2<br />

Carinthia 1.7 1.0<br />

Lower Austria 3.4 2.9<br />

Salzburg 2.0 1.1<br />

Styria 1.6 1.2<br />

Tyrol 0.4 0.1<br />

Upper Austria 5.2 3.7<br />

Vienna 9.4 7.5<br />

Vorarlberg 4.6 2.2<br />

Austria total 2.5 1.8<br />

Common ash has its optimum on moist, nutrient-rich sites, but in areas with<br />

limestone as geological bedrock it can also occur on drier sites (Mayer, 1984). It is<br />

mainly found from the lowlands up to elevations <strong>of</strong> 900 m asl. (Schadauer, 1994;<br />

Nationalpark Kalkalpen, 2007) In the Alps it rarely occurs at elevations higher than<br />

1200 m asl. (Mayer, 1984; Schadauer, 1994; Nationalpark Kalkalpen, 2007).<br />

Fraxinus excelsior occurs in a large number <strong>of</strong> forest types. It is a characteristic<br />

component <strong>of</strong> floodplain forests along big rivers, forests along streams and in<br />

glens, but also occurs in other hardwood-dominated forests on moist and<br />

sometimes drier sites (Jelem, 1974; Mayer, 1974; 1984; Nationalpark Kalkalpen,<br />

2007). For forest owners managing riparian forests, along the Danube for example,<br />

common ash is usually the most economically important timber species (Jelem,<br />

1974). In many areas, e<strong>special</strong>ly on sites at lower elevations it has received much<br />

attention for the production <strong>of</strong> valuable timber and as an alternative to Norway<br />

spruce (Picea abies) (Wolf and Jasser, 2003). It is also important for various<br />

ecosystem services, for example stabilization <strong>of</strong> riverbanks and slopes (Mayer,<br />

1984). Moreover, F. excelsior is an attractive and appreciated landscape tree in the<br />

countryside and shade tree in urban areas. Due to its nutrient-rich foliage it was the<br />

most preferred tree species for lopping in the Alps, providing fresh or dry fodder<br />

for cattle, sheep and other domestic animals.<br />

Narrow-leaved ash, Fraxinus angustifolia subsp. danubialis (hereafter referred<br />

to just as ‘F. angustifolia’), is mainly distributed in lowland areas <strong>of</strong> south-eastern<br />

Europe. In eastern Austria it reaches one <strong>of</strong> its distribution limits and is therefore<br />

rare (Jelem, 1974; 1975; Adler et al., 1994; Zukrigl, 1997). It forms part <strong>of</strong> riparian<br />

forest ecosystems along the lower reaches <strong>of</strong> the rivers March, Danube, Fischa<br />

100


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

(province Lower Austria) and Leitha (provinces Lower Austria and Burgenland)<br />

(Jelem, 1974; 1975; Adler et al., 1994; Zukrigl, 1997). Hybrids between<br />

F. angustifolia and F. excelsior have been reported to occur in areas in Austria,<br />

where the distribution ranges <strong>of</strong> the two species overlap (Jelem, 1974; 1975).<br />

While it is mainly a botanical curiosity and <strong>of</strong> interest for nature conservation, this<br />

ash species is an economical important timber species in floodplain forests along<br />

the March (Jelem 1975; Damm, 1997).<br />

Flowering ash mainly occurs in southern and south-eastern Europe, but its<br />

natural distribution range just reaches southern Austria (Mayer, 1974; Adler et al.,<br />

1994; Zukrigl, 1997). It is mainly found in southern and eastern Carinthia (Adler et<br />

al., 1994), where it usually occurs on steep, rocky, warm and dry sites on limestone<br />

(Zukrigl, 1997). Together with European hop-hornbeam, Ostrya carpinifolia,<br />

F ornus <strong>of</strong>ten forms dense bush forests on such sites (Mayer et al., 1974; Zukrigl,<br />

1997). Apart from Carinthia, flowering ash also occurs in other Austrian provinces<br />

(Eastern and Northern Tyrol, Styria, Lower Austria and Burgenland), where it is<br />

generally rare (Adler et al., 1994; Zukrigl, 1997). Most <strong>of</strong> the occurrences in these<br />

provinces are likely not native, but F. ornus has become naturalized in some areas<br />

(Zukrigl, 1997). This ash species has no importance for <strong>forestry</strong> in Austria, but is<br />

<strong>of</strong> interest for nature conservation, as it forms part <strong>of</strong> rare and ecologically valuable<br />

forest types. It is occasionally used as shade and ornamental tree and planted in<br />

shelterbelts (Zukrigl, 1974).<br />

Green ash (F. pennsylvanica) and white ash (F. americana), two introduced<br />

species from North America are occasionally planted as ornamentals (Adler et al.,<br />

1994). Some decades ago they also received some interest as plantation trees in<br />

floodplain forests, but as their growth potential and timber quality did not fulfill the<br />

expectations <strong>of</strong> foresters they are presently no longer planted. Fraxinus<br />

pennsylvanica has become naturalized in some riparian areas, e<strong>special</strong>ly in those<br />

along the lower reach <strong>of</strong> the Danube and the March (Essl and Rabitsch, 2002). It is<br />

therefore considered as alien invasive species in Austria (Essl and Rabitsch, 2002;<br />

Essl et al., 2006).<br />

3. SYMPTOMS OF ASH DIEBACK<br />

The symptoms <strong>of</strong> ash dieback share characteristics <strong>of</strong> a bark disease, a vascular<br />

wilt disease and a leaf disease (Figure 1 and 2; Thomsen et al., 2007; Halmschlager<br />

and Kirisits, 2008; Kirisits et al., 2008a; 2008b; Kowalski and Holdenrieder, 2008;<br />

Szabó, 2008; Bakys et al., <strong>2009</strong>b; Engesser et al, <strong>2009</strong>; Kirisits and Cech, <strong>2009</strong>;<br />

Kirisits et al., <strong>2009</strong>; Ogris et al., <strong>2009</strong>). Typical symptoms occur in the bark,<br />

phloem and wood <strong>of</strong> shoots, twigs, branches and stems as well as on leaves <strong>of</strong> ash<br />

trees. Necrotic lesions and wood discoloration can also extend into the roots, from<br />

where the fungus infects coppice sprouts, but the disease clearly starts at aboveground<br />

parts <strong>of</strong> the tree (Kowalski and Holdenrieder, 2008).<br />

101


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 1. A stand <strong>of</strong> common ash severely affected by shoot, twig and branch dieback<br />

(Laussa, Upper Austria, July 2007).<br />

The most obvious symptom is dieback <strong>of</strong> shoots, twigs and branches (Fig. 1).<br />

Shoot dieback is caused by localized necrotic lesions that initially are small, but as<br />

they expand they girdle the phloem and sapwood occlusion occurs, too. When<br />

phloem girdling and sapwood occlusion take place in winter time, shoots do not<br />

flush in spring, however, when they happen during the vegetation period,<br />

simultaneous wilting <strong>of</strong> leaves above the lesions occurs (Fig. 2A). Leaves then dry,<br />

turn brown to black and remain attached to the shoots for a long time. Elongated,<br />

<strong>of</strong>ten elliptical necrotic lesions and cankers in the bark are characteristic symptoms<br />

<strong>of</strong> ash dieback (Fig. 2B and 2C). These lesions either form around a dead side twig<br />

(Fig. 2B) or occur around a leaf scar (Fig. 2C). On larger shoots, twigs, branches as<br />

well as on younger stems, the tree <strong>of</strong>ten defends itself against the pathogen attack,<br />

at least for some time, leading to perennial cankers.<br />

102


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Necrotic lesions and cankers are usually accompanied by brownish to grayish<br />

discoloration <strong>of</strong> the wood (Fig. 2D) that frequently extends longitudinally beyond<br />

necrotic areas in the bark. Diseased trees react with prolific formation <strong>of</strong> epicormic<br />

shoots and the silhouettes <strong>of</strong> heavily affected trees look tousled and distorted<br />

(Fig. 1). Chalara fraxinea also causes symptoms on leaves, resulting from direct<br />

leaf infections (Kirisits et al., 2008b; Ogris et al., <strong>2009</strong>; Bakys et al., <strong>2009</strong>b). These<br />

symptoms include brown to blackish necrotic lesions on leaf rachises and leaflet<br />

veins, followed by wilting <strong>of</strong> parts <strong>of</strong> the leaves distal to the necrotic lesions. Early<br />

leaf shedding is <strong>of</strong>ten a consequence <strong>of</strong> these leaf symptoms.<br />

4. ASH DIEBACK IN AUSTRIA<br />

In Austria, first unambiguous observations <strong>of</strong> ash dieback were made in 2005,<br />

mainly on young trees (Cech, 2006a; 2006b). From 2006 to 2007 damage levels<br />

increased dramatically, particularly in the provinces Lower and Upper Austria as<br />

well as Styria (Cech and Hoyer-Tomiczek, 2007; Hagen, 2007; Fachabteilung<br />

Forstwesen-Forstdirektion, <strong>2009</strong>). In 2008 the phenomenon was widespread and<br />

symptoms were observed in all Austrian provinces (Kirisits et al., 2008a; Kirisits<br />

and Cech, <strong>2009</strong>).<br />

Prior to the widespread occurrence <strong>of</strong> dieback <strong>of</strong> shoots, twigs and branches,<br />

early leaf shedding on ash trees, occurring already in late August and early<br />

September, was observed in parts <strong>of</strong> the provinces Lower and Upper Austria in<br />

2005 and Styria in 2006 (Cech, 2005; Hagen, 2005; Fachabteilung Forstwesen-<br />

Forstdirektion, <strong>2009</strong>). In the following year, thus in 2006 in Lower and Upper<br />

Austria and in 2007 in Styria, ash dieback was for the first time recorded at high<br />

intensity (Cech, 2006b; Hagen, 2007; Fachabteilung Forstwesen-Forstdirektion,<br />

<strong>2009</strong>). Originally, this early shedding <strong>of</strong> F. excelsior leaves was thought to be<br />

caused by powdery mildews (Phyllactinia fraxini) and other micr<strong>of</strong>ungi (Cech,<br />

2005). However, as C. faxinea is now known to cause also symptoms on ash leaves<br />

(Thomsen et al., 2007; Bakys et al., <strong>2009</strong>b; Kirisits and Cech, <strong>2009</strong>; Ogris et al.,<br />

<strong>2009</strong>), it is likely that the episodes <strong>of</strong> early leaf shedding in 2005 and 2006 were<br />

the first obvious indications <strong>of</strong> ash dieback. In 2008 leaf symptoms and early leaf<br />

shedding occurred again, for example in parts <strong>of</strong> Styria (Fachabteilung Forstwesen-<br />

Forstdirektion, <strong>2009</strong>).<br />

Incidence and severity <strong>of</strong> ash dieback varies considerably in different parts <strong>of</strong><br />

the country. It appears to be most serious in the Northern Limestone Alps in the<br />

provinces Lower and Upper Austria, Styria and Salzburg. Likely, there are still<br />

areas, where the disease does not occur, e<strong>special</strong>ly in parts <strong>of</strong> the Alps, where<br />

common ash is present at a low density (Kirisits and Cech, <strong>2009</strong>). Apart from<br />

F. excelsior, dieback also occurs on narrow-leaved ash and weeping ash (Fraxinus<br />

excelsior ‘Pendula’), an ornamental variety <strong>of</strong> common ash (Kirisits et al., 2008a;<br />

<strong>2009</strong>; Kirisits and Cech, <strong>2009</strong>). No symptoms have thus-far been observed on<br />

103


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

F. ornus as well as the exotic F. pennsylvanica and F. americana (Kirisits, 2008;<br />

Kirisits et al., 2008a; Kirisits and Cech, <strong>2009</strong>).<br />

The disease occurs on ash trees <strong>of</strong> all ages, both on natural regeneration and<br />

planted trees and on the entire spectrum <strong>of</strong> sites and forest types, where ash is<br />

found (Cech, 2008). Ash dieback is damaging on both forest and shade trees and it<br />

also causes serious problems in nurseries, where rearing <strong>of</strong> healthy ash seedlings<br />

has become difficult, if not impossible. In heavily affected forests mortality is<br />

common amongst saplings and young trees. Moreover, in some areas dying <strong>of</strong><br />

mature trees has started to occur. It is expected that the future use <strong>of</strong> F. excelsior as<br />

economically and ecologically valuable noble hardwood species will be<br />

substantially impaired by this emerging disease.<br />

Figure 2. Symptoms <strong>of</strong> ash dieback: (A) Wilting <strong>of</strong> leaves due to girdling <strong>of</strong> the phloem<br />

and sapwood occlusion, (B) A necrotic lesion in the bark with a small dead twig in the<br />

centre, (C) A necrotic lesionr in the bark with a leaf scar (arrow) in the centre, (D)<br />

Discoloration <strong>of</strong> the wood.<br />

104


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

A number <strong>of</strong> surveys are presently underway to obtain more precise information<br />

on the geographical distribution, incidence, severity and temporal development <strong>of</strong><br />

ash dieback in Austria. In 2007 and 2008 monitoring plots have been established in<br />

Lower Austria (Cech, 2008) and additional plots in other provinces will be<br />

installed in <strong>2009</strong>. On these plots, disease intensity will be monitored on<br />

permanently marked ash trees over the next years. From <strong>2009</strong> onwards the disease<br />

is also included in the ‘Documentation <strong>of</strong> forest damage factors’ (German:<br />

‘Dokumentation der Waldschädigungsfaktoren – DWF’), a monitoring system that<br />

is based on expert opinions and appraisals <strong>of</strong> staff <strong>of</strong> the district forest authorities<br />

(Steyrer et al., 2008). Finally, assessments on the geographical distribution and<br />

incidence <strong>of</strong> ash dieback will be carried out as part <strong>of</strong> the field work <strong>of</strong> the Austrian<br />

Forest Inventory.<br />

5. Chalara fraxinea ASSOCIATED WITH ASH DIEBACK IN AUSTRIA<br />

Starting in June 2007, we aimed to examine the role <strong>of</strong> C. fraxinea in ash<br />

dieback in Austria. Shoots, twigs, branches, stems and leaves <strong>of</strong> young F. excelsior<br />

trees showing symptoms <strong>of</strong> the disease were collected in many different parts <strong>of</strong><br />

the country. From January 2008 onwards <strong>special</strong> attention was given to collect only<br />

samples from trees showing early symptoms <strong>of</strong> ash dieback, particularly shoots<br />

with small, localized necrotic phloem lesions. About 4 to 6 cm long segments,<br />

containing the transition between necrotic and healthy phloem tissues and/or<br />

discolored and healthy xylem were cut from symptomatic ash organs. These<br />

segments were surface sterilized as described by Kowalski (2006). Thereafter, the<br />

outer bark was carefully peeled <strong>of</strong>f and 3 to 10 mm small discs containing wood<br />

and phloem tissues were cut under aseptic conditions and put onto malt extract agar<br />

(MEA; 20 g malt extract, 16g agar, 1000 ml tap water supplemented after<br />

autoclaving with 100 mg streptomycin sulphate). In the earlier isolation series until<br />

January 2008, the outer bark was not peeled <strong>of</strong>f and not discs, but small pieces <strong>of</strong><br />

phloem or wood were removed and placed on MEA plates.<br />

Initially the isolation plates were incubated at room temperature (23-25°C), but<br />

from August 2008 onwards they were immediately stored at low temperatures<br />

(approximately +4°C) in refrigerators. At cool temperatures, the growth <strong>of</strong> many<br />

fungi competing with C. fraxinea is more inhibited than that <strong>of</strong> C. fraxinea itself.<br />

In addition, phialophore formation <strong>of</strong> C. fraxinea is greatly enhanced by low<br />

temperatures (Halmschlager and Kirisits, 2008; Kirisits et al., 2008a). Both factors<br />

increase the likelihood to detect the fungus. Chalara fraxinea was identified based<br />

on morphological characteristics (colony morphology, phialophores and conidia;<br />

Kowalski, 2006; Halmschlager and Kirisits, 2008; Kowalski and Holdenrieder,<br />

2008; Kirisits et al., 2008a).<br />

Chalara fraxinea was for the first time isolated in Austria in June 2007, at one<br />

locality in Upper Austria (Edt bei Lambach, 48°06'51'' N, 13°53'29'' E) and another<br />

one in Styria (Altaussee, 47°38'34'' N, 13°45'36'' E) (Halmschlager and Kirisits,<br />

105


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

2008; Kirisits and Halmschlager, 2008). Subsequent surveys have shown that the<br />

pathogen is widespread in the country and apparently occurs everywhere, where<br />

ash dieback is present. Until June <strong>2009</strong> the fungus was obtained from symptomatic<br />

ash trees at 82 localities in eight out <strong>of</strong> the nine Austrian provinces (Table 1). The<br />

differences in the number <strong>of</strong> records <strong>of</strong> C. fraxinea in the various provinces<br />

(Table 1) do not allow inferring about the intensity <strong>of</strong> ash dieback in various parts<br />

<strong>of</strong> Austria, but just reflect differences in the intensity <strong>of</strong> sampling. In the future it is<br />

planned to examine more samples and sites in those provinces, where extensive<br />

collections have thus-far not been conducted, e<strong>special</strong>ly in Tyrol, Vorarlberg,<br />

Burgenland and Carinthia<br />

In the early isolation series carried out in 2007 C. fraxinea was rarely isolated,<br />

because samples were mainly collected from ash trees showing relatively late<br />

symptoms <strong>of</strong> disease. We suppose that on such plant material the slow growing<br />

C. fraxinea is in most cases already outcompeted by fast-growing endophytic and<br />

saprotrophic fungi (Kowalski and Holdenrieder, 2008). However, when isolations<br />

were made from shoots, twigs and stems showing early symptoms <strong>of</strong> disease,<br />

C. fraxinea was the most consistently and most frequently isolated fungus and in<br />

most cases the only one that was recovered. For example, isolation frequencies <strong>of</strong><br />

C. fraxinea at ten localities in six Austrian provinces ranged from 81% to 100% <strong>of</strong><br />

the examined shoots and necrotic lesions (Table 2). Overall, the fungus was<br />

obtained from 94% <strong>of</strong> the samples, from which isolations were made. Isolation <strong>of</strong><br />

C. fraxinea was successful throughout the year, given that samples were collected<br />

from ash trees showing early disease symptoms (Table 2).<br />

Apart from F. excelsior, C. fraxinea was isolated from young, planted<br />

F. angustifolia trees in floodplain areas along the river Morava near<br />

Hohenau/March and from symptomatic seedlings <strong>of</strong> this species in a nursery in<br />

Lower Austria (Kirisits et al., <strong>2009</strong>; Table 2). In addition, it was obtained at a few<br />

localities from F. excelsior ‘Pendula’ (Table 2). To our knowledge, these are the<br />

first and thus-far only European records <strong>of</strong> the fungus from hosts other than<br />

F. excelsior. The fungal isolations from Fraxinus ssp. have shown that C. fraxinea<br />

is associated with early symptoms <strong>of</strong> ash dieback, as it is typical for the primary<br />

causal agent <strong>of</strong> a plant disease. These results agree well with other studies in<br />

Europe, in which this fungus was consistently isolated or detected with molecular<br />

markers from diseased ash trees (Kowalski, 2006; Bakys et al., <strong>2009</strong>b; Ioos et al.,<br />

<strong>2009</strong>; Chandelier et al., <strong>2009</strong>).<br />

In May 2008 potted, one-year-old common ash seedlings were woundinoculated<br />

with five C. fraxinea isolates and in June 2008 with five other strains.<br />

Similarly, potted, two-year-old narrow-leaved ash seedlings were inoculated with<br />

one C. fraxinea isolate in May 2008 (Kirisits et al., <strong>2009</strong>) and with two other<br />

isolates in June 2008. Inoculum consisted <strong>of</strong> small pieces <strong>of</strong> autoclaved F.<br />

excelsior phloem (approximately 10 x 4 x 2-3 mm) that had been placed for 15 to<br />

30 days on the various C. fraxinea cultures on MEA.<br />

106


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

107


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

108


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Inoculation <strong>of</strong> seedling stems was done by cutting an approximately 2 cm-long<br />

slit into the bark, to the level <strong>of</strong> the cambium, pulling the bark slightly away,<br />

inserting inoculum and wrapping parafilm around the wound to seal the bark flap<br />

back to the stem, to minimize contamination and to prevent rapid drying <strong>of</strong> the<br />

phloem and wood. Each isolate was inoculated onto 20 seedlings and twenty other<br />

seedlings <strong>of</strong> each ash species were inoculated with sterile pieces <strong>of</strong> ash phloem to<br />

serve as control. In each experiment, seedlings were observed for external<br />

symptoms during a period <strong>of</strong> approximately three months. Thereafter, they were<br />

dissected and lengths <strong>of</strong> necrotic lesions and longitudinal extension <strong>of</strong> wood<br />

discoloration were recorded.<br />

On both ash species all isolates tested caused symptoms virtually identical to<br />

those occurring on naturally infected trees: wilting <strong>of</strong> leaves and dieback (Fig. 3A)<br />

due to girdling <strong>of</strong> the phloem and sapwood occlusion around the inoculation site,<br />

necrotic lesions in bark (Fig. 3B), phloem (Fig. 3C) and cambium as well as<br />

brown-greyish discoloration in the wood (Fig. 3D) (Kirisits et al. 2008a; <strong>2009</strong>). On<br />

the control seedlings the inoculation wounds were partly or entirely closed and no<br />

necrotic lesions or wood discoloration occurred. The three isolates that were tested<br />

on both ash species caused more intensive symptoms (longer necrotic lesions, more<br />

plants displaying wilt and dieback) on F. angustifolia than on F. excelsior, which<br />

may indicate that the former species is more susceptible to C. fraxinea than the<br />

latter species. Chalara fraxinea was consistently re-isolated from the inoculated<br />

seedlings, while it was not recovered from any <strong>of</strong> the control plants. On<br />

F. excelsior the re-isolation rates <strong>of</strong> the various C. fraxinea isolates ranged from 25<br />

to 73 %. On F. excelsior and F. angustifolia Kochs postulates were thus fulfilled<br />

for C. fraxinea, clearly suggesting that this fungus is the primary causal agent <strong>of</strong><br />

ash dieback. This is in agreement with studies in Poland (Kowalski, 2006;<br />

Kowalski and Holdenrieder, 2008; <strong>2009</strong>a), Sweden (Bakys et al., <strong>2009</strong>b), Hungary<br />

(Szabó, 2008), Slovenia (Ogris et al., <strong>2009</strong>) and Norway (Talgø et al., <strong>2009</strong>).<br />

Flowering ash was also included in the inoculation experiments. Two C.<br />

fraxinea isolates, one in May 2008 and the other one in June 2008, were woundinoculated<br />

onto one-year-old seedlings <strong>of</strong> this ash species as described above. In<br />

both experiments C. fraxinea displayed pathogenicity to F. ornus. While in the first<br />

experiment some plants showed wilting <strong>of</strong> leaves and dieback, and necrotic lesions<br />

<strong>of</strong> similar size as those on F. excelsior developed, no dieback occurred in the<br />

second experiment and necrotic lesions were rather small. Flowering ash may be<br />

less susceptible to C. fraxinea than the other two European ash species which is<br />

supported by the fact that natural infections have thus-far not been observed.<br />

Further disease surveys and inoculation trials are, however, required to definitely<br />

appraise the susceptibility <strong>of</strong> F. ornus to the ash dieback pathogen.<br />

109


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 3. Symptoms on potted Fraxinus angustifolia seedlings following woundinoculation<br />

with Chalara fraxinea: (A) Wilting <strong>of</strong> leaves, (B) Superficially visible necrotic<br />

lesion in the bark, (C) Necrotic lesion in the phloem, (D) Discoloration <strong>of</strong> the wood (Bar<br />

for B, C and D [each showing the same inoculation point] = 1 cm). See Kirisits et al. (<strong>2009</strong>)<br />

for colored versions <strong>of</strong> the photographs.<br />

6. INFECTION BIOLOGY OF Chalara fraxinea AND HYPOTHETIC AL<br />

DISEASE CYCLE OF ASH DIEBACK<br />

Until recently, the infection biology <strong>of</strong> C. fraxinea was totally enigmatic. There<br />

were no published reports <strong>of</strong> the fungus sporulating on dead shoots, necrotic<br />

lesions or cankers and its mode <strong>of</strong> dispersal was unknown. The conidia <strong>of</strong><br />

C. fraxinea are sticky, accumulate in droplets on the top <strong>of</strong> phialophores and do not<br />

appear to be adapted to wind-dispersal (Kowalski, 2006; Kowalski and<br />

Holdenrieder, 2008). It was therefore speculated that the fungus is transmitted by<br />

animal vectors such as the ash bark beetle Leperesinus varius (Kowalski and<br />

Holdenrieder, 2008). No clear evidence was found, however, that vectors are<br />

involved and circumstantial evidence, for example the occurrence <strong>of</strong> the disease on<br />

trees <strong>of</strong> all age classes and the low degree or lack <strong>of</strong> association between insect<br />

infestations and ash dieback, made vector-dispersal <strong>of</strong> the fungus unlikely.<br />

110


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Figure 4. Apothecia <strong>of</strong> Hymenoscyphus albidus on leaf rachises <strong>of</strong> Fraxinus excelsior<br />

from the previous year (Neuwaldegg-Dornbach, Vienna, 14 and 16 June <strong>2009</strong>).<br />

The enigma, how the ash dieback pathogen is transmitted was solved, at least<br />

partly, by Kowalski and Holdenrieder (<strong>2009</strong>b) who discovered the teleophorph <strong>of</strong><br />

C. fraxinea and linked it to Hymenoscyphus albidus. Similar as in other<br />

ascomycetes, the ascospores <strong>of</strong> H. albidus are likely to be wind-dispersed<br />

(Kowalski and Holdenrieder, <strong>2009</strong>) and appear to play the key role in inciting<br />

infections <strong>of</strong> ash trees. Ascospore dispersal by wind would also explain the rapid<br />

spread <strong>of</strong> the ash dieback pathogen in Europe, if it had changed genetically or were<br />

an invasive alien organism (Kowalski and Holdenrieder, <strong>2009</strong>b). In contrast, we<br />

suppose that the spores <strong>of</strong> C. fraxinea are unable to cause infections and they may<br />

play a different, if any role in the biology <strong>of</strong> H. albidus. In May 2008 we<br />

inoculated ash shoots and leaves with suspensions <strong>of</strong> C. fraxinea spores, but no<br />

symptoms developed on any <strong>of</strong> the test seedlings. In addition, we repeatedly aimed<br />

to test the germination <strong>of</strong> spores <strong>of</strong> C. faxinea, but without any success. They did<br />

not germinate on MEA, V8 agar or an agar medium containing an extract from ash<br />

leaves that stimulated mycelial growth <strong>of</strong> C. fraxinea, but not conidial germination.<br />

The spores also did not germinate after in vitro inoculation <strong>of</strong> detached ash leaflets.<br />

We thus speculate that the spores <strong>of</strong> C. fraxinea are not conidia, but probably<br />

111


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

spermatia that play a role in exchanging nuclei and in fertilization <strong>of</strong> the fungus, if<br />

they are <strong>of</strong> any biological significance.<br />

Independent from the discovery <strong>of</strong> H. albidus by Kowalski and Holdenrieder<br />

(<strong>2009</strong>b), from September 2008 onwards we started to comprehend the importance<br />

<strong>of</strong> leaf rachises for the infection biology and epidemiology <strong>of</strong> the ash dieback<br />

pathogen. Fungal isolation from necrotic lesions on leaf rachises in September and<br />

October 2008 has shown that C. fraxinea is clearly associated with these leaf<br />

symptoms, as it was the most frequently isolated fungus and was <strong>of</strong>ten obtained in<br />

pure culture. Repeated isolations from shed leaf rachises collected from the forest<br />

floor in autumn, winter and spring confirmed that C. fraxinea persists and<br />

overwinters in these parts <strong>of</strong> ash trees. Given that isolation <strong>of</strong> C. fraxinea from<br />

dead shoots and necrotic lesions can be difficult (e. g. Bakys et al., <strong>2009</strong>a), it was<br />

surprising to isolate it as the most frequent fungus from decaying leaf remnants<br />

collected on the ground. Occasionally, also phialophores <strong>of</strong> C. fraxinea were<br />

found, sometimes abundantly. Furthermore, from late November 2008 onwards,<br />

most leaf rachises were covered by black, pseudosclerotial plates, resembling the<br />

structures <strong>of</strong>ten occurring in cultures <strong>of</strong> C. fraxinea, and now, with hindsight,<br />

known to be associated with H. albidus (Kowalski, 2006; Kowalski and<br />

Holdenrieder, 2008; <strong>2009</strong>b; Halmschlager and Kirisits, 2008; Kirisits et al., 2008a).<br />

While the discovery <strong>of</strong> H. albidus as teleomorph <strong>of</strong> C. fraxinea (Kowalski and<br />

Holdenrieder, <strong>2009</strong>b) came surprising for us, we were not so surprised that the<br />

apothecia <strong>of</strong> this fungus are predominantly formed on leaf rachises from the<br />

previous year. In spring <strong>2009</strong> we repeatedly inspected leaf rachises for the<br />

occurrence <strong>of</strong> H. albidus. At one site in Vienna, developing apothecia with unripe<br />

ascospores were first seen at the end <strong>of</strong> May, while in mid-June the first fully<br />

developed fruiting bodies with ripe, germinating ascospores occurred abundantly<br />

(Fig. 4). Since then, apothecia <strong>of</strong> H. albidus have been recorded at several sites in<br />

various parts <strong>of</strong> Austria, indicating that they occur widespread and in high<br />

numbers. Intriguingly, apothecia were observed much earlier in the year (June)<br />

than previously reported in the literature (Kowalsi and Holdenrieder, <strong>2009</strong>a and<br />

references therein).<br />

Based on the discovery <strong>of</strong> H. albidus as the teleomorph <strong>of</strong> C. fraxinea,<br />

published information on the disease as well as own observations and studies, we<br />

propose a hypothetical disease cycle for ash dieback (Fig. 5). This scheme (Fig. 5)<br />

shall also emphasize knowledge gaps and form the conceptual basis for future<br />

investigations on the infection biology and epidemiology <strong>of</strong> H. albidus/C. fraxinea.<br />

Infection <strong>of</strong> ash trees is thought to occur by wind-dispersed ascospores <strong>of</strong> H. albius<br />

which form mainly on leaf rachises from the previous year, lying on the forest floor<br />

(Kowalski and Holdenrieder, <strong>2009</strong>b; Fig. 5). Occasionally they also occur on dead<br />

shoots (Kowalski and Holdenrieder, <strong>2009</strong>b). Ascospores are likely released from<br />

June to early October (Fig. 5). The length <strong>of</strong> the infectious period will depend on<br />

the local climate and will likely vary from year to year according to the weather<br />

conditions.<br />

112


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

How infection by ascospores exactly takes place is unknown, but careful<br />

observations <strong>of</strong> symptoms suggest that leaves are an important target for infections<br />

(Fig. 5). In 2008 we observed leaf symptoms occasionally already in June and July,<br />

but they were most conspicuous from August onwards. In our opinion they can<br />

lead to early leaf shedding in late August and in September, as it has been<br />

repeatedly observed in Austria since 2005 (Cech, 2005; Hagen, 2005;<br />

Fachabteilung Forstwesen-Forstdirektion, <strong>2009</strong>). We suppose that H. albidus is<br />

able to grow from the leaves into the shoots <strong>of</strong> ash trees (Fig. 5), where it causes<br />

necrotic lesions and wood discoloration. When examining numerous young ash<br />

trees showing early stages <strong>of</strong> disease, necrotic lesions occurred either around leaf<br />

scars (Fig. 2B) or around dead side twigs (Fig. 2C), but lesions were never seen in<br />

other positions. Dead side twigs are for sure entrance points <strong>of</strong> the pathogen into<br />

main shoots, bigger branches and stems <strong>of</strong> ash trees and the location <strong>of</strong> lesions<br />

around leaf scars may support the suspicion that the pathogen can enter the phloem<br />

and xylem via leaves. Direct infections <strong>of</strong> shoots possibly also occur (Fig. 5).<br />

Whatever organ is concerned, we assume that wounding is not required for<br />

infection. Environmental factors, particularly high amounts <strong>of</strong> precipitation and<br />

high levels <strong>of</strong> air humidity are likely conducive for ascospore release and for<br />

infections to be successful.<br />

In 2008 small, localized necrotic lesions on shoots, <strong>of</strong>ten located around leaf scars<br />

(Fig. 5), where leaves had already been shed, were first observed in early August, but<br />

more commonly in September and October. These symptoms must have originated<br />

from current-years infections. Observations from 2007 to <strong>2009</strong> suggest, however, that<br />

many infections may remain latent for a while and that most <strong>of</strong> the host colonization<br />

and symptom progression takes place outside the vegetation period (Fig. 5). This view<br />

is supported by wound-inoculation <strong>of</strong> F. angustifolia with C. fraxinea in early<br />

December, resulting in dieback <strong>of</strong> many test seedlings in early spring. Because<br />

symptom development occurs to a large extent in autumn and winter, high damage<br />

levels become obvious in spring, when shoots do not flush and die, wilting <strong>of</strong> leaves<br />

occurs and trees show extensive dieback (Fig. 5). On bark tissues killed a while ago,<br />

saprotrophic or endophytic fungi sporulate (Fig. 5) and outcompete C. fraxinea,<br />

making isolation <strong>of</strong> the primary pathogen difficult or impossible. With the occurrence<br />

<strong>of</strong> H. albidus apothecia on leaf rachises on the forest floor (Fig. 5), the disease cycle<br />

starts again.<br />

7. RECOMMENDATIONS FOR DISEASE MANAGEMENT<br />

Ash dieback is another example <strong>of</strong> a serious disease damaging a valuable<br />

hardwood tree species in Europe, thereby causing problems for <strong>forestry</strong>, nature<br />

conservation and shade tree management. Common ash is not amongst the main<br />

timber species in Austria, but on many sites and forest types it is an economically<br />

and ecologically important species. On some sites there are hardly any or no<br />

attractive alternatives to ash for the selection <strong>of</strong> tree species. If ash dieback in<br />

Austria develops similar as in other countries, so that also old trees are seriously<br />

113


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

affected, considerable losses will occur and ash may lose much <strong>of</strong> its importance in<br />

silviculture.<br />

Figure 5. Hypothetical disease cycle <strong>of</strong> ash dieback caused by Hymenoscyphus albidus/<br />

Chalara fraxinea. See text for explanations<br />

Although many aspects are still unknown, much progress has recently been<br />

made to better understand ash dieback and, based on this knowledge, to<br />

recommend measures for disease management. However, this new phenomenon<br />

reminds us, how little can in most cases be done against emerging forest health<br />

problems. As a consequence <strong>of</strong> ash dieback the silvicultural characteristics <strong>of</strong> ash<br />

need to be re-appraised. While it used to be a ‘stable’ tree species that was little<br />

affected by diseases, insect pests and abiotic damaging factors, it is presently<br />

threatened by this new phenomenon. It is therefore recommended to plant common<br />

ash less extensively as before and mix it with other site-adapted tree species. Plants<br />

for planting should be carefully inspected for the occurrence <strong>of</strong> symptoms by<br />

nursery managers, forest owners and foresters. Likewise, it should be avoided to<br />

bring diseased seedlings into areas, where ash dieback has thus-far not been<br />

recorded. Wherever it is possible, leaves shed in autumn and leaf rachises on the<br />

114


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

ground prior to the occurrence <strong>of</strong> apothecia <strong>of</strong> H. albidus should be removed,<br />

ploughed into the soil or covered with soil. Such sanitation measures are probably<br />

possible and economic feasible in nurseries and urban areas. However, the<br />

dispersal distances <strong>of</strong> ascospores <strong>of</strong> H. albidus are presently unknown and it will<br />

depend on these distances, whether infections can be effectively prevented by local<br />

removal <strong>of</strong> leafs and leaf rachises. In nurseries fungicide application to protect<br />

plants from infections by ascospores may be another measure, but thus-far there is<br />

no experience regarding the fungicides to be effectively used as well as the precise<br />

timing <strong>of</strong> the applications. It is likely that infections can occur over a long period <strong>of</strong><br />

time, from June to early October (Fig. 5), which would make many applications<br />

necessary. While fungicide treatments may be a useful method to raise healthy ash<br />

seedlings, their general importance for disease management will be limited, as<br />

seedlings will become infected after having been planted in the field.<br />

Thus-far there are no reliable recommendations for the silvicultural treatment <strong>of</strong><br />

stands affected by ash dieback. It is, however, recommended not to abandon ash<br />

too early, and, if there are no other reasons, to harvest only dead and severely<br />

damaged trees. It has been observed that C. fraxinea can grow from epicormic<br />

shoots into the wood <strong>of</strong> ash stems and causes discoloration there (Kowalski and<br />

Holdenrieder, 2008; Thomsen et al., <strong>2009</strong>), thereby lowering the timber quality and<br />

value. To avoid such losses as well as damage caused by wood-decay fungi, timely<br />

felling <strong>of</strong> severely diseased trees is recommended. Ash dieback will likely weaken<br />

the populations <strong>of</strong> ash trees and secondary pathogens such as Armillaria spp. and<br />

ash bark beetles may become more important and need to be considered (Kowalski<br />

and Holdenrieder, 2008; Thomsen et al., <strong>2009</strong>). When intensive care is possible,<br />

individual trees can be rescued by cutting infected shoots, twigs and branches.<br />

Likewise, young trees can be cut to rescue the stump and root system, from where<br />

suckers will develop. However, in both cases, new infections <strong>of</strong><br />

H. albidus/C. fraxinea are likely to occur and thus, trees need to be inspected and<br />

treated repeatedly.<br />

The most promising potential option for disease management may be the<br />

existence <strong>of</strong> considerable levels <strong>of</strong> resistance or tolerance within the populations <strong>of</strong><br />

common ash. In heavily affected areas it is not rare to see severely diseased ash<br />

trees growing aside <strong>of</strong> still healthy or little affected trees. Likewise, investigations<br />

in seed plantations in Denmark strongly suggest that there are considerable<br />

differences in the resistance levels <strong>of</strong> ash clones towards ash dieback, with some<br />

clones hardly being affected (Olrik et al., 2007; Kjær et al., <strong>2009</strong>). Preliminary<br />

assessments in clonal seed orchards in Austria support this view (C. Freinschlag,<br />

C. Jasser, A. Gaisbauer and T. Kirisits, unpublished data). It is therefore<br />

recommended that forest owners and foresters record, mark and promote healthy<br />

and slightly diseased ash trees growing in severely affected stands and promote<br />

natural regeneration <strong>of</strong> these potentially resistant or tolerant trees. Thereby they<br />

may facilitate that resistance levels in the ash populations are maintained, get<br />

115


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

stabilized or even increase. Breeding for resistance may be another, more intensive<br />

option for disease management in the future.<br />

7. REFERENCES<br />

Adler, W., Oswald, K., Fischer, R., 1994. Exkursionsflora von Österreich [Excursion flora <strong>of</strong><br />

Austria]. Eugen Ulmer Verlag, Stuttgart, Wien, 1180 pp.<br />

Bakys, R., Vasaitis, R., Barklund, P., Thomsen, I.M., Stenlid, J., <strong>2009</strong>a. Occurrence and pathogenicity<br />

<strong>of</strong> fungi in necrotic and non-symptomatic shoots <strong>of</strong> declining common ash (Fraxinus<br />

excelsior) in Sweden. European Journal <strong>of</strong> Forest Research 128, 51-60.<br />

Bakys, R., Vasaitis, R., Barklund, P., Ihrmark, K., Stenlid, J, <strong>2009</strong>b. Investigations concerning the<br />

role <strong>of</strong> Chalara fraxinea in declining Fraxinus excelsior. Plant Pathology 58, 284-292.<br />

Cech, T.L., 2005. Blattkrankheiten und vorzeitiger Laubfall – eine Folge des kühlfeuchten Sommers<br />

2005 [Fungal diseases <strong>of</strong> leaves – a consequence <strong>of</strong> the cool and wet summer 2005].<br />

Forstschutz Aktuell 34, 11-12. http://bfw.ac.at/400/pdf/fsaktuell_34.pdf<br />

Cech, T.L., 2006a. Auffallende Schadfaktoren an Waldbäumen im Jahr 2005 [Striking damaging<br />

agents on forest trees in 2005]. Forstschutz Aktuell 35, 6-7.<br />

http://bfw.ac.at/400/pdf/fsaktuell_35.pdf<br />

Cech, T.L., 2006b. Eschenschäden in Österreich [Ash dieback and premature leaf shedding in<br />

Austria]. Forstschutz Aktuell 37, 18-20. http://bfw.ac.at/400/pdf/fsaktuell_37_8.pdf<br />

Cech, T.L., 2008. Eschenkrankheit in Niederösterreich – neue Untersuchungsergebnisse [Dieback <strong>of</strong><br />

ash in Lower Austria – new results]. Forstschutz Aktuell 43, 24-28.<br />

http://bfw.ac.at/400/pdf/fsaktuell_43_9.pdf<br />

Cech, T.L., Hoyer-Tomiczek, U., 2007. Aktuelle Situation des Zurücksterbens der Esche in<br />

Österreich [Actual situation <strong>of</strong> dieback <strong>of</strong> ash in Austria]. Forstschutz Aktuell 40, 8-10.<br />

http://bfw.ac.at/400/pdf/fsaktuell_40_3.pdf<br />

Chandelier, A., André, F., Laurent, F., <strong>2009</strong>. Detection <strong>of</strong> Chalara fraxinea in common ash (Fraxinus<br />

excelsior) using real time PCR. Forest Pathology, in press. DOI: 10.1111/j.1439-<br />

0329.<strong>2009</strong>.00610.x<br />

Damm, H.J., 1997. Seltene Eschen II – Fraxinus angustifolia im Auwald Hohenau [Rare ash species<br />

II – Fraxinus angustifolia in the floodplain forest Hohenau]. In: WWF Austria (Ed.),<br />

Proceedings <strong>of</strong> the symposium ‘Zukunft für gefährdete Baumarten? Rückbringung und<br />

Förderung seltener und gefährdeter Baum- und Straucharten’ [‘The Future for Endangered<br />

Tree Species. Preservation and Promotion <strong>of</strong> Rare and Endangered Tree and Shrub<br />

Species’], 1 October 1997, Vienna, Austria, pp. 11-17.<br />

Engesser, R., Queloz, V., Meier, F., Kowalski, T., Holdenrieder, O., <strong>2009</strong>: Das Triebsterben der<br />

Esche in der Schweiz [Ash dieback in Switzerland]. Wald und Holz 6/<strong>2009</strong>, 24-27.<br />

Essl, F., Rabitsch, W., 2002. Neobiota in Österreich [Neobiota in Austria]. Federal Environment<br />

Agency - Austria, 432 pp.<br />

Essl, F., Rabitsch, W., Obermayr, G. (general co-ordination), Freud, M. (translation), 2006: Austrian<br />

Action Plan on Invasive Alien Species. Austrian Federal Ministry <strong>of</strong> Agriculture,<br />

Forestry, Environment and Water Management (‘Lebensministerium’), 26 pp.<br />

http://www.umweltbundesamt.at/fileadmin/site/umweltthemen/naturschutz/Neobiota_Eng<br />

l.pdf<br />

116


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Fachabteilung Forstwesen-Forstdirektion, <strong>2009</strong>. Fortschutzbericht Steiermark 2008 [Report on Forest<br />

Health in Styria in 2008]. 22 pp.<br />

Hagen, R., 2005. NÖ Forstschutzsituation 2005. [Report on the Forest Health Situation in Lower<br />

Austria in 2008]. 3 pp.<br />

Hagen R, 2007. NÖ Forstschutzsituation 2006/2007 [Report on the Forest Health Situation in Lower<br />

Austria in 2006/2007]. 5 pp.<br />

Halmschlager, E., Kirisits, T., 2008. First report <strong>of</strong> the ash dieback pathogen Chalara fraxinea on<br />

Fraxinus excelsior in Austria. Plant Pathology 57, 1177 and New Disease Reports<br />

[http://www.bspp.org.uk/ndr/july2008/2008-25.asp], Volume 17 .<br />

Ioos, R., Kowalski, T., Husson, C., Holdenrieder, O., <strong>2009</strong>. Rapid in planta detection <strong>of</strong> Chalara<br />

fraxinea by a real-time PCR assay using a dual-labelled probe. European Journal <strong>of</strong> Plant<br />

Pathology, in press. DOI: 10.1007/s10658-009-9471-x.<br />

Jankovský, L., Palovčíková, Dvořák, M., 2008. Alien diseases <strong>of</strong> woody plants in the Czech<br />

Republic. In: Steyrer G, Tomiczek C, Lackner C (Eds.), Proceedings <strong>of</strong> the Second<br />

Meeting <strong>of</strong> Forest Protection and Forest Phytosanitary Specialists, 27-28 November 2007,<br />

Vienna, Austria. Forstschutz Aktuell 44, 32-34. .<br />

http://bfw.ac.at/400/pdf/fsaktuell_44_14.pdf<br />

Jelem, H., 1974. Die Auwälder der Donau in Österreich [The floodplain forests along the Danube in<br />

Austria]. Mitteilungen der Forstlichen Bundesversuchsanstalt Wien 109, 1-287.<br />

Jelem, H., 1975. Marchauen in Niederösterreich [Floodplain forests along the March in Lower<br />

Austria]. Mitteilungen der Forstlichen Bundesversuchsanstalt Wien 113, 1-93.<br />

Kirisits, T., 2008. Eschenpathogen Chalara fraxinea nun auch in Kärnten nachgewiesen [First report<br />

<strong>of</strong> the ash dieback pathogen Chalara fraxinea from Carinthia]. Forstschutz Aktuell 45,<br />

28-30. http://bfw.ac.at/400/pdf/fsaktuell_45_12.pdf<br />

Kirisits, T., Halmschlager, E., 2008. Eschenpilz nachgewiesen [Ash fungus detected]. Forstzeitung<br />

119 (02-2008), 32-33. .<br />

https://www.dafne.at/prod/dafne_plus_common/attachment_download/efd9e0d1ee75f887<br />

d883b3c6e597ba03/Kirisits%20and%20Halmschlager_Chalara%20fraxinea_FZ_119_02-<br />

2008.pdf<br />

Kirisits, T., Cech, T.L., <strong>2009</strong>. Eschentriebsterben: Ursache, Symptome, Empfehlungen [Ash dieback:<br />

cause, symptoms, recommendations]. Kärntner Forstverein Information 59/Juni <strong>2009</strong>, 17-<br />

19. http://www.verwaltung.ktn.gv.at/161547_DE-<br />

Kirisits, T., Matlakova, M., Mottinger-Kroupa, S., Halmschlager, E., 2008a. Verursacht Chalara<br />

fraxinea das Zurücksterben der Esche in Österreich? [Is Chalara fraxinea the causal agent<br />

<strong>of</strong> ash dieback in Austria?]. Forstschutz Aktuell 43, 29-34.<br />

http://bfw.ac.at/400/pdf/fsaktuell_43_10.pdf<br />

Kirisits, T., Matlakova, M., Mottinger-Kroupa, S., Halmschlager, E., 2008b. Involvement <strong>of</strong> Chalara<br />

fraxinea in ash dieback in Austria. In: Steyrer G, Tomiczek C, Lackner C (Eds.),<br />

Proceedings <strong>of</strong> the Second Meeting <strong>of</strong> Forest Protection and Forest Phytosanitary<br />

Specialists, 27-28 November 2007, Vienna, Austria. Forstschutz Aktuell 44, 16-18.<br />

http://bfw.ac.at/400/pdf/fsaktuell_44_7.pdf<br />

Kirisits, T., Matlakova, M., Mottinger-Kroupa, S., Halmschlager, E., Lakatos, F., <strong>2009</strong>. Chalara<br />

fraxinea associated with dieback <strong>of</strong> narrow-leafed ash (Fraxinus angustifolia). Plant<br />

Pathology 58, in press and New Disease Reports .<br />

(http://www.bspp.org.uk/publications/new-disease-reports/ndr.php?id=019043), Volume<br />

19.<br />

117


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Kjær, E.D., Nielsen, L.R., Thomsen, I.M., Skovsgaard, J.P., <strong>2009</strong>. Forskning om asketoptørre<br />

[Research on ash dieback]. Skoven 02/<strong>2009</strong>, 92-94.<br />

Kowalski, T., 2006. Chalara fraxinea sp. nov. associated with dieback <strong>of</strong> ash (Fraxinus excelsior) in<br />

Poland. Forest Pathology 36, 254-270.<br />

Kowalski, T., Holdenrieder, O., 2008. Eine neue Pilzkrankheit an Esche in Europa [A new fungal<br />

disease <strong>of</strong> ash in Europe]. Schweizerische Zeitschrift für Forstwesen 159, 45-50.<br />

Kowalski, T., Holdenrieder, O., <strong>2009</strong>a. Pathogenicity <strong>of</strong> Chalara fraxinea. Forest Pathology 39, 1-7.<br />

Kowalski, T., Holdenrieder, O., <strong>2009</strong>b. The teleomorph <strong>of</strong> Chalara fraxinea, the causal agent <strong>of</strong> ash<br />

dieback. Forest Pathology, in press. Doi: 10.111/j.1439-0329.2008.00589.x.<br />

Olrik, D.C., Kjær, E.D., Ditlevsen, B., 2007. Klonforskelle i angreb af asketoptørre [Clones differ in<br />

the infestation by ash dieback]. Skoven 11/2007, 522-525.<br />

Mayer, H., 1974. Wälder des Ostalpenraumes [Forests <strong>of</strong> the Eastern Alps]. Gustav Fischer Verlag,<br />

Stuttgart, 344 pp.<br />

Mayer, H., 1984. Waldbau auf soziologisch-ökologischer Grundlage. [Silviculture on Plant<br />

Sociological and Ecological Basis]. Gustav Fischer Verlag, Stuttgart, New York, 514 pp.<br />

Ogris, N., Hauptmann, T., Jurc, D., <strong>2009</strong>. Chalara fraxinea causing common ash dieback newly<br />

reported in Slovenia. Plant Pathology 58, in press and New Disease Reports<br />

(http://www.bspp.org.uk/publications/new-disease-reports/ndr.php?id=019015 ), Volume<br />

19.<br />

Lygis, V., Vasiliauskas, R., Larsson, K.H., Stenlid, J., 2005. Wood-inhabiting fungi in stems <strong>of</strong><br />

Fraxinus excelsior in declining ash stands <strong>of</strong> northern Lithuania, with particular reference<br />

to Armillaria cepistipes. Scandinavian Journal <strong>of</strong> Forest Research 20, 337-346.<br />

Nationalpark Kalkalpen, 2007. Schutzgüter im Nationalpark O.ö. Kalkalpen. Auswahl<br />

schützenswerter Tiere, Pflanzen und Lebensräume [Objects worth to be protected in the<br />

national park Upper Austrian Northern Limestone Alps. Selection <strong>of</strong> animals, plants and<br />

habitats worth to be protected], Schriftenreihe des Nationalparks Kalkalpen, Band 6, 127<br />

pp.<br />

Przybył, K., 2002. Fungi associated with necrotic apical parts <strong>of</strong> Fraxinus excelsior shoots. Forest<br />

Pathology 32, 387-392.<br />

Pukacki, P.M., Przybył, K., 2005. Frost injury as a possible inciting factor in bud and shoot necroses<br />

<strong>of</strong> Fraxinus excelsior L. Journal <strong>of</strong> Phytopathology 153, 512-516.<br />

Schadauer, K., 1994. Baumartenatlas für Österreich. Die Verbreitung der Baumarten nach Daten der<br />

Österreichischen Waldinventur. [Atlas <strong>of</strong> tree species in Austria. The distribution <strong>of</strong> tree<br />

species based on data <strong>of</strong> the Austrian Forest Inventory]. FBVA-Berichte 76, 1-157.<br />

Schumacher, J., Wulf, A., Leonhard, S., 2007. Erster Nachweis von Chalara fraxinea T. Kowalski sp.<br />

nov. in Deutschland – ein Verursacher neuartiger Schäden an Eschen [First record <strong>of</strong><br />

Chalara fraxinea T. Kowalski sp. nov. in Germany – a new agent <strong>of</strong> ash decline].<br />

Nachrichtenblatt des Deutschen Pflanzenschutzdienstes 59, 121-123.<br />

Skovsgaard, J.P., Thomsen, I.M., Skovgaard, I.M., Martinussen, T., <strong>2009</strong>. Associations among<br />

symptoms <strong>of</strong> dieback in even-aged stands <strong>of</strong> ash (Fraxinus excelsior L.). Forest<br />

Pathology, in press. DOI: 10.1111/j.1439-0329.<strong>2009</strong>.00599.x<br />

Steyrer, G., Krenmayer, W., Schaffer, H., 2008. Dokumentation der Waldschädigungsfaktoren<br />

(DWF) 2007 [Documentation <strong>of</strong> forest damage factors 2007]. Forstschutz Aktuell 42, 17-<br />

82. http://bfw.ac.at/400/pdf/fsaktuell_42_5.pdf<br />

118


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Szabó, I., 2008. First report <strong>of</strong> Chalara fraxinea affecting common ash in Hungary. Plant Pathology<br />

58, in press and New Disease Reports (http://www.bspp.org.uk/publications/new-diseasereports/ndr.php?id=018030),<br />

Volume 18.<br />

Talgø, V., Sletten, A., Brurberg, M.B., Solheim, H., Stensvand, A., <strong>2009</strong>. Chalara fraxinea isolated<br />

from diseased ash in Norway. Plant Disease 93, 548.<br />

Thomsen, I.M., Skovsgaard, J.P., Barklund, P., Vasaitis, R., 2007. Svampesygdom er årsag til<br />

toptørre i ask [A fungal disease is the cause <strong>of</strong> ash dieback]. Skoven 05/2007, 234-236.<br />

Thomsen, I.M., Skovsgaard, J.P., Kjær, E.D., Nielsen, L.R., <strong>2009</strong>. Status for asketopptørre i Danmark<br />

og Europa [The status <strong>of</strong> ash dieback in Denmark and Europe]. Skoven 02/<strong>2009</strong>, 87-91.<br />

Wolf, W., Jasser, C., 2003. Laubholz – Der richtige Weg zum Erfolg [Hardwood trees – the right way<br />

to success]. Forest Authority <strong>of</strong> the province Upper Austria, 3rd <strong>edition</strong>, 30 pp.<br />

Zukrigl, K., 1997. Seltene Eschen I [Rare ash species I]. In: WWF Austria (Ed.), Proceedings <strong>of</strong> the<br />

symposium ‘Zukunft für gefährdete Baumarten? Rückbringung und Förderung seltener<br />

und gefährdeter Baum- und Straucharten’ [‘The Future for Endangered Tree Species.<br />

Preservation and Promotion <strong>of</strong> Rare and Endangered Tree and Shrub Species’], 1 October<br />

1997, Vienna, Austria, pp. 7-10.<br />

Acknowledgements: The financial support by the Austrian Federal Ministry <strong>of</strong><br />

Agriculture, Forestry, Environment and Water Management (BMLFUW research<br />

project no. 100343, BMLFUW-LE.3.2.3/0001-IV/2/2008), the provincial<br />

governments <strong>of</strong> Lower Austria, Carinthia, Salzburg, Burgenland, Upper Austria<br />

and Styria as well as the Austrian Federal Forests (ÖBf AG) is gratefully<br />

acknowledged. We thank the Forest Authorities <strong>of</strong> all Austrian provinces as well as<br />

numerous District Forest Authorities and forest owners for the support <strong>of</strong> our<br />

research on ash dieback. The information on the occurrence <strong>of</strong> ash dieback in<br />

various European countries was collected as part <strong>of</strong> the Coordination Action<br />

‘European Network on emerging diseases and threats through invasive alien<br />

species in forest ecosystems (FORTHEATS)’, contract no. 044436, within the 6 th<br />

Framework Programme <strong>of</strong> the European Union.<br />

119


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 120-128<br />

DIEBACK ON Fraxinus ornus IN KONYA REGION<br />

Asko Lehtijärvi 1* , H.Tuğba Doğmuş-Lehtijärvi 1 , Mertcan Karadeniz 1 ,<br />

Mustafa Uygun 1<br />

ABSTRACT<br />

1 Süleyman Demirel University, Faculty <strong>of</strong> Forestry, 32260 Isparta, Turkey<br />

*asko@orman.<strong>sdu</strong>.edu.tr<br />

In many European countries, intensive dieback <strong>of</strong> ash has been observed in all age<br />

classes, independent <strong>of</strong> forest type and geographic position during the last 10 years. There<br />

are 3 species <strong>of</strong> Fraxinus in Turkey: Fraxinus excelsior L., F. angustifolia Vahl.. and F.<br />

ornus L.. They are fast growing trees and have valuable wood which is widely used in<br />

furniture industry. In addition, the trees are used in landscape architecture.<br />

In this study, existence and causal agents <strong>of</strong> dieback was investigated in F. ornus<br />

plantations located in Dutlukır (7.2 ha) and Altınapa Dam (6.4 ha) in Konya. Condition <strong>of</strong><br />

the shoots, characteristics <strong>of</strong> the cankers and lesions, height and diameter (at root collar) <strong>of</strong><br />

the trees, and signs <strong>of</strong> insect attacks were recorded. Almost all, 98.2%, <strong>of</strong> the sampled trees<br />

were bearing cankers, 4.1% had signs <strong>of</strong> insect attacks, and 24.7% <strong>of</strong> the trees had dry<br />

shoots.<br />

Keywords: Dieback, Fraxinus excelsior, Fraxinus angustifolia, Fraxinus ornus<br />

INTRODUCTION<br />

Dieback <strong>of</strong> ash has been one <strong>of</strong> the most important diseases on ash in European<br />

countries during the last 10 years. Recently the casual agent <strong>of</strong> the dieback was<br />

reported to be Chalara fraxinea T. Kowalski (Kowalski, 2006; Cech and Hoyer-<br />

Tomiczek, 2007; Kirisits et al., 2008; Halmschlager and Kirisits, 2008). Symptoms<br />

are necrosis <strong>of</strong> leaf rachises and leaflet veins, shoot, twig and branch dieback as<br />

well as necrotic lesions and cankers in the bark. Bark necrosis is <strong>of</strong>ten<br />

accompanied with brownish discolouration <strong>of</strong> the wood. Wilting <strong>of</strong> leaves can<br />

sometimes be seen on recently girdled shoots and twigs (Bakys et al., 2008;<br />

Kowalski, 2006; Kowalski and Holdenrieder, 2008; Halmschlager and Kirisits,<br />

2008; Krisits et al., 2008).<br />

There are three species and seven subspecies <strong>of</strong> Fraxinus distributed in<br />

Marmara, Black sea, Aegian and Mediterranean Region <strong>of</strong> Turkey (Yaltırık,<br />

1978). Fraxinus ornus (Manna ash) is a native species to southern Europe and<br />

south western Asia. Manna ash is an important tree species widely used in<br />

furniture industry and music instrument manufacturing. In addition, it is used in<br />

landscape architecture.<br />

120


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The aim <strong>of</strong> the present study was to investigate the frequency <strong>of</strong> dieback in<br />

Manna ash plantations located in Dutlukır and near Altınapa Dam in Konya<br />

province.<br />

MATERIALS AND METHODS<br />

Manna ash plantations located in Dutlukır and near Atınapa Dam in Konya<br />

province were investigated in April <strong>2009</strong>. The origin <strong>of</strong> the trees in the study areas<br />

is not known.<br />

The plantation site near Altınapa Dam, approximately 16.5 km west <strong>of</strong> the city<br />

<strong>of</strong> Konya, covers 6.4 ha on a minor slope facing the dam at an altitude <strong>of</strong> 1280 m<br />

a.s.l. The 7.2–hectare plantation in Dutlukır is located on flat terrain, 1075 m a.s.l.,<br />

approximately 9 km southwest <strong>of</strong> Konya. In Konya region, winters are cold and<br />

snowy and summers are hot and dry. In the study area, annual precipitation ranges<br />

from 300 to 400 mm (Fig. 1). The average number <strong>of</strong> days with precipitation per<br />

month is 9-11 from December to May, and 2-7 from June to November. Between<br />

July and September there are only 2-3 rainy days per month. During these three<br />

summer months the total precipitation is below 30 mm. The average annual<br />

temperature is +11.4˚C (Anonymous, <strong>2009</strong>).<br />

Average precipitation<br />

(mm)<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1 2 3 4 5 6 7 8 9 10 11 12<br />

121<br />

Months<br />

Average precipitation (mm) Average Temperature (°C)<br />

Figure 1: Monthly means for precipitation and temperature in Konya 1975-2007.<br />

The plantations were investigated using a systematic sampling strategy.<br />

Condition <strong>of</strong> the shoots, characteristics <strong>of</strong> the cankers and lesions, tree height and<br />

stem diameter at root collar, and signs <strong>of</strong> insect attacks were recorded for every 4 th<br />

tree in Dutlukır and for every 6 th tree in Altınapa plantation. Shoot and bark<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

Average temperature (°C)


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

samples were collected from trees bearing cankers and lesions; totally 340 samples<br />

were collected for fungal isolation. The samples were kept in paper bags at +4˚C<br />

until isolation.<br />

The isolations were made from cankers and lesions on the shoots and the stem<br />

(Fig. 2).<br />

Figure 2: Cankers and lesions on the stem and the shoot <strong>of</strong> F. ornus.<br />

After surface sterilization with 70% ethanol and removing the surface bark,<br />

pieces <strong>of</strong> shoots were removed and placed in petri dishes containing 2% malt<br />

extract agar. The cultures were incubated at room temperature in dark conditions.<br />

RESULTS & DISCUSSION<br />

The average diameter <strong>of</strong> the sampled trees was 3.2 cm in Dutlukır and 2.6 cm in<br />

Altınapa and the average height <strong>of</strong> the sampled trees was 235 cm in Dutlukır and<br />

233 cm in Altınapa.<br />

Almost all, 98.2% <strong>of</strong> the sampled trees were bearing cankers, 4.1% had signs <strong>of</strong><br />

insect attacks and 24.7% <strong>of</strong> the shoots had dried. In addition, wood discolouration<br />

and necrosis on the shoots were observed. The study is still going on and<br />

identification <strong>of</strong> the fungi is under progress.<br />

The high proportion <strong>of</strong> the trees bearing cankers and dry shoots indicates that<br />

the trees are either growing at an unsuitable site or that the trees are frequently<br />

attacked by a biotic agent. As the origin <strong>of</strong> the trees is not known, the possibility<br />

that they are poorly adapted to the local climate in the study area can not be<br />

excluded. The fact that the plantation sites are located outside <strong>of</strong> the natural<br />

distribution <strong>of</strong> F. ornus in Turkey (cf. Yaltırık, 1978) supports a hypothesis <strong>of</strong><br />

weather related damage.<br />

122


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Periods <strong>of</strong> low rainfall can be one <strong>of</strong> the abiotic factors correlated with the<br />

initiation <strong>of</strong> the disease (Hibben and Silverborg, 1978). However, there has not<br />

been any unusually dry periods, that could explain the high frequency <strong>of</strong> dead<br />

shoots, in Konya during the last 3 years. On the other hand, insufficient adaptation<br />

to the local climate could result in e.g. delayed winter hardening <strong>of</strong> the current–<br />

year shoots increasing susceptibility to frost damage or winter drought. The shoots<br />

had dried after bud formation indicating that the shoots had died between late<br />

summer and early spring. The frequency <strong>of</strong> dry shoots was similar in both sites.<br />

Therefore differences in microclimate between the sites, caused by the dam in<br />

Altınapa, seemed not to have had any effect on the occurrence <strong>of</strong> the damage.<br />

REFERENCES<br />

Anonymous, <strong>2009</strong>. Turkish State Meteorological Service (www.dmi.gov.tr).<br />

Bakys, R., Vasaitis, R., Barklund, P., Ihrmark, K. and Stenlid, J., 2008. Investigations concerning the<br />

role <strong>of</strong> Chalara fraxinea in declining Fraxinus excelsior. Plant Pathology (2008).<br />

Cech, T. L. and Hoyer-Tomiczek, U., 2007. Aktuelle Situation des Zurücksterbens der Esche in<br />

Österreich. Forstschutz Aktuell 40, 8-10.<br />

Halmschlager, E. and Kirisits, T., 2008. First report <strong>of</strong> the ash dieback pathogen Chalara fraxinea on<br />

Fraxinus excelsior in Austria. Plant Pathology (2008) 57, 1177.<br />

Hibben, C. R. and Silverborg, S. B., 1978. Severity and causes <strong>of</strong> ash dieback. Journal <strong>of</strong><br />

Arboriculture 4(12), 274-279.<br />

Kirisits, T., Matlakova, M. Mottinger-Kroupa, S., Harmschlager, E., 2008. Verursacht Chalara<br />

fraxinea das Zürücksterben der Esche in Österreich. Forstschuzt Aktuell 43-2008.<br />

Kowalski, T., 2006.Chalara fraxinea sp. nov. associated with dieback <strong>of</strong> ash (Fraxinus excelsior) in<br />

Poland. Forest Pathology, 36 (2006) 264–270, Blackwell Verlag, Berlin.<br />

Kowalski, T. and Holdenrieder, O., 2008. Pathogenicity <strong>of</strong> Chalara fraxinea. Forest Pathology 2008,<br />

Blackwell Verlag, Berlin.<br />

Yaltırık, F., 1978. Türkiye’deki Doğal Oleaceae Taksonlarının Sistematik Revizyonu. İstanbul<br />

Üniversitesi <strong>Orman</strong> <strong>Fakültesi</strong> Yayınları, I. U. Yayın No.: 2404, O. F. Yayın No.: 250.<br />

123


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 124-128<br />

ASH DIEBACK IN THE CZECH REPUBLIC<br />

Petr ŠŤASTNÝ 1 , Dagmar PALOVČÍKOVÁ 1 , Libor JANKOVSKÝ 1*<br />

1 Mendel University <strong>of</strong> Agriculture and Forestry, Faculty <strong>of</strong> Forestry and Wood Technology,<br />

Department <strong>of</strong> Forest Protection and Wildlife Management, Zemědělská 3, 613 00 Brno, Czech<br />

Republic<br />

ABSTRACT<br />

*jankov@mendelu.cz<br />

Ash dieback was observed in Baltic states since middle <strong>of</strong> 90`s, however the progress<br />

<strong>of</strong> disease was observed within last years. The new fungus Chalara fraxinea was described<br />

as a causal agent <strong>of</strong> ash dieback in 2006. The ash dieback was observed in some local<br />

areas in the Czech republic since middle <strong>of</strong> 90`s and it was connected mostly with extreme<br />

climatic conditions. Progress <strong>of</strong> disease was observed since 2003. Ash decline was<br />

observed practically in all regions <strong>of</strong> the CR, C. fraxinea were confirmed in the CR in<br />

September 2007. Occurrence <strong>of</strong> Chalara fraxinea was confirmed in Fraxinus excelsior<br />

and F. angustifolia. Perfect stage <strong>of</strong> C. fraxinea Hymenscyphus albidus is known as<br />

common saprophytes on leaf-stalks. In areas with esh dieback was observed aphid<br />

Prociphilus bumeliae, however the role <strong>of</strong> this insect is actually discus.<br />

Key words: ash decline, Chalara fraxinea, ash, ash dieback<br />

1. INTRODUCTION<br />

Common ash (Fraxinus excelsior) is threatened in large parts <strong>of</strong> Europe by ash<br />

dieback (eg. Cech, 2006; Lygis et al., 2005). Chalara fraxinea has been recently<br />

determined to be the causal agent <strong>of</strong> this disease (Kowalski, 2006; Kowalski &<br />

Holdenrieder, <strong>2009</strong>a; Bakys et al., <strong>2009</strong>).<br />

Chalara fraxinea was reported in Poland (Kowalski, 2006), Denmark (Thomsen<br />

et al., 2007), Germany (Schumacher et al., 2007), Austria (Halmschlager and<br />

Kirisits, 2008), Hungary (Szabo, 2008a; 2008b), Finland (EPPO, 2008a), Lithuania<br />

(R. Vasaitis, personal communication), Norway (EPPO, 2008b; H. Solheim,<br />

personal communication), Sweden (Bakys et al., <strong>2009</strong>), Switzerland (Engesser and<br />

Holdenrieder, unpublished) and France (Ioos, personal communication). Symptoms<br />

<strong>of</strong> ash dieback were also reported from Slovakia in 2008 (Kunca, personal<br />

communication); the disease was noted also in Slovenia (Ogris et al., <strong>2009</strong>) and<br />

Croatia (own observation) in 2008. From the Czech republic is pathogen reported<br />

from 2007 (Jankovský and Holdenrieder, <strong>2009</strong>). The ascomycete Hymenoscyphus<br />

albidus (Roberge ex Desm.) W. Phillips was identified as the teleomorph <strong>of</strong> C.<br />

fraxinea by Kowalski and Holdenrieder (<strong>2009</strong>b).<br />

124


2. MATERIAL AND METHODS<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Ash twigs (4-8 mm diameter) with dieback (comprising a necrotic distal portion and<br />

a proximal living portion) were collected and transferred to the laboratory. The surface<br />

were disinfected by spraying with concentrated ethanol and superficially dried on a<br />

clean bench. The tissue samples (diameter 2-5 mm, about 2-3 mm long), were<br />

dissected from the sapwood below necrotic lesions, after bark removal. Samples were<br />

then surface sterilized by immersion in sodium hypochlorite (7 - 10 %) for 60 - 90 sec,<br />

then immersed in 96% ethanol for 60 - 90 sec, washed in sterilized water and placed on<br />

the medium. Tissue samples were aseptically transferred on malt extract agar (MEA 3;<br />

30 g/L, peptone 5 g/L, agar 15 g/L) and, according to Kowalski (2006), streptomycin<br />

(100 mg/L) added after autoclaving. Identification was made on the bases <strong>of</strong> colony<br />

morphology and microscopic features.<br />

3. RESULTS AND DISCUSSION<br />

Ash dieback associated with bark necroses and withering <strong>of</strong> young shoots was<br />

recorded in several areas in the CR during 2004 – 2008. The locations affected<br />

include mostly all area <strong>of</strong> the CR, eg. Beskydy Mts., Jeseniky Mts., Giant Mts.,<br />

Bayerischer Wald Mts., Central Bohemia, Prague, Eastern Bohemia, Czech<br />

Moravian Highland, the area at the junction <strong>of</strong> the Thaya and Morava Rivers, and<br />

along the Czech, Austrian and Slovak borders. Ash dieback has extended across the<br />

entire country since 2004. The symptoms were also noted in nurseries, e<strong>special</strong>ly<br />

on saplings in South Moravia. However wooly aphids from genus Prociphilus spp.<br />

were observed in this areas abundantly as a important harmful factor in early<br />

spring.<br />

The first record <strong>of</strong> Chalara fraxinea in the CR originated from samples<br />

collected at Drahany Highland, Arboretum Krtiny, from Fraxinus excelsior and in<br />

some other locations (Jankovský and Holdenrieder <strong>2009</strong>). Ash diaback were<br />

observed on Fraxinus excelsior and its cultivars, e<strong>special</strong>ly sensible is F. excelsior<br />

cv. „Pendula“, and F. angustifolia.<br />

On the area <strong>of</strong> South Moravia samples has been taken e<strong>special</strong>ly from the<br />

district <strong>of</strong> Židlochovice forest enterprise, close to Austrian and Slovak borders.<br />

Health condition <strong>of</strong> young forest stands is serious on some locations and there is a<br />

danger <strong>of</strong> large damages <strong>of</strong> these young plantations. The culture <strong>of</strong> C. fraxinea<br />

origin from flood-plain forest by the village Tvrdonice in this area. It is located<br />

near <strong>of</strong> the boarder with Slovakia. The other samples were collected in the Ivaň<br />

area (adjacent to the Nove mlyny reservoir) and Soutok game preserve (National<br />

nature reserve Cahnov).<br />

The occurrence <strong>of</strong> ash decline is continually monitored, e<strong>special</strong>ly in the area <strong>of</strong><br />

south Moravia.<br />

According to the experience with symptoms <strong>of</strong> ash wilting we can assume that<br />

C. fraxinea is spread through the whole area <strong>of</strong> the CR.<br />

125


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Table 1. Occurrence <strong>of</strong> Ash dieback on some monitored plots.<br />

LOCALITY DATE OF COLLECT. CONCLUSION NOTICE POSITION<br />

1 Arboretum <strong>of</strong> Křtiny 26.09.2007<br />

2 Ochoz u Brna 29.09.2008 Pozitive<br />

126<br />

Pozitive Fraxinus excelsior<br />

"Pendula"<br />

3 Hradčany u Brna 06.10.2008 Pozitive<br />

Symptom <strong>of</strong> disease<br />

4 Brno Lesná - Soběšice 08.10.2008<br />

5 Lomnice u Tišnova (Sýkoř hill) 12.10.2008 Pozitive Shoots along the way<br />

Symptom <strong>of</strong> disease<br />

Regeneration under the<br />

6 LZ Židlochovice - NPR Cahnov 13.10.2008<br />

Symptom <strong>of</strong> disease<br />

cover <strong>of</strong> older stand<br />

7 District Lednice-Valtice area - Rendez vous 30.10.2008<br />

8 LZ Židlochovice - Ivaň 04.02.<strong>2009</strong><br />

9 Kroměříž - zámeček 07.02.<strong>2009</strong><br />

Symptom <strong>of</strong> disease<br />

Symptom <strong>of</strong> disease<br />

10 LZ Židlochovice – forest district Tvrdonice 18.02.<strong>2009</strong> Pozitive Compartment 909<br />

11 LZ Židlochovice – forest district Tvrdonice 18.02.<strong>2009</strong> Symptom <strong>of</strong> disease Compartment 904<br />

12<br />

Forest nursery Hadovna - region<br />

Kroměříž 30.05.<strong>2009</strong> Symptom <strong>of</strong> disease<br />

13 Arboeko co. - Smržice 30.04.<strong>2009</strong> Symptom <strong>of</strong> disease<br />

49°19'7"N/<br />

16°44'35"E<br />

49°15'24"N,<br />

16°43'56"E<br />

49°19'36"N,<br />

16°25'58"E<br />

49°14'37.5"N,<br />

16°37'5.728"E<br />

49°24'34.51"N,<br />

16°24'46.765"E<br />

48°39'17.223"N,<br />

16°56'32.378"E<br />

48°44'54.48"N,<br />

16°47'39.299"E<br />

48°55'18.304"N,<br />

16°34'56.446"E<br />

49°16'55.536"N,<br />

17°27'50.746"E<br />

48°47'29.587"N,<br />

17°4'21.828"E<br />

48°47'23.761"N,<br />

17°4'21.588"E<br />

49°18'24.285"N,<br />

17°39'4.909"E<br />

49°23'56.145"N,<br />

17°11'26.528"E<br />

Dry lesions show circular shape at first, they gradually change into elliptical<br />

oblong, depressed necrosis, which gradually extends on the surface; it was<br />

observed in most places. Below the bark, the dieback <strong>of</strong> the cambium is evident.<br />

The necrosis extends both into transpiration flow and assimilation flow. The<br />

infection gets from necrosis also into the wood part, which is discolored by grey<br />

brown. One year shoots die above the necrosis e<strong>special</strong>ly in autumn months (from<br />

end <strong>of</strong> August till October). Necrosis on 2 years old and perennial shoots can be<br />

occluded and there can be superficial cankers with callus on the edge created. Dark<br />

brown necrosis is formed on petioles and trees shed premature, from end <strong>of</strong> August<br />

till October, their still green leaves. One year old, rarely also older shoots die on<br />

older trees. Typical <strong>of</strong> this is a creation <strong>of</strong> compressed crown and disturbance <strong>of</strong><br />

parallel trunk.<br />

Hymenoscyphus albidus is noted by Kowalski and Holdenrieder (<strong>2009</strong>) as the<br />

teleomorph <strong>of</strong> Chalara fraxinea. The species is widespread in Europe. According<br />

to the literature, H. albidus occurs exclusively on ash petioles from ash litter<br />

(Breitenbach and Kranzlin, 1984; Ellis and Ellis, 1985). In the CR, it is well known<br />

as a saprophyte on leaf-stalks in the litter, it is considered to be a common species<br />

in myc<strong>of</strong>loristic research.


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Transfer <strong>of</strong> infection nor vector has not been clarified till now. One can assume<br />

that infection occurs via ascospores <strong>of</strong> H. albidus which are formed on leaf fall <strong>of</strong><br />

ash during summer months. It corresponds to time period <strong>of</strong> development <strong>of</strong> newly<br />

attacking infection. Rapid and e<strong>special</strong>ly sudden dissemination <strong>of</strong> this disease<br />

throughout Europe is not typical for tree disease, is it more typical for non biotic<br />

diseases or for insect outbreaks. E<strong>special</strong>ly an area, which is being occupied during<br />

relatively short period <strong>of</strong> time, is enormous. For pathogens, it is difficult to<br />

penetrate structural suberin barrier which is represented by unharmed bark. The<br />

influence <strong>of</strong> sucking insect can be discussed. Remarkable is the fact that lesions<br />

occur practically at the same place like the colony <strong>of</strong> woolly aphids Prociphilus<br />

spp. Aphid`s colonies are located on shoots and petioles <strong>of</strong> leaves. In nurseries,<br />

where chemical control against sucking insect was applied, symptoms <strong>of</strong> ash<br />

dieback disappeared. Woolly aphids are not reported to occur in all areas where ash<br />

decline was observed. Sucking insect can play a role <strong>of</strong> agent which creates a<br />

defect in the bark. In sucking locations, tissues suffer from necrosis and fungi like<br />

C. fraxinea penetrate through the tissues. There is an apparent separation <strong>of</strong> reason<br />

– sucking <strong>of</strong> aphids during spring time and formation <strong>of</strong> necrosis during summer<br />

time when the connection to sucking insect needs not to be evident.<br />

4. CONCLUSION<br />

The bionomy <strong>of</strong> the fungus, infection cycle, pathology, resp. pathogenicity <strong>of</strong><br />

the fungus and genetic structure are main topics <strong>of</strong> contemporary research.<br />

Although pathogenicity <strong>of</strong> the C. fraxinea has been proved, the mechanism <strong>of</strong><br />

penetration <strong>of</strong> fungus into host is not known yet. The role <strong>of</strong> sucking insect, mainly<br />

woolly aphids, should be crucial due to production <strong>of</strong> lesion around sucking points<br />

at the position <strong>of</strong> previous aphid colonies. The relationship <strong>of</strong> sucking insects and<br />

C.fraxinea in ash dieback pathology requires further investigations.<br />

5. ACKNOWLEDGEMENTS<br />

This project was supported by MSM 6215648902 and Grant foundation <strong>of</strong><br />

FFWT MUAF.<br />

6. REFERENCES<br />

Bakys, R., Vasaitis, R., Barklund, P., Ihrmark, K., Stenlid, J., <strong>2009</strong>. Investigations concerning the role<br />

<strong>of</strong> Chalara fraxinea in declining Fraxinus excelsior. Plant Pathology 58, 284 - 292.<br />

Cech, T., 2006. Eschenschäden in Österreich. Forstschutz Aktuell (Wien), 37, 18 - 20. (in German)<br />

Cech, T., Hoyer-Tomiczek, U., 2007. Aktuelle Situation des Zurücksterbens der Esche in Österreich.<br />

Forstschutz Aktuell (Wien) 40, 8 - 10.<br />

EPPO, 2007. Ash dieback in Europe and possible implication <strong>of</strong> Chalara fraxinea: addition to the<br />

EPPO Alert List. EPPO Reporting service 2007/179.<br />

EPPO, 2008a. First report <strong>of</strong> Chalara fraxinea in Norway. EPPO Reporting service 2008/181.<br />

127


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Halmschlager, E., Kirisits, T., 2008. First report <strong>of</strong> the ash dieback pathogen Chalara fraxinea on<br />

Fraxinus excelsior in Austria. Plant Pathology 57, 1177.<br />

Jankovský, L., Holdenrieder O., <strong>2009</strong>. Chalara fraxinea - ash dieback in the Czech Republic. Plant<br />

protection Science. (in press).<br />

Kirisits, T., Matlakova, M., Mottinger-Kroupa, S., Halmschlager, E., 2008. Verursacht Chalara<br />

fraxinea das Zurucksterben der Eshe in Osterreich. Forstschutz Aktuell (Wien) 43. 29 –<br />

34.<br />

Kowalski, T., 2006. Chalara fraxinea sp. nov. associated with dieback <strong>of</strong> ash (Fraxinus excelsior) in<br />

Poland. Forest Pathology, 36(4). 264 - 270.<br />

Kowalski, T., Holdenrieder, O., <strong>2009</strong>a. Pathogenicity <strong>of</strong> Chalara fraxinea. Forest Pathology 38. 1 - 7.<br />

Kowalski, T., Holdenrieder, O., <strong>2009</strong>. The teleomorph <strong>of</strong> Chalara fraxinea, the causal agent <strong>of</strong> ash<br />

dieback. Forest Pathology online publication doi: 10.1111/j.1439-0329.2008.00589.x.<br />

Lygis, V., Vasiliauskas, R., Larsson, K.H. & Stenlid J., 2005. Wood-inhabiting fungi in stems <strong>of</strong><br />

Fraxinus excelsior in declining ash stands <strong>of</strong> northern Lithuania, with particular reference<br />

to Armillaria cepistipes. Scandinavian Journal <strong>of</strong> Forest Research 20. 337-346.<br />

Ogris, N., Hauptman T., and Jurc D., <strong>2009</strong>. Chalara fraxinea causing common ash dieback newly<br />

reported in Slovenia. New Disease Reports 19, Feb <strong>2009</strong> to Aug <strong>2009</strong>.<br />

http://www.bspp.org.uk/publications/new-disease-reports/ndr.php?id=019015.<br />

Schumacher, J., Wulf A., Leonhard, S., 2007. Erster Nachweis von Chalara fraxinea T. Kowalski sp.<br />

nov. Deutschland - ein Verursacher neuartiger Schäden an Eschen. Nachrichtenblatt des<br />

Deutschen Pflanzenchutzdienstes (Braunschweig), 59(6). 121 - 123.<br />

Szabó, I., 2008b. First report <strong>of</strong> Chalara fraxinea affecting common ash in Hungary. New Disease<br />

Reports, 18. Aug 2008 to Jan <strong>2009</strong>. http://www.bspp.org.uk/publications/new-diseasereports/ndr.php?id=018030.<br />

Thomsen, I.M., Skovsgaard, J.P., Barklund, P., Vasaitis, R., 2007. Svampesygdom er arsag til<br />

toptorre i ask. Skoven (Skov & Landskab; Kobenhavns Universitet; Copenhagen, DK), 5:<br />

234 -236.<br />

128


Canker Diseases<br />

129


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

130


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 131-135<br />

HORSE CHESTNUT BLEEDING CANKER – BAGGING THE BUG !<br />

Sarah GREEN 1* , Bridget LAUE 1 , Grace MACASKILL 1 , Heather STEELE 1<br />

1 Forest Research, Northern Research Station, Roslin, Midlothian, Scotland EH25 9SY<br />

ABSTRACT<br />

*sarah.green@<strong>forestry</strong>.gsi.gov.uk<br />

Bleeding canker <strong>of</strong> horse chestnut, caused by the bacterial pathogen, Pseudomonas<br />

syringae pv. aesculi, has rapidly established and become widespread throughout parts <strong>of</strong><br />

northern Europe, including Great Britain, over the last five years. Despite the fact that this<br />

disease is having a devastating effect on an important amenity tree species, very little is<br />

known about the infection processes <strong>of</strong> this pathogen, or the genetic and physiological<br />

factors which cause it to be so highly damaging. One difficulty associated with studying<br />

this pathogen is the lengthy procedure required to confirm its presence on the host. Realtime<br />

PCR primers were developed based on gyrase B gene sequence data for the specific<br />

detection <strong>of</strong> P. syringae pv. aesculi in infected horse chestnut. This quantitative real-time<br />

PCR assay provides the facility to study several important aspects <strong>of</strong> the biology <strong>of</strong> P.<br />

syringae pv. aesculi on horse chestnut including its potential for epiphytic survival on<br />

healthy trees, as well as its dissemination in different environmental substrates, such as soil,<br />

water and infected tree debris. As part <strong>of</strong> ongoing work, the complete genome <strong>of</strong> P.<br />

syringae pv. aesculi is currently being sequenced to determine its complement <strong>of</strong> virulence<br />

and fitness genes. Molecular tools are also being used to determine the origin <strong>of</strong> P. syringae<br />

pv. aesculi, the location/s and time <strong>of</strong> its introduction and geographical routes <strong>of</strong> spread<br />

within Great Britain and Europe.<br />

Keywords: Pseudomonas syringae pv. aesculi, bleeding canker, Aesculus<br />

hippocastanum, real-time PCR.<br />

1. INTRODUCTION<br />

In the last 10-15 years, there has been an unprecedented increase in the numbers<br />

<strong>of</strong> hitherto unrecognised diseases attacking trees throughout the world. Many <strong>of</strong> the<br />

causal organisms have been inadvertently introduced into new ecosystems through<br />

the increase in global commerce, via pathways such as trade in live plants<br />

(including soils), poorly treated timber products and international trade in bonsai<br />

(Brasier, 2008). European horse chestnut (Aesculus hippocastanum), an important<br />

amenity tree species throughout much <strong>of</strong> Great Britain and northern Europe, is<br />

suffering from two major problems <strong>of</strong> recent introduction into the continent:<br />

bacterial bleeding canker and the horse chestnut leaf miner. Native to northern<br />

Greece and Albania (Phillips, 1978), horse chestnut was introduced into the UK in<br />

the late 16 th Century and was planted widely in parks and gardens in both urban<br />

and rural areas, <strong>of</strong>ten in avenues bordering roads. Horse chestnut is highly regarded<br />

for its qualities as a shade tree, for its showy white flowers in spring and the<br />

production <strong>of</strong> its fruits or ‘conkers’.<br />

131


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Since 2003, an epidemic <strong>of</strong> a bleeding canker disease <strong>of</strong> horse chestnut has<br />

become widespread across Great Britain as well as a number <strong>of</strong> other European<br />

countries, including the Netherlands, Belgium, France and Germany. The disease<br />

has been attracting a great deal <strong>of</strong> media attention in Great Britain over the last<br />

couple <strong>of</strong> years due to the severity <strong>of</strong> damage on affected trees. Symptoms <strong>of</strong> the<br />

disease include bleeding cankers located on the stem and branches, with rustcoloured<br />

liquid oozing from cracks in the bark, necrotic phloem, foliar<br />

discolouration, and crown dieback <strong>of</strong>ten leading to tree death. A Great Britain-wide<br />

survey <strong>of</strong> 2629 horse chestnut trees conducted in 2007 found that over 70% <strong>of</strong> trees<br />

surveyed in parts <strong>of</strong> England exhibited symptoms typical <strong>of</strong> bleeding canker, with<br />

36% and 42% <strong>of</strong> surveyed trees showing these symptoms in Wales and Scotland,<br />

respectively (Forestry Commission, 2008). The reason that bleeding canker disease<br />

<strong>of</strong> horse chestnut is currently <strong>of</strong> such high pr<strong>of</strong>ile is due to its dramatic impact on<br />

an important amenity tree species. The disease is now a key tree health issue in<br />

Great Britain, particularly in the context <strong>of</strong> ‘urban greening’ and with climate<br />

change in mind there is increasing recognition <strong>of</strong> the need to maintain healthy<br />

populations <strong>of</strong> shade trees within urban areas.<br />

The causal agent responsible for this new epidemic <strong>of</strong> bleeding canker disease<br />

<strong>of</strong> horse chestnut has only very recently been identified as a gram-negative<br />

fluorescent Pseudomonas syringae bacterium, which has an identical partial gyrase<br />

B gene sequence to a strain isolated from leaf lesions on Indian horse chestnut<br />

(Aesculus indica) in India in 1980 (Schmidt et al., 2008; Webber et al., 2008).<br />

There are at least 50 closely related pathovars <strong>of</strong> the species Pseudomonas syringae<br />

which can be distinguished by host range, and which infect a wide range <strong>of</strong><br />

herbaceous and woody plants. It is thought that P. syringae pv. aesculi may have<br />

originated from India, being native on Indian horse chestnut, and has been recently<br />

introduced to Great Britain and Europe, possibly via imported, infected nursery<br />

stock (Brasier, 2008). If it is a new introduction to Europe, P. syringae pv. aesculi<br />

has found a new host, European horse chestnut, on which it is considerably more<br />

aggressive than its native host on which it only causes leaf lesions. Britain’s forests<br />

are under continuous risk from new, exotic diseases as a result <strong>of</strong> the increased<br />

movement <strong>of</strong> plant stock between countries over ever greater distances. Bacterial<br />

diseases <strong>of</strong> trees present a particular risk to Britain’s forests and woodlands since<br />

their activity in northern Europe may be favoured by the milder, wetter winters<br />

predicted under future climate change scenarios and because so little is currently<br />

known about the routes <strong>of</strong> invasion, spread and survival <strong>of</strong> these pathogens on<br />

woody hosts.<br />

One <strong>of</strong> the obstacles to studying P. syringae pv. aesculi is the difficulty <strong>of</strong><br />

confirming its presence in host tissues. To date, this bacterium is detected on<br />

symptomatic trees by isolation and culturing, PCR amplification and sequencing <strong>of</strong><br />

the gyrase B gene using universal primers, and comparing sequence homology with<br />

a type strain <strong>of</strong> P. syringae pv. aesculi published in the US National Center for<br />

Biotechnology Information (NCBI) GenBank sequence database. However, this is<br />

132


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

time consuming, through the need to obtain pure bacterial cultures from diseased<br />

horse chestnut tissues which may be colonised by a range <strong>of</strong> bacterial genera<br />

(Green et al., <strong>2009</strong>). Recently, real-time PCR has proven to be a very useful tool<br />

for the detection <strong>of</strong> plant pathogenic fungi and bacteria, being highly sensitive,<br />

specific and rapid, with the added capacity for quantification <strong>of</strong> the pathogen in<br />

host tissues (Schaad and Frederick, 2002; Vandroemme et al., 2008). The first aim<br />

<strong>of</strong> this study was to develop and test a robust, reliable, quantitative real-time PCR<br />

assay which is specific to P. syringae pv. aesculi, and to demonstrate its use for<br />

detecting the bacterium in inoculated and naturally infected horse chestnut trees<br />

(Green et al., <strong>2009</strong>). Also, briefly discussed are ongoing projects aimed at i)<br />

characterising the key genetic and physiological factors determining virulence and<br />

epiphytic fitness <strong>of</strong> P. syringae pv. aesculi on European horse chestnut and ii)<br />

tracing the epidemiological origins <strong>of</strong> this devastating bacterium.<br />

2. MATERIALS AND METHODS<br />

For the development <strong>of</strong> the real-time PCR assay, a total <strong>of</strong> 65 bacterial isolates<br />

were collected from various parts <strong>of</strong> diseased or healthy horse chestnut in Britain,<br />

the DNA extracted and the partial gyrase B gene region amplified using universal,<br />

degenerate primers. The amplified fragments were sequenced and aligned with<br />

other bacterial gyrase B gene sequences available in GenBank to design real-time<br />

PCR primers specific to P. syringae pv. aesculi. The specificity and sensitivity <strong>of</strong><br />

the real-time PCR primers was tested using nine strains <strong>of</strong> P. syringae pv. aesculi,<br />

17 other strains <strong>of</strong> P. syringae, 11 other non-pathogenic Pseudomonas spp. and 14<br />

other species <strong>of</strong> bacteria isolated from horse chestnut trees. The ability <strong>of</strong> the realtime<br />

primers to amplify and quantify DNA <strong>of</strong> P. syringae pv. aesculi in diseased<br />

horse chestnut tissues was also demonstrated.<br />

3. RESULTS<br />

The real-time primer pair and reaction conditions developed to detect P.<br />

syringae pv. aesculi amplified all isolates <strong>of</strong> P. syringae pv. aesculi, but did not<br />

amplify the DNA extracted from the included reference bacteria or horse chestnut<br />

when 1 ng <strong>of</strong> DNA was used as the template. The real-time primers reliably<br />

amplified P. syringae pv. aesculi down to 1 pg <strong>of</strong> extracted DNA, with and without<br />

the presence <strong>of</strong> host DNA, and also amplified unextracted DNA in whole cells <strong>of</strong><br />

the bacterium down to at least 160 colony forming units. The real-time PCR assay<br />

detected and quantified DNA <strong>of</strong> P. syringae pv. aesculi in sixteen out <strong>of</strong> seventeen<br />

tissue samples collected from naturally infected or artificially inoculated horse<br />

chestnut, with generally good consistency between the two PCR runs in terms <strong>of</strong> Ct<br />

values and quantity <strong>of</strong> pathogen DNA detected.<br />

133


4. DISCUSSION<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

A novel, quantitative real-time PCR assay has now been developed to detect the<br />

pathogenic bacterium, P. syringae pv. aesculi, which is currently causing a severe<br />

bleeding canker disease <strong>of</strong> horse chestnut trees in several European countries<br />

(Green et al., <strong>2009</strong>). The real-time PCR primers were shown to be both highly<br />

specific, giving exponential amplification only with the target pathogen, and highly<br />

sensitive, allowing detection <strong>of</strong> P. syringae pv. aesculi down to 1 pg DNA in<br />

diseased horse chestnut tissues with the optimised reaction conditions used (Green<br />

et al., <strong>2009</strong>). This advance in methodology now provides a tool for the accurate and<br />

sensitive quantitative detection <strong>of</strong> P. syringae pv. aesculi in host tissues, as well as<br />

in different environmental substrates. This assay will be used to examine the ability<br />

<strong>of</strong> P. syringae pv. aesculi to survive epiphytically in host material, and its potential<br />

to contaminate rainwater and soil. The aim <strong>of</strong> this is to determine the routes for<br />

transmission <strong>of</strong> the pathogen, and to explain the reason for its high mobility.<br />

Research into P. syringae plant pathogens is currently <strong>of</strong> high pr<strong>of</strong>ile<br />

internationally due to their detrimental impact in the horticultural, agricultural and<br />

<strong>forestry</strong> sectors (Kennelly et al., 2007; Perez-Martinez et al., 2008). Ongoing<br />

research into the dentification <strong>of</strong> the key virulence and fitness genes <strong>of</strong> P. syringae<br />

pv. aesculi through genome sequencing, combined with in vivo studies on<br />

pathogenicity, will provide novel insights into the factors determining the<br />

pathogenicity and fitness <strong>of</strong> an important and newly damaging P. syringae<br />

pathovar on a woody host; this being an area for which information is currently<br />

scarce.<br />

The effectiveness <strong>of</strong> import and quarantine regulations and disease management<br />

strategies designed to protect the agricultural, horticultural and forest industries<br />

against exotic pathogens relies on a thorough understanding, based on sound<br />

scientific data, <strong>of</strong> the routes <strong>of</strong> introduction and spread <strong>of</strong> exotic pathogens in new<br />

locations. This is reflected in the current high priority given to these types <strong>of</strong><br />

studies within the European Union (EU), particularly given the recent, rapid<br />

expansion <strong>of</strong> the international plant trade and an appreciation <strong>of</strong> the threat that this<br />

presents. Current research on P. syringae pv. aesculi will identify the most suitable<br />

genetic markers, based on genome sequence data, for studying the recent<br />

evolutionary history <strong>of</strong> P. syringae pv. aesculi. These markers will then be used for<br />

phylogenetic analyses to elucidate whether this pathogen is a new introduction to<br />

Europe, and if so, when and from where it was introduced and routes <strong>of</strong> subsequent<br />

spread within regions. The research briefly described here will be essential for the<br />

development <strong>of</strong> more effective phytosanitary measures and disease management<br />

strategies to guard against future threats posed by exotic bacterial pathogens. The<br />

project will ultimately inform and guide management and mitigation strategies for<br />

an important bacterial disease <strong>of</strong> an amenity tree species through increased<br />

understanding <strong>of</strong> the causal agent.<br />

134


5. REFERENCES<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Brasier, C.M., 2008. The biosecurity threat to the UK and global environment from international trade<br />

in plants. Plant Pathology 57, 792-808.<br />

Forestry Commission, 2008. Report on the national survey to assess the presence <strong>of</strong> bleeding canker<br />

<strong>of</strong> horse chestnut trees in Great Britain. Forestry Commission, Edinburgh, UK.<br />

Green, S., Laue, B., Fossdal, C.G., A’Hara, S., Cottrell, J., <strong>2009</strong>. Infection <strong>of</strong> horse chestnut (Aesculus<br />

hippocastanum) by Pseudomonas syringae pv. aesculi and its detection by quantitative<br />

real-time PCR. Plant Pathology, In Press.<br />

Kennelly, M.M., Cazorla, F.M., de Vicente, A., Ramos, C., Sundin, G.W., 2007. Pseudomonas<br />

syringae diseases <strong>of</strong> fruit rees; Progress towards understanding and control. Plant Disease<br />

91, 4-16.<br />

Perez-Martinez, I., Zhao, Y., Murillo, J., Sundin, G.W., Ramos, C., 2008. Global genomic analysis <strong>of</strong><br />

Pseudomonas savastanoi plasmids. Journal <strong>of</strong> Bacteriology 190, 625-635.<br />

Philips, R., 1978. Trees in Britain, Europe and North America. Macmillan Reference.<br />

Schaad, N.W., Frederick, R.D., 2002. Real-time PCR and its application for rapid plant disease<br />

diagnostics. Canadian Journal <strong>of</strong> Plant Pathology 24, 250-8.<br />

Schmidt, O., Dujesiefken, D., Stobbe, H., Moreth, U., Kehr, R., Schröder, T., 2008. Pseudomonas<br />

syringae pv. aesculi associated with horse chestnut bleeding canker in Germany. Forest<br />

Pathology 38, 124-8<br />

Vandroemme, J., Baeyen, S., Van Vaerenbergh, J., De Vos, P., Maes, M., 2008. Senstitive real-time<br />

PCR detection <strong>of</strong> Xanthomonas fragariae in strawberry plants. Plant Pathology 57, 438-<br />

44<br />

Webber, J.F., Parkinson, N.M., Rose, J., Stanford, H., Cook, R.T.A., Elphinstone, J.G., 2008.<br />

Isolation and identification <strong>of</strong> Pseudomonas syringae pv. aesculi causing bleeding canker<br />

<strong>of</strong> horse chestnut in the UK. Plant Pathology 57, 368.<br />

135


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 136-140<br />

AN OVERVIEW OF POTENTIAL INFECTION COURTS FOR Neonectria<br />

fuckeliana, THE CAUSAL AGENT OF NECTRIA FLUTE CANKER IN<br />

Pinus radiata IN NEW ZEALAND<br />

ABSTRACT<br />

Anna J.M HOPKINS 1* , Patricia E. CRANE 1 , Margaret A. DICK 1<br />

1 Forest Protection, Scion, Private Bag 3020, Rotorua 3010, New Zealand<br />

* Anna.Hopkins@scionresearch.com<br />

Nectria flute canker is a disease <strong>of</strong> Pinus radiata stems in the southern parts <strong>of</strong> New<br />

Zealand caused by the pathogen Neonectria fuckeliana. Although tree crowns generally<br />

remain healthy, stem cankers and associated defect reduce the commercial value <strong>of</strong> the<br />

timber. In Northern Hemisphere countries where N. fuckeliana is endemic, open wounds,<br />

dead attached branches and branch stubs have been identified as the primary infection<br />

courts for N. fuckeliana. In New Zealand the development <strong>of</strong> the Nectria flute canker<br />

disease is primarily associated with pruned branch stubs however recent studies suggest<br />

that this is not the only possible infection court as the fungus has been found in trees prior<br />

to pruning. Three separate field trials were established to examine potential infection courts<br />

for N. fuckeliana in P. radiata in New Zealand. These infection courts included stem<br />

wounds, pruned stubs, branch crotches and branch collars. Stem depressions, the usual<br />

precursor to flute cankers, were created following inoculation <strong>of</strong> deep and shallow stem<br />

wounds and <strong>of</strong> some branch collars. Inoculation directly into pruned stubs resulted in only a<br />

few, small stem depressions and the fungus was largely contained within the branch trace.<br />

Infection through branch crotches was not successful. Both inoculation types tested<br />

(ascospores and conidia) resulted in similar canker development and fungal spread within<br />

the tree. The trials described in this paper are ongoing.<br />

Keywords: Neonectria fuckeliana, stem cankers, Pinus radiata.<br />

1. INTRODUCTION<br />

Pinus radiata D. Don is the most important plantation tree species grown in New<br />

Zealand, comprising more than 89% <strong>of</strong> the plantation estate (NZFOA, <strong>2009</strong>). Many <strong>of</strong><br />

the plantations are managed to produce clear, knot-free wood by pruning from one to<br />

three times during the rotation (NZFOA, <strong>2009</strong>). Stem cankers, <strong>of</strong>ten associated with<br />

pruned stubs, have become increasingly noticeable in some Pinus radiata plantations<br />

in the lower South Island <strong>of</strong> New Zealand over the last 15 years (Dick and Crane,<br />

<strong>2009</strong>; Gadgil et al., 2003). The long, narrow cankers, commonly referred to as “flute<br />

cankers” for their elongated appearance, can extend for several metres above and for a<br />

shorter distance below a pruned branch stub. Formation <strong>of</strong> cankers associated with<br />

natural injuries on the stem internodes has rarely been observed (Dick and Crane,<br />

<strong>2009</strong>). Although tree crowns generally remain healthy, affected trees are susceptible to<br />

decay, to wind breakage at infected whorls, and wood quality can be affected.<br />

136


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Neonectria fuckeliana (C. Booth) Castl. & Rossman (Nectria fuckeliana C.<br />

Booth) (Ascomycota: Nectriaceae) is the fungus most commonly found in<br />

association with the flute cankers (Dick and Crane, <strong>2009</strong>). This pathogen is thought<br />

to be endemic to Northern Europe, Scandinavia and North America where it has<br />

been recorded principally as a common wound invader or weak pathogen <strong>of</strong><br />

species <strong>of</strong> Picea and Abies (e.g. Roll-Hansen and Roll-Hansen, 1979; Schultz and<br />

Parmeter, 1990; Vasiliauskas and Stenlid, 1998). Pathogenicity <strong>of</strong> the fungus has<br />

been reported infrequently in Pinus spp., and this has been primarily as the result <strong>of</strong><br />

artifical inoculations (Smerlis, 1969). The pathogen has three spore states. In<br />

addition its teleomorph, in which ascospores are produced in perithecia (the<br />

Neonectria phase), two anamorphs are formed under certain conditions: an<br />

Acremonium state with unicellular spores and a Cylindrocarpon state<br />

(Cylindrocarpon cylindroides var. tenue Wollenweber) with multicellular spores.<br />

Vasiliauskas and Stenlid (1997) demonstrated that, in Europe, the N. fuckeliana<br />

ascospores are probably the major dispersal propagules. The importance <strong>of</strong> the<br />

anamorphs in the pathogen life cycle and in disease development is not fully<br />

understood.<br />

In the Northern Hemisphere, open wounds, dead attached branches and branch<br />

stubs have been identified as the primary infection courts for N. fuckeliana (Roll-<br />

Hansen and Roll-Hansen, 1979). In New Zealand, since the development <strong>of</strong> the<br />

Nectria flute canker disease is primarily associated with pruned branch stubs, it<br />

was assumed that these branch stubs were the primary infection court (e.g. Bulman,<br />

2007). Recent studies by Power and Ramsfield (2006, 2007) however, suggest that<br />

this is not the only possible infection court for N. fuckeliana. In a study <strong>of</strong> 90<br />

pruned and 90 unpruned trees, the pathogen N. fuckeliana was found in<br />

approximately 22% <strong>of</strong> trees and no significant difference in frequency <strong>of</strong> the<br />

pathogen was found between pruned and unpruned trees. None <strong>of</strong> the trees<br />

examined showed symptoms <strong>of</strong> Nectria flute canker. This suggests that N.<br />

fuckeliana is able to enter trees prior to pruning using some other infection court/s.<br />

This paper outlines a number <strong>of</strong> trials currently being undertaken in southern New<br />

Zealand to identify possible alternative infection courts for N. fuckeliana.<br />

2. INFECTION THROUGH STEM WOUNDS<br />

A field trial was undertaken to examine the importance <strong>of</strong> different wound types<br />

and inoculum sources for disease development and fungal infection. Specifically<br />

the trial aimed to determine whether pruned branch stubs were an effective<br />

infection court for N. fuckeliana and, following on from Vasiliauskas and Stenlid<br />

(1997), whether ascospores were the most effective inoculum source. Forty-five 6year-old<br />

Pinus radiata were subjected to one <strong>of</strong> three wound types (shallow stem<br />

wound, deep stem wound or pruned branch stub) and one <strong>of</strong> three inoculation types<br />

(ascospore inoculation, conidial inoculation or a water control). Trees were<br />

assessed for the formation <strong>of</strong> stem depressions (the typical precursor to flute<br />

cankers) after 6, 12 and 18 months, after which time they were harvested and the<br />

137


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

spread <strong>of</strong> N. fuckeliana was examined within the tree using isolations. Wound type<br />

was found to be very important for flute canker development, with trees with deep<br />

wounds showing largest stem depressions and most spread <strong>of</strong> N. fuckeliana within<br />

the tree. Shallow stem wounds also resulted in some stem depressions however<br />

these were usually small and there was only a little vertical movement <strong>of</strong> the<br />

fungus through the stem. Inoculation <strong>of</strong> branch stubs resulted in few or small stem<br />

depressions and the fungus was largely contained within the branch trace. Both<br />

inoculation types (ascospores and conidia) resulted in similar levels <strong>of</strong> stem<br />

depressions and fungal spread within the tree.<br />

3. INFECTION THROUGH BRANCH CROTCHES<br />

Pinus radiata grown in the southern regions <strong>of</strong> New Zealand is frequently<br />

subjected to snow events during the winter months. Due to the acute branching<br />

angle <strong>of</strong> P. radiata this <strong>of</strong>ten results in severe branch bending and can lead to<br />

branch breakage. This severe branch bending can lead to openings in bark <strong>of</strong> the<br />

branch crotch which may provide an infection court for fungal spores. Forty eightyear-old<br />

Pinus radiata trees were used to examine this theory. On each tree, six<br />

branches <strong>of</strong> similar size were selected on the same or adjacent whorls. Each branch<br />

was then subjected to one <strong>of</strong> two bending treatments (bent with string to simulate<br />

the weight <strong>of</strong> snow on the branch, or unbent) and one <strong>of</strong> three inoculum sources<br />

(control, Acremonium conidia on a colonised twig, or a piece <strong>of</strong> bark containing N.<br />

fuckeliana perithecia with ascospores). The inoculum source was glued or stapled<br />

directly above the branch crotch. Trees were checked after 6 and 12 months for any<br />

canker development and no change to the trees was observed. After 18 months, 20<br />

<strong>of</strong> the trees were felled and dissected through the branch traces. Isolations were<br />

undertaken to determine whether N. fuckeliana was now present within the stem<br />

tissue. No N. fuckeliana was isolated. The remaining 20 trees will be monitored for<br />

a further 6-12 months and may be felled if any external symptoms <strong>of</strong> Nectria flute<br />

canker develop.<br />

4. INFECTION THROUGH BRANCH COLLARS<br />

Although inoculations directly into pruned stubs were not always successful at<br />

initiating N. fuckeliana spread throughout the stem (see section 2), the symptoms <strong>of</strong><br />

Nectria flute canker are almost exclusively associated with pruned branches.<br />

Incidence <strong>of</strong> Nectria flute canker in a stand can be much higher than 22% (the<br />

proportion <strong>of</strong> colonisation recorded in unpruned trees) and so some infection may<br />

be occurring at the time <strong>of</strong> pruning. If the branch collar was damaged during<br />

pruning, it is possible that this branch collar may act as an infection court for N.<br />

fuckeliana. To investigate this, 41 eight-year-old P. radiata trees were pruned in<br />

the lower third <strong>of</strong> the stem and three branch collars on each tree were inoculated<br />

with an Acremonium spore suspension. Great care was taken to prevent the spore<br />

suspension spreading onto the rest <strong>of</strong> the pruned branch surface. The remaining<br />

138


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

branch stubs on each whorl were treated as controls and were not inoculated. After<br />

6 months, eight <strong>of</strong> the trial trees were showing very slight canker development<br />

from inoculated sites. Following 12 months stem depressions were recorded above<br />

17 <strong>of</strong> the 123 inoculated branch collars. This trial is ongoing and isolations will be<br />

made from trees felled after 18-24 months to determine whether N. fuckeliana had<br />

infected the stem and how far it has spread.<br />

5. DISCUSSION AND FUTHER RESEARCH<br />

The inoculation experiments described in this paper have given some insight<br />

into the potential infection mechanisms for N. fuckeliana in P. radiata in New<br />

Zealand. In the first trial, infection and symptom development was clearly<br />

demonstrated using both ascospores and conidia from the Acremonium stage. This<br />

indicates that spores from both these lifestages could potentially play a role in<br />

infection <strong>of</strong> this pathogen. In New Zealand however, although the Acremonium<br />

stage is produced in culture, it is rarely observed in the field. In contrast, perithecia<br />

producing ascospores <strong>of</strong> the teleomorph are frequently observed in the field and,<br />

due to their abundance, are much more likely to play a role in dispersal and<br />

infection <strong>of</strong> the pathogen.<br />

While both deep and shallow stem wounds in the first trial resulted in successful<br />

infection <strong>of</strong> N. fuckeliana and development <strong>of</strong> some stem depressions, it is unlikely<br />

that these infection mechanisms play an important role in infection in the field.<br />

Few wounds are found on P. radiata in plantations. Thus wounding is unlikely to<br />

play a role as a primary infection court for N. fuckeliana.<br />

Although no N. fuckeliana was isolated from inside the stems in the branch<br />

crotch trial this does not rule out branch crotches as an infection court for this<br />

pathogen. During the experiment, it was very difficult to simulate the effect <strong>of</strong><br />

snow weight on branches, particularly any repetitive opening and closing <strong>of</strong> the<br />

branch crotch associated with branch movement. As a result, the experimental<br />

conditions may not have been sufficient to allow penetration <strong>of</strong> spores into the<br />

stem. Further trials <strong>of</strong> this nature are planned.<br />

6. REFERENCES<br />

Bulman, L.S., 2007. Nectria pruning trial. Forest Health News 179,1-2. Available from<br />

http://www.ensisjv.com/NewsEventsandPublications/Newsletters/ForestHealthNews/Fore<br />

stHealthNewsArchive/tabid/179/Default.aspx [accessed 6 March <strong>2009</strong>].<br />

Dick, M.A. and Crane, P.E., <strong>2009</strong>. Neonectria fuckeliana is pathogenic to Pinus radiata in New<br />

Zealand. Australasian Plant Disease Notes 4, 12-14.<br />

Gadgil, P.D., Dick, M.A. and Dobbie, K., 2003. Fungi Silvicolae Novazelandiae: 4. New Zealand<br />

Journal <strong>of</strong> Forestry Science 33, 265-272.<br />

NZFOA (New Zealand Forest Owners Association). <strong>2009</strong>. New Zealand Forest Industry Facts and<br />

Figures 2008/<strong>2009</strong> booklet. Available from www.nzfoa.org.nz. [accessed 5 June <strong>2009</strong>].<br />

139


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Power, M.W.P. and Ramsfield, T.D., 2006. Nectria disease <strong>of</strong> radiata pine. Forest Health News<br />

165,1-2. Available from<br />

http://www.ensisjv.com/NewsEventsandPublications/Newsletters/ForestHealthNews/Fore<br />

stHealthNewsArchive/tabid/165/Default.aspx [accessed 6 March <strong>2009</strong>].<br />

Power, M.W.P. and Ramsfield, T.D., 2007. Nectria fuckeliana in pruned and unpruned trees. Forest<br />

Health News 175,1. Available from<br />

http://www.ensisjv.com/NewsEventsandPublications/Newsletters/ForestHealthNews/Fore<br />

stHealthNewsArchive/tabid/175/Default.aspx [accessed 6 March <strong>2009</strong>].<br />

Roll-Hansen, F. and Roll-Hansen, H., 1979. Micr<strong>of</strong>lora <strong>of</strong> sound-looking wood in Picea abies stems.<br />

European Journal <strong>of</strong> Forest Pathology 9, 308-316.<br />

Schultz, M.E. and Parmeter, J.R., 1990. A canker disease <strong>of</strong> Abies concolor caused by Nectria<br />

fuckeliana. Plant Disease 74,178-180.<br />

Smerlis, E., 1969. Pathogenicity tests <strong>of</strong> four pyrenomycetes in Quebec. Plant Disease Reporter<br />

53, 979-981<br />

Vasiliauskas, R. and Stenlid, J., 1997. Population structure and genetic variation in Nectria<br />

fuckeliana. Canadian Journal <strong>of</strong> Botany 75, 1707-1713.<br />

Vasiliauskas, R. and Stenlid, J., 1998. Fungi inhabiting stems <strong>of</strong> Picea abies in a managed stand in<br />

Lithuania. Forest Ecology and Management 109,119-126.<br />

140


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 141-149<br />

PRELIMINARY RESULTS OF MYCOFLORA ASSOCIATED WITH<br />

CANKERS ON Cupressus sempervirens var. horizontalis (Mill.) GORDON IN<br />

TURKEY<br />

ABSTRACT<br />

Asko Lehtijärvi 1* , H. Tugba Doğmuş- Lehtijärvi 1 , Funda Oskay 1 ,<br />

A. Gülden Aday 1<br />

1 Suleyman Demirel University, Faculty <strong>of</strong> Forestry, 32260 Isparta, Turkey<br />

*asko@orman.<strong>sdu</strong>.edu.tr<br />

Natural stands <strong>of</strong> C. sempervirens in Turkey are among the largest forests <strong>of</strong> this species<br />

in the world and are regarded as relicts <strong>of</strong> the centre <strong>of</strong> origin <strong>of</strong> var. horizontalis. However,<br />

phytopathological status <strong>of</strong> these stands was not investigated so far. In this study, canker<br />

formations were investigated in trees, saplings and seedlings <strong>of</strong> C. sempervirens var.<br />

horizontalis within natural stands located in Köprülü Kanyon National Park, Antalya and<br />

Aydıncık, Mersin. Incidence <strong>of</strong> the cankers and some other disease associated symptoms<br />

and signs <strong>of</strong> insect attacks on sampled trees were recorded. Isolations were made from<br />

cankers, and the obtained fungal cultures identified morphologically.<br />

Cankers were present on the trunk or branches <strong>of</strong> 34.4% <strong>of</strong> the totally 1023 trees<br />

sampled. Canker incidence was greater in Aydıncık than in Köprülü Kanyon, 38.0 % vs.<br />

29.8%, respectively. Totally 497 fungal isolates, representing 30 genera, were obtained.<br />

The most common species isolated from both Aydıncık and Köprülü Kanyon was:<br />

Phomopsis cf. occulta (44 and 22 %), unidentified coelomycete (22 and 6%), Alternaria<br />

spp. (5 and 13%), Cladosporium spp. (5 and 5%), Cytospora sp. (4 and 2%) and<br />

Pestalotiopsis funerea (1 and 37%, respectively). While P. cf. occulta was the most<br />

common species in Aydıncık, P. funerea - which was rare in Aydıncık - was the most<br />

frequently isolated species in Köprülü Kanyon. To our knowledge, with the exception <strong>of</strong> P.<br />

funerea, these species are new records on cypress in Turkey. Moreover, both P. occulta and<br />

P. funerea are reported to be pathogenic on cypress, e<strong>special</strong>ly under stress conditions.<br />

Keywords: Cupressus sempervirens var. horizontalis, Canker, Fungi<br />

1. INTRODUCTION<br />

Cupressus sempervirens L., the Mediterranean cypress or common cypress, is a<br />

long established cultivated forest tree species exterior to its natural geographic<br />

range. However its natural geographic distribution had been restricted to disjoint<br />

and <strong>of</strong>ten relict populations within Iran, Syria, Jordan, Lebanon, Libya, the<br />

Aegean Islands, Crete, Turkey, and Cyprus, which are thought to be being<br />

remnants <strong>of</strong> an extensive C. sempervirens forest, (Raddi and Sümer, 1999; Ducrey<br />

et al., 1999). The natural stands <strong>of</strong> Cupressus. sempervirens var. horizontalis<br />

(Mill.) Gordon in Turkey are considered among the most significant and largest<br />

natural Mediterranean cypress communities (Neyişçi, 1989; Özçelik, 2005), and<br />

141


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

regarded as the relicts <strong>of</strong> the original source <strong>of</strong> C. sempervirens var. horizontalis<br />

due to the high diversity observed among the populations (Koral et al., 1997; Raddi<br />

and Sümer, 1999; Ducrey et al., 1999). However, they constitute only 1392.5 ha <strong>of</strong><br />

forests within Turkey, where more than 75% <strong>of</strong> the total is degraded (Anonymous,<br />

2006). The largest pure and mixed stands (mixed with Pinus brutia Ten.) <strong>of</strong> the<br />

species are located in Köprülü Kanyon National Park, Antalya and in Aydıncık,<br />

Mersin. Other cypress forests in Turkey are located mostly in the south-western<br />

part <strong>of</strong> Turkey as small stands mainly mixed with Pinus brutia Ten. (Sabuncu,<br />

2004). As everywhere else within its distribution area, the Mediterranean cypress<br />

forests in Turkey have been frequently endangered by human activities, such as<br />

deforestation and wild fires, as well as by overgrazing.<br />

Several species <strong>of</strong> fungi have been associated with diseases <strong>of</strong> natural and<br />

cultivated varieties <strong>of</strong> C. sempervirens. Among these fungi, Seiridium cardinale<br />

(Wag.) Sutton and Gibson, the causal agent <strong>of</strong> the canker disease which had caused<br />

heavy damage in forests, nurseries and ornamental plantations, e<strong>special</strong>ly in the<br />

Mediterranean countries, has taken the most attention (Graniti, 1986, 1998). On the<br />

other hand, many pathogens, such as Diplodia pinea f.sp. cupressi Solel, Madar,<br />

Kimchi & Golan, Phomopsis occulta (Sacc.) Trav., Pestalotiopsis funerea (Desm.)<br />

Steyaert, Fusarium sp. Link, Cytospora sp. and Lasiodiplodia theobromae (Pat.)<br />

Griffon & Maubl., (syn: Botryodiplodia theobromae (Pat.)) have also been found to be<br />

the causal agents <strong>of</strong> the cankers on cypress (Solel et al., 1987; Bruck et al., 1990;<br />

Madar et al., 1991, 1996; Linde et al., 1997; Ducrey et al., 1999; Gonthier and<br />

Nicolotti, 2002; Bajo et al., 2008).<br />

Although there are many reports on the phytopathological problems – e<strong>special</strong>ly<br />

associated with Seiridium cardinale – <strong>of</strong> Mediterranean cypress where it has been<br />

introduced, information <strong>of</strong> the relict stands <strong>of</strong> C. sempervirens is available only for<br />

Greece and Cyprus (Xenopoulos and Diamandis, 1985; Tsopelas et al., 2007,<br />

2008). The only exception is the paper by Sümer (1987) reporting two pathogens in<br />

the Aegean coast <strong>of</strong> Turkey, S. cardinale (in Muğla) and P. funerea. No further<br />

studies have been performed since then.<br />

In this study, the presence <strong>of</strong> canker formations, top and crown diebacks,<br />

foliage symptoms, resin exudation from the trees and insect associations were<br />

investigated within natural cypress stands in two locations. In addition, isolation <strong>of</strong><br />

fungi from cankered tissues was tried.<br />

2. MATERIALS and METHODS<br />

2.1. Survey locations and sampling<br />

The surveys and samplings were conducted in two natural stands <strong>of</strong> C.<br />

sempervirens var. horizontalis located within Köprülü Kanyon National Park in<br />

Antalya province and in Aydıncık Forestry Enterprise in Mersin, during September<br />

and December 2008, respectively. The location, altitude, and some topographic and<br />

stand characteristics <strong>of</strong> the sampling plots are given in Table 1.<br />

142


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The Köprülü Kanyon National Park (37º01′N and 31º14′E) is located in the<br />

western part <strong>of</strong> the Taurus Mountains in Antalya province in southern Turkey. The<br />

annual average rainfall is 1140.5 mm and the annual average temperature is 18.3<br />

ºC. The cypress forest in the national park is composed <strong>of</strong> pure (195.6 ha) and<br />

mixed stands with P. brutia (255.6 ha) at elevations ranging from 650 to 950 m<br />

(Anonymous, 2008).<br />

Aydıncık (36º11′N and 33º27′E) is situated along hillsides <strong>of</strong> the central part <strong>of</strong><br />

the Taurus Mountains facing Mediterranean Sea coast in Mersin province, 325 km<br />

east <strong>of</strong> Antalya. The annual average rainfall is 936.2 mm and the annual average<br />

temperature is 19.1ºC. While degraded (292.5 ha) and normal (88.5 ha) stands<br />

comprised a total <strong>of</strong> 381 ha <strong>of</strong> cypress forest in Aydıncık, within the same basin the<br />

total cypress forests made up 1247.5 ha, at elevations from 20 to 120 m,<br />

representing the largest distribution area <strong>of</strong> this variety (Özçelik, 2005). The<br />

cypress forest within Aydıncık, composed <strong>of</strong> pure and mixed stands (P. brutia) was<br />

mostly spread along the valley <strong>of</strong> the Sipahili (Babadili) stream.<br />

Sampling<br />

plots<br />

Table 1: Characteristics <strong>of</strong> the sampling plots<br />

Location Latitude<br />

(N)<br />

Coordinates<br />

Longitude<br />

(E)<br />

Altitude<br />

(m a.s.l.)<br />

143<br />

Exposure Description<br />

KK1 Köprülü Kanyon 37 º12′ 31 º09′ 710 S Groups <strong>of</strong> seedlings and saplings along road side<br />

KK2 Köprülü Kanyon 37 º13′ 31 º08′ 844 S Groups <strong>of</strong> seedlings and saplings along road side<br />

KK3 Köprülü Kanyon 37 º13′ 31 º08′ 736 SE Groups <strong>of</strong> seedlings and saplings along road side<br />

KK4 Köprülü Kanyon 37 º13′ 31 º08′ 151 S Groups <strong>of</strong> seedlings and saplings along road side<br />

KK5 Köprülü Kanyon 37 º12′ 31 º09′ 740 SE Groups <strong>of</strong> seedlings and saplings along road side<br />

KK6 Köprülü Kanyon 37 º11′ 31 º11′ 737 S Groups <strong>of</strong> seedlings and saplings along road side<br />

M1 Aydıncık 36° 11' 33° 27' 49 W Degrade, Mixed with P. brutia<br />

M2 Büyükeceli 36 °11' 33° 27' 6 W Degrade, Mixed with P. brutia<br />

M3 Aydıncık, Babadili 36°12' 33° 27' 24 SE Degrade, Mixed with P. brutia, Gene protection forest<br />

M4 Aydıncık, Babadili 36°12' 33° 27' 29 E Degrade, Mixed with P. brutia, Gene protection forest<br />

M5 Aydıncık, Babadili 36°12' 33° 27' 172 E Pure, Gene protection forest<br />

M6 Aydıncık, Karaseki 36°10' 33°23' 61 NE Degrade, Pure, along stream<br />

M7 Aydıncık, Babadili 36° 12 ' 33° 27' 157 E Pure, Gene Protection Forest<br />

M8 Aydıncık, Babadili 36° 11' 33 °27' 114 NE Pure<br />

M9 Aydıncık, Babadili 36° 11' 33 °27' 72 E Degrade, Pure<br />

M10 Aydıncık, Duruhan 36°13' 33°17' 482 N<br />

Degrade, Mixed with P. brutia, along stream and<br />

agricultural fields<br />

In Köprülü Kanyon, seedlings and saplings <strong>of</strong> C. sempervirens along the road<br />

sides within 6 sampling plots, and in Aydıncık cypress trees <strong>of</strong> different age classes


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

within 10 sampling plots were investigated for the canker formations on branches<br />

and trunks. In Aydıncık, sampling plots were consist <strong>of</strong> four circular sub-sampling<br />

plots with a 10–m–radius, located 25 m to the north, east, south and west from the<br />

plot centre. While in Köprülü Kanyon the length <strong>of</strong> the cankers and the height <strong>of</strong><br />

the cankers above ground, as well as tree features were recorded, no such<br />

measurements were done in Aydıncık. Contrary to the survey conducted in<br />

Köprülü Kanyon, in Aydıncık trees were also assessed for top and crown dieback,<br />

resin exudation, foliage symptoms, and insect damage.<br />

2.2. Fungal isolation and identification<br />

Branch and trunk samples with cankers were collected from a number <strong>of</strong> trees<br />

within the sampling plots and transferred to the laboratory. The samples were<br />

surface sterilized with 70% ethanol, the outer tissues near the canker margins were<br />

removed with a sterile scalpel, and small pieces <strong>of</strong> the affected bark transferred<br />

onto potato dextrose agar plates (PDA; Merck, Darmstadt, Germany). Some <strong>of</strong> the<br />

samples were incubated in moist-chamber in order to induce formation <strong>of</strong> fruiting<br />

bodies on plant tissues. The cultures incubated at 25 ºC were identified according<br />

to their morphological characteristics.<br />

3. RESULTS AND DISCUSSION<br />

During the surveys, in Köprülü Kanyon and in Aydıncık, 449 and 574 trees,<br />

respectively, were examined (totally 1023 trees). Bark cracks and cankers, as well<br />

as resin drops and moderate to heavy resin flows were present on the trunk,<br />

branches and twigs at all sites. Of the 1023 trees examined, 34.4% (352) were<br />

bearing at least one canker. Canker incidence was greater in Aydıncık than in<br />

Köprülü Kanyon, 38.0 and 29.8%, respectively. Branch cankers were more<br />

frequent than trunk cankers in Aydıncık (31.7% and 17.9%, respectively), while in<br />

Köprülü Kanyon trunk cankers were significantly more common. Bark cracks<br />

exuding resin, e<strong>special</strong>ly on branches and twigs, were observed <strong>of</strong>ten in Aydıncık,<br />

where small perennial resin soaked cankers on declining lower branches were also<br />

found.<br />

While in Köprülü Kanyon the length <strong>of</strong> the cankers and the height <strong>of</strong> the<br />

cankers above ground, as well as the tree height and diameter were recorded (Table<br />

2), no such measurements were done in Aydıncık. In Köprülü Kanyon, among the<br />

seedlings and saplings investigated, the average height and diameter at breast<br />

height <strong>of</strong> 134 canker bearing individuals were 360.5 and 5.6 cm, respectively. In<br />

this site, majority <strong>of</strong> the cankers observed were on the road–facing parts <strong>of</strong> the<br />

saplings, mostly under breast height (mean 81.1cm). On the other hand, grazing<br />

damage was also remarkably more prevalent in Aydıncık, e<strong>special</strong>ly on lower<br />

branches where canker formations were also common. Therefore it is likely that<br />

wounds resulting from grazing have worked as suitable entrance points for canker–<br />

causing pathogens.<br />

144


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Table 2: Results <strong>of</strong> Köprülü Kanyon sampling area<br />

Sampling<br />

plot<br />

No <strong>of</strong> trees<br />

No <strong>of</strong><br />

trees with<br />

cankers<br />

Proportion<br />

<strong>of</strong> trees with<br />

cankers<br />

(%)<br />

145<br />

(trees with<br />

cankers)<br />

average values<br />

for; (cm)<br />

Height D1,3<br />

Average<br />

height <strong>of</strong><br />

cankers above<br />

ground<br />

(cm)<br />

KKK1 30 5 16.7 282.6 3.4 66.3<br />

KKK2 162 62 38.3 385.6 7.3 91.3<br />

KKK3 127 17 13.4 262.9 6.1 133.0<br />

KKK4 56 32 57.1 490.0 6.0 119.1<br />

KKK5 36 15 41.7 250.0 3.3 23.0<br />

KKK6 38 3 7.9 491.7 7.5 53.6<br />

Total 449 134 - - - -<br />

Avarage 74.83 22.33 29.2 360.5 5.6 81.1<br />

In Aydıncık, the number <strong>of</strong> trees per sampling plot varied from 28 to 86 with a<br />

mean value <strong>of</strong> 57.4 (±21SD). The average tree height and diameter in the whole<br />

Aydıncık sampling area were 586.7 (±368.8 SD) and 14.9 cm (±12.6 SD),<br />

respectively. There were significant differences in average diameter and height <strong>of</strong><br />

trees between sampling plots (Table 3). While the highest averages were in the<br />

sampling plot M7, the lowest ones were in the M8. On the other hand, correlations<br />

<strong>of</strong> symptom incidences with tree size (Table 4) were not significant, with the<br />

exception <strong>of</strong> the negative correlation between tree diameter and foliage symptoms.<br />

The highest canker incidence was in the M5 plot (52.3%), followed by the M8,<br />

M4, M9, M3 plots (range approximately 40-50%). The canker incidence (7.7%) as<br />

well as the other assessed symptoms had the lowest values in the M6 plot. The<br />

highest incidence <strong>of</strong> top dieback, crown dieback, resin exudation, and foliage<br />

symptoms was in the M9 (23.5%), M4 (18.1%), M8 (66.2%) and M10 plots<br />

(51.0%), respectively. Insect damage was the most frequent in the M8 plot (53.8%)<br />

and occasional in the M1 plot (3.6%).<br />

Table 3: Number and size <strong>of</strong> trees and incidence <strong>of</strong> symptoms in the sampling plots in<br />

Aydıncık.<br />

Sampling<br />

plot<br />

No <strong>of</strong> trees<br />

Diameter<br />

cm<br />

Tree sizes Number <strong>of</strong> individuals exhibiting symptoms <strong>of</strong>; .<br />

Height<br />

cm<br />

Canker<br />

Top<br />

dieback<br />

Crown<br />

dieback<br />

Resin<br />

exudation<br />

Foliage<br />

symptoms<br />

M1 28 17.6±11.9b 555.5±319.8b 6 1 1 4 1 1<br />

M2 53 15.0±10.2bc 572.8±374.7ab 16 3 1 10 5 1<br />

M3 83 15.4±11.9bc 609.3±380.3ab 33 8 11 47 5 9<br />

M4 83 15.2±12.7bc 634.8±415.8ab 36 9 15 50 17 20<br />

M5 86 13.0±10.0bc 583.3±307.9ab 45 5 7 50 17 22<br />

M6 39 17.3±11.8b 606.8±264.7ab 3 0 0 3 3 3<br />

M7 37 24.8±20.0a 734.4±480.1a 12 4 3 20 6 18<br />

M8 65 10.5±6.0c 497.4±353.1b 32 6 8 43 18 35<br />

M9 51 12.7±10.7bc 497.4±353.1b 21 12 9 24 23 15<br />

M10 49 14.6±16.8bc 576.8±481.4ab 14 4 2 27 25 18<br />

Σ 574 - - 218 52 57 278 120 142<br />

Avr.<br />

± SD<br />

57.4±20.9 14.9±12.6<br />

586.7±368.8<br />

Insect<br />

Damage<br />

21.8±14.0 5.2±3.7 5.7±5.1 27.8±18.7 12.0±8.9 14.2±10.8


Incidence (%)<br />

70,0<br />

65,0<br />

60,0<br />

55,0<br />

50,0<br />

45,0<br />

40,0<br />

35,0<br />

30,0<br />

25,0<br />

20,0<br />

15,0<br />

10,0<br />

5,0<br />

0,0<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

M1 M2 M3 M4 M5 M6 M7 M8 M9 M10 M<br />

Sampling Plots<br />

Canker Top dieback Crown dieback Resin exudation Foliage Symtomps Insect damage diameter height<br />

Figure 1: The mean incidences <strong>of</strong> canker, top and crown dieback, resin exudation,<br />

foliage symptoms and insect damage within sampling plots (M1-M10) and in the whole<br />

sampling area (M) in Aydıncık, as well as the mean diameter and height <strong>of</strong> the trees (the<br />

bars on height and diameter lines indicate the standard deviations).<br />

Top dieback, crown dieback and resin exudation were significantly correlated<br />

with canker formations (p


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

species isolated from both Köprülü Kanyon and Aydıncık were; P. cf. occulta (22<br />

and 44 %), unidentified coelomycete (6 and 22%), Alternaria spp. (13 and 5 %),<br />

Cladosporium spp. (5 and 5%), Cytospora sp. (2 and 4%) and P. funerea (37 and<br />

1%), respectively. While P. cf. occulta was the most common species in Aydıncık,<br />

P. funerea - which was rare in Aydıncık - was the most frequently isolated species<br />

in Köprülü Kanyon. Fungi isolated from both sampling area agrees slightly with<br />

those stated in previous studies (Munoz and Ruperez, 1980; Ducrey et al., 1999;<br />

Madar et al., 1991; Gonthier and Nicolotti, 2002; Bajo et al., 2008). Moreover, P.<br />

cf. occulta, P. funerea and Cytospora sp. are reported to be pathogenic on cypress,<br />

e<strong>special</strong>ly under stress conditions. However, neither S. cardinale nor other cypress<br />

canker related Seiridium species was recovered in this study. Nor were there well–<br />

known canker causing fungal pathogens <strong>of</strong> cypress, such as B. theobromae and D.<br />

pinea f.sp. cupressi, among the isolates. This indicates that either the pathogens<br />

were missing from the studied stands or replaced by other fungi in the sampled<br />

cankers. On the other hand, as the study areas were large, a sampling strategy<br />

focusing in sampling only the fresh cankers could have given different results.<br />

Nevertheless, the absence <strong>of</strong> e.g. S. cardinale in the study areas may not indicate<br />

that the native populations would be highly resistant against the pathogen, since the<br />

previously tested Turkish provenances (including Köprülü Kanyon –Zerk–, but not<br />

Aydıncık) tended to have only intermediate resistance against S. cardinale (Santini<br />

and Di Lonardo, 2000). The absence may be due to other factors including<br />

geographical barriers (Santini and Di Lonardo, 2000).<br />

In Köprülü Kanyon, P. funerea was isolated nearly from all cankers sampled. This<br />

species is endemic in Europe and is also present in the native areas <strong>of</strong> cypress, and<br />

therefore considered to have co-evolved with C. sempervirens (Santini and Di Lonardo,<br />

2000). It is considered a weak pathogen <strong>of</strong> a wide range <strong>of</strong> conifer hosts including<br />

species in the following genera: Cupressus, Pinus, Juniperus and Thuja (Madar et al.,<br />

1991; Sinclair et al., 1993; Santamaria et al., 2007). Moreover, it is also thought to be<br />

capable <strong>of</strong> replacing other pathogens such as S. cardinale (Sanches and Gibbs, 1995).<br />

During the survey in Köprülü Kanyon fruiting structures on plant tissues were not<br />

noticed, however in Aydıncık, there were unripe fruiting structures on the plants.<br />

Ecology <strong>of</strong> Cytospora sp., P. funerea and P. cf. occulta could differ between the two<br />

study areas, however, the differences could have been caused by the different sampling<br />

times or climate as well.<br />

In conclusion, canker incidence in natural C. sempervirens var. horizontalis<br />

stands in southern Turkey was relatively high. To our knowledge, all obtained<br />

fungal species, except P. funerea, are new records on cypress in Turkey. None <strong>of</strong><br />

the fungi is reported to be an aggressive pathogen, but could be harmful under<br />

stress conditions.<br />

ACKNOWLEDGEMENTS<br />

Financial support from the TUBITAK (Project No. TOVAG-108 O 287) is<br />

gratefully acknowledged. The authors are thankful to Dr. Gürsel Karaca for the<br />

identification <strong>of</strong> fungi.<br />

147


REFERENCES<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Anonymous, 2006. <strong>Orman</strong> Varlığımız, T. C. <strong>Orman</strong> Bakanlığı, <strong>Orman</strong> Genel Müdürlüğü, Ankara.<br />

Anonymous, 2008. Antalya İl ve Çevre <strong>Orman</strong> Müdürlüğü, Köprülü Kanyon Milli Parkı İşletme<br />

Şefliği Fonksiyonel <strong>Orman</strong> Amenajman Planı (2008-2017)<br />

Bruck, R.I., Solel, Z., Ben-Ze'ev, I.S., Zehavi, A., 1990, Diseases <strong>of</strong> Italian cypress caused by<br />

Botryodiplodia theobromae Pat. Eur. J. For. Path. 20:392-396.<br />

Bajo, J., Santamaría, O. and J. J. Diez, 2008. Cultural characteristics and pathogenicity <strong>of</strong><br />

Pestalotiopsis funerea on Cupressus arizonica. Forest Pathology, 38 (4), 263-274.<br />

Ducrey M. (ed.), Barthélémy D. (ed.), Pichot C. (ed.), Giannini R. (ed.), Raddi P. (ed.), Roques A.<br />

(ed.), Sales Luis J. (ed.), Thibaut B. (ed.), Teissier Du Cros E.. 1999. Cypress: a practical<br />

handbook. Florence, Studio Leonardo, 139 p.<br />

Graniti, A. 1986. Seiridium cardinale and other cypress cankers. Bull. EPPO/OEPP Bull. 16:479–<br />

486.<br />

Graniti, A., 1998. Cypress canker: A Pandemic in Progress. Annual Review <strong>of</strong> Phytopathology. 36:<br />

91- 114<br />

Gonthier, P., and G., Nicolotti, 2002. First report <strong>of</strong> Pestalotiopsis funerea on Cupressocyparis<br />

leylandii in Italy. Plant Dis. 86, 1402.<br />

Korol, L., Kara, N., Işık, K and Schiller, G., 1997. Genetic Differentiation Among and within Natural<br />

and planted Cupressus sempervirens L. Eastern Mediteranean populations. Silva<br />

Genetica. 46(2-3), p 151- 155.<br />

Linde, C., Kemp, G.H.J., Wingfield, M.J., 1997. First report <strong>of</strong> Sphaeropsls canker on cypress in<br />

South Africa. Eur. J. For. Path. 27:173-177<br />

Neyişçi, T., 1989. Beşkonak Saf Servi (Cupressus sempervirens L.) <strong>Orman</strong>ında Ekolojik<br />

Araştırmalar, <strong>Orman</strong>cılık Araştırma Enstitüsü Yayınları, Teknik Raporlar Serisi:43, s.49-<br />

76.<br />

Madar, Z., Solel, Z., Kimchi M., 1991. Pestalotiopsis canker <strong>of</strong> cypress in Israel, Phytoparasitica,<br />

Volume 19 (1);79-81<br />

Madar, Z., Kimchi M., Solel, Z., 1996. Fusarium canker <strong>of</strong> Italian cypress. Eur. J. For. Path. 26:107-<br />

112<br />

Munoz, C., Ruperez, A., 1980: Causas de la desaparición de los cipreses en España. Bol. Serv. Plagas,<br />

6: 95-104.<br />

Özçelik, R., 2005. Aydıncık Yöresi Doğal Dallı Servi (Cupressus sempervirens L. var. horizontalis<br />

(Mill) Gord) Meşcerelerinde Homojenlik Durumunun ve Gövde Niteliklerinin<br />

Değerlendirilmesi. Batı Akdeniz <strong>Orman</strong>cılık Araştırma Müdürlüğü Dergisi, Sayı:6 29-42.<br />

Raddi, S. and Sümer, S., 1999. Genetic diversity in natural Cupressus sempervirens L. populations in<br />

Turkey. Biochemical Systematics and Ecology 27. 799-814<br />

Sabuncu, R., 2004. Dallı Servide Fidan büyüme karakteristikleri Açısından Genetik Çeşitlilik, Batı<br />

Akdeniz <strong>Orman</strong>cılık Araştırma Müdürlüğü, Teknik Bülten No: 22, 28 s., Antalya.<br />

Sanchez, E. and Gibbs, J.N., 1995. The ecology <strong>of</strong> fungal cankers on Cupressus macrocarpa in<br />

southern England. Eur. J. For. Path. 25, 266-273<br />

148


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Santamaría, O., Bajo, J., Pajares, J.A. and Diez, J.J., 2007. Susceptibility <strong>of</strong> Cupressus arizonica and<br />

Pinus pinea to Pestalotiopsis funerea Isolates. Acta Silv. Lign. Hung., Spec. Edition, p.<br />

283<br />

Santini, A. and Di Lonarda, V., 2000. Genetic variability <strong>of</strong> the ‘bark canker resistance’ character in<br />

several natural provenances <strong>of</strong> Cupressus sempervirens. Forest Pathology. 30: 87-96.<br />

Sinclair, W. A.; Lyon, H. H.; Johnson, W. T., 1993. Diseases <strong>of</strong> Trees and Shrubs. New York:Cornell<br />

University Press.<br />

Solel, Z.,Madar, Z., Kimchi, M., Golan, Y., 1987. Diplodia canker <strong>of</strong> cypress. Can. J. Plant Pathol.<br />

9:115-118<br />

Sumer, S., 1987. The distribution <strong>of</strong> cypress (Cupressus L.) in Turkey and the current status in its<br />

pests and diseases, e<strong>special</strong>ly cypress canker disease. Istanbul Univ. <strong>Orman</strong> Fakultesi<br />

Dergisi 37:56-66.<br />

Tsopelas, P., Barnes, I., Wingfield M.J., Xenopoulus, S., 2007. Seiridium cardinale on Juniperus<br />

species in Greece. Forest Pathology. 37:338-347.<br />

Tsopelas, P., Angelopoulos, A., Nikolaou, K., 2008. Seiridium cardinale is a new threat to cypress<br />

trees in Cyprus. Plant Pathology 57(4), 784.<br />

Xenopoulos, S., And Diamandis, S., 1985. A Distribution Map For Seiridium cardinale Causing The<br />

Cypress Canker Disease In Greece, European Journal <strong>of</strong> Forest Pathology, V. 15(4) P.<br />

223-226<br />

149


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 150-161<br />

SOME MORPHOLOGICAL ASPECTS OF EUTYPELLA CANKER OF<br />

MAPLE (Eutypella parasitica)<br />

ABSTRACT<br />

Nikica OGRIS * , Barbara PIŠKUR, Dušan JURC<br />

Slovenian Forestry Institute, Večna pot 2, 1000 Ljubljana, Slovenia<br />

* nikica.ogris@gozdis.si<br />

Eutypella canker <strong>of</strong> maple originates from North America and was recently reported<br />

from Slovenia, Austria and Croatia. Disease distribution and frequency in the surveyed<br />

forest stands in Slovenia, Eutypella canker shape and its extent on the trunk, fungi present<br />

in discolored wood, perithecia density, and ascospore discharge were explored. Diseased<br />

maples were usually grouped into infection centers and the disease occurred on 3–5% <strong>of</strong> all<br />

maple trees at surveyed forest stands. However, incidence up to 30% was recorded. The<br />

canker was usually oval shaped and the average area <strong>of</strong> the canker was 48% <strong>of</strong> the affected<br />

part <strong>of</strong> the trunk. Canker width measured about a half (0.44) <strong>of</strong> canker length. 54.8% <strong>of</strong> all<br />

isolates (2,276) from discolored wood were identified as Eutypella parasitica. Perithecia<br />

covered an average <strong>of</strong> 32% <strong>of</strong> the total cankered area. A good correlation existed between<br />

the area with perithecia and the whole cankered area. In average, 647,000 perithecia per<br />

canker were found. The average discharge was 506,000 ascospores cm –2 h –1 . One Eutypella<br />

canker discharged from 65 million to 3.3 billion ascospores h –1 , with an average <strong>of</strong> 1.0<br />

billion ascospores hour –1 under favorable environmental conditions. The inoculation<br />

potential <strong>of</strong> the fungus is enormous but its rapid colonization <strong>of</strong> European forests is<br />

prevented by ineffective mode <strong>of</strong> transmission and slow development <strong>of</strong> the disease.<br />

Keywords: shape, area, perithecia density, ascospore discharge, frequency, Acer<br />

1. INTRODUCTION<br />

Eutypella canker <strong>of</strong> maple caused by the fungus Eutypella parasitica R. W.<br />

Davidson & R. C. Lorenz is a well-known disease in North America in the area<br />

around the Great Lakes where it was first found and described (Davidson and<br />

Lorenz, 1938). In Slovenia and Europe, the disease was not reported until 2005<br />

although it seems to have been present for some time prior to this (Jurc et al.,<br />

2006). The disease was also reported from Austria (Cech, 2007) and Croatia (Ogris<br />

et al., 2008). The means <strong>of</strong> the disease introduction into Slovenia is not known. The<br />

hosts <strong>of</strong> the disease are maples (Acer spp.). When a disease is introduced to new<br />

location, new hosts can emerge. Similar scenario was observed with Eutypella<br />

canker, when field maple (A. campestre L.) has been found to be a new host <strong>of</strong> E.<br />

parasitica (Ogris et al., 2005). About 35% <strong>of</strong> known cankered trees in Slovenia are<br />

field maples. The most susceptible maple in Slovenia is sycamore maple (A.<br />

150


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

pseudoplatanus L., 54% <strong>of</strong> cankered trees), while only 3.5% trees with Eutypella<br />

canker are Norway maple (A. platanoides L.). In North America, the disease is a<br />

common perennial canker on sugar maple (Acer saccharum Marsh.), red maple (A.<br />

rubrum L.), infrequently occuring on boxelder (A. negundo L.), Norway maple,<br />

silver maple (A. saccharinum L.), black maple (A. nigrum Mich.), sycamore maple,<br />

striped maple (A. pennsylvanicum L.), and bigleaf maple (A. macrophyllum Prush.)<br />

(Davidson and Lorenz, 1938; French, 1969; Kliejunas and Kuntz, 1972; 1974).<br />

One <strong>of</strong> the reasons for performing additional research on Eutypella canker was<br />

the outcome <strong>of</strong> a spread risk assessment <strong>of</strong> the disease for Europe (Ogris et al.,<br />

2006). 13% (1,404,000 km 2 ) <strong>of</strong> Europe's land area was found to be at very high risk<br />

for E. parasitica due to favorable host range and suitable climatic conditions.<br />

Regions with very high spread risk extends in the Balkans (Slovenia, Bosnia and<br />

Herzegovina, Serbia and Montenegro, Croatia), Southern Europe (some parts <strong>of</strong> the<br />

Apennines, the central part <strong>of</strong> the Pyrenees), Central Europe (all parts <strong>of</strong> Austria<br />

except the eastern part, the whole Czech Republic, northern and southern parts <strong>of</strong><br />

Slovakia, central and southern part <strong>of</strong> Germany, almost all <strong>of</strong> Poland except the<br />

northeastern part), Western Europe (northern half <strong>of</strong> Switzerland, eastern part <strong>of</strong><br />

France), and Eastern Europe (some parts <strong>of</strong> Moldova, eastern region <strong>of</strong> Ukraine,<br />

Caucasus). The sycamore maple, field maple, and Norway maple are predicted to<br />

be the most endangered.<br />

The aim <strong>of</strong> this work was to explore the morphological characteristics <strong>of</strong> the<br />

Eutypella canker that may explain the as yet unrevealed capacity <strong>of</strong> the disease to<br />

spread. This work represents a supplement to the current knowledge <strong>of</strong> the disease.<br />

Parts <strong>of</strong> the study are tests <strong>of</strong> validity that have been stated in different papers<br />

earlier but not yet tested; some hypotheses are checked again. This work supplies<br />

information about Eutypella canker shape, area, fungi present in discolored wood,<br />

perithecia density, ascospore discharge, the disease distribution tendencies and<br />

frequency in the stand.<br />

2. MATERIALS AND METHODS<br />

2.1. Description <strong>of</strong> the sites<br />

Samples were taken from six locations. (1) Rožnik hill is an urban forest just 1<br />

km from the center <strong>of</strong> Ljubljana, the capital city <strong>of</strong> Slovenia. Rožnik hill rises<br />

about 130 m above Ljubljana (429 m above sea level) and covers over 380<br />

hectares. For analysis, 16 diseased trees (13 sycamore maples, 3 field maples) were<br />

taken from three infection centers. (2) A single specimen on sycamore maple was<br />

collected at Topol at Medvode (650 m a.s.l.); this is one <strong>of</strong> the westernmost<br />

locations in Slovenia where Eutypella canker has been found and is 9 km far from<br />

Ljubljana. The other four sites are located in the eastern part <strong>of</strong> Slovenia. (3) Jelški<br />

hill (310 m a.s.l.) is 60 km east from Site 1 and is located near Sevnica rising over<br />

the Sava River. Two specimens on sycamore maple were found at this location and<br />

both trees were cut down and included in the analysis. (4) The fourth site is 7 km<br />

151


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

from Site 3 and is near to Sevnica. It is located on Kremenc hill (350 m a.s.l.)<br />

where two specimens <strong>of</strong> Eutypella canker on sycamore maple were found and<br />

collected on shady side <strong>of</strong> the hill. (5) A single diseased sycamore maple was<br />

found at Bohor hill (580 m a.s.l.), near the town <strong>of</strong> Kozje, which is located 22 km<br />

east <strong>of</strong> site 4. (6) Site 6 is located at Rogaška Slatina (230 m a.s.l.) where 55<br />

Eutypella cankers were found – all on field maple. This site is the most eastern<br />

place where Eutypella canker <strong>of</strong> maple was found in Slovenia. It is located about 1<br />

km from the border with Croatia. At Site 6 one field maple was gathered. Distance<br />

from Site 1 is 90 km. Altogether, <strong>of</strong> the 23 diseased trees collected, 19 trees were<br />

sycamore maple and 4 trees were field maple. Sites 1–5 had rich, rather moist soils<br />

in common which suits the sycamore maple’s demands for water, while Site 6<br />

parent material was sandstone, soils were quite warm and dry, and location was on<br />

sunny side <strong>of</strong> the hill.<br />

Site 1 and 6 were chosen to determine the frequency <strong>of</strong> Eutypella canker <strong>of</strong><br />

maple in forest stands.<br />

2.2. Common measurable characteristics<br />

Some basics measurements were made for all specimens collected. (1) The<br />

position <strong>of</strong> canker on trunk was measured from the ground to the center <strong>of</strong> the<br />

canker which is usually represented by a dead branch stub (Davidson and Lorenz,<br />

1938). When the centre <strong>of</strong> the canker was within arm’s reach the position was<br />

determined by tape measure, otherwise a Haglöf Sweden Vertex III height<br />

measurer was used. (2) The diameter <strong>of</strong> trees at breast height was measured by log<br />

calliper. (3) Canker circumference was determined by tape measure at the centre <strong>of</strong><br />

the canker. (4) Canker length was determined by tape measure.<br />

2.3. Area and shape <strong>of</strong> the canker on the trunk<br />

The canker area was measured and calculated for all 23 specimens collected.<br />

Area measurement was performed by fastening a transparent plastic foil to canker<br />

by pins and drawing the outline <strong>of</strong> the canker using permanent marker. Three<br />

different densities <strong>of</strong> perithecia, bark without perithecia, and areas where bark had<br />

already fallen <strong>of</strong>f were also marked on the foil. The foil was then put down on a<br />

flat surface with white paper for background and the contours were digitalized.<br />

Each digital photo was taken with a 10 cm long scale bar to allow for the correct<br />

calibration <strong>of</strong> the photos. The areas were calculated using S<strong>of</strong>t Imaging System<br />

analySIS ® Pro function for measuring arbitrary areas. The areas <strong>of</strong> different<br />

densities <strong>of</strong> perithecia were used later for calculating ascospore discharge at the<br />

level <strong>of</strong> the whole canker.<br />

2.4. Isolates<br />

Dissectional study enabled us to determine the places where the fungus actually<br />

lives and is active. It also enabled us to search for other species <strong>of</strong> fungi present in<br />

the canker. Trunk cross sections were taken at 10 cm intervals. The isolations were<br />

made as described below. Samples were taken within a period <strong>of</strong> 24 hours after<br />

152


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

dissecting the trunk into discs. Samples <strong>of</strong> discolored or decayed wood were taken<br />

from each second disc from eight cankered trunks. One trunk was analyzed in<br />

detail and samples were taken from each disc. At least three samples <strong>of</strong> healthy<br />

wood (showing no discoloration) from each trunk were taken as a control. When<br />

taking samples, attention was given to the margin <strong>of</strong> decayed wood, the various<br />

colors <strong>of</strong> decayed wood, the different phases <strong>of</strong> decay, and, in some cases, to the<br />

dried zone <strong>of</strong> wood observed around the margin <strong>of</strong> the decayed wood. The samples<br />

were taken with a boring machine with a wood borer <strong>of</strong> 12 mm diameter. Between<br />

each boring the borer was sterilized by flaming with 96% (v/v) ethanol and was<br />

then cooled down by dipping it in cool distilled water for few seconds. First, a 1–3<br />

mm deep hole was made to eliminate possible contaminants and after flame<br />

sterilization <strong>of</strong> the borer the sample for fungal isolation was taken. The boring<br />

chips were collected in sterile Petri dishes. Up to five samples were taken per disc.<br />

From each sample four isolations on 2% (w/v) malt extract agar (MEA) were made<br />

with two repetitions. The sample was represented by a boring chip 2–3 mm × 2–3<br />

mm in size. Altogether, 2276 isolations were made. Pure cultures were made <strong>of</strong><br />

every fungus growing from the sample. Cultures were incubated at 24°C for 4<br />

weeks.<br />

2.5. Density <strong>of</strong> perithecia and ascospore discharge<br />

Three different densities <strong>of</strong> perithecia were distinguished when the area <strong>of</strong> the<br />

Eutypella canker was measured. This enabled an assessment <strong>of</strong> ascospore<br />

production <strong>of</strong> the canker in a more accurate way. Three different densities <strong>of</strong><br />

perithecia were then sampled. Each sample was checked for the maturity <strong>of</strong> its<br />

perithecia before it was used in the test. The maturity test was performed on the<br />

bark area just next to the area <strong>of</strong> perithecia that was taken into the test <strong>of</strong> ascospore<br />

discharge. Mature perithecia are full <strong>of</strong> dark brown ascospores (Davidson and<br />

Lorenz, 1938), young perithecia are white inside, and old perithecia are empty<br />

when moistened. After the maturity test, the samples were cut to approximately 1<br />

cm 2 . The exact area <strong>of</strong> the sample was determined after the test using analySIS<br />

s<strong>of</strong>tware. Because the samples had been air-dried they had to be immersed in water<br />

for at least 30 min (Johnson and Kuntz, 1979). We immersed the samples in<br />

distilled water for 1 h, and then the bark samples with perithecia were put on moist<br />

filter paper for 3 hours, which is necessary for ascospore discharge to begin<br />

(Lachance, 1971; Johnson and Kuntz, 1979). The sample with underlaying moist<br />

filter paper was fixed to a rubber stopper with thin needle. Ascospores were<br />

discharged into test tubes with a 2 cm diameter to which 2 ml 0.01% (v/v)<br />

detergent Tween ® 80 solution had been added. Ascospores were allowed to<br />

discharge for 3 hours at room temperature about 22°C. Ascospores were counted<br />

using a Bürker-Türk counting chamber. The calculation <strong>of</strong> the number <strong>of</strong><br />

ascospores discharged per cm –2 h –1 was corrected using the exact area <strong>of</strong> the<br />

sample. The samples were also histologically examined and the number and<br />

maturity <strong>of</strong> the perithecia were determined.<br />

153


3. RESULTS<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

3.1. Incidence <strong>of</strong> the disease in forest stands<br />

Site 1 had the greatest number <strong>of</strong> diseased trees, altogether 70 trees. Eutypella<br />

cankers usually form infection centers, while single occurrences are rarely found<br />

(Kliejunas and Kuntz, 1974; Martinez, 2003) since ascospores are disseminated<br />

only over short distances (Johnson and Kuntz, 1979). There were four infection<br />

centers and six single trees distributed around the hill. A full inventory <strong>of</strong> one<br />

infection centre at Site 1 was performed. All trees were identified and measured for<br />

their diameter at breast height (DBH) within an area <strong>of</strong> 2.7 hectares. In the<br />

searched area, 20 tree species were determined. The most common species was the<br />

sycamore maple (41.7% <strong>of</strong> all trees), followed by the big leaf linden (Tilia<br />

platyphyllos Scop.) with 15.3%, the common hornbeam (Carpinus betulus L.) with<br />

10.6%, and the northern red oak (Quercus rubra L.) with 9.7%. Among 606<br />

sycamore maples in the 2.7 hectare area, 3.3% were diseased by E. parasitica. This<br />

is within the range <strong>of</strong> the usual incidence <strong>of</strong> Eutypella canker (Gross, 1984b).<br />

At Site 6 on the area <strong>of</strong> 7.2 ha, frequency <strong>of</strong> Eutypella canker was determined.<br />

28.6% <strong>of</strong> 192 field maples had canker.<br />

3.2. Common measurable characteristics<br />

Table 1 shows a summary <strong>of</strong> the statistics for four variables. The sample size was<br />

different for each variable measured. Eutypella cankers were found over the entire<br />

DBH range from thin to large, as previously reported by French (1969). The average<br />

trunk diameter at breast height was 25 cm. The relative frequency <strong>of</strong> cankered trees<br />

among 7 classes showed that 52% <strong>of</strong> diseased trees had a DBH between 11 and 22 cm.<br />

65% <strong>of</strong> all trees with Eutypella canker occurred on first 220 cm <strong>of</strong> trunk above<br />

the ground line, and 92% <strong>of</strong> all cankers occurred on the first 435 cm <strong>of</strong> trunk.<br />

These results are comparable to French (1969), who noted that 60% <strong>of</strong> all cankers<br />

occurred on the first 244 cm <strong>of</strong> trunk and 81.5% on the first 488 cm <strong>of</strong> trunk.<br />

33 cankers were measured for canker length. 85% <strong>of</strong> cankers did not exceed a<br />

length <strong>of</strong> 120 cm. Gross (1984a) measured 27 trees and the range <strong>of</strong> canker length<br />

was from 10 cm to 170 cm with an average <strong>of</strong> 67 cm. Wide range <strong>of</strong> canker length<br />

was directly related to the age <strong>of</strong> the canker and could be explained as such.<br />

Table 1: Summary statistics for common measurable characteristics <strong>of</strong> Eutypella canker<br />

Trunk DBH<br />

(cm)<br />

Vertical<br />

distribution (cm)<br />

154<br />

Length <strong>of</strong><br />

canker (cm)<br />

Width <strong>of</strong><br />

canker (cm)<br />

Sample size 38 37 33 31<br />

Minimum 11 35 30 23<br />

Maximum 61 1150 233 115<br />

Average 25 216 102 59<br />

95% confidence<br />

interval (±)<br />

3.6 73.9 18.6 9.5


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Canker width was correlated to canker length, which was due to canker shape,<br />

since the majority <strong>of</strong> cankers were oval shaped. Regression analysis with a linear<br />

model (canker width = 14.6 + 0.44 × canker length) gave correlation coefficient <strong>of</strong><br />

0.87. Coefficient 0.44 tells that canker width measured usually about a half <strong>of</strong><br />

canker length.<br />

Possible entry points for E. parasitica are branch stubs and bark wounds<br />

(French, 1969). Great majority <strong>of</strong> cankers were traced to branch stubs as the<br />

avenue <strong>of</strong> entry. Only 3 out <strong>of</strong> 89 cankers were associated with bark wounds and<br />

all three wounds resulted from logging operations.<br />

3.3. Area and shape <strong>of</strong> canker on the trunk<br />

The largest Eutypella canker measured 1.7 m 2 , although the average was much<br />

lower, i.e. 0.5 m 2 (Table 2). The canker area varied a lot because it depended upon<br />

the canker age. The cankered area ranged from 4% to 90% <strong>of</strong> the trunk area with<br />

the canker, with an average <strong>of</strong> 48%.<br />

The area without bark also depended upon the canker age and thus the whole<br />

canker area. In order to determine if in fact there was a relationship between the<br />

canker area without bark and whole canker area, regression analysis using a linear<br />

model was carried out. The linear model described a relationship between canker<br />

area without bark and whole canker area. The equation <strong>of</strong> the model was: area<br />

without bark = –2.32 + 0.25 × whole canker area. The high correlation coefficient<br />

(0.83) showed a good linear relationship between the canker area without bark and<br />

whole canker area. The coefficient <strong>of</strong> the equation tells that the area without bark<br />

represented usually a quarter <strong>of</strong> the whole canker area when the canker was old<br />

enough to bark fall <strong>of</strong>f. The bark begins to fall <strong>of</strong>f a trunk due to high degree <strong>of</strong><br />

decay or/and high activity <strong>of</strong> bark beetles.<br />

Perithecia do not usually develop nearer than the fourth or fifth annual callus<br />

ring from the margin <strong>of</strong> the cankers (Davidson and Lorenz, 1938). Therefore,<br />

Davidson and Lorenz (1938) stated that perithecia are confined to only a small<br />

central portion <strong>of</strong> large cankers. During this study, their hypothesis about the area<br />

size covered with perithecia was tested. The part <strong>of</strong> their hypothesis was rejected. It<br />

was confirmed that perithecia are usually located in the central portion <strong>of</strong> the<br />

canker but this area is not small. The total area <strong>of</strong> bark covered with perithecia<br />

could exceed 68% <strong>of</strong> the whole canker area. An average <strong>of</strong> 32% (±7% at 95%<br />

confidence level) <strong>of</strong> the canker area was covered with perithecia. Further analysis<br />

was performed and it was found that the correlation between the total area <strong>of</strong><br />

perithecia and the area <strong>of</strong> the whole canker was very good. Simple regression<br />

analysis produced a correlation coefficient <strong>of</strong> 0.94. A simple regression was made<br />

using a linear model: total area <strong>of</strong> perithecia = –3.93 + 0.42 × whole canker area.<br />

Three different densities <strong>of</strong> perithecia were delineated. The high densities <strong>of</strong><br />

perithecia were usually located nearer to the canker centre while low densities were<br />

located nearer to canker margin. It should be noted that a middle density <strong>of</strong><br />

perithecia could hardly be distinguished from high density. Therefore, some<br />

155


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

portion <strong>of</strong> the middle density <strong>of</strong> perithecia was included in the high density. It was<br />

suggested that about one-third <strong>of</strong> the total area covered with perithecia belonged to<br />

low density class, one third to middle density, and one third to high density class.<br />

20 out <strong>of</strong> 23 cankers were oval shaped, two cankers were triangular, and only<br />

one was shaped irregularly. All triangular shaped cankers had the centre <strong>of</strong> the<br />

canker close to the ground line. The irregular shape <strong>of</strong> the canker area has been<br />

already observed and explained (Davidson and Lorenz, 1938). The annual<br />

extension <strong>of</strong> the cankered areas is usually uniform and therefore the shape <strong>of</strong><br />

canker area is usually oval. However, this extension is occasionally stopped<br />

suddenly by some unknown cause, around either the entire margin or only a portion<br />

<strong>of</strong> it. Such sudden cessation <strong>of</strong> the extension gives the canker an irregular shape.<br />

The extension <strong>of</strong> the canker is faster in the longitudinal direction than in the<br />

transverse direction. This leads to the elongated oval shape <strong>of</strong> canker area. Cankers<br />

could be from 1.32 to 2.62 longer than they are wide while the average<br />

length/width proportion was 1.74 (± 0.18 at 95% confidence level).<br />

Within the canker shape analysis it was observed that the trunk gradually grows<br />

thicker up to canker centre and then the trunk thins down. The thickness <strong>of</strong> the<br />

canker was measured for 14 specimens at 10 cm intervals. It was determined that<br />

the thickness <strong>of</strong> the Eutypella canker was represented by a parabola very well.<br />

Polynomial regression using a second order polynomial model was performed for<br />

each <strong>of</strong> the 14 specimens. The R 2 was up to 97.8% and the average R 2 was 76.7%.<br />

This shows again that the canker shape is a regular oval.<br />

Table 2: Summary statistics for different kinds <strong>of</strong> areas <strong>of</strong> Eutypella cankers<br />

Canker area<br />

(dm 2 ) Whole Without bark low<br />

density<br />

156<br />

Perithecia (averages)<br />

middle<br />

density<br />

high<br />

density<br />

Minimum 12.3 0 0 0 0 0<br />

Maximum 168.0 43.7 14.9 16.6 45.8 64.7<br />

Mean 53.3 11.3 6.5 3.9 *<br />

95% confidence<br />

interval (±)<br />

Total<br />

10.9 20.5<br />

20.6 6.3 2.2 1.8 7.3 9.2<br />

* Middle density <strong>of</strong> perithecia could hardly be distinguished from high density.<br />

Therefore, some portion <strong>of</strong> the middle density <strong>of</strong> perithecia was included in the<br />

high density.<br />

3.4. Isolates<br />

Davidson and Lorenz (1938) stated that a single species <strong>of</strong> Eutypella has been<br />

consistently isolated from discolored wood from under the centre and from near the<br />

margins <strong>of</strong> the discoloration. Our study strongly supports this statement. From 9<br />

dissected trees, 97 discs and 237 boring holes, 1896 samples were cultivated on


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

MEA for possible fungal presence. The most frequent fungus isolated was E.<br />

parasitica which represented 54.8% <strong>of</strong> all isolates. The second most frequent<br />

organisms isolated from discolored wood were bacteria (27.6% <strong>of</strong> all isolates).<br />

17.6% <strong>of</strong> all isolates were represented by 25 macroscopically different species <strong>of</strong><br />

fungi among which were Cladosporium cladosporioides (Fresen.) G.A. de Vries,<br />

Mucor circinelloides Tiegh., Umbelopsis vinacea (Dixon-Stew.) Arx, Armillaria<br />

spp., Nectria spp., Pestalotia spp., Fusarium spp., and Pythium spp.<br />

The controls were made by cultivating 74 boring chips that were taken from<br />

healthy wood. The result <strong>of</strong> this test was that 65 <strong>of</strong> the isolations were sterile and<br />

only 9 Petri dishes were populated by bacteria and three different genera <strong>of</strong> fungi.<br />

No E. parasitica was present in controls.<br />

Dried healthy wood was present quite <strong>of</strong>ten at the center <strong>of</strong> the canker or near<br />

the margin <strong>of</strong> canker. 64 isolations were made for testing for the presence <strong>of</strong> E.<br />

parasitica in this area <strong>of</strong> the wood. After a month <strong>of</strong> incubation almost all<br />

isolations were negative and only seven had bacteria.<br />

As part <strong>of</strong> testing for the presence <strong>of</strong> E. parasitica in different parts <strong>of</strong> the<br />

wood, 420 isolates were made from the canker margin in the wood. E. parasitica<br />

was also the dominant fungus here (58.6%). The second most frequently found<br />

were bacteria. 23.6% <strong>of</strong> isolates contained eight macroscopically different fungi.<br />

17.9% <strong>of</strong> samples were sterile.<br />

3.5. Growth in pure culture<br />

Average radial growth in pure culture at room temperature <strong>of</strong> about 25°C is<br />

reported to be about 20 mm in seven days (Davidson and Lorenz, 1938). We<br />

measured the growth rate <strong>of</strong> 45 pure cultures on 1.5% (w/v) MEA at 24°C; 1712<br />

measurements were made. The average radial growth per day was 2.4 mm (± 0.1<br />

mm at 95% confidence level). The slowest measured average radial growth was 0.9<br />

mm per day and the fastest was 5.6 mm per day.<br />

3.6. Density <strong>of</strong> perithecia and ascospore discharge<br />

When measuring the area <strong>of</strong> the canker, three different densities <strong>of</strong> perithecia<br />

were determined (Table 2). The low density perithecia ranged from 181 to 210<br />

perithecia cm –2 with an average <strong>of</strong> 196 perithecia cm –2 . The middle density ranged<br />

from 234 to 269 perithecia cm –2 with an average <strong>of</strong> 255 perithecia cm –2 . The high<br />

density ranged from 301 to 378 perithecia cm –2 with an average <strong>of</strong> 348 perithecia<br />

cm –2 . These average densities <strong>of</strong> perithecia were used for calculating the number <strong>of</strong><br />

perithecia at the whole canker level and the total ascospore production per canker.<br />

The total number <strong>of</strong> perithecia per whole canker could be from 25,000 for young<br />

cankers to over 2,102,000 perithecia for old cankers. The average number <strong>of</strong><br />

peritecia per canker was 647,000 for the 23 Eutypella cankers analyzed. It should<br />

be noted that not all <strong>of</strong> these perithecia discharged ascospores because some <strong>of</strong><br />

them were still developing, while others were too old and empty. The maturity test<br />

was performed on those samples that were used in the study <strong>of</strong> ascospore<br />

157


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

discharge. Maturity <strong>of</strong> perithecia varied substantially. There were some samples<br />

that were 100% old and empty, then there were some samples with only young<br />

perithecia incapable <strong>of</strong> discharging the ascospores, and there were whole range <strong>of</strong><br />

samples that had a mixture <strong>of</strong> young, mature and old perithecia. Generally, the<br />

maturity tests showed that areas <strong>of</strong> canker with a lower density <strong>of</strong> perithecia have<br />

more young perithecia, while those with higher density have more old perithecia.<br />

Because the maturity test showed that there was no strict rule <strong>of</strong> maturity <strong>of</strong><br />

perithecia among different densities, the results <strong>of</strong> ascospore discharge could not be<br />

shown by low, middle, and high class densities <strong>of</strong> perithecia but only as average<br />

value. The range <strong>of</strong> ascospore discharge was from 360,000 to 718,000 ascospores<br />

cm –2 h –1 while the average discharge was 506,000 ascospores cm –2 h –1 . These<br />

results are comparable to the results <strong>of</strong> Lachance (1971) and Johnson and Kuntz<br />

(1979) who reported 450,000 ascospores cm –2 h –1 .<br />

On average, a perithecium discharged 2520 ascospores per hour i.e. 315 octads. The<br />

average ascospore discharge (506,000 ascospores cm –2 h –1 ) was used for calculating the<br />

total ascospore discharge <strong>of</strong> cankers which represents the potential <strong>of</strong> disease spread.<br />

One Eutypella canker could discharge from 65 million to 3.3 billion ascospores per<br />

hour with an average 1.0 billion ascospores per hour under favorable environmental<br />

conditions. This kind <strong>of</strong> ascospore production represents enormous inoculation<br />

potential.<br />

4. DISCUSSION<br />

This research pointed out some <strong>of</strong> morphological aspects <strong>of</strong> Eutypella canker <strong>of</strong><br />

maple that had not been analyzed before or were hypothesized and not checked, while<br />

others are checked again. A total <strong>of</strong> 23 Eutypella cankers were analyzed in detail. This<br />

is the required minimum for reliable statistics and it is comparable to other dissectional<br />

studies (Gross, 1984a).<br />

While ascospore discharge <strong>of</strong> Eutypella canker can be enormous under favorable<br />

environmental conditions, the question is raised as to why the disease is not more<br />

frequent. Specific demands <strong>of</strong> the fungus for successful colonization are the main<br />

reason for generally low occurrence <strong>of</strong> Eutypella canker. The fungus needs an exposed<br />

xylem for successful infection, i.e. branch stubs or bark wounds. Since the branch stubs<br />

represent a small area and ascospores are disseminated only over short distances (25<br />

m), it is hard for infection to take place despite enormous ascospore production. There<br />

are some other environmental limitations. Ascospore discharge is greatest at<br />

temperatures between 24 and 28 °C (Lachance, 1971; Johnson and Kuntz, 1979).<br />

Laboratory tests show no ascospore discharge and dissemination at temperatures below<br />

4 °C or higher than 36 °C. Free moisture (rainfall) induces mature perithecia to<br />

discharge the ascospores. At least 3 mm <strong>of</strong> rain has to penetrate the tree canopy to<br />

initiate discharge (Lachance, 1971; Johnson and Kuntz, 1979). Spore ejection begins<br />

about 2 hours after rain has started. High relative humidity alone is not sufficient to<br />

induce discharge <strong>of</strong> spores but it can influence the rate <strong>of</strong> drying <strong>of</strong> bark on cankers<br />

and prolongs discharge after periods <strong>of</strong> rainfall (Johnson and Kuntz, 1979).<br />

158


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The findings <strong>of</strong> the study show some morphological characteristics <strong>of</strong> Eutypella<br />

canker that help in understanding the disease spread. The very high average<br />

ascospore discharge <strong>of</strong> 506,000 ascospores cm –2 h –1 , or 1.0 billion ascospores h –1<br />

for the whole canker area is important for the success <strong>of</strong> disease spread. The study<br />

<strong>of</strong> the isolates showed that E. parasitica is the dominant fungus in infected and<br />

decayed wood. The fungus has proven its pathogenic ability with progressive<br />

attack tactics where the maple rarely has a chance to protect itself with callus or<br />

wound-wood. The Eutypella canker grows faster in longitudinal direction that in<br />

transverse direction. Therefore, the canker shape is in most cases almost regular or<br />

elongated oval. When the maple succeeds in defending itself from the fungus by<br />

forming wound-wood, the fungus can reinfect the wound-wood and bark from the<br />

underlaying wood. When the host is killed or snapped due to canker progress, the<br />

fungus acts as a saprobe and continues to produce ascospores for some years. The<br />

disadvantages <strong>of</strong> the fungus are its heavy ascospore octads which are disseminated<br />

only over short distances and the fact that they can infect only exposed xylem that<br />

occurs up to 4 m from the ground level.<br />

When a disease is introduced to a new distant location, e.g. another continent, it<br />

is hard to predict disease behavior and its spread speed. Environmental conditions<br />

could be more favorable for the disease and thus it could accelerate its spread rate.<br />

In a new environment new vectors <strong>of</strong> the disease could be present which could also<br />

add to the success <strong>of</strong> disease spread. A third factor that could help the disease to<br />

spread and establish are hosts. In a new environment new hosts are possible and<br />

host distribution can be over a wider area and contiguous. A new host, i.e. the field<br />

maple, was found and described in Slovenia (Ogris et al., 2005), but there are<br />

numerous other possible hosts (Acer spp.) present in Europe that have not yet been<br />

identified. There are at least 10 autochthonous maple species in Europe that could<br />

be the target <strong>of</strong> E. parasitica (Ogris et al., 2006). An inoculation study should be<br />

planned and carried out to test possible susceptibility. The results <strong>of</strong> that study<br />

would give needed information for more accurate assessment <strong>of</strong> the disease spread<br />

risk.<br />

Eutypella canker has been recognized as a new disease <strong>of</strong> maple in Europe.<br />

Despite intensive searching by the Slovenian Forestry Service only 225 Eutypella<br />

cankers were found from 2005–<strong>2009</strong>. We suspect that there are many undetected<br />

cases <strong>of</strong> Eutypella canker. Since it is assumed that cankers are present for a longer<br />

time period and they are dispersed over wide area we do not believe that the<br />

eradication <strong>of</strong> the disease is possible (Jurc et al., 2005). The disease can also have<br />

an economical impact. Incidence <strong>of</strong> the canker in North America is usually less<br />

than 5% (Gross, 1984b). Stands with over 10% <strong>of</strong> maples cankered were fairly<br />

common, and instances <strong>of</strong> over 40% cankering have been observed; in Slovenia<br />

there was a case with almost 29 % incidence. The disease frequently kills trees less<br />

than 7.5 cm in diameter and larger trees are predisposed to wind snap at the canker<br />

(French, 1969; Kliejunas and Kuntz, 1974). Cankered trees lose up to 50% <strong>of</strong> their<br />

merchantable cubic volume because most <strong>of</strong> the cankers occur below a height <strong>of</strong> 3<br />

159


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

m, which represents the most valuable portion <strong>of</strong> trunk (Gross, 1984b). Control<br />

measures are proposed within the framework <strong>of</strong> sanitation felling procedures. Part<br />

<strong>of</strong> the trunk with a canker at least 40 cm below and above the canker margin should<br />

be removed from the forest and destroyed by fire; when left in forest, the canker<br />

should be turned with face towards ground.<br />

5. ACKNOWLEDGEMENTS<br />

The research was funded by Ministry <strong>of</strong> Agriculture, Forestry and Food. Special<br />

thanks are due to staff <strong>of</strong> Slovenian Forestry Service: Mojca Bogovič, Marija<br />

Kolšek, Tibor Palfy, Dražen Drevenkar, Danijel Pavlin, Aleš Kolman, Aleš Šeško,<br />

Barbara Slobanja, Nataša Strle, and Jože Gobec who helped with finding and<br />

providing the cankers. We are most grateful to KPL d. d., Robert Rode for<br />

providing most <strong>of</strong> the cankers for the research. We would also like to thank Robert<br />

Krajnc and Zvone Kastelic for their technical help with the dissectional study, and<br />

Vesna Rajh for her help with culturing.<br />

6. REFERENCES<br />

Cech, T.L., 2007. Erstnachweis von Eutypella parasitica in Österreich. Forstschutz Aktuell 40, 10-13.<br />

Davidson, R.W., Lorenz, R.C., 1938. Species <strong>of</strong> Eutypella and Schizoxylon associated with cankers <strong>of</strong><br />

maple. Phytopathology 28, 733-745.<br />

French, W.J., 1969. Eutypella canker on acer in New York. Technical Publication 94, 56.<br />

Gross, H.L., 1984a. Defect associated with Eutypella canker <strong>of</strong> maple. Forestry Chronicle 60, 15-17.<br />

Gross, H.L., 1984b. Impact <strong>of</strong> Eutypella canker on the maple resource <strong>of</strong> the Owen Sound and<br />

Wingham forest districts. Forest Chronicle 60, 18-21.<br />

Johnson, D.W., Kuntz, J.E., 1979. Eutypella canker <strong>of</strong> maple: ascospore discharge and dissemination.<br />

Phytopathology 69, 130-135.<br />

Jurc, D., Ogris, N., Jakša, J., Jurc, M., 2005. Is an attempt to eradicate Eutypella canker <strong>of</strong> maple in<br />

Europe feasible? , EPPO conference on Phytophthora ramorum and other forest pests.<br />

EPPO, Falmouth, Cornwall, GB.<br />

Jurc, D., Ogris, N., Slippers, B., Stenlid, J., 2006. First report <strong>of</strong> Eutypella canker <strong>of</strong> Acer<br />

pseudoplatanus in Europe. Plant Pathology 55, 577.<br />

Kliejunas, J.T., Kuntz, J.E., 1972. Development <strong>of</strong> stromata and the imperfect state <strong>of</strong> Eutypella<br />

parasitica in maple. Canadian Journal <strong>of</strong> Botany 50, 1453-1456.<br />

Kliejunas, J.T., Kuntz, J.E., 1974. Eutypella canker, characteristics and control. The Forestry<br />

Chronicle 50, 106-108.<br />

Lachance, D., 1971. Discharge and germinatiion <strong>of</strong> Eutypella parasitica ascospores. Canadian Journal<br />

<strong>of</strong> Botany 49, 1111-1118.<br />

160


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Martinez, I., 2003. Eutypella canker clustering and volume loss in the western upper peninsula <strong>of</strong><br />

Michigan. Northern Journal <strong>of</strong> Applied Forestry 20, 186-187.<br />

Ogris, N., Diminić, D., Piškur, B., Kraigher, H., 2008. First report <strong>of</strong> Eutypella parasitica causing<br />

cankers on field maple (Acer campestre) in Croatia. Plant Pathology 57, 785.<br />

Ogris, N., Jurc, D., Jurc, M., 2005. Javorov rak (Eutypella parasitica: Ascomycota: Fungi) na<br />

gorskem javorju in maklenu: značilnosti in razlike. Gozdarski vestnik 63, 411-418.<br />

Ogris, N., Jurc, D., Jurc, M., 2006. Spread risk <strong>of</strong> Eutypella canker <strong>of</strong> maple in Europe. Bulletin<br />

OEPP/EPPO Bulletin 36, 475-485.<br />

161


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 162-167<br />

OCCURRENCE OF Pseudomonas syringae ON POPLAR DAMAGED BY<br />

NECROSIS AND CANKER<br />

ABSTRACT<br />

Irmtraut ZASPEL 1*<br />

1 Federal Research Institute for Rural Areas, Forestry and Fisheries (vTI)<br />

Institute <strong>of</strong> Forest Genetics Eberswalder Chausee 3A<br />

D-15377 Waldsieversdorf Germany<br />

*irmtraut.zaspel@vti.bund.de<br />

In poplar and aspen clones, established in a collection with more than 250 different<br />

genotypes, in nursery practice as well as in stock plant cultures necrosis and canker<br />

symptoms have appeared at 1 – 3 year-old shoots for some years. The injuries occur in<br />

winter season are characterized by ring necrosis at shoots, canker and cracking. The most<br />

frequent bacteria type that was isolated from damaged clones was identified as<br />

Pseudomonas syringae by means <strong>of</strong> 16S rDNA analysis. The pathogenicity <strong>of</strong> the isolates<br />

was evident at cuttings <strong>of</strong> eight several poplar clones tested under greenhouse conditions.<br />

The deleterious effect was visible at young shoots with symptoms <strong>of</strong> discoloration and leaf<br />

spots as well as blackening and wilting <strong>of</strong> buds and young shoots. The eight poplar<br />

genotypes differently responded to bacterial infection.<br />

Keywords: Populus, stem necroses, canker, Pseudomonas syringae<br />

1. INTRODUCTION<br />

Populus species are widely distributed and abundant in many different<br />

environments. A broad range <strong>of</strong> species and hybrids can be found as nursery crops<br />

grown for biomass production in short rotation coppice plantations and for<br />

reforestation and restoration purposes. In nurseries, plant material <strong>of</strong> the most<br />

poplar species is normally propagated with hardwood cuttings whereas aspen may<br />

be propagated from seeds.<br />

In poplar and aspen clones, established in a collection with more than 250<br />

different genotypes, in nursery practice as well as in stock plant cultures necrosis<br />

and canker symptoms had appeared at 1 – 3 year-old shoots in North-Eastern<br />

Germany for some years. The injuries were characterized by ring-shaped necroses<br />

with constrictions <strong>of</strong> the shoot, canker formations and spontaneous cracking. The<br />

extend <strong>of</strong> disease symptoms like necroses formation and shoot cracking increased<br />

during the winter season and disrupted in early spring. This pattern was different<br />

from the known damages caused by the more common pathogens like<br />

Aplanobacter populi and Cryptodiaporthe populea.<br />

162


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The frequency <strong>of</strong> necroses formation was changing depending on from the<br />

poplar genotype. At some susceptible genotypes, single shoots showed up to 10<br />

necroses. Some shoots with lesions that did not crack, dried above the damaged<br />

spots.<br />

Above and below the smaller necroses and lesions a wound periderm had<br />

developed, varying in thickness with the poplar genotype (fig. 1-3). Inside the<br />

shoots, tissue showed a hypertrophic ring-shaped zone with dark discolorations.<br />

2. METHODS AND RESULTS<br />

2.1. Bacteria identification<br />

During the search for the cause <strong>of</strong> losses, no indication for fungal pathogens or<br />

phytophagous insects could be found. Rather, investigations <strong>of</strong> damaged and<br />

broken shoots resulted in a range <strong>of</strong> bacterial strains, which were isolated directly<br />

from the margin <strong>of</strong> necrotic tissues and discoloured zones in winter season. The<br />

isolations followed a standard protocol using Yeast extract-Mannitol Agar (YMA)<br />

(Elkan and Bunn, 1992). The plates were incubated at 28 °C for three days. The<br />

most frequent type could be identified by means <strong>of</strong> 16S rDNA sequence analysis<br />

(primer pairs: fd1/ 926r and 799f/1492r, resp.) as Pseudomonas syringae pv.<br />

syringae. This species could be found only in the damaged annular areas with inner<br />

brown discolorations. The shoot segments between the lesions were free <strong>of</strong> P.<br />

syringae. Further species like Rahnella sp., Pantoea agglomerans,<br />

Pseudoclavibacter helvolus, Cellulomonas sp., and Microbacterium oxydans were<br />

found with lower frequency. Only one strain could be isolated from Xanthomonas<br />

sp. and X. campestris, resp., Curtobacterium flaccumfaciens, as well as some<br />

unspecified isolates.<br />

Figure 1, 2: Ring-shaped necrosis with wound periderm (left) and cracking <strong>of</strong> shoot<br />

(right)<br />

163


2.2. Pathogenicity testing<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 3: Cuttings with small necroses<br />

The pathogenicity <strong>of</strong> two strains <strong>of</strong> P. syringae was analysed at cuttings <strong>of</strong><br />

eight different poplar clones tested at 20-21°C under greenhouse conditions. The<br />

cuttings that were intact from outward appearance with a length for 12 – 15 cm<br />

were harvested after a frost period in winter, and were grown with illumination<br />

(35-60 µE m -2 s -1 ) in beakers with water for 8 days. After bud break, the cuttings<br />

were inoculated with bacterial suspension (10 8 cfu/ml) for two days, plugged into<br />

moist sand, and cultivated for further five weeks at 20-22°C with additional light<br />

<strong>of</strong> 4 h (180-200 µE m -2 s -1 ) in greenhouse. The deleterious effect caused by the<br />

bacteria was assessed by counting the cuttings with symptoms weekly. Symptoms<br />

that were visible on young leaves and shoots had started with discoloration and<br />

leaf spots after 3 weeks and resulted in blackening and wilting <strong>of</strong> buds and young<br />

shoots. In general, 37.8 % and 40.9 % <strong>of</strong> the cuttings (n=88 and 87, resp.), treated<br />

with two different P. syringae strains developed severe symptoms on young shoots<br />

(table 1). Most <strong>of</strong> the shoots <strong>of</strong> affected cuttings died <strong>of</strong>f within the observation<br />

time <strong>of</strong> five weeks, even though the eight poplar genotypes responded to bacterial<br />

infection quite differently. The P. maximowiczii hybrids as well as the aspen clone<br />

were more susceptible than the other clones tested. Within the observation time,<br />

some single control cuttings <strong>of</strong> three poplar genotypes (4 % <strong>of</strong> total) died <strong>of</strong>f by<br />

unknown causes.<br />

164


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Table 1: Pathogenicity <strong>of</strong> Pseudomonas syringae strains on poplar clone cuttings<br />

Species / Hybrid Clone<br />

165<br />

Cuttings with visible damages (%)<br />

P. syringae<br />

strain 47<br />

P. syringae<br />

strain 48<br />

control<br />

Populus trichocarpa 30 21 (14) 1 0 (11) 0 (13)<br />

Populus trichocarpa 872 33 (9) 25 (12) 0 (9)<br />

Populus trichocarpa 936 27 (11) 45 (11) 0 (10)<br />

Populus trichocarpa x P.<br />

deltoides<br />

Populus tremula x P.<br />

tremuloides<br />

Populus deltoides x P.<br />

maximowiczii<br />

Populus maximowiczii x P.<br />

berolinensis<br />

Populus maximowiczii x P.<br />

nigra<br />

995 29 (12) 18 (11) 0 (10)<br />

L 211 56 (16) 79 (14) 8 (13)<br />

960 56 (9) 67 (9) 13 (8)<br />

Geneva 37 (8) 33 (9) 0 (9)<br />

Rochester 44 (9) 60 (10) 11 (9)<br />

1 in brackets: number <strong>of</strong> cuttings used for average<br />

The limited number <strong>of</strong> cuttings as well as the fact that some cuttings did not<br />

root resulted in the different number <strong>of</strong> replications per clone and treatment. Only<br />

those cuttings were finally assessed regarding bacteria sensitivity, which showed a<br />

sufficient root system. The inadequate rooting ability <strong>of</strong> the poplar clones can<br />

could have induced water stress resulting in similar wilting symptoms like the<br />

disease symptoms.<br />

3. DISCUSSION AND CONCLUSION<br />

P. syringae is a known pathogen on woody plants, not only in poplars, but also<br />

in ash, willow, cherry, horse chestnut, and others. In recent time, to P. syringae<br />

more than 50 pathovars were assigned. On woody plants, the pathogen causes bark<br />

necroses with slime flux and canker. The newly-discovered pathovar <strong>of</strong> horsechestnut,<br />

P. syringae pv. aesculi can act as a cambium destroyer with severe<br />

dieback (Stobbe et al., 2008; Kaminski et al., 2007).


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

In poplars, the pathogen played only a secondary role for a long time because<br />

poplar cultures were less important for timber industry in Germany. In other<br />

countries like New Zealand, Sweden, and Romania, observations <strong>of</strong> P. syringae<br />

were made on poplars for a long time (Spiers, 1990; Ramstedt et al., 1994; Mocanu<br />

& Poleac, 1965). The pattern <strong>of</strong> damages that was described were spotted leaves,<br />

wilting <strong>of</strong> leaves and shoots in the whole, as well as bark necroses. In current<br />

investigations in a Chinese poplar breeding program P. syringae pv. syringae was<br />

identified as the reason for injuries like bark necroses on cuttings <strong>of</strong> various hybrid<br />

clones (Xiang et al., 2001). Moreover, this species could be found in symptomless<br />

hybrid clones <strong>of</strong> different poplar sections (Ulrich et al., 2008).<br />

The disease transmission <strong>of</strong> the pathogen is unclear until now. Possibly<br />

mechanical bark wounds induced by insects and other animal vectors are<br />

responsible, considerably earlier in summer. The invasion <strong>of</strong> bark tissue is also<br />

possible by bark micro-cracks following weather conditions like summer drought<br />

or autumn frost. The propagation <strong>of</strong> the pathogen takes place within a large<br />

temperature range and can still be found slightly above zero. Sequent changing <strong>of</strong><br />

warm and cold periods with freeze stress could have promoted the infection by the<br />

pathogen and their distribution within the tissue. The formation <strong>of</strong> necroses in the<br />

cold season can be caused by pathovars belonging to ice nucleation inducing<br />

strains <strong>of</strong> P. syringae. Those pathotypes are able to induce the formation <strong>of</strong> ice<br />

crystals within plant cells downward from -1°C. As a result, the cell walls burst and<br />

the protoplasm can serve as nutrient resources for further bacteria growth.<br />

Moreover, some pathovars <strong>of</strong> the P. syringae group are known as phytotoxine<br />

producers. Analysed compounds are the lipodepsipeptides Syringomycin and<br />

Syringopeptin, possessing a cell wall destroying influence (Hutchinson and Gross,<br />

1997).<br />

In recent time, poplars became more important as a fast growing tree species for<br />

biomass production in short rotation coppice plantations, and a high demand for<br />

suited propagation material exists. A basic precondition <strong>of</strong> utilizing poplar for this<br />

purpose is the provision <strong>of</strong> proved propagation material with high quality and free<br />

<strong>of</strong> harmful organisms. The use <strong>of</strong> cuttings grown in infected stock plant cultures<br />

can lead to a fast distribution <strong>of</strong> the disease. It is difficult to appreciate a possible<br />

role <strong>of</strong> P. syringae as a serious pathogen for poplar cultures in Germany.<br />

Therefore, work on epidemiology, vectors, and host-specific association <strong>of</strong> various<br />

pathovars is recommended as well as research referring to resistance /<br />

susceptibility <strong>of</strong> clones and cultivars <strong>of</strong> economical importance.<br />

166


4. LITERATURE CITED<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Elkan, G. H. & Bunn, C.R., 1992. The Procaryotes: A Handbook on the Biology <strong>of</strong> Bacteria –<br />

Ecophysiology, Isolation, Identification, Applications, pp 2197-2210. EDS. BY A. Balows,<br />

H.G. Trüper, M. Dworkin, W. Harder & K.H. Schleifer, NEW YORK<br />

Hutchinson, M.A. & Gross, D.C., 1997. Lipopeptide Phytotoxins Produced by Pseudomonas syringae<br />

pv. syringae: Comparison <strong>of</strong> the Biosurfactant and Ion Channel Forming Activities <strong>of</strong><br />

Syringopeptin and Syringomycin. MPMI 10:347-354<br />

Kaminski, K., Wagner, S. & Werres, S., 2007. Neuartige Krankheit an Rosskastanien. Stadt und Grün<br />

56:55-57<br />

Mocanu, V. & Poleac, E., 1965. Studies on bacterial canker <strong>of</strong> poplar. Stud. Cerc. Inst. For. Bucuresti<br />

25:211-229<br />

Spiers, A.G., 1990. Bacterial leaf spot and canker <strong>of</strong> poplar, willow, and alder. Forest Pathol. New<br />

Zealand 21:4-7<br />

Ramstedt, M., Aström, B. & Fircks, H. A. v. (1994): Dieback <strong>of</strong> poplar and willow caused by<br />

Pseudomonas syringae in combination with freezing stress. Eur. J. For. Path.24: 305-315<br />

Stobbe, H., Schmidt, O., Moreth, U., Kehr, R. & Dujesiefken, D., 2008. Pseudomonas: Derzeitige<br />

Verbreitung. AFZ/Der Wald 4/2008, pp 176-177<br />

Ulrich, K., Ulrich, A. & Ewald, D., 2008. Diversity <strong>of</strong> endophytic bacterial communities in poplar<br />

grown under field conditions. FEMS Microbiol. Ecol. 63:169-180<br />

Xiang, C.T., Dong, A.R., Liu, X.F., Li, C., Yuan, S.Z. & Zhang, J.H., 2001. Key factors for the<br />

causing poplar ice nucleation active bacterial canker and its control techniques. J. For.<br />

Research 12:157-164<br />

167


Rust Diseases<br />

169


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

170


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 171-176<br />

SEASONAL FRUITING AND SPORULATION OF THEKOPSORA AND<br />

CHRYSOMYXA CONE RUSTS IN NORWAY SPRUCE CONES AND<br />

ALTERNATE HOSTS IN FINLAND<br />

Juha KAITERA 1* , E. TILLMAN-SUTELA 1 and A. KAUPPI 2<br />

1 Finnish Forest Research Institute, Muhos Research Unit, FI-91500, Muhos, Finland. 2 University<br />

<strong>of</strong> Oulu, Department <strong>of</strong> Biology, P.O. Box 3000, FI-90014 Oulu<br />

ABSTRACT<br />

*Juha.kaitera@metla.fi<br />

Seasonal fruiting and sporulation <strong>of</strong> cone rusts were investigated in Norway spruce<br />

cones and alternate hosts in 2006-2008. Current-year and one-year-old cones and leaves <strong>of</strong><br />

alternate hosts, Pyrola spp. and Orthilia secunda, were collected from case-stands in<br />

southern and northern Finland bi-monthly or monthly and checked for rust fruitbodies.<br />

Fruitbodies and different fruiting structures <strong>of</strong> the rusts were examined using light<br />

microscopy (LM), and field emission scanning electron microscope (FESEM).<br />

Spermogonia <strong>of</strong> Chrysomyxa pirolata and aecia <strong>of</strong> Thekopsora areolata developed<br />

in current-year cones in June, while C. pirolata aecia developed and began to sporulate in<br />

July. Thekopsora areolata aecia sporulated mainly in previous year’s cones in May-August.<br />

Uredinia, telia and basidia <strong>of</strong> C. pirolata developed in overwintered Pyrola spp. and<br />

Orthilia secunda leaves in May, and sporulated in May-June. Uredinia <strong>of</strong> T. areolata<br />

developed in current-year P. padus leaves in June and sporulated in June-August. Telia <strong>of</strong><br />

T. areolata developed in late summer, but no basidia developed in overwintered P. padus<br />

leaves in March-May.<br />

Keywords: Cone rusts, Thekopsora areolata, Chrysomyxa pirolata, sporulation<br />

1. INTRODUCTION<br />

Good quality seed crops <strong>of</strong> Picea abies (L.) Karst. are irregular due to insects<br />

and pathogens reducing both the amount and quality <strong>of</strong> seed crop in seed orchards<br />

and natural forests throughout Finland (Kangas, 1940; Rummukainen, 1960;<br />

Nikula and Jalkanen, 1990; Tillman-Sutela et al., 2004)). Thekopsora areolata (Fr.)<br />

Magnus, and Chrysomyxa pirolata Wint., cause severe damage on Picea spp.<br />

throughout the northern hemisphere (Savile, 1950; Gäumann, 1959; Roll-Hansen,<br />

1965; Ziller, 1974). Thekopsora areolata infects Prunus spp. (Gäumann, 1959),<br />

while C. pirolata infects species in genera Pyrola, Moneses and Orthilia (Savile,<br />

1950; Gäumann, 1959; Ziller, 1974). Uredinia and telia develop on alternate hosts<br />

after aeciospore infection.<br />

In 2006, florescence and cone crop <strong>of</strong> Norway spruce were abundant, but due to<br />

fungal injuries the seed crop was severely reduced in some seed orchards. The aim<br />

<strong>of</strong> this study was to collect information <strong>of</strong> rust sporulation after a serious rust<br />

outbreak to improve disease control. For a thorough description <strong>of</strong> the study, see<br />

Kaitera et al. (<strong>2009</strong>).<br />

171


2. MATERIAL AND METHODS<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Norway spruce cones were collected in a seed orchard (Stand 1) in southern<br />

Finland, and in a naturally regenerated stand (Stand 2) in northern Finland. A<br />

random sample <strong>of</strong> previous year’s (2006) and current-year (2007) cones, were<br />

collected in Stand 1 in 5 times in 2007. Similar sampling was performed in Stand 2<br />

for the current-year cones in 5 times in 2007 and for the previous year’s cones in 8<br />

intervals in 2006-2007. The cones were cut from sample trees either using branch<br />

scissors (Stand 1) or a lifting cage (Stand 2).<br />

Overwintered and current-year leaves <strong>of</strong> Prunus padus, Pyrola sp. and Orthilia<br />

secunda were collected along the cone sampling in May-October 2007 in Stand 1,<br />

and 7-19 times in 2007-8 in Stand 2. In a third group <strong>of</strong> seed orchards (Stand 3) in<br />

southern Finland, leaves <strong>of</strong> P. padus and Pyrola sp. were collected in 8 times in<br />

2007. Overwintered P. padus leaves were collected in March-May and current-year<br />

leaves were collected between early May and early October.<br />

The occurrence, incidence and distribution <strong>of</strong> rust fruitbodies (spermogonia,<br />

aecia), and proportion <strong>of</strong> sporulating fruitbodies were recorded in cones under<br />

stereo microscope. Leaves <strong>of</strong> alternate hosts were investigated for the occurrence,<br />

incidence and stage <strong>of</strong> sporulation <strong>of</strong> rust uredinia, telia and basidia per leaf. Rust<br />

fruiting stages were also selected and further studied using a JEOL JSM 6300F<br />

field emission scanning electron microscope (FESEM).<br />

3. RESULTS<br />

3.1. Rust incidence and sporulation in current-year cones<br />

In Stand 1, cones collected in late May to mid-June bore no rust fruitbodies. In<br />

late June, about 2 % <strong>of</strong> the sample cones carried C. pirolata spermogonia and 5 %<br />

carried immature T. areolata aecia. None <strong>of</strong> the aecia were sporulating. The T.<br />

areolata aecia located on both sides <strong>of</strong> cone scales along the entire cone. No fungal<br />

structures resembling T. areolata spermogonia were observed. In the early August,<br />

7 % <strong>of</strong> the sample cones bore C. pirolata aecia that located on outer (abaxial) side<br />

<strong>of</strong> scales, being currently sporulating or already had finished sporulating and<br />

ruptured. Two percent <strong>of</strong> the sample cones carried T. areolata aecia that occurred<br />

in all the scales <strong>of</strong> cones. None <strong>of</strong> the T. areolata aecia sporulated yet. In the early<br />

October, 2 % <strong>of</strong> the cones carried both T. areolata and C. pirolata aecia in cone<br />

scales. Chrysomyxa pirolata aecia were ruptured with only individual aeciospores<br />

within aecia. Thekopsora areolata aecia were immature and non-sporulating.<br />

3.2. Rust incidence and sporulation in previous years’ cones<br />

3.2.1. Stand 1<br />

Only T. areolata aecia were observed in previous year’s cones. In late May, 35<br />

% <strong>of</strong> the sample cones bore T. areolata aecia with 92 % <strong>of</strong> the cone scales being<br />

172


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

infected. All observed aecia located on both sides <strong>of</strong> cone scales with 96 % <strong>of</strong> the<br />

aecia per cone being currently sporulating or had finished sporulating being<br />

ruptured. From the mid-June to early October, 22 % - 45 % <strong>of</strong> the sample cones<br />

carried T. areolata aecia. All infected cones carried aecia on both sides <strong>of</strong> cone<br />

scales with 93 % - 100 % <strong>of</strong> the cone scales per cone carrying aecia.<br />

3.2.2. Stand 2<br />

In cones collected in October 2006, 90 % carried T. areolata aecia and 10 %<br />

carried C. pirolata aecia. Aecia <strong>of</strong> T. areolata were non-sporulating and located in<br />

both sides <strong>of</strong> cone scales with 94 % <strong>of</strong> the scales per cone being infected.<br />

Chrysomyxa pirolata aecia had all finished sporulating, locating in 77 % <strong>of</strong> the<br />

infected cones on the outer side <strong>of</strong> cone scales.<br />

In late March in 2007, 90 % <strong>of</strong> the scales carried T. areolata aecia on both sides<br />

<strong>of</strong> the scales in previous year’s cones that were non-sporulating. About 2 % <strong>of</strong><br />

these cones carried already sporulated and ruptured C. pirolata aecia on outer side<br />

<strong>of</strong> the scales with 70 % <strong>of</strong> the scales carrying aecia per infected cone. In the early<br />

May <strong>of</strong> 2007, 43 % <strong>of</strong> the sample cones carried T. areolata aecia and 7 % carried<br />

C. pirolata aecia .<br />

From late May to late June, 95 % <strong>of</strong> the sample cones bore T. areolata aecia and<br />

15 % bore C. pirolata aecia. Aecia <strong>of</strong> T. areolata occurred on both sides <strong>of</strong> cone<br />

scales in most <strong>of</strong> the cones with 86 % - 93 % <strong>of</strong> the scales per cone carrying<br />

fruitbodies. All C. pirolata aecia had finished sporulating.<br />

In late July and early October, 92 % - 94 % <strong>of</strong> the sample cones carried T.<br />

areolata aecia, and 0 % - 19 % <strong>of</strong> them carried C. pirolata aecia. The average<br />

proportions <strong>of</strong> cone scales bearing T. areolata aecia were 89 % - 95 % per cone.<br />

On average, 18 % - 23 % <strong>of</strong> the scales per cone carried sporulating T. areolata<br />

aecia.<br />

3.3. Rust incidence and sporulation on alternate hosts<br />

3.3.1. Stand 1<br />

In late May 2007, 1 % <strong>of</strong> the overwintered O. secunda leaves bore C. pirolata<br />

uredinia and undifferentiated fruitbodies. Undifferentiated fruitbodies were<br />

common on O. secunda in late June, but after that they became rare. Uredinia were<br />

common between mid-June and early August, after which they became rare. No<br />

sporulation was observed in these uredinia until early August in 2007, after which<br />

uredinia finished sporulating and ruptured. None <strong>of</strong> the current-year leaves <strong>of</strong> P.<br />

padus bore any T. areolata fruitbodies in late May, but practically all leaves carried<br />

uredinia from mid-June on and telia without basidia since late-June.<br />

3.3.2. Stand 2<br />

In late March 2007, none <strong>of</strong> the overwintered O. secunda and Pyrola sp. leaves<br />

carried C. pirolata fruitbodies. Undifferentiated fruitbodies occurred first in early<br />

173


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

May, but they were most common between mid-May and late May in 2007-8.<br />

Uredinia appeared in mid-May, being most frequent in late May with low coverage<br />

per infected leaf (2 % - 10 %) in 2007. Uredinia were, however, most frequent<br />

fruitbodies in 2008 with their coverage ranging between 47 % - 83 %. Uredinia<br />

started to sporulate immediately after their formation and continued throughout the<br />

sample collection. Telia <strong>of</strong> C. pirolata were the most frequent fruitbodies in 2007,<br />

but they were less common than uredinia in 2008. Telia with basidiospores were<br />

observed throughout the collection, but they were most common between late May<br />

and mid-June in 2007 and between early and mid-June in 2008. In 2007 telia began<br />

to rupture from late July on.<br />

Practically all overwintered P. padus leaves bore T. areolata telia without<br />

external basidia between late March and late May in 2007-8. Uredinia <strong>of</strong> T.<br />

areolata occurred on the lower leaf surface with less frequency than telia on<br />

overwintered P. padus leaves. Uredinia started to sporulate immediately after<br />

formation in early June, and they began to rupture in early July. Until mid-<br />

September, all uredinia had finished sporulating.<br />

3.3.3. Stand 3<br />

In late April in 2007, no fruitbodies <strong>of</strong> C. pirolata occurred on the overwintered<br />

leaves <strong>of</strong> O. secunda or Pyrola sp. Undifferentiated fruitbodies occurred on all<br />

such leaves with the highest coverage <strong>of</strong> 70 % per leaf in mid-May, after which<br />

their coverage decreased. First C. pirolata uredinia were observed with a low<br />

coverage per infected leaf in mid-May. Thereafter, uredinia occurred in increasing<br />

frequency until early October with a coverage <strong>of</strong> uredinia per leaf lower than 40 %.<br />

Sporulation <strong>of</strong> uredinia started in mid-May and continued until early October.<br />

Since early July the proportion <strong>of</strong> uredinia that had finished sporulating increased,<br />

and after late July uredinia ruptured.<br />

Telia were the most common fruitbodies <strong>of</strong> C. pirolata in 2007. The first telia<br />

appeared in late May in almost all infected leaves with high (87 %) coverage per<br />

infected leaf. Telia with basidia were frequent until mid-June, when they began to<br />

rupture. Ruptured telia were common from early July to early October.<br />

Almost all <strong>of</strong> the overwintered P. padus leaves carried T. areolata telia without<br />

basidia in late April. Sporulated and ruptured uredinia occurred on 38 % <strong>of</strong> the<br />

current-year sample leaves with a low (6 %) coverage per infected leaf. A few<br />

uredinia occurred in 17 % <strong>of</strong> the infected leaves, which all sporulated in late May.<br />

From mid-May until early October, almost all P. padus leaves carried sporulating<br />

T. areolata uredinia, and the proportion <strong>of</strong> uredinia that had finished sporulation<br />

increased after early July. Telia were observed in all infected leaves between early<br />

July and early October without external basidia.<br />

174


4. DISCUSSION<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

In this study, T. areolata and C. pirolata, which are causes for severe rust<br />

epidemics in Finland, fruited and sporulated in young current-year cones the next<br />

year after a serious rust outbreak. The high incidence <strong>of</strong> T. areolata aecia in<br />

previous year’s cones confirmed that it was the main cause for the serious rust<br />

outbreak in the study areas in 2006. In all stands, T. areolata uredinia and telia<br />

were frequent on both overwintered and current-year P. padus leaves indicating<br />

that alternate host infection took frequently place both in 2006-2007. The high<br />

disease incidence coincided with the incidence <strong>of</strong> aecia in previous year’s cones,<br />

too. In this study, T. areolata aecia sporulated mainly in one-year-old cones as<br />

reported (Jørstad, 1925), but single sporulating aecia could be found already in late<br />

summer after infection. The observed aeciospore and urediniospore size and<br />

morphology was in accordance with the reports elsewhere (Gäumann, 1959; Saho<br />

and Takahashi, 1970). As practically all scales in infected cones bore T. areolata<br />

aecia, the rust was highly pathogenic and systemic in the cones and hindered<br />

efficiently seed formation.<br />

The occurrence and morphology <strong>of</strong> spermogonia, spermatia, aecia, aeciospores,<br />

uredinia, urediniospores, telia and basidia <strong>of</strong> C. pirolata corresponded well to<br />

earlier reports (Gäumann, 1959; Sutherland et al., 1984; Crane and Hiratsuka,<br />

2000). As most scales in infected cones were covered by C. pirolata aecia, the rust<br />

was highly pathogenic and systemic in cones. Therefore, the rust had a great<br />

impact on cone development causing seed deformation in diseased cones. The<br />

gelatinuous young fruiting structures, undifferentiated fruitbodies, corresponded to<br />

those described previously (Crane and Hiratsuka, 2000), and reported to develop<br />

either into uredinia or telia depending on the amount <strong>of</strong> moisture and free water.<br />

5. REFERENCES<br />

Crane, P.E., Hiratsuka, Y., 2000. Evidence for environmental determination <strong>of</strong> uredinia and telia<br />

production in Chrysomyxa pirolata (inland spruce cone rust). Canadian Journal <strong>of</strong> Botany<br />

78, 660-667.<br />

Gäumann, E., 1959. Die Rostpilze Mitteleuropas. Beiträge zur Kryptogamenflora der Schweiz 12, 1-<br />

1407.<br />

Jørstad, I., 1925. Norske skogsykdommer I. Nåletresykdommer bevirket av rustsopper, ascomyceter<br />

og fungi imperfecti. Meddelelser fra det Norske Skogforsøksvesen 62, 19-186.<br />

Kaitera, J., Tillman-Sutela, E., Kauppi, A., <strong>2009</strong>. Seasonal fruiting and sporulation <strong>of</strong> Thekopsora and<br />

Chrysomyxa cone rusts in Norway spruce cones and alternate hosts in Finland. Canadian<br />

Journal <strong>of</strong> Forest Research (in press)<br />

Kangas, E., 1940. Cone injuries and seed crop <strong>of</strong> Norway spruce in 1937. Communicationes Instuti<br />

Forestalis Fenniae 29, 1-36<br />

Nikula, A., Jalkanen, R., 1990. Kuusen käpytuholaisten ja –tautien esiintyminen Pohjois-Suomessa<br />

kesällä 1989. Metsäntutkimuslaitoksen Tiedonantoja 362, 83-89.<br />

175


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Roll-Hansen, F., 1965. Pucciniastrum areolatum on Picea engelmannii. Identification by<br />

spermogonia. Meddelelser fra det Norske Skogforsøksvesen 22, 389-397.<br />

Rummukainen, U., 1960. Abundance and quality <strong>of</strong> seed injuries <strong>of</strong> Norway spruce.<br />

Communicationes Instituti Forestalis Fenniae 52, 1-83.<br />

Saho, H., Takahashi, I., 1970. Notes on the Japanese rust fungi VI. Inoculation experiments on<br />

Thekopsora areolata (Fr.) Magnus, a cone rust <strong>of</strong> Picea spp. in Japan. Transactions <strong>of</strong> the<br />

Mycological Society <strong>of</strong> Japan 11, 109-112.<br />

Savile, D.B.O., 1950. North American species <strong>of</strong> Chrysomyxa. Canadian Journal <strong>of</strong> Research, C 28,<br />

318-330.<br />

Sutherland, J.R., Hopkins, S.J., Farris, S.H., 1984. Inland spruce cone rust, Chrysomyxa pirolata, in<br />

Pyrola asarifolia and cones <strong>of</strong> Picea glauca, morphology <strong>of</strong> the spore stages. Canadian<br />

Journal <strong>of</strong> Botany 62, 2441-2447.<br />

Tillman-Sutela, E., Kauppi, A., Hilli, A., Kaitera, J., 2004. Fungal injury to seed tissues <strong>of</strong> Norway<br />

spruce, Picea abies (L.) Karst. Trees 18, 151-156.<br />

Ziller, W.G., 1974. The tree rusts <strong>of</strong> western Canada. Canadian Forestry Service Publications 1329,<br />

1-272.<br />

176


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 177-181<br />

PRELIMINARY STUDIES ON GENETIC VARIATION IN<br />

Gymnosporangium fuscum in the LAKES DISTRICT OF TURKEY<br />

DETECTED WITH M13 MINISATELLITE MARKER<br />

ABSTRACT<br />

Asko LEHTIJÄRVI 1* , H. Tuğba DOĞMUŞ-LEHTIJÄRVI 1 ,<br />

A. GüldenADAY 1 , Funda OSKAY 1<br />

1 Suleyman Demirel University Faculty <strong>of</strong> Forestry 32260 CUNUR/ISPARTA<br />

*asko@orman.<strong>sdu</strong>.edu.tr<br />

Gymnosporangium fuscum infections on the trunk and branches <strong>of</strong> Juniperus excelsa<br />

are common in natural stands in the Lakes District <strong>of</strong> Turkey. In the present study, level <strong>of</strong><br />

genetic variation among G. fuscum isolates was estimated. Telial horns were obtained from<br />

trunk lesions in Sütçüler, Bucak-Aziziye and Beşkonak sites. From each telium DNA was<br />

extracted by using plant mini kit. PCR amplification pr<strong>of</strong>iles were run using the M13<br />

minisatellite core sequence. Preliminary results indicated low variation among the isolates.<br />

Key words: European pear rust, Turkey, Minisatellite, M13, genetic variation<br />

1. INTRODUCTION<br />

Crimean juniper (Juniperus excelsa Bieb) contributes to 5.6 % <strong>of</strong> forest area in<br />

Turkey. It grows on dry rocky slopes <strong>of</strong> hills and mountains at elevations ranging<br />

from 150 to 2700 m above sea level, and <strong>of</strong>ten forms the tree line in the Taurus<br />

Mountains. Crimean juniper has been under certain level <strong>of</strong> protection since 1996<br />

when all silvicultural treatments in the juniper forests were ceased due to the bad<br />

condition <strong>of</strong> the stands (Güner et al., 2000).<br />

European pear rust is caused by the fungus Gymnosporangium fuscum DC. like<br />

rusts in general, it alternates between two hosts: J. excelsa and Prunus spp. The<br />

infections are perennial on the coniferous host, on which in spring it develops the<br />

characteristic telial horns.<br />

European pear rust is widely distributed throughout Europe with observations<br />

(including) extending to Asia Minor (Lebanon, Syria and Turkey) and North Africa<br />

(Algeria and Morocco). The pathogen has also been introduced to North America<br />

(California, Washington, and British Columbia) probably through the importation<br />

<strong>of</strong> junipers from Europe (Laundon, 1977; Hollebone, 2006). In Turkey, perennial<br />

lesions caused by the rust are common on Crimean juniper in the Lakes district<br />

(Doğmuş-Lehtijärvi et al., 2008).<br />

177


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Minisatellite sequences are repeated over the genome and exist in most organisms<br />

providing an obtainable and highly variable probe and PCR amplification (Karlsson,<br />

1993; Högberg et al., 1995). In most cases they have been used to study variation<br />

among populations.<br />

In our study, we used M13 minisatellite core sequence as a primer in PCR based<br />

DNA fingerprinting technique to study level <strong>of</strong> genetic variation among G. fuscum<br />

populations.<br />

2. MATERIAL AND METHODS<br />

Telial horns were collected from trunk lesions in Sütçüler, Bucak-Aziziye and<br />

Beşkonak sites. In Sütçüler three stands 3-8 km apart were sampled. The stands in<br />

Beşkonak and Bucak-Aziziye were 40 and 70 km west from the Sütçüler stands,<br />

respectively.<br />

In each stand telia horns were sampled from ten trees, placed into plastic bags,<br />

and stored in the laboratory at -20 ºC until DNA extraction.<br />

Figure 1. Locations <strong>of</strong> the sampled stands (crosses).<br />

2.1. DNA extraction and PCR amplification<br />

Telia were grounded using liquid nitrogen and DNA was extracted with Qiagen<br />

DNeasy Plant Mini Kit. PCR amplification pr<strong>of</strong>iles were obtained using M13<br />

minisatellite core sequence as a primer. Reactions were performed in volumes <strong>of</strong><br />

50 μl containing Tris-HCl 10 mM, pH 8; dNTPs 0.2 mM; MgSO4 2 mM; Tsg<br />

polymerase 1,25 U; primer (M13) 2 μM; approximately 10 ng <strong>of</strong> template DNA.<br />

PCR was conducted using a hot start step at 95°C for 10 min, followed by 45<br />

cycles at 95°C for 1 min, 48°C for 30 s, 72°C for 2 min and a final extension at<br />

72°C for 10 min.<br />

178


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

2.2. Analyzes <strong>of</strong> amplification products<br />

Amplification products were separated by electrophoresis (90 V for 120 min) in<br />

1.5 % agarose gels in TAE buffer and length <strong>of</strong> the products were determined by<br />

using 100 bp DNA Ladder. The presence or absence <strong>of</strong> amplification products was<br />

scored to analyze the variation among the populations.<br />

3. RESULTS<br />

The amplifications yielded a total <strong>of</strong> 18 markers ranging from 300 to 1200 bp in<br />

size. The amplification pr<strong>of</strong>iles were very similar for most <strong>of</strong> the<br />

Gymnosporangium telial horns; the M13 primer seemed not detect any significant<br />

level <strong>of</strong> variation within 50 samples.<br />

Figure 2: Amplification products <strong>of</strong> G. fuscum isolates from different sites.<br />

4. DISCUSSION<br />

The very low level <strong>of</strong> variation detected by M13 amplifications was surprising.<br />

Within its 2- year life cycle G. fuscum produces relatively limited number <strong>of</strong><br />

genetically identical spores. Each aecidium on pear is initiated by a single<br />

basidiospore and therefore could be expected to be (at least in practise) genetically<br />

different from other aecidia. As the fungus lacks an uredinial stage on pear the<br />

number <strong>of</strong> infections on pear caused by single genets is not multiplied. As a result,<br />

genetic variation among aecidiospores infecting junipers should be high. As new<br />

infections on juniper occur annually, one could expect to find a high level <strong>of</strong><br />

genetic variation among G. fuscum isolates on that host.<br />

179


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

As the results are preliminary we can not exclude the possibility <strong>of</strong> non-optimal<br />

conditions in the PCR reactions. However, the amplicon pr<strong>of</strong>iles looked normal<br />

without signs <strong>of</strong> e.g. abnormally long fragments. Therefore other explanations are<br />

more likely.<br />

Arbitrarily primed PCR using M 13 core sequence as a primer detects variation<br />

in several <strong>of</strong> species <strong>of</strong> fungi growing on forest trees including such as;<br />

basidiomycetes Heterobasidion annosum s.l. (Fr) Bref. (Hantula et al., 1996;<br />

Stenlid et al., 1993)and Fomitopsis pinicola (Schwarts: Fr) Karst. (Högberg et al.,<br />

1995), ascomycetes Ophiostoma ulmi (Buism), Ceratocystis fimbriata f.sp. platani<br />

(Santini et al., 2000), and Nectria fuckeliana C. Booth (Vasiliauskas and Stenlid,<br />

1997) and Sphaeropsis sapinea Dyko& Sutton (Xiaoqin et al., 2007) among fungi<br />

imperfecti. Among biotrophs the primer has been used successfully for powdery<br />

mildews (ascomycete) but to our knowledge not for smuts and rusts<br />

(basidiomycetes). It is possible that annealing sites for the M13 primer within G.<br />

fuscum genome are few. Another reason for the low variation could be that the<br />

distances between the sampled stands were short.<br />

Table 1. Presence and absence vector <strong>of</strong> amplification products.<br />

Isolates 330bp 360bp 380bp 390bp 400bp 450bp 540bp 550bp 590bp 650bp 660bp 680bp 710bp 810bp 870bp 910bp 950bp 1200bp<br />

S1.1 1 1 0 0 0 0 0 0 0 1 0 0 0 1 0 0 0 0<br />

S1.2 1 1 0 0 0 0 1 1 0 0 0 0 0 0 0 0 0 0<br />

S1.3 1 1 0 0 0 0 1 1 0 1 0 0 0 0 0 0 0 0<br />

St2.1 1 0 1 0 0 0 1 0 0 1 0 0 0 0 0 1 1 1<br />

St2.2 1 1 0 0 0 0 1 1 0 1 0 0 0 1 0 0 0 0<br />

St2.3 1 1 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0<br />

St2.6 0 0 1 0 0 0 1 1 0 1 0 0 0 0 0 0 0 0<br />

B3.1 0 1 0 0 0 0 1 1 0 1 1 1 0 0 1 0 0 1<br />

B3.2 0 1 0 0 0 0 1 0 0 1 1 1 0 0 0 0 0 1<br />

B3.3 0 1 0 0 0 0 1 1 0 1 0 0 0 0 0 0 0 0<br />

B3.4 0 1 0 0 0 0 1 1 0 1 0 0 0 0 0 0 0 0<br />

B4.4 0 0 1 0 0 0 1 1 0 1 0 0 0 0 0 0 1 1<br />

Be4.5 0 0 1 0 0 0 1 0 0 1 0 0 0 0 0 0 1 1<br />

Be4.6 1 0 0 0 0 0 1 1 0 1 0 0 0 0 0 1 0 0<br />

Be5.8 1 1 0 0 0 0 1 1 0 0 0 0 0 0 0 1 0 0<br />

Be5.9 1 1 0 0 0 0 1 1 0 0 0 0 0 0 0 1 0 0<br />

Be5.10 1 1 0 0 0 0 1 1 0 0 0 0 0 0 0 1 0 0<br />

5. ACKNOWLEDGEMENTS<br />

This work was supported by State Planning Organisation <strong>of</strong> Turkish Republic<br />

(DPT–2003 K 1211020-7).<br />

180


6. REFERENCES<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Doğmuş-Lehtijärvi, H. T.; Lehtijärvi, A. ; Aday, A. G., <strong>2009</strong>. European pear rust on Juniperus<br />

excelsa L. in south-western Turkey. Volume 39 (1): 35-42.<br />

Güner, A.; Özhatay, N.; Ekim, T.; Başer, K. H. C. (Eds.), 2000. Flora <strong>of</strong> Turkey and the East Aegean<br />

Islands, Vol. XI, Supplement - II, University Press, Edinburgh. 654 p.<br />

Hintz, W.E., Jeng, R.S., Hubbes, M.M., and Horgren, P.A., 1991. Identification <strong>of</strong> three populations<br />

<strong>of</strong> Ophiostoma ulmi by mitochondrial DNA restriction-site mapping and nuclear DNA<br />

fingerprinting. Exp. Mycol. 15,316-325.<br />

Hollebone J.E., 2006. Domestic Regulations <strong>of</strong> Pear Trellis Rust, Gymnosporangium fuscum Hedw. f.<br />

[online], Canadian Food Inspection Agency, Directive D-97-01, cited: 21/12/06, revised:<br />

04/07/06, Plant Health Division, Ottawa, Ontario, Available at the Internet:


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 182-186<br />

FACTORS FAVOURING BROOM RUST INFECTION IN ADVANCE<br />

PLANTINGS OF Abies alba IN SW-GERMANY<br />

ABSTRACT<br />

Tilo PODNER 1 and Berthold METZLER 1*<br />

1 Forest Research Institute <strong>of</strong> Baden-Wuerttemberg<br />

Department <strong>of</strong> Forest Protection, D- 79100 Freiburg/Br., Germany<br />

*berthold.metzler@forst.bwl.de<br />

For ecological reasons, even aged stands <strong>of</strong> Norway spruce (Picea abies) are going to<br />

be converted to mixed forests, including silver fir (Abies alba) and other native tree species.<br />

In order to avoid clear cuttings, the alternative tree species are <strong>of</strong>ten introduced into the<br />

spruce stands by advance plantings. However, after about two decades <strong>of</strong> planting, several<br />

stands <strong>of</strong> planted silver fir turned out to be severely infected by Melampsorella<br />

caryophyllacearum. If the stems are affected, wood quality is deteriorated by burl<br />

formation and even more by the risk <strong>of</strong> secondary infection by wood decay fungi like<br />

Phellinus hartigii which makes the stems break in timber age.<br />

A study was performed in order to quantify the disease incidence in advance plantings<br />

<strong>of</strong> silver fir under Norway spruce by using circular sample plots. Brooms respective<br />

cankers formed after Melampsorella-infections in branches and in stems were counted<br />

separately. Amongst others the stand structure and the presence <strong>of</strong> caryophyllaceous host<br />

plants (mainly Stellaria nemorum) bearing the dicaryotic phase <strong>of</strong> the rust fungus in the<br />

neighbourhood <strong>of</strong> the plots were recorded. The presence <strong>of</strong> alternate host plants turned out<br />

to be the most crucial factor for disease incidence, whereas, there was no evidence that<br />

geology, altitude, or stand density could play a major role. Since Stellaria plants were<br />

mainly found along forest roads or in logging lines, it is recommended that silver fir not be<br />

planted right beside such places.<br />

1. INTRODUCTION<br />

More than 100.000 ha (7.6 %) <strong>of</strong> the forest area <strong>of</strong> the SW-German land Baden-<br />

Wuerttemberg is covered by silver fir Abies alba MILL. This is by far the biggest<br />

area <strong>of</strong> this tree species in Germany. In order to achieve a more nature-oriented<br />

forest, the state forest administration is going to raise the proportion <strong>of</strong> A. alba to<br />

11% and to diminish monocultures <strong>of</strong> Norway spruce (Picea abies), presently by<br />

far the most dominant forest tree species. In contrast to the latter species, silver fir<br />

is neither object <strong>of</strong> butt rot by Heterobasidion spp. nor endangered by severe bark<br />

beetle attacks.<br />

In order to avoid clear cuttings, silver fir is <strong>of</strong>ten introduced into the spruce<br />

stands by advance plantings in small gaps. However, after about two decades <strong>of</strong><br />

planting, several stands <strong>of</strong> planted silver fir turned out to be severely infected by<br />

Melampsorella caryophyllacearum SCHROET. Forest administration <strong>of</strong> Baden-<br />

182


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Wuerttemberg reports fir broom rust to be economically important on 575 ha<br />

(Schroeter et al., <strong>2009</strong>).<br />

The increment <strong>of</strong> the trees is not directly affected by the rust infection (Solla et<br />

al., 2006). However, if the stems are affected, wood quality is deteriorated by burl<br />

formation and even more by secondary infection by wood decay fungi, primarily<br />

Phellinus hartigii, which makes the stems break in timber age. Thus, besides the<br />

economical damage the rust fungus is driving higher biodiversity in silver fir stands<br />

(Holdenrieder, 1994).<br />

We performed a study in order to quantify the disease incidence in advance<br />

plantings <strong>of</strong> silver fir under Norway spruce. In the past, broom rust was designated<br />

as a major problem in silver fir forest (Heck, 1894), so it should be elucidated<br />

whether this is true under the given circumstances today.<br />

Figure 1. M. caryophyllacearum-stem canker<br />

(arrows) in two Abies alba plants.<br />

2. MATERIALS AND METHODS<br />

183<br />

Figure 2. Several cankers on the trunk and<br />

in branches closed to the trunk <strong>of</strong> a silver fir.<br />

Eleven forest stands with advance plantings and two stands with natural<br />

regeneration <strong>of</strong> silver fir were selected in the eastern slope <strong>of</strong> Black Forest in the<br />

SW-German districts <strong>of</strong> Breisgau-Hochschwarzwald, Schwarzwald-Baar, and Tuttlingen.<br />

Altitude and geology are listed in Table 1. Age class and density <strong>of</strong><br />

regeneration and canopy were evaluated according to forest inventory methods<br />

(Kramer and Akca, 1995). Three to seven circular sample plots <strong>of</strong> 25 m² each were<br />

placed at random within these stands.<br />

Presence <strong>of</strong> brooms respective cankers formed in consequence <strong>of</strong><br />

Melampsorella-infections in were counted separately for branches and trunks<br />

(Figure 2). Branch cankers which are closer than 10 cm to the trunk are expected to<br />

fuse with the stem in a few years and hence were regarded like stem cankers. The<br />

presence <strong>of</strong> caryophyllaceous host plants (mainly Stellaria nemorum) bearing the


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

dicaryotic phase <strong>of</strong> the rust fungus in the plots and in their neighbourhood were<br />

recorded. Silver fir can only be infected by the rusts basidiospores, exclusively<br />

originating from the alternate host.<br />

Table 1: List <strong>of</strong> monitoring plots(a.p.: advance planting; n.r. natural regeneration)<br />

Stand<br />

Numb<br />

er<br />

District Geology<br />

Forest<br />

Compartment<br />

184<br />

Regeneration<br />

type<br />

Altitude<br />

(m s l)<br />

N<br />

plots<br />

1 Schw.-Baar Braunjura XI_1 a.p. 808 4<br />

2 Schw.-Baar Keuper IV_2 a.p. 739 4<br />

3 Schw.-Baar Keuper IV_3 a.p. 737 4<br />

4 Br.-Hschw. Buntsandstein III_2 a.p. 1036 5<br />

5 Br.-Hschw. Muschelkalk XXV_6 a.p. 732 3<br />

6 Br.-Hschw. Muschelkalk XXV_7 a.p. 873 3<br />

7 Br.-Hschw. Muschelkalk XXV_9 a.p. 811 4<br />

8 Br.-Hschw. Muschelkalk XXVII_1 a.p. 722 3<br />

9 Schw.-Baar Braunjura X_6 n.r. 822 7<br />

10 Tuttlingen Braunjura Gew.Teilenw. n.r. 802 4<br />

11 Br.-Hschw. Granit/Gneis Urish<strong>of</strong> a.p. 1062 4<br />

12 Schw.-Baar Braunjura XI_1 a.p. 806 4<br />

13 Br.-Hschw. Granit/Gneis Urish<strong>of</strong> a.p 1050 4<br />

3. RESULTS<br />

Mean<br />

846<br />

Total<br />

53<br />

Incidence <strong>of</strong> broom rust in the stands (average <strong>of</strong> percentage infected firs per<br />

plot) stretched from 2.2 till 100 % (Table 2). In five stands, 100% were reached;<br />

the mean amount was 63.8%. The percentage <strong>of</strong> stem cankers was lower (in<br />

average by 35.5%) and decreased parallel with the general disease incidence. In<br />

seven stands, more than 50% <strong>of</strong> the trees bared stem cankers (mean 42.4 %).<br />

As alternate host <strong>of</strong> the rust fungus nearly exclusively S. nemorum was<br />

recorded. In most cases, it was found numerously at roadsides and in logging lines,<br />

in the vicinity <strong>of</strong> the plots rather than in the plots themselves. This plant species<br />

was found much more frequently in stands with higher disease incidence.<br />

Neither density <strong>of</strong> canopy and young stand nor geological factors nor age class<br />

<strong>of</strong> the plantings respective regenerations showed evidence to be related with the<br />

percentage <strong>of</strong> the diseased silver fir trees. Advance plantings seemed to be more<br />

infected than natural regenerations.


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Table 2: Characteristics and disease incidence <strong>of</strong> monitoring plots. Stands are given in<br />

the order <strong>of</strong> descending disease incidence. (I: seedlings; II: saplings, III: young pole stage)<br />

Stand<br />

Number<br />

n<br />

A. alba<br />

counted<br />

Density <strong>of</strong><br />

young<br />

stand<br />

Density <strong>of</strong><br />

canopy<br />

Age<br />

Class<br />

185<br />

Alternate<br />

host<br />

recorded<br />

Disease<br />

Incidence<br />

%*<br />

Stem<br />

cankers<br />

%*<br />

7 45 +++ 0 I ++ 100 100<br />

13 33 ++ 0 I ++ 100 93<br />

2 27 +++ 0 II + 100 76<br />

11 25 + 0 II ++++ 100 67<br />

1 14 ++ + II ++ 100 65<br />

12 15 ++ + I - II ++ 94 60<br />

5 7 ++ 0 III 0 83 28<br />

9 n.r. 127 + ++ I - II ++ 70 52<br />

3 46 +++ 0 I - II 0 39 2<br />

6 14 +++ 0 II + 22 8<br />

8 14 ++ 0 II 0 14 0<br />

4 16 ++ 0 II + 5 0<br />

10 n.r. 304 +++ +++ I + 2 0<br />

Total<br />

687<br />

4. DISCUSSION<br />

Mean<br />

63,8<br />

* mean <strong>of</strong> plots<br />

Mean<br />

42,4<br />

The results show the potential <strong>of</strong> the rust fungi to infect young plantations <strong>of</strong><br />

silver fir up to 100 % and in rare cases up to 100 % incidence <strong>of</strong> stem canker. Thus,<br />

M. caryophyllacearum may under certain circumstances still be a serious threat for<br />

high quality timber growth <strong>of</strong> A. alba under modern silviculture. However, a<br />

disease incidence up to 50% <strong>of</strong> the trees is mostly restricted to harmless branch<br />

cankers. To provoke a certain amount <strong>of</strong> stem cankers, the infection pressure must<br />

be higher. This occurs both in natural regeneration as well as in advance plantings,<br />

if a close association with an alternate caryophyllaceous host is given.<br />

Data shows that all age classes from seedling stage till pole stage can equally be<br />

infected. Roth (1955) found stem cankers between 1.5 and 15.2 m tree height<br />

(average 6.1 m) in Swiss fir forest. Heck (1894) found stem cankers between 0.5<br />

and 18.5 m (average 5.0 m) in SW-Germany. Both give evidence that even late<br />

pole stage firs can also be infected at the leaders or twigs in the upper crown.<br />

Nicolotti et al. (1994) and Solla and Camarero (2006) stated a positive<br />

correlation between disease incidence and shorter distance to rivers. This can be<br />

due to air humidity in sites or regions where humidity is a limiting factor for the<br />

infection process <strong>of</strong> the fungus. Furthermore, under German conditions, the main<br />

alternate host S. nemorum is rather linked with plant associations in sites close to<br />

ditches, rivers, or to other moist and nutrient rich sites (e.g. “Stellario-Alnetum”)


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

than to typical silver fir forest (Oberdorfer, 2001). Consequently, the distance to<br />

rivers may be identical with the source <strong>of</strong> spore dispersal from alternate host plants.<br />

Influencing the infection process by opening the canopy to achieve lower air<br />

moisture seems to be an option in Mediterranean countries rather than in Black<br />

Forest conditions where moisture is hardly a limiting factor.<br />

The silvicultural consequences from this study are considered as follows:<br />

a) Broom rust is a natural phenomenon in silver fir forests. Thus, in close to nature<br />

<strong>forestry</strong> some infections on branches must be taken into consideration.<br />

b) Stem canker incidence up to about 60% in the regeneration can be eliminated during<br />

successive regular thinnings. Premature elimination <strong>of</strong> infected trees (sanitation<br />

treatment) is not necessary, since infection from tree to tree does not take place.<br />

Furthermore, exaggerative opening <strong>of</strong> the canopy respective regeneration may promote<br />

damage by the aphid Dreyfusia nordmannianae (Schröter et al., <strong>2009</strong>).<br />

c) Expensive plantings <strong>of</strong> silver fir should not take place close to road sides, logging<br />

lines, or in other disturbed sites where alternate hosts are abundant. These sites should<br />

be planted with non-host trees like European beech or Norway spruce.<br />

d) Dense stands may prevent growth <strong>of</strong> herbaceous alternate hosts.<br />

e) Selective pruning <strong>of</strong> infected branches may be meaningful, before cankers are going<br />

to merge with stems. Too intensive pruning should be avoided since silver fir tends to<br />

grow epicormic shoots, which may give entrance to new broom rust infections close to<br />

the trunk.<br />

REFERENCES<br />

Heck, C.R., 1894. Der Weisstannenkrebs. Springer Berlin, 163 pp.<br />

Holdenrieder, O., 1994. Krankheiten der Tanne Abies spp.. Schweiz Beitr Dendrologie 43: 11-21.<br />

Nicolotti, G, Cellerino, G.P., Anslemi, N., 1994. Distribution and damage caused by Melampsorella<br />

cariophyllacearum in Italy. Proc. IUFRO Canker, Shoot and Foliage diseases. Vallombrosa 6.-<br />

11.6.1994: 289-291.<br />

Oberdorfer, E., 2001. Pflanzensoziologische Exkursionsflora für Deutschland und angrenzende<br />

Gebiete. Ulmer Verlag Stuttgart 8. Aufl. 1051 pp.<br />

Oliva, J., Colinas, C., 2007. Canopy openings may prevent fir broom rust (Melampsorella<br />

caryophyllacearum) infections. Eur J Forest Res 126: 507-512.<br />

Roth, C., 1955. Die Wachstumsgeschwindigkeit von Weißtannenkröpfen. Schweiz Z Forstwes 106:<br />

657-665.<br />

Schroeter, H., Delb, H., Metzler, B., <strong>2009</strong>. Waldschutzsituation 2008/<strong>2009</strong> in Baden-Wuerttemberg.<br />

Allgemeine Forstzeitschrift-Der Wald 64: 336-339.<br />

Solla, A. Camarero, J.J., 2006. Spatial patterns and environmental factors affecting the presence <strong>of</strong><br />

Melampsorella caryophyllacearum infections in an Abies alba forest in NE Spain. Forest<br />

Pathology 36: 165 – 175.<br />

Solla, A., Sanchez-Miranda, A., Camarero, J.C., 2006. Radial-growth and wood anatomical changes<br />

in Abies alba infected by Melampsorella caryophyllacearum: a dendroecological assessment <strong>of</strong><br />

fungal damage. Ann Forest Sci 63: 293-300<br />

186


Foliage Diseases <strong>of</strong> Hardwood<br />

187


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

188


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 189-193<br />

NON-NATIVE HOSTS AND CONTROL OF Rhytisma acerinum<br />

CAUSING TAR SPOT OF MAPLE.<br />

Tom HSIANG 1* , L.X. TIAN 1 , C. SOPHER 1<br />

Dept. Environmental Biology, Univ. Guelph, Guelph, Ontario, Canada.<br />

* thsiang@uoguelph.ca<br />

ABSTRACT<br />

Tar spot <strong>of</strong> maple is an increasingly common disease in eastern North America.<br />

Rhytisma acerinum, causing large tar spot, was apparently introduced from Europe, and<br />

causes the most extensive problems on introduced maple species such as Acer platanoides<br />

(Norway maple). However, this pathogen was recently confirmed via molecular methods on<br />

native North American species <strong>of</strong> maple, A. saccharum and A. negundo. This raises the<br />

possibility <strong>of</strong> pathogenic adaptation to native species that will result in more widespread<br />

epidemics. Ascospore production on Norway maple was observed to occur over a relatively<br />

short period annually (May 25 - June 22, 2006; May 25 - July 3, 2007; and May 21 - June<br />

13, 2008). The duration <strong>of</strong> spore dispersal was dependent on the frequency <strong>of</strong> rainfall from<br />

the start <strong>of</strong> dispersal. Field studies on fungicidal control <strong>of</strong> tar spot on Norway maple were<br />

conducted in summer 2008, using nine chemicals and a water control. All nine chemicals<br />

were found to be effective if applied between late May and early June. A single application<br />

was sufficient to control disease in summer 2008.<br />

Keywords: disease, fungi, Acer, fungicides<br />

1. INTRODUCTION<br />

Tar spot <strong>of</strong> maple is caused by species <strong>of</strong> the ascomycete genus Rhytisma, and<br />

has a worldwide distribution wherever maples are found. Tar spot has been<br />

increasing in abundance across Eastern North America in the last 15 years, with<br />

leaves <strong>of</strong> Norway maple (Acer platanoides) bearing multiple black spots. There has<br />

been relatively little research done on tar spot in North America. The only<br />

scientific reports have come from Connecticut (Waterman, 1941) and New York<br />

(Hudler et al., 1987; 1998). The most recent peer-reviewed research report is one<br />

from New York (Hudler et al., 1998), which found that the fungus Rhytisma<br />

acerinum is the cause <strong>of</strong> tar spot on Norway maple, both <strong>of</strong> which (host and<br />

pathogen) are immigrant species, while a native fungal species, R. americanum,<br />

occurs on the native red and silver maples (A. rubrum and A. saccharinum). This is<br />

probably the reason that a Norway maple may be heavily infected with tar spot<br />

while an adjacent red maple (A. rubrum) or silver maple (A. saccharinum) may<br />

have no spots. The work report here continues from the earlier report (Hsiang and<br />

Tian, 2008) which examined the spore dispersal and identity <strong>of</strong> the fungus on<br />

various maples in North America. The purpose <strong>of</strong> this work was to continue to<br />

examine the epidemiology <strong>of</strong> this disease, by gathering overwintered maple leaves<br />

from southern Ontario weekly from March through August 2007 and 2008, and<br />

189


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

inspecting the asci for the presence <strong>of</strong> filiform ascospores, which initiate infections.<br />

Another objective <strong>of</strong> this research was to confirm the genetic identity <strong>of</strong> the<br />

organism causing tar spot on a range <strong>of</strong> maples in Ontario. And the final objective<br />

was to look at fungicidal control <strong>of</strong> tar spot.<br />

2. METHODS<br />

2.1 Sporulation<br />

After snowmelt, overwintered leaves <strong>of</strong> Norway maple bearing tar spots caused<br />

by Rhytisma acerinum were collected from a copse <strong>of</strong> maples at the Guelph<br />

Turfgrass Institute, Guelph, Ontario every week from March through August in<br />

2007 and 2008. Diseased Norway maple leaves were soaked in distilled water for<br />

24 h to allow the apothecia to open, and many spots were examined, with several<br />

cross-sections per spot. The percent asci that were empty was estimated. Maple<br />

phenology and weather conditions were also recorded at each sampling.<br />

2.2 Genetic Identity<br />

Samples <strong>of</strong> tar spot from a variety <strong>of</strong> different maple species were collected<br />

from Ontario and Quebec, Canada in 2007 and 2008. We used the Qiagen<br />

DNAeasy kit (Qiagen Inc., Mississauga, Ontario, Canada), to extract DNA from<br />

these samples. This DNA was then amplified with conserved ITS primers which<br />

target the internal transcribed spacer region <strong>of</strong> ribosomal DNA spanning the 3' end<br />

<strong>of</strong> the 18S gene to the 5' end <strong>of</strong> the 28S gene. The primer pair, ITS1<br />

(TCCGTAGGTGAACCTGCGG) and ITS4 (TCCTCCGCTTATTGATATGC)<br />

were from White et al. (1991). The 12 5 ul reaction mixture for PCR amplification<br />

contained the following : 10 ng DNA, 1 DNA polymerase buffer, 0 5 ìm <strong>of</strong> each<br />

primer, and 1 U Tsg DNA polymerase (Biobasic, Scarborough, Ontario, Canada).<br />

Amplifications were performed in a GeneAmp PCR System 2400 (Perkin Elmer,<br />

Norwalk, CT, USA), with an initial denaturation step <strong>of</strong> 94 for 2 min, followed by<br />

35 cycles <strong>of</strong> 94 for 30 s, 55 for 1 min, and 72 for 2 min, and a final extension at 72<br />

for 7 min. These PCR reactions were sent for sequencing at the Laboratory<br />

Services Division, University <strong>of</strong> Guelph with both forward and reverse primers. At<br />

least two tar spot sequences from each maple species were used for analyses.<br />

2.3 Fungicidal Control<br />

The fungicide trials were conducted at the Guelph Turfgrass Institute, Guelph,<br />

Ontario, Canada on 2 m tall plants. These plants had been obtained from a local<br />

nursery as bareroot saplings over 2 m tall, and were planted in the local Fox Sandy<br />

Loam soil in early May 2007. The trees were placed in four rows adjacent to a<br />

older stand <strong>of</strong> Norway maple trees. There was 1.5 m between the rows and 1.2 m<br />

between the trees. Each row was considered a block and consisted <strong>of</strong> 10 trees.<br />

Treatments were applied to each tree at two-week intervals but to different<br />

branches or twigs on each <strong>of</strong> five dates in 2008: May 6, May 20, June 4, June 17,<br />

190


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

June 30. On each date, the number <strong>of</strong> leaves per twig was assessed, and twigs were<br />

sprayed until run<strong>of</strong>f with each treatment (up to 5 mL per twig, but depending on<br />

the number <strong>of</strong> leaves and leaf sizes). A plastic board was held behind each twig<br />

while spraying to prevent overspray onto other twigs. There were nine chemical<br />

treatments and a water control (Table 1).<br />

Table 1: Treatments applied to 2 m tall maple trees at the Guelph Turfgrass Institute,<br />

Guelph, Ontario, Canada, in Spring 2008 for control <strong>of</strong> tar spot.<br />

Treatment Common Name Product/ L<br />

Water control water 0 ml<br />

Banner MAXX propiconazole (15.6%) 0.245 mL<br />

Compass 50WG trifloxystrobin (0.16%) 0.175 g<br />

Daconil Ultrex chlorothalonil (82.5%) 1.5 g<br />

Dithane DG mancozeb (75%) 3 g<br />

Eagle WP (Nova) myclobutanil (40%) 0.34 g<br />

Heritage WG azoxystrobin (50%) 0.3 g<br />

Rovral Green GT iprodione (25%) 10 mL<br />

Senator WP thiophanate-methyl (70%) 1 g<br />

Sulfur sulfur (92%) 10 g<br />

The number <strong>of</strong> spots per leaf was assessed weekly from the start <strong>of</strong> the trial<br />

until the end <strong>of</strong> June, and then assessed monthly until the end <strong>of</strong> September. The<br />

morphology and phenological state <strong>of</strong> the maple leaves was also recorded during<br />

each assessment.<br />

3. RESULTS AND DISCUSSION<br />

As observed in previous years (Hsiang & Tian 2008), the first symptoms <strong>of</strong> tar<br />

spot on Norway maple in Southern Ontario appeared in late June as small, round,<br />

light green, chlorotic spots, 2 mm across. Spots enlarged to 15 mm by mid-August,<br />

and developed small black tar-like raised structures on the adaxial surface with a<br />

yellow margin. Conidia, which are considered non-infective and possibly<br />

spermatizing, appeared as a shiny layer on the black stroma at this time. By early<br />

September, the individual spots merged into a circular black spot up to 2 cm across.<br />

Overwintered Norway maple leaves collected in March 2007 and 2008, had<br />

stroma, paraphyses and asci (56-80 µm × 8.5-10.6 µm), but no ascospores were<br />

visible. By the middle <strong>of</strong> April, the asci were still undifferentiated, but were found<br />

to contain globular vacuoles or bodies. The asci became swollen as spores<br />

developed, and filiform ascospores were first observed in early May, averaging 55<br />

× 2.0 µm. By late May, after soaking in water, slits in the hysterothecia (modified<br />

191


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

apothecia) on the leaf surface opened, and contained a grayish milky substance. At<br />

this time, Norway maples were abundantly producing and shedding pollen, and<br />

small samaras were formed, with leaf sizes averaging 10 cm × 15 cm. By late May<br />

in 2007 and 2008, a few partially filled or empty asci were observed, with<br />

ascospore release through the tips, and paraphyses becoming curled beside asci<br />

after spore release. In early June, Norway maple leaves reached their full size (20<br />

cm × 24 cm), and 10% (2007) to 30% (2008) <strong>of</strong> the asci had fully discharged their<br />

spores. By the beginning <strong>of</strong> July 2008, nearly all the asci were empty and this<br />

sporulation period was longer than that observed in 2006 (Hsiang and Tian, 2008)<br />

because <strong>of</strong> much drier conditions. However in 2008, the sporulation period was<br />

shortened with full spore release by mid June because <strong>of</strong> much wetter conditions.<br />

In 2007, we found a few tar spots on trees tentatively identified as sugar maple<br />

(A. saccharum). We confirmed the identity <strong>of</strong> these maple trees based on DNA<br />

sequencing <strong>of</strong> a chloroplast gene. The fungal DNA was also sequenced, and it<br />

turned out to match R. acerinum, the European species. This was very surprising<br />

since the European fungal species should not occur on a native North American<br />

maple species. The same trees were visited again in 2008, and samples were<br />

collected during the growing season. These also yielded DNA confirmed as<br />

belonging to A. acerinum. In 2008, we also collected specimens <strong>of</strong> A. campestre<br />

and A. negundo from Ontario, and other samples <strong>of</strong> A. saccharum from Quebec.<br />

All specimens were infected with tar spot caused by R. acerinum as identified by<br />

sequences <strong>of</strong> the ITS region. This result was not unexpected for A. campestre since<br />

both the host and pathogen are European species. However, this was another<br />

unexpected result for A negundo, and A. saccharum. The first species is considered<br />

a weed, but the second species is the major source <strong>of</strong> maple syrup. The occurrence<br />

<strong>of</strong> a European tar spot species on North American maple species raises the<br />

possibility <strong>of</strong> pathogenic adaptation to native species that could result in epidemics<br />

more widespread than currently seen on Norway maples.<br />

In Hsiang & Tian (2008), we predicted that based on spore production periods <strong>of</strong> R.<br />

acerinum, "the practical implication is that fungicide protection against tar spot, if<br />

necessary, needs only to be applied during a very short period, which begins near the<br />

end <strong>of</strong> full leaf expansion in Norway maple". We tested this hypothesis in spring 2008<br />

with application <strong>of</strong> fungicides at different times. We found that a single fungicide<br />

application in late May or early June was efficacious in reducing the number <strong>of</strong> spots<br />

per leaf assessed on September 2, 2008, from over 20 to none for many <strong>of</strong> the<br />

fungicide tested. All fungicides were found to significantly suppress tar spot compared<br />

to the water control when applied at either <strong>of</strong> these two times (May 20 or June 4).<br />

Applications <strong>of</strong> fungicides on May 6 or June 30 were not effective in reducing tar spot,<br />

while the application on June 17, which was just at the end <strong>of</strong> the spore production<br />

period, was effective for some fungicides but not others. The implication <strong>of</strong> these<br />

results is that if fungicides are used, only a single application between late May and<br />

early June is needed for control <strong>of</strong> tar spot <strong>of</strong> Norway maple in Southern Ontario, but<br />

this may depend on weather and growth, with another application possibly necessary if<br />

early June weather is dry.<br />

192


4. REFERENCES<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Hsiang T and Tian XL, 2008. Sporulation and identity <strong>of</strong> tar spot <strong>of</strong> maple in Canada. Acta Silvatica<br />

& Lignaria Hungarica 2007, 71-74.<br />

Hudler GW, Jensen-Tracy S, Banik MT. 1998. Rhytisma americanum sp. nov.: a previously<br />

undescribed species <strong>of</strong> Rhytisma on maples (Acer spp.). Mycotaxon 68, 405-416.<br />

Waterman AM. 1941. Diseases <strong>of</strong> shade and ornamental trees: annotated list <strong>of</strong> specimens received in<br />

1940 at the New Haven Office, Division <strong>of</strong> Forest Pathology. Plant Dis. Rep. 25, 181-<br />

182.<br />

White TJ, Bruns T, Lee S, Taylor J. 1990. Amplification and direct sequencing <strong>of</strong> fungal ribosomal<br />

RNA genes for phylogenetics. In: Innis MA, Gelfand DA, Sninsky JJ, White TJ. (Eds.),<br />

PCR Protocols: a guide to methods and applications. Academic Press, San Diego, U.S.A.,<br />

pp. 315 - 322.<br />

5. ACKNOWLEDGEMENTS<br />

This study was funded by the Ontario Ministry <strong>of</strong> Agriculture, Landscape<br />

Ontario, and the National Sciences and Engineering Research Council <strong>of</strong> Canada.<br />

Trees were donated by Winkelmolen Nursery, Lynden, Ontario, Canada.<br />

193


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 194-199<br />

BIOLOGICAL CONTROL TRIALS OF BEECH BARK DISEASE<br />

UNDER LABORATORY CONDITIONS<br />

Gaston LAFLAMME 1* , Simon BOUDREAULT 1 , Robert LAVALLEE 1 ,<br />

Martine BLAIS 1 , Jean-Yves BLANCHETTE 2 .<br />

1 Natural Resources Canada, CFS, Laurentian Forestry Centre, Québec, QC, Canada G1V 4C7<br />

2 Université de Moncton, Edmundston, N.-B., Canada E3V 2S8<br />

* Gaston.Laflamme@NRCan-RNCan.gc.ca<br />

ABSTRACT<br />

Beech bark disease (BBD) causes mortality <strong>of</strong> American beech (Fagus grandifolia<br />

Ehrh.). BBD involves an attack by the beech scale insect Cryptococcus fagisuga Lind.<br />

followed by the native fungal pathogen Neonectria faginata (Lohman et al.) Cast. &<br />

Rossman. C. fagisuga was introduced into Halifax, Nova Scotia, from Europe through<br />

seedlings around 1890. Damage to American beech was observed 20 years later. Our<br />

objective is to use entomogenous fungi to control the insect. Lecanicillium muscarium<br />

(Petch) Zare & W. Gams, common in European infested sites, was retained as well as<br />

Beauveria bassiana (Bals.-Criv.) Vuill. Our first trials were done on non-crawling nymphal<br />

stage on bark disks, 24 mm in diameter, kept individually in Solo cups ® at 20°C or 25°C.<br />

To expose the insects, the “wool-like” wax covering the colony was removed. The<br />

treatment consisted <strong>of</strong> an application <strong>of</strong> 125 µL <strong>of</strong> 10 6 spores/mL <strong>of</strong> water and oil. A<br />

second trial was conducted by spraying spore suspensions <strong>of</strong> L. muscarium (100 µL) on<br />

eggs kept at 25°C. Both biological control agents reduced the crawlers’ population by 50%<br />

after 11 days. Eggs treated with L. muscarium showed low mortality, but their<br />

development was slowed down. Fungi seen on the surface <strong>of</strong> the eggs invaded the first<br />

instars. Field trials are underway.<br />

Keywords: Beech bark disease, Cryptococcus fagisuga, Neonectria faginata,<br />

Lecanicillium muscarium, entomogenous fungi, Beauveria bassiana<br />

1. INTRODUCTION<br />

American beech (Fagus grandifolia Ehrh.) forests suffer significant mortality<br />

caused by beech bark disease (BBD); this complex disease has a permanent<br />

negative impact on the forest ecosystem in North America. A symposium on this<br />

important disease was held in New York State to summarize the status <strong>of</strong> our<br />

knowledge on BBD, and to identify the most important knowledge gaps <strong>of</strong> this<br />

phenomenon (Evans et al., 2005). Rapidly spreading across eastern North<br />

America, BBD involves a preliminary attack by the beech scale insect<br />

Cryptococcus fagisuga Lind. and fungal species <strong>of</strong> the genus Nectria (Ehrlich,<br />

1934), now known under the genus Neonectria (Figure 1). The insect C. fagisuga<br />

was introduced into Canada from Europe, around 1890, through European beech<br />

(F. sylvatica L.) seedlings planted as ornamentals in Halifax, Nova Scotia, Canada.<br />

194


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Damage to our native American beech was first observed 20 years later. The<br />

pathogen was thought to be the exotic fungus Neonectria coccinea from Europe,<br />

but recent taxonomic studies showed that the new species Neonectria faginata<br />

(=Nectria coccinea var. faginata) is the pathogen <strong>of</strong> this BBD complex in North<br />

America (Castlebury et al., 2006). Also, N. ditissima (=Nectria galligena), a<br />

pathogen causing cankers on Acer spp. and Betula spp. in North America, is <strong>of</strong>ten<br />

involved in the BBD (Houston, 1994).<br />

A B<br />

Figure 1: A- The beech scale insect Cryptococcus fagisuga colonizing the bark <strong>of</strong> the<br />

American beech (Fagus grandifolia). B- Multiple small cankers caused by Neonectria<br />

faginata.<br />

The control <strong>of</strong> BBD is not an easy task. Different approaches have been<br />

considered. The selection and propagation <strong>of</strong> resistant beech is one <strong>of</strong> them (Koch<br />

and Carey, 2005; Loo et al., 2005). Using silviculture to improve the health status<br />

<strong>of</strong> beech populations in newly infested stands (Heyd, 2005) or aftermath forests<br />

(Ostr<strong>of</strong>sky, 2005) is a second approach. Biological control <strong>of</strong> the scale insect and<br />

fungal pathogens is a possibility that has been suggested in the past, but so far only<br />

preliminary trials have been reported (Lonsdale, 1983).<br />

In our project, we are testing different entomogenous fungi and hyperparasite<br />

fungi to control the insects and the pathogens involved in the BBD complex. The<br />

objective <strong>of</strong> this study is to use entomogenous fungi to reduce the population <strong>of</strong> C.<br />

fagisuga under laboratory conditions.<br />

2. MATERIALS AND METHODS<br />

Since the entomogenous fungus Lecanicillium muscarium (= Verticillium<br />

lecanii Vegas) was observed to be common in heavily infested sites in England<br />

(Lonsdale, 1983), it was retained in our first experiment on the non-crawling<br />

nymphal stage. This instar is a crawling immature but has its stylet fixed in the<br />

bark and produces the typical “wool-like” wax. We are using the isolate L.<br />

195


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

muscarium (Mycotal®). Another fungus, Beauveria bassiana (Bals.-Criv.) Vuill.<br />

currently being tested against other forest insect pests in our laboratory, has been<br />

added to part <strong>of</strong> this trial.<br />

The first experiment was done on 24 mm diameter bark disks colonized by C.<br />

fagisuga collected on different trees in the fall and winter. To expose the insect<br />

nymphs to the fungus, the white “wool-like” wax was gently removed with fine<br />

tools to facilitate observations. The treatment consisted <strong>of</strong> an application <strong>of</strong> 125 µL<br />

<strong>of</strong> 10 6 spores/mL <strong>of</strong> water (added with oil (0.5%) and Tween 80 (0.0025%)). Bark<br />

disks were kept individually in Solo cups at 20°C (65% HR) and/or 25°C (85%<br />

HR) and photoperiod <strong>of</strong> 16D:8L. A control was also used with the same<br />

formulation but without the spores. Observations were done regularly over a<br />

period <strong>of</strong> two weeks. Table 1 presents the total number <strong>of</strong> live scale insects<br />

observed on the bark disks in each treatment on samples collected in fall and<br />

winter.<br />

Table 1: Total number <strong>of</strong> scale insects used according to treatments and sampling<br />

seasons.<br />

Treatment Sampling season<br />

196<br />

Number <strong>of</strong> bark<br />

discs<br />

Total number <strong>of</strong> live<br />

non-crawling<br />

nymphal stage on<br />

bark discs<br />

Control-Oil Fall 15 88<br />

ycotal-Oil Fall 11 119<br />

Beauveria-Oil Fall 7 84<br />

Control-Oil Winter 12 70<br />

Mycotal-Oil Winter 17 99<br />

The second experiment was run using a spore suspension <strong>of</strong> L. muscarium (100<br />

µL <strong>of</strong> 10 6 spores/mL) with the same formulation as the first experiment, sprayed<br />

on eggs (61 eggs for the control and 74 for the treatment) and kept at 25°C (85%<br />

HR and photoperiod <strong>of</strong> 16D:8L). Observations were done regularly over a period<br />

<strong>of</strong> three weeks.<br />

3. RESULTS AND DISCUSSION<br />

Both biological control agents reduced the non-crawling nymphal stage<br />

population by 50% after 10-11 days, but L. muscarium (Mycotal ®) showed<br />

slightly better results (Figure 2). Mortality <strong>of</strong> each female was based on apparent<br />

drying symptoms associated with cuticle darkening and hyphal growth on the<br />

insect body.


Cumulative mortality (%)<br />

100<br />

50<br />

0<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

0 2 4 6 8 10 12<br />

Days<br />

197<br />

Control-OIL-25-fall<br />

Mycotal-OIL-25-fall<br />

Beauveria-OIL-25-fall<br />

Figure 2: Cumulative mortality <strong>of</strong> Cryptococcus fagisuga exposed to Lecanicillium<br />

muscarium (Mycotal®), Beauveria bassiana and control over a period <strong>of</strong> 11 days.<br />

A higher mortality level <strong>of</strong> scale insects treated with L. muscarium was<br />

observed on samples collected in winter, compared with the fall samples. The<br />

winter mortality may have been caused by very cold temperatures (-35°C)<br />

measured in February (Environment Canada). Also, the measured mortality rate<br />

with L. muscarium was increasing much faster, reaching 50% in only 7 days<br />

compared with 10 days in the previous fall tests (Figure 3). Our observations are in<br />

agreement with Crosby and Bjorkbom (1958), who consider that severe winter<br />

temperatures <strong>of</strong> -35 o F (-37 o C) or lower are lethal to beech scale insects.<br />

Eggs treated with L. muscarium showed very low mortality, but the treatment<br />

slowed down their development (Figure 4). Fifty percent hatching was reached<br />

after 8 and 12 days respectively for the L. muscarium treatment and control. The<br />

fungus was observed to colonize only the external surface <strong>of</strong> the egg chorion.<br />

However, when they hatch, first instars become rapidly infected by the fungus<br />

already present on the egg chorion.<br />

Cumulative mortality (%)<br />

100<br />

50<br />

0<br />

0 2 4 6 8 10 12<br />

Days<br />

Control-OIL-25-fall<br />

Mycotal-OIL-25-fall<br />

Control-OIL-25-winter<br />

Mycotal-OIL-25-winter<br />

Figure 3: Mortality <strong>of</strong> scale insects over a period <strong>of</strong> 11 days from samples collected in<br />

the fall and winter, after treatment with Lecanicillium muscarium (Mycotal®) and control.


Eggs hatching (%)<br />

100<br />

50<br />

0<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

0 2 4 6 8 10 12 14 16 18 20 22<br />

198<br />

days<br />

Control-25<br />

Mycotal-25<br />

Figure 4: Percentage <strong>of</strong> Cryptococcus fagisuga’s eggs hatching after treatment with<br />

Lecanicillium muscarium (Mycotal®) over a 21-day period.<br />

4. CONCLUSION<br />

The scale insect Cryptococcus fagisuga, exposed to Lecanicillium muscarium<br />

(Mycotal®) and Beauveria bassiana under laboratory conditions, reduced the noncrawling<br />

nymphal stage population by 50%. This percentage could have been<br />

higher, but the experiment on bark disks cannot last longer than 10 to 12 days;<br />

because <strong>of</strong> the high rate <strong>of</strong> humidity, fungal hyphae invade the disks making<br />

further observations impossible. The fall population treated with entomogenous<br />

fungi had a mortality rate reaching 50% in 10 days. Scale insects collected in<br />

winter show a higher rate <strong>of</strong> mortality after biological treatment; the difference<br />

with fall is probably the mortality caused by the very cold temperatures during the<br />

winter. Finally, if eggs are not directly invaded by fungal hyphae, the young<br />

crawlers are rapidly colonized by the fungi soon after hatching (data not shown).<br />

Field trials with these biological control agents will be the next step <strong>of</strong> this study.<br />

5. LITERATURE CITED<br />

Castlebury, L.A., Rossman A.Y., Hyten A.S., 2006. Phylogenetic relationships <strong>of</strong><br />

Neonectria/Cylindrocarpon on Fagus in North America. Can. J. Bot. 84, 1417-1433.<br />

Crosby, D., Bjorkbom J.C., 1958. Timely salvage can reduce losses from beech scale Nectria attack.<br />

U.S. Forest Serv., Northeastern Forest Experiment Station, Forest Research Note 82, 4p.<br />

Ehrlich, J., 1934. The beech bark disease. A Nectria disease <strong>of</strong> Fagus, following Cryptoccus fagi<br />

(Baer.). Can. J. Res. 10, 593-692.<br />

Evans, C.A., Lucas J.A., Twery M.J. (Eds.), 2005. Beech Bark Disease: Proceedings <strong>of</strong> the Beech<br />

Bark Disease Symposium. General Technical Report NE-331. Newtown Square PA, US<br />

Department <strong>of</strong> Agriculture Forest Service, Northern Research Station, 149 p.


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Heyd, R.L., 2005. Managing Beech Bark Disease in Michigan In: Evans, C.A., Lucas J.A., Twery<br />

M.J. (Eds.). Beech Bark Disease: Proceedings <strong>of</strong> the Beech Bark Disease Symposium.<br />

General Technical Report NE-331. Newtown Square PA, US Department <strong>of</strong> Agriculture<br />

Forest Service, Northern Research Station, pp. 128-132.<br />

Houston, D.R., 1994. Temporal and spatial shift within the Nectria pathogen complex associated with<br />

beech bark disease <strong>of</strong> Fagus grandifolia. Can. J. For. Res. 24, 960-968.<br />

Koch, J.L., Carey D.W., 2005. The Genetics <strong>of</strong> Resistance <strong>of</strong> American Beech to Beech Bark<br />

Disease: Knowledge through 2004. In: Evans, C.A., Lucas J.A., Twery M.J. (Eds.), 2005.<br />

Beech Bark Disease: Proceedings <strong>of</strong> the Beech Bark Disease Symposium. General<br />

Technical Report NE-331. Newtown Square PA, US Department <strong>of</strong> Agriculture Forest<br />

Service, Northern Research Station, pp. 98 – 105.<br />

Lonsdale, D., 1983. Fungal associations in the buildup and decline <strong>of</strong> Cryptococcus fagisuga<br />

populations. In: Proceedings, IUFRO Beech Bark Disease Working Party Conference,<br />

September 26 - October 8, 1982, Hamden, Connecticut. General Technical Report WO-<br />

37. Washington, DC: U.S. Department <strong>of</strong> Agriculture, Forest Service. pp. 99-104.<br />

Loo, J., Ramirez, M., Krasowski M., 2005. American Beech Vegetative Propagation and Genetic<br />

Diversity. In: Evans, C.A., Lucas J.A., Twery M.J. (Eds.), 2005. Beech Bark Disease:<br />

Proceedings <strong>of</strong> the Beech Bark Disease Symposium. General Technical Report NE-331.<br />

Newtown Square PA, US Department <strong>of</strong> Agriculture Forest Service, Northern Research<br />

Station, pp. 106 – 112.<br />

Ostr<strong>of</strong>sky 2005. Management <strong>of</strong> beech bark disease in aftermath forests. In Evans, C.A., Lucas J.A.,<br />

Twery M.J. (Eds.), 2005. Beech Bark Disease: Proceedings <strong>of</strong> the Beech Bark Disease<br />

Symposium. General Technical Report NE-331. Newtown Square PA, US Department <strong>of</strong><br />

Agriculture Forest Service, Northern Research Station, pp. 133-137.<br />

199


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 200-205<br />

PATHOGENICITY OF Fusarium circinatum NIREMBERG &<br />

O’DONNELL ON SEEDS AND SEEDLINGS OF RADIATA PINE<br />

Pablo Martínez-ÁLVAREZ 1 , Juan BLANCO 2 , Milagros de VALLEJO 2 ,<br />

Fernando M. Alves-SANTOS 1 , Julio J. DIEZ 1*<br />

1 Departamento de Producción Vegetal y Recursos Forestales. ETSIIAA Palencia. Universidad de<br />

Valladolid. Avenida Madrid, 57. 34004 Palencia. España.<br />

2 Sección de Producción y Mejora Forestal Servicio de Montes Dirección General de Biodiversidad.<br />

Gobierno de Cantabria. Calle Rodríguez, nº 5, 1º. 39.071 Santander<br />

ABSTRACT<br />

*jdcasero@pvs.uva.es<br />

Pathogenicity <strong>of</strong> seven Fusarium circinatum isolates from Northern Spain was<br />

evaluated on Monterey pine (Pinus radiata) seeds and seedlings. The objectives <strong>of</strong> our<br />

study were also to investigate emergence and post-emergence damping-<strong>of</strong>f damage, and to<br />

observe differences in pathogenicity among the isolates.<br />

The effect <strong>of</strong> F. circinatum on seed emergence was approximately 10 to 19% lower than<br />

the control treatment. However, all F. circinatum isolates severely affected pine seedlings,<br />

causing 63% to 90% mortality <strong>of</strong> plants 30-days post inoculation. Sixty days after<br />

inoculation, isolate FcCa7 killed all the seedlings, while the less aggressive FcCa2 affected<br />

79% <strong>of</strong> the plants. We believe this homogeneity in aggressiveness among Fusarium isolates<br />

may possibly be attributed to the recent introduction <strong>of</strong> the pathogen in this region.<br />

Key words. Pitch Canker, Biological control, Pinus radiata, Nurseries, Plant health<br />

care<br />

1. INTRODUCTION<br />

Fusarium circinatum is a pathogenic fungi with great virulence in species <strong>of</strong> the<br />

genus Pinus, causing a disease called pitch canker. It was first discovered as a<br />

pathogen in California during 1986 (Mccain et al., 1987). Since then, F. circinatum<br />

was also found in Mexico (Rodriguez, 1989), South Africa (Viljoen et al., 1994;<br />

Nirenberg and O'Donnell, 1998; Crous et al., 2000; Steenkamp et al., 2002;<br />

Coutinho et al., 2007), Japan (Aoki et al., 2001; Kobayashi, 2007), Chile<br />

200


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

(Wingfield et al., 2002; Jacobs et al., 2007), Korea (Cho and Shin, 2004), Italy<br />

(Carlucci et al., 2007) and Spain (Landeras et al., 2005; Perez-Sierra et al., 2007).<br />

One <strong>of</strong> the most important actions to control the disease is to have a better<br />

understanding about the populations <strong>of</strong> the fungus, and it behaviour over plants<br />

host. Thus, the aim <strong>of</strong> this work is to investigate the effect <strong>of</strong> F. circinatum over<br />

seeds and seedlings <strong>of</strong> Pinus radiata, and to observe the differences in<br />

pathogenicity between isolates.<br />

2. MATERIAL AND METHODS<br />

2.1. Fungal material<br />

Seven different isolates <strong>of</strong> Fusarium circinatum obtained from Pinus radiata<br />

plantations <strong>of</strong> the Autonomous Community <strong>of</strong> Cantabria, in the northern Spain,<br />

were used for this study. Malt extract media (20 g/l) was prepared to achieve the<br />

spore dissolution <strong>of</strong> the fungus used in the inoculation (50 ml <strong>of</strong> autoclaved media<br />

in a Erlenmeyer flask). Once esterilized, four pieces <strong>of</strong> fungal mycelium grown in<br />

PDA-S (potato-dextrose-agar with 0.5 g/l <strong>of</strong> streptomycin sulfate) were placed<br />

inside the flasks. Production <strong>of</strong> spores was induced by using an orbital shaker.<br />

After that, the media was filtered in order to collect only spores in the dissolution.<br />

In order to obtain the fitted concentration (10 6 spores/ml), we used a Thoma<br />

counting chamber.<br />

2.2. Plant material<br />

A total <strong>of</strong> 672 seeds were sown to observe the effect <strong>of</strong> the fungus over the plant<br />

material. Before sowing them, seeds were washed with sterile distilled water<br />

repeatedly and kept there for twelve hours. After that, seeds were maintained in<br />

hydrogen peroxide (3%) for 30 minutes. Finally they were washed twice with<br />

sterile distilled water to remove the remaining hydrogen peroxide permeating the<br />

seeds.<br />

2.3. Substrate. A mixture <strong>of</strong> peat and vermiculite at 50% were used for the<br />

experiment. Before filling the nursery trays, the substrates were autoclaved twice<br />

during one hour at 120 ºC.<br />

2.4. Seeds sowing<br />

Seeds were grown in nursery trays, placing four seeds in each hole. Twenty one<br />

holes were used for each isolate. After sowing, trays were covered with a<br />

transparent plastic paper to prevent the aerial contaminations. The assay was<br />

developed in controlled conditions <strong>of</strong> temperature (20 ºC) and photoperiod (16/8)<br />

inside a growth chamber (Figure 1). Seedlings were watered once a week, with<br />

twenty millilitres <strong>of</strong> sterile distilled water, and checked for the progress <strong>of</strong> the<br />

assay.<br />

201


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Figure 1: Growth chamber with the trays inside.<br />

2.5. Data capture<br />

Seed germination was measured once a week. Futhermore dead seedlings were<br />

counted ten weeks after the sowing. At the end <strong>of</strong> the experiment, attempts to<br />

reisolate the pathogen from the seedlings were made to verify it presence in the<br />

necrotic lesions.<br />

2.6. Statistical analysis<br />

A Kruskal-Wallis test was performed with Statgraphics Plus 5.1. to find<br />

differences between germination and mortality taxes <strong>of</strong> the different isolates.<br />

3. RESULTS AND DISCUSION<br />

3.1. Germination<br />

Despite genus Fusarium is considered as one <strong>of</strong> the most important causes <strong>of</strong><br />

pre-emergence and post-emergence damping-<strong>of</strong>f (Machon et al., 2006; Pinto et al.,<br />

2006), in our present assay fungus did not cause great damages over seeds but,<br />

decreasing slightly germination rates. As it can be observed in the Figure 2,<br />

CONTROL was the treatment where seeds were more germinated. There were<br />

significant differences between this and the others treatments in which F.<br />

circinatum was present. On the other hand, no differences were observed among<br />

the seven isolates <strong>of</strong> the fungus used in the assay.<br />

202


Germination (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

07/01/<strong>2009</strong><br />

14/01/<strong>2009</strong><br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

21/01/<strong>2009</strong><br />

28/01/<strong>2009</strong><br />

04/02/<strong>2009</strong><br />

203<br />

11/02/<strong>2009</strong><br />

FcCa 1 FcCa 2<br />

FcCa 3 FcCa 4<br />

FcCa 5 FcCa 6<br />

FcCa 7 CONTROL<br />

18/02/<strong>2009</strong><br />

25/02/<strong>2009</strong><br />

04/03/<strong>2009</strong><br />

Figure 2: Rates <strong>of</strong> germination <strong>of</strong> the seeds depending on the isolate <strong>of</strong> Fusarium<br />

circinatum used<br />

3.2. Virulence<br />

Living plants (% )<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

a<br />

b<br />

b<br />

b<br />

FcCa 0 FcCa 1 FcCa 2 FcCa 3 FcCa 4 FcCa 5 FcCa 6 FcCa 7<br />

Figure 3: Percentage <strong>of</strong> living plants ten weeks after the inoculation according to the<br />

treatment used<br />

b<br />

b<br />

b<br />

b


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

All isolates <strong>of</strong> F. circinatum were highly virulent and significant differences among<br />

them were not observed. Aegerter and Gordon (2006) obtained the rates <strong>of</strong> seedling<br />

mortality ranging from 3.5 to 52%, however in our investigation most <strong>of</strong> the seedlings died<br />

after ten weeks <strong>of</strong> the inoculation. On the other hand, 70% <strong>of</strong> seedlings were still alive in<br />

case <strong>of</strong> the treatment where the fungus was not present (p-value < 0.001).<br />

4. CONCLUSIONS<br />

1. The effect <strong>of</strong> Fusarium circinatum over seed germination was small,<br />

reducing the rate <strong>of</strong> germinated seeds in between 10 and 20%. No differences were<br />

seen in germination among the seven isolates used.<br />

2. Mortality <strong>of</strong> Pinus radiata seedlings inoculated with the Fusarium circinatum<br />

was high for the seven isolates. Four isolates killed all the P. radiata seedlings.<br />

3. No big differences were found in virulence among the seven isolates.<br />

5. ACKNOWLEDGEMENTS<br />

We thank the Government <strong>of</strong> the Autonomous Community <strong>of</strong> Cantabria, the<br />

Ministry <strong>of</strong> Agriculture, and the Ministry <strong>of</strong> Environment (INIA-CIFOR) for<br />

financial support. We also want to thank Mohammed Masum Ul Haque for the<br />

revision <strong>of</strong> the article.<br />

REFERENCES<br />

Aegerter, B.J., Gordon, T.R., 2006. Rates <strong>of</strong> pitch canker induced seedling mortality among Pinus<br />

radiata families varying in levels <strong>of</strong> genetic resistance to Gibberella circinata (anamorph<br />

Fusarium circinatum). Forest Ecology and Management 235, 14-17.<br />

Aoki, T., O'Donnell, K., Ichikawa, K., 2001. Fusarium fractiflexum sp. nov. and two other species<br />

within the Gibberella fujikuroi species complex recently discovered in Japan that form<br />

aerial conidia in false heads. Mycoscience 42, 461-478.<br />

Carlucci, A., Colatruglio, L., Frisullo, S., 2007. First report <strong>of</strong> pitch canker caused by Fusarium<br />

circinatum on Pinus halepensis and P. pinea in Apulia (Southern Italy). Plant Disease 91,<br />

1683-1683.<br />

Cho, W.D., Shin, H.D., 2004. List <strong>of</strong> plant diseases in Korea., Fourth <strong>edition</strong> ed.<br />

Coutinho, T.A., Steenkamp, E.T., Mongwaketsi, K., Wilmot, M., Wingfield, M.J., 2007. First<br />

outbreak <strong>of</strong> pitch canker in a South African pine plantation. Australasian Plant Pathology<br />

36, 256-261.<br />

Crous, P.W., Phillips, A.J.L., Baxter, A.P., 2000. Phytopathogenic Fungi from South Africa.<br />

University <strong>of</strong> Stellenbosch, Department <strong>of</strong> Plant Pathology Press.<br />

Jacobs, A., Coutinho, T.A., Wingfield, M.J., Ahumada, R., Wingfield, B.D., 2007. Characterization<br />

<strong>of</strong> the pitch canker fungus, Fusarium circinatum, from Chile. South African Journal <strong>of</strong><br />

Science 103, 253-257.<br />

Kobayashi, T., 2007. Index <strong>of</strong> fungi inhabiting woody plants in Japan. Host, Distribution and<br />

Literature. Zenkoku-Noson-Kyoiku Kyokai Publishing Co., Ltd.<br />

Landeras, E., García, P., Fernández, Y., Braña, M., Fernández-Alonso, O., Méndez-Lodos, S., Pérez-<br />

Sierra, A., León, M., Abad-Campos, P., Berbegal, M., Beltrán, R., García-Jiménez, J.,<br />

Armengol, J., 2005. Outbreak <strong>of</strong> pitch canker caused by Fusarium circinatum on Pinus<br />

spp. in Northern Spain. Plant Disease 89, 1015-1015.<br />

204


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Machón, P., Santamaría, O., Pajares, J.A., Alves-Santos, F.M., Diez, J.J., 2006. Influence <strong>of</strong> the<br />

ectomycorrhizal fungus Laccaria laccata on pre-emergence, post-emergence and late<br />

damping-<strong>of</strong>f by Fusarium moniliforme and F.oxysporum on Scots pine seedlings.<br />

Symbiosis 42, 153-160.<br />

Mccain, A.H., Koehler, C.S., Tjosvold, S.A., 1987. Pitch canker threatens Californian pines.<br />

Californian Agriculture 41, 22-23.<br />

Nirenberg, H.I., O'Donnell, K., 1998. New Fusarium species and combinations within the Gibberella<br />

fujikuroi species complex. Mycologia 90, 434-458.<br />

Peréz-Sierra, A., Landeras, E., León, M., Berbegal, M., García-Jiménez, J., Armengol, J., 2007.<br />

Characterization <strong>of</strong> Fusarium circinatum from Pinus spp. in northern Spain. Mycological<br />

Research 111, 832-839.<br />

Pinto, P.M., Alonso, J.A.P., Fernández, V.P., Casero, J.J.D., 2006. Fungi isolated from diseased<br />

nursery seedlings in Spain. New Forests 31, 41-56.<br />

Rodríguez, R.G., 1989. Pitch canker on Pinus douglasiana, pines indigenous to San Andrés Milpillas,<br />

Municipal <strong>of</strong> Huajicori, Nay., City <strong>of</strong> Juarez, Chihuahua. Mexico.<br />

Steenkamp, E.T., Wingfield, B.D., Desjardins, A.E., Marasas, W.F.O., Wingfield, M.J., 2002. Cryptic<br />

speciation in Fusarium subglutinans. Mycologia 94, 1032-1043.<br />

Viljoen, A., Wingfield, M.J., Marasas, W.F.O., 1994. 1st Report <strong>of</strong> Fusarium subglutinans f. sp. pini<br />

on Pine-Seedlings in South-Africa. Plant Disease 78, 309-312.<br />

Wingfield, M.J., Jacobs, A., Coutinho, T.A., Ahumada, R., Wingfield, B.D., 2002. First report <strong>of</strong> the<br />

pitch canker fungus, Fusarium circinatum, on pines in Chile. Plant Pathology 51, 397-<br />

397.<br />

205


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 206-215<br />

POWDERY MILDEW ON WOODY PLANTS IN THE CZECH<br />

REPUBLIC<br />

Dagmar PALOVČÍKOVÁ 1* , Hana DANČÁKOVÁ,<br />

Hana MATOUŠKOVÁ, Jindřiška JUNÁŠKOVÁ, Libor JANKOVSKÝ*<br />

1 Mendel University <strong>of</strong> Agriculture and Forestry, Faculty <strong>of</strong> Forestry and Wood Technology,<br />

Department <strong>of</strong> Forest Protection and Wildlife Management, Zemědělská 3, 613 00 Brno, Czech<br />

Republic<br />

ABSTRACT<br />

*palovcik@mendelu.cz<br />

There were identified 30 different species <strong>of</strong> powdery mildew on woody plants in the<br />

Czech Republic. From this number <strong>of</strong> identified Erysiphales is Phyllactinia roboris<br />

reported as a missing species actually. Eleven species <strong>of</strong> reported powdery mildews can be<br />

termed alien for the Czech Republic, including oak powdery mildew Erysiphe alphitoides<br />

(syn. Microsphaera alphitoides Griff.) which is naturalized species throughout Europe now.<br />

Newly recorded species are Erysiphe arguata, E. azaleae, E. elevata, E. flexuosa, Erysiphe<br />

palczewskii, Erysiphe syringae, Erysiphe vanbruntiana var. sambuci-racemosae. Erysiphe<br />

euonymi-japonici was reported from herbarium specimen in 1941 only.<br />

Keywords: Powdery mildews, Erysiphales, Alien species<br />

1. INTRODUCTION<br />

Several important alien diseases <strong>of</strong> woody plants were introduced in Europe<br />

within 20th century. The largest number <strong>of</strong> newly discovered alien diseases is<br />

belonging to the powdery mildew Erysiphales. Powdery mildew was observed by<br />

Theoprastis on roses 300 year B.C. already. The survey <strong>of</strong> taxonomy <strong>of</strong> this group<br />

is given eg. by Jaczewski (1927), Blumer (1967), Braun et al. (1978, 1981,<br />

1987, 2006, 2007), Zheng (1985), Gelyuta (1989), Takamatsu et al. (2007) etc.<br />

The main area <strong>of</strong> powdery mildew distribution is temperate zone <strong>of</strong> Northern<br />

Hemisphere. Occurrence in tropics and subtropics is rare, in conidial stage mostly.<br />

Some genera <strong>of</strong> powdery mildews occur also in boreal and arctic areas, eg. in<br />

Scandinavia, Northern America, polar areas <strong>of</strong> Russia, Island and Greenland<br />

(Braun, 1995).<br />

Preliminary list <strong>of</strong> alien disease in the Czech Republic (CR) include more than<br />

30 most important species, including 4 quarantine pests, most <strong>of</strong> them belonging to<br />

Erysiphales. The aim <strong>of</strong> this paper is to evaluate <strong>of</strong> spectrum <strong>of</strong> powdery mildews<br />

in the CR focused to alien species.<br />

206


2. MATERIAL AND METHODS<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Distribution, identification and pathology <strong>of</strong> individual species <strong>of</strong> powdery<br />

mildew have been studied on set <strong>of</strong> 240 examined samples collected on amenity<br />

trees and woody plants in parks in the CR within 2004 - 2008. Leaves infected with<br />

powdery mildew were collected in North and South Moravia and East Bohemia.<br />

Identification was provided according Braun (1987, 1995) and Paulech C. (1995),<br />

taxonomy was corrected by Takamasu (2007). Identification was made according<br />

to morphological features; critical findings were confirmed by H.D. Shin, new<br />

species as Erysiphe elevata and E. azaleae were confirmed on the bases <strong>of</strong> DNA.<br />

Herbarium specimens are deposited at Herbarium <strong>of</strong> Department <strong>of</strong> Forest<br />

Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood Technology.<br />

3. RESULTS, DISCUSSION<br />

There were identified 30 different species <strong>of</strong> powdery mildew on woody plants<br />

belonging to following genera: Erysiphe, Microsphaera, Phyllactinia,<br />

Podosphaera, Sawadaea, Sphaerotheca, Uncinula, Uncinuliella (Tab. 1). From this<br />

number <strong>of</strong> identified Erysiphales is Phyllactinia roboris reported as a missing<br />

species actually and Erysiphe euonymi-japonici (Vienn.-Bourg.) U. Braun & S.<br />

Takamatsu is known only from herbarium specimen collected in 1941; recent<br />

records are missing. Nearly 11 species <strong>of</strong> reported powdery mildews can be termed<br />

alien, although origin <strong>of</strong> several species is not clear. Newly recorded species are<br />

Erysiphe arguata, E. azaleae, E. elevata, E. flexuosa, Erysiphe palczewskii,<br />

Erysiphe syringae, Erysiphe vanbruntiana var. sambuci-racemosae.<br />

Table 1. List <strong>of</strong> powdery mildews observed in the CR<br />

Powdery<br />

Mildew<br />

Erysiphe adunca (Wallr.) Fr.<br />

Hosts Origin/<br />

distribution<br />

(syn. Uncinula adunca (Wallr.) Lév.)<br />

Salix appendiculata,<br />

S. caprea, S. renii<br />

Erysiphe adunca var. adunca<br />

(Wallr.) Fr.<br />

(syn. Uncinula adunca var. adunca<br />

(Wallr.) Lév.)<br />

Populus nigra<br />

Erysiphe<br />

alphitoides<br />

(Griffon &<br />

Maubl.) U.<br />

Braun & S.<br />

Takam.<br />

(syn.<br />

Microsphaera<br />

alphitoides var.<br />

Quercus robur,<br />

Q. petraea, Q.<br />

cerris,<br />

Q. robur<br />

'Fastigiata',<br />

Q. glandifera,<br />

Castanea sativa<br />

- / Europe,<br />

Asia, N.<br />

America<br />

- / all Europe,<br />

all Asia, N.<br />

America<br />

Asia ?/ today<br />

nearly global<br />

First<br />

record<br />

from<br />

Europe<br />

1876<br />

Portugal,<br />

1906/1907<br />

spreading<br />

throughou<br />

t Europe;<br />

207<br />

1909<br />

outbreak<br />

References First<br />

record<br />

in the<br />

CR<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

References<br />

2006 Palovčíková<br />

et al 2007<br />

2005 Palovčíková<br />

et al 2007<br />

1907 (?)<br />

Cejp et<br />

Skalický<br />

1954,<br />

Příhoda1959


Powdery<br />

Mildew<br />

alphitoides<br />

Griffon &<br />

Maubl.)<br />

Erysiphe arcuata<br />

U. Braun,<br />

V.P.Heluta and<br />

S. Takam.<br />

Erysiphe azaleae<br />

(U. Braun) U.<br />

Braun & S.<br />

Takam)<br />

(syn.<br />

Microsphaera<br />

azaleae U.<br />

Braun)<br />

Erysiphe<br />

berberidis D.C.<br />

(syn.<br />

Microsphaera<br />

berberidis (DC.)<br />

Lév.<br />

Erysiphe elevata<br />

(Burrill) U.<br />

Braun & S.<br />

Takam.<br />

(syn.<br />

Microsphaera<br />

elevata Burrill)<br />

Erysiphe<br />

euonymi (DC.)<br />

(syn.<br />

Microsphaera<br />

euonymi (DC.)<br />

Sacc.)<br />

Erysiphe<br />

euonymijaponici<br />

(Vienn.-<br />

Bourg.) U.<br />

Braun & S.<br />

Takamatsu<br />

(Microsphaera<br />

euonymijaponici<br />

Vienn.-<br />

Bourg.)<br />

Erysiphe<br />

flexuosa (Peck)<br />

U. Braun et S.<br />

Takamatsu<br />

(syn.<br />

Uncinuliella<br />

flexuosa (Peck)<br />

U. Braun)<br />

Hosts Origin/<br />

distribution<br />

Carpinus betulus<br />

(after revision;<br />

reported as a E.<br />

carpinicola<br />

previously)<br />

Rhododendron<br />

spp.<br />

Berberis<br />

thunbergii<br />

'Atropurpurea'<br />

B. vulgaris,<br />

B. vulgaris x<br />

rubra,<br />

Mahonia<br />

aquifolium<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Japan, Far<br />

East?<br />

North America<br />

- /all Europe,<br />

Central Asia,<br />

Turkey, Iran<br />

208<br />

First<br />

record<br />

from<br />

Europe<br />

Western<br />

Europe<br />

Germany,<br />

Hungary,<br />

Ukraine<br />

2006<br />

Germany<br />

1997<br />

References First<br />

record<br />

in the<br />

CR<br />

Braun et al.<br />

2006<br />

Pastirčáková et<br />

al. 2008<br />

Braun 1997,<br />

Inmann et al.<br />

2000<br />

Braun 1987<br />

Paulech 1995<br />

Catalpa<br />

bignonioides North America 2002 Vajna et al<br />

2003,<br />

Euonymus<br />

europaeus<br />

- /all Europe,<br />

Central Asia,<br />

Turkey<br />

Euonymus spp. Asia /Europe,<br />

Asia,<br />

N.America,<br />

S.America,<br />

Australia,<br />

Aesculus x<br />

carnea<br />

A.<br />

hippocastaneum,<br />

A. pavia<br />

New Zealand<br />

North<br />

America/<br />

North America<br />

Ale-Agha et al.<br />

2004<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

2000 Ale-Agha et al.<br />

2000,<br />

Ing et Spooner<br />

2002,<br />

Zimmermanno<br />

va-<br />

Pastircakova et<br />

al. 2002<br />

References<br />

2004 Palovčíková<br />

et al. 2007<br />

Palovčíková<br />

2003 et<br />

Dančáková<br />

2005,<br />

Lebeda et al.<br />

2007,<br />

Bacigálová,<br />

Marková<br />

2006<br />

2004 Palovčíková<br />

et<br />

Dančáková<br />

2005,<br />

2005 Ale-Agha et<br />

al. 2004,<br />

Palovčíková<br />

et al. 2007<br />

2005 Palovčíková<br />

et al.2007<br />

1931<br />

Piskoř<br />

by<br />

Prague<br />

Herb<br />

specimen M-<br />

0016258 The<br />

Erysiphales<br />

Collection at<br />

the<br />

Botanische<br />

Staatssamml<br />

ung<br />

München,<br />

2004 Palovčíková<br />

et<br />

Dančáková<br />

2005,<br />

Palovčíková<br />

et al 2007


Powdery<br />

Mildew<br />

Erysiphe<br />

hedwigii (Lév.)<br />

U. Braun & S.<br />

Takam.<br />

(syn.<br />

Microsphaera<br />

hedwigii Lév.)<br />

Erysiphe<br />

lonicerae (DC.)<br />

(syn.<br />

Microsphaera<br />

lonicerae (DC.)<br />

G. Winter)<br />

Erysiphe ornata<br />

var. europaea<br />

(U.Braun)<br />

U.Braun & S.<br />

Takam.<br />

(syn.<br />

Microsphaera<br />

ornata var.<br />

europaea<br />

U.Braun)<br />

Erysiphe<br />

palczewskii<br />

(Jacz.) Braun &<br />

Takamatsu<br />

(syn.<br />

Microsphaera<br />

palczewskii<br />

Jacz.)<br />

Erysiphe<br />

penicillata<br />

(Wallr.) Link<br />

(syn.<br />

Microsphaera<br />

penicillata<br />

(Wallr.) Lév.)<br />

Erysiphe<br />

syringae<br />

Schwein.<br />

(syn.<br />

Microsphaera<br />

syringae<br />

(Schwein) H.<br />

Magn.)<br />

Erysiphe tortillis<br />

(Wallr.) Link<br />

(syn.<br />

Microsphaera<br />

tortilis (Wallr.)<br />

Speer<br />

Erysiphe<br />

vanbruntiana<br />

var. sambuciracemosae<br />

(U.Braun) U.<br />

Braun & S.<br />

Takamatsu<br />

(syn.<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Hosts Origin/<br />

distribution<br />

Viburnum<br />

lantana<br />

-/ Europe,<br />

Armenia,<br />

Siberia<br />

Lonicera nigra - / all Europe,<br />

Central Asia,<br />

Japan<br />

Alnus glutinosa<br />

Betula pendula<br />

Crataegus<br />

monogyna<br />

- / all Europe,<br />

Central Asia<br />

Caragana spp. Asia/ Europe.<br />

Asia<br />

Alnus glutinosa - / all Europe,<br />

Asia (Iran,<br />

Siberia, East<br />

USSR, Japan),<br />

N.America<br />

Syringa vulgaris,<br />

S. chinensis<br />

Viburnum opulus<br />

Cornus<br />

sanguinea<br />

Sambucus<br />

racemosa<br />

N.America/<br />

N.America,<br />

Europa,<br />

Siberia,<br />

Australia<br />

- / all Europe<br />

- /<br />

North<br />

America, Asie<br />

First<br />

record<br />

from<br />

Europe<br />

209<br />

Hungary<br />

2005<br />

References First<br />

record<br />

in the<br />

CR<br />

References<br />

Braun 1987<br />

2004 Palovčíková<br />

et al. 2007<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun1995,<br />

Braun &<br />

Takamatsu<br />

2000, Gelyuta<br />

et Minter 1998,<br />

Vajna 2006b<br />

Braun 1987<br />

Sinclair 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

2006 Palovčíková<br />

et al 2007<br />

2005 Palovčíková<br />

et al 2007<br />

2006<br />

Lebeda et al.<br />

2008<br />

2006 Palovčíková<br />

et al 2007<br />

2004<br />

2005<br />

2005<br />

Palovčíková<br />

et<br />

Dančáková<br />

2005,<br />

Palovčíková<br />

et al 2007<br />

Palovčíková<br />

et al 2007<br />

Palovčíková<br />

et<br />

Dančáková<br />

2005,<br />

Palovčíková<br />

et al 2007


Powdery<br />

Mildew<br />

Microsphaera<br />

vanbruntiana<br />

var.sambuciracemosae<br />

U.Braun)<br />

Microsphaera sp.<br />

(Microsphaera<br />

penicillata s.l.)<br />

Phyllactinia<br />

fraxini (DC.)<br />

Fuss<br />

Phyllactinia<br />

guttata (Wallr.)<br />

Lév.<br />

Phyllactinia mali<br />

(Duby) U. Braun<br />

Phyllactinia<br />

roboris (Gachet)<br />

Blumer<br />

Podosphaera<br />

clandestina<br />

(Wallr.) Lév.<br />

Podosphaera<br />

clandestina var.<br />

clandestina<br />

(Wallr.) Lév.<br />

Podosphaera<br />

pannosa (Wallr.)<br />

de Bary<br />

(syn.<br />

Sphaerotheca<br />

pannosa (Wallr.)<br />

Lév.)<br />

Podosphaera<br />

tridactyla<br />

(Wallr.) de Bary<br />

Hosts Origin/<br />

distribution<br />

Sorbus<br />

intermedia<br />

F.excelsior,<br />

F.excelsior<br />

´Hessei´<br />

F. angustifolia<br />

Betula pendula,<br />

B.verrucosa,<br />

B.verrucosa<br />

'Yongii,,<br />

B. papyrifera,<br />

Cornus mas,<br />

Corylus avellana,<br />

C. avellana<br />

'Concorta‚<br />

C<br />

.avellana´Hetero<br />

phylla´,<br />

C. colurna, C.<br />

maxima<br />

'Purpurea',<br />

Crataegus<br />

monogyna ,<br />

Fagus sylvatica,<br />

Salix sp.<br />

Crataegus<br />

monogyna<br />

Quercus spp. –<br />

not confirmed<br />

within past years;<br />

in red list<br />

Sorbus<br />

intermedia<br />

Crataegus<br />

oxyacantha<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

- / all Europe,<br />

Turkey, Asia,<br />

N.America,<br />

N.Africa<br />

-/ global<br />

- / all Europe,<br />

Asia,<br />

N.America,<br />

N.Africa<br />

- / South<br />

Europe, Asia,<br />

S. America<br />

- / all Europe,<br />

Asia,<br />

N.America<br />

- / all Europe,<br />

Asia,<br />

N.America<br />

210<br />

First<br />

record<br />

from<br />

Europe<br />

1885 –<br />

Slovakia<br />

??<br />

References First<br />

record<br />

in the<br />

CR<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Rosa rugosa -/global Braun 1987<br />

Padus avium - / all Europe,<br />

Asia, America,<br />

Australia, New<br />

Zealand<br />

Sinclair 1987<br />

Braun 1987<br />

Paulech 1995<br />

References<br />

2004 Palovčíková<br />

et al 2007<br />

2005 Palovčíková<br />

et al 2007<br />

2006 Palovčíková<br />

et al 2007<br />

2004 Palovčíková<br />

et al 2007<br />

2005 Palovčíková<br />

et al 2007<br />

2006 Palovčíková<br />

et al 2007


Powdery<br />

Mildew<br />

Sawadaea<br />

bicornis (Wallr.)<br />

Homma<br />

Sawadaea<br />

tulasnei (Fuckel)<br />

Homma<br />

Uncinula<br />

prunastri var.<br />

prunastri (DC.)<br />

Sacc.<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Hosts Origin/<br />

distribution<br />

Acer campestre,<br />

A. ginnala, A.<br />

negundo,<br />

A. platanoides,<br />

A.<br />

pseudoplatanus,<br />

A. saccharinum<br />

Acer ginnala,<br />

A. palmatum<br />

'Dissectum',<br />

A. platanoides<br />

- / all Europe,<br />

Asia, New<br />

Zealand<br />

- / all Europe,<br />

Asia<br />

Prunus sp. - / Europe,<br />

Central Asia<br />

First<br />

record<br />

from<br />

Europe<br />

211<br />

References First<br />

record<br />

in the<br />

CR<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

Braun 1987<br />

Paulech 1995<br />

References<br />

2004 Palovčíková<br />

et<br />

Dančáková<br />

2005,<br />

2006 Palovčíková<br />

et al 2007<br />

2006 Palovčíková<br />

et al 2007<br />

Oak powdery mildew Erysiphe alphitoides (Griffon & Maubl.) U. Braun & S.<br />

Takam. (syn. Microsphaera alphitoides Griff.) is most important species for<br />

<strong>forestry</strong> as a naturalized species throughout Europe now. The origin <strong>of</strong> this species<br />

is unclear, although this species is widespread in Europe, Asia, North and South<br />

America, Australia and New Zealand. The first occurrence on this species is origin<br />

from limited area in Portugal from 1876 – 1877, in 1906 – 1907 were reported<br />

spreading <strong>of</strong> this species in many countries in Europe, in 1909 is mentioned<br />

outbreak <strong>of</strong> this species in Western Europe. Some sources note, that this species<br />

were introduced from Northern America (eg. Cejp et Skalický, 1954), some other<br />

authors assume, that this species were introduced to Asia, however recent genetic<br />

studies shows affinity to other species <strong>of</strong> powdery mildews in tropical areas in<br />

South East Asia (Limkaisang et al., 2006). Ufnalski et Przybyl (2004) show<br />

genetic diversity <strong>of</strong> oak powdery mildew.<br />

Rhododendron powdery mildew Erysiphe azaleae (U. Braun) U. Braun & S.<br />

Takam. (syn. Microsphaera azaleae U. Braun), probably introduced from North<br />

America or Asia (Inman et al., 2000), has been from Europe recorded in England,<br />

Germany, Switzerland (Inman et al., 2000), and Poland (Piatek, 2003; Shin &<br />

Mulenko, 2004) over recent years. Rhododendron powdery mildew has been firstly<br />

recorded in the CR by in 2003 by Lebeda et al. (2007) and furthermore reported by<br />

Bacigálová and Marková (2006). Species is widespread in parks and gardens across<br />

CR actually. One <strong>of</strong> reasons <strong>of</strong> spreading is trade with plant material with<br />

combination <strong>of</strong> favorable climatic conditions within past years.<br />

The catalpa powdery mildew Microsphaera elevata Burrill is a native species in<br />

North America (Braun 1987). From Europe was recorded in Europe by Ale-Agha<br />

et al. (2000). Actually is reported from the Germany, Hungary, Poland, Slovakia,


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Switzerland, United Kingdom, etc. (Ale-Agha et al., 2004; Vajna et al., 2004;<br />

Pastirčáková et al., 2006; etc.). Is it widespread in plantation in towns in the CR<br />

now (Ale-Agha et al., 2004; Palovčíková et al., 2007). Some latest records are<br />

origin from Bulgaria (Denchev et al., 2008), Romania and from Italy (own<br />

observations, not published). Reasons <strong>of</strong> spreading are the same – trade with plant<br />

material and favorable climate. Some other species on Catalpa Erysiphe catalpae<br />

Simonyanis was not reported from CR up to date. This species is confused with E.<br />

elevata in some European countries frequently (Ale-Agha et al., 2004).<br />

Horse chestnut mildew E. flexuosa (Peck) U. Braun & S. Takam (syn.<br />

Uncinuliella flexuosa (Peck) U. Braun.) on Aesculus spp., was firstly reported from<br />

Europe by Ale-Agha et al. (2000), Ing and Spooner (2002) and others. It is a<br />

common powdery mildew species infecting Aesculus trees in North America<br />

(Glawe and Dugan, 2007) and Europe - Croatia, France, Germany, Lithuania,<br />

Poland, Romania, Serbia, Slovakia, Slovenia, Switzerland, Ukraine, and United<br />

Kingdom (Braun, 1987; Heluta and Voytyuk, 2004; Pricop & Tănase, 2007;<br />

Zimmermannová-Pastirčáková et al., 2002; Kiss et al., 2004). From the Czech<br />

Republic is reported by Zimmermannová-Pastirčáková et al. (2002) and<br />

Palovčíková et al. (2007).<br />

Some other new species for the CR is Erysiphe arcuata U. Braun, V. P. Heluta<br />

and S. Takam. on hornbeams Carpinus betulus, previously (Palovčíková et al.,<br />

2007) reported this species as Erysiphe carpinicola (Hara) U. Braun & S. Takam.<br />

Braun et al. (2006) re-examined European powdery mildew collections on C.<br />

betulus (including the anamorph Oidium carpini) from Germany, Hungary and<br />

Ukraine, and described them as a new species E. arcuata, contrary to previous<br />

records. Pastirčaková et al. (2008) reports this species from Slovakia. Previous<br />

records <strong>of</strong> E. carpinicola from Hungary (Vajna, 2006a) and from Poland<br />

(Wołczanska, 2007; Piatek, 2004) probably regarded as E. arcuata as well.<br />

Caragana powdery mildew Erysiphe palczewskii Braun & Takamatsu (syn.<br />

Microsphaera palczewskii Jacz.) is native to Asia, however it has been introduced<br />

into many European countries (Gelyuta and Gorlenko, 1984; Braun, 1995; Gelyuta<br />

and Minter, 1998; Braun et al., 2006; Vajna, 2006b). During the summer <strong>of</strong> 2006<br />

were severe Erysiphe palczewskii recorded in the CR into area <strong>of</strong> Central<br />

Moravia as well (Lebeda et al., 2008).<br />

Erysiphe syringae Schwein. (syn. Microsphaera syringae (Schwein) H. Magn.)<br />

and Erysiphe vanbruntiana var. sambuci-racemosae (U. Braun) U. Braun & S.<br />

Takamatsu is common species across the CR actually and it was recorded in 2005<br />

(Palovčíková et al., 2007).<br />

Erysiphe euonymi-japonici (Vienn.-Bourg.) U. Braun & S. Takamatsu is<br />

reported from herbarium specimen No. M-0016258, collected in 1931 at Piskoř by<br />

Prague and specimen is deposited in powdery mildew collection at München,<br />

Germany (Botanische Staatssammlung München), although recent records are<br />

missing.<br />

212


4. CONCLUSIONS<br />

SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The largest number <strong>of</strong> newly discovered alien disease comes from the powdery<br />

mildew group Erysiphales. Newly recorded species are Erysiphe arguata, E.<br />

azaleae, E. elevata, E. flexuosa, Erysiphe palczewskii, Erysiphe syringae, Erysiphe<br />

vanbruntiana var. sambuci-racemosae. Erysiphe euonymi-japonici is reported<br />

from herbarium specimen in 1941 only.<br />

Explication <strong>of</strong> its spreading could be climatic conditions within past years and<br />

also by the interest about this group within past 10 years. The occurrence <strong>of</strong><br />

powdery mildew has not serious impact on health on observed woody plants<br />

actually. Powdery mildew is problem for nurseries and trade with plant material.<br />

5. ACKNOWLEDGEMENT<br />

Paper was supported by MSM 6215648902.<br />

6. REFERENCES<br />

Ale-Agha, N., Bolay, A., Braun, U., Feige, B., Jage, H., Kummer, K., Lebeda, A., Piątek, M., Shin,<br />

H.D., Zimmermannová-Pastirčaková, K., 2004. Erysiphe catalpae and Erysiphe elevata<br />

in Europe. Mycological Progress 3(4), 291 – 296.<br />

Ale-Agha, N., Braun, U., Feige, B., Jage, H., 2000. A new powdery mildew disease on Aesculus spp.<br />

introduced in Europe. Cryptogamie Mycologie 21, 89-92.<br />

Bacigálová, K., Marková, J., 2006. Erysiphe azaleae (Erysiphales) – a new species <strong>of</strong> powdery<br />

mildew for Slovakia and further records from the Czech republic. Czech Mycology 58(3-<br />

4), 198-199.<br />

Blumer, S., 1967. Echte Mehltaupilze (Erysiphaceae) ein Bestimmungsbuch für die in Europa<br />

vorkommenden Arten. Gustav Fischer Verl. Jena, pp. 436.<br />

Braun, U., 1978. Beiträg zur Systematik und Nomenklatur der Erysiphales. Feddes Report. 88, 655-<br />

665.<br />

Braun, U., 1981. Miscellaneous notes on Erysiphaceae II. Feddes Report. 92, 499-513.<br />

Braun, U., 1987. A Monograph <strong>of</strong> the Erysiphales (Powdery Mildews). Beihefte zur Nova Hedwigia<br />

89, 700.<br />

Braun, U., 1995. The Powdery Mildews (Erysiphales) <strong>of</strong> Europe. Gustav Fischer Verlag, Jena, pp.<br />

337.<br />

Braun, U., Heluta, V., Takamatsu, S., 2007. Validation <strong>of</strong> two species <strong>of</strong> the genus Erysiphe<br />

(Erysiphales). Schlechtendalia 16, 99–100.<br />

Braun, U., Takamatsu, S., Heluta, V., Limkaisang, S., Divirangkoon, R., Cook, R., Boyle, H., 2006.<br />

Phylogeny and taxonomy <strong>of</strong> powdery mildew fungi <strong>of</strong> Erysiphe sect. Uncinula on<br />

Carpinus species. Mycological Progress 5, 139–153.<br />

Cejp, K., Skalický, V., 1954. Plodná padlí na dubech v Československu, Microsphaera alphitoides<br />

GriffonetMaublanc a Phyllactinia roboris (Gachet) Blumer [Powdery mildews on oaks in<br />

213


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Czechoslovakia, Microsphaera alphitoides GriffonetMaublanc a Phyllactinia roboris<br />

(Gachet) Blumer]. Preslia 26, 43–54.<br />

Denchev, M, C. et al., 2008. New records <strong>of</strong> fungi, fungus-like organisms, and slime moulds from<br />

Europe and Asia: 1-6. Mycologia Balcanica 5 (93), 93–96.<br />

Gelyuta, V. P., Minter, D.W., 1998. Microsphaera palczewskii. IMI Descriptions <strong>of</strong> Fungi and<br />

Bacteria N0. 1375, Surrey, UK: CABI Bioscience.<br />

Gelyuta, V.P., 1989. Flora gribov Ukrajiny. Mučnisto-rosjanije griby. Naukova Dumka AN USSR,<br />

Kijev, pp. 256.<br />

Gelyuta, V.P., Gorlenko, M.V., 1984. Microsphaera palczewskii Jacz. in the USSR. Mikologia i<br />

Fitopatologia 18, 177-182.<br />

Glawe, D.A., Dugan, F.M., 2006. First report <strong>of</strong> Erysiphe (Uncinuliella) flexuosa in western North<br />

America. – Pacific Northwest Fungi 111, 1-11. [http://www.pnwfungi.org]<br />

Heluta, V.P., Voytyuk, S.O., 2004. Uncinula fl exuosa Peck, a new invasive species <strong>of</strong> the powdery<br />

mildew fungi (Erysiphales) in Ukraine. – Ukrainskiy Botanicheskiy Zhournal 61(5), 17-<br />

25. (In Ukrainian)<br />

Ing, B., Spooner, B., 2002. The Horse Chestnut powdery mildew Uncinula flexuosa in Europe. –<br />

Mycologist 16, 112-113.<br />

Inman, A.J., Cook, R.T.A., Beales, P.A., 2000. A contribution to the identity <strong>of</strong> Rhododendron<br />

powdery mildew in Europe. Journal <strong>of</strong> Phytopathology 148, 17-27.<br />

Jaczewski, A.A., 1927. Karmannyj opredeliteľ gribov 2. Mučnisto-rosjanije griby. Leningrad, pp.<br />

626.<br />

Kiss, L., Vajna, L., Fischl, G., 2004. Occurrence <strong>of</strong> Erysiphe flexuosa (syn. Uncinula flexuosa) on<br />

horse chestnut (Aesculus hippocastanum) in Hungary. Plant Pathology 53: 245. [online]<br />

<br />

Lebeda, A., Mieslerova, B., Sedlářová, M., 2008. First report <strong>of</strong> Erysiphe palczewskii on Caragana<br />

arborescens in the Czech Republic. Plant Pathology 57(4), 779.<br />

Lebeda, A., Sedlarova, M., Jankovsky, L., Shin, H.D., 2007. First report <strong>of</strong> rhododendron powdery<br />

mildew on Rhododendron spp. in the Czech Republic. Plant Pathology 56 (2), 354-354.<br />

Limkaisang, S., Cunnington, J.H., Wui, L.K., Salleh, B., Sato, Y., Divarangkoon, R., Fangfuk,<br />

W., To-anun, Ch., Takamatsu, S., 2006. Molecular phylogenetic analyses reveal a close<br />

relationship between powdery mildew fungi on some tropical trees and Erysiphe<br />

alphitoides, an oak powdery mildew. Mycoscience, 47 (6), 327-335.<br />

Palovčíková, D., Dančáková, H., 2005. Bionomie a zástupci čeledi padlí v České republice. [The<br />

bionomy and species <strong>of</strong> powdery mildew in the Czech Republic]. Lesnická práce 11,<br />

600–601.<br />

Palovčíková, D., Dančáková, H., Junášková, J., Matoušková, H., Jankovský, L., 2007. Druhové<br />

spektrum padlí na dřevinách v České republice, nové druhy padlí dřevin v ČR, [The<br />

powdery mildews on woody plants in the CZ. Some new species <strong>of</strong> powdery mildews in<br />

the CR]. In Kodrík, M. - Hlaváč, P. Ochrana lesa 2007. Zvolen: Technická univerzita vo<br />

Zvolene, katedra ochrana lesa a polovnictva, 2007, 71-79.<br />

Pastirčáková, K., Pastirčák, M. & Juhásová, G., 2006. The Catalpa powdery mildew Erysiphe elevata<br />

in Slovakia. Cryptogamie Mycologie 27, 31-34.<br />

Pastirčáková, K., Takamatsu, S., Shiroya, Y. and Pastirčák, M., 2008. European Hornbeam Powdery<br />

Mildew Erysiphe arcuata in Slovakia. J. Phytopathology 156, 597–601.<br />

214


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Paulech, C., 1995. Flóra Slovenska, X/1, Huby Múčnatkotvaré (Erysiphales). [Flora <strong>of</strong> Slovakia,<br />

Powdery mildews (Erysiphales)]. Veda, Bratislava, 291.<br />

Piatek, M., 2003. Erysiphe azaleae and Erysiphe syringae-japonicae introduced in Poland.<br />

Mycotaxon 87, 121-126.<br />

Piątek, M., 2004. First report <strong>of</strong> powdery mildew (Oidium carpini) on Carpinus betulus in Poland.<br />

Plant Pathology 53, 246. (First published online: New Disease Reports 8,<br />

http://www.bspp.org.uk/ndr/jan2004/2003-67.asp)<br />

Pricop, C., Tănase, C., 2007. Erysiphe flexuosa (Peck) U. Braun & S. Takamatsu (Fungi,<br />

Ascomycota) – a new species in Romania. – Analele Ştiinţifice ale Universităţii “Al. I.<br />

Cuza”, Iaşi, Biologie vegetală 532, 104-107.<br />

Příhoda, A., 1959. Lesnická fytopatologie. [Forest Pathology]. SZN Praha, 363.<br />

Shin, H.D., Mulenko, W., 2004. The record <strong>of</strong> Erysiphe azaleae (Erysiphales) from Poland and its<br />

anamorph. Mycobiology 32, 105-109.<br />

Sinclair, W.A., Lyon, H., 1987. Diseases <strong>of</strong> trees and shrubs. Comstock Publishing Associates, Ithaca<br />

and London, pp. 575.<br />

Takamatsu, S., Braun, U., Limkaisang, S., Kom-un, S., Sato, Y., Cunnington, J.H., 2007. Phylogeny<br />

and taxonomy <strong>of</strong> the oak powdery mildew Erysiphe alphitoides sensu lato. Mycological<br />

Research 111, 809–826.<br />

Ufnalski, K., Przybyl, K., 2004. Variability in morphology <strong>of</strong> Microsphaera alphitoides Griffon et<br />

Maubl. in Poland. Acta Soc. Bot. Pol. 73, 233–237.<br />

Vajna, L., 2006a. Powdery mildew caused by Erysiphe carpinicola on Carpinus betulus in Hungary:<br />

the first European report. Plant Pathology 55, 575. (First published online: New Disease<br />

Reports 12, http://www.bspp.org.uk/ndr/jan2006/2006-08.asp<br />

Vajna, L., 2006b. First report <strong>of</strong> powdery mildew on Caragana arborescens in Hungary caused by<br />

Erysiphe palczewskii. (First published online: New Disease Reports 8,<br />

http://www.bspp.org.uk/ndr/july2006/2006-22.asp).<br />

Vajna, L., Fischl, G., and Kiss, L., 2004. Erysiphe elevata (syn. Microsphaera elevata), a new North<br />

American powdery mildew fungus in Europe infecting Catalpa bignonioides trees. Plant<br />

Pathology 53, 244.<br />

Wołczanska, A., 2007. First report <strong>of</strong> Erysiphe carpinicola s.l. (perfect state) in Poland. Plant Pathol<br />

56, 354.<br />

Zheng, R.Y., 1985. Genera <strong>of</strong> the Erysiphaceae. Mycotaxon 22, 209-263.<br />

Zimmermannova-Pastircakova, K., Adamska, I., Blaszkowski, J., Bolay, A., Braun, U., 2002.<br />

Epidemic spread <strong>of</strong> Erysiphe flexuosa (North American powdery mildew <strong>of</strong> horsechestnut)<br />

in Europe. Schlechtendalia 8, 39-45.<br />

215


Abiotic Diseases And Other Diseases<br />

217


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

218


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 219-232<br />

URBAN TREE HEALTH OF 49 GREEN SPACES IN MADRID (SPAIN)<br />

Eva ALFONSO CORZO 1 *, M.J. GARCIA-GARCIA 1 , J.A. SAIZ de OMEÑACA 1<br />

ABSTRACT<br />

1 Technical University <strong>of</strong> Madrid, Madrid, Spain<br />

* alfonso_corzo@hotmail.com<br />

In order to improve the management <strong>of</strong> the urban tree health in Madrid, a sample <strong>of</strong> 49<br />

green spaces was evaluated. Data were obtained by visual assessment in 6 different districts<br />

<strong>of</strong> the city, during a period <strong>of</strong> 3 or 4 years. Diseases, pests and other problems were<br />

identified for each species in every green area, without considering the number <strong>of</strong> trees.<br />

The health status <strong>of</strong> the trees declined along the period and there were few differences<br />

among the districts. The main problems in trees were stem injury, dead branches, epicormic<br />

shoots and decay, and the most damaged species were Siberian elm (Ulmus pumila), Black<br />

locust (Robinia pseudoacacia), Box elder (Acer negundo), and Plane tree (Platanus<br />

hispanica). Decay was related with stem injury. Biotic diseases were encountered less<br />

frequently than abiotic, and they affected a smaller rank <strong>of</strong> species. The results showed a<br />

positive association between naturalness <strong>of</strong> species and their health status, with a higher<br />

damage risk among the exotic ones. Based on their health features, a ranking <strong>of</strong> the less<br />

suitable species to be grown in Madrid is given.<br />

Keywords: urban green space, tree health, shrub, naturalness, abiotic disease<br />

1. INTRODUCTION<br />

The city constitutes a hostile environment for ornamental vegetation because<br />

many adverse conditions may affect the plants living in it: from environmental<br />

factors such as urban heat island, air pollution or mechanical stem injury to biotic<br />

stresses such as pests and diseases. Several authors have reported different<br />

difficulties which are related with the urban environment (Impens and Delcarte,<br />

1979; Rocray, 1983; Berrang et al., 1985). These factors not only injure the plants,<br />

but, some <strong>of</strong> them also predispose plants to suffer from another diseases<br />

(Kozlowski, 1985), interacting synergistically.<br />

In order to improve vegetation management <strong>of</strong> the green spaces in Madrid, the<br />

city council commissioned a survey to the Forestry Pathology Department <strong>of</strong> the<br />

Technical University <strong>of</strong> Madrid, which was carried out from 2005 to 2008.<br />

The study has four principal objectives in trees, shrubs, vines and seasonal<br />

flowers populations:<br />

219


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Following up their health status.<br />

Identifying main problems and the most affected species.<br />

Evaluating if there would be an association between naturalness <strong>of</strong> trees<br />

species and their health condition.<br />

Proposing a ranking <strong>of</strong> the less suitable species, taking into consideration<br />

their health features.<br />

2. THE STUDY SITE<br />

The study was undertaken in six districts <strong>of</strong> Madrid. This city is the capital and the<br />

most populated city <strong>of</strong> Spain (about 3.2 million inhabitants). It is located at a latitude <strong>of</strong><br />

40°26′ N and a longitude <strong>of</strong> 3°41′ W with an altitude <strong>of</strong> 667m.a.s.l. The climate is<br />

temperate Mediterranean with a marked continentality. The monthly mean minimum<br />

temperature is 2.6º C in January and maximum, 31.2º C in July. August is the driest<br />

month with only 10mm <strong>of</strong> precipitation, being 436mm the annual value.<br />

Madrid has more than 1,500 public green spaces which occupy more than 9% <strong>of</strong><br />

its area. Besides, there are more than 300,000 street trees, without taking into<br />

account those <strong>of</strong> green spaces.<br />

3. METHODOLOGY<br />

The city is administratively divided into 21 districts, and the studied green<br />

spaces are distributed throughout six <strong>of</strong> them: Arganzuela, Barajas, Hortaleza,<br />

Retiro, Salamanca and Villa de Vallecas.<br />

A randomized representative sample <strong>of</strong> 25% <strong>of</strong> the green spaces in each district<br />

was taken. Samples were collected for each year along the period 2005-2008,<br />

although in 2007 only the green spaces <strong>of</strong> Arganzuela were examined.<br />

Consequently, as a result <strong>of</strong> the random process, only a few green spaces were<br />

evaluated all the years. An interesting temporary evolution was expected, so, only<br />

the green spaces which were common to all the years in the period were analyzed<br />

in this survey.<br />

Sampling is a more and more used technique because <strong>of</strong> the expensiveness <strong>of</strong><br />

complete inventories, and it may be an accurate method for revealing the general<br />

patterns and trends in street tree populations (Jaenson et al., 1992). Complete<br />

inventories, such as the one which was carried out with 81,000 trees in Brussels<br />

(Impens and Delcarte, 1979), are much more precise, but, they would not be<br />

necessary if the management and conservation <strong>of</strong> green spaces was the target, since<br />

“all data collected must be related to the goals <strong>of</strong> the inventory” (Smiley and<br />

Baker, 1988).<br />

To sum up, 49 green spaces were examined in the whole <strong>of</strong> the six districts and<br />

they represented about 10% <strong>of</strong> the total green spaces (Figure 1).<br />

220


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

A periodicity <strong>of</strong> one or two years was thought as adequate (Harris et al., 2004).<br />

The seasons during which the inventories were compiled depended on the district,<br />

but in all cases, an inspection was conducted in autumn and, another one, in spring,<br />

so as to find out the different features than could be observed whether plants keep<br />

their foliage or not.<br />

4470000 4480000 4490000<br />

¨<br />

Scale 1: 400.000<br />

0 2.500 5.000 10.000 15.000<br />

Legend<br />

Number <strong>of</strong> examined<br />

green spaces<br />

Arganzuela: 11<br />

Barajas: 9<br />

Hortaleza: 13<br />

Retiro: 5<br />

Salamanca: 4<br />

Villa de Vallecas: 7<br />

m<br />

221<br />

4<br />

5<br />

11<br />

410000 420000 430000 440000 450000<br />

Figure 1. Number <strong>of</strong> examined green spaces in the inventoried districts.<br />

Data were collected by pr<strong>of</strong>essors and collaborators <strong>of</strong> the Forestry Pathology<br />

Department. They were trained so that they could recognize the major health<br />

problems <strong>of</strong> the ornamental species, before conducting the visual inspection in each<br />

area. Just in case they were not able to recognize something, both a photograph and<br />

a vegetal sample were taken, in order to identify them in lab, with the aid <strong>of</strong><br />

bibliography. The information recorded in each green space and in each year<br />

consisted <strong>of</strong>: date; name <strong>of</strong> the green space; district and ward where it was located;<br />

problems affecting the soil; problems which were related with watering; all tree,<br />

shrub, vine or seasonal flowers species which were found in the green space; and<br />

all the diseases, pest and other problems or disturbances which were observed for<br />

each species.<br />

Visual assessment was chosen as a tool for predicting the health status <strong>of</strong> the<br />

plants. Visual evaluation is frequently a controversial issue since some authors<br />

defend the idea that it can be a reliable means to predict internal decay (Kennard et<br />

al., 1996), whereas others think that this procedure is insufficient (Dunster, 1996).<br />

Therefore, the present study intended only to be a first approximation <strong>of</strong> the plant<br />

health condition, trying to predict failure situations, so that a possible further<br />

13<br />

7<br />

9


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

survey could be focused on those riskiest cases. Nevertheless, predicting failure<br />

situations is a difficult task, as most guidelines have been developed after the<br />

failure has already occurred, and few analyses <strong>of</strong> the ability to predict have been<br />

performed (Harris et al., 2004).<br />

All the green spaces were thoroughly inspected and the collected data were<br />

computerized in spreadsheets in order to facilitate their analysis. The unit for this<br />

analysis was defined as SYG, which stands Species-Year-Green space. It referred to<br />

one species which was observed in a certain year (2005, 2006, 2007 or 2008) and in a<br />

specific green space. From now on, it will be mentioned by simply using SYG.<br />

4. RESULTS AND DISCUSSION<br />

4.1 Tree and shrub condition declines.<br />

The health condition <strong>of</strong> trees and shrubs declines. In 2005, 52.5% <strong>of</strong> SYGs were<br />

damaged, while in 2008 it accounted for 82.6% (Table 1). The figures for shrubs<br />

were smaller than those for trees, but they also increased, from 25.1% in 2005 to<br />

58.1% in 2008 (Table 2). A rise in the number <strong>of</strong> the disturbances in vegetation is<br />

seen as the most likely cause in both cases. In fact, the number <strong>of</strong> these<br />

disturbances increased by four times in trees and three times in shrubs during the<br />

period. However, this higher level <strong>of</strong> disturbances is not attributed to more types <strong>of</strong><br />

disturbances, since these did not increase in tune with the others. Therefore, some<br />

disturbances have become more frequent in the last years.<br />

Regarding data collected in each district, there was a general decreasing<br />

tendency similar in the percentage <strong>of</strong> undamaged SYGs.<br />

Table 1. Temporary evolution <strong>of</strong> damaged tree SYGs. Undamaged, damaged and<br />

percentage <strong>of</strong> damaged tree SYGs, in each district, for each year, and in all the inventoried<br />

green spaces.<br />

Tree SYGs<br />

Undamaged<br />

2005 2006 2007 2008<br />

Damaged<br />

% <strong>of</strong> damaged<br />

Undamaged<br />

Damaged<br />

Arganzuela 33 33 50.0 29 39 57.4 23 47 67.1 10 56 84.8<br />

Barajas 25 24 49.0 20 46 69.7 13 56 81.2<br />

Hortaleza 45 51 53.1 47 50 51.5 16 96 85.7<br />

Retiro 19 11 36.7 14 12 46.2 7 21 75.0<br />

Salamanca 4 9 69.2 5 9 64.3 2 13 86.7<br />

Villa de<br />

Vallecas<br />

222<br />

% <strong>of</strong> damaged<br />

18 31 63.3 14 44 75.9 14 53 79.1<br />

TOTAL 144 159 52.5 129 200 60.8 23 47 67.1 62 295 82.6<br />

Undamaged<br />

Damaged<br />

% <strong>of</strong> damaged<br />

Undamaged<br />

Damaged<br />

% <strong>of</strong> damaged


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Table 2. Temporary evolution <strong>of</strong> damaged shrub SYGs. Undamaged, damaged and<br />

percentage <strong>of</strong> damaged shrub SYGs, in each district, in each year, and in all the inventoried<br />

green spaces.<br />

Shrub<br />

SYGs<br />

Undamaged<br />

2005 2006 2007 2008<br />

Damaged<br />

% <strong>of</strong> damaged<br />

Undamaged<br />

Damaged<br />

Arganzuela 53 10 15.9 49 30 38.0 55 30 35.3 29 60 67.4<br />

Barajas 38 13 25.5 21 42 66.7 33 34 50.7<br />

Hortaleza 79 52 39.7 68 56 45.2 62 92 59.7<br />

Retiro 17 3 15.0 17 14 45.2 18 10 35.7<br />

Salamanca 24 2 7.7 16 10 38.5 9 20 69.0<br />

Villa de<br />

Vallecas<br />

223<br />

% <strong>of</strong> damaged<br />

51 8 13.6 23 45 66.2 31 36 53.7<br />

TOTAL 262 88 25.1 194 197 50.4 55 30 35.3 182 252 58.1<br />

4.2 Few differences are encountered among the districts.<br />

There are few differences in the average percentage <strong>of</strong> damaged tree or shrub<br />

SYGs among the districts (Table 3). A ji-squared test shows that these figures are<br />

not significantly different at greater than the 90% level. This result was expected<br />

since there are not big differences in climate or in soil composition among the<br />

districts.<br />

Table 3. Mean percentage <strong>of</strong> damaged SYGs. Mean figures <strong>of</strong> undamaged, damaged<br />

and percentage <strong>of</strong> damaged tree SYGs, in each district, and in all the green spaces.<br />

SYGs<br />

Undamaged<br />

Tree mean<br />

2005-2008<br />

Damaged<br />

Undamaged<br />

Damaged<br />

% <strong>of</strong> damaged<br />

Undamaged<br />

Shrub mean<br />

2005-2008<br />

Arganzuela 24 44 64.8 47 33 41.1<br />

Barajas 19 42 68.5 31 30 49.2<br />

Hortaleza 36 66 64.6 70 67 48.9<br />

Retiro 13 15 52.4 17 9 34.2<br />

Salamanca 4 10 73.8 16 11 39.5<br />

Villa de Vallecas 15 43 73.7 35 30 45.9<br />

All green spaces 111 219 66.3 216 178 45.3<br />

% <strong>of</strong> damaged<br />

Undamaged<br />

Damaged<br />

Damaged<br />

% <strong>of</strong> damaged<br />

% <strong>of</strong> damaged


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Analysis <strong>of</strong> means shows that the districts which differ the most <strong>of</strong> the global<br />

mean are Retiro and Salamanca. In both <strong>of</strong> them, the areas <strong>of</strong> the studied green<br />

spaces reached lower figures than the rest <strong>of</strong> the districts. A possible explanation is<br />

that the bigger the area <strong>of</strong> green spaces observed is, the more representative <strong>of</strong> the<br />

district it becomes; however it is likely that an area threshold exists, so from this<br />

point the results will be stable.<br />

4.3 Main problem in trees is stem wounds.<br />

The major problem in trees is stem wounds, followed by dead branches,<br />

epicormic shoots and decay (Table 4).<br />

Table 4. Percentage <strong>of</strong> affected SYGs by 4 main disturbances. In each year, during the<br />

whole period and in all the inventoried green spaces.<br />

224<br />

% <strong>of</strong> SYGs<br />

Disturbances 2005 2006 2007 2008 All years<br />

Stem wounds 13.2 22.8 31.4 36.4 25.2<br />

Dead branches 10.2 8.2 18.6 33.6 18.0<br />

Epicormic shoots 4.3 12.2 1.4 31.9 15.9<br />

Decay 13.5 17.9 17.1 12.6 14.8<br />

Stem wounds is also the most important problem identified in some districts,<br />

such as Arganzuela, Barajas, Hortaleza and Villa de Vallecas. The cause <strong>of</strong> this<br />

high frequency was mostly unknown, but some SYGs with stem wounds had, at the<br />

same time, sunburns lesions (11.2%) or human damages (9.0%) like those which<br />

were caused by lawn mowers, vehicles, tree shelters, stakes, and so forth.<br />

Therefore, these disturbances may have caused some <strong>of</strong> the stem wounds.<br />

Stem wounds were also very common in several urban tree surveys (Rocray,<br />

1983; Jaenson et al., 1992; Chacalo et al., 1994, Fostad and Pedersen, 1994;<br />

Cumming et al., 2001; Ayuntamiento de Madrid, <strong>2009</strong>), besides, in the last three<br />

ones, the injuries were mainly caused by mechanized machinery and automobiles.<br />

Neither <strong>of</strong> them mentions any kind <strong>of</strong> sunburn lesion on the bark, nor the one<br />

conducted in street trees <strong>of</strong> Madrid.<br />

Another consideration is the importance <strong>of</strong> SYGs which were damaged by<br />

decay. More than 60% <strong>of</strong> the SYGs with decay had, at the same time, stem<br />

wounds, which showed a possible relationship between these two variables. The<br />

explanation is that wounds can be penetrated by organisms which produce decay,<br />

such us fungi (Agrios, 2005). Therefore, the abundance <strong>of</strong> these two factors could<br />

indicate bigger internal defects <strong>of</strong> the trees and, trees could turn into hazardous<br />

because some decays might end up in failure. Besides, the strength loss due to<br />

decay is greater when it is produced by peripheral wounds, as in this survey (Kane<br />

et al., 2001). Any conifer SYG was affected by decay, maybe because <strong>of</strong> the<br />

existence <strong>of</strong> resin, as it was explained by Rodríguez Barreal et al referring to<br />

cypress (Rodríguez Barreal et al., 2000).


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

4.4 Main problem in shrubs, vines and seasonal flowers is dead plants.<br />

The mayor problem is the high death rate. In shrubs, it is followed by aphids,<br />

powdery mildew and decaying plants.<br />

4.5 The most damaged tree species are Siberian elm, black locust, box elder<br />

and plane tree.<br />

Siberian elm (Ulmus pumila), black locust (Robinia pseudoacacia), box elder<br />

(Acer negundo) and plane tree (Platanus x hispanica) are the species which reach<br />

the greatest proportions <strong>of</strong> observed disturbances (Table 5). In fact, only the<br />

problems <strong>of</strong> 16 species accounted for 70% <strong>of</strong> the total. That would imply that some<br />

species are far more prone to problems; they would be the “key plants” (Raupp et<br />

al, 1985). Hence, these species require the biggest efforts in maintenance and<br />

conservation.<br />

Table 5. Disturbances for the 20 most affected tree and shrub species. Relative abundance,<br />

percentage <strong>of</strong> total disturbances and frequency <strong>of</strong> SYGs with disturbances in each species, for all the<br />

years and for all the inventoried green spaces. (*) accumulated percentage until that species.<br />

Tree species<br />

% <strong>of</strong><br />

total<br />

SYGs<br />

% <strong>of</strong> total<br />

disturbances<br />

% <strong>of</strong> SYGs<br />

with<br />

disturbances<br />

Ulmus pumila 6.3 11.0 83.6<br />

Robinia<br />

pseudoacacia<br />

6 8.9 85.9<br />

Acer negundo 4.7 8.2 88.0<br />

Platanus x<br />

hispanica<br />

Prunus<br />

cerasifera var.<br />

atropurpurea<br />

Populus alba<br />

var. fastigiata<br />

Tilia<br />

platyphillos<br />

4.5 6.8 85.4<br />

225<br />

Shrub species<br />

Nerium<br />

oleander<br />

Cotoneaster<br />

sp.<br />

Pittosporum<br />

tobira<br />

Eonymus<br />

europaeus<br />

% <strong>of</strong><br />

total<br />

SYGs<br />

% <strong>of</strong> total<br />

disturbances<br />

% <strong>of</strong> SYGs<br />

with<br />

disturbances<br />

5.3 10.5 73.1<br />

6 7.5 60.0<br />

4.8 6.6 55.0<br />

3 6.2 84.2<br />

6.5 5.3 71.0 Rosa sp. 4.8 5.7 60.7<br />

2.2 4.8 95.7<br />

Viburnum<br />

tinus<br />

5.5 5.2 42.0<br />

2.1 3.8 81.8 Pyracantha sp. 3.8 4.8 58.3<br />

Cedrus sp. 4.7 3.6 66.0 Berberis sp. 2.1 4.1 70.4<br />

Populus nigra 2.1 3.6 100.0<br />

Morus alba 2.8 3.4 73.3<br />

Sophora<br />

japonica<br />

Rosmarinus<br />

<strong>of</strong>ficinalis<br />

Mahonia<br />

aquifolia<br />

4.4 3.7 49.1<br />

2.2 3.0 50.0<br />

2.5 3.1 77.8 Laurus nobilis 1.3 2.9 70.6<br />

Acer sp. 2 3.0 85.7 Juniperus sp. 4.5 2.7 36.8<br />

Cupressus sp. 6.4 2.7 38.2<br />

Olea europaea 1.9 2.6 70.0<br />

Gleditsia<br />

triacanthos<br />

1.8 2.4 84.2<br />

Photinia<br />

serrulata<br />

Lavandula<br />

latifolia<br />

Cotoneaster<br />

horizontalis<br />

2.3 2.4 44.8<br />

3.3 2.3 41.5<br />

2.5 2.0 35.5<br />

Pinus pinea 4.7 2.2 (75.5% * ) 46.0 Arbutus unedo 1.6 1.9 (71.7% * ) 35.0<br />

Eleagnus<br />

angustifolia<br />

0.9 2.2 100.0<br />

Populus alba 1.5 2.0 75.0<br />

Cercis<br />

siliquastrum<br />

1.2 1.9 92.3<br />

Salix sp. 1.4 1.7 93.3<br />

Spirea<br />

hypericifolia<br />

Hibiscus<br />

syriacus<br />

Prunus<br />

laurocerasus<br />

Escallonia<br />

rubra<br />

1.3 1.8 52.9<br />

1.5 1.8 52.6<br />

2.5 1.8 34.4<br />

2.3 1.7 41.4


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

There are some differences in health condition among tree species. For<br />

example, Siberian elm and black locust are very damaged as well as very abundant,<br />

whereas cypress (Cupressus sp.), umbrella pine (Pinus pinea) and cedar (Cedrus<br />

sp.) are very frequent but not so damaged.<br />

The most important problem for Siberian elm is bad shaped trees. This term is<br />

used to describe the plants with bad structured crowns or trunks, for example, those<br />

trees with hanging small branches or leaning stems. Hanging branches seem to be<br />

very common in this species (Saiz de Omeñaca and Prieto Rodríguez, 2004).<br />

Dead branches are frequent for black locust, and this was also showed in<br />

another survey (Rodríguez Barreal et al., 2000).<br />

Box elder has stem wounds as its main problem. If this problem was associated<br />

with decay it could make trees fail, since the wood <strong>of</strong> box elder is very susceptible<br />

to decay and has a practical inability to close its wounds (Saiz de Omeñaca and<br />

Prieto Rodríguez, 2004). Powdery mildew is found as well in some green spaces,<br />

which is ordinary in this species (Saiz de Omeñaca and Prieto Rodríguez, 2004).<br />

Anthracnose is observed in most <strong>of</strong> the green spaces where plane trees are<br />

grown. This fungal disease is produced by Apiognomonia veneta (Sporonema<br />

platani) and their first symptoms are not identified until the beginning <strong>of</strong> the 1970s<br />

in central Spain (Tello et al., 2000). Since that moment, the presence <strong>of</strong> the fungi<br />

has been known about in street trees in Madrid (Rodríguez Barreal, 1986). It<br />

remains unclear whether it affects less its precursor Plantanus orientalis (Villalva<br />

Quintana, 2005) than to the own hybrid (Rodríguez Barreal, 1986), although it<br />

seems that “the different genotypes <strong>of</strong> the hybrids allow differences in disease<br />

severity among trees growing under the same environmental conditions, though<br />

eventually all trees are affected to some degree or another” (Tello et al., 2000).<br />

The abundance <strong>of</strong> epicormic shoots and adventitious branches (3 rd and 4 th problem)<br />

could be related to this disease, since the infestation <strong>of</strong> the fungi in adventitious<br />

buds produces a much greater number <strong>of</strong> them (Tello et al.2000), however, it could<br />

be also associated with the tendency to create them when the planting conditions <strong>of</strong><br />

plane trees are not appropriate (Saiz de Omeñaca and Prieto Rodríguez, 2004).<br />

4.6 The most damaged shrub species are oleander, cotoneaster, Japanese<br />

cheesewood and European spindle.<br />

Oleander (Nerium oleander), cotoneaster (Cotoneaster sp.), Japanese<br />

cheesewood (Pittosporum tobira) and European spindle (Euonymus europaeus) are<br />

the most damaged species (Table 5).<br />

Main problem <strong>of</strong> oleander is the bacteria Pseudomonas syringae subsp.<br />

savastanoi, which affected about half <strong>of</strong> green spaces every year. The former<br />

produces galls in flower buds, so those plants become weaker and with aesthetic<br />

harms (Villalva Quintana, 2005). Dead plants are the principal problem affecting<br />

cotoneaster and Japanese cheesewood, the second regarding European spindle, and<br />

the third concerning oleander. Aphids were in more than 24% <strong>of</strong> SYGs <strong>of</strong> oleander,<br />

226


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

cotoneaster and Japanese cheesewood, and their importance was lower in European<br />

spindle, with an affected 10.5%. The latter species was affected by powdery<br />

mildew in more than 70% <strong>of</strong> the SYGs where European spindle was found. This<br />

disease is induced by the fungi Microsphaera euonymi-japonici and it has a high<br />

incidence among this species in cold areas <strong>of</strong> the Iberian Peninsula (Villalva<br />

Quintana, 2005).<br />

4.7 The most damaged vine species is ivy.<br />

Ivy (Hedera helix) is the most affected vine species as well as the most<br />

abundant vine. Its most frequent problems are dead plants and sunburn lesions on<br />

its leaves. Sunburn lesions appeared because, even though it is a shade tolerant<br />

species which grows in the understorey layer in nature, in many <strong>of</strong> the cases it grew<br />

in green spaces under full sun situations.<br />

4.8 Biotic diseases are encountered less frequently than abiotic in trees,<br />

although the result was the opposite in shrubs.<br />

In this survey, biotic diseases were due to organisms such as viruses, bacteria,<br />

pests and fungi. On the contrary, abiotic diseases consisted <strong>of</strong> problems caused by<br />

meteorological and physical agents, by human activities and, by other factors<br />

whose origin is a priori unknown, such as cracks, stem wounds, tumours, etc. Dead<br />

plants were not included in any <strong>of</strong> the former types.<br />

In trees, abiotic diseases comprised about 60% <strong>of</strong> total disturbances in each year<br />

(Table 6). The mayor contributions to the high incidence <strong>of</strong> abiotic diseases are due<br />

to stem wounds, dead branches and epicormic shoots, since they are the most<br />

frequent problems for trees. Similar results were also obtained in a survey in<br />

Quebec (Rocray, 1983).<br />

Table 6. Percentage <strong>of</strong> abiotic diseases <strong>of</strong> the total <strong>of</strong> disturbances. In each year in trees<br />

and shrubs <strong>of</strong> all the inventoried green spaces.<br />

% <strong>of</strong> abiotic diseases 2005 2006 2007 2008<br />

Trees 58.8 69.1 56.6 79.7<br />

Shrubs 19.7 47.9 27.6 23.7<br />

The principal problem in green spaces is the abiotic stresses when they are<br />

compared with biotic diseases. This fact should lead to the consideration <strong>of</strong> the<br />

maintenance and conservation practise which is being applied mainly to the<br />

vegetation <strong>of</strong> green spaces, because most damages could be avoided if some<br />

recommendations were followed. Various efforts were observed about this issue in<br />

this survey, for instance, the use <strong>of</strong> stakes and tree shelters in new plantations,<br />

however, in many cases, the achieved effect was just the opposite <strong>of</strong> the expected<br />

one, since stems were leaned and the friction between the trunk and the object<br />

caused the wounds which were intended to be avoided. Not only the staff in charge<br />

<strong>of</strong> the conservation <strong>of</strong> green spaces are responsible for these problems, but also the<br />

227


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

designers or planners <strong>of</strong> the green spaces, since some disturbances such as sunburn<br />

lesions or leaning stems by light deficit are generally due to inadequate designs. In<br />

fact, in natural environment most <strong>of</strong> these problems are very unusual. For all these<br />

reasons, the municipal green spaces management programmes should be look<br />

through in order to adapt conservation strategies to deal better with abiotic<br />

problems.<br />

Besides, abiotic diseases were much less frequent in shrubs, with 20-48% <strong>of</strong><br />

total disturbances depending on the considered year. Aphids were responsible <strong>of</strong> a<br />

great number <strong>of</strong> the biotic diseases, since it was the second problem, after dead<br />

plants. However, this result would have been pretty different if dead plants had<br />

been considered as an abiotic or biotic agent.<br />

4.9 Abiotic diseases affect a greater rank <strong>of</strong> tree and shrubs species.<br />

Abiotic diseases affect a greater variety <strong>of</strong> species (Table 7), which is<br />

reasonable if it is taken into account the fact that they are prompted by human<br />

activities or physical agents. However, biotic diseases seemed to be more specific<br />

in some plant species; for example, lace bugs (Corythuca ciliata), leaf beetle<br />

Galerucella luteola and Pseudomonas syringae subsp. savastanoi affected only one<br />

species (plane tree, Siberian elm and oleander, respectively). Concerning pests, this<br />

result would confirm the idea held by some authors who maintain that most<br />

herbivorous pests are <strong>special</strong>ized and only eat few taxa <strong>of</strong> plants (Rocray, 1983;<br />

Galvin, 1999; Raupp et al., 2001).<br />

Table 7. Affected species by biotic and abiotic diseases. Tree and shrub species affected<br />

by biotic and abiotic diseases and the percentage <strong>of</strong> the total species, in all the years and in<br />

all the inventoried green spaces.<br />

Trees Total<br />

Affected by<br />

biotic<br />

diseases<br />

Affected by<br />

abiotic<br />

diseases<br />

228<br />

Shrubs Total<br />

Affected by<br />

biotic<br />

diseases<br />

Affected by<br />

abiotic<br />

diseases<br />

Species 73 38 57 Species 77 32 40<br />

% <strong>of</strong><br />

species<br />

100 52.1 78.1<br />

% <strong>of</strong><br />

species<br />

100 41,6 51,9<br />

4.10 Positive association between naturalness <strong>of</strong> tree species and their<br />

health status.<br />

Pyšek’s criteria are used to define the naturalness <strong>of</strong> the tree (Pyšek , 1995). It is<br />

considered the national scale, therefore, all those species defined as natural <strong>of</strong><br />

Spain according to those criteria, are classified as native in this survey.<br />

Bibliography is used to make easier this controversial task (López González, 2002;<br />

Real Jardín Botánico, <strong>2009</strong>).<br />

A ji-squared test on the data shows that the relationship between health status<br />

(damaged or undamaged SYG) and naturalness (native or exotic SYG) is<br />

significant at 95% level (Table 8). The probability for an exotic SYG to be<br />

damaged is greater (0.68) than the probability for a native SYG (0.59). However,


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

although these proportions were significantly different for a hypothesis test at 95%<br />

level, with a greater level, they were not different. Thus, the risk <strong>of</strong> being damaged<br />

is only slightly higher for exotic SYGs than for native SYGs.<br />

Table 8. 2x2 contingency table <strong>of</strong> tree SYGs.<br />

SYGs<br />

Naturalness<br />

229<br />

Health status<br />

Undamaged Damaged<br />

Native 92 134<br />

Exotic 266 567<br />

This seems to be a coherent result since native species are expected to be better<br />

adapted, although there are authors who maintain that exotic species do better than<br />

native ones because their pests and diseases have not yet arrived from their home<br />

country (Schimdt and Kerenyine-Nemstothy, 1999). Similarly, Harris explains that<br />

native species sometimes do not perform as well as exotic ones (Harris et al.,<br />

2004). Therefore, as it remains unclear what kind <strong>of</strong> species should be planted<br />

more frequently, it should be consider that exotic flora affects native and overall<br />

species richness throughout the globe (Alvey, 2006) while native species improve<br />

the sustainability <strong>of</strong> urban forests (Clark et al., 1997). In any case, the<br />

recommendation <strong>of</strong> not planting invasive exotic species should be followed (Alvey,<br />

2006).<br />

4.11 Ranking <strong>of</strong> less suitable species to be grown in Madrid.<br />

An algorithm to establish which tree and shrub species are less adapted to the<br />

urban environment <strong>of</strong> Madrid is created. On one hand, a quantitative scale is<br />

established to asses the severity <strong>of</strong> the observed disturbances, assigning greater<br />

coefficients to the most serious disturbances (Table 9). On the other, a qualitative<br />

scale takes into consideration: a) the number <strong>of</strong> disturbances which appeared<br />

within each species, and b) the proportion <strong>of</strong> damaged SYGs. The result <strong>of</strong><br />

multiplying each severity coefficient by the number <strong>of</strong> disturbances (a) is weighted<br />

with the proportion <strong>of</strong> damaged SYGs (b) within each species. The outcome is a<br />

ranking with the least suitable species, where the species with the highest grade<br />

appear on top (Table 10 and Table 11).<br />

Table 9. Severity classes <strong>of</strong> disturbances with coefficients.<br />

Severity classes <strong>of</strong> disturbances Coefficients<br />

Hazardous 0.5<br />

Pretty serious 0.3<br />

Less serious 0.2<br />

Acceptable 0.1<br />

This kind <strong>of</strong> species suitability ranking has been observed in other surveys<br />

(Impens and Delcarte, 1979; Raupp and Noland, 1984; Nielsen et al., 1985; Ball,


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

1987; Rodríguez Barreal et al., 2000), although in some <strong>of</strong> them the way in which<br />

the results are obtained is not specified, so comparison is difficult. In spite <strong>of</strong> this,<br />

in the research carried out in the street trees <strong>of</strong> Madrid (Rodríguez Barreal et al.,<br />

2000), black locust appears as not very desirable species. Otherwise, this was<br />

considered very advisable species in another survey (Raupp and Noland, 1984)<br />

since it did not suffer from pest problems unlike the present survey, where many<br />

black locust SYGs were affected by aphids and wood borer insects.<br />

Table 10. Ranking <strong>of</strong> ten least suitable tree and shrub species. Percentage <strong>of</strong> damaged<br />

SYGs, severity <strong>of</strong> disturbances (coefficient*number <strong>of</strong> disturbances) and final value, in all<br />

the years and in all the inventoried green spaces.<br />

Order<br />

1<br />

Tree species<br />

Robinia<br />

pseudoacacia<br />

% <strong>of</strong><br />

damaged<br />

SYGs<br />

Severity <strong>of</strong><br />

disturbances<br />

valuation<br />

Final<br />

value<br />

230<br />

Tree species<br />

% <strong>of</strong><br />

damaged<br />

SYGs<br />

Severity <strong>of</strong><br />

disturbances<br />

valuation<br />

Final<br />

value<br />

85,9 64,3 55,3 Nerium oleander 73,1 20,6 15,1<br />

2 Ulmus pumila 83,6 64,0 53,5 Cotoneaster sp. 60,0 20,9 12,5<br />

3 Acer negundo 88,0 47,0 41,4<br />

4<br />

5<br />

Platanus x<br />

hispanica<br />

Populus alba var.<br />

fastigiata<br />

Eonymus<br />

europaeus<br />

84,2 12,2 10,3<br />

85,4 37,6 32,1 Pyracantha sp. 58,3 16,4 9,6<br />

95,7 32,0 30,6<br />

Pittosporum<br />

tobira<br />

55,0 16,7 9,2<br />

6 Populus nigra 100,0 24,2 24,2 Rosa sp. 60,7 14,6 8,9<br />

7<br />

Prunus cerasifera<br />

var. atropurpurea<br />

71,0 27,3 19,4<br />

8 Tilia platiphyllos 81,8 23,2 18,0<br />

Berberis<br />

thunbergii<br />

‘Atropurpurea’<br />

Rosmarinus<br />

<strong>of</strong>ficinalis<br />

70,4 11,8 8,3<br />

49,1 14,5 7,1<br />

9 Sophora japonica 77,8 22,0 18,0 Viburnum tinus 42,0 16,6 7,0<br />

10 Acer sp. 85,7 19,8 17,0 Mahonia aquifolia 50,0 7,9 4,0<br />

5. REFERENCES<br />

Agrios, G.N., 2005. Plant Pathology. 5 th Ed. San Diego, Ca: Elsevier Academic Press Publications,<br />

922 pp. ISBN 0-12-044565-4, p. 88<br />

Alvey, A.A., 2006. Promoting and preserving biodiversity in the urban forest. Urban Forestry &<br />

Urban Greening, 5, 195-201<br />

Ball, J., 1987. Efficient monitoring for an urban IPM program. Journal <strong>of</strong> Arboriculture, 13:7, 174-<br />

177<br />

Berrang, P., Karnosky, D.F. and Stanton, B.J., 1985. Environmental factors affecting tree health in<br />

New York City. Journal <strong>of</strong> Arboriculture, 11:6, 185-189<br />

Chacalo, A., Aldama A., Grabinsky, J., 1994. Street tree inventory in Mexico City. Journal <strong>of</strong><br />

Arboriculture, 20:6, 222-226<br />

Clark, J.R. et al., 1997. A model <strong>of</strong> urban forest sustainability. Journal <strong>of</strong> Arboriculture, 23:1, 17-30<br />

Cumming, A.B. et al., 2001. Forest Health Monitoring Protocol applied to roadside trees in Maryland.<br />

Journal <strong>of</strong> Arboriculture, 27:3, 126-138


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Dunster, A.J., 1996. Hazard tree assessments: developing a species pr<strong>of</strong>ile for western hemlock.<br />

Journal <strong>of</strong> Arboriculture, 22:1, 51-57<br />

Fostad, O. and Pedersen, P.A., 1997. Vitality, variation, and causes <strong>of</strong> decline <strong>of</strong> trees in Oslo Center<br />

(Norway). Journal <strong>of</strong> Arboriculture, 23:4, 155-165<br />

Galvin, M.F., 1999. A methodology for assessing and managing biodiversity in street tree<br />

populations: a case <strong>of</strong> study. Journal <strong>of</strong> Arboriculture, 25:3, 124-128<br />

Harris, R.W., Clark, J.R., Matheny, N.P., 2004. Arboriculture: Integrated management <strong>of</strong> landscape<br />

trees, shrubs and vines. Fourth <strong>edition</strong>. Prentice Hall, pp. 123, 406, 408<br />

Impens, R.A., Delcarte, E., 1979. Survey <strong>of</strong> urban trees in Brussels, Belgium. Journal <strong>of</strong><br />

Arboriculture, 5:8, 169-176<br />

Jaenson, R. et al., 1992. A statistical method for the accurate and rapid sampling <strong>of</strong> urban street tree<br />

populations. Journal <strong>of</strong> Arboriculture, 18:4, 171-183<br />

Kane, B., Ryan D., Bloniarz, D.V., 2001. Comparing formulae that assess strength loss due to decay<br />

in trees. Journal <strong>of</strong> Arboriculture, 2001, 27:2, 78-87<br />

Kennard, D.K.; Putz, F.E.; Niederh<strong>of</strong>er, M., 1996. The predictability <strong>of</strong> tree decay based on visual<br />

assessment. Journal <strong>of</strong> Arboriculture, 22:5, 249-254<br />

Kozlowski, T.T., 1985. Tree growth in response to environmental stresses. Journal <strong>of</strong> Arboriculture,<br />

11:4, 97-111<br />

López González, G., 2002. Guía de los árboles y arbustos de la Península Ibérica y Baleares. Madrid:<br />

Ediciones Mundiprensa, 894 pp. ISBN: 84-8476-050-2<br />

Nielsen, D.G. et al., 1985. Common street trees and their pests problems in the North Central United<br />

States. Journal <strong>of</strong> Arboriculture, 11:8, 225-232<br />

Proyecto Madrid: un alcorque, un árbol, <strong>2009</strong> [on line]. Ayuntamiento de Madrid. Área Temática de<br />

Medio Ambiente. Madrid. [19 th <strong>of</strong> February, <strong>2009</strong>].<br />

<br />

Pyšek, P., 1995. Alien and native species in Central European floras: a quantitative comparison.<br />

Journal <strong>of</strong> Biogeography, 25, 155-163<br />

Raupp, M.J. and Noland, R.M., 1984. Implementing landscape plant management programs in<br />

institutional and residential settings. Journal <strong>of</strong> Arboriculture, 10:6, 161-169<br />

Raupp, M.J. et al., 1985. The concept <strong>of</strong> Key plants in Integrated Pest Management for landscapes.<br />

Journal <strong>of</strong> Arboriculture, 11:11, 317-322<br />

Raupp, M.J. et al., 2001. Plant species diversity and abundance affects the number <strong>of</strong> arthropod pests<br />

in residential landscape. Journal <strong>of</strong> Arboriculture, 27:4, 222-229<br />

Real Jardín Botánico. Flora Ibérica [on line]. Centro Superior de Investigaciones Científicas. 4th <strong>of</strong><br />

February, <strong>2009</strong>, [8th <strong>of</strong> March, <strong>2009</strong>]. <br />

Rocray, D., 1983. Problems affecting urban trees in Quebec City. Journal <strong>of</strong> Arboriculture, 9:6, 167-<br />

169<br />

Rodríguez Barreal, J.A. et al., 2000. Estudio fitopatológico del arbolado de alineación urbana de<br />

varios distritos de Madrid. In: Comunicaciones XXVII Congreso nacional de parques y<br />

jardines públicos, 22-24 November 2000, Sevile, Spain, pp. 213-223<br />

231


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Rodríguez Barreal, J.A., 1986. Principales enfermedades sobre el Platanus hybrida Brot. en las<br />

grandes ciudades. Montes. Revista de Ámbito Forestal, 12, 43-50<br />

Saiz de Omeñaca, J. A. y Prieto Rodríguez, A., 2004. Arboricultura Urbana. Centro de Estudios De<br />

Técnicas Aplicadas del CEDEX, pp. 53-57<br />

Schimdt, G. and Kerenyine-Nemstothy, K., 1999. Urban Tree Health in Budapest. In: Lemattre, M.,<br />

Lemattre, P., Lemaire, F. (Eds.), Proceedings <strong>of</strong> the International Symposium on Urban<br />

Tree Health, 1 September 1999, Paris, France, pp. 183-188<br />

Smiley, E.T., and Baker, F.A., 1988. Options in street tree inventories. Journal <strong>of</strong> Arboriculture, 14:2,<br />

36-42<br />

Tello, M.L., Redondo, C. and Mateo-Sagasta E., 2000. Health status <strong>of</strong> plane trees (Platanus ssp.) in<br />

Spain. Journal <strong>of</strong> Arboriculture, 26:5, 246-254<br />

Villalva Quintana, S., 2005. Plagas y enfermedades de jardines. Editorial Mundi-Prensa, pp. 110, 132<br />

232


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 233-237<br />

CHARACTERISATION OF CZECH Ophiostoma novo-ulmi ISOLATES<br />

Milon DVOŘÁK 1* , Libor JANKOVSKÝ 1 , J . KRAJŇÁKOVÁ 2<br />

1 Department <strong>of</strong> Forest Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood<br />

Technology, Mendel University <strong>of</strong> Agriculture and Forestry, Brno, Zemedelska 3, 613 00 Brno,<br />

Czech Republic<br />

2 Department <strong>of</strong> Biology and Plant Protection, Unit <strong>of</strong> Plant Biology, University <strong>of</strong> Udine, Via<br />

Cotonificio 108, 33100 Udine, Italy<br />

ABSTRACT<br />

* klobrc@centrum.cz<br />

Ophiostoma novo-ulmi was recorded for the first time in the area <strong>of</strong> the Czech Republic<br />

in 2006, although it was suspected since early 1960´s. During the years 2006 and 2007, 58<br />

isolates were collected. Isolated strains <strong>of</strong> the DED causal agent were determined by PCR<br />

and RFLP molecular methods. DNA <strong>of</strong> each sample was isolated from cultured mycelium,<br />

amplified at cu, and col1 gene regions and restricted by Hph I and Bfa I endonucleases.<br />

Strains were determined according to both <strong>of</strong> these two gene regions. An old species<br />

Ophiostoma ulmi (Buism.) Nannf. was not noticed in any case. Every sample belongs to the<br />

species Ophiostoma novo-ulmi Bras., and both known subspecies americana (5 isolates)<br />

and novo-ulmi (29 isolates) were present. Additionally, 13 strains <strong>of</strong> Ophiostoma novo-ulmi<br />

were more particularly tested for vegetative compatibility type, mating type, fertility with<br />

both subspecies and cerato-ulmin production. Among these 13 strains only 3 seemed to be<br />

non-hybridized ssp. novo-ulmi and only one non-hybridized ssp. americana, the remainder<br />

(9 strains) was intraspecific hybrids. 5 <strong>of</strong> the tested strains were <strong>of</strong> mating type A and<br />

mating type B occurred 8-times. None <strong>of</strong> these strains was compatible with any other. High<br />

frequency <strong>of</strong> intraspecific hybrids (24 isolates) is remarkable and shows on a frequent<br />

occurrence <strong>of</strong> sexual hybridization. Cerato-ulmin production did not show any significant<br />

differences according to strain subspecies. Virulence <strong>of</strong> one O. novo-ulmi strain (M3) was<br />

tested on two Dutch resistant cultivars, 13-year-old elms ´Groeneveld´ and ´Dodoens´. Only<br />

5 weeks after inoculation were sufficient period for 30 – 90% defoliation <strong>of</strong> cultivar<br />

´Groeneveld´, ´Dodoens´ with only 0 – 5% seemed to be much more resistant.<br />

Keywords: Dutch Elm Disease, PCR-RFLP, elm, infection tests<br />

1. INTRODUCTION<br />

Ophiostoma ulmi and O. novo-ulmi – causal agents <strong>of</strong> Dutch elm disease (DED)<br />

has been threatening elms in the Czech Republic since 1930's, as well as in the<br />

entire Northern Hemisphere. The first occurrence <strong>of</strong> Dutch elm disease in the<br />

Czech Republic caused by Ophiostoma ulmi (Buism.) Nannf. was noted by pr<strong>of</strong>.<br />

Peklo (Polák, 1932) who found infected trees in elm alleys in Prague and<br />

Poděbrady in 1932.<br />

233


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

A newer and more aggressive form was since 1970´s distinguished into two<br />

races - an Eurasian (EAN), probably originated in Moldavia and the Ukraine, and a<br />

North American race isolate (NAN). In 1991, the more aggressive form was<br />

described by Brasier as a new species O. novo-ulmi Bras (Brasier 1991). Ten years<br />

later, Brasier et al. (2001) designated races EAN and NAN as subspecies <strong>of</strong> O.<br />

novo-ulmi ssp. novo-ulmi and O. novo-ulmi ssp. americana. Specific group <strong>of</strong> DED<br />

pathogens are hybrids <strong>of</strong> O. novo-ulmi subspecies.<br />

O. novo-ulmi, its subspecies and their hybrids were for the first time recorded<br />

in the Czech Republic by Dvořák et al. (2006, 2007) . The aim <strong>of</strong> present study is<br />

a complex characterisation <strong>of</strong> strains, which are actually spreading in the area <strong>of</strong><br />

the Czech Republic.<br />

2. MATERIALS AND METHODS<br />

2.1. Isolation <strong>of</strong> the pathogen: Coordinates <strong>of</strong> position and altitude <strong>of</strong> every<br />

infected elm with inner DED symptoms was achieved by GPS. Samples <strong>of</strong> twigs<br />

were cut from different parts <strong>of</strong> drying crowns and cultivated on cycloheximide 3%<br />

MEA medium. Each strain was deposited into the culture collection <strong>of</strong> MUAF<br />

Brno.<br />

2.2. Molecular-biological characterisation: DNA was obtained from pure<br />

MEA cultures by PowerSoil DNA Kit (Mo Bio). Identification <strong>of</strong> species and<br />

subspecies was performed by the mean <strong>of</strong> Polymerase Chain Reaction (PCR) and<br />

Restriction Fragment Length Polymorphism (RFLP) using two gene regions. The<br />

former is a cerato-ulmin (cu) gene region, described in more detail in Pipe et al.<br />

(1997) and the latter gene region is called col1 and encodes colony type (Konrad et<br />

al., 2002). Amplified fragments <strong>of</strong> these two gene regions were restricted by<br />

endonucleases, Hph I used for cu gene region and Bfa I for col1 region (Konrad et<br />

al., 2002). RFLP fragments were visualized on 3% agarose gel and evaluated.<br />

Sequencing was provided in a few cases and submitted to Gen Bank.<br />

2.3. Production <strong>of</strong> cerato-ulmin: Amount <strong>of</strong> cerato-ulmin produced by liquid<br />

cultures <strong>of</strong> 15 isolates was determined by chromatography (Scala et al., 1994).<br />

The tested group <strong>of</strong> isolates was composed from 4 cultures <strong>of</strong> ssp. novo-ulmi, 2<br />

cultures <strong>of</strong> ssp. americana and 9 strains <strong>of</strong> genetic hybrids <strong>of</strong> precedent strains. In<br />

this tested group, well known isolates H328 ssp. novo-ulmi and 182E ssp.<br />

americana were used as reference strains. Liquid cultures <strong>of</strong> dilutions 1:1, 1:2, 1:4<br />

and 1:8 were stirred to obtain turbid milky solutions. The turbidity <strong>of</strong> the samples<br />

at the various dilutions was immediately assayed for optical density at 400 nm.<br />

Results <strong>of</strong> the dilution with the most variable values represent the best<br />

measurement for comparing the cu-production <strong>of</strong> the strains.<br />

2.4. Mating type, vegetative compatibility and fertility tests: General<br />

characterisation <strong>of</strong> 13 isolates has been performed according to Brasier (1981). A<br />

<strong>special</strong> medium – Elm Sapwood Agar was used. Occurrence <strong>of</strong> elm sapwood in the<br />

234


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

medium is necessary for obtaining synnemata and perithecia, essentially important<br />

for classifying into mating type (A or B), vegetative compatibility groups and<br />

fertility tests (identification <strong>of</strong> subspecies).<br />

2.5. Infection tests: Isolate M3 was inoculated into sapwood <strong>of</strong> 10 living elms<br />

which represents in vitro cloned progenies <strong>of</strong> two elm hybrids – ´Groeneveld´ and<br />

´Dodoens´ (Krajnakova and Longauer, 1996). The experimental plot is situated<br />

near to Banská Štiavnica (coordinates: N 48°28´ E 18°58´, 590 m a. s. l.) 5<br />

representative trees per each hybrid were inoculated and observed according to<br />

Solla et al. (2005).<br />

3. RESULTS<br />

3.1. Ratios <strong>of</strong> O. novo-ulmi subspecies: Altogether 58 isolates were achieved<br />

from nearly the entire area <strong>of</strong> the Czech Republic. The isolated strains were<br />

identified as O. novo-ulmi and both subspecies novo-ulmi (29) and americana (5)<br />

were found, as well as their hybrids (24). Decreasing trend in occurrence <strong>of</strong><br />

Ophiostoma ulmi confirms the results <strong>of</strong> other authors (Brasier, 1991; Hoegger et<br />

al., 1996; Konrad et al., 2002). They suppose that Ophiostoma ulmi disappeared by<br />

the end <strong>of</strong> the 1970's.<br />

3.2. Production <strong>of</strong> cerato-ulmin: Absorbances <strong>of</strong> the samples were very<br />

variable. Dilution with most variable results was 1:4. Statistical analysis (z-test)<br />

did not show any differences in cerato-ulmin production between subspecies and<br />

hybrids. Interesting exception is M9, which produces no cerato-ulmin, although it<br />

was isolated from infected branch. Culture <strong>of</strong> M9 seems to be d-infected.<br />

3.3. Mating type, vc-type and fertility response variability: Among 13<br />

isolates investigated both mating types occur; 9 isolates are <strong>of</strong> B type, 4 <strong>of</strong> A type.<br />

Vc-tests showed on high diversity <strong>of</strong> strains. Any isolate was not compatible with<br />

any other, although two <strong>of</strong> them were collected from the same tree. Fertility tests<br />

revealed 3 subspecies americana and 10 ssp. novo-ulmi. In comparison with<br />

identification by molecular tools there is some relation with cu-region, isolate M10<br />

is the only exception.<br />

3.4. Infection tests: Isolate M3 caused extremely high defoliation on resistant<br />

elm hybrid ´Groeneveld´ and low defoliation on the hybrid ´Dodoens´.<br />

4. DISCUSSION AND CONCLUSION<br />

Methods which have been used proved to be useful in the recognition <strong>of</strong><br />

diversity in DED causal agent populations. From the distribution <strong>of</strong> isolates is<br />

evident that we cannot draw any borders between areas which are occupied only by<br />

ssp. novo-ulmi or by ssp. americana. In many cases such areas were observed<br />

where both subspecies and their hybrids are overlapping and where more than one<br />

strain <strong>of</strong> pathogen is responsible for dying <strong>of</strong> a tree. This phenomenon was<br />

235


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

concluded by negative vc-type tests <strong>of</strong> two isolates from one tree. This is<br />

confirmed by many authors (Brasier et al., 1998; Santini, unpublished). High<br />

variability <strong>of</strong> mating types and subspecies is an evidence <strong>of</strong> postepidemic phase <strong>of</strong><br />

DED (Konrad et al., 2002; Brasier et al., 2004). Differences in cerato-ulmin<br />

production among subspecies which are described by Brasier et al. (2001) and<br />

others were not confirmed. Production <strong>of</strong> cerato-ulmin by hybrids and ssp. novoulmi<br />

is not different, production by ssp. americana is also not different, but there<br />

were only two isolates ssp. americana tested. Infection tests on living trees are<br />

useful for assessment <strong>of</strong> the resistance <strong>of</strong> each species or elm hybrid and are much<br />

more reliable by use <strong>of</strong> older plant material than common 3-year-old trees. (Solla et<br />

al., 2005). Cultivar ´Groeneveld´ showed very low resistance against currently<br />

spread strain <strong>of</strong> Ophiostoma novo-ulmi.<br />

5. ACKNOWLEDGEMENTS<br />

The author wishes to thank Drs. Alberto Santini and Alejandro Solla for<br />

sharing their experience with laboratory and field work.<br />

The results presented herein were obtained from projects supported by MSM<br />

6215648902.<br />

6. REFERENCES<br />

Brasier, C. M., 1981. Laboratory investigation <strong>of</strong> Ceratocystis ulmi, Compendium <strong>of</strong> Elm Diseases, 76<br />

– 79.<br />

Brasier, C. M., 1991. Ophiostoma novo-ulmi sp. nov., Causative Agent <strong>of</strong> Current Dutch Elm Disease<br />

Pandemics. Mycopathologia, 115(3), 151 – 161.<br />

Brasier, C. M., Kirk, S. A., Pipe, N. D., et al., 1998. Rare interspecific hybrids in natural populations<br />

<strong>of</strong> the Dutch elm disease pathogens Ophiostoma ulmi and O. novo-ulmi. Mycol. Res.,<br />

102(1), 45 – 57.<br />

Brasier, C. M., Kirk S. A., 2001. Designation <strong>of</strong> the EAN and NAN Races <strong>of</strong> Ophiostoma novo-ulmi<br />

as Subspecies. Mycological Research, 105, 547 - 554.<br />

Brasier, C., Buck, K., Paoletti, M., Crawford, L., Kirk, S., 2004. Molecular analysis <strong>of</strong> evolutionary<br />

changes in populations <strong>of</strong> Ophiostoma novo-ulmi. Invest. Agrar.: Sist. Recur. For., 13(1),<br />

93 – 103.<br />

Dvořák, M., Palovčíková, D., Jankovský, L., 2006. The occurrence <strong>of</strong> endophytic fungus<br />

Phomopsis oblonga. Journal <strong>of</strong> Forest Science, 52(11), 531 – 535.<br />

Dvořák, M., Tomšovský, M., Novotný, L. et al., 2007. Contribution to identify the causal agents <strong>of</strong><br />

Dutch elm disease in the Czech Republic. Plant Protection Science, 43(4), 142 - 145.<br />

Hoegger, P. J., Binz, T., Heiniger, U., 1996. Detection <strong>of</strong> genetic variation between Ophiostoma ulmi<br />

and the NAN and EAN races <strong>of</strong> O. novo-ulmi in Switzerland using RAPD markers.<br />

European Journal <strong>of</strong> Forest Pathology, 26. 57 - 68.<br />

236


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Konrad, H., Kirisits, T., Riegler et al., 2002. Genetic evidence for natural hybridization between the<br />

Dutch elm disease pathogens Ophiostoma novo-ulmi ssp. novo-ulmi and O. novo-ulmi ssp.<br />

americana. Plant Pathology, 51. 78 - 84.<br />

Krajňáková, J. and Longauer, R., 1996. Culture initiation, multiplication and identification <strong>of</strong> in vitro<br />

regenerants <strong>of</strong> resistant hybrid elms. Lesníctví - Forestry 42(6). 261 – 270.<br />

Pipe, N. D., Buck, K. W., Brasier, C. M., 1997. Comparison <strong>of</strong> the cerato-ulmin (cu) gene sequences<br />

<strong>of</strong> the Himalayan Dutch elm disease fungus Ophiostoma himal-ulmi with those <strong>of</strong> O. ulmi<br />

and Ophiostoma novo-ulmi suggests that the cu gene <strong>of</strong> O. novo-ulmi is unlikely to have<br />

been acquired recently from O. himal-ulmi. Mycol. Res., 101(4). 415 – 421.<br />

Polák, O., 1932. Ohrožení našich jilmů houbou Graphium ulmi (Referát o přednášce p. pr<strong>of</strong>. Ph. Dr.<br />

Jar. Pekla). Československý les, 12(11). 87 – 89.<br />

Scala, A., Tegli, S., Comparini, C. et al., 1994. Influence <strong>of</strong> fungal inoculum on cerato-ulmin<br />

production; purification <strong>of</strong> cerato-ulmin and detection in elm sucker cuttings. Petria, 4. 57<br />

– 67.<br />

Solla, A., Bohnens, J., Collin et al., 2005. Screening European Elms for Resistance to Ophiostoma<br />

novo-ulmi. Forest Science, 51(2). 134 – 141.<br />

237


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 238-240<br />

HAIL DAMAGE OF FOREST TREES IN WESTERN CANADA<br />

ABSTRACT<br />

Yasuyuki HIRATSUKA 1*<br />

1 Northern Forestry Centre, Canadian Forest Service<br />

Edmonton, Alberta CANADA<br />

* yahirats@nrcan.gc.ca<br />

Hail and hail storms damage agricultural and horticultural crops, as well as vehicles,<br />

houses and other buildings; these weather events also cause various degrees <strong>of</strong> damages to<br />

forest trees. Western Canada, with its diverse topography and varied weather systems, is<br />

known to have frequent hail storms. In the Canadian Rockies in the fall <strong>of</strong> 2008, a severe<br />

discoloration <strong>of</strong> foliage and many dead branches were observed over several hectares <strong>of</strong> a<br />

subalpine fir (Abies laciocarpa (Hook.)Nutt.) stand. After examining the symptoms and<br />

topographical features <strong>of</strong> the affected area, it was concluded that the damage was caused by<br />

a hail storm. Hail damage as well as other weather-related damage to forest trees, such as<br />

frost damage and drought damage, are <strong>of</strong>ten predisposing factors <strong>of</strong> branch and foliage<br />

symptoms associated with pathogenic fungi belonging to such genera as Nectria,<br />

Cytospora, and Sphaeropsis, and this weather-related damage should be considered when<br />

diagnosing forest tree health problems.<br />

Keywords: hail damage, western Canada, non-biotic tree diseases, predisposition<br />

1. INTRODUCTION<br />

Hail or hail storms usually occur at the front <strong>of</strong> storm systems from spring to<br />

fall but rarely during cold winter weather. Hail is formed when updrafts in<br />

thunderclouds carry rain drops upward into extremely cold areas <strong>of</strong> the cloud and<br />

form ice particles; updrafts carry the ice particles to the cold region <strong>of</strong> the storm<br />

clouds, where the size <strong>of</strong> the ice particles increases. When the ice particles become<br />

too heavy to be supported by the updraft, they fall to the ground at speeds <strong>of</strong> up to<br />

100 km/h. Hail storms can cause damage to crops, houses, and vehicles as well as<br />

injuries to people and animals. Hail damages to forest trees have <strong>of</strong>ten been<br />

reported (Benjamin, 1957; Grayburn, 1957; Hiratsuka and Zalasky, 1993; Laut and<br />

Elliott, 1966; Nelson, <strong>2009</strong>; Riley, 1953).<br />

Hail storms occur more frequently along mountain ranges because winds<br />

(moving in a horizontal direction) react to the change in topography, shifting<br />

upwards within thunderstorms and creating favorable conditions for the creation <strong>of</strong><br />

hail and hail storms. Hail storms are known to occur throughout many parts <strong>of</strong> the<br />

238


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

world mostly in temperate zones. Bangladesh has reported some <strong>of</strong> the largest<br />

hailstones ever measured and more hail-related human deaths than anywhere else<br />

in the world. In Canada, the foothills area <strong>of</strong> the Rocky Mountains, e<strong>special</strong>ly in<br />

the province <strong>of</strong> Alberta, is known to have more incidences <strong>of</strong> hail storms. Hail<br />

damage can be localized in small areas but areas as big as 10-km wide are also<br />

recorded.<br />

2. OBSERVATIONS<br />

In the fall <strong>of</strong> 2008, at Mount Assineboine Provincial Park, in British Columbia,<br />

Canada, I observed wide areas <strong>of</strong> natural forests <strong>of</strong> mature subalpine fir (Abies<br />

laciocarpa (Hook.)Nutt.) that were severely discoloured with many dead branches.<br />

No fungal or other biotic signs were present. After close examination <strong>of</strong> the<br />

damaged branches and topographical features <strong>of</strong> the area, I concluded that the<br />

symptoms were most likely caused by hail that occurred early in the summer, a few<br />

months before my visit. Symptoms included half- healed wounds on the upper side<br />

<strong>of</strong> the branches and many fallen branches on the ground. Needles positioned<br />

beyond the wounds on branches were <strong>of</strong>ten discoloured. Also, the damaged stands<br />

were facing a lake and the damage occurred mostly on the exposed side <strong>of</strong> the<br />

stand, indicating that the hail storm hit sideways from the direction <strong>of</strong> the lake.<br />

3. DISCUSSION AND CONCLUSION<br />

Hail damage as well as other harmful non-biotic conditions, such as severe<br />

drought and frost damage, will serve as predisposition factors to fungal and<br />

bacterial shoot and stem diseases. Trees in the affected areas need to be carefully<br />

examined and diagnosed. These climatic and physiological damages to trees<br />

initially do not have signs <strong>of</strong> fungi or bacteria, but they are <strong>of</strong>ten invaded by certain<br />

fungi or bacteria later, and they are <strong>of</strong>ten misdiagnosed as caused by secondary<br />

fungal invaders. Genera <strong>of</strong> fungi such as Cytospora, Nectria, and Sphaeropsis are<br />

known to be in this category <strong>of</strong> fungi on stems and branches <strong>of</strong> forest trees<br />

(Blodget et al., 1997; Zwolinski et al., 1995). In our recent publication <strong>of</strong> forest<br />

tree diseases <strong>of</strong> west-central Canada (Hiratsuka et al., 2004; Hiratsuka, 1987) we<br />

used expressions such as “Nectria and Cytospora associated with canker <strong>of</strong> broad<br />

leaf trees—” rather than “Nectria canker” or “Cyotspora canker”. So called<br />

“Screloderris canker” caused by Gremmeniella abiteina is also likely triggered by<br />

one or more non-biotic predisposition factor.<br />

It is important to consider non-biotic factors, including hail damage, as<br />

important predisposing conditions when we diagnose forest tree health problems,<br />

e<strong>special</strong>ly in cold climatic regions such as those in Canada.<br />

239


4. LITERATURE CITED<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Benjamin, D.M., 1957. Hail damage to five-year-old red pine in northern Wisconsin. For. Sci. 3:249-<br />

252.<br />

Blodget, J.T., Stanosz, G.R., Kruger, E.L.,1997. Sphaeropsis sapinea and water stress in a red pine<br />

plantation in central Wisconsin. Phytopathology 87:429-434.<br />

Grayburn, A.W., 1957. Hail damage to exotic forests in Canterbury. New Zealand J. For. 7: 50-57.<br />

Hiratsuka, Y., 1987. Forest tree diseases <strong>of</strong> the prairie provinces. Can. For. Serv., North. For. Cent.<br />

Inf. Rep. NOR-X-286. 142 p.<br />

Hiratsuka, Y., Langor, D.W., Crane, P.E., 2004. Field guide to forest insects and diseases <strong>of</strong> Canadian<br />

prairie provinces. Second Edition. Can. For. Serv., North. For. Cent. Spec. Publ. 297 p.<br />

Hiratsuka, Y, Zalasky, H, 1993. Frost and other climate-related damage <strong>of</strong> forest trees in the Prairie<br />

Provences. Can. For. Serv., North. For. Cent. Inf. Rep. NOR-X-331.<br />

Laut, J.G., Elliott, K.R., 1966. Extensive hail damage in northern Manitoba. For. Chron. 42:198.<br />

Nelson, S., <strong>2009</strong>. Some tree troubles not caused by insects or diseases. Sustainable horticultural<br />

information. http://gardenline.usask.ca/trees/some.html<br />

Riley, C.G., 1953. Hail damage in forest stands. For. Chron. 29:139-143.<br />

Zwolinski, J.B., Swart, W.J., Wingfield, M.J., 1995. Association <strong>of</strong> Sphaeropsis sapinea with insect<br />

infestation following hail damage <strong>of</strong> Pinus radiata. For. Ecol. Manag. 72:2<br />

240


Extended Abstracts<br />

241


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

242


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 243-244<br />

THREATENING TREE DISEASE IN EAST AFRICA<br />

Pia BARKLUND 1* , Jane NJUGUNA 1,2 , Abdella GURE 3 , Philip NYEKO 4 ,<br />

Katarina IHRMARK 1 and Jan STENLID 1<br />

1 Department <strong>of</strong> Forest Mycology and Pathology SLU, Box 7026,75007 Uppsala Sweden.<br />

2 Kenya Forest Research Institute, P.O. Box 20412 Nairobi City Square 00200, Kenya.<br />

3 Wondo Genet College <strong>of</strong> Forestry and Natural Resources, Hawassa University, P.O. Box 128<br />

Shashemane Ethiopia<br />

4 Department <strong>of</strong> Forest Biology and Ecosystems Management, Makerere University,P.O Box<br />

7062, Kampala, Uganda.<br />

ABSTRACT<br />

*pia.barklund@mykopat.slu.se<br />

Severe and extensive outbreaks <strong>of</strong> a dieback and canker disease have recently<br />

been observed on Grevillea robusta in Kenya, Uganda and to a lesser extent in<br />

Ethiopia. Grevillea is an excellent agr<strong>of</strong>orestry tree species grown intensively in<br />

east Africa to improve agricultural land use and rural livelihoods, and provide food<br />

security. Our recent studies on the disease indicate that 50-80% tree mortality<br />

occurs on severely infected farms. It is caused by Botryosphaeria spp., a fungal<br />

genus containing many species and more than one pathogenic species can occur in<br />

diseased trees. Samples were taken from Grevillea trees growing in different<br />

agroecological zones and from some other tree species with similar symptoms.<br />

Morphological and molecular methods were used to identify species and to study<br />

differences between populations in different agroecological zones as well as<br />

countries.<br />

The disease is more severe in dry areas than wet ones, emphasizing the need<br />

for proper species-site matching. Several other tree species, including indigenous<br />

and exotics, were found infected by Botryosphaeria in the region. E<strong>special</strong>ly<br />

alarming is the attack on different Eucalyptus species.<br />

Such disease outbreaks may be attributed to increased tree planting in<br />

agr<strong>of</strong>orestry and commercial tree plantations in the region. Increased acreage and<br />

number <strong>of</strong> trees/ha leads to an enlarged number <strong>of</strong> potential hosts, and a larger<br />

population size for pathogens to evolve genetically into more aggressive<br />

genotypes. Moreover, complex threats can arise when previously isolated fungal<br />

species brought together by human interference hybridize posing threats to tree<br />

hosts previously immune from their effects. Implications <strong>of</strong> the dieback and canker<br />

disease on the scaling up <strong>of</strong> agr<strong>of</strong>orestry technologies and commercial <strong>forestry</strong> in<br />

the region are discussed.<br />

Kewords: Grevillea robusta, Eucalyptus spp. Botryosphaeria, dieback and<br />

canker disease<br />

243


References:<br />

SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Toljander, Y. Toljander ,Y.K., Nyeko, P., Stenström, E., Ihrmark, K. and Barklund, P. 2007. First<br />

Report <strong>of</strong> Canker and Dieback Disease <strong>of</strong> Grevillea robusta in East Africa Caused by<br />

Botryosphaeria spp. Plant Disease 91(6):773 + suppl.<br />

Abdella, G., Slippers, B. Stenlid, J., 2005. Seed-borne Botryosphaeria spp. from native Prunus and<br />

Podocarpus trees in Ethiopia, with a description <strong>of</strong> the anamorph Diplodia rosulata sp.<br />

nov. Mycol. Res. 109 (9): 1005–1014.<br />

Gezahgne, A., Roux, J and Wingfield, M., 2003. Diseases <strong>of</strong> exotic Eucalyptus and pinus species in<br />

Ethiopian plantations. S. Afr. J. Sci. 99:29.<br />

244


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 245<br />

PREMATURE DEFOLIATION OF Cedrus libani IN SOUTH-WESTERN<br />

TURKEY<br />

Asko Lehtijärvi 1* and H. Tuğba Doğmuş-Lehtijärvi 1<br />

1 Süleyman Demirel University, Faculty <strong>of</strong> Forestry, 32260 Isparta, Turkey<br />

*asko@orman.<strong>sdu</strong>.edu.tr<br />

Cedrus libani A. Rich forests are presently found mainly in the Taurus<br />

Mountains <strong>of</strong> Turkey while only small populations <strong>of</strong> the once extensive and<br />

magnificent cedar forests remain in Lebanon and Syria. There are some 600 000 ha<br />

<strong>of</strong> C. libani forests in Turkey, most <strong>of</strong> which are mixed stands. Wood <strong>of</strong> C. libani<br />

is economically important as it is highly resistant to decay, durable and easy to<br />

process by hand tools and machines. In addition to economical value C. libani<br />

forests are significant from historical, cultural and aesthetic point <strong>of</strong> view, and<br />

therefore actions for sustainable usage and restoring degraded stands have been<br />

initiated.<br />

In some C. libani stands in the lakes district <strong>of</strong> Turkey browning <strong>of</strong> needles<br />

occurred in spring, e<strong>special</strong>ly in the lower part <strong>of</strong> the canopy. The disease was<br />

observed both on saplings growing as understory in mixed forest as well as on<br />

approximately 10–m–tall trees in an even–aged stand. A close investigation<br />

revealed ascomata on 1–year–old needles still attached to the twigs. Often only the<br />

ascomata-bearing part <strong>of</strong> the needle was brown while other parts had remained<br />

green. In the lower part <strong>of</strong> the canopy <strong>of</strong> a sapling major part <strong>of</strong> the needles could<br />

be diseased. On a single branch showing signs <strong>of</strong> serious premature defoliation<br />

proportion <strong>of</strong> needles bearing ascomata was 49.3 % <strong>of</strong> the total dry weight <strong>of</strong> the<br />

needles, while the corresponding value for more or less healthy needles was 37.4%.<br />

Chlorotic or damaged needles without ascomata made up 13.2% <strong>of</strong> the total needle<br />

dry weight.<br />

The morphological characteristics <strong>of</strong> the ascomata were very similar to those <strong>of</strong><br />

Ploioderma cedri S. Singh, S.N. Khan & B. Misra occurring on C. deodara in<br />

India, but asci and ascospores were somewhat larger. The frequent fruiting on dead<br />

parts <strong>of</strong> an otherwise green needle indicates that the fungus is the causal agent <strong>of</strong><br />

the disease.<br />

245


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 246-251<br />

CONTRIBUTIONS TO THE PHYLOGENY OF EUROPEAN<br />

Porodaedalea SPECIES (BASIDIOMYCETES, HYMENOCHAETALES)<br />

Michal TOMŠOVSKÝ 1 , Libor JANKOVSKÝ 1*<br />

1 Faculty <strong>of</strong> Forestry and Wood Technology, Mendel University <strong>of</strong> Agriculture and Forestry in<br />

Brno, Zemědělská 3, CZ – 613 00 Brno, Czech Republic; e-mails: tomsovsk@mendelu.cz,<br />

ABSTRACT<br />

*jankov@mendelu.cz.<br />

The genus Porodaedalea is a taxonomically difficult complex <strong>of</strong> morphologically<br />

similar species causing white pocket rot <strong>of</strong> living conifers. The evolutionary relationships<br />

<strong>of</strong> European species were examined using sequences <strong>of</strong> the internal transcribed spacer<br />

(ITS) region <strong>of</strong> the nuclear ribosomal DNA and <strong>of</strong> translation elongation factor 1 alpha<br />

(tefa). Our results confirm the occurrence <strong>of</strong> Porodaedalea chrysoloma, P. pini and P.<br />

laricis in Europe. P. laricis is newly reported in Fennoscandia on Picea and in the Central<br />

European mountains (Alps, High Tatras, and Bohemian Forest) on Larix and Pinus spp.<br />

These specimens had been previously identified as Porodaedalea chrysoloma or Phellinus<br />

vorax (an invalidly described species). Although frequently confused, P. chrysoloma and P.<br />

laricis can be distinguished on the basis <strong>of</strong> pore morphology. We also report our finding <strong>of</strong><br />

P. pini on Larix. In general, the tefa sequences are more variable than the ITS sequences<br />

and reveal the remarkable affinity <strong>of</strong> some Scandinavian and Central European specimens<br />

to those from Central Asia.<br />

1. INTRODUCTION<br />

The genus Porodaedalea includes parasites on conifers, causing white pocket<br />

rot. The genus belongs to one <strong>of</strong> the most taxonomically difficult groups <strong>of</strong><br />

hymenochaetoid pore fungi. The so-called Phellinus pini group was raised to the<br />

generic level by Fiasson and Niemelä (1984). The basidiocarps are perennial,<br />

effused-reflexed to pileate, solitary to imbricate, and corky to woody hardness. The<br />

colour is rust brown to dark grey on the upper surface, while the poroid surface is<br />

ochre brown or rust brown to umbre brown, and more or less shining. The pores<br />

are circular to angular, tending to split and becoming irregular to daedaleoid and<br />

labyrinthine. Setae are commonly present in the hymenium. In some areas the<br />

species are reported to be economically important pathogens <strong>of</strong> conifers<br />

(Lannenpaa et al., 2008; Černý, 1989; Lehtijärvi et al., 2007). The genus has<br />

previously been treated as part <strong>of</strong> a broadly conceived genus Phellinus s.l.<br />

(Ryvarden and Gilbertson, 1994), but molecular studies have revealed the<br />

heterogeneity <strong>of</strong> that genus (Wagner and Fischer, 2002). Therefore smaller, more<br />

homogeneous genera are currently accepted.<br />

246


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

The genus Porodaedalea is comprised <strong>of</strong> a rather small number <strong>of</strong> species.<br />

Nevertheless, the substrate specificity and exact distribution <strong>of</strong> each species are<br />

poorly known due to the lack <strong>of</strong> microscopic characters suitable for exact<br />

identification at the species level (Fischer, 1996). In Europe, two species are<br />

traditionally recognized, Porodaedalea pini (Brot.) Murrill and Porodaedalea<br />

chrysoloma (Fr.) Fiasson & Niemelä, with the former species restricted to Pinus<br />

and the latter species restricted to Picea as specific hosts. A third species,<br />

Porodaedalea laricis (Jacz. ex Pilát) Niemelä (Niemelä et al., 2005), is distributed<br />

on Larix from the European part <strong>of</strong> Russia to Siberia and the Russian Far East and<br />

is probably present in China (Dai, 1999). Fischer (2000) discovered this species on<br />

a non-indigenous Larix trees in Southern Finland and proposed for it the new name<br />

Porodaedalea niemelaei M. Fischer. This name has been synonymised with P.<br />

laricis by Niemelä et al. (2005).<br />

Indigenous stands <strong>of</strong> Larix decidua, Pinus cembra and Pinus mugo in the<br />

mountains <strong>of</strong> Central Europe (Alps, High Tatras) are inhabited by species differing<br />

from P. chrysoloma and P. pini (Černý, 1985; Fischer, 2000). These populations<br />

may belong to P. laricis or an undescribed species. The name Phellinus vorax has<br />

been applied to specimens from this region. However, the combination Phellinus<br />

vorax is based on incorrectly published basionym Daedalea vorax Harkness.<br />

Therefore, the name is unavailable, although it has been commonly used<br />

(Breitenbach and Kränzlin, 1986). The fungus named Daedalea vorax, a Western<br />

American species growing on Pseudotsuga menziesii, was later correctly described<br />

as Phellinus gilbertsonii M.J. Larsen (Larsen, 2000).<br />

Molecular taxonomy methods are frequently used as tools for the identification<br />

<strong>of</strong> fungal taxa. Such methods as RFLP (Fischer, 1996) and sequencing <strong>of</strong> various<br />

regions <strong>of</strong> nuclear ribosomal DNA (Fischer, 2000; Wagner and Fischer, 2002) have<br />

been used to reveal the evolutionary relationships <strong>of</strong> the species. The sequence <strong>of</strong><br />

the ITS region <strong>of</strong> the ribosomal DNA also completes the neotypification <strong>of</strong><br />

Phellinus chrysoloma (Larsen and Stenlid, 1999). The aim <strong>of</strong> the present study is to<br />

elucidate the identification <strong>of</strong> Porodaedalea specimens occurring in various parts<br />

<strong>of</strong> Europe, using sequences <strong>of</strong> the ITS region <strong>of</strong> the nuclear ribosomal DNA (ITS)<br />

and <strong>of</strong> translation elongation factor 1 alpha (tefa). The study is focused mainly on<br />

specimens that have been previously identified as Phellinus vorax.<br />

2. MATERIALS AND METHODS<br />

In total, 30 specimens <strong>of</strong> Porodaedalea were included in the study. Either<br />

fungal cultures or herbarium specimens were used for DNA analyses.<br />

DNA was isolated from dried fungal material or from fresh cultures that were<br />

grown on Petri dishes with MEA medium (3% Malt extract, 0.5% peptone, 1.5%<br />

agar; Himedia, Mumbai, India) using the PowerSoil DNA Isolation Kit (Mo-Bio,<br />

Carlsbad, USA). PCR reactions were set up according to standard protocols,<br />

247


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

supplemented with 5% (v/v) bovine serum albumin (BSA) as a PCR enhancer.<br />

DNA fragments encompassing the ITS and tefa DNA regions were amplified using<br />

the primer combinations ITS1/ITS4 and EF595F/EF1160R, respectively (White et<br />

al., 1990; Kauserud and Schumacher, 2001). The DNA was PCR-amplified as in<br />

previous studies (Tomšovský et al., 2006) using a Mastercycler® ep thermocycler<br />

(Eppendorf, Hamburg, Germany). In cases when amplification <strong>of</strong> the ITS and tefa<br />

regions was difficult, the primer pairs ITS1F/ITS4B and EF-526F/EF-1567R,<br />

respectively, were used in nested PCR (for primer sequences, see Gardes and<br />

Bruns, 1993; O’Donnell et al., 1998).<br />

The amplified fragments were sequenced by the DNA Sequencing Service <strong>of</strong><br />

Macrogen Inc. (Seoul, Korea). We added an ITS sequence from the Porodaedalea<br />

chrysoloma neotype (Genbank acc. no. AF123440; Larsen and Stenlid, 1999) to the<br />

data set. ITS and tefa sequences <strong>of</strong> Onnia leporina were chosen as outgroups based<br />

on the results <strong>of</strong> Wagner and Fischer 2002.<br />

Sequences <strong>of</strong> each individual marker were aligned using the Clustal W<br />

algorithm in BioEdit and adjusted manually. To determine whether the datasets<br />

from different genetic markers were in significant conflict, partition homogeneity<br />

tests were performed between the markers in all possible pair-wise combinations.<br />

The tests were done in PAUP 4.0b10 using 100 replicates and the heuristic general<br />

search option. The null hypothesis <strong>of</strong> congruence was rejected if p < 0.01<br />

Phylogenies were generated in MrBayes version 3.1.2.The best-fit model and<br />

parameters given by MrModeltest were used in the analyses. Markov chains were<br />

initiated from a random tree and were run for 2,000.000 generations; the samples<br />

were taken every 100th generation. Posterior probabilities (PP) were used as<br />

Bayesian branch supports on the consensus trees. In addition, bootstrap branch<br />

support values (BP) were estimated in PAUP 4.0b10 under the maximum<br />

parsimony criterion using 1000 replicate datasets with random sequence addition<br />

during each heuristic search.<br />

3. RESULTS<br />

Partition homogeneity tests showed significant conflict between the two genetic<br />

markers used (p ≤ 0.01). This did not allow us to perform a combined analysis <strong>of</strong><br />

the ITS and tefa sequence data. The results <strong>of</strong> the phylogenetic analyses <strong>of</strong> the two<br />

datasets are ambiguous. The ITS phylogram shows three main groups, while the<br />

tefa phylogram shows four. In ITS data set, the most basal lineage within the<br />

ingroup is composed <strong>of</strong> P. chrysoloma specimens from the Czech Republic,<br />

Estonia, Romania, and Southern Sweden, including the P. chrysoloma neotype.<br />

This P. chrysoloma group is also consistent in the tefa phylogram.<br />

The second group includes P. pini specimens from the Czech Republic, Estonia,<br />

Croatia, Lithuania, and Sweden, including a specimen growing on Larix. The<br />

position <strong>of</strong> this group is variable; it forms a well supported terminal clade in the<br />

ITS phylogram, but it is placed in the centre <strong>of</strong> the tefa phylogram.<br />

248


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

In the ITS phylogram, the third clade is composed <strong>of</strong> P. laricis specimens from<br />

European part <strong>of</strong> Russia, Kazakhstan and the Russian Far East, Fennoscandia, and<br />

the Central European mountains (Alps, High Tatras, Bohemian Forest). However,<br />

this group is not consistent in both analyses. In the tefa phylogram, five sequences<br />

from the Czech Republic, Norway, Sweden, and Kazakhstan are excluded from this<br />

group due to the presence <strong>of</strong> unique nucleotide substitutions at positions 166 and<br />

408 in the alignment. Nevertheless, this distant clade is poorly supported.<br />

4. DISCUSSION<br />

The topologies <strong>of</strong> the two gene regions (ITS, tefa) delimiting Porodaedalea<br />

species are inconcordant, so they do not follow phylogenetic species recognition<br />

(according to Taylor et al., 2000). Nevertheless, the phylogram topologies <strong>of</strong> these<br />

two regions have been found to be incongruent in other studies (Kauserud et al.,<br />

2007; Ota and Hattori, 2008). The most surprising result in our study is the division<br />

<strong>of</strong> the homogeneous P. laricis ITS group into two groups in the tefa dataset.<br />

However, the PP and BP values <strong>of</strong> the P. laricis B group in the tefa phylogram are<br />

low. Therefore, we suggest that the name P. laricis is adopted for all specimens<br />

included in the P. laricis ITS group that were originally identified as P.<br />

chrysoloma. Almost all sequenced specimens from Fennoscandia that were<br />

identified as Phellinus chrysoloma are unrelated to the neotype <strong>of</strong> that species and<br />

most likely belong to P. laricis instead. Our results resemble those <strong>of</strong> Černý (1985,<br />

1989), who placed these Fennoscandian specimens growing on Picea in Phellinus<br />

vorax, an incorrectly published name synonymous with Phellinus laricis (Černý,<br />

1985). In any case, the distribution <strong>of</strong> genuine P. chrysoloma in North<br />

Fennoscandia is questionable. The species occurs without doubt in southern<br />

Sweden and Finland (Larsen and Stenlid 1999; Fischer 2000), but the northern<br />

limit <strong>of</strong> its distribution is unclear.<br />

The occurrence and host affinity <strong>of</strong> Porodaedalea is worth a detailed<br />

discussion. Niemelä et al. (2005) assumed strict host specificity <strong>of</strong> Porodaedalea<br />

species, and therefore they set the westernmost limits <strong>of</strong> P. laricis in accordance<br />

with the natural occurrence <strong>of</strong> Larix sibirica in Russia. Although Fischer (2000)<br />

examined a specimen growing on artificially planted Larix from Finland, it was<br />

reported to be an introduced fungus due to the non-indigenous state <strong>of</strong> its host.<br />

According to our results, P. laricis is widely distributed in Fennoscandia, using<br />

spruce as its host.<br />

In Central Europe, the elevation (≥ 1400 m in the High Tatras in Slovakia) and<br />

native distribution <strong>of</strong> the hosts (Pinus and Larix) are reported to be crucial for the<br />

occurrence <strong>of</strong> Phellinus vorax (= P. laricis) (Černý, 1989). Our observations<br />

support this belief; P. laricis growing on Pinus mugo in the Bohemian Forest (=<br />

Šumava Mts., Czech Republic) occurs under different climatic conditions than P.<br />

pini inhabiting Pinus sylvestris in the adjacent area. While the only collection <strong>of</strong> P.<br />

laricis there was recorded at ca. 1300 m, the highest elevation recorded for P. pini<br />

249


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

is 825 m, and the centre <strong>of</strong> distribution <strong>of</strong> P. pini is around 700 m or below<br />

(Tomšovský, 2002). Kotlaba (1984) confirms the distribution <strong>of</strong> P. pini between<br />

155-780 m in elevation in the former Czechoslovakia. Jahn (1963) points out the<br />

occurrence <strong>of</strong> P. chrysoloma (under the name <strong>of</strong> Phellinus pini var. abietis) not<br />

only on Picea, but also on Abies and Larix. These host species are also cited by<br />

more recent publications (Kotlaba, 1984; Černý, 1989; Ryvarden and Gilbertson,<br />

1994), but we did not obtain a specimen from Abies. Our results confirm that fungi<br />

on Larix may belong to either P. laricis or P. pini. The native status <strong>of</strong> the host<br />

trees, correlated with elevation, is crucial for occurrence <strong>of</strong> the respective species.<br />

Collections <strong>of</strong> P. pini from Larix have been mentioned previously by Černý<br />

(1985, 1989) and Kotlaba (1984), while Ryvarden and Gilbertson (1994) reported<br />

Pinus as the only host genus <strong>of</strong> this fungus. Therefore, the host spectrum <strong>of</strong> P. pini<br />

might extend to non-indigenous conifers.<br />

The complex phylogeographical structure <strong>of</strong> Euroasian Porodaedalea deserves<br />

to be studied in detail. Questions about the histories <strong>of</strong> local populations are posed<br />

by the occurrence <strong>of</strong> a common nucleotide substitution among geographically<br />

distant specimens from Scandinavia, the Czech Republic, and Kazakhstan (i.e., the<br />

"P. laricis B" group). Historical changes in host distributions probably led to geneflow<br />

and introgression <strong>of</strong> new genotypes that may have survived in local areas.<br />

5. ACKNOWLEDGEMENTS<br />

This work was supported by the Czech Science Foundation, project no.<br />

521/07/J039.<br />

6. REFERENCES<br />

Breitenbach, J., Kranzlin, F., 1986. Fungi <strong>of</strong> Switzerland, vol. 2, Heterobasidiomycetes,<br />

Aphyllophorales, Gasteromycetes. Mykologia, Lucerne<br />

Černý, A., 1985. Taxonomic study in the Phellinus pini group. Česká mykologie 39: 71-84<br />

Černý, A., 1989. Parazitické dřevokazné houby [Parasitic wood-decaying fungi]. Státní zemědělské<br />

nakladatelství, Praha<br />

Dai, Y.C., 1999. Phellinus sensu lato (Aphyllophorales, Hymenochaetaceae) in East Asia. Acta<br />

Botanica Fennica 166: 1-115<br />

Fischer, M., 1996. Molecular and microscopical studies in the Phellinus pini group. Mycologia 88:<br />

230-238<br />

Fischer, M., 2000. Porodaedalea (Phellinus pini group, Basidiomycetes) in Europe: a new species on<br />

Larix sibirica, P. niemelaei. Karstenia 40: 43-48<br />

Gardes, M., Bruns, T.D., 1993. ITS primers with enhanced specificity for basidiomycetes-application<br />

to the identification <strong>of</strong> mycorhizae and rusts. Molecular Ecology 2: 113-118<br />

Jahn, H., 1963. Mitteleuropäische Porlingen (Polyporaceae s. lato) und ihr Vorkommen in Westfalen.<br />

Westfälische Pilzbriefe 4: 91-112<br />

250


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Kauserud H, Schumacher T (2001) Outcrossing or inbreeding: DNA Markers provide evidence for<br />

type <strong>of</strong> reproductive mode in Phellinus nigrolimitatus (Basidiomycota). Mycol. Res. 105:<br />

676-683<br />

Kauserud, H., H<strong>of</strong>ton, T.H., Saetre, G.P., 2007. Pronounced ecological separation between two<br />

closely related lineages <strong>of</strong> the polyporous fungus Gloeoporus taxicola. Mycological<br />

Research 111: 778-786.<br />

Kotlaba, F., 1984. Zeměpisné rozšíření a ekologie chorošů (Polyporales s.l.) v Československu<br />

[Geografical distribution and ecology <strong>of</strong> polypores (Polyporales s.l.) in Czechoslovakia].<br />

Academia, Praha.<br />

Larsen, M.J., 2000. Phellinus gilbertsonii sp. nov. from western North America causing heart-rot <strong>of</strong><br />

coastal Douglas-fir. Folia Cryptog Estonica 37: 51-54<br />

Larsen, M.J., Stenlid, J., 1999. Neotypification <strong>of</strong> Phellinus chrysoloma. Folia Cryptog Estonica 34:<br />

9-13<br />

Lannenpaa, A., Aakala, T., Kauhanen, H., Kuuluvainen, T., 2008. Tree mortality agents in pristine<br />

Norway spruce forests in northern Fennoscandia. Silva Fennica 42: 151-163<br />

Lehtijärvi, A., Lehtijärvi, H.T.D., 2007. Occurrence <strong>of</strong> Porodaedalea pini (Brot.: Fr.) Murr. in pine<br />

forests <strong>of</strong> the lake district in south-western Turkey. Phytopathologia Mediterranea 46:<br />

316-319<br />

Niemelä, T., Kinnunen, J., Larsson, K.H., Schigel, D.S., Larsson, E., 2005. Genus revisions and new<br />

combination <strong>of</strong> some North European polypores. Karstenia 45: 75-80<br />

O'Donnell, K., Kistler, H.C., Cigelnik, E., Ploetz, R.C., 1998. Multiple evolutionary origins <strong>of</strong> the<br />

fungus causing Panama disease <strong>of</strong> banana: Concordant evidence from nuclear and<br />

mitochondrial gene genealogies. Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong><br />

USA 95: 2044–2049<br />

Ota, Y., Hattori, T., 2008. Relationships among three Japanese Laetiporus taxa based on phylogenetic<br />

analysis and incompatibility tests. Mycoscience 49: 168-177.<br />

Ryvarden, L., Gilbertson, R.L., 1994. European polypores, vol.2. Synopsis Fungorum<br />

6.Fungiflora, Oslo.<br />

Taylor, J.W., Jacobson, D.J., Kroken, S., Kasuga, T., Geiser, D.M., Hibbett, D.S., Fisher, M.C., 2000.<br />

Phylogenetic species recognition and species concepts in fungi. Fungal Genetics and<br />

Biology 31: 21-32.<br />

Tomšovský, M., 2006. Molecular phylogeny <strong>of</strong> European Trametes (Basidiomycetes, Polyporales)<br />

species based on LSU and ITS (nrDNA) sequences. Nova Hedwigia 82: 269-280.<br />

Wagner, T., Fischer, M., 2002. Proceedings towards a natural classification <strong>of</strong> the worldwide taxa<br />

Phellinus s.l. and Inonotus s.l., and phylogenetic relationships <strong>of</strong> allied genera. Mycologia<br />

94: 998-1016<br />

White, T.J., Bruns, T.D., Lee, S.B., Taylor, J.W., 1990. Amplification and direct sequencing <strong>of</strong> fungal<br />

ribosomal RNA genes for phylogenetics. Academic Press, New York<br />

251


Abstracts<br />

253


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

254


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 255<br />

EXAMINING THE GEOGRAPHIC DISTRIBUTION OF Diplodia pinea<br />

AND D. scrobiculata:<br />

A CASE STUDY FROM MINNESOTA, USA<br />

J. S. ALBERS 1 , Denise R. SMITH 2 , Glen R. STANOSZ 2*<br />

1 Division <strong>of</strong> Forestry, Minnesota Department <strong>of</strong> Natural Resources, Grand Rapids, MN, 55744, USA<br />

2 Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Wisconsin-Madison, WI, 53706, USA<br />

* grs@plantpath.wisc.edu<br />

Although the shoot blight and canker pathogen Diplodia pinea is more<br />

commonly reported and is distributed in native and exotic pine stands in much <strong>of</strong><br />

the world, a second very similar and closely related fungus, D. scrobiculata, has<br />

been detected in the USA, Mexico, France, Israel, Italy, and Spain, and is likely<br />

present in other countries. Both <strong>of</strong> these fungi are associated with red pine (Pinus<br />

resinosa) and jack pine (P. banksiana) in the northcentral and northeastern USA.<br />

Their distribution in Minnesota was studied by examination <strong>of</strong> seed cones (on<br />

which these fungi sporulate). 100 cones collected from the forest floor <strong>of</strong> each <strong>of</strong><br />

109 red pine stands and 28 jack pine stands were visually examined for Diplodia<br />

pycnidia and conidia. At least one <strong>of</strong> these fungi was detected from 106 <strong>of</strong> 109 red<br />

pine stands and from all jack pine stands. Mean frequencies <strong>of</strong> positive red and<br />

jack pine cones, respectively, were 27% (range 0-84%) and 12% (range 2-41%).<br />

PCR assays confirmed pathogen identity for subsets <strong>of</strong> cones. D. pinea was<br />

detected from cones collected at 102 <strong>of</strong> 109 red pine stands (69% <strong>of</strong> red pine cones<br />

tested), and 18 <strong>of</strong> 28 jack pine stands (18% <strong>of</strong> jack pine cones tested). In contrast,<br />

D. scrobiculata was detected from cones collected at only 26 <strong>of</strong> 109 red pine<br />

stands (7% <strong>of</strong> red pine cones tested), but 26 <strong>of</strong> 28 jack pine stands (79% <strong>of</strong> jack<br />

pine cones tested). These fungi sometimes co-occurred in stands <strong>of</strong> either host, and<br />

occasionally both were detected from individual cones <strong>of</strong> either host. Although<br />

differences between D. pinea and D. scrobiculata in host association, presence at a<br />

given location, and frequency <strong>of</strong> occurrence at a given location were apparent, each<br />

was found across the entire area surveyed.<br />

255


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 256<br />

FOREST INVASIVE ALIEN FUNGAL SPECIES PRESENT IN LIVE<br />

PLANT MATERIAL<br />

Jean A. BERUBE 1*<br />

1 Canadian Forest Service, Natural Resources Canada, Ste-Foy, Qc, Canada G1V 4C7.<br />

*jberube@cfl.<strong>forestry</strong>.ca<br />

An early warning system based on a random sampling <strong>of</strong> asymptomatic live<br />

plant material arriving in Canada is used to detect alien fungal pests. Forty-six<br />

sample lots collected by Canadian Food Inspection Agency (CFIA) inspectors from<br />

the province <strong>of</strong> Quebec were analyzed by cloning the fungal ribosomal ITS present<br />

in the plant tissues. We obtained 101 fungal species associated with 36 different<br />

host plants from the USA, France, the Netherlands and Thailand. Six fungal<br />

species found in this study could have a low to moderate potential impact and 11<br />

could have a low potential impact for Canadian forests. Another 14 species could<br />

not be assessed given the limited scientific information available. In all cases, the<br />

potential impact evaluations <strong>of</strong> these 31 species originate from the fact that these<br />

species are new to science and/or belong to genera and families where pathogenic<br />

species are common. The alien fungal introductions with a potential to affect<br />

Canadian forests were found at a significant frequency (12.4%) and were present in<br />

every sample lot sent by CFIA. The 70 other species found in this study were nonpathogenic<br />

fungi; weak to moderately virulent, common and cosmopolitan species;<br />

or virulent species found on tropical hosts only.<br />

256


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 257<br />

NEW ADVANCES IN THE STUDY OF THE TAXONOMY OF THE<br />

EUROPEAN RACE OF GREMMENIELLA ABIETINA<br />

Leticia BOTELLA 1 , Julio Javier DIEZ 1 , Jarkko HANTULA 2*<br />

1 Laboratorio de Entomología y Patología Forestal, Departamento de Producción Vegetal<br />

Recursos Forestales, E.T.S.I.I.A.A., Universidad de Valladolid, “Campus La Yutera”, Avda. Madrid,<br />

44, 34004, Palencia, Spain.<br />

2 Finnish Forest Research Institute (METLA), Unit <strong>of</strong> Vantaa. Joikiniemenkuja, 1, PL 18, FI-<br />

0130. Finland.<br />

* jarkko.hantula@metla.fi<br />

In order to investigate the taxonomy <strong>of</strong> the European race <strong>of</strong> Gremmeniella<br />

abietina var. abietina ten Spanish isolates were randomly chosen among 91 to be<br />

analyzed genetically and compared with 7 Swiss isolates biotype Alpine, 10<br />

Finnish isolates biotype A and 10 Finnish isolates biotype B randomly chosen in<br />

the same way. RAMS markers CCA, CGA and GAAA1000 supported by sequence<br />

analysis <strong>of</strong> the locus GAAA1000 were used in this report to study the genetic<br />

variability <strong>of</strong> Spanish and Swiss isolates in respect <strong>of</strong> the rest <strong>of</strong> European biotypes<br />

and, thus, to establish what biotype they belong to. Philogenic relationships among<br />

A, B, Alpine and Spanish sequences analyzed based on the neighbour-joining<br />

method showed that Alpine type, that is, Swiss sequences, is more closely related<br />

to B type, and Spanish isolates appear clearly separated from the rest <strong>of</strong> the<br />

biotypes.<br />

Keywords: Brunchorstia pinea – European race – Pinus halepensis – RAMS<br />

– taxonomy.<br />

257


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 258<br />

STUDIES ON THE SIGNIFICANCE, CAUSAL AGENTS AND<br />

CONTROL METHODS OF DAMPING- OFF DISEASE IN FOREST<br />

NURSERIES OF AEGEAN AND LAKES DISTRICT<br />

H. Tuğba DOĞMUŞ LEHTİJÄRVİ 1* and Gülay TURHAN 2<br />

1 Suleyman Demirel University, Faculty <strong>of</strong> Forestry, 32260 Isparta, Turkey<br />

2 Aegean University, Faculty <strong>of</strong> Plant Protection, 35100, Bornova- İzmir, Turkey<br />

*tugba@orman.<strong>sdu</strong>.edu.tr<br />

The survey <strong>of</strong> containerized and bare rooted seedlings <strong>of</strong> Pinus brutia (Ten.),<br />

Pinus nigra subsp. pallasiana (Lamb.)Holmboe, Cedrus libani (A. Rich), Pinus<br />

pinea (L.) and Ailanthus glandulosa (Desf.) showed that damping- <strong>of</strong>f intensity<br />

was higher in the nurseries <strong>of</strong> Lakes District than that <strong>of</strong> Aegean District. The most<br />

prevalent fungi were found Fusarium spp., Rhizoctonia solani (Kühn.), Pythium<br />

spp., Alternaria spp. and Macrophomina phaseolina with the rates <strong>of</strong> 53, 19, 10, 10<br />

and 6 %, respectively.<br />

Five hundred fifty fungal isolates were collected from diseased seedlings and<br />

108 <strong>of</strong> them were tested for their pathogenicity. Total 283 fungi and 200 bacteria<br />

were isolated from natural forest soil and among them, Gliocladium virens,<br />

Trichoderma koningii, Penicillium ademetzi, Myrothecium verrucaria,<br />

Paecilomyces lilacinus, 2 Streptomyces and 3 unidentified bacteria were found to<br />

have strong antibiotic activity against the selected pathogenic isolates. Five<br />

fungicides, propomocarp, hymexazole, thiram, PCNB and propineb, were<br />

evaluated under in vitro against the antagonists. Antagonistic isolates showed a<br />

very high degree <strong>of</strong> sensitivity to the fungicides with the exception <strong>of</strong><br />

promomocarp and hymexazole.<br />

In vivo studies were carried out as biological, chemical and integrated control <strong>of</strong><br />

5 damping- <strong>of</strong>f pathogens on P. brutia and P. nigra. After soil infestation with<br />

pathogens, the fungicides were applied by soil drenching and the antagonist<br />

application was achived by seed coating. Propomocarp and hymexazole were<br />

found more effective than the other fungicides and gave the best results when they<br />

were used with the antagonists in combination.<br />

Key words: Pythium spp. , Macrophomina phaseolina, Rhizoctonia solani,<br />

Alternaria alternata, Fusarium spp., Chemical control, biological control.<br />

258


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 259<br />

A FOLIAR DISEASE OF Celtis glabrata IN THE LAKES REGION<br />

Gürsel KARACA 2* , Tuğba DOĞMUŞ-LEHTİJÄRVI 1 , Hüseyin FAKIR 1<br />

1 University <strong>of</strong> Süleyman Demirel, Faculty <strong>of</strong> Forestry, Isparta<br />

2 University <strong>of</strong> Süleyman Demirel Faculty <strong>of</strong> Agriculture, Isparta<br />

*gkaraca@ziraat.<strong>sdu</strong>.edu.tr<br />

Celtis species are widely distributed in Turkey and has 4 species. One <strong>of</strong> them is<br />

C. glabrata (Steven ex Planchon) and is found in the Lakes region. A fungal<br />

disease was commonly observed on Celtis leaves with irregular, large, black spots<br />

on both sides and velvety apperance on the lower side. Disease incidence was so<br />

high that all the leaves <strong>of</strong> the trees showed symptoms with varying degrees. As a<br />

result <strong>of</strong> the microscopic examination <strong>of</strong> the lower leaf surfaces, the fungus was<br />

identified as Sirosporium celtidis (Biv., Bern. Ex Sprengel) M. B. Ellis. Since C.<br />

glabrata is not grown in forest nurseries in our country, the fungus is not<br />

economically important. However, it can cause defoliation and poor growth <strong>of</strong><br />

trees in natural ecosystems.<br />

259


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 260<br />

DETERMINATION OF MACROMYCETES IN THE REGION OF<br />

KOCAELI<br />

Ayhan KARAKAYA¹*<br />

1 Poplar and Fast Growing Forest Trees Research Institute<br />

P.O. Box 93 41001 Kocaeli/Turkey<br />

* ayhankarakaya68@hotmail.com<br />

This study entitled “Determination <strong>of</strong> Macromycetes in the Region <strong>of</strong> Kocaeli”<br />

was carried out to identify macr<strong>of</strong>ungi species growing in the province <strong>of</strong> Kocaeli.<br />

The aim <strong>of</strong> these types <strong>of</strong> studies can be summarized as to identify species, to<br />

determine its rank in the systematic and to delineate the species’in distribution<br />

area.<br />

In the first step, the literature on the macr<strong>of</strong>ungi was studied, and the relevant<br />

information was collected. The samples were collected following determining the<br />

field sites . All <strong>of</strong> the relevant information on the samples collected was recorded.<br />

Each sample was later carefully excavated, bagged separately and transferred to the<br />

lab.<br />

The internal morphological characteristics <strong>of</strong> the reproductive organs including<br />

the cap and lamella <strong>of</strong> the samples were determined. The cap characteristics<br />

including shape, dimension, color, color change, centrality, and lamella or pore<br />

structure, and the odor and degree <strong>of</strong> density <strong>of</strong> the succulent part were determined.<br />

Various features <strong>of</strong> the stem including length, color, color change, and the<br />

existence <strong>of</strong> stem as well as the ring and residuals, if existed, were also noted.<br />

Moreover, in order to determine spore print, one reproductive organ was<br />

sampled for each sample brought to the lab. The porous part <strong>of</strong> the cap was placed<br />

on a white sheet and protected from air currents that would sweep the fallen<br />

pollens. At least 12 hours later, spore print’s color was captured on the sheet.<br />

The macr<strong>of</strong>ungi was identified after the information collected from the field<br />

sites and the lab study had been compared with the literature. At the end <strong>of</strong> the<br />

study, 89 macr<strong>of</strong>ungi species, 79 <strong>of</strong> which belonged to the Basidiomycetes class<br />

and 10 <strong>of</strong> which were <strong>of</strong> the Ascomycetes class were indentified in the province <strong>of</strong><br />

Kocaeli.<br />

Keywords: Macr<strong>of</strong>ungi, Kocaeli, forest, systematic, phytopatology.<br />

260


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 261<br />

ARE SUBPOPULATIONS OF Heterobasidion parviporum<br />

DIFFERENTIATED BY LOCAL CLIMATE?<br />

Michael M. MÜLLER 1* , Nicola LA PORTA 2 , Jaana EKOJÄRVI 1 , Jarkko<br />

HANTULA 1 , Kari KORHONEN 1<br />

1 Finnish Forest Research Institute, Box 18, 01301 Vantaa, Finland<br />

2 IASMA Research and Innovation Centre, Fondazione Edmund Mach, Environment and<br />

Natural Resources Area, Via E. Mach 1, S. Michele all'Adige 38010, Italy<br />

* Michael.mueller@metla.fi<br />

We determined the decomposition rate (DR) <strong>of</strong> spruce wood by representatives<br />

<strong>of</strong> different subpopulations <strong>of</strong> Heterobasidion parviporum at various temperatures.<br />

Sixty three H. parviporum isolates originating from geographically distant and<br />

climatically varying environments (Finland, Denmark, Italy and Central Siberia)<br />

were cultivated at eight temperatures between 6 o C and 33 o C on Norway spruce saw<br />

dust as the only substrate. Decomposition activity was determined as the<br />

production <strong>of</strong> carbon dioxide. The optimal temperature for decomposition varied<br />

considerably between the isolates and ranged between 20 o C and 30 o C. The activity<br />

<strong>of</strong> all isolates decreased drastically at temperatures from 30 o C to 33 o C, being at<br />

33 o C only 7 % <strong>of</strong> that at 30 o C. The highest between-isolate variations in DR were<br />

at the extremes <strong>of</strong> the applied temperature scale, at 33°C and 6°C.<br />

The DR <strong>of</strong> the four subpopulations did not differ significantly from each other<br />

at any temperature, neither was found any variation according to the age <strong>of</strong> the<br />

cultures (0.3 - 16 y) (ANOVA). However, the Italian and Siberian isolates were<br />

collected from several locations in which the climate varied considerably, and the<br />

highest monthly average temperature <strong>of</strong> each district explained partly the DR <strong>of</strong> the<br />

isolates at 6 o C (p = 0.017). When only the Italian isolates were included in the<br />

ANOVA, a similar significant variation was found (p = 0.043, n = 15). The highest<br />

monthly average temperature <strong>of</strong> the location correlated negatively with the DR <strong>of</strong><br />

H. parviporum at 6 o C. Hence, local climate affects significantly the DR <strong>of</strong> H.<br />

parviporum.<br />

Gene variation <strong>of</strong> different isolates was studied with six microsatellites and by<br />

determining the DNA sequences <strong>of</strong> three sequence characterized amplified regions.<br />

Interestingly, no significant genetic variation was found between Italian, Danish<br />

and Finnish isolates. This suggests that there is significant gene flow between these<br />

subpopulations <strong>of</strong> H. parviporum, and that the variation in DR at 6 o C between<br />

isolates from different localities may be a consequence <strong>of</strong> other factors than<br />

restricted gene flow.<br />

261


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 262<br />

ATTEMPTS TO NATURALLY REGENERATE RED PINE CAN BE<br />

THREATENED BY<br />

DIPLODIA SHOOT BLIGHT DAMAGE TO UNDERSTORY SEEDLINGS<br />

B. W. OBLINGER 1 , Denise R. SMITH 1 , Glen R. STANOSZ 1*<br />

Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Wisconsin-Madison, WI, 53706, USA<br />

* grs@plantpath.wisc.edu<br />

Changes in red pine (Pinus resinosa) management, due to aesthetic and<br />

biodiversity concerns, include creation <strong>of</strong> harvest units with irregular edges, long<br />

borders <strong>of</strong> mature trees, and retention <strong>of</strong> some overstory trees within a harvested<br />

area. Also, in contrast to traditional even-aged management in which trees <strong>of</strong> one<br />

age class are grown, clearcut at final harvest, and replaced by planted seedlings,<br />

there is increasing interest in natural regeneration and developing multi-aged red<br />

pine stands. However, crowns <strong>of</strong> red pines can be sources <strong>of</strong> abundant inoculum <strong>of</strong><br />

the shoot blight pathogen Diplodia pinea. To determine if Diplodia shoot blight<br />

threatens young, naturally regenerated red pine in the understory, six replicate plots<br />

were established in each <strong>of</strong> four mature plantations in central Wisconsin. The<br />

frequency <strong>of</strong> standing, dead seedlings bearing shoot blight symptoms or signs <strong>of</strong><br />

the pathogen, and the incidence and severity <strong>of</strong> shoot blight damage to live<br />

seedlings were recorded. Mean seedling mortality ranged from 13-30% and mean<br />

incidence <strong>of</strong> blighted living seedlings ranged from 94-100% at all sites. The mean<br />

frequency <strong>of</strong> live seedlings with their terminal leaders killed in the past was from<br />

55-94%. Mean severity <strong>of</strong> damage to live seedlings, on a 0-3 scale, was ÿ2.16 at<br />

all sites. Results <strong>of</strong> a PCR assay confirmed pathogen identity. These results<br />

support previous research and concern that shoot blight pathogens threaten young<br />

red pines in the understory.<br />

262


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 263<br />

SOME FUNGAL SPECIES ON Pinus pinaster Ait. AND Pinus radiata<br />

D. Don PLANTATIONS IN MARMARA REGION OF TURKEY<br />

Fazıl Selek 1*<br />

1 Poplar and Fast Growing Forest Trees Research Institute<br />

P.O. Box 93 41001 Kocaeli/Turkey<br />

* selek@kavak.gov.tr<br />

The global forest areas decrease with the growing world population. The<br />

protection <strong>of</strong> the natural forest is necessary and the most <strong>of</strong> wood production<br />

should be done from outside <strong>of</strong> the natural forest.<br />

The industrial plantation give yields about 13,8-55,9 m3/ha/year for rotation<br />

between 10-30 years. The industrial plantations entrepreneurship gains importance<br />

as an effective sector in the countriers such as Chile, Brasil, USA, South African,<br />

New Zealand, Australia, South Korea<br />

Regular afforestation began in 1955 in Turkey. The plantation <strong>of</strong> Forestry began<br />

in 1963. The studies and researchs on Fast Growing Forest trees were given to the<br />

Poplar Research Institute from General Directorate <strong>of</strong> Forestry in 1968<br />

Marmara Region is located on the the Northwest <strong>of</strong> Turkey. The East<br />

Longitudes <strong>of</strong> region are 25 0 50 ’ -30 0 55 ’ , the North latitudes <strong>of</strong> region are 39 0 06 ’ -<br />

42 0 05 ’<br />

The area <strong>of</strong> the Marmara Region is 67300 Km 2<br />

Totally in the Turkey, there are 42 185 hectares plantation areas <strong>of</strong> Pinus<br />

pinaster and 949 hectares plantation areas <strong>of</strong> Pinus radiata<br />

The aim <strong>of</strong> this study is to determine the harmful fungi on P. pinaster and P.<br />

Radiata plantations <strong>of</strong> the Marmara Region. The observation in plantation were<br />

done in 2003. Until now, the following fungal species have been identified: Pluteus<br />

plautus, Lenzites striata, Pleurotus astreatus, Schizophyllum commune and Streum<br />

purpuretum<br />

Keywords: Pinus pinaster, Pinus radiata plantations, harmful fungi.<br />

263


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 264<br />

RESPONSE OF Alnus tenuifolia TO INOCULATION WITH Valsa<br />

melanodiscus<br />

Glen R. STANOSZ 1* , L. M. TRUMMER 2 , J. K. ROHRS-RICHEY 3 , G. C.<br />

ADAMS 4 , J. T. WORRALL 5<br />

1 Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Wisconsin-Madison, WI, 53706, USA<br />

2 USDA Forest Service, Forest Health Protection, Anchorage, AK, 99503, USA<br />

3 Department <strong>of</strong> Biology and Institute <strong>of</strong> Arctic Biology, University <strong>of</strong> Alaska Fairbanks, AK, 99775,<br />

USA<br />

4 Department <strong>of</strong> Plant Pathology, Michigan State University, East Lansing, MI, 48824, USA<br />

5 USDA Forest Service, Forest Health Management, Gunnison, CO, 81230, USA<br />

* grs@plantpath.wisc.edu<br />

Valsa melanodiscus (anamorph Cytospora umbrina) is associated with cankered<br />

and killed alder (Alnus) stems in western North America from Colorado to Alaska.<br />

The responses <strong>of</strong> thinleaf alder (A. tenuifolia) stems to inoculation with each <strong>of</strong> two<br />

isolates <strong>of</strong> V. melanodiscus were studied in south-central Alaska. At each <strong>of</strong> two<br />

sites, eight stems per isolate were wounded to expose both inner bark and sapwood<br />

and inoculated in early May 2007 by placing a colonized agar plug over the wound.<br />

Sterile agar plugs were applied to wounded control stems. Sunken, elongated<br />

cankers similar to those with which V. melanodiscus has been associated resulted<br />

on inoculated stems. In contrast, wounded control stems exhibited strong callus<br />

production and wound closure. In September 2007, cankers were harvested and<br />

lengths were recorded. Mean canker lengths measured externally (data for both<br />

isolates pooled) at the two sites were 45 (range 20-156) mm and 73 (range 22-201)<br />

mm. Analysis <strong>of</strong> variance <strong>of</strong> log transformed data revealed strong support for effect<br />

<strong>of</strong> location (P = 0.04), but not that <strong>of</strong> isolate (P = 0.12) or interaction (P = 0.20) on<br />

canker length. The fungus was reisolated from each inoculated stem, but not from<br />

any control stem. The ability <strong>of</strong> V. melanodiscus to cause cankers on thinleaf alder<br />

stems is confirmed, and these results support the conclusion that this pathogen is a<br />

cause <strong>of</strong> alder dieback in western North America.<br />

264


SDU Faculty <strong>of</strong> Forestry Journal<br />

Serial: A, Number: Special Issue, Year: <strong>2009</strong>, ISSN: 1302-7085, Page: 265<br />

GREMMENIELLA INFECTIONS ON SEEDLINGS AFTER<br />

REPLANTING SEVERELY INFECTED PINE FOREST<br />

Elna STENSTRÖM 1* , Maria JONSSON 1 , Kjell WAHLSTRÖM 1<br />

1 Department <strong>of</strong> Forest Mycology and Pathology, Swedish University <strong>of</strong> Agricultural Sciences,<br />

Uppsala, Sweden<br />

*elna.stenstrom@mykopat.slu.se<br />

During 1999 and 2001 the most severe Gremmeniella abietina epidemic ever<br />

appeared in Sweden. More than 300 000 ha forest were severely attacked and the<br />

forest industry lost milliards <strong>of</strong> SEK. Big forest areas, with at least 50 000 ha,<br />

needed to be clear cut in advance followed by replanting.<br />

For the out-planting experiment pine seedlings were planted on three different<br />

locations in Dalarna in Sweden, that were clear-cut in advance due to severe<br />

Gremmeniella infection. The forest had been clear-cut in 2001 and this study was<br />

conducted during 2002-2005. Each site contained one clear cut area and a nearby<br />

Gremmeniella infected forest. Seedlings were planted on clear-cut areas with and<br />

without remaining twigs and branches, at the edge <strong>of</strong> the clear-cut areas, as well as<br />

in the adjacent forest. In total at least 200 seedlings were planted on each area. The<br />

areas were replanted every year with new one year old seedlings received from a<br />

forest nursery. The disease incident was determined visually the year after<br />

plantation and then the infection was confirmed with PCR using Gremmeniella<br />

specific primers.<br />

For seedling planted at the clear cut areas, the infection decreased from 50 -<br />

90% infected seedlings planted one year after felling to 0 - 55% planted the second<br />

year after felling and to 0 -38% for seedling planted three years after felling. After<br />

four years there was almost no infection on the clear-cut areas. The variation was<br />

big between the sites. Two and three years after felling there were almost no<br />

differences between seedlings planted on areas with or without twigs and branches.<br />

However seedling planted in the adjacent diseased forest became much more<br />

infected then seedlings planted on the clear cut areas. For seedlings planted in the<br />

forest one year after felling almost all seedlings became infected. Two and three<br />

years after felling they became infected to up to 50 % and even four years after<br />

felling 15-40 % <strong>of</strong> the seedlings became infected. The seedlings planted close to<br />

the forest edge were always more infected then the seedlings out on the clear-cut<br />

areas but less infected than the seedlings in the forest.<br />

It is remarkable that the disease incidence on seedlings in the forest still is high<br />

after 3-4 years but we assume that the snow cover during winter have promoted the<br />

infections in small seedlings. From inventory in mature forest in the same areas it<br />

265


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

has been found that the disease incidence decreased after some years after the<br />

severe outbreaks in 1999 and 2001. In spite <strong>of</strong> this it means that several years after<br />

a severe infection there is still a lot <strong>of</strong> viable spores left, particularly in the forest.<br />

More then two years after felling it seems to be a trend that seedlings planted in the<br />

middle <strong>of</strong> the clear-cut areas were less infected then seedling planted close to the<br />

forest. The variation between infections in the forest on the different sites was low<br />

compared to infections on the clear cut areas. The general conclusion is that it is<br />

advisable to wait with replanting with pine seedlings at least two years after felling<br />

severely Gremmeniella infected stands. Then the risk <strong>of</strong> infection form twigs and<br />

branches left on the clear cut area also is smaller.<br />

266


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

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

Anna Hopkins<br />

Scion (New Zealand Forest research Institute)<br />

Scion, Private Bag 3020, Rotorua 3010, New Zealand<br />

Anna.Hopkins@scionresearch.com<br />

Antti Uotila<br />

Helsinki University, Hyytiälä Forestry Field Station<br />

Hyytiäläntie 124, 35500 Korkeakoski, Finland<br />

antti.uotila@helsinki.fi<br />

Asko Lehtijärvi<br />

Suleyman Demirel University<br />

SDU, Faculty <strong>of</strong> <strong>forestry</strong>, Dept. <strong>of</strong> Botany, 32260, Cunur, Isparta, Turkey<br />

asko@orman.<strong>sdu</strong>.edu.tr<br />

Ayhan Karakaya<br />

Poplar and Fast Growing Forest Trees Research Institute<br />

Kavakçılık Araştırma Müdürlüğü 41001 Başiskele, Kocaeli, Türkiye<br />

ayhankarakaya68@hotmail.com<br />

Ayşe Gülden Aday<br />

Suleyman Demirel University<br />

SDU, Faculty <strong>of</strong> <strong>forestry</strong>, Dept. Of siviculture, 32260, Cunur, Isparta, Turkey<br />

guldenaday@orman.<strong>sdu</strong>.edu.tr<br />

Banu Karabıyık<br />

T. C. <strong>Orman</strong> Genel Müdürlüğü<br />

<strong>Orman</strong> İdaresi ve Planlama Dairesi Başkanlığı, 7 nolu Bina <strong>Orman</strong> Genel<br />

Müdürlüğü Tesisleri Gazi Ankara, Türkiye<br />

bnkarabiyik@hotmail.com<br />

Berthold Metzler<br />

Forest Research Institute Baden-Wuerttemberg<br />

Wonnhaldestr, 4D-79100 Freiburg/Br. Germany<br />

berthold.metzler@forst.bwl.de<br />

Dagmar Palovcikova<br />

Dpt. <strong>of</strong> Forest Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood<br />

Technology, Mendel University <strong>of</strong> Agriculture and Forestry, Brno<br />

Zemedelska 3, 613 00 Brno, Czech Republic<br />

palovcik@mendelu.cz<br />

Denise Smith<br />

University <strong>of</strong> Wisconsin-Madison<br />

Dept. <strong>of</strong> Plant Pathology, 1630 Linden Drive, Madison, WI 53706<br />

dzs@plantpath.wisc.edu<br />

267


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Elna Stenström<br />

Department <strong>of</strong> forest mycology and pathology<br />

SLU, Box 7026, S-75007 Uppsala, Sweden<br />

elna.stenstrom@mykopat.slu.se<br />

Erhard Halmschlager<br />

İnstitute <strong>of</strong> forest Entomology, Forest Pathology and Forest Protection, Dept. Of<br />

Forest and Soil Sciences, University <strong>of</strong> Natural Resources and Applied Life<br />

Sciences Vienna (BOKU), Vienna<br />

Hasenauerstraße 38 A-1190 Wien<br />

erhard.halmschlager(at)boku.ac.at<br />

Eva Alfonso Corzo<br />

Universidad Politécnica de Madrid, Escuela Técnica Superior de Ingenieros de<br />

Montes Plaza de la Constitución, 8 Bajo Derecha Alcorcón, MADRID (Spain)<br />

alfonso_corzo@hotmail.com<br />

Faruk Ş. Özay<br />

Poplar And Fast Growing Forest Tree Researc İnstitude<br />

Kavakçılık Araştırma Müdürlüğü 41001 Başisleke /Kocaeli, Türkiye<br />

Fax: 0262 311 69 72<br />

Faruk@kavak.gov.tr<br />

Fazıl SELEK<br />

Poplar and Fast Growing Forest Trees Research Institute<br />

Kavakçılık Araştırma Müdürlüğü 41001 Başiskele/ KOCAELİ, TÜRKİYE<br />

selek@kavak.gov.tr<br />

Funda Oskay<br />

Suleyman Demirel University<br />

SDU, Faculty <strong>of</strong> <strong>forestry</strong>, Dept. <strong>of</strong> Botany, 32260, Cunur, Isparta, Turkey<br />

foskay@orman.<strong>sdu</strong>.edu.tr<br />

Gaston Laflamme<br />

Canadian Forest Service<br />

1055 rue du P.E.P.S. Quebec, Qc Canada GIV 4C7<br />

Laflamme@rncan.gc.ca<br />

Gürsel Karaca<br />

Suleyman Demirel University<br />

SDU, Faculty <strong>of</strong> Agriculture, Dept. Of plant protection, 32260, Cunur, Isparta,<br />

Turkey<br />

gkaraca@ziraat.<strong>sdu</strong>.edu.tr<br />

Irmtraut Zaspel<br />

Institute <strong>of</strong> Forest Genetics<br />

Eberswalder Chausee 3A, 15377 Waldsieversdorf, Germany<br />

irmtraut.zaspel@vti.bund.de<br />

268


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Juha Kaitera<br />

Finnish Forest Research Institute<br />

Muhos Research Unit, Kirkkosaarentie 7, FI-91500 Muhos, Finland<br />

juha.kaitera@metla.fi<br />

Julio Javier Diez Casero<br />

University Of Valladolid, Spain<br />

Dpt. Producción Vegetal, Campus “La Yutera” Edif.<br />

E. Av. Madrid, 44, 44004. Palencia. Spain<br />

Jdcasero@Pvs.Uva.Es<br />

Kazım ULUER<br />

Poplar and Fast Growing Forest Trees Research Institute<br />

Kavakçılık Araştırma Müdürlüğü 41001 Başiskele/ KOCAELİ, TÜRKİYE<br />

uluer@kavak.gov.tr<br />

Leticia Botella Sánchez<br />

University Of Valladolid, Spain<br />

Dpt. Producción Vegetal, Campus “La Yutera” Edif.<br />

E. Av. Madrid, 44, 44004. Palencia. Spain<br />

Lbotella@pvs.uva.es<br />

Libor Jankovsky<br />

Dpt. <strong>of</strong> Forest Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood<br />

Technology, Mendel University <strong>of</strong> Agriculture and Forestry, Brno<br />

Zemedelska 3, 613 00 Brno, Czech Republic<br />

jankov@mendelu.cz<br />

Maria Jesus Garcia-Garcia<br />

Technical University <strong>of</strong> Madrid, Madrid, Spain<br />

Dpto. Proyectos y Planificacion Rural. E.T.S.I. de<br />

Montes. Universidad Politecnica de Madrid. Ciudad Universitaria s/n - 28040. Madrid.<br />

Spain<br />

mariajesus.garcia.garcia@upm.es<br />

Martti Vuorinen<br />

The Finnish forest Research Institute<br />

Metla, Suonenjoki Research Unit, Juntintie 154, 77600 Suonenjok, Finland<br />

martti.vuorinen@metla.fi<br />

Mertcan Karadeniz<br />

Suleyman Demirel University<br />

SDU, Faculty <strong>of</strong> <strong>forestry</strong>, Dept. <strong>of</strong> Botany, 32260, Cunur, Isparta, Turkey<br />

karadenizmertcan@gmail.com<br />

Michael Müller<br />

The Finnish forest Research Institute<br />

Finnish Forest Resaerch Institute, P. O. Box 18, 01301 Vantaa, Finland<br />

Michael.Mueller@metla.fi<br />

269


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

Michal Tomsovsky<br />

Dpt. <strong>of</strong> Forest Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood<br />

Technology, Mendel University <strong>of</strong> Agriculture and Forestry, Brno Zemedelska 3, 613<br />

00 Brno, Czech Republic<br />

tomsovsk@mendelu.cz<br />

Milon Dvorak<br />

Dpt. <strong>of</strong> Forest Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood<br />

Technology, Mendel University <strong>of</strong> Agriculture and Forestry, Brno<br />

Zemedelska 3, 613 00 Brno, Czech Republic<br />

klobrc@centrum.cz<br />

Mustafa Uygun<br />

Konya <strong>Orman</strong> Bölge Müdürlüğü<br />

Konya, Turkey<br />

mustafauygun@ogm.gov.tr<br />

Nevzat Gürlevik<br />

Suleyman Demirel University<br />

SDU, Faculty <strong>of</strong> <strong>forestry</strong>, Dept. Of siviculture, 32260, Cunur, Isparta, Turkey<br />

gulevik@orman.<strong>sdu</strong>.edu.tr<br />

Nikica Ogris<br />

Slovenian Forestry Institute<br />

Slovenian Forestry Institute, Nikica Ogris, Vecna pot 2, 1000 Ljubljana, Slovenia<br />

nikica.ogris@gozdis.si<br />

Oscar Santamaría<br />

University <strong>of</strong> Extremadura<br />

Escuela de Ingenierías Agrarias. Ctra. de Cáceres s/n, 06071 Badajoz, SPAIN<br />

osantama@unex.es<br />

Pablo Martínez Álvarez<br />

University <strong>of</strong> Valladolid<br />

Departamento de Producción Vegetal y Silvopascicultura. Campus La Yutera.<br />

Edificio E. Avda. Madrid 44 34004 Palencia (Spain)<br />

pmtnez@pvs.uva.es<br />

Paolo Capretti<br />

DiBA Department <strong>of</strong> Agricultural Biotechnology Sect. <strong>of</strong> Plant Pathology.<br />

Piazzale delle Cascine, 28 50144 FIRENZE ITALY<br />

paolo.capretti@unifi.it<br />

Petr Stastny<br />

Dpt. <strong>of</strong> Forest Protection and Wildlife Management, Faculty <strong>of</strong> Forestry and Wood<br />

Technology, Mendel University <strong>of</strong> Agriculture and Forestry, Brno<br />

Zemedelska 3, 613 00 Brno, Czech Republic<br />

svezi.mirka@email.cz<br />

270


SDÜ ORMAN FAKÜLTESİ DERGİSİ<br />

Pia Barklund<br />

Department <strong>of</strong> Forest Mycology and Pathology, SLU<br />

Box 7026, SE-75007 Uppsala, Sweden<br />

Pia.Barklund@mykopat.slu.se<br />

Sabrina Reignoux<br />

The University <strong>of</strong> Edinburgh Institute <strong>of</strong> Evolutionary Biology<br />

West Main Road, Edinburgh<br />

EH9 3JT<br />

s0676163@sms.ed.ac.uk<br />

Sarah Green<br />

Forest Research<br />

Northern Research Station, Roslin, Midlothian, Scotland EH25 9SY<br />

sarah.green@<strong>forestry</strong>.gsi.gov.uk<br />

Tom Hsiang<br />

Department <strong>of</strong> Environmental Biology, University <strong>of</strong> Guelph, Guelph, Ontario,<br />

Canada<br />

thsiang@uoguelph.ca<br />

Tuğba Doğmuş-Lehtijärvi<br />

Suleyman Demirel University<br />

SDU, Faculty <strong>of</strong> <strong>forestry</strong>, Dept. <strong>of</strong> Botany, 32260, Cunur, Isparta, Turkey<br />

tugba@orman.<strong>sdu</strong>.edu.tr<br />

Yasuyuki Hiratsuka<br />

Canadian Forest service<br />

Nothern Forestry Centre 12304 66A Avenue, Edmonton, Alberta T6H 1Z3, Canada<br />

Yasu.Hiratsuka@NRCan-RNCan.gc.ca<br />

271


SDÜ Faculty <strong>of</strong> Forestry Journal<br />

272

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!