04.03.2014 Views

Proceedings of the meeting - IOBC-WPRS

Proceedings of the meeting - IOBC-WPRS

Proceedings of the meeting - IOBC-WPRS

SHOW MORE
SHOW LESS

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

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

<strong>IOBC</strong> / <strong>WPRS</strong><br />

Working Group „Integrated Control in Cereal Crops“<br />

OILB / SROP<br />

Groupe de Travail „Lutte Intégrée en Céréales“<br />

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

at<br />

Gödöllö, Hungary<br />

9 - 12 September, 1999<br />

editors:<br />

Christian Borgemeister & Hans-Michael Poehling<br />

<strong>IOBC</strong> wprs Bulletin<br />

Bulletin OILB srop Vol. 24 (4) 2001


The <strong>IOBC</strong>/<strong>WPRS</strong> Bulletin is published by <strong>the</strong> International Organization for Biological and Integrated<br />

Control <strong>of</strong> Noxious Animals and Plants, West Palearctic Regional Section (<strong>IOBC</strong>/<strong>WPRS</strong>)<br />

Le Bulletin OILB/SROP est publié par l‘Organisation Internationale de Lutte Biologique et Intégrée<br />

contre les Animaux et les Plantes Nuisibles, section Regionale Ouest Paléarctique (OILB/SROP)<br />

Copyright: <strong>IOBC</strong>/<strong>WPRS</strong> 2001<br />

The Publication Commission <strong>of</strong> <strong>the</strong> <strong>IOBC</strong>/<strong>WPRS</strong>:<br />

Horst Bathon<br />

Federal Biological Research Center<br />

for Agriculture and Forestry (BBA)<br />

Institute for Biological Control<br />

Heinrichstr. 243<br />

D-64287 Darmstadt (Germany)<br />

Tel +49 6151 407-225, Fax +49 6151 407-290<br />

e-mail: h.bathon.biocontrol.bba@t-online.de<br />

Luc Tirry<br />

University <strong>of</strong> Gent<br />

Laboratory <strong>of</strong> Agrozoology<br />

Department <strong>of</strong> Crop Protection<br />

Coupure Links 653<br />

B-9000 Gent (Belgium)<br />

Tel +32-9-2646152, Fax +32-9-2646239<br />

e-mail: luc.tirry@rug.ac.be<br />

Address General Secretariat:<br />

INRA – Centre de Recherches de Dijon<br />

Laboratoire de recherches sur la Flore Pathogène dans le Sol<br />

17, Rue Sully, BV 1540<br />

F-21034 DIJON CEDEX<br />

France<br />

ISBN 92-9067-135-8


i<br />

Preface<br />

This bulletin contains most <strong>of</strong> <strong>the</strong> papers presented on <strong>the</strong> last biennial <strong>meeting</strong> <strong>of</strong> <strong>the</strong><br />

<strong>IOBC</strong>/<strong>WPRS</strong> working group “Integrated Control in Cereal Crops”. The <strong>meeting</strong> took place at<br />

Gödöllö University <strong>of</strong> Agricultural Sciences, Faculty <strong>of</strong> Agricultural Sciences, Department <strong>of</strong><br />

Plant Protection, Hungary, from 09 to 12 September 1999. The <strong>meeting</strong> was hosted by Pr<strong>of</strong>.<br />

Dr. Jozef Kiss. First <strong>of</strong> all on <strong>the</strong> behalf <strong>of</strong> our group I want to express my deepest gratitude to<br />

Jozef Kiss and his staff who organised an excellent <strong>meeting</strong> and a very pleasant stay for us in<br />

Gödöllö. Also thanks to <strong>the</strong> supporting organisations, i.e. <strong>IOBC</strong>, Gödöllö University <strong>of</strong><br />

Agricultural Sciences, The Ministry <strong>of</strong> Agriculture & Regional Development <strong>of</strong> Hungary and<br />

OMFB (National Committee for Technological Development <strong>of</strong> Hungary). More than 30<br />

participants attended <strong>the</strong> <strong>meeting</strong> and spent fruitful days to present papers, discuss scientific<br />

progress and establish new contacts. Particularly <strong>the</strong> increasing number <strong>of</strong> participants from<br />

eastern and sou<strong>the</strong>rn European countries was stimulating. However, quite obviously<br />

integrated control <strong>of</strong> cereal pests is no longer a major research topic in many middle and<br />

nor<strong>the</strong>rn European countries that previously dominated our working group. This situation<br />

needs to be critically discussed on <strong>the</strong> next <strong>meeting</strong> to assure a new orientation <strong>of</strong> <strong>the</strong> working<br />

group.<br />

Finally I am extremely thankful to Christian Borgemeister for his tremendous effort in<br />

compiling <strong>the</strong>se proceedings.<br />

Hans-Michael Poehling, Convenor<br />

Hannover University


ii<br />

List <strong>of</strong> <strong>the</strong> Participants<br />

AFONINA, V.M.<br />

ALBAJES, R.<br />

BASKY, Z.<br />

BERECS-BAHDI, G.<br />

FIEBIG, M.<br />

FREIER, B.<br />

GALLER, M.<br />

GOSSELKE, U.<br />

GUEORGUIEVA, T.<br />

GOTLIN CULIAK, T.<br />

Moscow State University<br />

Faculty <strong>of</strong> Biology, Dept. Entomology<br />

Moscow 119899, RUSSIA<br />

Centre UdL-IRTA.<br />

Area de Protecció de Conreus<br />

Rovira Roure, 177<br />

E-25198 Lleida, SPAIN<br />

Plant Protection Institute<br />

Hungarian Academy <strong>of</strong> Sciences<br />

P.O.Box 102<br />

H-1525 Budapest, HUNGARY<br />

Plant Health and Soil Conservation Station<br />

H-2100 Gödöllő, HUNGARY<br />

Institut für Pflanzenkrankheiten und Pflanzenschutz<br />

Universität Hannover<br />

Herrenhäuser Str. 2<br />

D-30419 Hannover, GERMANY<br />

BBA - Institut für integrierten Pflanzenschutz<br />

Stahnsdorfer Damm 81<br />

D-14532 Kleinmachnow, GERMANY<br />

Institut für Pflanzenkrankheiten und Pflanzenschutz<br />

Universität Hannover<br />

Herrenhäuser Str. 2<br />

D-30419 Hannover, GERMANY<br />

BBA - Institut für integrierten Pflanzenschutz<br />

Stahnsdorfer Damm 81<br />

D-14532 Kleinmachnow, GERMANY<br />

Agricultural University, Dept. <strong>of</strong> Entomology<br />

12 Mendeleev St.<br />

4000 - Plovdiv, BULGARIA<br />

Faculty <strong>of</strong> Agriculture<br />

Dept. <strong>of</strong> Zoology<br />

Svetosimunska 25<br />

10000 Zagreb, CROATIA


iii<br />

HATVANI, A.<br />

HULLÉ, M.<br />

HUUSELA-VEISTOLA, E.<br />

IGRC-BARČIĆ, J.<br />

KISS, J.<br />

KOZMA, E.<br />

KROMP, B.<br />

LHALOUI, S.<br />

LUCZA, Z.<br />

MAKKÓ, V.<br />

MATEEVA-RADEVA, A.<br />

MEINDL, P.<br />

MIHÁLY, B.<br />

PAPP, E.<br />

University <strong>of</strong> Horticulture<br />

Erdei F. tér 1-3<br />

H-6000 Kecskemét, HUNGARY<br />

INRA, Laboratoire de Zoologie<br />

Domaine de la Motte-au-Vicomte, BP 29<br />

F-356530 Le Rheu Cedex, FRANCE<br />

Agricultural Research Centre <strong>of</strong> Finland<br />

Plant Protection Research<br />

FIN-31600 Jokioinen, FINLAND<br />

Faculty <strong>of</strong> Agriculture<br />

Dept. <strong>of</strong> Zoology<br />

Svetosimunska 25<br />

10000 Zagreb, CROATIA<br />

Szent István University<br />

Páter K u. 1<br />

H-2100 Gödöllő, HUNGARY<br />

University <strong>of</strong> Agricultural Sciences<br />

Dept. <strong>of</strong> Plant Protection<br />

Pater U u. 1<br />

H-2100 Gödöllő, HUNGARY<br />

L. Boltzmann Institut für biologische Landwirtschaft<br />

Rinnboeckstr. 15<br />

A-1110 Wien, AUSTRIA<br />

INRA-CRPA-Settat<br />

P.O. Box 589<br />

Settat, MOROCCO<br />

Plant Health and Soil Conservation Station Sapitol<br />

H-1119 Budapest, HUNGARY<br />

University <strong>of</strong> Horticulture, Dept. <strong>of</strong> Entomology<br />

Mènes. ut 44<br />

H-1118 Budapest, HUNGARY<br />

Agricultural University, Dept. <strong>of</strong> Entomology<br />

12 Mendeleev St.<br />

4000 - Plovdiv, BULGARIA<br />

L. Boltzmann Institut für biologische Landwirtschaft<br />

Rinnboeckstr. 15<br />

A-1110 Wien, AUSTRIA<br />

Institute for Nature Conservation <strong>of</strong> <strong>the</strong><br />

Institute for Environmental Management<br />

Költő u. 21<br />

H-1121 Budapest, HUNGARY<br />

Plant Health and Soil Conservation Station<br />

H-1119 Budapest, HUNGARY


iv<br />

POEHLING, H.-M.<br />

PONS, X.<br />

SAMU, F.<br />

SZAKÀL, M.<br />

SZENTKIRÁLYI, F.<br />

TÓKÉS, G.<br />

TÓTH, F.<br />

TRILTSCH, H.<br />

TSHERNYSHEV, W.<br />

VÖRÖS, G.<br />

WINKLER, I.<br />

Institut für Pflanzenkrankheiten und Pflanzenschutz<br />

Universität Hannover<br />

Herrenhäuser Str. 2<br />

D-30419 Hannover, GERMANY<br />

Centre UdL-IRTA<br />

Area de Protecció de Conreus<br />

Rovira Roure, 177<br />

E-25198 Lleida, SPAIN<br />

Hungarian Academy <strong>of</strong> Sciences<br />

Dept. <strong>of</strong> Plant Protection<br />

PO Box 102<br />

H-1525 Budapest, HUNGARY<br />

Plant Health and Soil Conservation Station<br />

Kòtlán ε u. 3<br />

H-2100 Gödöllő, HUNGARY<br />

Institute <strong>of</strong> Plant Protection H.A.S.<br />

Dept. <strong>of</strong> Zoology<br />

P.O. Box 102<br />

H-1525 Budapest, HUNGARY<br />

Plant Health and Soil Conservation Station<br />

Budaörsi út 141<br />

H-1118 Budapest, HUNGARY<br />

Szent István University<br />

Páter K u. 1<br />

H-2100 Gödöllő, HUNGARY<br />

BBA - Institut für integrierten Pflanzenschutz<br />

Stahnsdorfer Damm 81<br />

D-14532 Kleinmachnow, GERMANY<br />

Moscow State University<br />

Faculty <strong>of</strong> Biology, Dept. Entomology<br />

Moscow 119899, RUSSIA<br />

Tolna County Plant Health and Soil<br />

Conservation Station, SZEKSZARD<br />

Plant Health and Soil Conservation Station<br />

Kòtlán ε u. 3<br />

H-2100 Gödöllő, HUNGARY


v<br />

Contents<br />

Variability in <strong>the</strong> timing <strong>of</strong> sexual morph production in <strong>the</strong> aphid Rhopalosiphum padi<br />

Hullé, M., D. Maurice, V. Stevoux, J. Bonhomme, C. Rispe & J.-Chr. Simon ................. 1<br />

Biotypic variation <strong>of</strong> Diuraphis noxia (Homoptera: Aphididae) between South Africa<br />

and Hungary<br />

Basky, Z. & J. Jordaan ...................................................................................................... 9<br />

Impact <strong>of</strong> barley yellow dwarf virus infection on physiological conditions <strong>of</strong> wheat and<br />

<strong>the</strong> consequences for cereal aphids attack<br />

Fiebig, M. & H.-M. Poehling .......................................................................................... 25<br />

The Russian wheat aphid on barley in Morocco:<br />

survey and identification <strong>of</strong> new sources <strong>of</strong> resistance<br />

Lhaloui, S., M. El Bouhssini, S. Ceccarelli, S. Grando & A. Amri ................................. 33<br />

Effects <strong>of</strong> induced tolerance and induced resistance against aphids in wheat<br />

Galler, M. & H.-M. Poehling .......................................................................................... 39<br />

How does a ladybird respond to aphids?<br />

Triltsch, H., G. Hechenthaler, U. Gosselke & B. Freier ................................................. 49<br />

Computer simulations on <strong>the</strong> efficiency <strong>of</strong> cereal aphid predators in winter wheat<br />

Gosselke, U., D. Roßberg, H. Triltsch & B. Freier ......................................................... 59<br />

Potentials and limitations <strong>of</strong> long-term field data to identify numerical and functional<br />

responses <strong>of</strong> predators to aphid density in wheat<br />

Freier, B., Triltsch, H. & U. Gosselke............................................................................. 65<br />

Density <strong>of</strong> epigeal predators on maize plants untreated and treated with imidacloprid<br />

Pons, X. & R. Albajes ...................................................................................................... 73<br />

Arthropod natural enemies <strong>of</strong> <strong>the</strong> cereal leaf beetle (Oulema melanopus L.) in organic<br />

winter wheat fields in Vienna, Eastern Austria<br />

Meindl, P., B. Kromp, B. Bartl & E. Ioannidou .............................................................. 79<br />

Habitat preference <strong>of</strong> carabids (Coleoptera: Carabidae) in Central Hungary in winter<br />

wheat field and in adjacent habitats<br />

Hatvani, A., F. Kádár, J. Kiss & G. Péter....................................................................... 87<br />

Role <strong>of</strong> field margin in <strong>the</strong> winter phenophase <strong>of</strong> Carabid beetles (Coleoptera:<br />

Carabidae) in winter wheat field<br />

Péter, G., F. Kádár, J. Kiss & F. Tóth............................................................................. 91<br />

Insect pests <strong>of</strong> cereals in Croatia<br />

Igrc-Barčić, J. & T. Gotlin Culjak................................................................................... 95<br />

The Hessian fly in Morocco: Surveys, loss assessment, and genetic resistance in bread<br />

wheat<br />

Lhaloui, S., M. El Bouhssini & A. Amri......................................................................... 101


vi<br />

Ecological Pest Management (EPM): General Problems<br />

Tshernyshev, W.B........................................................................................................... 109<br />

Hessian fly (Mayetiola destructor Say) damage in relay intercropping <strong>of</strong> cereals in<br />

Finland<br />

Huusela-Veistola, E., A. Vasarainen & J. Grahn.......................................................... 113<br />

Results <strong>of</strong> a nation-wide survey <strong>of</strong> spider assemblages in Hungarian cereal fields<br />

Samu, F., F. Tóth, C. Szinetár, G. Vörös & E. Botos .................................................... 119<br />

Agrobiological and biocenological study <strong>of</strong> winter oats (A. sativa L.)<br />

Gueorguieva, T. & A. Mateeva...................................................................................... 129<br />

Side effects <strong>of</strong> some pesticides on aphid specific predators in winter wheat<br />

Mateeva, A., M. Vassileva & T. Gueorguieva ............................................................... 139<br />

Influence <strong>of</strong> <strong>the</strong> stubble burning on some pests and earth-worms density<br />

Mateeva, A., D. Svetleva, D. Andonov & St. Stratieva .................................................. 143<br />

A study <strong>of</strong> aphid predation by Coccinella septempunctata L. (Coleoptera:<br />

Coccinellidae) using gut dissection<br />

Triltsch, H...................................................................................................................... 147<br />

Arthropod complex <strong>of</strong> winter wheat crops and its seasonal dynamics<br />

Afonina, V.M., W.B. Tshernyshev, I.I. Soboleva-Dokuchaeva, A.V. Timokhov,<br />

O.V. Timokhova & R.R. Seifulina.................................................................................. 153<br />

Studies <strong>of</strong> <strong>the</strong> pests <strong>of</strong> Canary-grass (Phalaris canariensis L.)<br />

Kozma, E., G. Gólya & Z. Záhorszki.............................................................................. 165


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 1 - 7<br />

Variability in <strong>the</strong> timing <strong>of</strong> sexual morph production<br />

in <strong>the</strong> aphid Rhopalosiphum padi<br />

Maurice Hullé, Damien Maurice, Véronique Stevoux, Joel Bonhomme, Claude Rispe<br />

and Jean-Christophe Simon<br />

INRA, Laboratoire de Zoologie, 35653 Le Rheu Cedex, France<br />

Summary<br />

In host-alternating aphid species, <strong>the</strong> latest a clone switches to sexual production <strong>the</strong> highest is its rate<br />

<strong>of</strong> increase because <strong>the</strong> par<strong>the</strong>nogenetic phase is longer. Never<strong>the</strong>less clones are constrained to lay<br />

eggs before leaf fall <strong>of</strong> <strong>the</strong> primary host. In this paper, we have studied <strong>the</strong> variation in sexual morph<br />

production <strong>of</strong> several cyclical par<strong>the</strong>nogenetic clones <strong>of</strong> <strong>the</strong> aphid Rhopalosiphum padi which<br />

alternate between cereals and Prunus padus, <strong>the</strong> bird cherry. Ten clones <strong>of</strong> this species differing by<br />

<strong>the</strong>ir geographic origin (five from western and five from eastern France) were placed in two<br />

laboratory-simulated environments, mimicking <strong>the</strong> changes <strong>of</strong> photo-period and <strong>the</strong>rmo-period<br />

occurring naturally from <strong>the</strong> end <strong>of</strong> summer and during <strong>the</strong> autumn in oceanic and continental<br />

conditions. The analysis <strong>of</strong> clonal responses in both climatic conditions showed (i) no geographic<br />

adaptation among clones, (ii) an earlier production <strong>of</strong> sexuals in continental conditions and (iii) a<br />

higher production <strong>of</strong> males in oceanic conditions. Fur<strong>the</strong>rmore, we have compared <strong>the</strong> dates <strong>of</strong> first<br />

appearance <strong>of</strong> sexuals in our experiments with those occurring in <strong>the</strong> field based on suction trap<br />

database. Sexuals were observed in nature at least four weeks earlier than in <strong>the</strong> lab. Placed in seminatural<br />

conditions, <strong>the</strong> responses <strong>of</strong> <strong>the</strong> clones were similar to those observed under field conditions.<br />

These results underline <strong>the</strong> need for a better understanding <strong>of</strong> <strong>the</strong> influence <strong>of</strong> <strong>the</strong> whole array <strong>of</strong><br />

environmental factors, inducing <strong>the</strong> transition from par<strong>the</strong>nogenetic to sexual reproduction in aphids.<br />

Key words: Rhopalosiphum padi, geographic clones, overwintering, reproductive strategy, <strong>the</strong>rmoperiod,<br />

photo-period<br />

Introduction<br />

In cyclical par<strong>the</strong>nogenetic populations <strong>of</strong> many host-alternating species, <strong>the</strong> par<strong>the</strong>nogenetic<br />

phase occurs during <strong>the</strong> spring and summer on herbaceous plants (<strong>the</strong> secondary hosts), and<br />

sexual reproduction occurs on a woody plant (<strong>the</strong> primary host). At <strong>the</strong> end <strong>of</strong> summer and<br />

<strong>the</strong> beginning <strong>of</strong> autumn, two morphs are involved in <strong>the</strong> return flight to <strong>the</strong> primary host:<br />

gynoparae and males. Gynoparae are winged par<strong>the</strong>nogenetic females, giving birth on <strong>the</strong><br />

primary host to oviparous sexual females. The timing <strong>of</strong> <strong>the</strong> switch to <strong>the</strong> sexual phase is<br />

important for a clone to maximise its fitness. The later cyclical par<strong>the</strong>nogenetic clones switch<br />

to sexual reproduction to increase <strong>the</strong>ir fitness, because <strong>the</strong> growth season is longer.<br />

Never<strong>the</strong>less, <strong>the</strong>se clones are constrained to switch to <strong>the</strong> sexual phase before <strong>the</strong> leaf fall <strong>of</strong><br />

<strong>the</strong> primary host on which eggs are laid. All individuals involved in sexual reproduction have<br />

to be produced synchronously between clones to ensure <strong>the</strong> mating rendezvous. Toge<strong>the</strong>r,<br />

short day-lengths and low temperatures induce <strong>the</strong> development <strong>of</strong> both gynoparae and males<br />

in species like Rhopalosiphum padi (L.) (Dixon & Glen, 1971), Myzus persicae (Sulzer)<br />

(Blackman, 1975), Dysaphis plantaginea (Passerini) (Bonnemaison, 1970). Aphids may,<br />

however, show a variability in <strong>the</strong>ir responses to <strong>the</strong>se environmental cues.<br />

Rhopalosiphum padi, which is a major vector <strong>of</strong> BYDV on winter cereals (Dedryver &<br />

Gelé, 1982; Lea<strong>the</strong>r et al. 1989) alternates between Poaceae (secondary hosts) and <strong>the</strong> bird<br />

1


2<br />

cherry, Prunus padus (primary host). In <strong>the</strong> autumn, <strong>the</strong> date <strong>of</strong> gynoparae and males<br />

production may influence <strong>the</strong> date <strong>of</strong> departure from <strong>the</strong> secondary hosts and <strong>the</strong>refore <strong>the</strong><br />

risk <strong>of</strong> BYDV transmission. Regional variation in <strong>the</strong> timing <strong>of</strong> <strong>the</strong> first males <strong>of</strong> R. padi<br />

caught in suction traps were reported in Great Britain and were related to photoperiodic<br />

differences among sites. Males were recorded earlier in <strong>the</strong> north <strong>of</strong> <strong>the</strong> UK (Tatchell, 1988),<br />

and <strong>the</strong>se observations were confirmed experimentally. Clones from nor<strong>the</strong>rn UK switch to<br />

sexual reproduction earlier than clones from sou<strong>the</strong>rn regions (Austin et al., 1996). Results<br />

from <strong>the</strong>se authors suggested also that <strong>the</strong> effect <strong>of</strong> photo-period was modulated by<br />

temperature.<br />

Concerning R. padi, several questions remain open. When should a clone switch to sexual<br />

reproduction in <strong>the</strong> field? Is <strong>the</strong>re any clonal variability in <strong>the</strong> responses to conditions<br />

inducing <strong>the</strong> production <strong>of</strong> sexuals? Is <strong>the</strong>re any geographic adaptation <strong>of</strong> <strong>the</strong> responses to<br />

<strong>the</strong>se inducing conditions? In order to answer <strong>the</strong>se questions, several clones <strong>of</strong> R. padi,<br />

differing by <strong>the</strong>ir geographic origin, were placed in two laboratory environments, mimicking<br />

<strong>the</strong> change <strong>of</strong> photo-period and <strong>the</strong>rmo-period, occurring naturally from <strong>the</strong> end <strong>of</strong> summer in<br />

oceanic and continental conditions, respectively. Experimental results concerning <strong>the</strong> timing<br />

<strong>of</strong> sexual morph production <strong>of</strong> clones were compared with field observations and semi-natural<br />

experiment.<br />

Materials and methods<br />

Aphid clones<br />

Two groups <strong>of</strong> five cyclical par<strong>the</strong>nogenetic clones <strong>of</strong> R. padi were tested. These clones were<br />

collected on <strong>the</strong> primary host, and before <strong>the</strong> spring migration in two regions <strong>of</strong> France<br />

differing in winter climate. The first five clones were collected in a continental region<br />

(Colmar, eastern France), and <strong>the</strong> last five in an oceanic region (Rennes, western France).<br />

These two regions were 800 km apart.<br />

After collection and before <strong>the</strong> experiments, <strong>the</strong> 20 clones were maintained at 20°C and a<br />

light regime <strong>of</strong> L:D 16:8 on wheat seedlings (cv. Arminda) to ensure continuous<br />

par<strong>the</strong>nogenetic reproduction (Simon et al., 1991).<br />

Experiments<br />

In a first experiment and in order to mimic changes in photo-period and <strong>the</strong>rmo-period at <strong>the</strong><br />

end <strong>of</strong> summer and during autumn, light and temperature daily regime recorded for a mild<br />

season at Rennes (oceanic condition) and a cold season at Colmar (continental condition)<br />

were simulated in two programmable cabinets from August 15 th (before <strong>the</strong> beginning <strong>of</strong><br />

sexuals production) to November 30 th (after <strong>the</strong> end <strong>of</strong> sexuals production) (Fig. 1). The<br />

simulated photo-periodic decrease corresponded to <strong>the</strong> latitude <strong>of</strong> Rennes and Colmar (48°07'<br />

North). Diurnal and nocturnal temperatures were applied during photo-phase and scoto-phase<br />

respectively.<br />

Both experiments, simulating oceanic and continental conditions, were started with ten<br />

fourth instar alatiform larvae <strong>of</strong> each <strong>of</strong> <strong>the</strong> 10 clones. Aphids were placed in Perspex boxes<br />

and reared on wheat seedlings (cv. Arminda). Twice a week, all mature winged aphids were<br />

removed from Perspex boxes with a fine brush, and identified as winged virginoparae, males<br />

or gynoparae. Gynoparae and virginoparae were distinguished with <strong>the</strong> squash blot test<br />

(Lowles, 1995).<br />

In a second experiment, and in order to monitor <strong>the</strong> timing <strong>of</strong> sexual morphs under<br />

natural conditions, <strong>the</strong> same clones were placed outdoor from <strong>the</strong> summer solstice, each in a<br />

cage containing wheat and a P. padus sapling. First gynoparae and males, which typically fly


3<br />

to P. padus, were recorded for each clone. During that time, <strong>the</strong>se clones were also placed in<br />

cabinets under experimental conditions, mimicking changes in <strong>the</strong>rmo-period and photoperiod<br />

from <strong>the</strong> summer solstice. In cabinets, <strong>the</strong>rmo-period was intermediary between<br />

Rennes and Colmar.<br />

Field observations<br />

Field observations came from <strong>the</strong> suction traps network Agraphid, which has been operating<br />

in France since 1978 (Hullé, 1991). Because <strong>of</strong> practical reasons, gynoparae were not<br />

distinguished from winged virginoparae in trap catches. The comparison between<br />

experimental results and field observations was <strong>the</strong>refore limited to male catches. Data<br />

collected from 1978 to 1995 at Colmar and Rennes, corresponding respectively to <strong>the</strong> eastern<br />

and western regions, were used in this study.<br />

30°C<br />

25°C<br />

20°C<br />

Oceanic diurnal temperature<br />

Oceanic night temperature<br />

Continental diurnal temperature<br />

Continental night temperature<br />

Photophase<br />

16:OO<br />

14:OO<br />

12:OO<br />

10:OO<br />

15°C<br />

8:OO<br />

10°C<br />

6:OO<br />

4:OO<br />

5°C<br />

2:OO<br />

0°C<br />

15.8<br />

19.8<br />

23.8<br />

27.8<br />

31.8<br />

4.9<br />

8.9<br />

12.9<br />

16.9<br />

20.9<br />

24.9<br />

28.9<br />

2.10<br />

6.10<br />

10.10<br />

14.10<br />

18.10<br />

22.10<br />

26.10<br />

30.10<br />

3.11<br />

7.11<br />

11.11<br />

15.11<br />

19.11<br />

23.11<br />

27.11<br />

0:OO<br />

Fig. 1. Simulated temperature and photo-period conditions. Diurnal temperature was applied<br />

during photo-phase and night temperature during scoto-phase.<br />

Results<br />

The different sequences <strong>of</strong> sexual morph production<br />

Sequences <strong>of</strong> sexual morph production were similar whatever <strong>the</strong> geographic origin <strong>of</strong> clones.<br />

The transition between par<strong>the</strong>nogenetic and sexual phase was complete. The production <strong>of</strong><br />

par<strong>the</strong>nogenetic virginoparae stopped when <strong>the</strong> production <strong>of</strong> gynoparae started. Differences<br />

in <strong>the</strong> number <strong>of</strong> gynoparae and males were not significant (Tab. 1). The timing <strong>of</strong> sexual<br />

morphs production was similar for both group <strong>of</strong> clones. First gynoparae were produced by<br />

<strong>the</strong> end <strong>of</strong> September (September 21 st for eastern clones and September 25 th for western<br />

clones) and first males one month latter (October 26 th and October 24 th , respectively) (Tab. 1).


4<br />

Tab. 1. Effect <strong>of</strong> geographic origin on sexual morphs production in R. padi: Date <strong>of</strong> first<br />

appearance and number (means with <strong>the</strong> same letter are not significantly different).<br />

Geographic origin<br />

East West<br />

1 st gynoparae 21 Sept (a) 25 Sept (a)<br />

1 st male 26 Oct (a) 24 Oct (a)<br />

No. <strong>of</strong> gynoparae 740 (a) 1012 (a)<br />

No. <strong>of</strong> males 146 (a) 326 (a)<br />

In contrast sequences <strong>of</strong> sexual morphs production were different between <strong>the</strong>rmoperiodic<br />

regimes (Tab. 2). Gynoparae were produced earlier in continental than in oceanic<br />

conditions (September 17 th and September 28 th , respectively), while males appeared at <strong>the</strong><br />

same time in both conditions. Number <strong>of</strong> gynoparae was similar in both conditions but<br />

number <strong>of</strong> males was lower in continental than in oceanic conditions.<br />

Tab. 2. Effect <strong>of</strong> experimental condition on sexual morphs production in R. padi: date <strong>of</strong> first<br />

appearance and number (means with <strong>the</strong> same letter are not significantly different)<br />

Thermoperiodic regime<br />

Continental Oceanic<br />

1st gynoparae 17 Sept (a) 28 Sept (b)<br />

1st male 27 Oct (a) 21 Oct (a)<br />

No. <strong>of</strong> gynoparae 879 (a) 974 (a)<br />

No. <strong>of</strong> males 104 (a) 388 (b)<br />

Tab. 3. Effect <strong>of</strong> experimental condition on sexual morphs production in R. padi:<br />

temperature and photo-phase corresponding to <strong>the</strong> date <strong>of</strong> first appearance<br />

Experimental condition<br />

Continental Oceanic<br />

Day °C 1 st gynoparae 20 22<br />

1 st male 11 17<br />

Night °C 1 st gynoparae 9 12<br />

1 st male 3 9<br />

Photo-phase 1 st gynoparae 12h25 11h47<br />

1 st male 10h09 10h29<br />

Mean conditions <strong>of</strong> temperature and photo-period for sexual morph production<br />

First gynoparae appeared when <strong>the</strong> day and <strong>the</strong> night temperature decreased below 20 and<br />

9°C, respectively in continental conditions, and below 22 and 12°C in oceanic conditions<br />

(Tab. 3). First males were recorded at a lower temperature in continental than in oceanic<br />

conditions (11 during <strong>the</strong> day and 3°C during <strong>the</strong> night, versus 17 and 9°C). The photo-phase<br />

corresponding to <strong>the</strong> first record <strong>of</strong> gynoparae was longer under continental than oceanic<br />

conditions (12h25 versus 11h47).


5<br />

Comparison with field data<br />

The first appearance <strong>of</strong> males ranged from August 12 th to September 24 th in <strong>the</strong> eastern<br />

continental region (Colmar suction trap), and from August 17 th to October 14 th in <strong>the</strong> western<br />

oceanic region (Rennes suction trap) (Tab. 4). In both sites, <strong>the</strong>se dates were earlier than in<br />

experimental results: The mean date <strong>of</strong> first male was seven weeks earlier in <strong>the</strong> Colmar<br />

suction trap than in continental conditions (Sept 4 th versus Oct 27 th ), and four weeks earlier in<br />

<strong>the</strong> Rennes suction trap than in oceanic conditions (Sept 23 rd versus Oct 22 nd ).<br />

Tab. 4. Date <strong>of</strong> <strong>the</strong> first males caught in suction trap from 1978 to 1995 compared to<br />

experimental results<br />

Earliest<br />

first catch<br />

Latest<br />

first catch<br />

Mean<br />

first catch<br />

Continental Suction trap 12 Aug 24 Sept 4 Sept<br />

(Colmar)<br />

Experiment 16 Oct 21 Nov 27 Oct<br />

Oceanic Suction trap 17 Aug 14 Oct 23 Sept<br />

(Rennes)<br />

Experiment 17 Oct 11 Nov 21 Oct<br />

Timing <strong>of</strong> sexual morph production in semi-natural conditions<br />

Under semi-natural conditions, dates <strong>of</strong> sexual morphs production were similar to field<br />

observations: First gynoparae were recorded on P. padus between August 18 th and September<br />

15 th , and males between September 22 nd and September 28 th . Conversely, under experimental<br />

conditions starting from summer solstice <strong>the</strong> dates <strong>of</strong> first appearance <strong>of</strong> sexual morphs were<br />

still late and similar to those <strong>of</strong> previous experiment: First gynoparae were produced between<br />

<strong>the</strong> end <strong>of</strong> September and mid October, depending on <strong>the</strong> clone and <strong>the</strong> first males two weeks<br />

latter (Tab. 5).<br />

Tab. 5. Date <strong>of</strong> first appearance <strong>of</strong> sexual morphs under experimental and semi-natural<br />

conditions starting from summer solstice<br />

Earliest<br />

first appearance<br />

Latest<br />

first appearance<br />

Gynoparae Outdoor 18 Aug 15 Sept<br />

Experiment 22 Sept 14 Oct<br />

Male Outdoor 22 Sept 28 Sept<br />

Experiment 9 Oct 1 Nov<br />

Discussion<br />

The same order <strong>of</strong> progeny sequence (par<strong>the</strong>nogenetic females?gynoparae?males) was<br />

achieved by all cyclical par<strong>the</strong>nogenetic clones. This order was similar to a pattern already<br />

described for cyclical par<strong>the</strong>nogenetic clones <strong>of</strong> R. padi but obtained in constant short days<br />

and low temperature (Dixon & Glen, 1971; Simon et al., 1991). No geographic adaptation in<br />

<strong>the</strong> timing <strong>of</strong> sexual morph production was found among cyclical par<strong>the</strong>nogenetic clones.<br />

However, <strong>the</strong>se clones were all collected at <strong>the</strong> same latitude contrary to <strong>the</strong> work <strong>of</strong> Austin et


6<br />

al. (1996) showing an effect <strong>of</strong> latitude on <strong>the</strong> progeny sequences <strong>of</strong> R. padi clones from<br />

different sites in Great Britain, and <strong>the</strong> experiments <strong>of</strong> Lushai et al. (1996), which showed an<br />

effect <strong>of</strong> latitude on photo-periodic responses for sexual morph production.<br />

In our experiments, <strong>the</strong> simulated photo-phase decrease was <strong>the</strong> same in both oceanic and<br />

continental conditions. There was only a difference in temperature which was lower and<br />

decreased more rapidly in continental than in oceanic conditions. Gynoparae were produced<br />

earlier in continental conditions at a mean day temperature <strong>of</strong> 14.5°C, and a photo-phase <strong>of</strong><br />

12h25 and later in oceanic conditions at a mean day temperature <strong>of</strong> 16.9°C, and a photo-phase<br />

<strong>of</strong> 11h47. These results suggest, that sexuals production depends on a combination <strong>of</strong> both<br />

factors which varies with geographic location. In this way and as long as temperature is high<br />

enough, aphids should sustain par<strong>the</strong>nogenetic reproduction. This strategy represents a<br />

selective advantage because it increases <strong>the</strong> fitness <strong>of</strong> clones. The reasoning was here applied<br />

to production date <strong>of</strong> gynoparae, as <strong>the</strong>y should precede males to maximise <strong>the</strong> success <strong>of</strong><br />

mating (Ward & Wellings, 1994). It could be applied just as well to <strong>the</strong> males and probably to<br />

previous generations because parents are sensitive to photo-period during <strong>the</strong> beginning <strong>of</strong><br />

<strong>the</strong>ir nymphal life (Dixon & Dewar, 1974).<br />

The number <strong>of</strong> males produced in continental conditions seemed very low. Since males<br />

are produced after gynoparae, <strong>the</strong> unexpected earlier mortality <strong>of</strong> parents may explain a<br />

deficit in males. This might have been <strong>the</strong> case in our experiment, because <strong>the</strong> daily<br />

temperatures were 7°C lower in continental than in oceanic conditions, suggesting that<br />

complete sequences could not be achieved in a colder climate.<br />

The time lag <strong>of</strong> nearly 4 weeks observed between <strong>the</strong> first sexual morphs in experimental<br />

conditions and in field or semi-natural observations underline <strong>the</strong> need for a better<br />

understanding <strong>of</strong> <strong>the</strong> influence <strong>of</strong> <strong>the</strong> whole array <strong>of</strong> environmental factors inducing <strong>the</strong><br />

transition from par<strong>the</strong>nogenetic to sexual reproduction, especially with <strong>the</strong> aim <strong>of</strong> predicting<br />

<strong>the</strong> timing <strong>of</strong> autumnal migration <strong>of</strong> virus vectors.<br />

References<br />

Austin, A.B., Tatchell, G.M., Harrington, R. & Bale, J.S., 1996: Adaptative significance <strong>of</strong><br />

changes in morph production during <strong>the</strong> transition from par<strong>the</strong>nogenetic to sexual<br />

reproduction in <strong>the</strong> aphid Rhopalosiphum padi (Homoptera: Aphididae). Bulletin <strong>of</strong><br />

Entomological Research 86: 93-99.<br />

Blackman, R.L., 1975: Photoperiodic determination <strong>of</strong> <strong>the</strong> male and female sexual morphs <strong>of</strong><br />

Myzus persicae. Journal <strong>of</strong> Insect Physiology 21: 435-453.<br />

Bonnemaison, L., 1970: Action de la photopériode sur la production des gynopares ailées de<br />

Dysaphis plantaginea Pass. Ann. Zool. Ecol. Anim. 2: 523-554.<br />

Dedryver, C.A. & Gelé, A., 1982: Biologie des pucerons des céréales dans l'Ouest de la<br />

France IV. - Etude de l'hivernation de populations anholocycliques Rhopalosiphum padi<br />

L., Metopolophium dirhodum Wlk. et Sitobion avenae F. sur repousses de céréales, dans<br />

trois stations de Bretagne et du Bassin parisien. Acta Oecologica Oecologia Applicata 3:<br />

321-342.<br />

Dixon, A.F. & Dewar, A.M., 1974: The time <strong>of</strong> determination <strong>of</strong> gynoparae and males in <strong>the</strong><br />

bird cherry-oat aphid, Rhopalosiphum padi. Annals <strong>of</strong> Applied Biology 78: 1-6.<br />

Dixon, A.F. & Glen, D.M., 1971: Morph determination in <strong>the</strong> bird cherry-oat aphid,<br />

Rhopalosiphum padi L. Annals <strong>of</strong> Applied Biology 68: 11-21.<br />

Hullé, M., 1991: Agraphid, un réseau de surveillance des populations de pucerons : base de<br />

données associée et domaines d'application. Annales ANPP 2: 103-113.


Lea<strong>the</strong>r, S.R., Walters, K.F. & Dixon, A.F., 1989: Factors determining <strong>the</strong> pest status <strong>of</strong> <strong>the</strong><br />

bird cherry-oat aphid, Rhopalosiphum padi (L.) (Hemiptera: Aphididae), in Europe: a<br />

study and a review. Bulletin <strong>of</strong> Entomological Research 79: 345-360.<br />

Lowles, A., 1995: A quick method for distinguishing between <strong>the</strong> two autumn winged female<br />

morphs <strong>of</strong> <strong>the</strong> aphid Rhopalosiphum padi. Entomologia Experimentalis et Applicata 74:<br />

95-99.<br />

Lushai, G., Hardie, J. & Harrington, R., 1996: Diapause termination and egg hatch in <strong>the</strong> bird<br />

cherry aphid, Rhopalosiphum padi. Entomologia Experimentalis et Applicata 81: 113-<br />

115.<br />

Simon, J.C., Blackman, R.L. & Le Gallic, J.F., 1991: Local variability in <strong>the</strong> life cycle <strong>of</strong> <strong>the</strong><br />

bird cherry-oat aphid, Rhopalosiphum padi (Homoptera: Aphididae) in western France.<br />

Bulletin <strong>of</strong> Entomological Research 81: 315-322.<br />

Tatchell, G.M., 1988: Regional adaptations in <strong>the</strong> phenology <strong>of</strong> a migrant aphid<br />

Rhopalosiphum padi (Homoptera: Aphididae) (Abstract). Proceeding <strong>of</strong> <strong>the</strong> 18th<br />

International Congress <strong>of</strong> Entomology, 182.<br />

7


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 9 - 23<br />

Biotypic variation <strong>of</strong> Diuraphis noxia (Homoptera: Aphididae)<br />

between South Africa and Hungary<br />

Zuzsa Basky 1 and Jorrie Jordaan 2<br />

1<br />

Plant Protection Institute, Hungarian Academy <strong>of</strong> Sciences, P.O. Box 102, Budapest 1525,<br />

Hungary<br />

2<br />

Sensaco Cooperative, Ltd., P.O. Box 566, Bethlehem 9700, Republic <strong>of</strong> South Africa<br />

Summary<br />

Impact <strong>of</strong> Diuraphis noxia (Mordvilko) from South Africa and Hungary was measured on susceptible<br />

and resistant South African wheat cultivars and a susceptible Hungarian barley cultivar. The effect <strong>of</strong><br />

aphid infestation on fresh plant weight over time differed between Hungarian and South African<br />

D. noxia strains for <strong>the</strong> susceptible barley cultivar Isis, and <strong>the</strong> resistant wheat cultivar SST 333.<br />

Aphid infestation significantly affected leaf area between South Africa and Hungary for susceptible<br />

wheat cultivar Betta and resistant wheat cultivar SST 333. In Hungary D. noxia feeding resulted in leaf<br />

rolling and chlorotic spots and stripes on resistant wheat cultivar SST 333 grown in growth chamber.<br />

Fur<strong>the</strong>rmore, <strong>the</strong> dynamics <strong>of</strong> <strong>the</strong> number <strong>of</strong> aphids per plant differed between South Africa and<br />

Hungary for Betta and SST 333. Infestation by Hungarian D. noxia reduced plant fresh weight and<br />

leaf area <strong>of</strong> <strong>the</strong> resistant wheat cultivars SST 333 and PI 262660 as much as <strong>the</strong> susceptible wheat<br />

cultivar Betta regardless <strong>of</strong> <strong>the</strong> growing conditions (growth chamber or greenhouse). In addition, water<br />

imbalance occurred in resistant wheat cultivars SST 333 and PI 262660 in <strong>the</strong> greenhouse. These<br />

differences between Hungarian and South African D. noxia strains suggest genetic differences<br />

between <strong>the</strong>se populations. Our results support <strong>the</strong> idea that resistant plant germplasm has<br />

geographical limits because <strong>of</strong> geographical variation in pest species.<br />

Key words: Diuraphis noxia, plant weight, leaf area, biotypic variation, South Africa, Hungary<br />

Introduction<br />

The Russian wheat aphid, Diuraphis noxia (Mordvilko) (Homoptera: Aphidaidae), was first<br />

recorded as a pest <strong>of</strong> cereals by Mokrzhetsky (1901). It has not subsequently been a persistent<br />

pest in Eurasia, its area <strong>of</strong> origin, although short-lived outbreaks have been reported (e.g.,<br />

Grossheim, 1914; Tuatay & Remaudiére, 1964; Dyadechko & Ruban, 1975; Fernández et al.,<br />

1992). However, after its discovery in South Africa in 1978 and in <strong>the</strong> United States in 1986,<br />

it became a major pest <strong>of</strong> cereals in <strong>the</strong>se countries (Du Toit & Walters, 1984; Brooks et al.,<br />

1994). Diuraphis noxia was first detected in Hungary in 1989 by Basky and Eastop (1991),<br />

but has not become a pest.<br />

To reduce D. noxia damage, intensive resistance breeding programs were undertaken in<br />

South Africa and <strong>the</strong> United States (Du Toit, 1989; Webster et al., 1987; Webster, 1990;<br />

Miller et al., 1994). Biotypic variation can affect <strong>the</strong> success <strong>of</strong> such breeding programs (see<br />

Diehl & Bush, 1984 for review). Puterka et al. (1992) shoved variation in damage to resistant<br />

wheat entries among eight D. noxia collections from several regions throughout <strong>the</strong> world,<br />

suggesting that biotypic variation may exist in D. noxia. Here, we report experiments on<br />

differences in plant development as a result <strong>of</strong> attack by South African and Hungarian<br />

D. noxia strains. Because <strong>of</strong> <strong>the</strong> risk <strong>of</strong> introducing pest biotypes into South Africa or<br />

Hungary, we could not transfer D. noxia between <strong>the</strong> two countries. Therefore in South Africa<br />

and in Hungary we measured <strong>the</strong> effect <strong>of</strong> infestation by locally collected D. noxia specimens<br />

9


10<br />

on fresh plant weight and leaf area <strong>of</strong> South African wheat cultivars, susceptible and resistant<br />

to D. noxia, and a Hungarian barley cultivar susceptible to D. noxia. The experimental<br />

conditions were as close as possible in <strong>the</strong> two countries.<br />

Materials and methods<br />

Sources <strong>of</strong> aphids<br />

In South Africa, D. noxia viviparous apterae and nymphs were collected at <strong>the</strong> beginning <strong>of</strong><br />

November from wheat at Zadoks growth stage 65-69 (an<strong>the</strong>sis half way, an<strong>the</strong>sis complete)<br />

(Tottman & Broad, 1987) in <strong>the</strong> main South African wheat growing area near Bethlehem,<br />

Orange Free State. Before our experiments, <strong>the</strong> aphids were reared for 3 wks on seedlings <strong>of</strong><br />

wheat variety Betta at 20°C and photo-period ~14:10 (L:D) h at 5.000-15.000 light intensity<br />

at <strong>the</strong> Sensaco Cooperative Breeding Station.<br />

In Hungary, D. noxia fundatrices were collected in <strong>the</strong> middle <strong>of</strong> April from wheat at<br />

growth stage 30-35 (stem elongation) near Szolnok, which is at <strong>the</strong> centre <strong>of</strong> <strong>the</strong> main wheat<br />

growing area in Hungary. Before our experiments, <strong>the</strong> aphids were reared for about 3 months<br />

on wheat variety Bezoshtaja in an environmental growth chamber at 20°C and photo-period<br />

14:10 (L:D) h at 7.500-8.500 light intensity.<br />

Treatments<br />

To test differences between South African and Hungarian D. noxia strains, experiments were<br />

done in each country with <strong>the</strong> South African winter wheat cultivars Betta, which is<br />

susceptible to D. noxia, and SST 333, which is resistant to D. noxia, and <strong>the</strong> Hungarian spring<br />

barley variety Isis, which is susceptible to D. noxia.<br />

Because D. noxia feeding resulted <strong>the</strong> occurrence <strong>of</strong> classic leaf rolling and streaks on <strong>the</strong><br />

resistant South African wheat cultivar SST 333 in environmental growth chamber in Hungary,<br />

an additional experiment was done in Hungary to compare effects <strong>of</strong> Hungarian D. noxia on<br />

Betta, SST 333, and on ano<strong>the</strong>r D. noxia resistant wheat line PI 262660. PI 262660, in which<br />

resistance originates from <strong>the</strong> single dominant gene Dn2 (Du Toit, 1989), was <strong>the</strong> source <strong>of</strong><br />

resistance on SST 333.<br />

In each experiment, sixteen seeds were sown in each <strong>of</strong> eight pots (17 cm diameter in<br />

Bethlehem and 15 cm in Hungary) for each variety. After emergence, seedlings were thinned<br />

to 12 seedlings per pot. Six days after emergence, i.e. at growth stage 11, <strong>the</strong> plants in half <strong>of</strong><br />

<strong>the</strong> pots were infested with one D. noxia apterae (7 day-old) each. In South Africa, insects and<br />

plants were kept in a controlled-temperature greenhouse at 20-14°C (day-night) with a photoperiod<br />

<strong>of</strong> ~ 14:10 (L:D) h. In South Africa eight pots were placed into one box, <strong>the</strong> front and<br />

sides <strong>of</strong> <strong>the</strong> boxes was glass, <strong>the</strong> top and <strong>the</strong> back <strong>of</strong> <strong>the</strong> boxes was fine mesh material. The<br />

infested and non-infested plants were placed into separate boxes, but within <strong>the</strong> boxes pots <strong>of</strong><br />

different cultivars were randomly designed. The plants were irrigated by an automated system<br />

with 40 ml water per pot three times per day. In Hungary in <strong>the</strong> first experiment, insects and<br />

plants were kept in a growth chamber at 20-14°C (day-night) with a photo-period <strong>of</strong> 14:10<br />

(L:D) h. The light intensity in <strong>the</strong> growth chamber varied between 7.500-8.500 lux. The<br />

plants were irrigated manually with 90-100 ml water per pot once per day. To avoid <strong>the</strong><br />

possible confounding effect <strong>of</strong> greenhouse vs. growth chamber, <strong>the</strong> trial was repeated in<br />

Hungary in <strong>the</strong> greenhouse and in an environmental growth chamber. The growing conditions<br />

in <strong>the</strong> Hungarian greenhouse were <strong>the</strong> following: photoperiod ~14:10 (L:D) h ~20-14 o C. The<br />

temperature regime actually varied between 13-24°C because <strong>the</strong> greenhouse was not<br />

temperature-controlled, we tried to maintain <strong>the</strong> desired temperature with manual ventilation.<br />

The light intensity varied between 5.000 and 15.000 lux. The plants were watered twice a day<br />

with 60-70 ml water each time. The higher water requirement <strong>of</strong> <strong>the</strong> plants was due to <strong>the</strong>


11<br />

higher temperature in <strong>the</strong> greenhouse. In Hungary, both <strong>the</strong> infested and non-infested pots<br />

were covered with 20 cm high transparent cages. The ventilation holes and <strong>the</strong> top <strong>of</strong> <strong>the</strong><br />

cages were covered with fine mesh organsa material. In in <strong>the</strong> environmental growth chamber<br />

and in <strong>the</strong> greenhouse plants were randomly arranged.<br />

Measurements<br />

In each experiment, four plants were destructively sampled from each plot at 7, 10 and 14 d<br />

after infestation. Thus, for each variety, infestation level, and sample date, four plants from<br />

each <strong>of</strong> <strong>the</strong> four pots were cut at soil surface. For infested plants, <strong>the</strong> numbers <strong>of</strong> D. noxia per<br />

plant were counted and <strong>the</strong>reafter <strong>the</strong> aphids were removed. Cut plants were weighted, and<br />

placed flat between two transparent sheets for photocopying. The areas <strong>of</strong> <strong>the</strong> photocopied<br />

plants were measured with a computer, using <strong>the</strong> Vidas Processing System (Kontron Image<br />

Analysis Division, Neufahrn, Germany). Although <strong>the</strong> visual scale <strong>of</strong> leaf rolling proposed by<br />

Webster et al. (1991) is widely used, leaf area was measured instead in order to avoid<br />

subjectivity (Webster et al., 1987), and to distinguish differences more precisely, and to allow<br />

detection <strong>of</strong> stunting.<br />

Data analyses<br />

For analysis, data from <strong>the</strong> experiment in South Africa and <strong>the</strong> first experiment in Hungary<br />

was combined. Repeated measures analysis <strong>of</strong> variance (ANOVA) was used to test <strong>the</strong> effects<br />

<strong>of</strong> country, infestation with D. noxia, sample date, and <strong>the</strong>ir interactions on plant weight and<br />

leaf area <strong>of</strong> each cultivar (Statistica Stat S<strong>of</strong>t). The same type <strong>of</strong> analysis was used to test <strong>the</strong><br />

effects <strong>of</strong> country, sample date, and <strong>the</strong>ir interactions on number <strong>of</strong> aphids per plant for each<br />

cultivar. Because <strong>the</strong> objective was to compare damage between countries and not between<br />

plant cultivars or species, separate analyses were done for each cultivar. Pots were <strong>the</strong><br />

smallest experimental units; <strong>the</strong> four plants sampled from each pot on each date were subsamples.<br />

Therefore mean plant weight, leaf area, and number <strong>of</strong> aphids for <strong>the</strong> four plants<br />

sampled from a pot on a date were <strong>the</strong> observations for analysis. Comparison <strong>of</strong> means <strong>of</strong><br />

plant weight and leaf area between infested and non-infested plants for each country and<br />

within each <strong>of</strong> <strong>the</strong>se categories between countries for <strong>the</strong> third sample date (14 days after<br />

infestation) was done by using t-test.<br />

For analysis <strong>of</strong> <strong>the</strong> second experiment in Hungary, data from <strong>the</strong> experiment in<br />

environmental growth chamber and greenhouse was combined. Repeated measures ANOVA<br />

was used to test <strong>the</strong> effects <strong>of</strong> growing conditions (greenhouse, growth chamber), infestation<br />

with D. noxia, sample date, and <strong>the</strong>ir interactions on plant weight and leaf area <strong>of</strong> each<br />

cultivar. The same type <strong>of</strong> analysis was used to test <strong>the</strong> effects <strong>of</strong> growing conditions, sample<br />

date, and <strong>the</strong>ir interactions on number <strong>of</strong> aphids per plant for each cultivar.<br />

Means <strong>of</strong> plant weight and leaf area between infested and non-infested plants for growing<br />

conditions and within each <strong>of</strong> <strong>the</strong>se categories between environmental growth chamber and<br />

greenhouse for <strong>the</strong> third sample date (14 days after infestation) were compared by using t-test<br />

(Statistica Stat S<strong>of</strong>t).<br />

Results<br />

Hungary versus South Africa<br />

Plant damage. The effect <strong>of</strong> aphid infestation on fresh plant weight over time differed<br />

between Hungary and South Africa for resistant wheat cultivar SST 333, and susceptible<br />

barley cultivar Isis, while leaf area differed between South Africa and Hungary for<br />

susceptible wheat cultivar Betta, and for resistant wheat cultivar SST 333 (Tab. 1).


12<br />

Tab. 1. Analysis <strong>of</strong> variance for effects <strong>of</strong> country (South Africa, Hungary) infestation (with<br />

or without D. noxia) time (7, 10,14 days after infestation), and <strong>the</strong>ir interactions on<br />

plant weight and leaf area <strong>of</strong> two wheat cultivars (Betta, SST 333) and one barley<br />

cultivar (Isis).<br />

Plant weight<br />

Leaf area<br />

Variety Factor df F P df F P<br />

Betta Country 1, 48 2.02 0.249 1, 48 13.84 0.033*<br />

Infestation 1, 48 22.83 0.017* 1, 48 79.78 0.002*<br />

Date 2, 48 22.89 0.001* 2, 48 14.95 0.004*<br />

Country * Infestation 1, 48 38.04 0.008* 1, 48 4.23 0.131<br />

Country * Date 2, 48 11.09 0.009* 2, 48 1.36 0.324<br />

Infestation * Date 2, 48 11.55 0.008* 2, 48 12.76 0.006*<br />

Country * Infestation * Date 2, 48 6.56 0.030* 2, 48 0.93 0.443<br />

SST 333 Country 1, 48 19.09 0.022* 1, 48 17.55 0.024*<br />

Infestation 1, 48 49.88 0.005* 1, 48 179.97 0.000*<br />

Date 2, 48 623.77 0.000* 2, 48 128.39 0.000*<br />

Country * Infestation 1, 48 13.16 0.036* 1, 48 230.35 0.000*<br />

Country * Date 2, 48 7.35 0.024* 2, 48 1.90 0.228<br />

Infestation * Date 2, 48 11.04 0.009* 2, 48 19.93 0.002*<br />

Country * Infestation * Date 2, 48 6.67 0.029* 2, 48 7.71 0.021*<br />

Isis Country 1, 48 43.58 0.007* 1, 48 2.27 0.228<br />

Infestation 1, 48 207.06 0.000* 1, 48 148.30 0.001*<br />

Date 2, 48 177.03 0.000* 2, 48 165.54 0.000*<br />

Country * Infestation 1, 48 51.77 0.005* 1, 48 43.46 0.007*<br />

Country * Date 2, 48 4.28 0.069 2, 48 2.51 0.161<br />

Infestation * Date 2, 48 70.18 0.000* 2, 48 47.25 0.000*<br />

Country * Infestation * Date 2, 48 15.99 0.003* 2, 48 68.24 0.000*<br />

By <strong>the</strong> time <strong>of</strong> <strong>the</strong> third sample date when <strong>the</strong> duration <strong>of</strong> infestation was long enough<br />

and damage caused by D. noxia became obvious, <strong>the</strong> mean fresh plant weight and mean plant<br />

area <strong>of</strong> <strong>the</strong> infested plants was significantly lower both in South Africa and in Hungary at<br />

each cultivar (Tab. 2, Fig. 1, 2).<br />

The fresh plant weight and leaf area <strong>of</strong> infested Isis barley cultivar did not differ<br />

significantly in South Africa and Hungary. There was no significant difference between <strong>the</strong><br />

fresh weight <strong>of</strong> infested Betta, however, <strong>the</strong> leaf area was significantly different at infested<br />

Betta between South Africa and Hungary, with a reduction <strong>of</strong> 71.4 and 49.1%, respectively.<br />

Both South African and Hungarian D. noxia reduced <strong>the</strong> fresh weight <strong>of</strong> SST 333, but <strong>the</strong><br />

percent reduction was greater in Hungary (45.4%) than in South Africa (24.1%) (Fig. 1). The<br />

leaf area <strong>of</strong> resistant SST 333 was reduced as a result <strong>of</strong> D. noxia feeding both in South Africa<br />

and in Hungary, with a leaf area reduction <strong>of</strong> 29.1 and 47.0%, respectively (Tab. 2, Fig. 2). In<br />

spite <strong>of</strong> <strong>the</strong> reduction in plant fresh weight and leaf area <strong>the</strong>re was no sign <strong>of</strong> D. noxia feeding<br />

on <strong>the</strong> leaves <strong>of</strong> SST 333 in South Africa, while characteristic leaf rolling and chlorotic spots<br />

and stripes developed on <strong>the</strong> D. noxia infested plants in Hungary. Although we tried to keep<br />

<strong>the</strong> plants at <strong>the</strong> two sites under <strong>the</strong> as much as possible <strong>the</strong> same conditions, <strong>the</strong> fresh plant<br />

weight and leaf area <strong>of</strong> <strong>the</strong> non-infested plant was significantly different in South Africa and<br />

in Hungary.


13<br />

Tab. 2. Comparisons <strong>of</strong> means <strong>of</strong> plant weight and leaf area between infested and non-infested plants in South Africa and Hungary and<br />

between South Africa and Hungary for infested and non-infested plants for two wheat varieties (Betta, SST 333) and one barley<br />

cultivar Isis 14 d after infestation.<br />

Treatment<br />

Weight Leaf area<br />

mean ± SEM<br />

SST 333 ISIS BETTA<br />

mean ± SEM<br />

SST 333<br />

BETTA<br />

ISIS<br />

South Africa<br />

infested vs 0.236 ± 0.022a 0.466 ± 0.044a* 0.373 ± 0.040a 704.76 ± 48.67a 2054.10 ± 103.0a 1339.68 ± 79.16a<br />

uninfested 0.347 ± 0.017b 0.631 ± 0.066b 0.871 ± 0.045b 2466.53 ± 165.2b 2962.59 ± 124.4b 3031.52 ± 84.36b<br />

t values 3.81 2.06 8.19 8.46 4.01 8.86<br />

Hungary<br />

infested vs 0.229 ± 0.016a 0.261 ± 0.025a 0.375 ± 0.024a 844.70 ± 87.98a 1168.65±112.77a 1347.27 ± 27.36a<br />

uninfested 0.427 ± 0.020b 0.478 ± 0.022b 0.486 ± 0.031b 1661.26 ± 72.75b 2213.29 ±102.97b 2093.56 ± 195.6b<br />

t values 7.61 6.43 2.78 6.47 6.84 4.72<br />

Infested<br />

South Africa vs 0.236 ± 0.022a 0.466 ± 0.044a 0.373 ± 0.040a 704.76 ± 48.67a 2054.10 ±103.04a 1339.68 ± 79.16a<br />

Hungary 0.229 ± 0.016a 0.261 ± 0.025b 0.375 ± 0.024a 844.70 ± 87.98b 1168.65 ±112.77b 1347.27 ±127.36a<br />

t values 0.223 6.06 0.02 8.46 4.04 0.05<br />

Uninfested<br />

South Africa vs 0.347 ± 0.017a 0.631 ± 0.066a 0.871 ± 0.045a 2466.53 ± 165.2b 2962.59±124.47a 3031.52 ± 84.36a<br />

Hungary 0.427 ± 0.020b 0.478 ± 0.022b 0.486 ± 0.031b 1661.26 ± 72.75b 2213.29 ±102.97b 2093.56 ± 195.6b<br />

t values 5.02 6.43 2.78 2.10 3.84 3.51<br />

Means ± SEM within pairs followed by different letters are significantly different (P


14<br />

14<br />

DAYS AFTER INFECTION<br />

Fig. 1. Mean fresh plant weight <strong>of</strong> three wheat cultivars (Betta, SST 333, PI 262660) and one barley cultivar (Isis) with and without D. noxia<br />

in South Africa (SA) in greenhouse and in Hungary (H) in environmental growth chamber (EGC) and in greenhouse (GH) sampled at<br />

three dates after aphid infestation.


15<br />

DAYS AFTER INFECTION<br />

Fig. 2. Mean leaf area mm 2 <strong>of</strong> three wheat cultivars (Betta, SST 333, PI 262660) and one barley cultivar (Isis) with and without D. noxia in<br />

South Africa in greenhouse and in Hungary in environmental growth chamber and in greenhouse sampled at three dates after aphid<br />

infestation.<br />

15


16<br />

Fig. 3. Mean number <strong>of</strong> D. noxia per plant for three wheat cultivars (Betta, SST 333, PI<br />

262660) and one barley cultivar (Isis) in South Africa in <strong>the</strong> greenhouse and in<br />

Hungary in an environmental growth chamber and in <strong>the</strong> greenhouse, sampled at<br />

three dates after aphid infestation.<br />

Aphid numbers. The dynamics <strong>of</strong> <strong>the</strong> number <strong>of</strong> aphids per plant differed between<br />

Hungary and South Africa for <strong>the</strong> susceptible wheat cultivar Betta and <strong>the</strong> resistant wheat<br />

cultivar SST 333 (Tab. 3, Fig. 3a). In both countries aphids reproduced on <strong>the</strong> resistant SST<br />

333. The dynamics <strong>of</strong> <strong>the</strong> aphid populations were similar on <strong>the</strong> two wheat cultivars aphid<br />

numbers and did not differ between South Africa and Hungary at days 7 and 10. However, by<br />

day 14, aphid densities were 460 and 135% higher in South Africa than in Hungary on Betta<br />

and SST 333, respectively. The difference between countries supports <strong>the</strong> observation that<br />

South African D. noxia has a higher net reproductive rate than Hungarian D. noxia (Basky &


17<br />

Jordaan, 1997). Surprisingly, <strong>the</strong> number <strong>of</strong> aphids per plant on <strong>the</strong> susceptible barley variety<br />

Isis did not vary significantly with country, sample date, or <strong>the</strong>ir interaction (Tab. 3).<br />

Tab. 3. Analysis <strong>of</strong> variance for effects <strong>of</strong> country (South Africa, Hungary), sample date (7,<br />

10, 14 days after infestation), an <strong>the</strong>ir interaction on number <strong>of</strong> aphids per plant<br />

Aphids per plant<br />

Cultivar Factor df F P<br />

Betta Country 1,24 31.2 0.000*<br />

Sample date 2,24 32.7 0.000*<br />

Country x Sample date 2,24 21.8 0.000*<br />

SST 333 Country 1,24 11.4 0.010*<br />

Sample date 2,24 25.9 0.000*<br />

Country x Sample date 2,24 11.1 0.001*<br />

Isis Country 1,24 2.3 0.17<br />

Sample date 2,24 3.5 0.06<br />

Country x Sample date 2,24 2.6 0.10<br />

Environmental growth chamber versus greenhouse<br />

Plant damage. The significant difference <strong>of</strong> non-infested plant fresh weight and leaf area<br />

between South Africa and Hungary rose <strong>the</strong> question whe<strong>the</strong>r <strong>the</strong> difference between South<br />

African and Hungarian D. noxia damage was only due to different growing conditions<br />

(greenhouse in South Africa and environmental growth chamber in Hungary). The effect <strong>of</strong><br />

aphid infestation on plant fresh weight and leaf area over time differed between growth<br />

chamber and greenhouse for resistant wheat cultivar SST 333 and for susceptible barley<br />

cultivar Isis. The growth chamber vs. greenhouse significantly affected <strong>the</strong> leaf area <strong>of</strong><br />

susceptible wheat cultivar Betta and <strong>the</strong> weight <strong>of</strong> resistant wheat cultivar PI 262660 (Tab. 4).<br />

By <strong>the</strong> time <strong>of</strong> <strong>the</strong> third assessment <strong>the</strong> mean fresh plant weight and leaf area <strong>of</strong> infested<br />

plants was significantly lower for each cultivar, both in <strong>the</strong> growth chamber and <strong>the</strong> in<br />

greenhouse (Tab. 5, Fig. 1 and 2). However, <strong>the</strong> percent reduction was higher in <strong>the</strong><br />

greenhouse than in <strong>the</strong> growth chamber for each wheat cultivar. The plant fresh weight<br />

reduction <strong>of</strong> Betta, Isis, SST 333 and PI 262660 in <strong>the</strong> growth chamber were 34,6, 22.5, 35<br />

and 40%, while in <strong>the</strong> greenhouse <strong>the</strong>y were 62.2, 66.7, 65.4 and 67.9% respectively. The<br />

percent reduction <strong>of</strong> infested leaf area <strong>of</strong> Betta, Isis, SST 333 and PI262660 were 46.2, 35.6,<br />

56 and 46.6% in <strong>the</strong> growth chamber and 77, 70.6, 74.1 and 69.2% in <strong>the</strong> greenhouse,<br />

respectively (Tab. 5, Fig. 1 and 2). The effect <strong>of</strong> aphid infestation on fresh plant weight by <strong>the</strong><br />

time <strong>of</strong> <strong>the</strong> third sample date differed between <strong>the</strong> growth chamber and <strong>the</strong> greenhouse only<br />

for resistant wheat cultivar SST 333. While <strong>the</strong> leaf area <strong>of</strong> infested susceptible Betta and<br />

resistant SST333 were significantly different between <strong>the</strong> growth chamber and <strong>the</strong><br />

greenhouse, <strong>the</strong> differences were not significant for non-infested plants. At <strong>the</strong> same time<br />

<strong>the</strong>re were significant differences between growth chamber and greenhouse both for plant<br />

fresh weight and for leaf area <strong>of</strong> non-infested PI 262660, while <strong>the</strong>se characters did not vary<br />

significantly in infested plants.


18<br />

Tab. 4. Analysis <strong>of</strong> variance for effects <strong>of</strong> growing conditions (environmental growth<br />

chamber, greenhouse), infestation (with and without D. noxia), time (7, 10, 14 days<br />

after infestation), and <strong>the</strong>ir interactions on plant weight and leaf area <strong>of</strong> three wheat<br />

cultivars (Betta, ST 333, PI 262660) and one barley cultivar (Isis).<br />

Plant weight<br />

Leaf area<br />

Variety Factor df F P df F P<br />

Betta Location 1, 48 6.28 0.087 1, 48 43.84 0.007*<br />

Infestation 1, 48 144.81 0.007* 1, 48 652.11 0.000*<br />

Date 2, 48 99.32 0.000* 2, 48 128.75 0.000*<br />

Location * Infestation 1, 48 0.30 0.619 1, 48 6.52 0.08<br />

Location * Date 2, 48 21.78 0.001* 2, 48 8.63 0.017*<br />

Infestation * Date 2, 48 155.41 0.000* 2, 48 58.41 0.000*<br />

Location*Infestation*Date 2, 48 4.05 0.076 2, 48 10.76 0.010*<br />

SST 333 Location 1, 48 17.39 0.025* 1, 48 46.48 0.006*<br />

Infestation 1, 48 148.33 0.001* 1, 48 199.81 0.000*<br />

Date 2, 48 540.52 0.000* 2, 48 143.32 0.000*<br />

Location * Infestation 1, 48 79.57 0.002* 1, 48 18.16 0.023*<br />

Location * Date 2, 48 12.69 0.006* 2, 48 8.58 0.017*<br />

Infestation * Date 2, 48 82.45 0.000* 2, 48 178.95 0.000*<br />

Location*Infestation*Date 2, 48 12.97 0.006* 2, 48 14.67 0.004*<br />

Isis Location 1, 48 463.54 0.000* 1, 48 38.57 0.000*<br />

Infestation 1, 48 231.30 0.000* 1, 48 146.32 0.000*<br />

Date 2, 48 298.79 0.000* 2, 48 126.5 0.000*<br />

Location * Infestation 1, 48 56.61 0.004* 1, 48 6.36 0.08<br />

Location * Date 2, 48 56.59 0.000* 2, 48 8.32 0.016*<br />

Infestation * Date 2, 48 24.96 0.001* 2, 48 62.79 0.000*<br />

Location*Infestation*Date 2, 48 186.20 0.000* 2, 48 2.72 0.006*<br />

PI262660 Location 1, 48 141.45 0.001* 1, 48 0.6 0.478<br />

Infestation 1, 48 259.33 0.000* 1, 48 132.2 0.001*<br />

Date 2, 48 213.11 0.000* 2, 48 64.26 0.000*<br />

Location * Infestation 1, 48 41.91 0.007* 1, 48 7.09 0.076<br />

Location * Date 2, 48 30.09 0.000* 2, 48 22.53 0.001*<br />

Infestation * Date 2, 48 424.32 0.000* 2, 48 3.24 0.000*<br />

Location*Infestation*Date 2, 48 16.03 0.003* 2, 48 13.76 0.005*<br />

The plant growth was higher in <strong>the</strong> greenhouse than in <strong>the</strong> growth chamber, coupled by<br />

higher aphid damage in all cultivars (Tab. 5). The aphid damage was so severe on resistant<br />

SST 333 and PI 262660 that Russian wheat aphid damage induced visible water imbalance in<br />

<strong>the</strong> greenhouse. Signs <strong>of</strong> water imbalance, i.e. visible loss <strong>of</strong> turgor, was detected ten days<br />

after D. noxia infestation for both resistant cultivars (SST 333 and PI 262660), but no signs <strong>of</strong><br />

water imbalance were observed on <strong>the</strong> susceptible wheat cultivar Betta and <strong>the</strong> susceptible<br />

barley cultivar Isis.<br />

Aphid numbers. The dynamics <strong>of</strong> aphid densities per plant did not differ between<br />

growth chamber and greenhouse for <strong>the</strong> susceptible wheat and barley cultivars Betta and Isis,<br />

but significantly differed between <strong>the</strong> resistant SST 333 and PI 262660 cultivars (Tab. 6).


19<br />

Aphid numbers were higher in growth chamber than in greenhouse on all cultivars, and<br />

particularly remarkable on PI 262660 where <strong>the</strong> number <strong>of</strong> aphids 14 days after infestation<br />

was 55.4% higher in <strong>the</strong> growth chamber than in <strong>the</strong> greenhouse. The lower aphid<br />

reproduction on <strong>the</strong> resistant cultivars may be related to aphid induced water imbalances.<br />

Reduction <strong>of</strong> aphid reproduction occurred on SST 333 between 10 and 14 days after<br />

infestation (Fig. 3b).<br />

Table 6. Analysis <strong>of</strong> variance for effects <strong>of</strong> location (environmental growth chamber,<br />

greenhouse), sample date (7, 10, 14 days after infestation), an <strong>the</strong>ir interaction on<br />

number <strong>of</strong> aphids per plant.<br />

Aphids per plant<br />

Cultivar Factor df F P<br />

Betta Location 1, 24 0.15 0.71<br />

Sample date 2, 24 22.59 0.001*<br />

Location x Sample date 2, 24 1.56 0.28<br />

SST 333 Location 1, 24 13.28 0.035*<br />

Sample date 2, 24 12.56 0.007*<br />

Location x Sample date 2, 24 1.09 0.39<br />

Isis Location 1, 24 0.005 0.944<br />

Sample date 2, 24 1.48 0.299<br />

Location x Sample date 2, 24 1.39 0.318*<br />

PI 262660 Location 1, 24 16.79 0.026*<br />

Sample date 2, 24 28.31 0.000*<br />

Location x Sample date 2, 24 21.02 0.001*<br />

Discussion<br />

The differences between Hungarian and South African D. noxia strains in <strong>the</strong>ir effects on<br />

plant fresh weight and leaf area suggest genetic differences between <strong>the</strong>se aphid populations.<br />

O<strong>the</strong>r observations during this experiment also indicate different biotypes: infested Betta and<br />

Isis had rolled leaves and chlorotic streaks in both South Africa and Hungary, whereas<br />

infested SST 333 had <strong>the</strong>se symptoms in Hungary in a growth chamber, but not in <strong>the</strong> South<br />

African greenhouse. Despite differences in visual symptoms between countries, aphid<br />

infestation reduced fresh plant weight and leaf area <strong>of</strong> SST 333 in both countries. This effect<br />

on weight without visual symptoms corroborates results by Bush et al. (1989) and Scott et al.<br />

(1990), who found that reduction in plant weight can occur in a line when a visual damage<br />

rating indicates a high level <strong>of</strong> resistance. At least for plant fresh weight and leaf area,<br />

resistant wheat SST 333 and PI262660 appeared to suffer as much reduction from infestation<br />

by Hungarian D. noxia as <strong>the</strong> susceptible cultivar Betta, beside <strong>the</strong> quantitative changes<br />

yellow spots and stripes and leaf rolling, characteristic to D. noxia damage appeared on <strong>the</strong><br />

resistant cultivars.


20<br />

20<br />

Tab. 5. Comparisons <strong>of</strong> means <strong>of</strong> plant weight and leaf area between infested and non-infested plants in <strong>the</strong> growth chamber and <strong>the</strong><br />

greenhouse, and between <strong>the</strong> growth chamber and <strong>the</strong> greenhouse in Hungary for infested and non-infested plants for three wheat<br />

varieties (Betta, SST 333, PI 262660) 14 d after infestation.<br />

Treatment<br />

BETTA<br />

Weight<br />

mean ± SEM<br />

SST 333 PI 262660 BETTA<br />

Leaf area<br />

mean ± SEM<br />

SST 333 PI 262660<br />

Growth chamber<br />

Infested vs 0.300 + 0.029a 0.432 + 0.016a 0.297 + 0.031a 1060.75 + 87.60a 1305.03 + 22.55a 1041.90 + 111.56a<br />

Uninfested 0.459 + 0.009b 0.664 + 0.021b 0.494 + 0.041b 1969.75 + 87.62b 2965.38 + 78.74b 1951.65 + 144.88b<br />

t-values 4.04 6.88 3.08 5.73 16.21 3.97<br />

Greenhouse<br />

Infested vs 0.253 + 0.010a 0.318 + 0.028a 0.254 + 0.028a 517.97 + 56.89a 822.92 + 61.56a 811.45 + 65.322a<br />

Uninfested 0.670 + 0.017b 0.918 + 0.022b 0.791 + 0.030b 2242.05 + 64.03b 3166.4 + 124.62b 2630.12 + 87.428b<br />

t-values 16.10 13.18 10.24 16.10 12.06 13.33<br />

Infested<br />

Growth chamb vs 0.300 + 0.029a 0.432 + 0.016a 0.297 + 0.031a 1060.75 + 87.60a 1305.03 + 22.55a 1041.90 + 111.56a<br />

Greenhouse 0.253 + 0.010a 0.318 + 0.028b 0.254 + 0.028a 517.97 + 56.89b 822.92 + 61.56b 811.45 + 65.32a<br />

t values 1.18 2.75 0.80 4.14 5.88 1.42<br />

Uninfested<br />

Growth chamb vs 0.459 + 0.009a 0.664 + 0.021a 0.494 + 0.040a 1969.75 + 87.62a 2965.38 + 78.74a 1951.65 + 144.88a<br />

Greenhouse 0.670 + 0.017b 0.918 + 0.022b 0.791 + 0.030b 2242.05 + 64.03a 3166.4 + 124.62a 2630.12 + 87.42 b<br />

t values 8.41 6.54 4.707 1.95 0.98 3.200<br />

Means ± SEM within pairs followed by different letters are significantly different (P


21<br />

It is likely that damage by D. noxia was <strong>of</strong>ten not significant 7-10 d after infestation in<br />

our experiments (data not shown) because one adult and its progeny were too few aphids to<br />

cause detectable damage during this period. Webster et al. (1987) suggested that an initial<br />

infestation <strong>of</strong> at least 10 D. noxia per seedling is required for evaluation <strong>of</strong> resistance.<br />

Westhuizen and Bota (1993) found that D. noxia infestation induced quantitative differences<br />

between <strong>the</strong> polypeptide pr<strong>of</strong>iles <strong>of</strong> resistant and susceptible wheat leaves. Aphid infestation<br />

induces accumulation <strong>of</strong> specific proteins in <strong>the</strong> intercellular fluid <strong>of</strong> resistant varieties only<br />

(Nagel et al., 1994), but production <strong>of</strong> <strong>the</strong>se proteins requires sufficient aphid numbers (A. J<br />

van der Westhuizen, personal communication).<br />

Higher plant weight and leaf area in South African greenhouse than in Hungarian<br />

environmental growth chamber for non-infested plants may have resulted from differences<br />

from light level or irrigation systems.<br />

When <strong>the</strong> trial was repeated in <strong>the</strong> growth chamber and <strong>the</strong> greenhouse on D. noxia<br />

resistant SST 333 and PI262660 cultivars, plant fresh weight and leaf area <strong>of</strong> D. noxia<br />

infested plants significantly decreased and typical D. noxia damage symptoms, such as<br />

chlorotic spots and stripes, leaf rolling, were formed on <strong>the</strong> resistant plants. Moreover, signs<br />

<strong>of</strong> water imbalance occurred on <strong>the</strong> resistant cultivars 10 and 14 days after infestation in <strong>the</strong><br />

greenhouse. Miller et al. (1994) reported symptoms <strong>of</strong> susceptibility to D. noxia in barley,<br />

indicating alterations in <strong>the</strong> water status <strong>of</strong> <strong>the</strong> leaf. Infested susceptible barley took up less<br />

water than non-infested plants. Burd and Burton (1992), characterising <strong>the</strong> D. noxia damage,<br />

pointed out that “The prevention <strong>of</strong> unfolding <strong>of</strong> new leaves and reduction on leaf size caused<br />

by Russian wheat aphid feeding apparently results from <strong>the</strong> reduction <strong>of</strong> leaf turgor below <strong>the</strong><br />

threshold for elongation and cell wall extensibility.” Burd et al. (1993) found significantly<br />

lower leaf turgor for infested susceptible triticale Beagle 82 and susceptible wheat TAM W<br />

101 and resistant wheat PI 372129 compared to <strong>the</strong> non-infested control. The leaf turgor <strong>of</strong><br />

resistant D. noxia infested triticale cultivars Okay R and PI 386148 did not differ from <strong>the</strong><br />

non-infested control. The water imbalance in most cases, except for resistant wheat PI 372129<br />

(Burd et al., 1993), occurred at susceptible plant entries. In our experiment visible water<br />

imbalance occurred on resistant cultivars only (SST 333 and PI 262660), even when ample<br />

moisture was provided to <strong>the</strong> roots. The occurrence <strong>of</strong> water imbalance, toge<strong>the</strong>r with<br />

characteristic leaf rolling and yellow spots and stripes on cultivars which are resistant to<br />

South African D. noxia suggests biotypic differences in D. noxia between South Africa and<br />

Hungary.<br />

Our data support <strong>the</strong> idea that resistant plant germplasm has geographical limits because<br />

<strong>of</strong> geographical variation in pest species (Puterka et al., 1992). This means that, in an<br />

aggressive breeding program, resistance should be identified against collections <strong>of</strong> pests from<br />

throughout <strong>the</strong> whole region <strong>of</strong> crop production. Fur<strong>the</strong>rmore, stacking genes for resistance in<br />

a variety should prove to be a more durable strategy on <strong>the</strong> long run.<br />

Acknowledgements<br />

We thank Léan van der Westhuizen (University <strong>of</strong> Orange Free State, Dep. <strong>of</strong> Entomology)<br />

for suggestions on experimental design, Tanya Saayman (Agricultural Research Council,<br />

Plant Protection Research Institute Pretoria) for measuring <strong>the</strong> leaf area, Willie Maree<br />

(Sensaco Cooperative, Ltd.) for providing research facilities, <strong>the</strong> Cereal Research Institute,<br />

Szeged, Hungary for supplying germplasm, Keith R. Hopper (Beneficial Insect Introduction<br />

Research Laboratory, ARS, USDA Newark), Arpad Szentesi and Ferenc Kadar ( Plant


22<br />

Protection Institute <strong>of</strong> Hungarian Academy <strong>of</strong> Sciences) for suggestions on <strong>the</strong> statistical<br />

analyses. This research was funded by Sensaco Cooperative, Ltd.<br />

References<br />

Basky, Z. & Eastop, V.F., 1991: Diuraphis noxia in Hungary. Newslett. Barley Yellow Dwarf<br />

4: 34.<br />

Basky, Z. & Jordaan, J., 1997: Comparison <strong>of</strong> <strong>the</strong> development and fecundity <strong>of</strong> Russian<br />

wheat aphid (Homoptera: Aphididae) in South Africa and in Hungary. J. Econ. Entomol.<br />

90: 623-627.<br />

Brooks, L., Hein, G., Johnson, G., Legg, D., Massey, B., Morrison, P., Weiss M. & Peairs, F.,<br />

1994: Economic impact <strong>of</strong> <strong>the</strong> Russian wheat aphid in <strong>the</strong> western United States. 1991 -<br />

1992. Great Plains Agric. Council Pub. No. 147: 250-268, In: <strong>Proceedings</strong> 6 th Russian<br />

Wheat Aphid Workshop, 23-25 January 1994 Fort Collins Colorado.<br />

Bush, L., Slosser, J. E., & Worall, W.D., 1989: Variations in damage to wheat caused by<br />

Russian wheat aphid (Homoptera: Aphididae) in Texas. J. Econ. Entomol. 82: 466-471.<br />

Burd, J.D. & Burton, R.L., 1992: Characterization <strong>of</strong> plant damage caused by Russian wheat<br />

aphid (Homoptera: Aphididae. J. Econ. Entomol. 85: 2017-2022.<br />

Burd, J.D., Burton, R.L. & Webster, J.A., 1993: Evaluation <strong>of</strong> Russian wheat aphid (Homoptera:<br />

Aphididae) damage on resistant and susceptible hosts with comparisons <strong>of</strong> damage<br />

ratings to quantitatve plant measurements. J. Econ. Entomol. 86: 974-980.<br />

Diehl, S.R. & Bush, G.L., 1984: An evolutionary and applied perspective <strong>of</strong> insect biotypes.<br />

Ann. Rev. Entomol. 29: 1251-1253.<br />

Du Toit, F. 1989: Inheritance <strong>of</strong> resistance in two Triticum aestivum lines to Russian wheat<br />

aphid (Homoptera: Aphididae). J. Econ. Entomol. 82: 1251-1253.<br />

Du Toit, F. & Walters, M.C., 1984: Damage assessment and economic threshold values for<br />

<strong>the</strong> chemical control <strong>of</strong> <strong>the</strong> Russian wheat aphid, Diuraphis noxia (Mordvilko) on winter<br />

wheat. Technical Communication <strong>of</strong> <strong>the</strong> Department <strong>of</strong> Agriculture, Republic <strong>of</strong> South<br />

Africa 191: 58-62.<br />

Dyadechko, N.P. & Ruban, M.B., 1975: The harmfulness <strong>of</strong> cereal aphid. Zashch. Rast.<br />

(Mosc.) 12: 17-18.<br />

Fernandez, V.N., Perez, E.N., Santero, E.D.& Nafria, J.M.N., 1992: Situacion en el norte de<br />

espana del pulgon ruso del trigo, Diuraphis noxia (Mordvilko) (Homoptera, Aphididae).<br />

Georgica 1: 9-24.<br />

Grossheim, N.A. 1914: The barley aphid, Brachycolus noxius Mordwilko. Mem. Nat. Hist.<br />

Mus. Zemstwo Province Tavria 3: 35-78. (In Russian; English translation by Poprawski,<br />

T.J., Wraight, S.P. & Peresypkina, S. In: Morrison, W.P. (ed.), Proc. 5 th Russian Wheat<br />

Aphid Conference, 26-28 January 1992. Great Plains Agricultural Council Publication<br />

142: 34-55.).<br />

Miller, H., Porter, D.R., Burd, J.D., Mornhinweg, D.W. & Burton, R.L., 1994: Physiological<br />

effects <strong>of</strong> Russian wheat aphid (Homoptera: Aphididae) on resistant and susceptible<br />

barley. J. Econ. Entomol. 87: 493-499.<br />

Mokrzhetsky, K.A., 1901: (Animal and Plant Pests <strong>of</strong> Crimea in 1900.) Simferopol; cited in<br />

Kovalev, O.V., Poprawski, T.J., Stekolshchikov, A.V., Vereshchagina, A.B. &<br />

Gandrabur, S.A. 1991: Diuraphis Aizenberg (Hom., Aphididae) key to apterous females,<br />

and review <strong>of</strong> Russian language literature on <strong>the</strong> natural history <strong>of</strong> Diuraphis noxia<br />

(Kurdjumov, 1913 ) J. Appl. Entomol. 112: 425-436.


Nagel, M.A., Pretorius, C.Z., Botha, A.M. & van der Westhuizen, A.J., 1994: Russian wheat<br />

aphid resistance markers in wheat. Abstracts <strong>of</strong> papers and posters presented at <strong>the</strong><br />

Twentieth Annual Congress <strong>of</strong> <strong>the</strong> South African Association <strong>of</strong> Botanists. Witwatersrand<br />

10-14 January 1994: 69.<br />

Puterka, G.J., Burd, J.D. & Burton, R.L., 1992: Biotypic variation in a worldwide collection<br />

<strong>of</strong> Russian wheat aphid (Homoptera: Aphididae) J. Econ. Entomol. 85: 1497-1506.<br />

Scott, R.A., Worrall, W.D. & Frank, W.A., 1990: Comparison <strong>of</strong> three techniques for<br />

measuring antibiosis to Russian wheat aphid. Southwestern Entomologist. 15: 439-446.<br />

Statistica. 1994: Statistica for <strong>the</strong> Windows TM StatS<strong>of</strong>t, Inc., Tulsa, OK.<br />

Tottman, D.R. & Broad, H., 1987: Decimal code for <strong>the</strong> Growth Stage <strong>of</strong> cereals. Ann. Appl.<br />

Biol. 110: 683-687.<br />

Tuatay, N., & Remaudiére, G., 1964: Premiere contribution an catalogue des Aphididae<br />

(Hom.) de la Turquie. Rev. Pathol. Veg. Entomol. Agronom. Rf. 43: 237-278.<br />

Vidas Processing System, 1993: (Kontron Image Analysis Division) Neufahrn, Germany<br />

Webster, J.A. 1990: Resistance in Triticale to <strong>the</strong> Russian wheat aphid (Homoptera:<br />

Aphididae) J. Econ. Entomol. 83: 1091-1095.<br />

Webster, J.A., Starks, K.J. & Burton, R.L., 1987: Plant resistance studies with Diuraphis<br />

noxia (Homoptera Aphididae), a new United States wheat pest. J. Econ. Entomol 80:<br />

944-949.<br />

Webster, J.A., Baker, C.A. & Porter, D.R., 1991: Detection and mechanisms <strong>of</strong> Russian wheat<br />

aphid (Homoptera: Aphididae) resistance in barley. J. Econ. Entomol. 84: 669-673.<br />

van der Westhuizen, A.J. & Botha, F.C., 1993: Effect <strong>of</strong> <strong>the</strong> Russian Wheat Aphid on <strong>the</strong><br />

composition and syn<strong>the</strong>sis <strong>of</strong> water soluble proteins in resistant and susceptible wheat. J.<br />

Agronomy & Crop Science 170: 322-326.<br />

23


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 25 - 31<br />

Impact <strong>of</strong> barley yellow dwarf virus infection on physiological conditions<br />

<strong>of</strong> wheat and <strong>the</strong> consequences for cereal aphids attack<br />

M. Fiebig and H.-M. Poehling<br />

Institut für Pflanzenkrankheiten und Pflanzenschutz, Universität Hannover, Herrenhäuser<br />

Str. 2, D-30419 Hannover, Germany<br />

Background and introduction<br />

The BYDV (Barley Yellow Dwarf Virus) pathosys<strong>the</strong>m has three biotic elements: <strong>the</strong><br />

luteoviruses, that form <strong>the</strong> BYDV complex, <strong>the</strong> different aphid species, which are virus<br />

vectors and <strong>the</strong> grass and o<strong>the</strong>r plant species, which are hosts <strong>of</strong> both, <strong>the</strong> virus and <strong>the</strong>ir<br />

vectors.<br />

As plants may harbour virus without contributing to its fur<strong>the</strong>r spreading, vectors are<br />

responsible for <strong>the</strong> virus spread. A vector must move <strong>the</strong> virus from <strong>the</strong> reservoir to o<strong>the</strong>r<br />

hosts. Therefore <strong>the</strong> spread <strong>of</strong> <strong>the</strong> vectors is important for <strong>the</strong> epidemiology <strong>of</strong> BYDV and<br />

influences spatial and temporal pattern <strong>of</strong> infection in <strong>the</strong> field.<br />

BYDV causes physiological and biochemical changes in host plants (Fereres et al., 1990;<br />

Jensen & Sambeek, 1972; Jensen, 1968; 1970; Orlob & Arny, 1961). Often <strong>the</strong> development<br />

<strong>of</strong> <strong>the</strong> plants is retarded after a BYDV infection.<br />

It is supposed that luteoviruses do not replicate in <strong>the</strong>ir vectors and do not directly<br />

influence aphid performance (Montllor & Gildow, 1986); <strong>the</strong>refore effects <strong>of</strong> BYDV on its<br />

vectors reflect more induced changes in <strong>the</strong> host plant physiology. Particularly <strong>the</strong><br />

development <strong>of</strong> phloem feeding aphids depends on <strong>the</strong> quality <strong>of</strong> <strong>the</strong> phloem sap (Douglas,<br />

1993). Weibull (1987) found a close relation between performance and population growth <strong>of</strong><br />

cereal aphids to <strong>the</strong> nutritional status <strong>of</strong> <strong>the</strong> host plant and <strong>the</strong> turgescence <strong>of</strong> <strong>the</strong> plant tissues.<br />

Aphids use carbohydrates derived from <strong>the</strong> phloem as <strong>the</strong>ir predominant energy resource and<br />

amino acids, e.g. for <strong>the</strong>ir protein metabolism (Rhodes et al., 1996). In preceding experiments<br />

we found that infection <strong>of</strong> <strong>the</strong> summer wheat cultivar Tinos with BYD-MAV and BYD-PAV–<br />

isolates caused changes in host selection behaviour, feeding intensity and growth rate <strong>of</strong><br />

Sitobion avenae F. (Homoptera: Aphididae). For that reason we hypo<strong>the</strong>sise that BYDV<br />

infection changes host suitability by altering concentrations <strong>of</strong> <strong>the</strong>se compounds within <strong>the</strong><br />

phloem sap to influence aphid development and behaviour which may favour secondary<br />

spread. The following questions have arisen:<br />

- Does a BYDV infection change food quality <strong>of</strong> <strong>the</strong> host?<br />

- Are <strong>the</strong>re changes in food utilisation by S. avenae?<br />

Measurements <strong>of</strong> sugar and amino acid contents in <strong>the</strong> phloem sap <strong>of</strong> infected and noninfected<br />

wheat plants were used to document BYDV-induced changes <strong>of</strong> main nutrients in <strong>the</strong><br />

phloem sap, which may be responsible for effects on aphid feeding behaviour and<br />

development. Quantifying <strong>the</strong> honeydew composition <strong>of</strong> sugar and amino acids and <strong>the</strong><br />

amount <strong>of</strong> secretion, possible changes in food utilisation by S. avenae should be recorded,<br />

particularly <strong>the</strong> amount <strong>of</strong> ingested phloem sap and <strong>the</strong> efficiency <strong>of</strong> resorption <strong>of</strong> different<br />

components from <strong>the</strong> phloem sap. The summer wheat cultivar Tinos, BYD-PAV and MAV<br />

virus isolate and a synchronised culture <strong>of</strong> a virus-free green clone <strong>of</strong> S. avenae, which was<br />

also reared on <strong>the</strong> wheat cultivar Tinos before, were used for <strong>the</strong> experiments.<br />

25


26<br />

Methods<br />

For phloem sap collection stylectomy was performed with a high frequency microcautery unit<br />

(type CA-50, 48 MHz, circa 25 Watt, Syntech). A 0.2 mm tungsten wire prepared as needle<br />

lead <strong>the</strong> amputation <strong>of</strong> aphids stylets, when <strong>the</strong>y showed <strong>the</strong> phloem sap ingestion pattern E2<br />

by <strong>the</strong> electronic monitoring system (EPG, Tjallingii, 1988).<br />

For honeydew collection specific cages (6 cm x 4 cm x 2 cm) were prepared allowing <strong>the</strong><br />

aphids to settle above a glass object slide, where <strong>the</strong> honey drops could be collected. The<br />

sugar analysis was performed according to a modified method <strong>of</strong> Fretz (1970), Juvik et al.<br />

(1982) and Nikolov and Reilly (1983) with a Hewlett Packard gas chromatograph and <strong>the</strong><br />

amino acid analysis with <strong>the</strong> amino-DABS method (Sykam) in a high pressure liquid<br />

chromatograph.<br />

The daily sequence <strong>of</strong> honeydew excretion by <strong>the</strong> aphids was measured with honeydew<br />

clocks. To calibrate <strong>the</strong> relationship between spot diameter and droplet volume, different<br />

known volumes <strong>of</strong> a 20 % sucrose solution were spotted with microcapillaries onto a prepared<br />

foil. The amount and diameter <strong>of</strong> excreted honeydew droplets per day were measured during<br />

<strong>the</strong> larval development <strong>of</strong> <strong>the</strong> aphids (L1-ecdysis).<br />

Results and discussion<br />

Content <strong>of</strong> sugar and amino acid in <strong>the</strong> phloem sap<br />

Whereas <strong>the</strong> BYDV infection <strong>of</strong> <strong>the</strong> wheat variety Tinos causes no differences in <strong>the</strong> sugar<br />

concentration <strong>of</strong> <strong>the</strong> phloem sap, <strong>the</strong>re was a quantitative reduction in <strong>the</strong> amino acids<br />

concentration in <strong>the</strong> phloem sap <strong>of</strong> BYDV-infected wheat (Tab. 1)<br />

Tab. 1. Total sugar and amino acid concentration in <strong>the</strong> phloem sap <strong>of</strong> BYDV-infected and<br />

non-infected wheat Tinos. Different letters indicate significant differences, Scheffétest,<br />

p < 0.05.<br />

Treatment<br />

total concentration <strong>of</strong><br />

sugar<br />

(µg / 100 µg)<br />

amino acid<br />

Nmol / µl<br />

in <strong>the</strong> phloem sap<br />

Non-infected 21.54 ± 0.73 a 223.63 ± 70.45 a<br />

(8) (9)<br />

BYD-MAV 22.09 ±0.97 a 74.83 ± 34.95 b<br />

(9) (12)<br />

BYD-PAV 21.85 ± 1.44 a 78.84 ± 11.23 b<br />

(11) (14)<br />

Changes in <strong>the</strong> relative concentration <strong>of</strong> single amino acids in <strong>the</strong> phloem sap <strong>of</strong> BYD-MAV<br />

and BYD-PAV wheat compared to non-infected wheat showed a strong reduction for most <strong>of</strong><br />

<strong>the</strong> amino acids after a BYDV infection; except for <strong>the</strong> amino acid methionine, where <strong>the</strong><br />

concentration increased after a BYDV infection (Fig. 1). These findings correspond with<br />

results <strong>of</strong> several investigations, which describe substantial differences in <strong>the</strong> free amino acid<br />

pool <strong>of</strong> virus infected plants compared to healthy plants (Lowe & Strong, 1963; Blua et al.,<br />

1994).


27<br />

For example <strong>the</strong> three amino acids aspartic acid, glutamic acid and glutamine (Asp, Glu, Gln)<br />

are important for <strong>the</strong> nutritional supply <strong>of</strong> aphids and <strong>of</strong> basic importance for amino acid and<br />

protein metabolism (Febvay et al.,1988). In our experiments all three amino acids were <strong>the</strong><br />

most abundant amino acids in <strong>the</strong> phloem sap in all treatments independent <strong>of</strong> a virus<br />

infection. However, <strong>the</strong> relative concentration <strong>of</strong> <strong>the</strong>m were reduced in <strong>the</strong> phloem sap <strong>of</strong><br />

BYDV-infected wheat.<br />

% 132.1<br />

100<br />

80<br />

BYD-MAV<br />

BYD-PAV<br />

60<br />

40<br />

20<br />

0<br />

* * *<br />

* * *<br />

* *<br />

* * * * *<br />

-20<br />

-40<br />

-60<br />

-80<br />

-100<br />

Asp Glu Gln Ser The Gly Ala Arg Amib Pro Val Met Try Ileu Leu Phe CySH Lys His Tyr<br />

Fig. 1. Changes in <strong>the</strong> relative concentration <strong>of</strong> single amino acids in <strong>the</strong> phloem sap <strong>of</strong><br />

BYDV-infected wheat compared to non-infected wheat Tinos. * indicates significant<br />

differences, Scheffé-test, p < 0.05.<br />

Content <strong>of</strong> sugar and amino acid in <strong>the</strong> honeydew<br />

In addition to <strong>the</strong> phloem sap analysis, <strong>the</strong> hypo<strong>the</strong>sis <strong>of</strong> nutritional deficiencies on virus<br />

infected plants is supported by <strong>the</strong> honeydew experiments. In honeydew <strong>of</strong> S. avenae <strong>the</strong><br />

carbohydrate concentration was slightly higher when feeding on BYD-MAV and BYD-PAV<br />

infected plants in contrast to <strong>the</strong> total amino acid concentration which was marginally reduced<br />

on BYDV-infected wheat compared to non-infected wheat (tab. 2).<br />

Tab. 2. Total sugar and amino acid concentration in <strong>the</strong> honeydew <strong>of</strong> S. avenae on BYDVinfected<br />

and non-infected wheat Tinos. Different letters indicate significant<br />

differences, Scheffé-test, p < 0.05.<br />

Treatment<br />

total concentration <strong>of</strong><br />

sugar<br />

(µg / 100 µg)<br />

amino acid<br />

Nmol / µl<br />

in <strong>the</strong> honeydew<br />

Non-infected 12.57 ± 0.97 a 14.59 ± 1.54 a<br />

(49) (59)<br />

BYD-MAV 14.70 ±1.44 a 10.99 ± 0.97 a<br />

(40) (46)<br />

BYD-PAV 14.84 ± 0.73 a 12.99 ± 1.08 a<br />

(62) (73)


28<br />

Also in <strong>the</strong> honeydew most amino acids were still reduced on both BYDV treatments<br />

except for aspartic acid, glutamic acid and methionine (fig. 2). The differences in<br />

concentration <strong>of</strong> several amino acids were not homogenous between BYD-MAV or BYD-<br />

PAV and healthy plants. Whereas <strong>the</strong> amino acid arginine was significantly reduced in <strong>the</strong><br />

honeydew from S. avenae feeding on BYD-MAV plants, <strong>the</strong> amino acids serine and proline<br />

were significantly reduced on BYD-PAV plants. Honeydew from aphids on BYDV-MAV<br />

plants contained much lower relative amounts (compared to virus free plants) <strong>of</strong> cysteine,<br />

histidine and tyrosine than that <strong>of</strong> BYD-PAV infected plants.<br />

%<br />

100<br />

80<br />

BYD-MAV<br />

BYD-PAV<br />

60<br />

40<br />

20<br />

0<br />

*<br />

*<br />

*<br />

-20<br />

-40<br />

-60<br />

-80<br />

-100<br />

Asp Glu Gln Ser Thr Gly Ala Arg Amib. Pro Val Met Try Ileu Leu Phe Cys Lys His Thy<br />

Fig. 2. Changes in relative concentration <strong>of</strong> single amino acids in honeydew <strong>of</strong> S. avenae on<br />

BYDV-infected wheat compared to non-infected wheat Tinos. * indicates significant<br />

differences, Scheffé-test, p < 0.05.<br />

Comparing <strong>the</strong> amino acid concentration in <strong>the</strong> phloem sap with <strong>the</strong> concentration in <strong>the</strong><br />

honeydew <strong>of</strong> S. avenae, it could be concluded that <strong>the</strong> aphids absorbed and <strong>the</strong>refore utilised<br />

more amino acids from phloem sap on <strong>the</strong> non-infected plants. As aphids are strongly<br />

nitrogen dependent, <strong>the</strong>ir ability to efficiently exploit <strong>the</strong> amino acids <strong>of</strong> <strong>the</strong>ir diet is a crucial<br />

factor for <strong>the</strong>ir development and reproduction.<br />

Honeydew excretion<br />

A compensatory effect by an increased phloem feeding on infected plants could not be<br />

observed. In our experiments a BYDV infection with <strong>the</strong> MAV- or <strong>the</strong> PAV-strain caused<br />

even a reduction in honeydew excretion by S. avenae during <strong>the</strong> whole larval development.<br />

In fact, <strong>the</strong> droplet size increased in all three treatments with proceeding larval<br />

development due to <strong>the</strong> increasing size <strong>of</strong> <strong>the</strong> aphids independent <strong>of</strong> an virus infection.<br />

Simultaneously <strong>the</strong> number <strong>of</strong> honeydew droplets per day decreased in all treatments.<br />

Particularly in <strong>the</strong> second half <strong>of</strong> <strong>the</strong> larval development, <strong>the</strong> number <strong>of</strong> excreted honeydew<br />

droplets per day was higher on non-infected than on BYDV-infected wheat. This entailed an<br />

increased amount <strong>of</strong> honeydew production on non-infected wheat during <strong>the</strong> larval<br />

development which could also be observed on BYD-MAV-infected wheat, but with a much<br />

lower rate. In total <strong>the</strong> excreted honeydew volume was significantly higher on non-infected<br />

compared to BYDV-infected wheat (fig. 3).


29<br />

As honeydew excretion corresponds to phloem ingestion (Auclair, 1963; Rhodes et al.,<br />

1996), S. avenae seems to reduce <strong>the</strong> amount <strong>of</strong> phloem ingestion on BYDV-infected wheat.<br />

According to Prado and Tjallingii (1999), a reduced phloem ingestion bases mainly on <strong>the</strong><br />

pressure and <strong>the</strong> viscosity <strong>of</strong> <strong>the</strong> phloem sap, because aphids ingest phloem sap passively. No<br />

compensatory feeding behaviour has also been observed by Weibull (1987). He found that <strong>the</strong><br />

ingestion rate was related to <strong>the</strong> nutrient quality <strong>of</strong> <strong>the</strong> phloem sap, particularly <strong>the</strong> level <strong>of</strong><br />

some amino acids. Therefore Weibull (1987)concluded that individual amino acids or groups<br />

<strong>of</strong> <strong>the</strong>m may also regulate <strong>the</strong> feeding rate, but that <strong>the</strong> distinction between phagostimulatory<br />

and compensatory mechanisms are difficult to make.<br />

4,0<br />

amount <strong>of</strong> honeydew (µl) / aphid and day<br />

3,5<br />

3,0<br />

2,5<br />

2,0<br />

1,5<br />

1,0<br />

0,5<br />

non-infected BYD-MAV BYD-PAV<br />

non-infected<br />

BYD-PAV<br />

BYD-MAV<br />

0,0<br />

0 1 2 3 4 5 6 7 8 9 10<br />

age <strong>of</strong> aphids (in days)<br />

Fig. 3. Amount <strong>of</strong> honeydew excretion (µl) per day by S. avenae during <strong>the</strong>ir larval<br />

development on BYDV-infected and non infected wheat Tinos.<br />

Comparing <strong>the</strong> reduced reproduction capacity and growth rate <strong>of</strong> S. avenae on BYDVinfected<br />

to non-infected wheat (Fiebig & Poehling, 1998) one can suppose that <strong>the</strong> nutritional<br />

quality <strong>of</strong> BYDV-infected wheat plants was lower for <strong>the</strong> aphids compared to uninfected<br />

ones, whereas phloem location depending on sugar concentration seems to be no problem for<br />

S. avenae on BYDV-infected wheat, as shown by electronic monitoring <strong>of</strong> stylet penetrations<br />

(Fiebig & Poehling, 1998). Mittler and Meikle (1991) found that aphids need to ingest a<br />

sufficiently large volume <strong>of</strong> phloem sap for <strong>the</strong>ir growth and reproduction in order to obtain<br />

adequate amounts <strong>of</strong> o<strong>the</strong>r less abundant nutrients, particularly amino acids. Therefore <strong>the</strong><br />

nutrient supply for S. avenae on <strong>the</strong> virus treatments was ra<strong>the</strong>r low, because <strong>of</strong> <strong>the</strong> reduced<br />

amount <strong>of</strong> phloem ingestion by S. avenae on BYDV-infected wheat (Fiebig & Poehling,<br />

1998), combined with <strong>the</strong> low total amino acid concentration in this treatment.<br />

In general our findings confirm that <strong>the</strong> disease induced changes in host plant suitability<br />

for sap-feeding insects on virus infected plants. At <strong>the</strong> same time this reflects <strong>the</strong> crucial role<br />

<strong>of</strong> <strong>the</strong> multitrophic interaction pattern between plants, virus and vectors. The efficiency <strong>of</strong><br />

aphids as vectors depends in part on <strong>the</strong>ir physiological and behavioural responses to <strong>the</strong><br />

virus-induced biochemical or physical changes in host plants. Behavioural responses <strong>of</strong><br />

aphids on infected plants include attraction, settling or feeding on <strong>the</strong> plants. Such responses<br />

can strongly influence vector distribution and movement and are particularly important<br />

because <strong>of</strong> <strong>the</strong>ir potential effects on <strong>the</strong> spread <strong>of</strong> <strong>the</strong> virus in <strong>the</strong> field. In addition to <strong>the</strong><br />

spatial dimension, nutritional quality triggers aphid development and is related to settling,


30<br />

wing polymorphism and fitness. In our experiments BYD-MAV or BYD-PAV infection not<br />

only reduced host suitability and induced spread <strong>of</strong> wingless morphs, but additionally<br />

promoted <strong>the</strong> production <strong>of</strong> alatae progeny as shown by Fiebig and Poehling (1998). Hence,<br />

both factors could be favourable for <strong>the</strong> mutualistic effects leading to virus-induced vector<br />

spread.<br />

References<br />

Auclair, J.L., 1963: Aphid feeding and nutrition. Ann. Rev. Entomol. 8: 439-490.<br />

Blua, M.J., Perring, P.A. & Madore, M.A., 1994: Plant virus-induced changes in aphid<br />

population development and temporal fluctuations in plant nutrients. J. Chem. Ecol.<br />

20(3): 691-707.<br />

Douglas, A.E., 1993: The nutritional quality <strong>of</strong> phloem sap utilised by natural aphid<br />

populations. Ecol. Entomol. 18: 31-38.<br />

Fiebig, M. & Poehling, H.-M., 1998: Host-plant selection and population dynamics <strong>of</strong> <strong>the</strong><br />

grain aphid Sitobion avenae (F.) on wheat infected with Barley Yellow Dwarf Virus.<br />

Integrated control in cereal crops. <strong>IOBC</strong>/wprs Bull. 21(8): 51-62.<br />

Febvay, G., Bonnin, J., Rabhé, Y., Bournoville, R., Dolret, S. & Bonnemain, J.L., 1988:<br />

Resistance <strong>of</strong> different lucerne cultivars to <strong>the</strong> pea aphid Acyrthosiphon pisum: influences<br />

<strong>of</strong> phloem composition on aphid fecundity. Entomol. Exp. Appl. 48: 127-134.<br />

Fereres, A., Araya, J.E., Housley, T.L. & Foster, J.E., 1990: Carbohydrate composition <strong>of</strong><br />

wheat infected with barley yellow dwarf virus. Z. Pflanzenkrankh. Pflanzenschutz 97(3):<br />

600-608.<br />

Fretz, T.A., Dunham, C.W. & Woodmansee, C.W., 1970: A gas chromatographic procedure<br />

for determining soluble carbohydrates extracted from leaf tissue <strong>of</strong> Ilex opaca Ait. Cv.<br />

´Miss Helen`. J. Amer. Soc. Hort. Sci. 95(1): 99-102.<br />

Jensen, S.G., 1968: Photosyn<strong>the</strong>sis, respiration and o<strong>the</strong>r physiological relationships in barley<br />

infected with barley yellow dwarf virus. Phytopathology 58: 204-208.<br />

Jensen, S.G., 1970: Metabolism and carbohydrate composition in barley yellow dwarf virusinfected<br />

wheat. Phytopathology 62: 587-592.<br />

Jensen, S.G. & van Sambeek, J.W., 1972: Differential effects <strong>of</strong> barley yellow dwarf virus on<br />

<strong>the</strong> physiology <strong>of</strong> tissues <strong>of</strong> hard red spring wheat. Phytopathology 62: 290-293.<br />

Juvik, J.A., Stevens, M.A., Rick, C.M., 1982: Survey <strong>of</strong> <strong>the</strong> genus Lycopersicon for<br />

variability in α-tomatin content. Hort. Science 17(5): 764-766.<br />

Lowe, S. & Strong, F.E., 1963: The unsuitability <strong>of</strong> some viruliferous plants as hosts for <strong>the</strong><br />

green peach aphid, Myzus persicae. J. Econ. Entomol. 56(3): 307-309.<br />

Mittler, T.E. & Meikle, T., 1991: Effects <strong>of</strong> dietary sucrose concentration on aphid honeydew<br />

carbohydrate level and rate <strong>of</strong> excretion. Entomol. Exp. Appl. 59: 1-7.<br />

Montllor, C.B., Gildow, F.E., 1986: Feeding responses <strong>of</strong> two grain aphids to barley yellow<br />

dwarf virus-infected oats. Entomol. Exp. Appl. 42: 63-69.<br />

Nikolov, Z.L. & Reilly, P.J., 1983: Iso<strong>the</strong>rmal Capillary Column Gas Chromatography <strong>of</strong><br />

trimethylsilyl disaccharides. J. Chromatography 254: 157-162.<br />

Orlob, G.B. & Arny, D.C., 1961: Some metabolic changes accompanying infection by barley<br />

yellow dwarf virus. Phytopathology 51: 768-775.<br />

Prado, E. & Tjallingii, W.F., 1999: Effects <strong>of</strong> experimental stress factors on probing<br />

behaviour by aphids. Entomol. Exp. Appl. 90: 289-300.<br />

Rhodes, J.D., Croghan, P.C. & Dixon, A.F.G., 1996: Uptake, excretion and respiration <strong>of</strong><br />

sucrose and amino acids by <strong>the</strong> pea aphid Acyrthosiphon pisum. J. Exp. Biol. 199: 1269-<br />

1276.


Tjallingii, F., 1988: Electrical recording <strong>of</strong> stylet penetration activities. In: Minks, A.K. &<br />

Harrewijn, P. (eds.): Aphids: Their Biology, Natural Enemies and Control. Vol. B.,<br />

Elsevier, Amsterdam, <strong>the</strong> Ne<strong>the</strong>rlands: 95-107<br />

Weibull, J., 1987: Seasonal changes in <strong>the</strong> free amino acids <strong>of</strong> oat and barley phloem sap in<br />

relation two plant growth stage and growth <strong>of</strong> Rhopalosiphum padi. Ann. Appl. Biol.<br />

111: 729-737.<br />

31


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 33 - 37<br />

The Russian wheat aphid on barley in Morocco:<br />

survey and identification <strong>of</strong> new sources <strong>of</strong> resistance<br />

Lhaloui, S. 1 , M. El Bouhssini 2 , S. Ceccarelli 2 , S. Grando 2 and A. Amri 2<br />

1 Institut National de la Recherche Agronomique, Centre Aridoculture, P.O. Box 589, Settat,<br />

Morocco<br />

2 International Center for Agricultural Research in <strong>the</strong> Dry Areas, P.O. Box 5466, Aleppo,<br />

Syria<br />

Summary<br />

Recently <strong>the</strong> Russian wheat aphid (Diuraphis noxia Mordvilko) has become a serious problem for<br />

production <strong>of</strong> cereals in Morocco. The damage has first been limited to high altitude regions, but<br />

lately, <strong>the</strong> pest has spread to all <strong>the</strong> cereal growing areas <strong>of</strong> <strong>the</strong> country. The heaviest populations are<br />

still found in <strong>the</strong> mountain regions where our surveys <strong>of</strong> 1996 and 1997 showed that 100% <strong>of</strong> <strong>the</strong><br />

barley fields were infested, with a mean <strong>of</strong> 46% <strong>of</strong> plants showing feeding symptoms <strong>of</strong> <strong>the</strong> pest. As<br />

part <strong>of</strong> a pest management package, large collections <strong>of</strong> barley lines were screened for resistance to <strong>the</strong><br />

pest. Six plant introductions were identified with very high levels <strong>of</strong> tolerance (PI366447, PI366449,<br />

PI366450, PI366453, Ciho1412, and Ciho9897). Recently, <strong>the</strong> screening <strong>of</strong> an ICARDA Barley<br />

nursery, both in <strong>the</strong> field and under controlled environmental conditions, showed that 17 entries were<br />

highly resistant (100%), 2 o<strong>the</strong>rs had 95% <strong>of</strong> <strong>the</strong>ir plants resistant, and a third one had an 85%<br />

resistance. These lines, as well as <strong>the</strong> selected plant introductions, have been incorporated into <strong>the</strong><br />

Moroccan barley breeding program in order to transfer this resistance into adapted cultivars.<br />

Key words: Russian wheat aphid, survey, incidence, barley, plant resistance.<br />

Introduction<br />

The Russian wheat aphid (RWA), Diuraphis noxia (Mordvilko), a pest <strong>of</strong> wheat and barley, is<br />

native to Russia and its neighbouring countries, but has now spread to most <strong>of</strong> <strong>the</strong> cereal<br />

growing regions <strong>of</strong> <strong>the</strong> world, including Africa (Hewitt et al., 1988) and America (Gilchrist et<br />

al., 1986).<br />

In Morocco, damage due to this pest has first been noticed during <strong>the</strong> 80’s, mainly in high<br />

altitude regions, but recently, <strong>the</strong> pest has become more serious, and spread to all <strong>the</strong> cereal<br />

growing areas <strong>of</strong> <strong>the</strong> country. The heaviest populations are still found in <strong>the</strong> mountain regions<br />

where our surveys <strong>of</strong> 1996 and 1997 (Tab. 1) showed that 100% <strong>of</strong> <strong>the</strong> barley fields were<br />

infested, with a mean <strong>of</strong> 46% <strong>of</strong> <strong>the</strong> plants showing aphid-feeding symptoms (Lhaloui et al.,<br />

1998).<br />

Damage caused by this pest presents very typical symptoms, i.e. long longitudinal<br />

chlorotic white to yellow streaks on <strong>the</strong> leaves, frequently associated with leaf rolling. This<br />

damage results in severe dry matter and grain yield loss when infestations are heavy.<br />

The best and most efficient way <strong>of</strong> controlling this pest in Morocco will be by genetic<br />

resistance through <strong>the</strong> development <strong>of</strong> improved resistant cultivars. Resistance is environmentally<br />

safe, compatible with o<strong>the</strong>r biological control agents, does not involve any extra cost<br />

for <strong>the</strong> farmer (same cost as for non-resistant seed, and no extra cost as compared to buying<br />

pesticides), and does not require any sophisticated equipment for its application in <strong>the</strong> field.<br />

The first screening for resistance to this pest in Morocco were conducted by Boulemane<br />

33


34<br />

(1995). He identified six plant introductions, that showed very adequate levels <strong>of</strong> resistance.<br />

These entries have been included into ICARDA’s Barley Improvement Program, and<br />

selections coming out <strong>of</strong> crosses with <strong>the</strong>se lines are now in an advanced stage.<br />

Tab. 1 Russian wheat infestation levels on barley in <strong>the</strong> highland regions <strong>of</strong> Morocco. 1996<br />

and 1997.<br />

Regions<br />

Number fields<br />

sampled<br />

% fields<br />

infested<br />

% plants<br />

infested<br />

Annoceur 7 100 88<br />

Guigou 7 100 35<br />

Midelt 4 100 30<br />

El Ksibah 7 100 34<br />

Azilal 4 100 44<br />

Total/Mean 29 100 46<br />

The objective <strong>of</strong> this study were to screen diverse germplasm collections in order to<br />

identify new sources <strong>of</strong> resistance to <strong>the</strong> RWA that can be incorporated into <strong>the</strong> barley<br />

breeding program <strong>of</strong> Morocco, and serve as a gene bank for an eventual change in <strong>the</strong> pest<br />

genetics, causing new biotypes to develop.<br />

Material and methods<br />

Greenhouse screening<br />

A RWA nursery (31 entries) obtained from ICARDA was screened for resistance to this pest<br />

<strong>the</strong> Aridoculture centre, Morocco. Five <strong>of</strong> <strong>the</strong>se entries have already been selected as resistant<br />

in Morocco, and have been included as checks. They originated from crosses made with <strong>the</strong><br />

barley sources selected by Boulmane (1995). A local susceptible cultivar ‘Tamelalt’ was also<br />

included as a check. The study was conducted in a growth chamber adjusted to 20±2°C and a<br />

photo-period <strong>of</strong> 16h:8h (light:dark). Entries were planted in standard greenhouse wooden<br />

flats, containing a mixture <strong>of</strong> 2/3 soil and 1/3 peat, at a rate <strong>of</strong> four seeds per hill. Each flat<br />

contained 12 entries including a check. When plants were at <strong>the</strong> two-leaf stage, <strong>the</strong>y were<br />

thinned to three plants per hill, <strong>the</strong>n infested with RWA adults at a rate <strong>of</strong> five individuals per<br />

plant. Then, flats were covered with plastic cages having cheese cloth tops for ventilation.<br />

Insects used in <strong>the</strong> test were randomly collected from a culture maintained at <strong>the</strong> centre on <strong>the</strong><br />

susceptible cultivar ‘Tamelalt’, and originally started with aphids collected from neighbouring<br />

fields.<br />

Scoring method<br />

Scorings were conducted one, two, and three weeks after infestation. Resistance was<br />

evaluated using <strong>the</strong> damage score <strong>of</strong> 1-6 rating system developed by DuToit (1987), where 1<br />

is <strong>the</strong> most resistant and 6 is <strong>the</strong> most susceptible. This rating system is based on <strong>the</strong> presence<br />

or absence <strong>of</strong> chlorotic spots, streaking, and leaf rolling. Plant damage is rated visually by<br />

scoring individual entries on <strong>the</strong> 1-6 damage rating scale:<br />

1. small isolated chlorotic spots on <strong>the</strong> leaves: highly resistant plant.<br />

2. larger isolated chlorotic spots on <strong>the</strong> leaves: resistant plant.<br />

3. chlorotic spots tend to become streaky: moderately resistant plant.


35<br />

4. mild streaks visible and leaves tend to roll lengthwise: moderately susceptible plant.<br />

5. prominent white-yellow streaks present and leaves tightly rolled: susceptible plant.<br />

6. severe white-streaks, leaves tightly rolled and starting to die from <strong>the</strong> tips: highly<br />

susceptible plant.<br />

Tab. 2 Reaction <strong>of</strong> ICARDA barley Russian wheat aphid nursery for resistance to this pest<br />

in Morocco.<br />

Entry number/name Resistance Resistance<br />

reaction (GC) reaction (Field)<br />

R001 R R<br />

STARS-9577B=R006 HR R<br />

R034 HR R<br />

R011 HR R<br />

RO13 HR R<br />

RO15 HR R<br />

RO16 HR R<br />

RO17 HR R<br />

R018 R R<br />

R022 HR R<br />

R023 HR R<br />

R024 HR R<br />

STARS-9301B=RO27 HR HR<br />

R028 HR R<br />

R029 MR R<br />

R031 HR R<br />

RWA.M46 HR R<br />

RWA.M53 HR R<br />

RWA.M54 S S<br />

RWA.M55 HR R<br />

RWA.M56 HR R<br />

Mo.B1337/WI2291//Bonita/Weeah S S<br />

Mo.B1337/W12291//Stirling<br />

/FNCI-22 NE417/Arta S MS<br />

H.spont.41-1//ER/Apm S MS<br />

H.spont.41-1//ER/Apm S S<br />

ER/Apm//Lignee131/4/ER/Apm/3/<br />

Arr/Esp//Alger/Ceres-362-1-1 MS S<br />

Arar/H.spon.19-15//Arta MS S<br />

H.spont.38-3/6/Pld10342//Cr.115/<br />

Por/3/Bahtim 9/4/Ds/Apro/5/WI2291 S S<br />

1 SLB 45-40/H.spont.41-5 S S<br />

2 WI2269/Lignee131/3/SB73358-<br />

B-104-16-1-3//ER/Apm MS S<br />

3 Susceptible check ‘Tamelalt’ S S<br />

CG= growth chamber, S = susceptible, MS = moderately susceptible, R = resistant, MR = moderately<br />

resistant, and HR = highly resistant.


36<br />

Field screening<br />

To confirm <strong>the</strong> results <strong>of</strong> <strong>the</strong> screening under controlled environmental conditions, <strong>the</strong> same<br />

nursery was seeded at Annoceur Experimental Station, situated in a high elevation region<br />

where RWA population levels are very high. Entries were seeded in one-meter long lines and<br />

replicated four times. A susceptible check was included in each replication. Also, a Moroccan<br />

barley land race collection <strong>of</strong> 759 entries, obtained from ICARDA, was screened for<br />

resistance to RWA at this same location. Entries were also seeded in one-meter long lines, but<br />

not replicated. For both collections, <strong>the</strong> damage scoring was done when symptoms became<br />

clearly visible on <strong>the</strong> susceptible lines.<br />

Results and discussion<br />

The RWA population levels were very high in <strong>the</strong> field during late spring, and allowed for<br />

excellent infestations <strong>of</strong> <strong>the</strong> field screened material. The Moroccan barley land race collection<br />

did not exhibit any significant levels <strong>of</strong> resistance to <strong>the</strong> RWA in <strong>the</strong> field. All entries were<br />

severely damaged, with ratings going from susceptible to highly susceptible plants. Thus, this<br />

collection is considered as not carrying any sources <strong>of</strong> resistance to this pest, <strong>the</strong>refore no<br />

fur<strong>the</strong>r controlled screening will be carried on it. Inversely, <strong>the</strong> results <strong>of</strong> <strong>the</strong> RWA nursery<br />

screening were very striking; both field and greenhouse results showed <strong>the</strong> presence <strong>of</strong><br />

excellent sources <strong>of</strong> resistance to this pest. The growth chamber results showed that 17 entries<br />

were highly resistant, with almost no feeding symptoms on <strong>the</strong> leaves, and two o<strong>the</strong>rs were<br />

resistant. Of <strong>the</strong> remaining entries, one was moderately resistant, three were moderately<br />

susceptible, and 8 were susceptible (Table 2). The results <strong>of</strong> <strong>the</strong> field test confirmed those<br />

obtained in <strong>the</strong> laboratory. All <strong>the</strong> lines that were resistant in <strong>the</strong> growth chamber were also<br />

resistant in <strong>the</strong> field. A very slight difference in reaction was however noticed for line number<br />

15, which scored moderately resistant in <strong>the</strong> laboratory, but resistant in <strong>the</strong> field; and for lines<br />

23, 24, and 25 which scored susceptible in <strong>the</strong> laboratory, but moderately susceptible in <strong>the</strong><br />

field, and lines 27, 28, and 31 which scored moderately susceptible in <strong>the</strong> laboratory but<br />

susceptible in <strong>the</strong> field (Table 2). These differences may be due to <strong>the</strong> plants having a better<br />

tolerance in <strong>the</strong> field and/or to <strong>the</strong> presence <strong>of</strong> some genetic variability between <strong>the</strong><br />

populations <strong>of</strong> <strong>the</strong> aphid reared in <strong>the</strong> laboratory and that <strong>of</strong> <strong>the</strong> field.<br />

Fur<strong>the</strong>r research will be carried out to characterise <strong>the</strong> resistance in <strong>the</strong> selected lines<br />

(antibiosis, antixenosis, or tolerance), and to study <strong>the</strong> eventual genetic variability within and<br />

among <strong>the</strong> RWA populations <strong>of</strong> different regions <strong>of</strong> Morocco.<br />

Conclusion<br />

These results indicate that more than 50% <strong>of</strong> <strong>the</strong> tested lines carry very good levels <strong>of</strong><br />

resistance, and because <strong>the</strong>y have good agronomic criteria, <strong>the</strong>y can readily be incorporated<br />

into <strong>the</strong> breeding program to transfer this resistance into Moroccan adapted cultivars.<br />

References<br />

Boulemane, H., 1995: Evaluation de materiel végétal de blé, orge, et triticale pour la<br />

résistance au puceron russe au Maroc. Memoire de fin d’etude pour l’obtention du<br />

diplôme d’ingenieur d’état de l’école d’agriculture de Meknes, Maroc.<br />

Dutoit, F., 1987: Resistance in wheat (Triticum aestivum L.) to Diuraphis noxia (Homoptera:<br />

Aphididae). Cereal Res. Commun. 15:175-178.


Hewitt, P.H., 1988: The south African experience with <strong>the</strong> Russian wheat aphid. P.1-3. In:<br />

Peairs, F.B. & Pilcher, S.D. (eds.): Proc. 2 nd Russian wheat aphid Workshop, Denver, Co.<br />

11-12 Oct. 1988. Colorado State Univ. Ft. Collins, Co, USA.<br />

Lhaloui, S., El Bouhssini, M. & Amri, A., 1998: Survey <strong>of</strong> <strong>the</strong> Russian wheat aphid<br />

populations in <strong>the</strong> Highland regions <strong>of</strong> Morocco. INRA-CRRA Settat annual research<br />

Report.<br />

37


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 39 - 47<br />

Effects <strong>of</strong> induced tolerance and induced resistance against aphids in wheat<br />

Martina Galler and Hans-Michael Poehling<br />

Institut für Pflanzenkrankheiten und Pflanzenschutz, Universität Hannover, Herrenhäuser<br />

Str. 2, D-30419 Hannover, Germany<br />

Summary<br />

In field- and greenhouse experiments resistance and tolerance in wheat induced by application <strong>of</strong> <strong>the</strong><br />

microbial inducer B50 and <strong>the</strong> syn<strong>the</strong>tic inducer Bion ® against <strong>the</strong> grain aphid Sitobion avenae were<br />

examined. B50 induced both tolerance as well as resistance to S. avenae. Resistance was <strong>the</strong>reby based<br />

on <strong>the</strong> mechanism <strong>of</strong> antibiosis. A tolerance effect <strong>of</strong> Bion ® could only be estimated on yield<br />

parameters in <strong>the</strong> field, but no tolerance effect <strong>of</strong> this inducer could be measured on wheat seedlings<br />

with a tolerance index. Fur<strong>the</strong>r, no significant systemic resistance effects could be observed after a<br />

Bion ® treatment. First results <strong>of</strong> honeydew excretion <strong>of</strong> S. avenae indicated that <strong>the</strong> application <strong>of</strong> B50<br />

changed <strong>the</strong> nutritional quality <strong>of</strong> <strong>the</strong> plant for <strong>the</strong> aphids.<br />

Key words: S. avenae, wheat, induced tolerance, induced resistance<br />

Introduction<br />

The terms “induced tolerance” and “induced resistance” are used in <strong>the</strong> literature with<br />

different meanings and assumptions. Karban and Baldwin (1997) defined “induced<br />

resistance” as regarded from <strong>the</strong> herbivore’s point <strong>of</strong> view and “induced tolerance” as<br />

regarded from <strong>the</strong> plant’s point <strong>of</strong> view. Many insect-induced plant responses against<br />

herbivores have been shown (Brody & Karban, 1989; Wool & Hales, 1996; Agrawal, 1998),<br />

and some <strong>of</strong> <strong>the</strong> underlying mechanism have already been described (McCloud & Baldwin,<br />

1997; Trewhella et al., 1997; Agrawal et al., 1999). In addition, resistance and/or tolerance<br />

reactions in plants can also be triggered by chemical substances, so called defence-inducers,<br />

that can be <strong>of</strong> natural or syn<strong>the</strong>tic origin. Effects <strong>of</strong> <strong>the</strong>se substances on pathogens are well<br />

documented (Steiner, 1989; Schneider & Ullrich; 1994; Kogel et al., 1995; Görlach et al.,<br />

1996). However, in only few experiments effects on foliage feeding or sucking insects have<br />

so far been studied (Wittmann & Schönbeck, 1996; Stout et al., 1999). In <strong>the</strong> present study we<br />

investigated two chemical inducers, <strong>the</strong> microbial inducer B50 and <strong>the</strong> syn<strong>the</strong>tic inducer<br />

Bion ® for <strong>the</strong>ir potential to induce resistance and/or tolerance to S. avenae in wheat.<br />

Material and methods<br />

Induction<br />

Plants were treated with <strong>the</strong> microbial inducer B50, metabolites <strong>of</strong> Bacillus subtilis,<br />

(Schönbeck et al., 1980), or <strong>the</strong> syn<strong>the</strong>tical inducer Bion ® (BTH; benzo(1,2,3)thiadiazole-7-<br />

carbothioic acid S-methyl ester), a commercial product from Novartis. Both inducers were<br />

sprayed on <strong>the</strong> plants. Control plants were treated with water.<br />

Field experiments<br />

The field trials were conducted in 1997 and 1998 with summer wheat Triticum aestivum<br />

L.(cv. Remus) on <strong>the</strong> campus <strong>of</strong> <strong>the</strong> Faculty <strong>of</strong> Horticulture at Hannover University. A 6 × 6<br />

Latin square design was used with a plot size <strong>of</strong> 17.5 m 2 . The plants were fertilised with<br />

39


40<br />

120 kg N/ha (‘Nitrophoska’; N:P:K:Mg = 12:12:17:2). In addition, all plots were treated with<br />

<strong>the</strong> herbicide ‘Pointer’ (Tribenuron) at EC22 and <strong>the</strong> fungicide ‘Opus top’ (a.i.<br />

Epoxiconazol/Fenpropimorph) at EC65. The inducers were applied with a portable<br />

compression sprayer, B50 was sprayed twice at EC32 and EC61 (500 l/ha; 1:5 dilution <strong>of</strong> its<br />

normal concentration) and Bion ® once at EC32 (60g/ha). In each plot, 10 tillers <strong>of</strong> first order<br />

were marked and <strong>the</strong> population densities <strong>of</strong> S. avenae were monitored weekly by counting<br />

<strong>the</strong> number <strong>of</strong> individuals per ear from May to July. Since natural infestation was low, at<br />

EC69 five adult aphids from a laboratory culture were added to each marked ear. At EC71,<br />

<strong>the</strong> insecticide ‘Pirimor’ (a.i. Pirimicarb) was applied to three plots <strong>of</strong> each treatment to<br />

achieve aphid free plants. After ripening, each marked ear was harvested separately to<br />

determine <strong>the</strong> direct yield parameters. For quantification <strong>of</strong> tolerance effects aphid infestation<br />

was expressed as aphid index, which considers <strong>the</strong> number <strong>of</strong> aphids and <strong>the</strong> time <strong>of</strong><br />

infestation (Wratten & Lee, 1979).<br />

Greenhouse experiments<br />

Experiments in <strong>the</strong> greenhouse were carried out to clarify whe<strong>the</strong>r tolerance could also be<br />

measured in young plants. Fur<strong>the</strong>rmore, <strong>the</strong> potential <strong>of</strong> <strong>the</strong> inducers to evoke antibiosis and<br />

antixenosis effects was investigated.<br />

Plant rearing, cultivation and inducer application<br />

Summer wheat Triticum aestivum L cv. Remus was used in <strong>the</strong> experiments. The plants were<br />

grown in a greenhouse chamber at 20 ± 4 °C, 50-80 % relative humidity and 16:18h<br />

photoperiod (L:D). They were sown in commercial soil (Fruhstorfer Einheitserde) in pots <strong>of</strong><br />

11 cm diameter and watered sufficiently. In <strong>the</strong> host preference experiment, plants were<br />

grown in soil in Petri dishes <strong>of</strong> 12 cm diameter. The inducer solutions (B50 1:3 dilution <strong>of</strong><br />

culture filtrate, 0,1 g/l Bion ® ) were applied using a laboratory spraying device.<br />

1. Determination <strong>of</strong> tolerance in young plants<br />

In <strong>the</strong>se experiments <strong>the</strong> effects <strong>of</strong> <strong>the</strong> inducer B50 and Bion ® were tested separately. They<br />

were sprayed on <strong>the</strong> plants at EC12 and <strong>the</strong> control plants were treated with water. Four days<br />

after <strong>the</strong> application, half <strong>of</strong> <strong>the</strong> plants were infested with five synchronised adults <strong>of</strong><br />

S. avenae, which were attached in clip-cages to <strong>the</strong> second leaf <strong>of</strong> <strong>the</strong> plant. Hence, <strong>the</strong>re were<br />

four treatments: control plants, B50 or Bion ® treated plants, with and without aphids. One<br />

experimental unit consisted <strong>of</strong> six replicates <strong>of</strong> each treatment, a total <strong>of</strong> nine units were<br />

investigated. Seven days after infestation, half <strong>of</strong> <strong>the</strong> plants were harvested and weighed and<br />

<strong>the</strong> whole trial was terminated 14 days after infestation. According to Reese et al. (1994), <strong>the</strong><br />

regression between <strong>the</strong> weight <strong>of</strong> non-infested and infested plants was used as tolerance<br />

index. To investigate potential effects <strong>of</strong> <strong>the</strong> inducers on aphid biomass, <strong>the</strong> aphids on each<br />

plant were counted, dried and weighed.<br />

2. Determination <strong>of</strong> resistance<br />

Antixenosis. The host preference experiment was carried out in cylinders (inner diameter:<br />

58 cm, 100 cm high) as a free-choice test. A total <strong>of</strong> 16 cylinders were used. Ten plants were<br />

sown in one Petri dish at a time and fertilised twice with 0.8 % ‘Nitrophoska’ solution.<br />

Induction took place at EC11. Three days after inducer application, four Petri dishes <strong>of</strong> each<br />

treatment were placed in a random order in one cylinder and 100 winged aphids from <strong>the</strong><br />

laboratory culture were added. Thus, <strong>the</strong> aphids were allowed to choose between control<br />

plants, B50 and Bion ® treated plants. Six days later, aphid preference and colonisation were<br />

assessed, by determining <strong>the</strong> proportion <strong>of</strong> infested plants <strong>of</strong> each treatment and <strong>the</strong> number<br />

<strong>of</strong> aphids per plant.


41<br />

Antibiosis - systemic effects. The antibiosis effects <strong>of</strong> <strong>the</strong> inducers on S. avenae were tested on<br />

plants at EC61. The inducers were sprayed only on <strong>the</strong> lower part <strong>of</strong> <strong>the</strong> plants, but not on <strong>the</strong><br />

ear and <strong>the</strong> flag leaf. Four days after <strong>the</strong> application, three synchronised apterous adults <strong>of</strong><br />

S. avenae were placed in one clip-cage that was attached to <strong>the</strong> flag leaf <strong>of</strong> each plant. Adults<br />

were kept in <strong>the</strong> cages for 24h and <strong>the</strong>n removed. Two L 1 per plant were allowed to complete<br />

<strong>the</strong>ir development. The weight <strong>of</strong> <strong>the</strong> larvae, <strong>the</strong> adult weight and developmental time were<br />

measured to calculate <strong>the</strong> relative growth rate (RGR) according to Howard and Dixon (1995).<br />

Then <strong>the</strong> adults were transferred individually to <strong>the</strong> flag leaf <strong>of</strong> new plants that had <strong>the</strong> same<br />

developmental stage and that had been induced 3-5 days prior to <strong>the</strong> infestation. The prereproductive<br />

time (d) and <strong>the</strong> effective fecundity were taken, to calculate <strong>the</strong> intrinsic rate <strong>of</strong><br />

natural increase (r m ) according to Wyatt and White (1977). Fur<strong>the</strong>r, <strong>the</strong> weight <strong>of</strong> <strong>the</strong> larvae<br />

which were produced in a time span equivalent to <strong>the</strong> pre-reproductive time was determined<br />

within <strong>the</strong> first 24 h after birth. Sample size per treatment ranged from 10 to 20 replicates.<br />

3. Physiological aspects<br />

Honeydew-excretion. Induced responses <strong>of</strong> plants to herbivores involve multiple mechanisms;<br />

i.e. modification <strong>of</strong> secondary metabolite concentrations, rearrangement <strong>of</strong> resources etc.<br />

(Karban & Myers, 1989). In <strong>the</strong> present study, honeydew excretion <strong>of</strong> S. avenae on induced<br />

wheat plants were investigated to ga<strong>the</strong>r information on potential effects on food ingestion <strong>of</strong><br />

aphids. Frequency and amount <strong>of</strong> honeydew excretion are suitable parameters to calculate <strong>the</strong><br />

feeding rate and <strong>the</strong> total amount <strong>of</strong> food ingested (Mittler & Sylvester, 1961). In <strong>the</strong>se<br />

investigations, only B50 treated plants were tested in comparison to water treated plants. Four<br />

days after application, one synchronised apterous adult <strong>of</strong> S. avenae was fixed in a clip cage<br />

to <strong>the</strong> bottom side <strong>of</strong> he third leaf <strong>of</strong> each plant for 24 h. To measure <strong>the</strong> quantity <strong>of</strong><br />

honeydew excreted, <strong>the</strong> frequency and size <strong>of</strong> honeydew droplets were estimated, using<br />

honeydew-clocks: droplets were recorded on a disc <strong>of</strong> paper treated with 4 mg Bromocresol<br />

green/1 ml 80% EtOH as an indicator. This indicator changed its colour from yellow to blue<br />

when in contact with honeydew. After removal <strong>of</strong> <strong>the</strong> clip cages, <strong>the</strong> discs were placed<br />

beneath <strong>the</strong> sucking aphids and slowly turned by a clockwork drive (one turn in 24 h).The<br />

number <strong>of</strong> droplets was counted and <strong>the</strong> diameter <strong>of</strong> <strong>the</strong> droplets was measured using an<br />

ocular micrometer.<br />

Data analysis<br />

In <strong>the</strong> field experiments, <strong>the</strong> yield parameters, <strong>the</strong> grain number per ear and <strong>the</strong> grain yield per<br />

ear, were analysed using Kruskal-Wallis analysis <strong>of</strong> variance (ANOVA). Medians were<br />

compared with <strong>the</strong> Mann-Whitney U-test. One-way-ANOVA was used to test <strong>the</strong> effects <strong>of</strong><br />

<strong>the</strong> treatments on <strong>the</strong> population development <strong>of</strong> S. avenae. Fur<strong>the</strong>rmore <strong>the</strong> effects on <strong>the</strong><br />

parameters <strong>of</strong> host preference were tested. Posteriori comparisons <strong>of</strong> means were conducted<br />

using <strong>the</strong> Tukey-test. The relationship between <strong>the</strong> weight <strong>of</strong> non-infested and infested plants<br />

was analysed by regression analysis. Life table characteristics <strong>of</strong> S. avenae were analysed by<br />

applying a pair-wise t-test, <strong>the</strong>reby testing B50 and Bion ® separately in comparison to <strong>the</strong><br />

control. The pair-wise t-test was also used to analyse <strong>the</strong> effects <strong>of</strong> B50 on <strong>the</strong> honeydew<br />

excretion.<br />

Results<br />

Field experiments<br />

The results indicated that both inducers, B50 and Bion ® , induced tolerance against S. avenae<br />

on wheat In <strong>the</strong> control, aphid infestation significantly reduced <strong>the</strong> yield per ear (Fig. 1),<br />

whereas no significant reduction was observed after <strong>the</strong> treatment with B50 and/or Bion ® . No


42<br />

significant differences in <strong>the</strong> infestation intensity (aphid index) between <strong>the</strong> treatments were<br />

found. The aphid index was 377 ± 218 on <strong>the</strong> control plants, 320 ± 160 on <strong>the</strong> B50 treated<br />

plants and 385 ± 172 on <strong>the</strong> Bion ® treated plants. The difference in <strong>the</strong> infestation intensity<br />

between <strong>the</strong> control and <strong>the</strong> B50 treated plants was too small to have an effect on <strong>the</strong> yield. A<br />

significant reduction in <strong>the</strong> yield could only be observed if great differences in <strong>the</strong> infestation<br />

intensity occur (Wratten & Lee, 1979; Nieh<strong>of</strong>f & Stäblein, 1998).<br />

Figure 2 shows <strong>the</strong> population development <strong>of</strong> S. avenae on <strong>the</strong> ear in 1997 and 1998. In<br />

both years, no significant differences were observed between <strong>the</strong> treatments. However, a trend<br />

towards lower aphid infestation in B50 treated plots was visible in both years<br />

Fig. 1. Effect <strong>of</strong> B50 and Bion ® on <strong>the</strong> number <strong>of</strong> grains per ear (left) and on <strong>the</strong> grain yield<br />

per ear (right) <strong>of</strong> summer wheat cv. Remus after an infestation with S. avenae in <strong>the</strong><br />

field experiment 1997 (Box-Whisker-Plot with median as localisation parameter;<br />

different letters indicate significant differences; Kruskal-Wallis-test with U-Test <strong>of</strong><br />

Mann-Whitney, P < 0,05; n = 60).<br />

number <strong>of</strong> aphids / ear<br />

number <strong>of</strong> aphids / ear<br />

grains / ear<br />

grain yield / ear [g]<br />

Fig. 2. Population development <strong>of</strong> S. avenae on summer wheat cv. Remus on untreated and<br />

induced plants (B50 and Bion ® ) in field experiments 1997 and 1998 (mean ±<br />

standard deviation; n = 60).


43<br />

Greenhouse experiments<br />

1. Determination <strong>of</strong> tolerance in young plants<br />

The relationship between infested and control seedling weights differed in both inducers<br />

treatments (Fig. 3). In comparison to <strong>the</strong> control, <strong>the</strong> B50 treatment resulted in a significantly<br />

steeper slope, indicating increased tolerance. In contrast, no effect <strong>of</strong> <strong>the</strong> Bion ® treatment<br />

could be observed. No significant differences in <strong>the</strong> dry weight <strong>of</strong> aphids per plant, used for<br />

an estimation <strong>of</strong> <strong>the</strong> infestation density, could be observed. In <strong>the</strong> experiment with B50 <strong>the</strong><br />

aphid dry weight per plant was 4.79 ± 1.35 mg in <strong>the</strong> control , and 4.71 ± 0.93 mg in <strong>the</strong> B50<br />

treatment. In <strong>the</strong> experiment with Bion ® <strong>the</strong> dry weight <strong>of</strong> <strong>the</strong> aphids was 4.6 ± 1,53 mg in<br />

<strong>the</strong> control, and 4.3 ± 1,3 mg in <strong>the</strong> Bion ® treatment.<br />

fresh weight <strong>of</strong> infested plants [g]<br />

fresh weight <strong>of</strong> infested plants [g]<br />

Fig. 3. Effect <strong>of</strong> an inducer treatment on <strong>the</strong> relationship between <strong>the</strong> fresh weight <strong>of</strong> noninfested<br />

and infested seedlings <strong>of</strong> summer wheat cv. Remus when infested with<br />

S. avenae (each point is <strong>the</strong> mean <strong>of</strong> each unit; n = 6).<br />

2. Determination <strong>of</strong> resistance<br />

The resistance effects <strong>of</strong> <strong>the</strong> inducers were investigated in a host preference experiment in<br />

order to examine <strong>the</strong> antixenotic properties and <strong>the</strong> growth parameters <strong>of</strong> S. avenae and to<br />

determine whe<strong>the</strong>r <strong>the</strong> inducers have a systemic effect on <strong>the</strong> development <strong>of</strong> S. avenae. No<br />

significant differences in <strong>the</strong> proportion <strong>of</strong> infested plants and in <strong>the</strong> number <strong>of</strong> aphids per<br />

plant could be detected among <strong>the</strong> various treatments (Tab. 1). However, <strong>the</strong> relative growth<br />

rate <strong>of</strong> S. avenae and also <strong>the</strong> weight <strong>of</strong> <strong>the</strong> larvae was significantly reduced on B50 treated<br />

plants (Tab. 2).<br />

3. Physiological aspects<br />

Honeydew-excretion<br />

The amount <strong>of</strong> honeydew excretion indicated that apterous imagines <strong>of</strong> S. avenae produced<br />

more honeydew on B50 treated plants compared with aphids on control plants.<br />

Sitobion avenae excreted significantly fewer drops <strong>of</strong> honeydew on control plants than on<br />

B50 treated plants, but <strong>the</strong> size <strong>of</strong> <strong>the</strong> honeydew drops was not affected (fig. 4).


44<br />

Tab. 1. Effect <strong>of</strong> B50 and Bion ® on <strong>the</strong> host preference <strong>of</strong> S. avenae on summer wheat cv.<br />

Remus (means ± standard deviation; n=16).<br />

Treatment Control B50 Bion ®<br />

Proportion <strong>of</strong> infested<br />

plants [%]<br />

Number <strong>of</strong> aphids /<br />

plant<br />

35.6 ± 6.1 31.4 ± 8.0 33.0 ± 7.5<br />

5.4 ± 2.1 3.6 ± 1.9 4.5 ± 2.3<br />

Tab. 2. Systemic effect <strong>of</strong> an inducer treatment on growth, development and reproduction <strong>of</strong><br />

S. avenae on summer wheat cv. Remus; RGR (relative growth rate), M d (effective<br />

fecundity in number <strong>of</strong> <strong>of</strong>fspring), r m (intrinsic rate <strong>of</strong> natural increase); (mean ±<br />

standard deviation; means followed by different letters differ significantly; pair-wise<br />

t-test, P< 0,05).<br />

Treatment (n>10) Control B50 Bion ®<br />

Adult weight (mg) 0.93 ± 0.23a 0.87 ± 0.27a 0.84 ± 0.20a<br />

Developmental time (d) 7.01 ± 0.83a 7.55 ± 0.71a 7.27 ± 0.70a<br />

RGR 0.42 ± 0.08a 0.36 ± 0.07b 0.37 ± 0.06a<br />

Prereproductive time (d) 8.46 ± 0.93a 8.82 ± 0.60a 8.67 ± 0.78a<br />

M d 32.82 ± 5.29a 29.36 ± 8.91a 32.71 ± 7.21a<br />

r m 0.31 ± 0.03a 0.28 ± 0.04a 0.30 ± 0.02a<br />

L 1 -weight 0.05 ± 0.02a 0.04 ± 0.02b 0.06 ± 0.02a<br />

number <strong>of</strong> honeydew drops / h [n]<br />

size <strong>of</strong> honeydew drops [mm]<br />

Fig. 4. Influence <strong>of</strong> a B50 treatment on <strong>the</strong> number and size <strong>of</strong> honeydew drops produced by<br />

apterous imagines <strong>of</strong> S. avenae on summer wheat cv. Remus (different letter differ<br />

significantly, pair-wise t-test, P< 0.05; n = 7)


45<br />

Discussion<br />

The experiments showed that <strong>the</strong> microbial inducer B50 and <strong>the</strong> syn<strong>the</strong>tic inducer Bion ® , both<br />

developed to control plant diseases like powdery mildew on cereals, can have also affect<br />

plant-aphid (S. avenae) interactions. B50 mainly induces resistance and/or tolerance in cereal<br />

plants to obligate biotrophic pathogens (Oerke et al., 1989; Kehlenbeck & Schönbeck, 1995).<br />

Bion ® , a commercial product, is characterised as a resistance inducer with a broader spectrum<br />

<strong>of</strong> disease control (Friedrich et al., 1996; Sticher et al., 1997). In <strong>the</strong> present study, both B50<br />

and Bion ® induced tolerance to S. avenae in <strong>the</strong> field. A similar tolerance induction by B50 to<br />

<strong>the</strong> bird cherry aphid Rhopalosiphum padi (L.) was already described by Wittmann (1996),<br />

indicating that this effect is not strictly species specific. As determination <strong>of</strong> tolerance in<br />

terms <strong>of</strong> yield in field experiments is time consuming, we investigated whe<strong>the</strong>r it is possible<br />

to already asses <strong>the</strong> effect with young plants. Commonly, to separate tolerance and antibiosis<br />

components, <strong>the</strong> reduction in seedling weight in relation to aphid units is used as a tolerance<br />

index (Dixon et al., 1990; Robinson et al., 1991; Lamb & MacKay, 1995). Reese (1994)<br />

proposed to use <strong>the</strong> slope <strong>of</strong> <strong>the</strong> regression between <strong>the</strong> weight <strong>of</strong> non-infested and infested<br />

plants for measuring tolerance effects. By this method, <strong>the</strong> problem <strong>of</strong> quantifying tolerance<br />

when only small differences in weight between infested and non-infested plants were<br />

measured could be overcome. For this reason, in <strong>the</strong> present study <strong>the</strong> index <strong>of</strong> Reese (1994)<br />

was used. Using this laboratory based test system tolerance could only be demonstrated for<br />

B50. The reasons for <strong>the</strong> different results with young and old plants remain unclear because<br />

knowledge on <strong>the</strong> mechanisms <strong>of</strong> induced tolerance is still limited. Tolerance induction after a<br />

B50 treatment has been shown to be associated with changes in <strong>the</strong> sink-source-relationship<br />

<strong>of</strong> wheat plants infested by R. padi or powdery mildew (Wittmann, 1996). Fur<strong>the</strong>r, Gernns<br />

(1999) demonstrated that tolerance induction had an effect on <strong>the</strong> fructan reserve pools <strong>of</strong><br />

powdery mildew infected barley plants, and hence on <strong>the</strong> quality and quantity <strong>of</strong> <strong>the</strong> grain.<br />

A resistance effect <strong>of</strong> B50 was shown in <strong>the</strong> field by reduction <strong>of</strong> population development<br />

<strong>of</strong> S. avenae. In greenhouse experiments we examined whe<strong>the</strong>r resistance was caused by<br />

antibiosis or antixenosis effects. Sitobion avenae showed no preference for any <strong>of</strong> <strong>the</strong> treated<br />

plants, suggesting that resistance was mainly caused by antibiosis effects. In addition, <strong>the</strong><br />

results <strong>of</strong> <strong>the</strong> antibiosis experiments indicated a systemic effect <strong>of</strong> <strong>the</strong> B50 treatment on <strong>the</strong><br />

growth parameters <strong>of</strong> <strong>the</strong> aphids, whereas Bion ® had no significant influence. This<br />

corroborates earlier reports which already showed local resistance induction with B50 against<br />

S. avenae (Galler & Poehling, 1998). Systemic effects <strong>of</strong> B50 on pathogens were already<br />

described by Kraska (1996). Similar effects on pathogens and insect pest are not <strong>the</strong> rule: for<br />

instance, Stout et al. (1999) demonstrated induced systemic resistance on tomatoes to a<br />

bacterial pathogen, Pseudomonas syringae pv. tomato by benzothiadiazole, while at <strong>the</strong> same<br />

time suitability <strong>of</strong> leaflets for larvae <strong>of</strong> <strong>the</strong> chewing caterpillar, Helicoverpa zea was<br />

improved. This indicates that <strong>the</strong> induced response is inducer specific and depends on <strong>the</strong><br />

interacting organisms.<br />

Concerning mechanisms <strong>of</strong> resistance, antibiosis could be due to <strong>the</strong> absence or<br />

differences in concentrations <strong>of</strong> nutritional components and/or stimulants, as well as due to<br />

<strong>the</strong> presence <strong>of</strong> deterrents and/or toxic metabolites (Niraz et al., 1985). Brody and Karban<br />

(1989) suggested that induced mite resistance in cotton was caused by <strong>the</strong> degradation <strong>of</strong> <strong>the</strong><br />

nutritional quality <strong>of</strong> <strong>the</strong> plant.<br />

First attempts to clarify <strong>the</strong> underlying mechanisms centred on induced differences in<br />

nutritional components that are important for aphid development. Honeydew analysis is a<br />

convenient tool to quantify <strong>the</strong> amount <strong>of</strong> phloem sap taken up, as well as to describe diet<br />

substances ingested by <strong>the</strong> aphids. The first results <strong>of</strong> honeydew excretion <strong>of</strong> S. avenae


46<br />

showed an increase in honeydew production on B50 treated plants. This may indicate that <strong>the</strong><br />

B50 treatment reduced <strong>the</strong> nutritional value <strong>of</strong> <strong>the</strong> phloem sap which <strong>the</strong> aphids tried to<br />

compensate by increased feeding activity. Fur<strong>the</strong>r work is in progress to analyse more in<br />

detail <strong>the</strong> effects <strong>of</strong> inducer treatments on concentration and composition <strong>of</strong> carbohydrates<br />

and amino acids in <strong>the</strong> phloem sap and honeydew <strong>of</strong> S. avenae.<br />

References<br />

Agrawal, A.A., 1998: Induced responses to herbivory and increased plant performance.<br />

Science 279: 1201-1202.<br />

Agrawal, A.A., Tuzun, S. & Bent, E., 1999: Induced plant defenses against pathogens and<br />

herbivores. American Phytopathological Society Press, Minnesota.<br />

Brody, A.K. & Karban, R., 1989: Demographic analysis <strong>of</strong> induced resistance against spider<br />

mites (Acari: Tetranychidae) in cotton. J. Econ. Entomol. 82: 462-465.<br />

Dixon, A.G.O., Bramel-Cox, P.J., Reese, J.C. & Harvey, T.L., 1990: Mechanisms <strong>of</strong><br />

resistance and <strong>the</strong>ir interactions in twelve sources <strong>of</strong> resistance to biotype E greenbug<br />

(Homoptera: Aphididae) in sorghum. J. Econ. Entomol. 83: 234-240.<br />

Friedrich, L., Lawton, K., Ruess, W., Masner, P., Specker, N., Gut Rella, M., Meier, B.,<br />

Dincher, S., Staub, T., Uknes, S., Métraux, J.-P., Kessmann, H. & Ryals, J., 1996: A<br />

benzothiadiazole derivate induces systemic acquired resistance in tobacco. The Plant<br />

Journal 10: 61-70.<br />

Galler, M. & Poehling, H.-M., 1998: Induced tolerance and induced resistance against<br />

biotrophic pathogens and cereal aphids in wheat. <strong>IOBC</strong>/<strong>WPRS</strong> Bull. 21(8): 193-199.<br />

Gernns, H. 1999: Mechanismen der induzierten Kompensationsfähigkeit von Pflanzen<br />

gegenüber Schäden durch den obligat biotrophen Pilz Erysiphe graminis DC. f.sp. hordei<br />

Ém. Marchal. Ph.D. <strong>the</strong>sis, Hannover University, Germany.<br />

Görlach, J., Volrath, S., Knauf-Beiter, G., Hengy, G, Beckhove, U., Kogel, K.-H.,<br />

Oostendorp, M., Staub, T., Ward, E., Kessmann, H. & Ryals, J., 1996: Benzothiadiazole,<br />

a novel class <strong>of</strong> inducers <strong>of</strong> systemic acquired resistance, activates gene expression and<br />

disease resistance in wheat. Plant Cell 8: 629-643.<br />

Howard, M.T. & Dixon, A.F.G., 1995: Factor determining <strong>the</strong> pest status <strong>of</strong> <strong>the</strong> rose-grain<br />

aphid, Metopolophium dirhodum (Walker), on winter barley in <strong>the</strong> United Kingdom.<br />

Ann. appl. Biol. 127: 1-10.<br />

Karban, R. & Myers, J.H., 1989: Induced plant responses to herbivory. Annual Review <strong>of</strong><br />

Ecology and Systematics 20: 331-348.<br />

Karban, R. & Baldwin, I.T., 1997: Induced responses to herbivory. Univ. <strong>of</strong> Chicago Press,<br />

Chicago.<br />

Kehlenbeck, H. & Schönbeck, F., 1995: Effects <strong>of</strong> induced resistance on disease<br />

severity/yield relations in mildewed barley. J. Phytopath. 143: 561-567.<br />

Kogel, K.H., Ortel, B., Jarosch, B., Atzorn, R., Schiffer, R. & Wasternack, K., 1995:<br />

Resistance in barley against <strong>the</strong> powdery mildew fungus (Erysiphe graminis f.sp. hordei)<br />

is not associated with enhanced levels <strong>of</strong> endogenous jasmonates. J. Plant Pathol. 101:<br />

319-332.<br />

Kraska, T., 1996: Vergleichende Untersuchungen von Resistenzinduktoren, deren Wirkungsweise<br />

und dem Einfluß auf die DNA-Methylierung. Ph.D. <strong>the</strong>sis, Hannover University,<br />

Germany.<br />

Lamb, R.J. & MacKay, P.A., 1995: Tolerance <strong>of</strong> antibiotic and susceptible cereal seedlings to<br />

<strong>the</strong> aphids Metopolophium dirhodum and Rhopalosiphum padi. Ann. Appl. Biol. 127:<br />

573-583.


McCloud, E.S. & Baldwin, I.T., 1997: Herbivory and caterpillar regurgitants amplify <strong>the</strong><br />

wound-induced increases in jasmonic acid but not nicotine in Nicotiana sylvestris. Planta<br />

203: 430-435.<br />

Mittler, T.E. & Sylvester, E.S., 1961: A comparison <strong>of</strong> <strong>the</strong> injury to alfalfa by <strong>the</strong> aphids<br />

Therioaphis maculata and Macrosiphon pisi. J. Econ. Entomol. 54: 615-622.<br />

Nieh<strong>of</strong>f, B. & Stäblein, J., 1998: Vergleichende Untersuchungen zum Schadpotential der<br />

Getreideblattlausarten Metopolophium dirhodum (Wlk.) und Sitobion avenae (F.) in<br />

Winterweizen. J. Appl. Ent. 122: 223-229.<br />

Niraz, S., Leszczynski, B., Ciepiela, A. & Urbanska, A., 1985: Biochemical aspects <strong>of</strong> winter<br />

wheat resistance to aphids. Insect Sci. Applic. 6: 253-257.<br />

Oerke, E.-C., Steiner, U. & Schönbeck, F., 1989: Zur Wirksamkeit der induzierten Resistenz<br />

unter praktischen Anbaubedingungen. V. Mehltaubefall und Ertrag von Winter- und<br />

Sommergerste in Abhängigkeit von der Stickst<strong>of</strong>fdüngung. J. Plant Dis. Prot. 96: 140-<br />

153.<br />

Reese, J.C., Schwenke, J.R., Lamont, P.S. & Zehr, D.D., 1994: Importance and quantification<br />

<strong>of</strong> plant tolerance in crop pest management programs for aphids: Greenbug resistance in<br />

sorghum. Journal <strong>of</strong> Agricultural Entomology 11: 255-270.<br />

Robinson, J., Vivar, H.E., Burnett, P.A., Calhoun, D.S., 1991: Resistance to Russian wheat<br />

aphid (Homoptera: Aphididae) in barley genotypes. J. Econ. Entomol. 84: 674-679.<br />

Schneider, S. & Ullrich, W.R., 1994: Differential induction <strong>of</strong> resistance and enhanced<br />

enzyme activities in cucumber and tobacco caused by treatment with various abiotic and<br />

biotic inducers. Physiol. Mol. Plant Pathol., 45: 291-304.<br />

Schönbeck, F., Dehne, H.-W. & Beicht, W., 1980: Untersuchungen zur Aktivierung unspezifischer<br />

Resistenzmechanismen in Pflanzen. J. Plant Dis. Prot. 87: 654-666.<br />

Steiner, U., 1989: Zum Einfluß induzierter Resistenz auf den Wirt-Parasit-Komplex Gerste<br />

Echter Mehltau: Sortenabhängige Resistenzreaktionen und Befalls-Verlust-Relationen.<br />

Ph.D. <strong>the</strong>sis, Hannover University, Germany.<br />

Sticher, L., Mauch-Mani, B. & Métraux, J.P., 1997: Systemic acquired resistance. Ann. Rev.<br />

Phytopathol. 35: 235-270.<br />

Stout, M.J., Fidantsef, A.L., Duffey, S.S. & Bostock, R.M., 1999: Signal interactions in<br />

pathogen and insect attack: systemic plant-mediated interactions between pathogens and<br />

herbivores <strong>of</strong> <strong>the</strong> tomato, Lycopersicon esculentum. Physiological and Molecular Plant<br />

Pathology 54: 115-130.<br />

Trewhella, K.E., Lea<strong>the</strong>r, S.R. & Day, K.R., 1997: Insect induced resistance in Lodgepole<br />

pine: effects on two pine feeding insects. J. Appl. Ent. 121: 129-136.<br />

Wittmann, J. & Schönbeck, F., 1996: Studies <strong>of</strong> tolerance induction in wheat infested with<br />

powdery mildew or aphids. J. Plant Dis. Prot. 103: 300-309.<br />

Wool, D. & Hales, D.F., 1996: Previous infestation affects recolonization <strong>of</strong> cotton by Aphis<br />

gossypii: Induced resistance or plant damage? Phytoparasitica 24: 39-48.<br />

Wratten, S.D. & Lee, G., 1979: Duration <strong>of</strong> cereal aphid populations and <strong>the</strong> effects on wheat<br />

yield and quality. Proc. BCPC - Pests and Diseases, 1-8.<br />

Wyatt, I.J. & White, P.F., 1977: Simple estimation <strong>of</strong> intrinsic rates for aphids and tetranychid<br />

mites. Appl. Ecol. 14: 757-766.<br />

47


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 49 - 58<br />

How does a ladybird respond to aphids?<br />

Holger Triltsch, Gunnar Hechenthaler, Uwe Gosselke and Bernd Freier<br />

BBA, Institute for Integrated Plant Protection, Stahnsdorfer Damm 81,<br />

D-14532 Kleinmachnow, Germany<br />

Summary<br />

Three different experimental designs were used to investigate prey searching behaviour <strong>of</strong> adult<br />

Coccinella septempunctata and Propylea quatuordecimpunctata under laboratory conditions: 1. Y-<br />

tube-airflow-olfactometer, 2. static 4-chamber olfactometer, and 3. semi-field bioassay-arena. Y-<br />

olfactometer bioassays were performed with both ladybird species. No significant attraction to odours<br />

<strong>of</strong> aphids or aphids on host plants was observed in olfactometer bioassays. The spatial distribution <strong>of</strong><br />

ten C. septempunctata females was examined in a 2.65m² semi-field bioassay-arena which was<br />

subdivided into 64 panels. Different coloured panels or Petri dishes containing aphids were used to<br />

simulate prey and non-prey patches. Ladybird females did not respond to aphid occurrence but<br />

accumulated in panels <strong>of</strong> high brightness. Although different experimental designs were used it was<br />

not possible to detect any response to olfactory cues under laboratory conditions. Field data on <strong>the</strong><br />

spatial distribution <strong>of</strong> aphids and ladybirds in an agricultural area <strong>of</strong> 4 km² were obtained by sweep<br />

netting at two locations from March to July in 1998 and in 1999. The spatial distribution <strong>of</strong> C.<br />

septempunctata was correlated only at certain time periods with aphid occurrence. Moreover, it was<br />

observed that C. septempunctata adults tend to migrate into fields with relatively low aphid densities.<br />

They promptly disappeared at higher aphid densities and abundant occurrence <strong>of</strong> conspecific larvae.<br />

These observations indicate a behaviour in accordance with <strong>the</strong> optimal foraging <strong>the</strong>ory <strong>of</strong> ladybirds.<br />

In P. quatuordecimpunctata distribution was generally not related to aphid density.<br />

Key words: ladybird, Coccinella septempunctata, Propylea quatuordecimpunctata, prey searching,<br />

bioassay, dispersal, numerical response, optimal foraging <strong>the</strong>ory<br />

Introduction<br />

Ladybirds are considered to be important antagonists, regulating aphid population dynamics<br />

in cereals (Sunderland et al., 1983; Poehling, 1988; Freier et al., 1998). Two ecological terms<br />

describe <strong>the</strong> efficiency <strong>of</strong> a predator: 1. numerical response and 2. functional response (e.g.<br />

Gutierrez et al., 1984; Wellings & Dixon, 1987). An efficient predator should be able to<br />

aggregate in prey patches and should increase its predation rate with increasing prey density.<br />

However, in a number <strong>of</strong> field investigations it was shown that ladybird occurrence was<br />

not at all related to aphid density or only in certain crops or years (Honek, 1985; Hemptinne<br />

et al., 1994; Triltsch & Freier, 1998). Moreover, <strong>the</strong> mechanisms a ladybird uses to locate<br />

aphid infested patches are still unknown. Several authors have reported that ladybirds search<br />

randomly and discover <strong>the</strong>ir prey only by physical contact (Hodek, 1970; Frazer & Gilbert,<br />

1976). Recent findings, however, indicate that coccinellids recognise aphids visually and<br />

respond to aphid odours (Nakamuta, 1984a & b; Sengonca & Liu, 1994), but <strong>the</strong> perception<br />

distance <strong>of</strong> ladybirds was found to be very small and <strong>the</strong>se findings cannot explain why<br />

ladybirds decide to enter or leave certain fields.<br />

The key stimuli ladybirds use during <strong>the</strong>ir movement within an agricultural area, as well<br />

as <strong>the</strong>ir aphid consumption rate under natural conditions are still a field <strong>of</strong> speculation.<br />

49


50<br />

We used three different experimental designs to investigate prey searching <strong>of</strong> adults <strong>of</strong><br />

<strong>the</strong> two ladybird species Coccinella septempunctata L. and Propylea quatuordecimpunctata<br />

(L.). In order to ga<strong>the</strong>r more information about how <strong>the</strong>se ladybirds respond to aphid<br />

occurrence under field conditions, we also investigated <strong>the</strong> spatial distribution <strong>of</strong> ladybirds<br />

and aphids in two agricultural areas from March to July in 1998 and in visually 1999.<br />

Materials and methods<br />

Laboratory experiments<br />

Y-tube-airflow-olfactometer. These experiments took place at <strong>the</strong> end <strong>of</strong> September<br />

1997 and in March 1998. Adults <strong>of</strong> C. septempunctata were tested in spring and autumn,<br />

whereas P. quatuordecimpunctata adults were investigated in autumn only. Ladybird adults<br />

were collected from different hibernation sites in February and in an orchard in autumn. They<br />

were <strong>the</strong>n held in plastic boxes (5,000 cm³) under laboratory conditions (20±3 °C, 60-70 %<br />

rh) with aphid surplus. Before testing, <strong>the</strong> ladybirds were separated and kept without any food<br />

for 12-20 hours. The experiments were performed using <strong>the</strong> protocol developed by Ru<strong>the</strong>r<br />

and Thiemann (1997). During <strong>the</strong> experiments temperature ranged between 20 and 25 °C. The<br />

ladybirds were allowed to move within <strong>the</strong> Y-tube for three minutes. The two legs <strong>of</strong> <strong>the</strong> Y-<br />

tube were connected via silicon tubes with two-part glass vessels (bioassays with AFB and<br />

RPW) or wash bottles (bioassays with AF and C7F). One <strong>of</strong> <strong>the</strong> bottles (or glass vessels)<br />

contained <strong>the</strong> test material, <strong>the</strong> o<strong>the</strong>r one was empty (control). The test materials used in <strong>the</strong><br />

bioassays were:<br />

AFB: potted bean plants (DC 16) infested with Aphis fabae (Scop.) (≈500<br />

aphids/pot);<br />

RPW: potted winter wheat plants (DC 12-20) infested with Rhopalosiphum padi (L.)<br />

(≈500-1000 aphids/pot);<br />

AF: A. fabae (≈400 aphids, partly smashed);<br />

C7F: C. septempunctata adults (5 females).<br />

Static 4-chamber olfactometer. In May and June 1999 prey searching behaviour <strong>of</strong> P.<br />

quatuordecimpunctata females and males was investigated in a circular arena (diameter:<br />

200 cm) consisting <strong>of</strong> a glass ring (5 cm high) covered at both openings with gauze netting.<br />

The gauze covering at <strong>the</strong> bottom <strong>of</strong> <strong>the</strong> experimental arena was divided into four uniform<br />

areas. Each area was equipped underneath with a small vial, containing ei<strong>the</strong>r 5 aphids (test)<br />

or none (control). In <strong>the</strong>se experiments we used <strong>the</strong> cereal aphid Sitobion avenae (Fabr.). The<br />

first experiments consisted only <strong>of</strong> one test area and three control areas. In <strong>the</strong> following<br />

experiments we used two test and two control areas. The ladybirds were captured with a<br />

sweep net at a forest edge in April 1999. Before testing, all ladybirds were kept without food<br />

for about 20-24 hours and <strong>the</strong>n placed into <strong>the</strong> centre <strong>of</strong> <strong>the</strong> bioassay arena. The movement <strong>of</strong><br />

each ladybird individual was recorded for 10 minutes. The experiments took place under<br />

laboratory conditions with 22±2 °C and 60-70 % RH.<br />

Semifield bioassay-arena. A 2.65 m² quadratic bioassay arena was constructed to<br />

investigate searching behaviour <strong>of</strong> C. septempunctata adults under semi-field conditions. The<br />

bottom <strong>of</strong> this arena was subdivided into 64 uniform panels and <strong>the</strong> top was covered with<br />

gauze netting. In 16 out <strong>of</strong> <strong>the</strong> 64 panels one gauze-covered Petri dish, containing five aphids<br />

(A. fabae), was placed simulating an aphid infested field patch. Movement <strong>of</strong><br />

C. septempunctata adults within this arena was studied during January – March 1999. The


51<br />

ladybirds for this experiment were collected from four different hibernation sites and were fed<br />

ad libitum with A. fabae. Only females were chosen. They were kept 24 hours without any<br />

food before testing. Ten females were <strong>the</strong>n introduced into <strong>the</strong> arena at 07.00 a.m. and <strong>the</strong>ir<br />

distribution was recorded every 15 minutes for <strong>the</strong> next 10 hours. Temperature increased<br />

during each experiment from 16 °C in <strong>the</strong> morning to 23 °C in <strong>the</strong> afternoon. The experiment<br />

was repeated four times and <strong>the</strong> whole arena was rotated by 90° after each experiment.<br />

In ano<strong>the</strong>r experiment four different coloured panels (white, yellow, orange, dark green)<br />

were placed at <strong>the</strong> bottom <strong>of</strong> <strong>the</strong> bioassay-arena to investigate any visual orientation <strong>of</strong><br />

C. septempunctata females.<br />

Field work<br />

At <strong>the</strong> two agricultural locations, i.e. Berlin-Staaken (BS) and Nor<strong>the</strong>rn Flaeming (NF), 15-17<br />

sampling points were adopted within an quadratic area <strong>of</strong> 4 km² size. Ten sampling points<br />

were located in different field crops, <strong>the</strong> o<strong>the</strong>r points were situated at forest edges or hedges.<br />

All sampling points were arranged in order to cover a high number <strong>of</strong> different crops and<br />

habitats. From March/April to July in 1998 and 1999 sweep net catches (250 catches per<br />

sampling point) were undertaken in time distances <strong>of</strong> about ten days to estimate ladybird and<br />

aphid abundance within <strong>the</strong> two agricultural areas.<br />

Results<br />

Laboratory experiments<br />

Y-tube-airflow-olfactometer.<br />

Coccinella septempunctata adults were not significantly attracted to odours <strong>of</strong> aphids on host<br />

plants, nei<strong>the</strong>r in spring nor in autumn. In addition, <strong>the</strong>re was no indication <strong>of</strong> any attraction<br />

to odours <strong>of</strong> conspecific females (Fig. 1). Adults <strong>of</strong> P. quatuordecimpunctata were tested only<br />

in autumn and no attraction to aphids or aphids on host plants was observed. There was a<br />

slight tendency to avoid aphids on host plants during one experiment (Fig. 1).<br />

Static 4-chamber olfactometer.<br />

During <strong>the</strong>se experiments again no significant attraction to aphid odours could be observed.<br />

Although during one experiment with female P. quatuordecimpunctata a slight preference <strong>of</strong><br />

<strong>the</strong> aphid odour contaminated area was visible, in <strong>the</strong> following experiments with two test<br />

areas no such preference was recorded (Fig. 1).<br />

Semifield bioassay-arena.<br />

Results from <strong>the</strong> third type <strong>of</strong> laboratory experiments, intended to investigate prey searching<br />

<strong>of</strong> ladybird adults, indicate <strong>the</strong> same non-preference previously observed during <strong>the</strong><br />

olfactometer experiments. Females <strong>of</strong> C. septempunctata did not prefer or aggregate in <strong>the</strong><br />

aphid infested area (Fig. 2). We only found a significant reaction to different coloured panels.<br />

Coccinella septempunctata females showed a photo-positive orientation, whereby 47 % <strong>of</strong> all<br />

ladybird records belonged to <strong>the</strong> white area. The corresponding values for <strong>the</strong> o<strong>the</strong>r three<br />

areas were 20 % (yellow), 18 % (orange), and 15 % (dark green).<br />

Field work<br />

In 1998, 1,085 ladybird adults were collected at location BS, and 683 adults at location NF. In<br />

<strong>the</strong> following year <strong>the</strong> numbers were in <strong>the</strong> same range, with 959 collected adults at BS and<br />

883 at NF. Coccinella septempunctata and P. quatuordecimpunctata were <strong>the</strong> most dominant<br />

ladybird species representing 17-37 % and 13-32 %, respectively, <strong>of</strong> all ladybird adults found.


52<br />

Aphid occurrence differed between locations and years. At location BS winter rye, oats,<br />

spring wheat, and winter barley were heavily infested with aphids in 1998. Aphid densities<br />

higher than 500 aphids per sample first occurred in winter rye (beginning <strong>of</strong> May), later in<br />

spring cereals (end <strong>of</strong> May) and <strong>the</strong>n in winter barley (June). At location NF aphids were<br />

abundant in winter rye, oats, peas, winter wheat and hedges in 1998. In contrast to location<br />

BS, aphid numbers <strong>of</strong> <strong>the</strong> different crops/habitats at NF were highest in mid June. In 1999<br />

aphids were not as frequent as in <strong>the</strong> preceding year and built up <strong>the</strong>ir populations relatively<br />

late in June and July. Higher aphid densities, i.e. >500 aphids/sample, were only observed in<br />

oats (BS, NF) and pea (NF). At location BS it was again evident that aphid population first<br />

peaked in winter rye (beginning <strong>of</strong> June) and later in oats (end <strong>of</strong> June) and in winter wheat<br />

(July).<br />

100%<br />

80%<br />

Y-tube airflow-olfactometer<br />

n. sign.<br />

n = 60<br />

spring<br />

n. sign.<br />

n = 39<br />

n. sign.<br />

n = 44<br />

n. sign.<br />

n = 43<br />

autumn<br />

P


53<br />

2.65m 2 semifield bioassay-arena<br />

30% 26%<br />

24% 20%<br />

1-5 6 - 10 11 - 20 21 - 40<br />

A. fabae<br />

Fig. 2. Mean distribution <strong>of</strong> ten Coccinella septempunctata females in a semi-field<br />

bioassay-arena with aphid infested and non-infested patches.<br />

To a certain extend <strong>the</strong> described cycle in occurrence <strong>of</strong> aphids at different crops was also<br />

observable in C. septempunctata but not in P. quatuordecimpunctata. At <strong>the</strong> beginning <strong>of</strong> <strong>the</strong><br />

observation period adults <strong>of</strong> C. septempunctata were found only at forest edges and in hedges.<br />

At <strong>the</strong> end <strong>of</strong> April and <strong>the</strong> beginning <strong>of</strong> May <strong>the</strong> ladybirds were most frequently found in<br />

fallow lands or in early cereal crops, e.g. winter rye and winter barley. At <strong>the</strong> same time <strong>the</strong><br />

first aphid colonies appeared <strong>the</strong>re. In <strong>the</strong> second half <strong>of</strong> May ladybird distribution changed<br />

remarkably. Despite <strong>of</strong> an increased aphid density in those early crops adults <strong>of</strong><br />

C. septempunctata disappeared and were subsequently very frequently found in oats and<br />

spring wheat. At this time, only few initial aphid colonies were present in spring cereal fields.<br />

In <strong>the</strong> first half <strong>of</strong> June <strong>the</strong> ladybird adults preferred o<strong>the</strong>r crops, e.g. peas and winter wheat.<br />

Now aphids occurred in higher numbers in nearly all crops but most attractively to <strong>the</strong> adults<br />

<strong>of</strong> C. septempunctata were fields with only few conspecific larvae. An example <strong>of</strong> such an<br />

aggregation <strong>of</strong> ladybird adults in a winter barley field, which was previously not used as<br />

breeding site and was <strong>the</strong>refore occupied only by few ladybird larvae, is shown in Figure 3.


54<br />

In contrast, adults <strong>of</strong> P. quatuordecimpunctata preferred winter cereals and to a lesser<br />

extend hedges during <strong>the</strong> whole investigation period. Remarkable changes in habitat<br />

preferences as described for C. septempunctata adults were not observed in P. quatuordecimpunctata.<br />

Berlin-Staaken<br />

June 02 1998<br />

winter rye, winter barley<br />

spring wheat, oats<br />

potato<br />

1 - 50<br />

51 - 100<br />

101 - 500<br />

501 - 1000<br />

1001 - 5000<br />

> 5000<br />

1<br />

2 - 5<br />

6 - 15<br />

16 - 50<br />

> 50<br />

hedge<br />

fallow<br />

grassland<br />

1<br />

2 - 5<br />

6 - 15<br />

16 - 50<br />

forest<br />

village<br />

maize<br />

> 50<br />

Individuals / 250 sweeps<br />

Fig. 3. Sweep net catches <strong>of</strong> Coccinella septempunctata and aphids at Berlin-Staaken


55<br />

In order to analyse <strong>the</strong> relation between ladybird distribution and aphid density data we<br />

used correlation coefficients. Density data <strong>of</strong> C. septempunctata adults (X 1 ) and aphids (X 2 )<br />

from all field sampling points were chosen and correlation coefficients were calculated for<br />

each sampling date (Fig. 4). Considering all four investigated seasons, only in 10 out <strong>of</strong> 27<br />

cases <strong>the</strong> occurrence <strong>of</strong> C. septempunctata adults was positively related to <strong>the</strong> aphid density.<br />

But, as shown in Figure 3, a significant relation occurred at certain periods within <strong>the</strong><br />

investigated time interval, namely from late April to <strong>the</strong> beginning <strong>of</strong> May and in <strong>the</strong> first half<br />

<strong>of</strong> June. Because <strong>of</strong> that periodically existent aphid related and non-related occurrence <strong>of</strong> C.<br />

septempunctata adults, which corresponded with <strong>the</strong> above described habitat/crop preferences<br />

<strong>of</strong> this ladybird species we defined five distribution phases (Tab. 1).<br />

In P. quatuordecimpunctata adult density was generally not correlated with aphid density<br />

in all investigations.<br />

Location<br />

Year<br />

April May June July<br />

BS 1998 0.17 0.81 0.23 0.11 0.20 0.14 0.74<br />

BS 1999 0.91 0.38 0.15 0.86 0.14 0.08<br />

NF 1998 0.24 0.49 0.09 0.02 0.40 0.09 0.18<br />

NF 1999 0.03 0.59 0.04 0.09 0.64 0.29 0.58<br />

Fig. 4<br />

Correlation coefficients <strong>of</strong> Coccinella septempunctata adult density related to aphid<br />

density data (data arranged according to <strong>the</strong> time scale, significant values in bold<br />

letters, P < 0.05).<br />

Tab. 1 Observed phases <strong>of</strong> distribution <strong>of</strong> Coccinella septempunctata adults within an<br />

agricultural area at two localities in 1998 - 1999<br />

Defined<br />

Phase<br />

Date<br />

Preferred<br />

crops/habitats<br />

I April forest edges,<br />

hedges<br />

II late April- winter rye,<br />

early May winter barley,<br />

III<br />

IV<br />

V<br />

second half<br />

<strong>of</strong> May<br />

first half<br />

<strong>of</strong> June<br />

end <strong>of</strong> June<br />

to July<br />

Observed distribution <strong>of</strong><br />

Coccinella septempunctata adults<br />

ladybird adult activity is restricted to<br />

<strong>the</strong> hibernacula<br />

ladybirds are present in early crops<br />

accompanied with initial aphid<br />

fallow land colonies<br />

oats, ladybird adults preferred spring<br />

spring wheat cereals with sparse vegetation and<br />

only initial aphid colonies<br />

winter wheat, most ladybird adults are found in<br />

winter barley, crops infested with aphids but<br />

pea<br />

previously not used as breeding site,<br />

i.e. only few conspecific larvae<br />

spring cereals disappearance <strong>of</strong> overwintered<br />

adults due to <strong>the</strong> end <strong>of</strong> <strong>the</strong>ir life<br />

span and emergence <strong>of</strong> <strong>the</strong> new<br />

adult generation<br />

Ladybird<br />

density / aphid<br />

density<br />

--<br />

related<br />

not related<br />

related<br />

not related


56<br />

Discussion<br />

Laboratory experiments<br />

Despite <strong>the</strong> fact that we used three different experimental designs to investigate searching<br />

behaviour <strong>of</strong> ladybird adults, we were not able to find some clear indications that<br />

aphidophagous ladybirds respond to olfactory cues <strong>of</strong> <strong>the</strong>ir essential prey. These results are<br />

contradictory to those findings <strong>of</strong> Sengonca and Liu (1994) who investigated<br />

C. septempunctata adults and larvae in an airflow olfactometer and found significant reactions<br />

to aphid odours. In contrast, Nakamuta (1991) found that <strong>the</strong> aphids alarm pheromone was<br />

not an orientation cue for C. septempunctata bruckii (Muls.), whereas Obata (1997) observed<br />

that adults <strong>of</strong> Harmonia axyridis (Pallas) were attracted to bags containing aphid infested<br />

leaves and it was suggested that <strong>the</strong>y use olfactory cues to detect prey.<br />

Our results suggest that ladybird adults search randomly. They probably can recognise<br />

olfactory cues over very short distances or even by physical contact with <strong>the</strong>ir maxillary palps<br />

or o<strong>the</strong>r sensory organs (Ferran & Dixon, 1993; Hodek & Honek, 1996).<br />

Field work<br />

Field investigations at two different agricultural locations indicate that C. septempunctata<br />

adults are very mobile predators. Their distribution within <strong>the</strong> agricultural area completely<br />

changed several times during <strong>the</strong> investigation period. The observed changes in <strong>the</strong>ir<br />

preference <strong>of</strong> certain crop and non-crop habitats were only sometimes explicable with<br />

differences in aphid occurrence. Occurrence <strong>of</strong> C. septempunctata adults was only<br />

periodically related to aphid density. It was observed that <strong>the</strong> ladybird adults prefer cereals<br />

and o<strong>the</strong>r crops with only initial aphid colonies at a relatively early stage <strong>of</strong> plant<br />

development. At higher aphid densities C. septempunctata adults disappeared from <strong>the</strong>se<br />

fields and at <strong>the</strong> end <strong>of</strong> <strong>the</strong> season <strong>the</strong>y aggregated in fields with low densities <strong>of</strong> conspecific<br />

larvae. These observations could confirm <strong>the</strong> optimal foraging <strong>the</strong>ory (Hemptinne et al.,<br />

1992; Hemptinne et al., 1995). According to this <strong>the</strong>ory, <strong>the</strong>re exists only a short period at <strong>the</strong><br />

beginning <strong>of</strong> aphid population development which is suitable for ladybird egg laying. Because<br />

<strong>of</strong> <strong>the</strong>ir relatively long larval developmental period, ladybirds should lay <strong>the</strong>ir eggs in<br />

younger aphid colonies to maximise survival <strong>of</strong> <strong>the</strong>ir larvae. Moreover, ladybird adults should<br />

emigrate after egg deposition to avoid competition with <strong>the</strong>ir own <strong>of</strong>fspring. A behaviour,<br />

confirming <strong>the</strong> optimal foraging <strong>the</strong>ory, was found in a number <strong>of</strong> investigated field crops and<br />

an example is given in Figure 5. Adults <strong>of</strong> C. septempunctata migrated into <strong>the</strong> oat field at an<br />

early stage <strong>of</strong> aphid population build up. With increasing aphid density <strong>the</strong> number <strong>of</strong><br />

ladybird larvae increased, but adult density decreased. Therefore a negative response <strong>of</strong><br />

ladybird adults to aphid density existed.<br />

Propylea quatuordecimpunctata, <strong>the</strong> o<strong>the</strong>r very common aphidophagous ladybird<br />

species, differed in its behaviour from C. septempunctata. Adult density <strong>of</strong> P.<br />

quatuordecimpunctata was generally not related to aphid density. Propylea<br />

quatuordecimpunctata adults preferred winter cereals and to a certain degree hedges during<br />

<strong>the</strong> whole investigation period. Perhaps P. quatuordecimpunctata adults are not as mobile as<br />

C. septempunctata adults. Never<strong>the</strong>less, <strong>the</strong>re is ano<strong>the</strong>r striking difference between <strong>the</strong>se two<br />

ladybird species: Whereas C. septempunctata like to bask in <strong>the</strong> sun on top <strong>of</strong> <strong>the</strong> vegetation<br />

which makes sweep netting very efficiently, P. quatuordecimpunctata occurred more hidden<br />

in lower strata <strong>of</strong> <strong>the</strong> vegetation.


57<br />

Aphid density<br />

HEMPTINNE et al. (1992)<br />

egg laying<br />

larval development<br />

<br />

<br />

developmental period<br />

Aphid density<br />

[Individuals/250 sweeps]<br />

2000<br />

oats (Nor<strong>the</strong>rn Flaeming 1998)<br />

Ladybird density<br />

79 163<br />

20<br />

1500<br />

1000<br />

aphids<br />

C. 7punctata<br />

larvae<br />

overwintered adults<br />

newly emerged adults<br />

10<br />

500<br />

0<br />

07.05. 18.05. 27.05. 05.06. 15.06. 26.06.<br />

0<br />

Fig. 5 Occurrence <strong>of</strong> Coccinella septempunctata and aphids in oats at location NF 1998<br />

compared with <strong>the</strong> optimal foraging <strong>the</strong>ory (Hemptinne et al., 1992).<br />

Acknowledgements<br />

We would like to acknowledge <strong>the</strong> “Deutsche Forschungsgemeinschaft” (DFG) for funding<br />

<strong>the</strong> essential part <strong>of</strong> all investigations and <strong>the</strong> Federal Biological Research Centre for<br />

Agriculture and Forestry for facilitating large parts <strong>of</strong> <strong>the</strong> research. We are very grateful to<br />

Pr<strong>of</strong>. M. Hilker for assistance regarding <strong>the</strong> olfactometer experiments and <strong>the</strong> possibility to<br />

use experimental equipment at her Institute for Applied Zoology/Animal Ecology, Freie<br />

Universität Berlin. We also thank Pr<strong>of</strong>. J.-L. Hemptinne for his helpful comments and<br />

suggestions to <strong>the</strong> field investigations.


58<br />

References<br />

Ferran, A. & Dixon, A.F.G., 1993: Foraging behaviour <strong>of</strong> ladybird larvae (Coleoptera:<br />

Coccinellidae). Eur. J. Entomol. 90: 383-402.<br />

Frazer, B.D. & Gilbert, N., 1976: Coccinellids and aphids: A quantitative study <strong>of</strong> <strong>the</strong> impact<br />

<strong>of</strong> adult ladybirds (Col., Coccinellidae) preying on field populations <strong>of</strong> pea aphids (Hom.,<br />

Aphididae). J. Entomol. Soc. Br. Columbia 73: 33-56.<br />

Freier, B., Möwes, M. & Triltsch, H., 1998: Beneficial thresholds for Coccinella 7-punctata<br />

L. (Col., Coccinellidae) as a predator <strong>of</strong> cereal aphids in winter wheat – results <strong>of</strong><br />

population investigations and computer simulations. J. Appl. Entomol. 122: 213-217.<br />

Gutierrez, A.P., Baumgartner, J.U. & Summers, C.G., 1984: Multitrophic models <strong>of</strong> predatorprey<br />

energetics. Can. Entomologist 116: 923-963.<br />

Hemptinne, J.-L., Dixon, A.F.G. & C<strong>of</strong>fin, J., 1992: Attack strategy <strong>of</strong> ladybird beetles<br />

(Coccinellidae): factors shaping <strong>the</strong>ir numerical response. Oecologia 90: 238-245.<br />

Hemptinne, J.-L., Doucet, J.-L. & Gaspar, C., 1994: How do ladybirds and syrphids respond<br />

to aphids in <strong>the</strong> field? <strong>IOBC</strong>/wprs Bull. 17(4): 101-111.<br />

Hemptinne, J.-L., Doumbia, M. & Gaspar, C., 1995: The reproductive strategy <strong>of</strong> predators is<br />

a major constraint to <strong>the</strong> implementation <strong>of</strong> biological control in <strong>the</strong> field. Med. Fac.<br />

Landb. Toegep. Biol. Wetensch. Univ. Gent 60: 735-741.<br />

Hodek, I., 1970: Coccinellids and modern pest management. Bioscience 20: 543-552.<br />

Hodek, I. & Honek, A., 1996: Ecology <strong>of</strong> coccinellidae. Kluwer Academic Publ., Dordrecht:<br />

464 pp.<br />

Honek, A., 1985: Habitat preferences <strong>of</strong> aphidophagous coccinellids (Coleoptera). Entomophaga<br />

30: 253-264.<br />

Nakamuta, K., 1984a: Visual orientation <strong>of</strong> a ladybeetle, Coccinella septempunctata L. (Col.,<br />

Coccinellidae), toward its prey. Appl. Entomol. Zool. 19: 82-86.<br />

Nakamuta, K., 1984b: Aphid body fluid stimulates feeding <strong>of</strong> a predatory ladybeetle<br />

Coccinella septempunctata L. (Col., Coccinellidae). Appl. Entomol. Zool. 19: 123-125.<br />

Nakamuta, K., 1991: Aphid alarm pheromone component (E)-Beta-farnesene and local search<br />

by a predatory lady beetle Coccinella septempunctata bruckii Mulsant (Col., Coccinellidae).<br />

Appl. Entomol. Zool. 26: 1-17.<br />

Obata, S., 1997: The influence <strong>of</strong> aphids on <strong>the</strong> behaviour <strong>of</strong> adults <strong>of</strong> <strong>the</strong> ladybird beetle,<br />

Harmonia axyridis (Col.: Coccinellidae). Entomophaga 42: 103-106.<br />

Poehling, H.-M., 1988: Zum Auftreten von Syrphiden- und Coccinellidenlarven in Winterweizen<br />

von 1984-1987 in Relation zur Abundanz von Getreideblattläusen. Mitt. Dtsch.<br />

Ges. allg. ang. Entomol. 6: 248-254.<br />

Ru<strong>the</strong>r, J. & Thiemann, K., 1997: Response <strong>of</strong> <strong>the</strong> pollen beetle Melige<strong>the</strong>s aeneus to<br />

volatiles emitted by intact plants and conspecifics. Entomol. Exp. Appl. 84: 183-188.<br />

Sengonca, C. & Liu, B., 1994: Response <strong>of</strong> <strong>the</strong> different instar predator, Coccinella septempunctata<br />

L. (Col., Coccinellidae), to <strong>the</strong> kairomones produced by <strong>the</strong> prey and non-prey<br />

as well as <strong>the</strong> predator itself. Z. Pflanzenkrankh. Pflanzenschutz 101: 173-177.<br />

Sunderland, K.D., Chambers, R.J., Stacey, D.L. & Wyatt, I.J., 1983: Predators prevent aphid<br />

outbreaks. AFRC News 1983: 12.<br />

Triltsch, H. & Freier, B., 1998: Investigations on differences between ladybird populations<br />

(Coleoptera, Coccinellidae) at three localities. <strong>IOBC</strong>/wprs Bull. 21(8): 113-124.<br />

Wellings, P.W. & Dixon, A.F.G., 1987: The role <strong>of</strong> wea<strong>the</strong>r and natural enemies in<br />

determining aphid outbreaks. In: Barbosa, P. & Schultz, J.C. (eds.): Insect Outbreaks.<br />

Academic Press, London: 313-346.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 59 - 64<br />

Computer simulations<br />

on <strong>the</strong> efficiency <strong>of</strong> cereal aphid predators in winter wheat<br />

Uwe Gosselke 1 , Dietmar Roßberg 2 , Holger Triltsch 3 and Bernd Freier 3<br />

1 Humboldt-University, Institute for Phytomedicine/Applied Entomology, Berlin, Germany<br />

2 BBA, Institute for Technology Assessment in Plant Protection, Kleinmachnow, Germany<br />

3 BBA, Institute for Integrated Plant Protection, Stahnsdorfer Damm 81, D-14532<br />

Kleinmachnow<br />

Summary<br />

Efficiency <strong>of</strong> predators in <strong>the</strong> tritrophic system winter wheat- cereal aphids – antagonists was<br />

investigated with computer simulations in combination with real field data. In scenario runs with <strong>the</strong><br />

model GTLAUS99 real conditions, based on field data <strong>of</strong> population dynamics <strong>of</strong> aphids and <strong>the</strong>ir<br />

antagonists were simulated. Simulated data did not differ from field data by more than 20%. The<br />

corresponding model run without any predators was calculated afterwards, and <strong>the</strong> difference in aphid<br />

infestation was interpreted as a predator effect. Our study indicates that without antagonists aphid<br />

infestation would cause economically important yield loss in about 2/3 <strong>of</strong> all cases. The economic<br />

value per single predator unit seemed to be small, but <strong>the</strong> entire predator community was responsible<br />

for a notable benefit up to 100 DM per ha.<br />

Key words: computer simulations, aphid predators, cereal aphids, winter wheat, predator efficiency,<br />

economic predator effect<br />

Introduction<br />

Within <strong>the</strong> concept <strong>of</strong> integrated plant protection in arable farming <strong>the</strong> consideration <strong>of</strong><br />

ecological self-control mechanisms like natural enemies is an important goal. Therefore,<br />

several attempts were made to get a realistic picture about <strong>the</strong> effect <strong>of</strong> beneficial arthropods.<br />

In many cases <strong>the</strong> tritrophic interactions cereals – aphids – predators was used as a model<br />

system. A large number <strong>of</strong> feeding experiments with predatory arthropods were performed,<br />

providing data about <strong>the</strong>ir feeding capacity under laboratory conditions (e.g. Sundby, 1966;<br />

Hämäläinen et al., 1975; Chiverton, 1988; Winder et al., 1994). Because <strong>of</strong> <strong>the</strong> well known<br />

difficulties <strong>of</strong> transferring such data to complex natural conditions, many investigations<br />

concentrate on field work. Certain predator-prey constellations were studied in experiments<br />

with field cages (e.g. Rautapää, 1975; Ghanim, 1981; Wetzel et al., 1981). In o<strong>the</strong>r<br />

experiments field barriers were introduced into <strong>the</strong> field to study population dynamics <strong>of</strong><br />

aphids with and without <strong>the</strong> impact <strong>of</strong> predatory ground dwelling arthropods (Chiverton,<br />

1987). Never<strong>the</strong>less, in many field investigations data obtained from detailed field counts<br />

were used for different calculations <strong>of</strong> <strong>the</strong> efficiency <strong>of</strong> natural enemies (e.g. Wratten, 1987;<br />

Freier et al., 1999). Because <strong>of</strong> mainly methodical problems investigating <strong>the</strong> complex nature<br />

<strong>of</strong> predator-prey interactions under natural conditions such an approach is not likely to<br />

provide quantitative data on <strong>the</strong> efficiency <strong>of</strong> natural enemies. Hence, economic calculations<br />

<strong>of</strong> <strong>the</strong> effect <strong>of</strong> predators are still rare (Sterling et al., 1992).<br />

In recent years simulation models became useful tools for studies on <strong>the</strong> population<br />

dynamics <strong>of</strong> aphids and antagonists, <strong>the</strong>ir trophic interactions and dependency on driving<br />

forces (Freier et al., 1996; Skirvin et al., 1997). These models could <strong>of</strong>fer <strong>the</strong> chance to<br />

investigate agro-ecosystems in a more complex way.<br />

59


60<br />

The aim <strong>of</strong> <strong>the</strong> present study was to quantify <strong>the</strong> efficiency <strong>of</strong> <strong>the</strong> most important cereal<br />

aphid predators within <strong>the</strong> tritrophic interactions <strong>of</strong> winter wheat – aphids – antagonists. A<br />

combination <strong>of</strong> detailed field counts with computer simulations was used and <strong>the</strong><br />

quantification should <strong>the</strong>n allow to calculate <strong>the</strong> economical benefit <strong>of</strong> <strong>the</strong> antagonists.<br />

Material and methods<br />

We used <strong>the</strong> following methodological approach in this study: Field data on aphid and<br />

antagonist densities were utilised as database for computer simulations. Model runs started<br />

with initial values observed in <strong>the</strong> field. Simulated aphid density was <strong>the</strong>n compared with<br />

field data and <strong>the</strong> run was termed “successful” if simulated values differed from field values<br />

by less <strong>the</strong>n 20%. Then <strong>the</strong> corresponding model run without any predators was calculated,<br />

and <strong>the</strong> difference in aphid infestation was interpreted as predator effect. Aphid infestation<br />

was measured as number <strong>of</strong> individuals per m², and aphid days per tiller (aphid index). The<br />

entire predator community was measured by calculating predator units, which allows to add<br />

<strong>the</strong> different predator fractions according to <strong>the</strong>ir potential food uptake at 20°C (Freier et al.,<br />

1997, 1998).<br />

Simulation model<br />

In our study we used <strong>the</strong> model GTLAUS 99 which is a deterministic, discrete simulation<br />

model written in Borland-Pascal 7.0. GTLAUS 99 contains submodels for winter wheat,<br />

cereal aphids (Sitobion avenae, Rhopalosiphum padi, Metopolophium dirhodum), and aphid<br />

predators (Coccinella septempunctata, Propylea quatuordecimpunctata, syrphids). The<br />

impact <strong>of</strong> o<strong>the</strong>r aphid antagonists, i.e. chrysopids, carabids, staphylinids, spiders, parasitoids<br />

and entomopathogenic fungi, was considered using special regression functions. GTLAUS<br />

was previously validated with field data and already used in different investigations, e.g.<br />

about climate change and predator-prey interaction (Freier & Triltsch, 1996; Triltsch &<br />

Roßberg, 1997) or beneficial thresholds (Freier et al., 1998). A detailed description <strong>of</strong> <strong>the</strong><br />

current model is provided in Gosselke et al. (2001).<br />

The GTLAUS model has a compartment structure and each compartment contains<br />

different age classes. Each model run begins with a range <strong>of</strong> starting values <strong>of</strong> density,<br />

structure and immigration type <strong>of</strong> <strong>the</strong> aphid population and <strong>the</strong>ir antagonists at a specific date<br />

during wheat flowering. Then daily density values <strong>of</strong> cereal aphids and antagonists were<br />

calculated by <strong>the</strong> model.<br />

Field counts<br />

Investigations were carried out in unsprayed winter wheat fields at two different localities,<br />

Flaeming (F) and Magdeburger Boerde (M) from 1993 to 1999. During eight weeks field<br />

counts were performed weekly between wheat growth stages (BBCH) 49 and 87 (Meier,<br />

1997). Densities <strong>of</strong> aphids and antagonists were recorded at 2x5 counting points at distances<br />

<strong>of</strong> 20, 40, 60, 80, and 100 m, respectively, from one field margin. At every point wheat tillers<br />

were examined for 3 m along a row.<br />

Results<br />

Population dynamics <strong>of</strong> aphids and <strong>the</strong>ir antagonists were simulated, using initial density<br />

values according to field data. In figure 1 a comparison between simulated aphid infestation<br />

and aphid density data observed during field counts is shown. In all 14 cases, e.g.


61<br />

Magdeburger Boerde and Flaeming 1993 to 1999, simulated aphid data correlate with field<br />

data. When running aphid index data, model runs and field data differed by less than 20%,<br />

e.g. 9.2% (S.D. 6.08) at Magdeburger Boerde and 9.2% (S.D. 5.78) at Flaeming.<br />

2500 Aphids/m² Aphid index during simulation<br />

Flaeming 1996<br />

(Aphid days/tiller)<br />

2000<br />

Field count 70.2<br />

Simulation 83.4<br />

∗<br />

1500<br />

∗<br />

1000<br />

500<br />

∗<br />

∗<br />

∗ ∗<br />

∗<br />

0 ∗<br />

∃<br />

11. 6. 18. 6. 25. 6. 2. 7. 9. 7. 16. 7. 23. 7. 30. 7. 6. 8. 13. 8.<br />

Fig. 1. Example <strong>of</strong> a computer simulation (curve) in comparison to field data (dots).<br />

800<br />

Aphid days/tiller<br />

600<br />

400<br />

200<br />

0<br />

93 94 95 96 97 98 99 93 94 95 96 97 98 99<br />

Magdeburger Boerde<br />

Flaeming<br />

Fig. 2. Results <strong>of</strong> computer simulations with (dark) and without (light) cereal aphid<br />

predators.


62<br />

According to our methodological concept <strong>the</strong> next step was to exclude aphid antagonists<br />

impact by simulating all cases without predators, i.e. without coccinellids, syrphids,<br />

chrysopids, and polyphagous predators. An overview <strong>of</strong> that is given in figure 2. The<br />

consequence <strong>of</strong> excluding all antagonists was an increased cereal aphid infestation, which was<br />

assumed to be <strong>the</strong> impact <strong>of</strong> <strong>the</strong> predators. The average aphid index <strong>of</strong> <strong>the</strong> seven investigated<br />

seasons increased at <strong>the</strong> Flaeming locality by 262.3 aphid days per tiller (S.D.: 359.44). At <strong>the</strong><br />

second site (Magdeburger Boerde) considerably lower values were recorded (97.0, S.D.:<br />

31.61). Remarkable differences between <strong>the</strong> different study years as well as between <strong>the</strong> two<br />

localities were observed. Values ranged from 36 (1994) and 976 aphid days per tiller (1998)<br />

at Flaeming to 58 (1996) and 139 (1998) aphid days per tiller at Magdeburger Boerde.<br />

Assuming an aphid damage threshold level <strong>of</strong> 150 aphid days per tiller, without predator<br />

impact cereal aphid infestation would have caused damage <strong>of</strong> economical importance in 9 out<br />

<strong>of</strong> 14 investigated cases (Flaeming: 3 and Magdeburger Boerde: 6). Efficacy <strong>of</strong> <strong>the</strong><br />

antagonists can be measured in terms <strong>of</strong> prevented increase in aphid infestation. Thereby a<br />

certain aphid infestation above 150 aphid days per tiller is related to a measurable loss in<br />

grain yield (in dt per ha). The economical predator effect can <strong>the</strong>n be calculated (Tab. 1). At a<br />

wheat price level <strong>of</strong> 20 DM per dt average predator effect ranged from 2.99 to 109.25 DM per<br />

ha (mean: 28.16, S.D.: 33.65). The corresponding economical value per single predator unit<br />

was in <strong>the</strong> range between 8 x 10 -5 and 361 x 10 -5 DM (mean: 73 x 10 -5 , S.D.: 115 x 10 -5 ).<br />

Tab. 1. Calculation <strong>of</strong> <strong>the</strong> economical benefit <strong>of</strong> cereal aphid predators in winter wheat based<br />

on field data and computer simulations.<br />

Location<br />

Potential Predator Economic benefit<br />

yield loss occurrence<br />

(dt/ha) (PU/m²) (DM/ha) (10 -5 DM/PU)<br />

Flaeming 1995 0.30 10.63 6.90 6.5<br />

Flaeming 1996 1.63 3.66 37.49 102.4<br />

Flaeming 1998 4.75 3.03 109.25 360.6<br />

Magdeburg 1994 0.24 5.99 5.52 9.2<br />

Magdeburg 1995 1.74 3.55 40.02 112.7<br />

Magdeburg 1996 0.14 5.23 3.22 6.2<br />

Magdeburg 1997 0.13 3.53 2.99 8.5<br />

Magdeburg 1998 1.08 10.77 24.84 23.1<br />

Magdeburg 1999 1.01 6.40 23.23 36.3<br />

Mean (S.D.) 28.16 73.9<br />

(33.65) (115.1)<br />

Discussion<br />

Despite <strong>of</strong> a large number <strong>of</strong> investigations estimating <strong>the</strong> potential benefit <strong>of</strong> natural enemies<br />

on pest population dynamics, most <strong>of</strong> <strong>the</strong>se attempt were accompanied by more or less<br />

significant shortcomings. Data based on laboratory experiments or even on experiments using<br />

field cages and exclusion barriers were <strong>of</strong>ten criticised because <strong>of</strong> artificial conditions. On <strong>the</strong>


63<br />

o<strong>the</strong>r hand calculations on <strong>the</strong> efficiency <strong>of</strong> certain predators based on detailed field counts<br />

are not applicable without any assumptions. However, assumptions and simplifications are<br />

also necessary when using computer simulation models. The model GTLAUS, used in this<br />

study, dates back 15 years now, and has been several times revised, improved and validated<br />

with new field data. Though during its last validation with long term field data from four<br />

different locations with 35 investigated fields in two cases <strong>the</strong> model failed to generate a<br />

realistic simulation. The advantage <strong>of</strong> simulation models is <strong>the</strong> possibility to take<br />

systematically into account large volumes <strong>of</strong> published data as well as field data and to<br />

simulate many different situations within a short time (Freier et al., 1996).<br />

Exclusion <strong>of</strong> aphid predators resulted in a remarkable increase in aphid infestation. The<br />

aphid index was up to 17 times higher in simulations without predators. Our study indicates<br />

that without antagonists aphid infestation would cause economically important yield losses in<br />

about 2/3 <strong>of</strong> all cases. Although <strong>the</strong> economic value per single predator unit seems to be<br />

ra<strong>the</strong>r small, <strong>the</strong> entire predator community was responsible for a notable benefit <strong>of</strong> up to 100<br />

DM per ha. We found remarkable differences between different locations and years with<br />

regard to aphid infestation pattern and effects <strong>of</strong> predators. Aphid population dynamics do not<br />

only depend on <strong>the</strong> influence <strong>of</strong> antagonists, but also on a number <strong>of</strong> o<strong>the</strong>r environmental and<br />

aphid population inherent factors (Freier et al., 1999). Additionally, <strong>the</strong> effectiveness <strong>of</strong> aphid<br />

antagonists is very variable, e.g. depending strongly on temperature (Triltsch et al., 1996),<br />

predator-prey-ratio (Rautapää, 1975) and synchronisation between predator and prey<br />

(Hemptinne et al., 1995).<br />

Acknowledgement<br />

We would like to acknowledge <strong>the</strong> Federal Ministry for Agriculture and Forestry for partly<br />

funding <strong>the</strong> research.<br />

References<br />

Chiverton, P.A., 1987: Effects <strong>of</strong> exclusion barriers and inclusion trenches on polyphagous<br />

and aphid specific predators in spring barley. J. Appl Ent. 103: 193-203.<br />

Chiverton, P.A., 1988: Searching behaviour and cereal aphid consumption by Bembidion<br />

lampros and Pterostichus cupreus, in relation to temperature and prey density. Ent. Exp.<br />

Appl. 47: 137-182.<br />

Freier, B. & Triltsch, H., 1996: Climate chamber experiments and computer simulations on<br />

<strong>the</strong> influence <strong>of</strong> increasing temperature on wheat-aphid-predator interaction. Asp. Appl.<br />

Biol. 45: 293-298.<br />

Freier, B., Triltsch, H. & Roßberg, D., 1996: GTLAUS – A model <strong>of</strong> wheat-cereal aphidpredator<br />

interaction and its use in complex agroecological studies. Z. Pflanzenkrankh.<br />

Pflanzenschutz 103: 543-554.<br />

Freier, B., Triltsch, H., Möwes, M. & Rappaport, V., 1997: Der relative Wert von Prädatoren<br />

bei der natürlichen Kontrolle von Getreideblattläusen und die Verwendung von Prädatoreinheiten.<br />

Nachrichtenbl. Deut. Pflanzenschutzd. 49: 215-222.<br />

Freier, B., Möwes, M. & Triltsch, H., 1998: Beneficial threshold for Coccinella 7-punctata L.<br />

as predator <strong>of</strong> cereal aphids in winter wheat – results <strong>of</strong> population investigations and<br />

computer simulations. J. Appl. Ent. 121: 213-217.<br />

Freier, B., Triltsch, H. & Gosselke, U., 1999: Die Dimension der natürlichen Kontrolle von<br />

Getreideblattläusen durch Prädatoren. Gesunde Pflanzen 51: 65-71.


64<br />

Ghanim, A., 1981: Untersuchungen über den Einfluss von Parasiten und Prädatoren auf die<br />

Entstehung von Gradationen der Getreideblattläuse. PhD Thesis, Martin-Lu<strong>the</strong>r-<br />

Universität Halle, 136 pp.<br />

Gosselke, U., Roßberg, D., Freier, B. & Triltsch, H., 2001: GTLAUS99 – <strong>the</strong> latest version <strong>of</strong><br />

a model for simulation <strong>of</strong> wheat-aphid-predator interaction. Ecol. Modelling 99: in press.<br />

Hämäläinen, M., Markkula, M. & Raij, T., 1975: Fecundity and larval voracity <strong>of</strong> four<br />

ladybird species (Col., Coccinellidae). Ann. Ent. Fenn. 41: 124-127.<br />

Hemptinne, J.-L., Doumbia, M. & Gaspar, C., 1995: The reproductive strategy <strong>of</strong> predators in<br />

a major constraint to <strong>the</strong> implementation <strong>of</strong> biological control in <strong>the</strong> field. Med. Fac.<br />

Landb. Toeg. Biol. Wetensch. 60: 735-741.<br />

Meier, U., 1997: Growth stages <strong>of</strong> mono- and dicotyledonous plants. BBCH-Monograph.<br />

Berlin, Wien: Blackwell, 622 pp.<br />

Rautapää, J., 1975: Control <strong>of</strong> Rhopalosiphum padi (L.) (Hom., Aphididae) with Coccinella<br />

septempunctata L. (Col., Coccinellidae) in cages, and effect <strong>of</strong> late aphid infestation on<br />

barley yield. Ann. Agric. Fenniae 14: 231-239.<br />

Skirvin, D.J., Perry, J.N. & Harrington, R., 1997: A model describing <strong>the</strong> population<br />

dynamics <strong>of</strong> Sitobion avenae and Coccinella septempunctata. Ecol. Modelling 96: 29-39.<br />

Sundby, R.A., 1966: A comparative study <strong>of</strong> <strong>the</strong> efficiency <strong>of</strong> three predatory insects,<br />

Coccinella septempunctata L. (Col., Coccinellidae), Chrysopa carnea St. (Neuropt.,<br />

Chrysopidae) and Syrphus ribesii (Dipt., Syrphidae) at two different temperatures.<br />

Entomophaga 11: 395-404.<br />

Sterling, W.L., Dean, A. & Abd el Salam, N.M., 1992: Economic benefits <strong>of</strong> spiders<br />

(Araneae) and insects (Hemiptera: Miridae) predators <strong>of</strong> cotton leafhoppers. J. Econ. Ent.<br />

85: 52-57.<br />

Triltsch, H., Freier, B. & Roßberg, D., 1996: Temperatur – Schlüsselfaktor für Nützlingsleistungen<br />

im Winterweizen? Mitt. Biol. Bundesanstalt 321: 447.<br />

Triltsch, H. & Roßberg, D., 1997: Cereal aphid predation by <strong>the</strong> ladybird Coccinella septempunctata<br />

L. (Col.: Coccinellidae) – Including its simulation in <strong>the</strong> model GTLAUS. Acta<br />

Jutlandica 72: 259-270.<br />

Wetzel, T., Ghanim, A. & Freier, B., 1981: Zur Bedeutung von Prädatoren und Parasiten für<br />

die Überwachung und Bekämpfung von Blattläusen in Getreidebeständen. Nachrichtenbl.<br />

Pflanzenschutz DDR 35: 239-244.<br />

Winder, L.D., Hirst, J., Carter, N., Wratten, S. D. & Sopp, P.I., 1994: Estimating predation <strong>of</strong><br />

<strong>the</strong> grain aphid Sitobion avenae by polyphagous predators. J. Appl. Ecol. 31: 1-12.<br />

Wratten, S.D., 1987: The effectiveness <strong>of</strong> native natural enemies. In: Integrated Pest<br />

Management, Academic Press: 89-112.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 65 - 71<br />

Potentials and limitations <strong>of</strong> long-term field data to identify numerical and<br />

functional responses <strong>of</strong> predators to aphid density in wheat<br />

Freier 1 , B., Triltsch 1 , H. and U. Gosselke 2<br />

1<br />

Federal Biological Research Centre for Agriculture and Forestry, Institute <strong>of</strong> Integrated<br />

Plant Protection, D-14532 Kleinmachnow, Germany<br />

2<br />

Humboldt University, Institute <strong>of</strong> Horticultural Sciences, FG Phytomedicine/Applied<br />

Entomology, D-14195 Berlin, Germany<br />

Summary<br />

Long-term field studies on population dynamics <strong>of</strong> aphids and <strong>the</strong> community <strong>of</strong> predators have been<br />

performed in wheat fields at two extremely different sites since 1993 to investigate <strong>the</strong> dimension,<br />

stability and variance <strong>of</strong> natural control. To identify density feed backs <strong>of</strong> <strong>the</strong> predators, regression<br />

analyses <strong>of</strong> <strong>the</strong> 7-year data were performed and <strong>the</strong> equation with <strong>the</strong> highest significance (P


66<br />

BBCH 49 and 87 (growth stages according to Meier 1997) were recorded. To get a sufficient<br />

sample, each survey included 2 x 5 counting on an approximately 3 m drilling row within a 2<br />

ha patch <strong>of</strong> one side <strong>of</strong> <strong>the</strong> test field. The total area studied per sampling date was 4 m², that is<br />

2,400 tillers. The average densities <strong>of</strong> aphids and beneficials per m² were used for statistical<br />

analyses.<br />

The computer calculations included <strong>the</strong> analysis <strong>of</strong> <strong>the</strong> following data:<br />

– Data pairs <strong>of</strong> aphid and predator densities obtained on <strong>the</strong> same sampling date (both sites<br />

toge<strong>the</strong>r and separately); symbols in Table 1: FM, F and M.<br />

– Data pairs <strong>of</strong> aphid density observed at a given sampling date and predator density one<br />

week later (d i+1 ) as an indicator <strong>of</strong> delayed responses (both sites toge<strong>the</strong>r and separately);<br />

symbols in Table 1: FM (d i+1 ), F (d i+1 ) and M (d i+1 ).<br />

– Data pairs for seasonal average densities <strong>of</strong> aphids and predators to estimate fundamental<br />

responses (both sites toge<strong>the</strong>r and separately); symbols in Table 1: FM (season),<br />

F (season) and M (season).<br />

Fur<strong>the</strong>rmore, numerical interactions between fractions within <strong>the</strong> predator community<br />

were also investigated. Special analyses <strong>of</strong> functional responses were made, especially <strong>the</strong><br />

influence <strong>of</strong> <strong>the</strong> aphid infestation level on <strong>the</strong> larvae-egg abundance ratio for coccinellids,<br />

syrphids and chrysopids at a given date. The occurrence <strong>of</strong> individuals in <strong>the</strong> overall predator<br />

community were considered by calculating predator units (PU) derived from surplus feeding<br />

rates <strong>of</strong> each predator fraction at 20°C (Freier et al., 1998).<br />

Results and discussion<br />

Table 1 documents all results <strong>of</strong> <strong>the</strong> statistical analyses, starting with <strong>the</strong> reaction <strong>of</strong> <strong>the</strong> whole<br />

predator community and including aphid specific predators and all important fractions <strong>of</strong><br />

predators up to <strong>the</strong> chrysopid larvae.<br />

The results indicate varying close relationships between population densities <strong>of</strong> cereal<br />

aphids and predators, although <strong>the</strong> significance <strong>of</strong> tendencies was ra<strong>the</strong>r weak. However, prior<br />

to any evaluation some methodological problems need to be considered. Firstly, counts do not<br />

represent <strong>the</strong> real densities <strong>of</strong> <strong>the</strong> predators. For instance <strong>the</strong> abundance <strong>of</strong> polyphagous<br />

arthropods is nearly always underestimated (Möwes et al,. 1997). Moreover, o<strong>the</strong>r<br />

antagonists, such as parasitoids and entomopathogenic fungi, and o<strong>the</strong>r factors, such as <strong>the</strong><br />

wea<strong>the</strong>r and wheat plant acting as <strong>the</strong> aphid host, cause variable multiple effects and thus<br />

cause additional problems. When aphid and predator occurrence under field conditions<br />

exceeds a specific level, <strong>the</strong> infestation-reducing effects <strong>of</strong> <strong>the</strong> predators overlap and<br />

counteract with <strong>the</strong>ir numerical and functional responses. However, <strong>the</strong> present 7-year<br />

investigation on <strong>the</strong> occurrence <strong>of</strong> aphids and predators, defined as <strong>the</strong> number <strong>of</strong> individuals<br />

per m², document a general lack <strong>of</strong> high-density data for both, aphids and predators.<br />

Predator community, aphid specific predators<br />

The working hypo<strong>the</strong>sis was that rising aphid numbers lead to an increasing predator<br />

potential. This could be confirmed in our analyses. The responses <strong>of</strong> <strong>the</strong> overall predator<br />

community and aphid-specific predator potential to increasing levels <strong>of</strong> aphid infestation were<br />

significant, though ra<strong>the</strong>r weak. They also did not become stronger when delayed reactions<br />

were taken into consideration. The clearer tendencies were always observed at <strong>the</strong> high-input<br />

site in Magdeburg (M). The season averages did not exhibit any corresponding tendencies,<br />

that is, aphid years did not necessarily correspond to predator years or vice versa. By using<br />

predator units to summarise different predator fractions according to <strong>the</strong>ir aphid consumption,<br />

<strong>the</strong> significance <strong>of</strong> relationships was higher than a simple addition <strong>of</strong> individual numbers.


67<br />

Coccinellids<br />

Apart from Coccinella 7-punctata at Flaeming (F), no relationships or delayed responses were<br />

observed. This suggests that <strong>the</strong> adults <strong>of</strong> both C. 7-punctata and Propylaea 14-punctata<br />

colonised <strong>the</strong> winter wheat fields ra<strong>the</strong>r independently <strong>of</strong> <strong>the</strong> aphid infestation level.<br />

Only a weakly positive influence <strong>of</strong> aphid infestation on coccinellid egg occurrence was<br />

detected. Comparing both species, <strong>the</strong>se tendencies were more clearly pronounced in P. 14-<br />

puctata at Flaeming (F) than in Magdeburger Boerde (M). However, no delayed responses or<br />

seasonal effects were observed in ei<strong>the</strong>r species.<br />

Tab. 1 Results <strong>of</strong> regression analyses on numerical responses <strong>of</strong> predators to <strong>the</strong> infestation<br />

level <strong>of</strong> cereal aphids in winter wheat fields at Flaeming (F) and Magdeburger<br />

Boerde (M) from 1993-1999.<br />

Predators Regression r² n Significance<br />

Predators units<br />

FM degressive increase 0.2465 109 P < 0.05<br />

F degressive increase 0.1593 54 P < 0.05<br />

M degressive increase 0.2894 55 P < 0.05<br />

FM (d i + 1) degressive increase 0.3217 95 P < 0.05<br />

F (d i + 1) linear increase 0.1975 47 P < 0.05<br />

M (d i + 1) degressive increase 0.3578 48 P < 0.05<br />

FM (season), F (season), M (season)<br />

no<br />

Aphid-specific predators<br />

FM degressive increase 0.2404 109 P < 0.05<br />

F linear increase 0.1112 54 P < 0.05<br />

M degressive increase 0.3167 55 P < 0.05<br />

FM (d i + 1) degressive increase 0.4027 95 P < 0.05<br />

F (d i + 1) linear increase 0.2127 47 P < 0.05<br />

M (d i + 1) degressive increase 0.5100 48 P < 0.05<br />

FM (season), F (season), M (season)<br />

no<br />

Coccinellid adults<br />

FM, F, M<br />

no<br />

FM (d i + 1), F (d i + 1), M (d i + 1)<br />

no<br />

FM (season), F (season), M (season)<br />

no<br />

Coccinella 7-punctata adults/Propylaea 14-punctata adults<br />

FM<br />

no<br />

F degressive increase 0.1166 54 P < 0.05<br />

M (only C. 7-punctata) no<br />

FM (d i + 1), F (d i + 1), M (d i + 1)<br />

no<br />

FM (season), F (season), M (season)<br />

no<br />

Coccinellid eggs<br />

FM linear increase 0.0799 109 P < 0.05<br />

F linear increase 0.1810 54 P < 0.05<br />

M linear increase 0.1194 55 P < 0.05<br />

FM (d i + 1), F (d i + 1), M (d i + 1)<br />

no<br />

FM (season), F (season), M (season)<br />

no


68<br />

Predators Regression r² n Significance<br />

Coccinella 7-punctata eggs<br />

FM linear increase 0.0572 109 P < 0.05<br />

F linear increase 0.0911 54 P < 0.05<br />

M<br />

no<br />

FM (d i + 1), F (d i + 1), M (d i + 1)<br />

no<br />

FM (season), F (season), M (season)<br />

no<br />

Propylaea 14-punctata eggs<br />

FM<br />

no<br />

F linear increase 0.1536 54 P < 0.05<br />

M linear increase 0.1969 55 P < 0.05<br />

FM (d i + 1), F (d i + 1), M (d i + 1)<br />

no<br />

FM (season), F (season), M (season)<br />

no<br />

Coccinellid larvae<br />

FM, F, M<br />

no<br />

FM (d i + 1) degressive increase 0.1977 P < 0.05<br />

F (d i + 1) linear increase 0.0866 P < 0.05<br />

M (d i + 1) progressive increase 0.6591 P < 0.05<br />

FM (season)<br />

no<br />

F (season)<br />

no<br />

M (season) progressive increase 0.6467 7 P < 0.05<br />

Coccinella 7-punctata larvae<br />

FM, F, M<br />

no<br />

FM (d i + 1) progressive increase 0.4912 95 P < 0.05<br />

F (d i + 1) no<br />

M (d i + 1) progressive increase 0.5269 48 P < 0.05<br />

FM (season) progressive increase 0.3617 14 P < 0.05<br />

F (season)<br />

no<br />

M (season)<br />

no<br />

Propylaea 14-punctata larvae<br />

FM<br />

no<br />

F<br />

no<br />

M linear increase 0.1821 55 P < 0.05<br />

FM (d i + 1)<br />

no<br />

F (d i + 1) no<br />

M (d i + 1) degressive increase 0.3011 48 P < 0.05<br />

FM (season), F (season), M (season)<br />

no<br />

Syrphid eggs<br />

FM degressive increase 0.3302 109 P < 0.05<br />

F degressive increase 0.1800 54 P < 0.05<br />

M degressive increase 0.4806 55 P < 0.05<br />

FM (d i + 1), F (d i + 1), M (d i + 1)<br />

no<br />

FM (season), F (season), M (season)<br />

no


69<br />

Predators Regression r² n Significance<br />

Syrphid larvae<br />

FM progressive increase 0.3690 109 P < 0.05<br />

F linear increase 0.2020 54 P < 0.05<br />

M progressive increase 0,4147 55 P < 0.05<br />

FM (d i + 1) degressive increase 0.4489 95 P < 0.05<br />

F (d i + 1) linear increase 0.4809 47 P < 0.05<br />

M (d i + 1) degressive increase 0.4346 48 P < 0.05<br />

FM (season) progressive increase 0.5051 14 P < 0.05<br />

F (season)<br />

no<br />

M (season) progressive increase 0.6365 7 P < 0.05<br />

Chrysopid eggs<br />

FM linear increase 0.1678 109 P < 0.05<br />

F<br />

no<br />

M linear increase 0.1494 55 P < 0.05<br />

FM (d i + 1) linear increase 0.1497 95 P < 0.05<br />

F (d i + 1) no<br />

M (d i + 1) linear increase 0.15137 48 P < 0.05<br />

FM (season), F (season), M (season)<br />

no<br />

Chrysopid larvae<br />

FM linear increase 0.1060 109 P < 0.05<br />

F linear increase 0.1133 54 P < 0.05<br />

M linear increase 0.0841 55 P < 0.05<br />

FM (d i + 1) degressive increase 0.2747 95 P < 0.05<br />

F (d i + 1) no<br />

M (d i + 1) degressive increase 0.2994 48 P < 0.05<br />

FM (season), F (season), M (season)<br />

no<br />

This is surprising because laboratory experiments suggested an increasing egg production<br />

<strong>of</strong>, e.g. C. 7-punctata feeding on Aphis gossypii in cotton (Xia et al., 1999). Our statistical<br />

results show that coccinellids, particularly C. 7-punctata, seem to immigrate well satiated and<br />

with a certain fecundity fundament from <strong>the</strong> surrounding areas, so that <strong>the</strong> females produce<br />

eggs more or less independently <strong>of</strong> aphid population density in wheat fields.<br />

For coccinellid larvae, <strong>the</strong> aphid infestation had no immediate influence. The exception<br />

was P. 14-punctata at <strong>the</strong> high-input site Magdeburger Boerde (M). At this location, <strong>the</strong><br />

P. 14-punctata larvae seemed to respond directly to an increasing aphid infestation. However,<br />

delayed responses <strong>of</strong> larvae to aphid occurrence one week before has been clearly established,<br />

particularly at <strong>the</strong> high-input site Magdeburger Boerde (M). Laboratory studies showed that<br />

prey density determines <strong>the</strong> survival rate <strong>of</strong> C. 7-punctata larvae (e.g. Xia et al,. 1999). Using<br />

<strong>the</strong> mean seasonal data, only weak density relationships were found. This means that years <strong>of</strong><br />

aphid infestation do not necessarily lead to higher numbers <strong>of</strong> coccinellid larvae.<br />

These results corroborate previous investigations by Hemptinne et al. (1994) and Triltsch<br />

et al. (2001), who suggest that both C. 7-punctata and P. 14-punctata do not show a clear and<br />

simple pattern <strong>of</strong> response to <strong>the</strong> presence <strong>of</strong> aphids.


70<br />

Syrphids<br />

The results in table 1 show a clear positive influence <strong>of</strong> aphid infestation on syrphid egg<br />

occurrence at both locations, particularly at <strong>the</strong> high-input site Magdeburger Boerde (M). The<br />

response in egg production was immediate, and delayed density responses <strong>of</strong> egg abundance<br />

were not observed. Surprisingly <strong>the</strong> data analyses did not reveal a relationship between aphids<br />

and syrphid eggs upon seasonal analysis. The mechanisms <strong>of</strong> synchronisation <strong>of</strong> aphid<br />

occurrence and syrphid egg laying has been investigated in different studies in winter wheat<br />

crops (Poehling & Borgemeister, 1989; Krause, 1997) and in bean fields (Hemptinne et al.,<br />

1994).<br />

Also, a clear density feedback on aphid infestation was observed in syrphid larvae at both<br />

locations. In contrast to <strong>the</strong> eggs, clear delayed effects and seasonal reactions were established<br />

in <strong>the</strong> case <strong>of</strong> syrphid larvae.<br />

Chrysopids<br />

The aphid infestation seemed to cause a certain immediate and delayed positive influence on<br />

chrysopid egg and larvae occurrence, particularly at <strong>the</strong> high-input site Magdeburger Boerde<br />

(M). However, when evaluated according to seasonal averages, no clear tendency was<br />

observed.<br />

The evidence <strong>of</strong> additional relationships between <strong>the</strong> occurrence <strong>of</strong> <strong>the</strong> specialised aphid<br />

predator coccinellids and syrphid larvae was an interesting finding. Despite intra-guild<br />

competition, both predator groups seem to pr<strong>of</strong>it by increasing aphid infestation at <strong>the</strong> same<br />

level.<br />

The ability to identify density feedbacks in long-term studies is quite limited. We know<br />

that counteractive infestation-reducing effects <strong>of</strong> <strong>the</strong> predators clearly overlap <strong>the</strong><br />

phenomenon <strong>of</strong> density responses. Also, <strong>the</strong> different level <strong>of</strong> additional parasitoid and fungal<br />

attacks in aphid populations can modify <strong>the</strong> results extremely.<br />

In contrast to numerical responses, <strong>the</strong> functional responses (change in attack rate <strong>of</strong> an<br />

individual) could not be identified directly from <strong>the</strong> field counts, but indirectly by <strong>the</strong><br />

assumption that increasing feeding rates lead to rising egg production (coccinellids) and<br />

increasing survival rates <strong>of</strong> larvae (coccinellids, syrphids, chrysopids). Such relationships<br />

were identified only in coccinellid larvae and syrphid eggs and larvae.<br />

The effect <strong>of</strong> aphid infestation level on <strong>the</strong> larvae-egg abundance ratio for coccinellids,<br />

syrphids and chrysopids has not yet been established, although it must be assumed that a high<br />

ratio reflects a good survival and is <strong>the</strong>refore an indicator <strong>of</strong> increasing food uptake.<br />

Acknowledgements<br />

We are grateful to Mrs. B. Schlage for <strong>the</strong> joint performance <strong>of</strong> all field investigations,<br />

computer documentation, and statistical analysis <strong>of</strong> field data. Fur<strong>the</strong>rmore, we thank <strong>the</strong><br />

farmers J. Grabo (Pflügkuff, site Flaeming) and U. Hartmann (Ochtmersleben, site<br />

Magdeburger Boerde) who made <strong>the</strong>ir unsprayed wheat fields available for field counts<br />

throughout all <strong>the</strong> years <strong>of</strong> <strong>the</strong> study. Our special thanks to Mrs. S. Wandrey, who had <strong>the</strong><br />

kindness to edit <strong>the</strong> English manuscript with enduring patience.<br />

References<br />

Freier, B.; Möwes, M., Triltsch, H. & Rapperport, V., 1998: Predator units - an approach to<br />

evaluate coccinellids within <strong>the</strong> aphid predator community in winter wheat, <strong>IOBC</strong>/<strong>WPRS</strong><br />

Bull. 21 (8): 103-111.


Hassell, M.P., 1978: The dynamics <strong>of</strong> arthropod predator prey systems. Monographs in<br />

Population Biology 13, Princeton Univ. Press.<br />

Hemptinne, J.-L., Doucet, J.-L. & Gaspar, C., 1994: How do ladybirds and syrphids respond<br />

to aphids in <strong>the</strong> field? <strong>IOBC</strong>/<strong>WPRS</strong> Bull: 17(4): 101-111.<br />

Krause, U., 1997: Populationsdynamik und Überwinterung von Schwebfliegen (Diptera,<br />

Syrphidae) in zwei unterschiedlich strukturierten Agrarlandschaften Norddeutschlands.<br />

Agrarökologie 22: 1-150.<br />

Möwes, M., Freier, B., Kreuter, Th. & Triltsch, H., 1997: Halmzählung oder<br />

Parzellentotalernte - wie genau sind Prädatorbonituren im Winterweizen? Anz.<br />

Schädlingskunde Pflanzenschutz Umweltschutz 70: 121-126<br />

Poehling, H.M. & Borgemeister, C., 1989: Abundance <strong>of</strong> coccinellids and syrphids in relation<br />

to cereal aphid density in winter wheat fields in nor<strong>the</strong>rn Germany. <strong>IOBC</strong>/wprs Bull.<br />

12(1): 99-107.<br />

Solomon, M.E., 1949: The natural control <strong>of</strong> animal populations. J. Animal Ecology 18, 1-35.<br />

Triltsch, H., Hechenthaler, G., Gosselke, U. & Freier, B., 2001: How does a ladybird respond<br />

to aphids? <strong>IOBC</strong>/wprs Bull. 24(6): 49-58.<br />

Xia, J.Y, van der Werf, W. & Rabbinge, R., 1999: Temperature and prey density on<br />

bionomics <strong>of</strong> Coccinella septempunctata (Coleoptera: Coccinellidae) feeding on Aphis<br />

gossypii (Homoptera: Aphididae) on cotton. Environ. Ent. 28: 307-314.<br />

71


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 73 - 78<br />

Density <strong>of</strong> epigeal predators on maize plants untreated and treated with<br />

imidacloprid<br />

Xavier Pons and Ramon Albajes<br />

Universitat de Lleida, Centre UdL-IRTA, Rovira Roure 177, 25198 Lleida, Spain<br />

Summary<br />

In 1997 and 1998, densities <strong>of</strong> aphids and predators on maize plants were monitored weekly in<br />

commercial plots treated or not treated with imidacloprid from plant emergence to flowering. The<br />

insecticide significantly reduced <strong>the</strong> aphid density during <strong>the</strong> whole sampling period. Orius spp.,<br />

Nabis provencallis, Coccinella septempunctata, Adonia variegata, Demetrias atricapillus, Tachyporus<br />

spp., earwigs and spiders were <strong>the</strong> main predators found. Densities <strong>of</strong> Syrphidae and Chrysopidae<br />

were very low. Numbers <strong>of</strong> individuals per plant <strong>of</strong> Orius spp. were significantly higher in untreated<br />

plots than in plots treated with imidacloprid in both years, whereas density <strong>of</strong> arachnida and D.<br />

atricapillus were lower in treated plots only in 1998.<br />

Key words: aphids, polyphagous predators, maize, imidacloprid<br />

Introduction<br />

Maize is <strong>the</strong> main summer cereal grown in Catalonia, and it covers nearly 45,000 ha, more<br />

than 50% <strong>of</strong> which is in <strong>the</strong> Lleida Basin. Pests attacking maize crops in <strong>the</strong> region can be<br />

divided into: soil pests (wireworms and cutworms), corn borers (Mediterranean Sesamia and<br />

European corn borers), sap sucking insects (aphids and leafhoppers), leaf and grain feeders<br />

and mites (Piqué et al., 1998).<br />

Aphids have a high damaging potential in Catalonia, especially in <strong>the</strong> first half <strong>of</strong> <strong>the</strong><br />

crop season when <strong>the</strong>y cause direct damage and transmit viruses. The most frequent and<br />

abundant species are Rhopalosiphum padi (L.), Sitobion avenae (Fabricius) and<br />

Metopolophium dirhodum (Walker) (Pons et al., 1994).<br />

Methylcarbamates, applied at sowing, have been one <strong>of</strong> <strong>the</strong> most common type <strong>of</strong><br />

insecticides used to prevent aphid damages. However, <strong>the</strong>y are being replaced by<br />

imidacloprid applied as a seed treatment.<br />

In 1997 we began a three-year research program to study <strong>the</strong> effects <strong>of</strong> imidacloprid on<br />

aphid populations and on natural enemies present in maize ecosystems. Here we present <strong>the</strong><br />

results <strong>of</strong> 1997 and 1998 on aphid populations and epigeal predators.<br />

Materials and methods<br />

The study was carried out on a commercial farm located 35 km to <strong>the</strong> west <strong>of</strong> Lleida. A<br />

complete randomised block design with four replications was used. Each block consisted <strong>of</strong><br />

two plots, one treated with imidacloprid and <strong>the</strong> o<strong>the</strong>r untreated. Seeds were treated with<br />

imidacloprid (Gaucho 35 FS ® , Bayer) at 490 g a.i. per 100 kg seed.<br />

Plot size varied from 0.25 to 1 ha. Maize was cultivated under conventional tillage and<br />

was sown on 9 and 13 May in 1997 1998, respectively. Two days after sowing, plots were<br />

sprayed with a mixture <strong>of</strong> 35 % alachlor + 25 % atrazine (Primdal ® , Agrodan) at 6 l/ha.<br />

73


74<br />

Plots were monitored weekly from maize emergence until flowering. Twenty-five plants<br />

per plot were visually sampled each sampling day. The species, number and morph <strong>of</strong> aphids,<br />

and species and number <strong>of</strong> predators on each plant were recorded.<br />

Data were analysed using <strong>the</strong> SAS statistics package (SAS Institute Inc., 1989). An<br />

analysis <strong>of</strong> variance was performed considering sampling dates, treatment, block and <strong>the</strong><br />

interaction sampling date*treatment as sources <strong>of</strong> variation. The association between aphids<br />

and <strong>the</strong> most frequent and abundant predators was additionally analysed using <strong>the</strong> Jacard's<br />

(Ludwig & Reynolds, 1988) and Horn's (Horn, 1981) indices.<br />

Results<br />

Aphids were found on maize since plant emergence. There was a significantly lower density<br />

<strong>of</strong> aphids in plots treated with imidacloprid during <strong>the</strong> whole sampling period (Fig. 1). Aphid<br />

populations in treated plots consisted almost exclusively by alatae that did not reproduce,<br />

especially in <strong>the</strong> first half <strong>of</strong> <strong>the</strong> sampling period.<br />

The predators found on maize plants during <strong>the</strong> two-year study were: Heteroptera (Orius<br />

spp., Nabis provencallis Remane and o<strong>the</strong>rs, less abundant), Dermaptera, Neuroptera<br />

(Chrysoperla carnea Stephens and Hemerobiidae), Coccinellidae (Coccinella septempunctata<br />

L, Adonia variegata (Goeze), Propylea quatuordecimpunctata (L.) , Scymnus spp. and o<strong>the</strong>rs,<br />

less abundant), Carabidae (Demetrias atricapillus L., Bembidion spp. and o<strong>the</strong>rs, much less<br />

abundant), Staphylinidae (Tachyporus spp. and o<strong>the</strong>rs) and Arachnida (Spiders, Trombidids<br />

and o<strong>the</strong>rs less abundant). Their relative abundance is shown in Table 1.<br />

Tab. 1. Occurrence <strong>of</strong> groups <strong>of</strong> predators found on maize plants in surveys <strong>of</strong> 1997 and<br />

1998 taken as a whole, in untreated plots and plots treated with imidacloprid.<br />

Predator<br />

Untreated<br />

(%)<br />

Treated<br />

(%)<br />

Orius spp. 27.5 19.2<br />

N. provencallis 3.1 3.7<br />

O<strong>the</strong>r Heteroptera 0.4 0.2<br />

C. 7-punctata 1 0.7<br />

A. variegata 2.7 3.2<br />

Larvae Coccinellidae 2.8 2.8<br />

O<strong>the</strong>r Coccinellidae 1 0.7<br />

D. atricapillus 9.6 7.9<br />

Bembidion spp. 1.6 0.9<br />

O<strong>the</strong>r Carabidae 0.5 0.4<br />

Staphylinidae 5 6.2<br />

Dermaptera 0 0.9<br />

Arachnida 43.9 52.8<br />

O<strong>the</strong>rs 0.7 0.5<br />

The same pattern was found in treated as in untreated plots. Arachnida and Orius spp.<br />

were <strong>the</strong> most common predators. The carabid D. atricapillus was also quite abundant in both


75<br />

years, and its relative abundance was slightly higher than that <strong>of</strong> <strong>the</strong> Coccinellidae taken as a<br />

group. Staphylinidae was <strong>the</strong> next group in abundance. Lacewings and hoverflies showed low<br />

presence. Cecidomids were not recorded.<br />

Plots treated with imidacloprid showed lower densities <strong>of</strong> Orius spp. than untreated plots<br />

in both study years, whereas D. atricapillus and arachnids were more abundant in untreated<br />

plots only in 1998 (Tab. 2). No significant differences (P < 0.05) were found for any o<strong>the</strong>r<br />

predator species or group considered (Tab. 2). Jacard's and Horn's indices between aphids and<br />

Orius spp., arachnids and D. atricapillus always yielded non-significant values.<br />

Discussion<br />

Although little is known about <strong>the</strong> effects <strong>of</strong> imidacloprid on aphids or predators in maize, it<br />

has been reported to reduce aphid densities and <strong>the</strong> spread <strong>of</strong> virus in winter cereals (Gourmet<br />

et al., 1994; Gray et al., 1996). The results <strong>of</strong> our study show that <strong>the</strong> seed treatment was<br />

effective in controlling aphids until maize flowering. The aphid population structure in treated<br />

plots suggests that <strong>the</strong> effect on aphids was fast, and that <strong>the</strong>ir reproduction was prevented but<br />

not <strong>the</strong> risk <strong>of</strong> maize dwarf mosaic virus (MDMV) transmission, due to its non-persistent<br />

nature.<br />

The relative abundance <strong>of</strong> predators found in this study agrees with that reported by Asín<br />

and Pons (1998a) and confirms that, in Catalonia, polyphagous predators are <strong>the</strong> main natural<br />

control agents in maize ecosystems.<br />

The lower densities <strong>of</strong> Orius spp., arachnids and D. atricapillus in treated plots could be<br />

due to:<br />

(i) An indirect effect <strong>of</strong> imidacloprid through its action on aphid populations, so plots<br />

with lower aphid densities would have lower predator densities. However, <strong>the</strong>re was no<br />

association between aphids and <strong>the</strong>se predators. Asin and Pons (1998b) also found poor<br />

correlations between densities <strong>of</strong> aphids and predators and between relative rates <strong>of</strong> increase<br />

<strong>of</strong> aphids and predators, and imputed this fact to <strong>the</strong> polyphagy <strong>of</strong> predators.<br />

(ii) A direct detrimental effect <strong>of</strong> imidacloprid on predators. This phenomenon would be<br />

more plausible in species that may feed on maize plant fluids or pollen, such as Orius spp.<br />

Seed treatment with imidacloprid does not appear to have any negative effect on:<br />

coccinellids, nabids, staphylinids, dermapters or o<strong>the</strong>r carabids preying on maize plants.<br />

This study indicates that:<br />

a) Maize seed dressing with imidacloprid is very effective in controlling aphids.<br />

b) Aphids are probably not <strong>the</strong> only determinant <strong>of</strong> <strong>the</strong> abundance <strong>of</strong> polyphagous predators<br />

in maize ecosystems, and <strong>the</strong>y can prey on o<strong>the</strong>r arthropods living on maize plants which<br />

may be less affected or unaffected by <strong>the</strong> seed treatment with imidacloprid.


76<br />

a) 1997<br />

aphids/plant<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

* *<br />

*<br />

* *<br />

*<br />

* * * * *<br />

13/5 19/5 26/5 2/6 9/6 16/6 23/6 30/6 7/7 21/7 28/7<br />

sampling date<br />

Untreated<br />

Treated<br />

b) 1998<br />

25<br />

20<br />

aphids/plant<br />

15<br />

10<br />

5<br />

0<br />

*<br />

*<br />

*<br />

*<br />

* *<br />

*<br />

*<br />

* *<br />

22/5 27/5 1/6 8/6 15/6 23/6 29/6 9/7 14/7 20/7 27/7<br />

sampling date<br />

Untreated<br />

Treated<br />

Fig. 1. Aphid densities on untreated plots and plots treated with imidacloprid from maize<br />

emergence (first sampling date) to flowering (last sampling date) in a) 1997 and b)<br />

1998. Asterisks upon bars mean that significant differences between treatments were<br />

found (P


77<br />

77<br />

Tab. 2. Mean seasonal densities <strong>of</strong> different groups <strong>of</strong> predators in maize plots untreated and treated with imidacloprid. (NS: non significant<br />

differences between plots; **: P


78<br />

Acknowledgements<br />

This research was funded by <strong>the</strong> Spanish Inter-Ministry <strong>of</strong> Science and Technology<br />

Committee (Comisión Interministerial de Ciencia y Tecnología - CICYT), project AGF96-<br />

0482.<br />

References<br />

Asín, L. & Pons, X. 1998 a: Aphid predators in maize fields. <strong>IOBC</strong>/<strong>WPRS</strong> Bull. 21 (8): 163-<br />

170.<br />

Asín, L. & Pons, X. 1998 b: Role <strong>of</strong> predators on maize aphid populations. In Aphids in<br />

natural and managed ecosystems (Nieto Nafría, J.M. & Dixon, A.F.G. (eds.). Universidad<br />

de León (Secretariado de Publicaciones). León (Spain). pp: 505-511.<br />

Gourmet, C., Hewings, A.D., Kolb, F.L. & Smyth, C.A. 1994: Effect <strong>of</strong> imidacloprid on<br />

nonflight movement <strong>of</strong> Rhopalosiphum padi and <strong>the</strong> subsequent spread <strong>of</strong> barley yellow<br />

dwarf virus. Plant. Dis. 78: 1098-1101.<br />

Gray, S.M., Bergstrom, G.C., Vaughan, R., Smith, D.M. & Kalb, D.W. 1996: Insecticidal<br />

control <strong>of</strong> cereal aphids and its impact on <strong>the</strong> epidemiology <strong>of</strong> <strong>the</strong> barley yellow dwarf<br />

luteoviruses. Crop Prot. 15: 687-697.<br />

Horn, D.J. 1981: Effect <strong>of</strong> weedy backgrounds on colonization <strong>of</strong> collards by green peach<br />

aphid, Myzus persicae, and its major predators. Environ. Entomol. 10: 285-289.<br />

Ludwig, J.A. & Reynolds, J.F. 1988: Statistical Ecology. A primer on methods and<br />

computing. John Wiley & Sons. New York. 337 pp.<br />

Piqué, J., Eizaguirre, M. & Pons, X. 1998: Soil insecticide treatments against maize soil pests<br />

and corn borers in Catalonia under traditional crop conditions. Crop Prot. 17: 557-561.<br />

Pons, X., Asín, L., Comas, J. & Albajes, R. 1994: Las especies de pulgones del maíz. Inv.<br />

Agr. Prod. Prot. Veg. Fuera de Serie 2: 125-129.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 79 - 86<br />

Arthropod natural enemies <strong>of</strong> <strong>the</strong> cereal leaf beetle (Oulema melanopus L.)<br />

in organic winter wheat fields in Vienna, Eastern Austria<br />

Peter Meindl, Bernhard Kromp, Barbara Bartl and Eleni Ioannidou<br />

L.-Boltzmann-Institute for Biological Agriculture and Applied Ecology, Rinnboeckstr. 15,<br />

A-1110 Vienna, Austria<br />

Summary<br />

In organic winter wheat fields in Vienna <strong>the</strong> species composition and rates <strong>of</strong> parasitism <strong>of</strong> <strong>the</strong> cereal<br />

leaf beetle (CLB) (Oulema melanopus) by egg- and larval parasitoids were investigated. Up to 95.7%<br />

<strong>of</strong> <strong>the</strong> eggs were parasitised by Anaphes flavipes (Hymenoptera, Mymaridae), averaging 67.3%. These<br />

rates were apparently not well synchronised with CLB egg density dynamics.<br />

Diaparsis carinifer (Hymenoptera, Ichneumonidae) was <strong>the</strong> dominant larval parasitoid, with rates<br />

<strong>of</strong> parasitism averaging 28%.<br />

In a laboratory feeding experiment we tested 28 different arthropod species from 12 families (e.g.<br />

Carabidae, Staphylinidae, Coccinellidae, Miridae, Nabidae, Araneae), sampled from <strong>the</strong> fields for <strong>the</strong>ir<br />

readiness to consume CLB eggs and larvae; <strong>the</strong> majority accepted CLB developmental stages as a<br />

food source.<br />

In an outdoor cage experiment, <strong>the</strong> impact <strong>of</strong> three predatory beetle species (Poecilus cupreus,<br />

Philonthus cognatus, Coccinella septempunctata) on mortality <strong>of</strong> CLB eggs was investigated.<br />

Key words: Oulema melanopus, cereal leaf beetle, parasitoids, predators<br />

Introduction<br />

Cereal leaf beetles (CLB) (Oulema ssp.) are pests <strong>of</strong> small grains and may cause<br />

economically important yield losses by larval feeding on <strong>the</strong> flag leaves, mainly in Central<br />

and South-eastern Europe. In <strong>the</strong> United States, CLB threatened grain production as an<br />

introduced pest in <strong>the</strong> 1960s, spreading rapidly due to <strong>the</strong> absence <strong>of</strong> natural antagonists<br />

(Haynes & Gage, 1981).<br />

An increase in CLB population density accompanied by an increase in damage and yield<br />

losses has recently been reported from several European countries (Heyer & Wetzel, 1990;<br />

Kaniuczak, 1994; Jossi & Bigler, 1996). For Eastern Austria, CLB is also recorded as a<br />

harmful pest (Stangelberger, 1995), although no quantifiable long-term data on importance<br />

and population increase <strong>of</strong> Oulema spp. are available. Never<strong>the</strong>less, based on extension<br />

services and farmers, CLB is considered <strong>the</strong> most important cereal pest in Austria, mainly in<br />

oats and winter wheat.<br />

Since pesticides are largely banned in organic agriculture (see EC-Regulation 2092/91,<br />

Appendix II B), pest control <strong>the</strong>re mainly depends on preventive measures. Unfortunately,<br />

little information on <strong>the</strong> applicability and efficacy <strong>of</strong> preventive, indirect control measures is<br />

available (Kromp et al., 1999).<br />

From 1996 until 1998, we conducted a research project on non-chemical control<br />

measures against CLB. One <strong>of</strong> <strong>the</strong> objectives was to estimate <strong>the</strong> potential <strong>of</strong> naturally<br />

occurring antagonists <strong>of</strong> CLB in organic cereal fields. In this paper, <strong>the</strong> species composition<br />

and rates <strong>of</strong> parasitism <strong>of</strong> CLB eggs and <strong>the</strong> impact <strong>of</strong> larval parasitoids are reported, along<br />

with <strong>the</strong> results <strong>of</strong> a screening <strong>of</strong> predatory arthropods for <strong>the</strong>ir readiness to feed on CLB<br />

79


80<br />

developmental stages. Finally, data from an outdoor cage experiment, estimating <strong>the</strong> impact<br />

<strong>of</strong> three predatory coleopteran species on CLB egg mortality is presented.<br />

Material and methods<br />

Egg parasitoids<br />

To determine rates <strong>of</strong> parasitism <strong>of</strong> CLB eggs, eggs were sampled in two winter wheat<br />

cultivars (cv. Capo and Spartacus). The eggs were transferred directly from <strong>the</strong> leaves onto a<br />

sticky tape. The tapes were placed in Petri dishes and stored in <strong>the</strong> laboratory (23° C, 70-80%<br />

relative humidity). On each sampling date (May 16, 23, June 2, 11, 1997), 37-110 eggs were<br />

collected in each cultivar. The eggs were controlled daily under a dissecting microscope.<br />

Hatched parasitoids or larvae, stuck to <strong>the</strong> tape beside <strong>the</strong> egg in which <strong>the</strong>y had developed,<br />

were recorded. The emerged egg parasitoids were removed from <strong>the</strong> tape and later determined<br />

by <strong>the</strong> mymarid specialist J. Huber, Ontario, Canada.<br />

Larval parasitoids<br />

Species composition and rates <strong>of</strong> parasitism by larval parasitoids were determined by<br />

collecting 264 CLB 4 th instar larvae in a winter wheat and in an oat field on June 9 and 12,<br />

1997, respectively. In <strong>the</strong> laboratory, <strong>the</strong> larvae were placed in plastic dishes filled with sand.<br />

Cereal leaves were put in a test-tube accompanied with moistened filter paper and sealed with<br />

adhesive tape, in order to supply food for <strong>the</strong> larvae if <strong>the</strong>y were not yet ready to pupate in <strong>the</strong><br />

sand. After hatching <strong>of</strong> <strong>the</strong> CLB adults, <strong>the</strong> sand was sieved and <strong>the</strong> remaining cocoons were<br />

dissected in order to obtain parasitoids. Parasitoid larvae were determined according to <strong>the</strong><br />

keys by Montgomery and Dewitt (1975) and Haeselbarth (1989).<br />

Laboratory feeding experiment<br />

In addition, 28 different predatory arthropods were screened for <strong>the</strong>ir acceptance <strong>of</strong> CLB eggs<br />

and larvae as prey (for a complete list, see results section). The predatory arthropods had been<br />

previously collected from wheat fields by pitfall trapping and sweep netting, were <strong>the</strong>n<br />

transferred to <strong>the</strong> laboratory and kept individually in Petri dishes. One CLB larva or three<br />

eggs were added to each Petri dish. In case eggs and larvae were consumed by <strong>the</strong> predators,<br />

fresh ones were introduced. The experiments were replicated three to six times.<br />

Outdoor cage experiment<br />

The possible impact <strong>of</strong> three predatory beetle species (Philonthus cognatus [Staphylinidae];<br />

Coccinella septempunctata [Coccinellidae]; P. cupreus [Carabidae]) on egg mortality in <strong>the</strong><br />

field, gauze cages (40 cm in diameter, mesh width 1 mm) were placed over five stems <strong>of</strong><br />

winter wheat and buried 20 cm deep into <strong>the</strong> ground around <strong>the</strong> stems. After removing all<br />

arthropods from inside <strong>the</strong> cage, a few CLB adults were introduced into <strong>the</strong> cage for egg<br />

laying. Thereafter <strong>the</strong> CLB adults were removed, <strong>the</strong> total number <strong>of</strong> eggs and <strong>the</strong>ir position<br />

recorded and two predators per cage inserted. The experiment was replicated 6-fold. The eggs<br />

were individually controlled by means <strong>of</strong> a portable dissecting microscope.<br />

Results and discussion<br />

Egg parasitism<br />

The most important egg parasitoid <strong>of</strong> Oulema spp. is Anaphes flavipes (Hymenoptera,<br />

Mymaridae), which is only 0.6 mm in size (Huber, 1992). The biology <strong>of</strong> A. flavipes and its<br />

significance as a natural control agent <strong>of</strong> CLB has been reported during <strong>the</strong> attempts to<br />

establish <strong>the</strong> mymarid in <strong>the</strong> United States in <strong>the</strong> 1960s (Anderson & Paschke, 1968; Maltby<br />

et al., 1971). For Austria, first data on rates <strong>of</strong> parasitism <strong>of</strong> Oulema spp. eggs by A. flavipes


81<br />

were reported by Bartl (1997). Average parasitism was around 40% with maximal rates <strong>of</strong><br />

87%, regardless whe<strong>the</strong>r eggs had been sampled in <strong>the</strong> centre <strong>of</strong> <strong>the</strong> fields or close to field<br />

margins.<br />

In our study, eggs were collected from two winter wheat cultivars (cv. Capo and<br />

Spartacus) that were part <strong>of</strong> a cultivar plot trial.<br />

When comparing egg parasitism and CLB egg densities, A. flavipes development clearly<br />

lagged behind <strong>the</strong> development <strong>of</strong> its host. Figure 1 shows a decrease <strong>of</strong> CLB eggs per stem<br />

from mid-May onwards, accompanied by an increase in egg parasitism. At <strong>the</strong> time <strong>of</strong><br />

maximal CLB egg density (0.55 eggs per stem) <strong>the</strong> rate <strong>of</strong> parasitism was still below 20%.<br />

Never<strong>the</strong>less, A. flavipes played a role in CLB egg mortality, as indicated by <strong>the</strong> rapid<br />

increase in egg parasitism from May 16 onwards. Moreover, rates <strong>of</strong> parasitism remained at a<br />

constant high level until <strong>the</strong> end <strong>of</strong> <strong>the</strong> CLB egg-laying period.<br />

Ellis and Kormos (1988) attributed an outbreak <strong>of</strong> CLB in Ontario in 1987 to <strong>the</strong> absence<br />

<strong>of</strong> parasitoids; in <strong>the</strong>ir study only 0.1% <strong>of</strong> Oulema spp. eggs were parasitised by A. flavipes.<br />

Heyer (1992) recorded no parasitism in 323 CLB eggs, sampled during surveys between 1988<br />

–and 1990, possibly due intensive, conventional agricultural practices in <strong>the</strong> study area.<br />

Enhancing <strong>the</strong> impact <strong>of</strong> A. flavipes could be <strong>the</strong> most promising venue for future research on<br />

biological control <strong>of</strong> CLB.<br />

eggs/stem x 100<br />

parasitation rate %<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

May 14 May 16 May 20 May 23 May 26 June 2 June 11<br />

parasitation rate<br />

egg density<br />

Fig. 1. Synchronisation <strong>of</strong> Oulema spp. egg density and rates <strong>of</strong> parasitism egg by A.<br />

flavipes.<br />

Larval parasitism<br />

Diaparsis carinifer, Lemophagus curtus (both Hymenoptera, Ichneumonidae) and<br />

Tetrastichus julis (Hymenoptera, Chalcidoidea) are known as larval parasitoids <strong>of</strong> O.<br />

melanopus (Haynes & Gage, 1981). Among o<strong>the</strong>r countries, <strong>the</strong>y have been reported from<br />

Austria during a Europe-wide survey designed to introduce <strong>the</strong>se parasitoids into <strong>the</strong> United<br />

States (Dysart et al., 1973). Their potential for controlling CLB in <strong>the</strong> United States has been


82<br />

discussed (Dysart et al., 1973; Ellis & Kormos, 1988). Recent data from Austria, obtained<br />

from organic cereal fields around Vienna, showed average rates <strong>of</strong> larval parasitism by D.<br />

carinifer <strong>of</strong> 5.4% (maximum 12.9%) and by L. curtus <strong>of</strong> 0.4% (Bartl, 1997).<br />

In our samples from 1997, we could only rear D. carinifer from Oulema spp. larvae. On<br />

two sampling dates in winter wheat and oats, rates <strong>of</strong> parasitism ranged between 21.8% and<br />

33%, with an average <strong>of</strong> 28% (Tab. 1).<br />

Tab. 1: Larval parasitism <strong>of</strong> Oulema melanopus by <strong>the</strong> ichneumonid Diaparsis carinifer,<br />

Vienna, 1997.<br />

Grain Sampling date Nr. <strong>of</strong> Oulema larvae Rates <strong>of</strong> parasitism<br />

(%)<br />

Winter wheat June 9 77 21.8<br />

June 12 76 29.2<br />

Oats June 9 81 27.9<br />

June 12 30 33<br />

Rates <strong>of</strong> parasitism by D. carinifer <strong>of</strong> 6.7% and 6.2% were reported from central Lower-<br />

Austria in 1968 and 1969, respectively (Dysart et al., 1973). Our data indicates that D.<br />

carinifer can be a significant mortality factor for O. melanopus larvae. CLB-larvae collected<br />

later in <strong>the</strong> season were heavier parasitised than those collected earlier in <strong>the</strong> season. Dysart et<br />

al. (1973) reported that D. carinifer can parasitise all larval instars <strong>of</strong> CLB, but prefers<br />

smaller larvae as hosts.<br />

Feeding experiment<br />

The effects <strong>of</strong> predatory insects and spiders on <strong>the</strong> population dynamics <strong>of</strong> several arthropod<br />

pests have been reported in <strong>the</strong> literature. Carabids, staphylinids and coccinellids, for instance,<br />

were able to reduce cereal aphid populations significantly (Chambers et al., 1983; Scheller,<br />

1984; Andersen, 1992). Their contribution on mortality <strong>of</strong> CLB, however, has been largely<br />

neglected. Mamedov (1995) tested <strong>the</strong> ability <strong>of</strong> certain predators to feed on eggs and larvae<br />

<strong>of</strong> O. melanopus in <strong>the</strong> laboratory. Schärer (1993) observed coccinellids (mainly<br />

C. septempunctata) feeding on CLB eggs, Nabis spp. sucking on CLB eggs and smaller<br />

larvae, and Chrysopa carnea larvae feeding on CLB larvae.<br />

In our study, we tried to ga<strong>the</strong>r preliminary data on arthropods that might accept different<br />

developmental stages <strong>of</strong> CLB as prey. Table 2 lists all arthropod species used in our<br />

laboratory feeding experiment along with <strong>the</strong>ir feeding rates, i.e. <strong>the</strong> quotient <strong>of</strong> consumed<br />

and presented prey.<br />

With a few exceptions (spiders, cantharid beetles), CLB eggs were accepted by all tested<br />

arthropods. Miridae, Nabidae, <strong>the</strong> staphylinid Ph. Cognatus, as well as <strong>the</strong> carabid P. cupreus<br />

showed <strong>the</strong> highest feeding rates. Since some carabid species such as Harpalus rufipes and P.<br />

cupreus have been reported to climb grain stems at night when searching for prey (Loughride<br />

& Luff, 1983; Lövei & Szentkiralyi, 1984; Chiverton 1988), we also included some frequent<br />

epigaeic carabids (e.g. Bembidion lampros) that had not yet been observed preying in <strong>the</strong><br />

vegetation layer. Spiders did not feed on CLB eggs, probably because - with <strong>the</strong> exception <strong>of</strong><br />

<strong>the</strong> thomisid Xysticus kochi - <strong>the</strong> tested species were web-weavers following a „sit and wait“<br />

(i.e. stationary) predation strategy whose prime requirement is <strong>the</strong> mobility <strong>of</strong> <strong>the</strong> potential<br />

prey (Nyffeler, 1999).


83<br />

Tab. 2: Average feeding rate <strong>of</strong> arthropod antagonists on eggs and larvae <strong>of</strong> <strong>the</strong> cereal leaf<br />

beetle in <strong>the</strong> laboratory.<br />

Order Family Species Feeding<br />

rate eggs<br />

Feeding<br />

rate larvae<br />

Araneae Thomisidae Xysticus kochi 0 0.8<br />

Theridiidae Neottinura bimaculata 0 0<br />

Tetragnathidae Tetragnatha pinicola 0 0.35<br />

Araneidae Mangora acalypha 0.03 0.33<br />

Philodromidae Tibellus oblongus 0 0.75<br />

Heteroptera Miridae Adelphocoris 4-punctata – 0.55<br />

Calocoris sp. 0.6 –<br />

Leptoderna dolobrata – 0.33<br />

Lygus spp. 0 –<br />

Miris striatus 0.8 –<br />

Nabidae Nabis spp. 0.8 0.32<br />

Nabis pseud<strong>of</strong>erus 0.5 0.87<br />

Nabis rugosus – 0.91<br />

Aptus mirmicoides 0.92 0.96<br />

Coleoptera Carabidae Amara plebeja 0.4 0.75<br />

Asaphidion flavipes 0.45 0.79<br />

Agonum dorsale 0.6 0.96<br />

Bembidion lampros 0.39 0.91<br />

Harpalus tardus 0.18 0.8<br />

H. distinguendus 0.71 0.98<br />

H. rufipes 0.62 1<br />

Poecilus cupreus 0.98 1<br />

Staphylinidae Tachyporus hypnorum 0.45 0.5<br />

Philonthus cognatus 1 1<br />

Cantharidae Cantharis rustica 0 0.55<br />

Coccinellidae Coccinella 7-punctata 0.44 0.56<br />

Propylea 14-punctata 0.5 0.25<br />

Malachiidae Malachius bipustulatus 0 0<br />

Surprisingly, CLB larvae were accepted by almost every predator tested, although <strong>the</strong><br />

larvae are covered by <strong>the</strong>ir own faeces, probably to protect <strong>the</strong>m from being attacked. As in<br />

<strong>the</strong> case <strong>of</strong> CLB eggs, Nabidae, Carabidae and Ph. cognatus showed high feeding rates on<br />

CLB larvae. Amara plebeja, feeding also on CLB larvae, was found in sweep net samples<br />

from wheat; it has previously not been known to occur in <strong>the</strong> vegetation layer. Coccinelids,


84<br />

usually known as aphid antagonists, also readily accepted eggs and larvae <strong>of</strong> CLB as prey,<br />

corroborating earlier reports by Mamedov (1995) and Schärer (1993).<br />

Since most <strong>of</strong> <strong>the</strong> 28 tested predators accepted different developmental stages <strong>of</strong> CLB in<br />

<strong>the</strong> laboratory as prey, <strong>the</strong>ir actual impact on Oulema spp. mortality in <strong>the</strong> field was tested in<br />

a preliminary field trial.<br />

Cage experiments<br />

Cage experiments are helpful tools to generate quantitative data on <strong>the</strong> efficacy <strong>of</strong> predators in<br />

<strong>the</strong> field (Chambers et al., 1983; Hance, 1987; Andersen, 1992). However, little data on field<br />

mortality <strong>of</strong> O. melanopus due to arthropod predators is available. Schärer (1993) reported<br />

that in early May 50% <strong>of</strong> marked eggs in a winter wheat field were destroyed or had<br />

disappeared. In cage experiments by Mamedov (1995), 35% <strong>of</strong> Oulema spp. eggs were<br />

consumed by predatory arthropods, mainly by <strong>the</strong> staphylinid Tachyporus hypnorum.<br />

To evaluate <strong>the</strong> impact <strong>of</strong> <strong>the</strong> three beetle predators, i.e. P. cupreus, Ph. cognatus and C.<br />

septempunctata (selected after initial laboratory feeding experiments), we conducted an<br />

outdoor cage experiment in an organic winter wheat field in 1997. Figure 2 shows mortality<br />

data <strong>of</strong> Oulema spp. eggs in <strong>the</strong> cages. Average egg mortality was 92.7, 95.3, and 96.2% after<br />

introduction <strong>of</strong> Ph. cognatus, C. septempunctata and P. cupreus, respectively. In <strong>the</strong> predatorfree<br />

control on average 71.7% egg mortality was recorded. No significant differences in egg<br />

mortality was recorded between <strong>the</strong> three predator treatments; however, in all predator<br />

treatments significant higher mortality values compared to <strong>the</strong> untreated control were<br />

observed (P < 0.0001 for C. septempunctata and P. cupreus; P < 0.001 for Ph. cognatus).<br />

The high egg mortality values in <strong>the</strong> predator-free control were due to unsuccessful egg<br />

development and by parasitism by A. flavipes, which due tom its small size could not fully be<br />

excluded from <strong>the</strong> cages. In a laboratory trial Heyer (1992) also recorded high egg mortality<br />

in CLB (42 - 52.9%) because <strong>of</strong> unsuccessful egg development.<br />

100<br />

80<br />

egg mortality %<br />

60<br />

40<br />

20<br />

0<br />

P. cupreus C. septempunctata P. cognatus Control<br />

Fig. 2: Average egg mortality by three predators (control without predators) in a cage<br />

experiment in a winter wheat field, Vienna 1997.


85<br />

Conclusions<br />

Research on non-chemical control strategies for <strong>the</strong> cereal leaf beetle in organic grain<br />

indicates some potential <strong>of</strong> both choice <strong>of</strong> cultivar and reduced seeding rates for preventing<br />

<strong>the</strong> establishment <strong>of</strong> high CLB population densities in <strong>the</strong> early phase <strong>of</strong> population build-up<br />

(Meindl & Kromp, 1999). In addition, CLB populations might be effectively controlled by<br />

arthropod antagonists provided that <strong>the</strong>y occur in sufficiently high numbers in <strong>the</strong> field. Our<br />

study showed high parasitism by <strong>the</strong> egg parasitoid A. flavipes. As <strong>the</strong> egg parasitoid is an<br />

important factor <strong>of</strong> CLB, in future research <strong>the</strong> habitat requirements <strong>of</strong> <strong>the</strong> mymarid should<br />

elucidated with aiming at suitable habitat management strategies (for a review see Pickett &<br />

Bugg, 1998). The same is true for larval parasitoids <strong>of</strong> CLB. Moreover, <strong>the</strong> potential <strong>of</strong><br />

polyphagous predatory arthropods as CLB control agents needs to be evaluated for <strong>the</strong>ir<br />

efficacy in controlling CLB larvae both in <strong>the</strong> vegetation and on <strong>the</strong> surface <strong>of</strong> <strong>the</strong> soil.<br />

Acknowledgements<br />

These investigations were part <strong>of</strong> a research project funded by <strong>the</strong> Austrian Ministry <strong>of</strong><br />

Agriculture and Forestry and <strong>the</strong> Research Fund <strong>of</strong> <strong>the</strong> Austrian National Bank. We thank M.<br />

Stachowitsch for <strong>the</strong> linguistic revision.<br />

References<br />

Andersen, A., 1992: Predation by selected carabid and staphylinid species on <strong>the</strong> aphid<br />

Rhopalosiphum padi in laboratory and semifield experiments. Norwegian Journal <strong>of</strong><br />

Agricultural Sciences 6: 265-273.<br />

Anderson, R.C. & Paschke, J.D., 1968: The biology and ecology <strong>of</strong> Anaphes flavipes<br />

(Hymenoptera, Mymaridae), an exotic egg parasite <strong>of</strong> <strong>the</strong> cereal leaf beetle. Ann.<br />

Entomol. Soc. Amer. 61: 1-5.<br />

Bartl, B., 1997: Zum Antagonistenkomplex (Parasitoide und Prädatoren) der Eier und Larven<br />

der Getreidehähnchen Oulema sp. (Coleoptera, Chrysomelidae) auf biologisch bewirtschafteten<br />

Feldern in der Oberen Lobau, Wien. Diploma <strong>the</strong>sis, University <strong>of</strong> Vienna.<br />

Chambers, R.J., Sunderland, K.D., Wyatt, K.D., Vickermann, G.P., 1983: The effect <strong>of</strong><br />

predator exclusion and caging on cereal aphids in winter wheat. Journal <strong>of</strong> Applied<br />

Ecology 20: 209-224.<br />

Chiverton, P.A., 1988: Searching behaviour and cereal aphid consumption by Bembidion<br />

lampros and Pterostichus cupreus, in relation to temperature and prey density. Ent. Exp.<br />

Appl. 47: 173-182.<br />

Dysart, R.J., Maltby, H.L., Brunson, M.H., 1973: Larval parasites <strong>of</strong> Oulema melanopus in<br />

Europe and <strong>the</strong>ir colonization in <strong>the</strong> United States. Entomophaga 18: 133-167.<br />

Ellis, C.R. & Kormos, B., 1988: Absence <strong>of</strong> parasitism in an outbreak <strong>of</strong> <strong>the</strong> cereal leaf<br />

beetle, Oulema melanopus (Coleoptera, Chrysomelidae), in <strong>the</strong> central tobacco growing<br />

area <strong>of</strong> Ontario. <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> Entomol. Soc. <strong>of</strong> Ontario 119: 43-46.<br />

Haeselbarth, E., 1989: Über einige Schlupfwespen (Hymenoptera) als Parasiten des Weizenhähnchens<br />

Oulema lichenis (Voet) (Coleoptera: Chrysomelidae) in Südbayern. J. Appl.<br />

Ent. 107: 493-507.<br />

Hance, T., 1987: Predation impact <strong>of</strong> carabids at different population densities <strong>of</strong> Aphis fabae<br />

development in sugar beet. Pedobiologia 30: 251-262.<br />

Haynes, D.L. & Gage, S.H., 1981: The cereal leaf beetle in North America. Ann. Rev. Ent.<br />

26: 259-287.


86<br />

Heyer, W. & Wetzel, T., 1990: Zum Auftreten des Getreidehähnchen (Oulema melanopus L.<br />

und O. lichenis Voet) und zur Aktualisierung des Bekämpfungsrichtwertes. Nachr.-Bl.<br />

Pflanzenschutzdienst DDR. 44: 226-230.<br />

Heyer, W., 1992: Zur Parasitierung der Getreidehähnchen Oulema spp. im Gebiet von Halle.<br />

Mitt. Dtsch. Ges. allg. angew. Ent. 8: 87-89.<br />

Huber, J.T., 1992: The subgenera, species groups and synonyms <strong>of</strong> Anaphes (Hymenoptera,<br />

Mymaridae) with a review <strong>of</strong> <strong>the</strong> described nearctic species <strong>of</strong> <strong>the</strong> fuscipennis group <strong>of</strong><br />

Anaphes s.s. and <strong>the</strong> described species <strong>of</strong> Anaphes (Yungaburra). Proc. Ent. Soc. Ont.<br />

123: 23-110.<br />

Jossi, W. & Bigler, F., 1996: Getreidehähnchen: Befall und Resistenz bei Winterweizen<br />

(Schweiz). Agrarforschung (Switzerland) 3(3): 117-119.<br />

Kaniucziak, Z., 1994: Injuriousness <strong>of</strong> Oulema in winter wheat. Ochrana-Roslin.38(7): 3-4.<br />

Kromp, B., Meindl, P. & Harris, P.J.C. (eds.), 1999: Entomological research in organic<br />

agriculture. A B Academic Publishers, England.<br />

Loughridge, A.H. & Luff, M.L., 1983: Aphid predation by Harpalus rufipes (Degeer) (Col.,<br />

Carab.) in <strong>the</strong> laboratory and field. J. Appl. Ecol. 20: 451-462.<br />

Lövei, G. & Szentkiralyi, F., 1984: Carabids climbing maize plants. Z. Angew. Ent. 97: 107-<br />

110.<br />

Maltby, H.L., Stehr, F.W., Anderson, R.C., Moorehead, G.E., Barton, L.C., Paschke, J.D.,<br />

1971: Establishment in <strong>the</strong> United States <strong>of</strong> Anaphes flavipes, an egg parasite <strong>of</strong> <strong>the</strong><br />

cereal leaf beetle. J. Econom. Entomol. 64: 693-697.<br />

Mamedov, A.A., 1995: On quantitative estimation <strong>of</strong> <strong>the</strong> effectiveness <strong>of</strong> natural enemies <strong>of</strong><br />

Oulema melanopus L. (Coleoptera, Chrysomelidae) on winter wheat. Entomological<br />

Review 74 (2): 1-9.<br />

Meindl, P. & Kromp, B., 1999: Vorbeugende Maßnahmen zur Kontrolle des<br />

Getreidehähnchens Oulema melanopus L. im ökologischen Anbau. In: Beiträge zur 5.<br />

Wissenschaftstagung zum Ökologischen Landbau „Vom Rand zur Mitte“. H<strong>of</strong>fmann, H.<br />

& Müller, S. (eds.), Verlag Dr. Köster, Berlin: 521-524.<br />

Montgomery, V.E. & Dewitt, P.R., 1975: Morphological differences among immature stages<br />

<strong>of</strong> three genera <strong>of</strong> exotic larval parasitoids attacking <strong>the</strong> cereal leaf beetle in United<br />

States. Ann. Entomol. Soc. Amer. 68 (3): 574-578.<br />

Nyffeler, M., 1999: Prey selection <strong>of</strong> spiders in <strong>the</strong> field. The Journal <strong>of</strong> Arachnology 27:<br />

317-324.<br />

Pickett, C.h.H. & Bugg, R.L. (eds.), 1998: Enhancing Biological Control. Habitat<br />

Management to Promote Natural Enemies <strong>of</strong> Agricultural Pests. University <strong>of</strong> California<br />

Press, 422 pp.<br />

Schärer, P., 1993: Analyse dichtebeeinflussender Faktoren beim Getreidehähnchen (Oulema<br />

sp., Coleoptera, Chrysomelidae). PhD <strong>the</strong>sis, University Bern, Switzerland.<br />

Scheller, H.V., 1984: The role <strong>of</strong> groundbeetles (Carabidae) as predators on early populations<br />

<strong>of</strong> cereal aphids in spring barley. Z. Angew. Ent. 97: 451-463.<br />

Stangelberger, J., 1995: Bericht über Witterungsverlauf und bemerkenswertes Schadauftreten<br />

an Kulturpflanzen in Österreich im Jahr 1994. Pflanzenschutzberichte 55(2): 108-128.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 87 - 90<br />

Habitat preference <strong>of</strong> carabids (Coleoptera: Carabidae) in Central<br />

Hungary in winter wheat field and in adjacent habitats<br />

A. Hatvani 1 , F. Kádár 2 , J. Kiss 3 and G. Péter 3<br />

1<br />

Horticult. High School, 6000 Kecskemét, Erdei F. tér 1-3, Hungary<br />

2<br />

Plant Protect. Inst., Hung. Acad. Sci., 1525 Budapest, P.O. Box 102, Hungary<br />

3<br />

Dept. <strong>of</strong> Plant Protect, Szent István University Gödöllő, 2100 Gödöllő, Hungary<br />

Summary<br />

A pitfall trap study was carried out in three habitats (winter wheat field, its margin and in nearby tree<br />

rows), in 1995, in Central Hungary. A total <strong>of</strong> 1,586 individuals <strong>of</strong> 61 carabid species were collected<br />

in <strong>the</strong> three habitats. The most frequent species was Amara tricuspidata. Amara tricuspidata,<br />

Bembidion properans, Calathus ambiguus preferred <strong>the</strong> wheat field, whereas Amara anthobia,<br />

Calathus fuscipes and Harpalus tardus preferred <strong>the</strong> tree rows, and Harpalus serripes <strong>the</strong> field<br />

margin. The field margin was favourable also for C. fuscipes and H. tardus, while Amara aenea,<br />

Harpalus distinguendus, Microlestes minutulus, Poecilus cupreus, P. punctulatus, P. sericeus<br />

favoured fields. Harpalus rufipes was found in <strong>the</strong> wheat field and in <strong>the</strong> margin but showed a<br />

preference for <strong>the</strong> margin.<br />

Key words: carabid beetles, habitat preference, Hungary<br />

Introduction<br />

Carabids are potentially important natural pest control agents (Kromp, 1999), thus are focal<br />

points <strong>of</strong> many research related to IPM in Hungary. Investigations on species composition,<br />

abundance, habitats and swarming phenology <strong>of</strong> carabid beetles have been a part <strong>of</strong> <strong>the</strong> IPM<br />

activities in winter wheat in Hungary (see Kiss et al., 1993). In <strong>the</strong> present study we present<br />

data <strong>of</strong> <strong>the</strong> surveys conducted on <strong>the</strong> Great Plain <strong>of</strong> Central Hungary, an area which was not<br />

investigated in our previous studies (Kiss et al., 1994, 1998). Our objectives were to collect<br />

large numbers <strong>of</strong> adult ground beetles and study <strong>the</strong>ir biology, and especially <strong>the</strong>ir habitat<br />

preferences.<br />

Material and methods<br />

Surveys were carried out on <strong>the</strong> Great Plain in Central Hungary (near <strong>the</strong> town Nagykőrös) in<br />

a large scale winter wheat field (133 ha), in its field margins (4 m wide), and in tree rows (25<br />

m wide), next to <strong>the</strong> field margins. The soil was a sandy one. The plant cover <strong>of</strong> <strong>the</strong> study<br />

sites could be described as follows:<br />

Field margin<br />

Agropyron repens, Cynodon dactylon, Artemisia vulgaris, Melandrium album, Polygonum<br />

aviculare, Arctium lappa, Taraxacum <strong>of</strong>ficinale, Medicago sativa, Cannabis sativa,<br />

Amaranthus retr<strong>of</strong>lexus, Apera spica-venti, Echinochloa crus-galli, Tragopogon sp.,<br />

Amorpha fruticosa.<br />

87


88<br />

Tree rows<br />

Robinia pseudoacacia, Alnus sp., Artemisia vulgaris, Chenopodium album, Agropyron<br />

repens, Sambucus ebulus, Cannabis sativa, Lamium purpureum, Galium aparine, Arctium<br />

lappa, Rubus sp.<br />

Winter wheat field<br />

Triticum aestivum cv. Mv-19, Portulaca oleracea and Stellaria media.<br />

The winter wheat was sown in early October 1994, and harvested in July 1995. The<br />

previous crop was also winter wheat. Fifteen plastic pitfall traps were used for sampling <strong>of</strong><br />

carabids from end <strong>of</strong> April until end <strong>of</strong> December 1995. Traps contained 2 % <strong>of</strong> a formalin<br />

solution, were emptied weekly and were operated in three rows as follows: five traps were<br />

placed in <strong>the</strong> field margin, five were situated in <strong>the</strong> tree rows, and five were placed in <strong>the</strong><br />

wheat field (30 m from <strong>the</strong> margin). The distance between <strong>the</strong> traps in one row was 5 m.<br />

In order to estimate <strong>the</strong> effect <strong>of</strong> <strong>the</strong> winter wheat harvest on <strong>the</strong> carabid assemblage we<br />

illustrate <strong>the</strong> % <strong>of</strong> weekly catches <strong>of</strong> <strong>the</strong> 3 rd week, 2 nd week and 1 st week prior to and after <strong>the</strong><br />

harvest as % <strong>of</strong> <strong>the</strong> total 6-weeks catch. For <strong>the</strong> evaluation <strong>of</strong> habitat preferences α indices<br />

were used (Manly et al., 1972; Chesson, 1978, 1983).<br />

N<br />

∑<br />

α = n i / n i<br />

i=<br />

1<br />

, where n i is <strong>the</strong> value for <strong>the</strong> given species on i place.<br />

If preference was shown for more habitats, Yates corrected χ 2 -values were calculated on<br />

<strong>the</strong> basis <strong>of</strong> contingence tables for deciding <strong>the</strong> primary habitat.<br />

Tab. 1. List <strong>of</strong> carabids abundant in <strong>the</strong> sampling area<br />

Species<br />

Number <strong>of</strong><br />

α values*<br />

individuals<br />

Field Margin Tree rows<br />

Amara aenea (De Geer) 65 4 2 0.92<br />

A. anthobia (Villa) 0 3 57 0.95<br />

A. tricuspidata Dejean 348 49 2 0.87<br />

Bembidion properans (Stephens) 96 6 0 0.94<br />

Calathus ambiguus (Paykull) 204 21 17 0.84<br />

C. fuscipes (Goeze) 36 86 127 0.51<br />

Harpalus distinguendus<br />

49 7 2 0.84<br />

(Duftschmid)<br />

H. rufipes (De Geer) 23 38 2 0.60<br />

H. serripes (Quensel) 0 104 26 0.80<br />

H. tardus (Panzer) 1 81 101 0.55<br />

Zabrus tenebrioides (Goeze) 68 14 9 0.75<br />

* for explanation refer to chapter Material and Methods and <strong>the</strong> text.<br />

Results<br />

A total <strong>of</strong> 1,586 adult individuals <strong>of</strong> 61 carabid species were caught during <strong>the</strong> survey.<br />

Species and individual numbers in <strong>the</strong> wheat field were 39 and 811, in <strong>the</strong> field margin 45 and


89<br />

411, while in <strong>the</strong> tree rows 28 and 294. Eighteen species were found in each <strong>of</strong> <strong>the</strong> three<br />

habitats. Most abundant species are listed in Table 1. Amara tricuspidata was <strong>the</strong> dominant<br />

carabid species regarding its total individual number for <strong>the</strong> three habitats.<br />

Ground surface activity <strong>of</strong> carabids showed a strong seasonal activity with a smaller peak<br />

in May-June and with a higher one in August-September. Lowest numbers <strong>of</strong> beetles were<br />

caught during July. Though total catch in <strong>the</strong> wheat field decreased in July, density in <strong>the</strong><br />

field increased after <strong>the</strong> harvest (Tab. 1).<br />

Habitat preferences <strong>of</strong> carabid adults (α-values, Table 1) were assumed as follows:<br />

- Preference for wheat field: A. tricuspidata, B. proberans, C. ambiguus<br />

- Preference for grassy field margin: H. serripes<br />

- Preference for tree rows: A. anthobia<br />

Calathus fuscipes and H. tardus have affinity to each sampled habitat according to <strong>the</strong>ir<br />

lower α-value, but <strong>the</strong>y preferred tree rows (χ 2 values = 39.28, and = 53.51, respectively, P <<br />

0.05). Accordingly, A. aenea, H. distinguendus, M. minutulus (Goeze), P. cupreus (Linnaeus),<br />

P. puntulatus (Schaller) and P. sericeus Fischer proved to prefer wheat field (α > 0.8 in each<br />

case). Harpalus rufipes showed a preference for <strong>the</strong> margin (χ 2 value=25.95, P < 0.05).<br />

Discussion<br />

In our survey, A. tricuspidata was <strong>the</strong> most dominant carabid species in <strong>the</strong> winter wheat<br />

field. However, no data on its biology, and habitat preference is available. Its is in <strong>the</strong> carabid<br />

occurrence list <strong>of</strong> Lövei & Sárospataki (1990) though with low score. Frequently observed<br />

field species in our earlier investigations like M. minutulus, P. cupreus, P. punctulatus or<br />

P. sericeus (Kiss et al., 1998) occurred in <strong>the</strong> present study only in low numbers. Thus,<br />

influence <strong>of</strong> <strong>the</strong> local conditions on carabid assemblages is <strong>of</strong> great importance.<br />

Higher species and lower individual number <strong>of</strong> carabids in <strong>the</strong> margin compared to wheat<br />

field were found in <strong>the</strong> present study, corroborating earlier results from surveys in nor<strong>the</strong>rn<br />

Hungary (Kiss et al. 1998).<br />

50<br />

40<br />

FIELD<br />

MARGIN<br />

TREE ROWS<br />

30<br />

%<br />

20<br />

10<br />

0<br />

-3 -2 -1 +1 +2 +3<br />

Fig. 1. Change <strong>of</strong> catches <strong>of</strong> carabid beetles prior to and after <strong>the</strong> harvest <strong>of</strong> winter wheat


90<br />

Decreasing individual number <strong>of</strong> carabids at harvest time is a consequence <strong>of</strong> <strong>the</strong>ir<br />

reproduction cycle and not caused by <strong>the</strong> harvest itself. The explanation for this is that<br />

individual number was higher during <strong>the</strong> first week after <strong>the</strong> harvest than before harvest (Tab.<br />

1). On <strong>the</strong> o<strong>the</strong>r hand two peaks <strong>of</strong> dominant species (like A. tricuspidata) prior and after <strong>the</strong><br />

harvest consist <strong>of</strong> adults <strong>of</strong> two different generations with low catches at harvest time.<br />

Though carabid species in our study clearly belonged to one <strong>of</strong> <strong>the</strong> preference groups<br />

(field or margin species), <strong>the</strong> α values indicate <strong>the</strong>ir shift to habitats o<strong>the</strong>r than <strong>the</strong> preferred<br />

one. Many <strong>of</strong> <strong>the</strong> carabid species in agricultural landscape are moving between cultivated<br />

fields and <strong>the</strong>ir margins (Desender et al., 1981; So<strong>the</strong>rton, 1985; Wallin, 1985). From <strong>the</strong><br />

species in Table 1 data <strong>of</strong> C. fuscipes, H. tardus, H. rufipes and Zabrus tenebriodes indicate<br />

such dispersion pattern. Strong affinity <strong>of</strong> <strong>the</strong> three Poecilus species for winter wheat has<br />

been demonstrated in our study, confirming results <strong>of</strong> earlier studies (Szél et al., 1997; Kiss et<br />

al., 1998). Thus habitat preferences <strong>of</strong> carabid species in various regions in Hungary are very<br />

similar, though <strong>the</strong> influence <strong>of</strong> crop type to <strong>the</strong> activity and distribution <strong>of</strong> carabid species<br />

were demonstrated in o<strong>the</strong>r studies Cárcamo & Spence (1994).<br />

References<br />

Cárcamao, H.A. & Spence, J.R. 1994: Crop type effects on <strong>the</strong> activity and distribution <strong>of</strong><br />

ground beetles (Coleoptera: Carabidae). Environ. Entomol. 23: 684-692.<br />

Chesson, J. 1978: Measuring preference in selective predation. Ecology 59: 211-215.<br />

Chesson, J. 1983: The estimation and analysis <strong>of</strong> preference and its relationship to foraging<br />

models. Ecology 64: 1297-1304.<br />

Desender, K., Maelfait, P.-P., D’Hulster, M. & VanHercke, L. 1981: Ecological and faunal<br />

studies on Coleoptera in agricultural land. I. Seasonal occurrence <strong>of</strong> Carabidae in <strong>the</strong><br />

grassy edge <strong>of</strong> a pasture. Pedobiologia 22: 379-384.<br />

Kiss, J., Kádár, F., Kozma, E. & Tóth, I. 1993: Importance <strong>of</strong> various habitats in agricultural<br />

landscape related to integrated pest management: a preliminary study. Landscape and<br />

Urban Planning 27: 191-198.<br />

Kiss, J., Kádár, F., Tóth, I., Kozma, E. & Tóth, F. 1994: Occurrence <strong>of</strong> predatory arthropods<br />

in winter wheat field and in <strong>the</strong> field edge. Écologie 25: 127-132.<br />

Kiss, J., Kádár, F., Tóth, F., Barth, R. & Hatvani, A. 1998: Predatory arthropods sampled in<br />

pitfall traps in winter wheat in nor<strong>the</strong>rn Hungary. <strong>IOBC</strong> Bull. 21: 81-90.<br />

Kromp, B. 1999: Carabid beetles in sustainable agriculture: a review on pest control efficacy,<br />

cultivation impacts and enhancement. Agriculture, Ecosystems and Environment 74: 187-<br />

228.<br />

Lövei, G.L. & Sárospataki, M. 1990: Carabid beetles in agricultural fields in eastern Europe.<br />

In: The Role <strong>of</strong> Ground Beetles in Ecological and Environmental Studies, Intercept Ltd.<br />

Andover, ed. Stork: 87-93.<br />

Manly, B.F.J., Miller, P. & Cook, L.M. 1972: Analysis <strong>of</strong> selective predation experiments.<br />

Amer. Nat. 106: 719-736.<br />

So<strong>the</strong>rton, N.W. 1985: The distribution and abundance <strong>of</strong> predatory Coleoptera overwintering<br />

in field boundaries. Ann. Appl. Biol. 106: 17-21.<br />

Szél, Gy., Kádár, F. & Faragó, S. 1997: Abundance and habitat preference <strong>of</strong> some adultoverwintering<br />

ground beetle species in crops in western Hungary (Coleoptera:<br />

Carabidae). Acta Phytopath. Entom. Hung. 32: 369-376.<br />

Wallin, H. 1985: Spatial and temporal distribution <strong>of</strong> some abundant carabid beetles (Coleoptera:<br />

Carabidae) in cereal fields and adjacent habitats. Pedobiologia 28: 19-34.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 91 - 94<br />

Role <strong>of</strong> field margin in <strong>the</strong> winter phenophase <strong>of</strong> Carabid beetles<br />

(Coleoptera: Carabidae) in winter wheat field<br />

G. Péter 1 , F. Kádár 2 , J. Kiss 1 and F. Tóth 1<br />

1<br />

Department <strong>of</strong> Plant Protection, Szent István University, Gödöllő, H-2100 Gödöllő, Hungary<br />

2<br />

Plant Protection Research Institute <strong>of</strong> <strong>the</strong> Hungarian Academy <strong>of</strong> Sciences, Budapest,<br />

Hungary<br />

Summary<br />

Earlier studies on <strong>the</strong> ecological background <strong>of</strong> integrated pest management <strong>of</strong> winter wheat<br />

performed by <strong>the</strong> Department <strong>of</strong> Plant Protection at Szent István University, have identified <strong>the</strong> most<br />

important Carabid and spider species and revealed <strong>the</strong> potential importance <strong>of</strong> field margins as <strong>the</strong>ir<br />

habitat in Hungary. However, no evidence was available for <strong>the</strong> overwintering <strong>of</strong> Carabids in <strong>the</strong><br />

wheat field and in <strong>the</strong> field margins. Soil samples were taken on József-major farm, near Kartal<br />

village, nor<strong>the</strong>rn Hungary in December 1997 and 1998. Ten soil samples were taken from <strong>the</strong> field<br />

margin, 5-5 samples were dig in <strong>the</strong> wheat field at 1 m and 20 m distance from <strong>the</strong> margin. Soil was<br />

ga<strong>the</strong>red from an area <strong>of</strong> 31,5 x 31,5 cm per sample. In 1997 soil was removed to a depth <strong>of</strong> 25 cm. In<br />

1998, excavation depth was decreased to 15 cm. Arthropods and organic matter was removed from <strong>the</strong><br />

soil by floating <strong>the</strong> soil samples in water. Majority <strong>of</strong> Carabid beetles in our samples occurred in <strong>the</strong><br />

field margin both winters. In <strong>the</strong> wheat field, 1 m from <strong>the</strong> margin substantially lower numbers <strong>of</strong><br />

individuals and species were observed than in <strong>the</strong> field margin. No Carabids were found in samples<br />

taken 20 m from <strong>the</strong> margin. Thus, field margins serve as overwintering habitats for Carabids, which<br />

colonise <strong>the</strong> fields from <strong>the</strong>re.<br />

Key words: Carabid beetles, winter wheat, winter phenophase, field margin<br />

Introduction<br />

Winter wheat is one <strong>of</strong> Hungary’s major crop plants, covering about 25% <strong>of</strong> <strong>the</strong> country’s<br />

arable land (app. 1,1 million ha). This cereal growing area represents a wide spread artificial<br />

habitat that has its own characteristics on <strong>the</strong> basis <strong>of</strong> <strong>the</strong> arthropod community (Kiss et al.,<br />

1994). Integrated farming (as phrased by <strong>IOBC</strong> guidelines) and integrated pest control relies<br />

substantially on natural enemies, <strong>of</strong> which Carabid beetles represent a major group.<br />

Therefore, protection <strong>of</strong> predators by means <strong>of</strong> creating and maintaining appropriate habitats<br />

plays a key role in integrated pest management (IPM). Establishment <strong>of</strong> regional IPM<br />

guidelines needs regional data that will support <strong>the</strong> adaptation <strong>of</strong> <strong>WPRS</strong> guidelines in<br />

Hungary.<br />

Earlier studies on <strong>the</strong> ecological background <strong>of</strong> IPM <strong>of</strong> winter wheat carried out by <strong>the</strong><br />

Department <strong>of</strong> Plant Protection <strong>of</strong> Szent István University, have identified <strong>the</strong> most important<br />

Carabid and spider species and revealed <strong>the</strong> important role <strong>of</strong> field margins in Hungary (Kiss<br />

et al., 1994). However, to date, no evidence is available for <strong>the</strong> role <strong>of</strong> field margins in <strong>the</strong><br />

winter phenophase <strong>of</strong> predator populations. In our two-year survey we were focusing on <strong>the</strong><br />

winter phenophase <strong>of</strong> Carabid beetles in winter wheat fields and <strong>the</strong> adjacent field margins.<br />

91


92<br />

Material and methods<br />

The sampling area was located near Kartal village, about 40 km east <strong>of</strong> Budapest, Hungary,<br />

on <strong>the</strong> experimental farm <strong>of</strong> <strong>the</strong> Szent István University. The winter wheat fields sampled in<br />

December 1997 and January 1999 both occupied an area <strong>of</strong> 60 ha. The field margins were 3-5<br />

m wide, with dense herbaceous undergrowth coverage and sporadic Robinia trees, showing a<br />

typical view <strong>of</strong> field margins in <strong>the</strong> region. Dominant weed species <strong>of</strong> <strong>the</strong> margins were<br />

Galium aperine, Hordeum vulgare and Calamagrostis epigeios. Both fields were treated with<br />

Logran 75 WG herbicide (9 g triasulfuron ai/ha) in early spring and Bion 50 WG fungicide<br />

(30 g bendicar ai/ha) in late spring. No insecticides were applied to <strong>the</strong> winter wheat fields.<br />

Each winter 10 soil samples were taken from <strong>the</strong> field margins, 5-5 samples were dig in<br />

<strong>the</strong> wheat field at 1 m and 20 m distance from <strong>the</strong> margin. Soil was ga<strong>the</strong>red from an area <strong>of</strong><br />

31,5 x 31,5 cm per sample. Thus, total sampled area was 1 square meter both in <strong>the</strong> margin<br />

and in <strong>the</strong> field. In 1997 soil was removed to a depth <strong>of</strong> 25 cm. In 1998, building on<br />

experiences <strong>of</strong> <strong>the</strong> previous year, excavation depth was decreased to 15 cm. Arthropods and<br />

organic matter was removed from <strong>the</strong> soil by floating <strong>the</strong> soil samples in water. Later<br />

arthropods were separated manually from <strong>the</strong> debris.<br />

Tab. 1. Carabid species and <strong>the</strong>ir individual number in soil samples in winter, 1997/1998<br />

List <strong>of</strong> species and <strong>the</strong>ir Margin 1 m infield 20 m infield<br />

individual number<br />

Pterostichus sericeus<br />

38 2 0<br />

(Fischer von Waldheim)<br />

Pterostichus cupreus L. 21 1 0<br />

Agonum dorsale<br />

21 0 0<br />

Pontoppidan<br />

Microlestes minutulus<br />

8 0 0<br />

(Goeze)<br />

Calosoma auropunctatum 2 0 0<br />

(Herbst)<br />

Harpalus distinguendus 8 0 0<br />

Duftschmidt<br />

Harpalus tardus (Panzer) 3 0 0<br />

Harpalus azureus (Fab.) 2 0 0<br />

Brachinus explodens<br />

6 0 0<br />

Duftschmidt<br />

Amara similata (Gyllenhal) 4 1 0<br />

Total 113 4 0<br />

Results and discussion<br />

Winter <strong>of</strong> 1997/1998<br />

A total number <strong>of</strong> 117 specimens <strong>of</strong> Carabid beetles were found in <strong>the</strong> soil samples in <strong>the</strong><br />

winter <strong>of</strong> 1997/1998, which were distributed among 10 species (Tab. 1). Pterostichus sericeus<br />

(Fischer von Waldheim) proved to be superdominant (33.6 % in total number <strong>of</strong> specimens),


93<br />

Pterostichus cupreus L. and Agonum dorsale Pontoppidan appeared to be dominant (both<br />

18.6 % in total number <strong>of</strong> specimens).<br />

The number <strong>of</strong> sampled species and specimens was substantially lower in <strong>the</strong> infield<br />

samples than those <strong>of</strong> <strong>the</strong> field margin. No Carabid beetles were found in any <strong>of</strong> <strong>the</strong> 20 m<br />

samples. No species was present exclusively in <strong>the</strong> winter wheat field.<br />

In 1997 one soil sample from all three sets were taken in a way that <strong>the</strong> upper 15 cm and<br />

<strong>the</strong> lower 10 cm layers could be investigated separately. Since no Carabid beetles were found<br />

in <strong>the</strong> lower 10 cm <strong>of</strong> <strong>the</strong> soil samples, <strong>the</strong>y are not presented in Tab. 1. During <strong>the</strong> next<br />

winter sampled soil was removed from sampling areas to a depth <strong>of</strong> 15 cm only.<br />

Tab. 2. Carabid species and <strong>the</strong>ir individual number in soil samples in winter, 1998/1999<br />

List <strong>of</strong> species and <strong>the</strong>ir Margin 1 m infield 20 m infield<br />

individual number<br />

Pterostichus sericeus<br />

25 1 0<br />

(Fischer von Waldheim)<br />

Pterostichus cupreus L. 14 0 0<br />

Agonum dorsale<br />

12 0 0<br />

Pontoppidan<br />

Microlestes minutulus<br />

6 0 0<br />

(Goeze)<br />

Calosoma auropunctatum 1 0 0<br />

(Herbst)<br />

Brachynus explodens<br />

1 0 0<br />

Duftschmidt<br />

Harpalus rufipalpis Sturm 10 2 0<br />

Microlestes maurus (Sturm) 3 0 0<br />

Harpalus rufipes (De Geer) 2 0 0<br />

Total 74 3 0<br />

Winter <strong>of</strong> 1998/1999<br />

A total number <strong>of</strong> 77 specimens <strong>of</strong> Carabid beetles were sampled in <strong>the</strong> winter <strong>of</strong> 1998/1999,<br />

which were distributed among nine species (Tab. 2).<br />

Pterostichus sericeus proved to be superdominant again (33.8 % in total number <strong>of</strong><br />

specimens), P. cupreus and A. dorsale appeared to be dominant (18.9 and 16.2 %<br />

respectively, in total number <strong>of</strong> specimens). No species was present exclusively in <strong>the</strong> winter<br />

wheat field. The number <strong>of</strong> collected species and specimens was substantially lower in <strong>the</strong><br />

infield samples than those <strong>of</strong> <strong>the</strong> field margin. No Carabid beetles were found in any <strong>of</strong> <strong>the</strong> 20<br />

m samples.<br />

The majority <strong>of</strong> Carabid species and specimens were found in <strong>the</strong> field margin in both<br />

winter periods (Tab. 1 and Tab. 2). In <strong>the</strong> winter <strong>of</strong> 1997/1998 96.6 %, in 1998/1999 96.1 %<br />

<strong>of</strong> total specimens were sampled in <strong>the</strong> field margin. Given this overwhelming rate <strong>of</strong><br />

Carabids’ presence it can be concluded that field margins play a key role in <strong>the</strong> overwintering<br />

period <strong>of</strong> ground beetles. This finding, which correlate with that <strong>of</strong> Welling’s (1994), can be<br />

best explained by that larvae and adults developing after <strong>the</strong> winter wheat harvest at early<br />

summer find suitable food in <strong>the</strong> field margins. Weed coverage <strong>of</strong> <strong>the</strong> field margins and


94<br />

grassy lanes also provide shelter and rest-places for Carabids (Kiss et al., 1993). Later in <strong>the</strong><br />

year, as temperatures decrease considerably, ground beetles tend to move to field margins for<br />

overwintering. According to <strong>the</strong> investigations <strong>of</strong> So<strong>the</strong>rton (1984), Carabids can stay in <strong>the</strong><br />

field in <strong>the</strong> winter phenophase. This phenomenon was not observed during our investigations.<br />

Winter wheat, planted in early October, is not tall enough to provide sufficient shelter for<br />

Carabids.<br />

According to our results in <strong>the</strong> winter <strong>of</strong> 1997/1998 3.4 %, and in <strong>the</strong> winter <strong>of</strong><br />

1998/1999 3.9 % <strong>of</strong> total specimens were found 1 m infield from <strong>the</strong> margin. Possibly some<br />

Carabid species are still active in late autumn and early winter, and some - for example P.<br />

cupreus, a dominant species in <strong>the</strong> samples - show activity even in December (Kiss, pers.<br />

com.). Therefore, some individuals can be found overwintering in <strong>the</strong> field, close to <strong>the</strong><br />

margin. No Carabids were collected 20 m infield during both winters, probably because soil<br />

and plant structure are not suitable for late autumn feeding and overwintering. The substantial<br />

difference between <strong>the</strong> number <strong>of</strong> collected individuals <strong>of</strong> <strong>the</strong> two sampling period is an<br />

outcome <strong>of</strong> 1998 being a year with higher than average precipitation, which caused severe<br />

frost in <strong>the</strong> upper layer <strong>of</strong> <strong>the</strong> soil both infield and in <strong>the</strong> margin.<br />

O<strong>the</strong>r investigations showed that a number <strong>of</strong> Carabid species overwinter in <strong>the</strong> field<br />

margins and invade cultivated lands in spring, corroborating earlier reports <strong>of</strong> Desender<br />

(1982) and So<strong>the</strong>rton (1984). Agonum dorsale, a dominant species in our samples is listed<br />

among <strong>the</strong>se Carabids.<br />

References<br />

Desender, K. 1982: Ecological and faunal studies on Coleoptera in agricultural land, II.<br />

Hibernation <strong>of</strong> Carabidae in agroecosystems. Pedobiologia 23: 295-303.<br />

Desender, K. et al. 1989: Field edges and <strong>the</strong>ir importance for polyphagous predatory<br />

arthropods. Med. Fac. Landbouww. Rijksuniv. Gent 54(3a): 823-833.<br />

Horvatovich, S. & Szarukán, I. 1986: Faunal investigations <strong>of</strong> ground beetles (Carabidae) in<br />

<strong>the</strong> arable soils <strong>of</strong> Hungary. Acta Agronomica Hung. 35: 107-123.<br />

Kiss, J. et al. 1993: Importance <strong>of</strong> various habitats in agricultural landscape related to<br />

integrated pest management: a preliminary study. Landscape and urban planning 27: 191-<br />

198.<br />

Kiss, J. et al. 1994: Occurrence <strong>of</strong> predatory arthropods in winter wheat and in <strong>the</strong> field edge.<br />

Ecologie 25(2): 127-132.<br />

So<strong>the</strong>rton, N.W. 1984: The distribution and abundance <strong>of</strong> predatory arthropods overwintering<br />

on farmland. Ann. Appl. Biol. 105: 423-429.<br />

Wallin, H. 1985: Spatial and temporal distribution <strong>of</strong> some abundant carabid beetles<br />

(Coleoptera: Carabidae) in cereal fields and adjacent habitats. Pedobiologia 28: 19-34.<br />

Welling, M. 1990: Dispersal <strong>of</strong> ground beetles (Coleoptera: Carabidae) in arable land. Med.<br />

Fac. Landbouww. Rijksuniv. Gent 55(2b): 483-491.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 95 - 99<br />

Insect pests <strong>of</strong> cereals in Croatia<br />

Jasminka Igrc-Barčić and Tanja Gotlin Culjak<br />

Faculty <strong>of</strong> Agriculture, Department <strong>of</strong> Zoology, Svetosimunska 25 10000 Zagreb, Croatia<br />

Short communication<br />

Wheat is sown in Croatia on an average <strong>of</strong> 220,000 ha each year. Barley, rye, oats and triticale<br />

are sown on additional 50,000 ha.<br />

The average yields on big farms are between 5.5 and 6 t/ha, with a maximum <strong>of</strong> 8 t/ha.<br />

Small farms harvest between 4 and 4.5t/ha. In <strong>the</strong> year 1999 <strong>the</strong> yields were 20% lower than<br />

on average.<br />

According to Maceljski (1995) in Croatia damage <strong>of</strong> cereals caused by insect pests are on<br />

average 7 %, by diseases 9 % and by weeds 11 %. In addition, approximately 5 % <strong>of</strong> <strong>the</strong><br />

stored cereals are destroyed by post-harvest pests.<br />

Among many insect species feeding on cereals we will mention <strong>the</strong> most important ones.<br />

At least three species <strong>of</strong> Thysanoptera are damaging cereals in Croatia. Among <strong>the</strong>m<br />

Haplothrips tritici Kurdyumov is <strong>the</strong> most important species. High densities <strong>of</strong> H. tritici are<br />

recorded once in ten years. According to previous studies by Beš (1967) an economical<br />

damage is caused when more than 18 specimens are present on one ear. Such population<br />

densities are very rare, and hence <strong>the</strong>re is no need for control <strong>of</strong> this insect pest.<br />

In 1964 and 1965 <strong>the</strong> first strong appearance <strong>of</strong> bugs <strong>of</strong> <strong>the</strong> genus Eurygaster (Het.:<br />

Scutelleridae) was recorded. This period followed <strong>the</strong> large introduction <strong>of</strong> combines for<br />

harvesting <strong>the</strong> cereals. As known, <strong>the</strong> combines are used in average ten days later than <strong>the</strong><br />

former harvest techniques. Therefore <strong>the</strong> bugs were now able to feed ten days longer on <strong>the</strong><br />

ears and finish <strong>the</strong>ir metamorphosis.<br />

Thereafter Eurygaster spp. were present each year on cereals in Croatia, but never<br />

reached <strong>the</strong> high densities <strong>of</strong> 1964 and 1965. Only in some years on some fields in east<br />

Slavonia control measures were needed. We believe that Hymenopteran egg parasitioids are<br />

largely controlling <strong>the</strong> populations <strong>of</strong> cereal bugs, mainly because <strong>of</strong> <strong>the</strong> very restricted use <strong>of</strong><br />

insecticides on <strong>the</strong> main field crops, i.e. wheat and maize.<br />

Aphids<br />

In Croatia first real investigations on <strong>the</strong> aphid fauna in cereals started in <strong>the</strong> eighties. The<br />

results showed that contrary to previous reports, in which Schizaphis graminum Rondani was<br />

always mentioned as <strong>the</strong> most important aphid species on cereals, <strong>the</strong> following cereal aphids<br />

were collected and identified:<br />

Sitobion avenae F., Metopolophium dirhodum Walk., Rhopalosiphum padi R., R. maidis<br />

Fitch, R. insertum Walk., M. festucae Theob., S. graminum Rond., Macrosiphum euphorbiae<br />

Th., Tetraneura spp., Anoecia spp. (Igrc, 1985; 1989).<br />

The most important species among <strong>the</strong>m are S. avenae, M. dirhodum and R. padi. Very<br />

<strong>of</strong>ten <strong>the</strong>se three species appear toge<strong>the</strong>r infesting leaves, stalks and ears. Only in some years<br />

<strong>the</strong> attack <strong>of</strong> aphids is causing economic damages. Aphid infestations can be harmful when<br />

<strong>the</strong>ir attack occur early as <strong>the</strong>ir multiplication factor during one month can reach 200. In<br />

Croatia Igrc (in Maceljski, 1999) recommends <strong>the</strong> following decision threshold for <strong>the</strong> use <strong>of</strong><br />

insecticides:<br />

95


96<br />

- <strong>the</strong> expected yield should be more than 6 t/ha,<br />

- more than 60 % infested ears at <strong>the</strong> begin <strong>of</strong> flowering,<br />

- more than 70% during flowering, and<br />

- more than 80% <strong>of</strong> infested ears during <strong>the</strong> milky seed stage.<br />

The infestation <strong>of</strong> ears include aphids on stalks and leaves when <strong>the</strong> predator/aphid ratio is<br />

greater than 1:40 (sometimes predators can prevent <strong>the</strong> increase <strong>of</strong> aphid population at a ratio<br />

until 1:1000).<br />

For <strong>the</strong> control <strong>of</strong> cereal aphids many organophosphourous insecticides, some pyrethroids<br />

and <strong>the</strong> carbamate pirimicarb are registered. Among <strong>the</strong>m <strong>the</strong> more selective ones are<br />

recommended, i.e. pirimicarb, demetonmethyl and thiometon. The newly developed aphicides<br />

like imidacloprid, thiamethoxam, fipronil etc. have not yet been registered for control <strong>of</strong><br />

cereal aphids in Croatia.<br />

Since 1988 we are monitoring aphids, using <strong>the</strong> standard stationary Agraphid suction trap<br />

at <strong>the</strong> Agricultural Faculty in Zagreb. Additionally, we are using a yellow water trap, placed<br />

25 m apart from <strong>the</strong> suction trap.<br />

Here we present results <strong>of</strong> <strong>the</strong> monitoring activities during 1996 and 1998 for all aphids<br />

species in general and for R. padi in particular, by <strong>the</strong> Agraphid and <strong>the</strong> yellow traps.<br />

Figure 1 shows data on trap catches (both Agraphid suction and yellow traps) <strong>of</strong> all aphid<br />

species between 1996 and 1998. The year 1997 was a real aphid outbreak year, with<br />

approximately tenfold higher trap catches than in <strong>the</strong> two o<strong>the</strong>r years.<br />

In Figure 2 trap catches <strong>of</strong> R. padi are shown. Interestingly no R. padi were caught in <strong>the</strong><br />

yellow traps in <strong>the</strong> spring, but only in autumn. Moreover, certain discrepancies were recorded<br />

in <strong>the</strong> total number <strong>of</strong> R. padi caught with <strong>the</strong> two trapping devices.<br />

In Croatia R. padi has it's maximum flight activity in September, and in October still<br />

many aphids fly when temperatures are above 11 °C. These findings underline <strong>the</strong> potential<br />

danger <strong>of</strong> R. padi as a vector for BYDV.<br />

The main cereal insect pests in Croatia are <strong>the</strong> cereal leaf beetles Oulema melanopus L.<br />

and O. gallaeciana (Heyden) (formerly O. lichenis [Voet]) (both Col.: Chrysomelidae), with<br />

O. melanopus being <strong>the</strong> more important species. During June and July <strong>of</strong> each study year, a<br />

strong migration <strong>of</strong> young beetles to maize was observed, leading to locally high damages<br />

levels. The degree <strong>of</strong> this damage depends mainly upon <strong>the</strong> proportion <strong>of</strong> fields on which<br />

cereals o<strong>the</strong>r than maize and maize is grown.<br />

We estimate that cereal leaf beetles are controlled on 5 to 40 % <strong>of</strong> fields in each year.<br />

Many organophosphourous and pyrethroid insecticides are registered for this use. Moreover<br />

bensultap is registered for <strong>the</strong> control <strong>of</strong> cereal leaf beetle larvae. In our trial some chitin<br />

syn<strong>the</strong>sis inhibitors showed a satisfactory efficacy, but Bt insecticides failed to be efficient.<br />

As a first control measure we propose to control overwintering adults by treating narrow<br />

border stripes <strong>of</strong> fields where more than 25 adult beetles per square meter were encountered.<br />

This measure should be conducted before <strong>the</strong> adults disperse into <strong>the</strong> whole field.<br />

For <strong>the</strong> control <strong>of</strong> <strong>the</strong> cereal leaf beetles larvae we propose <strong>the</strong> following decision<br />

threshold:<br />

- if a yield <strong>of</strong> less than 5 t/ ha is expected, <strong>the</strong> insecticides should be applied when more<br />

than two larvae on <strong>the</strong> flag leaf are present,<br />

- if a yield between 5 and 6 t/ ha is expected, <strong>the</strong> insecticides should be applied when more<br />

than 1-1.5 larvae on <strong>the</strong> flag leaf are present,<br />

- if <strong>the</strong> yield is expected to be higher than 6 t/ ha, <strong>the</strong> insecticides should be applied if more<br />

than 0.5-1 larvae on one flag leaf are present.


97<br />

1996<br />

700<br />

600<br />

500<br />

400<br />

N°<br />

300<br />

200<br />

100<br />

0<br />

III IV V VI VII VIII IX X XI<br />

YELLOW WATER TRAP<br />

AGRAPHID<br />

1997<br />

6000<br />

5000<br />

4000<br />

N°<br />

3000<br />

2000<br />

1000<br />

0<br />

III IV V VI VII VIII IX X XI<br />

YELLOW WATER TRAP<br />

AGRAPHID<br />

1998<br />

400<br />

350<br />

300<br />

250<br />

N°<br />

200<br />

150<br />

100<br />

50<br />

0<br />

III IV V VI VII VIII IX X XI<br />

YELLOW WATER TRAP<br />

AGRAPHID<br />

Fig. 1. The flight activity <strong>of</strong> aphids in Zagreb 1996-1998, recorded in suction and yellow<br />

traps.


98<br />

R. padi 1996<br />

160<br />

140<br />

24.10.<br />

120<br />

100<br />

N°<br />

80<br />

60<br />

40<br />

20<br />

0<br />

26.09.<br />

07.05.<br />

20.11.<br />

V VI VII VIII IX X XI<br />

AGRAPHID<br />

YELLOW WATER TRAP<br />

R. padi 1997<br />

480<br />

17.09.<br />

400<br />

320<br />

N°<br />

240<br />

160<br />

80<br />

0<br />

07.05.<br />

11.11.<br />

12.11.<br />

V VI VII VIII IX X XI XII<br />

AGRAPHID<br />

YELLOW WATER TRAP<br />

35<br />

R. padi 1998<br />

30<br />

25<br />

N°<br />

20<br />

15<br />

03.09.<br />

10<br />

5<br />

03.05.<br />

04.11<br />

0<br />

12.11.<br />

V VI VII VIII IX X XI<br />

AGRAPHID<br />

YELLOW WATER TRAP<br />

Fig. 2. The flight activity <strong>of</strong> Rhopalosiphum padi in Zagreb 1996-1998, recorded in suction<br />

and yellow traps.


99<br />

In Croatia 10-20 % <strong>of</strong> cereals are sown on fields where cereals had been grown <strong>the</strong> preceding<br />

year. However, damage due to Zabrus tenebrioides Goeze (Col. Carabidae) is very rare. In<br />

our conditions, a three year monoculture <strong>of</strong> cereals is needed for developing economically<br />

harmful densities <strong>of</strong> Z. tenebrioides. In such cases <strong>the</strong> greatest problem is to persuade farmers<br />

to treat in autumn and not to wait until spring when <strong>the</strong> use <strong>of</strong> insecticides can not prevent<br />

high damages. In <strong>the</strong> third year <strong>of</strong> a cereal monoculture an autumn application <strong>of</strong> soil or foliar<br />

insecticides should be regularly carried out. There are many species <strong>of</strong> dipterous larvae<br />

damaging cereals in Croatia. This pest complex was only partly investigated. According to<br />

Kovačević (1952) Mayetiola destructor Say, Sitodiplosis mosellana (Géhin), Contarinia<br />

tritici (Kirby) (all Dipt.: Cecidomyiidae), Oscinella frit L. and Chlorops pumilionis<br />

(Bjerkander) (both Dipt: Chloropidae) are present in Croatia, but we have never registered<br />

any economic damages <strong>of</strong> <strong>the</strong>se species except for frit fly on maize. We have recorded strong<br />

attack and damages by Haplodiplosis marginata Roser (Dipt.: Cecidomyiidae) in wheat fields<br />

in <strong>the</strong> mountain region <strong>of</strong> Poakakotlina. Some authors recently mentioned <strong>the</strong> presence <strong>of</strong><br />

Delia coarctata (Fallén) (Dipt.: Anthomyiidae) and Phorbia fumigata (Meigen) (both Dipt.:<br />

Anthomyiidae), though no data on economic impact is available. We have recorded a very<br />

wide distribution <strong>of</strong> D. platura (Meigen) on many host plants, with some damage on wheat.<br />

We suspect that this poorly studied insect pests is much more important not only on cereals,<br />

but also on o<strong>the</strong>r commonly planted crops in Croatia. Each year some species <strong>of</strong> leaf miners<br />

<strong>of</strong> <strong>the</strong> family Agromyzidae were found damaging <strong>the</strong> tops <strong>of</strong> wheat leaves, though not<br />

causing any economic damage.<br />

Concluding we can state that at <strong>the</strong> moment insect pest do not cause important damage to<br />

cereal crops in Croatia. The greatest potential danger are aphids especially if <strong>the</strong> importance<br />

<strong>of</strong> BYDV and o<strong>the</strong>r viruses will increase. Due to relatively large areas on which in Croatia<br />

cereals are grown we consider that any insecticide applications should be restricted as much<br />

as possible to prevent harm to natural enemies <strong>of</strong> insect pests and to <strong>the</strong> environment.<br />

References<br />

Beš, A. 1967: Prilog poznavanju biologije i ekonomskog značaja pšeničnog tripsa –<br />

Haplothrips tritici Kurd. na području Bosne i Hercegovine, doktorska disertacija.<br />

Igrc, J. 1985: Važnost i potreba suzbijanja lisnih uši (Aphididae) strnih žita, Agronomski<br />

glasnik, 3-4: 109-118.<br />

Igrc, J. 1989: Lisne uši strnih žitarica, Glasnik zaštite bilja, 338-342.<br />

Igrc, J in Maceljski, M. 1999: Poljoprivredna entomologija, 65-111.<br />

Kovačević, Ž. 1952: Primijenjena entomologija, Poljoprivredni štetnici.<br />

Maceljski, M. 1995: Štete od štetoèinja u Hrvatskoj, Glasnik zaštite bilja, 6: 261-265.


100


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 101 - 107<br />

The Hessian fly in Morocco:<br />

Surveys, loss assessment, and genetic resistance in bread wheat<br />

Lhaloui, S. 1 , M. El Bouhssini 2 and A. Amri 2<br />

1<br />

Institut National de la Recherche Agronomique, Centre Aridoculture, P.O. Box 589, Settat,<br />

Morocco<br />

2<br />

International Center for Agricultural Research in <strong>the</strong> Dry Areas, ICARDA, Aleppo, Syria<br />

Summary<br />

Hessian fly, Mayetiola destructor (Say) is <strong>the</strong> most destructive insect pest <strong>of</strong> wheat (Triticum spp) in<br />

<strong>the</strong> major cereal growing regions <strong>of</strong> <strong>the</strong> world. It is believed to have originated in West Asia which is<br />

<strong>the</strong> centre <strong>of</strong> diversity <strong>of</strong> wheat, and has spread out to Europe, North Africa, and North America.<br />

Heavy populations are detected regularly in North America and western Mediterranean countries. In<br />

North Africa, especially in Morocco, grain yield losses have been estimated to 42 and 36 % in bread<br />

wheat using an insecticide control method, and near isogenic resistant and susceptible lines<br />

respectively (Lhaloui et al., 1992; Amri et al., 1992). As a comparison, losses have been estimated to<br />

32 % in durum wheat (Lhaloui et al., 1992).<br />

The most efficient and economic method <strong>of</strong> control <strong>of</strong> this pest is <strong>the</strong> genetic resistance via <strong>the</strong> use<br />

<strong>of</strong> resistant cultivars. In <strong>the</strong> USA, 29 genes <strong>of</strong> resistance have been named and characterised. Ten <strong>of</strong><br />

<strong>the</strong>se genes (H5, H7H8, H11, H13, H14H15, H21, H22, H23, H25, and H26) were selected as<br />

conferring resistance to Hessian fly in <strong>the</strong> field in Morocco (El Bouhssini et al., 1996). Most <strong>of</strong> <strong>the</strong>m<br />

are located on <strong>the</strong> A or <strong>the</strong> D genomes. The H5 gene was located in a South Dakota bread wheat<br />

germplasm, and was released as <strong>the</strong> first variety resistant to Hessian fly in Morocco, under <strong>the</strong> name <strong>of</strong><br />

‘Saada’ in 1988. The H22 gene has successfully been introgressed into an adapted Moroccan bread<br />

wheat cultivar, and has recently been released under <strong>the</strong> name <strong>of</strong> ‘Aguilal’. Also, <strong>the</strong> CIMMYT bread<br />

wheat resistant line L222 has been registered in <strong>the</strong> <strong>of</strong>ficial catalogue, under <strong>the</strong> name ‘Arrihane’.<br />

In addition to <strong>the</strong>se known genes <strong>of</strong> resistance, we have identified a large number <strong>of</strong> sources in <strong>the</strong><br />

wild wheat relatives. These sources are very useful because <strong>the</strong>y increase <strong>the</strong> gene bank available to<br />

breeders to choose <strong>the</strong> resistance from, and <strong>the</strong>y also widen <strong>the</strong> genetic base.<br />

Key words: Hessian fly, infestations, losses, plant resistance, bread wheat.<br />

Introduction<br />

Bread wheat is one <strong>of</strong> <strong>the</strong> major cereal crops in Morocco. It is grown over 1.5 Million ha<br />

annually. However, <strong>the</strong> production potential <strong>of</strong> this crop is limited by several biotic and<br />

abiotic stress factors. Hessian fly, Mayetiola destructor (Say) is <strong>the</strong> most destructive insect<br />

pest <strong>of</strong> wheat (Triticum spp) in <strong>the</strong> major cereal growing regions <strong>of</strong> <strong>the</strong> world. It is believed to<br />

have originated in West Asia which is <strong>the</strong> centre <strong>of</strong> diversity <strong>of</strong> wheat, and has spread out to<br />

Europe, North Africa, and North America. Heavy populations are detected regularly in North<br />

America and western Mediterranean countries. In North Africa, especially in Morocco, this<br />

insect is <strong>the</strong> focal research subject in <strong>the</strong> program <strong>of</strong> protecting cereals against damaging<br />

pests.<br />

The objective <strong>of</strong> this paper is to give an overview <strong>of</strong> what has been accomplished in <strong>the</strong><br />

field <strong>of</strong> host plant resistance to this pest in Morocco, and <strong>the</strong> plans for future research.<br />

101


102<br />

Surveys and diagnosis<br />

1: In <strong>the</strong> cereal growing plains<br />

A five year survey was conducted over <strong>the</strong> seven biggest cereal production provinces <strong>of</strong><br />

Morocco, to determine <strong>the</strong> intensity and distribution <strong>of</strong> <strong>the</strong> pest infestations, compare <strong>the</strong><br />

infestations on bread wheat to those on durum wheat, observe annual variations in <strong>the</strong><br />

infestations levels, and estimate yield losses due to <strong>the</strong>se infestations.<br />

This study showed that Hessian fly was very common throughout <strong>the</strong> major cereal<br />

growing areas <strong>of</strong> Morocco. It was also fairly persistent over all <strong>the</strong> five year surveys. Eightyeight<br />

and 85 % <strong>of</strong> bread wheat and durum wheat fields respectively were infested.<br />

Tab. 1. Mayetiola spp infestations in fields <strong>of</strong> bread wheat (BW), as compared to durum<br />

wheat (DW) surveyed over five years in Morocco (cf. Lhaloui et al., 1992).<br />

Years<br />

Number <strong>of</strong><br />

fields<br />

sampled<br />

Infested<br />

Percent fields<br />

>20% tillers<br />

infested<br />

>50% tillers<br />

infested<br />

BW DW BW DW BW DW BW DW<br />

1986 61 57 87 88 72 61 36 30<br />

1987 56 36 79 67 25 22 9 3<br />

1988 59 36 86 79 66 41 27 7<br />

1989 63 62 94 90 92 89 30 39<br />

1990 73 69 93 94 67 49 32 26<br />

Mean/Total 312 285 88 85 65 55 27 22<br />

Tab. 2. Mayetiola spp infestations in fields <strong>of</strong> bread and durum wheat surveyed in seven<br />

provinces over five years in Morocco.<br />

Provinces<br />

Number <strong>of</strong><br />

fields<br />

sampled<br />

Infested<br />

Percent fields<br />

>20% tillers<br />

infested<br />

>50% tillers<br />

infested<br />

BW DW BW DW BW DW BW DW<br />

Settat 44 44 100 98 91 75 48 36<br />

El Jadida 41 40 98 98 90 83 46 35<br />

Safi 47 47 100 96 74 70 45 45<br />

Marrakesh 33 25 64 68 33 24 6 4<br />

El Kalaa 58 58 90 83 60 43 16 9<br />

Beni Mellal 54 40 80 78 59 48 17 10<br />

Khouribga 35 31 80 65 40 26 11 6<br />

Mean/Total 312 285 88 85 65 55 27 22


103<br />

Economic levels <strong>of</strong> infestations (20% <strong>of</strong> tillers infested, cf. Lafever et al. 1980) were<br />

observed in 65% as compared to 55% <strong>of</strong> bread wheat and durum wheat fields, respectively<br />

(Tab. 1).<br />

The percent fields with economic infestations remained stable across <strong>the</strong> study years.<br />

Severe infestations (over 50% tillers infested) were observed in 27 and 22 % <strong>of</strong> bread wheat<br />

and durum wheat fields, respectively (Tab. 2) (cf. Lhaloui et al., 1992a).<br />

These results clearly indicate that Hessian fly is very wide spread in Morocco, with high<br />

infestation levels in all <strong>the</strong> regions, and that bread wheat is <strong>the</strong> most infested cereal.<br />

2: In <strong>the</strong> regions <strong>of</strong> high altitude<br />

A smaller scale survey, over two years only, was conducted in <strong>the</strong> Highland regions <strong>of</strong> <strong>the</strong><br />

Atlas mountains; this study showed that 72% <strong>of</strong> <strong>the</strong> bread wheat fields were infested with<br />

Hessian fly, and that <strong>the</strong> fields presented up to 27% <strong>of</strong> tillers infested. For durum wheat, <strong>the</strong><br />

percent <strong>of</strong> fields infested was equivalent to that <strong>of</strong> bread wheat, but tiller infestation was<br />

lower (Tab. 3).<br />

Tab. 3. Levels <strong>of</strong> Infestations <strong>of</strong> cecidomyiids on bread wheat (BW) as compared to durum<br />

wheat (DW) in <strong>the</strong> regions <strong>of</strong> High elevation in Morocco (cf. Lhaloui et al., 1998).<br />

Regions<br />

number<br />

fields<br />

sampled<br />

% fields<br />

infested<br />

% plants<br />

infested<br />

DW BW DW BW DW BW<br />

Annoceur 2 7 100 100 37 54<br />

Guigou 3 7 67 71 8 12<br />

Midelt 8 4 56 50 6 15<br />

El Ksibah 7 4 48 74 14 30<br />

Azilal 7 3 74 67 14 25<br />

Mean/Total 27 25 69 72 16 27<br />

Tab. 4. Mean percent grain yield loss due to Hessian fly infestations on mid season plantings<br />

estimated over three years in three different regions.<br />

Crop<br />

Percent grain yield loss<br />

Bread Wheat 42<br />

Durum Wheat 32<br />

Barley 45<br />

Estimates <strong>of</strong> yield losses<br />

Grain yield losses have been estimated to be 42 and 36 % in bread wheat using an insecticide<br />

control method, and near isogenic resistant and susceptible lines respectively (Lhaloui et al.,<br />

1992; Amri et al., 1992). In durum wheat, losses have been estimated to be 32 % (Lhaloui et<br />

al., 1992), which is equivalent to losses estimated on bread wheat.


104<br />

The estimate <strong>of</strong> yield loss test showed that yield increases between chemically protected<br />

and non-protected varieties, averaged over a four year period <strong>of</strong> <strong>the</strong> test, 42, 32, and 45% for<br />

bread wheat, durum wheat and barley, respectively (Tab. 4). In an average rainfall season, <strong>the</strong><br />

percentage <strong>of</strong> loss was estimated to be 35% <strong>of</strong> <strong>the</strong> yield. This indicates that losses caused by<br />

this pest on durum wheat are equivalent to those caused on bread wheat (Lhaloui et al.,<br />

1992b).<br />

Genetic control:<br />

identification <strong>of</strong> sources <strong>of</strong> resistance and development <strong>of</strong> resistant cultivars<br />

Host plant resistance has been widely utilised to limit <strong>the</strong> damage caused by insect pests over<br />

<strong>the</strong> world. Sources <strong>of</strong> resistance have been continuously sought and introgressed into adapted<br />

cultivars. This method presents many advantages:<br />

1. It is economical for <strong>the</strong> farmer, as compared to <strong>the</strong> repeated use <strong>of</strong> chemicals, it does<br />

not cost him anything; everything is incorporated into <strong>the</strong> seed <strong>of</strong> <strong>the</strong> resistant cultivar.<br />

2. It is safe for man and animals and does not present any toxicity.<br />

3. It is environmentally sound and safe, and thus does not pollute <strong>the</strong> environment.<br />

1: Methodology <strong>of</strong> screening<br />

In <strong>the</strong> greenhouse and <strong>the</strong> growth chambers, <strong>the</strong> screening procedure we followed is similar to<br />

that described by Cartwright and LaHue (1944). The genetic material consists <strong>of</strong> collections<br />

<strong>of</strong> wild wheat relatives, and nurseries generated by breeders from crosses between identified<br />

sources <strong>of</strong> resistance and adapted Moroccan varieties. The number <strong>of</strong> entries can vary from<br />

5,000 to 1,000 per year. Lines are seeded in standard greenhouse flats, containing a mixture <strong>of</strong><br />

2/3 soil and 1/3 peat, at a rate <strong>of</strong> one row per line, and 25 seeds per row. Each flat contains a<br />

susceptible (cv. Nesma) and a resistant (cv. Saada ) check. When plants are in <strong>the</strong> two-leaf<br />

stage, flats are covered with a cheesecloth tent, and about one hundred females <strong>of</strong> newly<br />

emerged and mated Hessian flies are released in each flat. Females are allowed to lay eggs for<br />

two days. Three weeks later, plants are removed from <strong>the</strong> flats and checked for <strong>the</strong>ir<br />

resistance reaction. Plants that are stunted and had a dark green colour were considered as<br />

susceptible, those that have normal growth, with a light green colour are considered as<br />

resistant. Resistance is fur<strong>the</strong>r confirmed by <strong>the</strong> presence <strong>of</strong> dead first instar larvae at <strong>the</strong><br />

bases <strong>of</strong> stems.<br />

Advanced breeding material is screened in <strong>the</strong> field, usually at <strong>the</strong> Jemaa shaim<br />

experimental station, which is considered as <strong>the</strong> hot spot <strong>of</strong> Hessian fly occurrence in<br />

Morocco. They are seeded in 3 m lines, 50 cm spaced. The seeding date is chosen in late<br />

December to allow plants to get <strong>the</strong> maximum infestation by <strong>the</strong> second generation <strong>of</strong> adult<br />

Hessian flies. The screening is done visually in February after <strong>the</strong> infestations have occurred<br />

and plants had shown resistance or susceptibility reaction. The selection is made in April, at<br />

crop maturity, in collaboration with <strong>the</strong> breeding program.<br />

2: Major results and accomplishments<br />

In <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> program, we screened al <strong>the</strong> existing Moroccan germplasm.<br />

Unfortunately it turned out to be all susceptible.<br />

Then we screened <strong>the</strong> North American uniform Hessian fly nursery in 1984, and <strong>the</strong><br />

results showed that three genes <strong>of</strong> this nursery provided a high level <strong>of</strong> resistance to <strong>the</strong><br />

Moroccan populations <strong>of</strong> this pest. These were H5, H11, and H13.<br />

The H5 gene existed in a south Dakota spring bread wheat germplasm SD8036. This line<br />

was tested in Morocco for yield performance, and released in 1989 as <strong>the</strong> first cultivar


105<br />

resistant to Hessian fly in Morocco, under <strong>the</strong> name <strong>of</strong> Saada. Later, in 1994, <strong>the</strong> first<br />

Moroccan bread wheat cultivar, Massira, tolerant to <strong>the</strong> fly, was released.<br />

The most efficient and economic method <strong>of</strong> control <strong>of</strong> this pest is <strong>the</strong> genetic resistance<br />

via <strong>the</strong> use <strong>of</strong> resistant cultivars. In <strong>the</strong> USA, 27 genes <strong>of</strong> resistance have been characterised<br />

and named. Ten <strong>of</strong> <strong>the</strong>se genes (H5, H7H8, H11, H13, H14H15, H21, H22, H23, H25, and<br />

H26) were selected as conferring resistance to Hessian fly in <strong>the</strong> field in Morocco (El<br />

Bouhssini et al., 1996). Most <strong>of</strong> <strong>the</strong>m are located on <strong>the</strong> A or <strong>the</strong> D genomes. The H22 gene<br />

has been successfully introgressed into an adapted Moroccan bread wheat cultivar. Also, <strong>the</strong><br />

CIMMYT bread wheat resistant line L222 has been registered in <strong>the</strong> <strong>of</strong>ficial catalogue. Both<br />

cultivars are now in <strong>the</strong> seed increase phase.<br />

The search for new sources <strong>of</strong> resistance continued, and we now have identified a total <strong>of</strong><br />

12 genes(H5, H7H8, H11, H13, H14H15, H21, H22, H23, H25, and H26).<br />

Three o<strong>the</strong>r bread wheat lines were selected as presenting very high levels <strong>of</strong> resistance<br />

to <strong>the</strong> fly; <strong>the</strong>se were L222, L254, and ADC14. Last year, <strong>the</strong> L222 was registered as a<br />

cultivar under <strong>the</strong> name <strong>of</strong> Arrihane, and a bread wheat line carrying <strong>the</strong> H22 was released<br />

under <strong>the</strong> name <strong>of</strong> Aguilal. Both lines are now in <strong>the</strong> seed increase phase, and will reach <strong>the</strong><br />

farmer in <strong>the</strong> year 2000.<br />

Tab. 5. Major wheat genes for resistance to Hessian fly in Morocco (cf. El Bouhssini et al.,<br />

1996).<br />

Gene Designation Source Genome /<br />

Chromosome location<br />

H5 T. aestivum 1A<br />

H11 T. turgidium 1A<br />

H13 A. squarrosa 6DL<br />

H14 H15 T. turgidium 5A<br />

H21 S. cereale 2BS.2RL<br />

H22 A. squarrosa 1D<br />

H23 A. squarrosa 6D<br />

H25 S. cereale 4AS.4AL-6RL.4AL<br />

H26 A. squarrosa 4D<br />

Tab. 6. Cultivars and lines with resistance to Hessian fly in Morocco.<br />

Line/cultivar Source Mechanism<br />

<strong>of</strong> resistance<br />

Saada T. aestivum Antibiosis<br />

Massira T. aestivum Tolerance<br />

L222/L254 T. aestivum Antibiosis<br />

ADC 14 T. aestivum Antibiosis<br />

Aguilal T. aestivum Antibiosis<br />

Arrihane T. aestivum Antibiosis


106<br />

In addition, we have identified a large number <strong>of</strong> sources <strong>of</strong> resistance in wild relatives;<br />

among this germplasm, There are many accessions <strong>of</strong> Triticum, Aegilops, and Hordeum.<br />

These constitute a very rich gene bank that we can turn to when <strong>the</strong> characterised genes are<br />

no longer effective, and widen <strong>the</strong> genetic base <strong>of</strong> resistance used in <strong>the</strong> breeding program.<br />

Conclusion<br />

Overall, we have identified a large number <strong>of</strong> sources <strong>of</strong> resistance, both in adapted wheat<br />

lines, and in several wild sources. The resistance that was identified is mostly <strong>of</strong> <strong>the</strong> antibiosis<br />

type; plants were not stunted, had normal growth with light green colour, and contained dead<br />

first instars <strong>of</strong> Hessian fly at <strong>the</strong> bases <strong>of</strong> <strong>the</strong>ir stems, which indicates that <strong>the</strong>se larvae died<br />

when <strong>the</strong>y started feeding on <strong>the</strong> plants (antibiosis reaction). Antibiosis is <strong>the</strong> most desirable<br />

form <strong>of</strong> resistance in <strong>the</strong> case <strong>of</strong> host plant resistance to Hessian fly as it is controlled by a<br />

gene for gene relationship.<br />

The selected breeding lines were homogeneous for resistance and presented good<br />

agronomic characters. This indicates, that lots <strong>of</strong> progress has been achieved as far as<br />

resistance to Hessian fly in both bread and durum wheat is concerned in Morocco.<br />

However, because <strong>of</strong> <strong>the</strong> nature <strong>of</strong> Hessian fly genetics, and its ability to develop new<br />

virulent biotypes, we need to stay continuously alerted, and ahead <strong>of</strong> biotype development. In<br />

fact, <strong>the</strong> speed <strong>of</strong> rise <strong>of</strong> new more virulent biotypes depend on <strong>the</strong> size <strong>of</strong> <strong>the</strong> pest population,<br />

<strong>the</strong> number <strong>of</strong> generations, and acreage on which <strong>the</strong> resistant gene has been deployed.<br />

Hessian fly has large populations, with two to three generations per year. Thus a large genetic<br />

variability among individuals <strong>of</strong> <strong>the</strong> population which is <strong>the</strong> material needed for <strong>the</strong> selection<br />

pressure to act on. The development <strong>of</strong> <strong>the</strong>se biotypes is even faster if <strong>the</strong> cultivar carrying<br />

<strong>the</strong> resistance is grown over a large acreage. This will be expected, as improved more<br />

productive varieties have always been adopted by farmers, and replaced older varieties. The<br />

durum and bread wheat improvement program <strong>of</strong> Morocco have to stay ahead <strong>of</strong> new<br />

biotypes by conducting surveys <strong>of</strong> deployed cultivars, estimating <strong>the</strong> % susceptibility in <strong>the</strong><br />

varieties; in o<strong>the</strong>r terms, <strong>the</strong> percent <strong>of</strong> virulence in <strong>the</strong> population.<br />

Also, <strong>the</strong> search for new sources <strong>of</strong> resistance to <strong>the</strong> developing biotype has to start very<br />

early and not wait until <strong>the</strong> biotype erupt. Cultures <strong>of</strong> <strong>the</strong> new virulent individuals have to be<br />

kept in <strong>the</strong> laboratory, and used in <strong>the</strong> search for new sources <strong>of</strong> resistance.<br />

References<br />

Amri, A., El Bouhssini, M., Lhaloui, S., Cox, T.S. & Hatchett, J.H. 1992: Estimates <strong>of</strong> yield<br />

loss due to Hessian fly (Diptera: Cecidomyiidae) on bread wheat using near isogenic<br />

lines. Al Awamia 77: 75-87.<br />

Cartwright, W.B. & LaHue, D.W. 1944: Testing wheats in <strong>the</strong> greenhouse for Hessian fly<br />

resistance. J. Econ. Entomol. 37: 385-387.<br />

El Bouhssini, M., Lhaloui, S., Amri, A., Jlibene, M., Hatchett, J.H., Nssarellah, N. & Nachit,<br />

M.M. 1996: Wheat genetic control <strong>of</strong> Hessian fly (Diptera: Cecidomyiidae) in Morocco.<br />

Field Crops Research 45: 111-114.<br />

Lafever, H.N., Sosa, O., Gallun, R.L., Foster, J.E. & Kuhn, R.C. 1980: Survey monitors <strong>of</strong><br />

Hessian fly population in Ohio wheat. Ohio Report on Research and Development;<br />

Agriculture, Home Economics, Natural Resources 64(4): 51-53.


Lhaloui, S., Buschman, L., El Bouhssini, M., Amri, A., Hatchett, J.H., Keith, D., Starks, K. &<br />

El. Houssaini, K. 1992 a: Infestations <strong>of</strong> Mayetiola spp. (Diptera: Cecidomyiidae) in<br />

bread wheat, durum wheat, and barley: Results <strong>of</strong> five annual surveys in <strong>the</strong> major cereal<br />

growing regions <strong>of</strong> Morocco. Al Awamia 77: 21-53.<br />

Lhaloui, S., Buschman, L., El Bouhssini, M., Starks, K., Keith, D.L. & El. Houssaini, K.<br />

1992 b: Control <strong>of</strong> Mayetiola species (Diptera: Cecidomyiidae) with carb<strong>of</strong>uran in bread<br />

wheat, durum wheat, and barley, with yield loss assessment and its economic analysis. Al<br />

Awamia 77: 55-73.<br />

Lhaloui, S., El Bouhssini, M. & Amri, A. 1998: Survey <strong>of</strong> insect pests <strong>of</strong> cereals in <strong>the</strong><br />

Highlands <strong>of</strong> Morocco in 1996 and 1997. 1986 Annual Research Report. Centre<br />

Aridoculture. Settat, Morocco.<br />

107


108


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 109 - 112<br />

Ecological Pest Management (EPM): General Problems<br />

Wladimir B. Tshernyshev<br />

Department <strong>of</strong> Entomology, Faculty <strong>of</strong> Biology, Moscow State University, Moscow 119899,<br />

Russia<br />

Summary<br />

Strategies <strong>of</strong> EPM and IPM are principally different. EPM uses only such methods which do not upset<br />

<strong>the</strong> natural balance <strong>of</strong> <strong>the</strong> agro-ecosystem and rejects both chemical pesticides and mass inundation by<br />

artificially reared natural enemies. EPM cannot use Economic Injury Level directly. Certain<br />

combinations <strong>of</strong> two parameters, density <strong>of</strong> pest and density <strong>of</strong> its natural enemies (<strong>the</strong> probability <strong>of</strong><br />

pest outbreak), should be a token for short-termed measures in EPM. The advantages, shortcomings<br />

and basic tasks <strong>of</strong> EPM are discussed.<br />

Key words: pests, integrated control, ecological control, natural enemies<br />

What is Ecological Pest Management?<br />

Integrated Pest Management (IPM) incorporates all methods <strong>of</strong> plant protecting, including<br />

both ecological and anti-ecological ones. Ecological Pest Management is <strong>of</strong>ten considered as<br />

only a part <strong>of</strong> IPM (Kogan, 1998).<br />

Proposed by us (Tshernyshev, 1994, 1995, 1996) Ecological Pest Management (EPM) is<br />

a direct result <strong>of</strong> IPM development and includes <strong>the</strong> same methods <strong>of</strong> plant protection except<br />

those which can upset <strong>the</strong> natural balance <strong>of</strong> <strong>the</strong> agro-ecosystem. One would think that we<br />

could better name it Ecological Integrated Pest Management (EIPM - Kozar, 1992; Sokolov<br />

et al., 1994; Sugonyaev & Monastyrsky, 1997) or Ecologically Based Integrated Pest<br />

Management (EBIPM - National Research Council, 1996). Strictly speaking EPM is really an<br />

integration <strong>of</strong> many methods <strong>of</strong> plant protection. However, <strong>the</strong> general strategies <strong>of</strong> EPM and<br />

<strong>the</strong> various types <strong>of</strong> IPM are quite different.<br />

Firstly, EPM bases on keeping <strong>the</strong> natural stability <strong>of</strong> agro-ecosystem, and rejects any<br />

chemical pesticides and o<strong>the</strong>r methods <strong>of</strong> plant protection which can upset natural balance.<br />

One should keep in mind that IPM considers all methods <strong>of</strong> plant protection which are able to<br />

increase <strong>the</strong> yield (Food and Agriculture Organisation, 1975; Kogan, 1998). It is well known<br />

that every hard chemical pesticide is harmful to a certain extent to <strong>the</strong> environment.<br />

Secondly, EPM is not able to use directly <strong>the</strong> backbone <strong>of</strong> IPM, i.e. <strong>the</strong> concept <strong>of</strong><br />

Economic Injury Threshold, because when pest densities have already reached such high<br />

levels, only hard measures are able to save <strong>the</strong> yield. The decision to take measures should be<br />

based on two separate parameters: <strong>the</strong> abundance <strong>of</strong> <strong>the</strong> pest and that <strong>of</strong> its natural enemies.<br />

Every combination <strong>of</strong> <strong>the</strong>se parameters corresponds to definite level <strong>of</strong> probabilities that <strong>the</strong><br />

economic injury level will soon be reached (i.e. possible outbreak development). The<br />

necessary table <strong>of</strong> combinations may be derived from long-term monitoring data <strong>of</strong><br />

arthropods. The admissible risk will depend on <strong>the</strong> kind <strong>of</strong> crop and geographical region.<br />

109


110<br />

EPM and IPM from <strong>the</strong> point <strong>of</strong> view <strong>of</strong> population dynamics<br />

Accordingly to <strong>the</strong> general <strong>the</strong>ory <strong>of</strong> population dynamics (Victorov, 1976; Isayev et al.,<br />

1984), <strong>the</strong> level <strong>of</strong> any population density usually varies in definite limits, being regulated by<br />

predators and non-specialised parasites. Sometimes <strong>the</strong>se predators and parasites can not hold<br />

<strong>the</strong> reproduction <strong>of</strong> pest population and so-called "slipping out" takes place (pest outbreak).<br />

Specialised parasites are usually effective only at <strong>the</strong> highest levels <strong>of</strong> <strong>the</strong> pest populations.<br />

The level <strong>of</strong> an economic injury threshold is usually higher than <strong>the</strong> level <strong>of</strong> slipping out.<br />

The general task <strong>of</strong> EPM is to prevent <strong>the</strong> pest population from slipping out <strong>of</strong> <strong>the</strong> control <strong>of</strong><br />

<strong>the</strong>ir natural enemies. If <strong>the</strong> natural enemies are always abundant we can forget about<br />

pesticides. In IPM we also use so-called « levels <strong>of</strong> natural enemies efficacy». It means that if<br />

<strong>the</strong>re are no more than «n» specimens <strong>of</strong> a given pest species per one predator (parasite) in <strong>the</strong><br />

field, you should not use pesticides even when <strong>the</strong> abundance <strong>of</strong> <strong>the</strong> pest exceeds <strong>the</strong><br />

economic injury threshold. However, <strong>the</strong> backbone <strong>of</strong> every IPM concept is <strong>the</strong> economic<br />

injury level. It means that in practice, we may wait until <strong>the</strong> pest density is very high and <strong>the</strong>n<br />

use pesticides.<br />

The management <strong>of</strong> natural enemies and <strong>the</strong>ir alternative prey species<br />

The main measure in EPM is a management <strong>of</strong> natural enemies. Grassy margins around <strong>the</strong><br />

field, meadows, meliorate ditches, hedges, forest belts, forest edges and fields <strong>of</strong> perennial<br />

herbs may be good refuges for predators and parasites. We should note that only species<br />

inhabiting relatively open spaces (usually ecotones) are able to colonise <strong>the</strong> field, forest<br />

species cannot do it. The stability <strong>of</strong> an agro-ecosystem can also be increased by <strong>the</strong><br />

introduction <strong>of</strong> plant polycultures, i.e. by sowing different plants (especially flowering<br />

cultural plants) on separate strips and by placing "cassettes" with any artificially reared «food<br />

insects» in <strong>the</strong> field to attract natural enemies (Kovalenkov & Tjurina, 1993). If <strong>the</strong> slipping<br />

out <strong>of</strong> any pest is outlined, a partial mowing <strong>of</strong> <strong>the</strong> nearest grassy field margins, meadows or<br />

perennial herbs can increase <strong>the</strong> natural enemies density in <strong>the</strong> field (Nyazov, 1992;<br />

Khamrayev, 1992).<br />

All natural enemies should be provided with food during periods when <strong>the</strong>re are no pests<br />

in <strong>the</strong> field. However, <strong>the</strong> alternative prey species also need <strong>the</strong>ir host plants. Therefore we<br />

have to regulate complexes <strong>of</strong> wild plants near <strong>the</strong> crop fields.<br />

Short-term measures in EPM<br />

In EPM we have at least four barriers for pests in <strong>the</strong> field. 1) Natural predators (insects,<br />

spiders, mites, may be harvested). They are universal and can effectively work at every level<br />

<strong>of</strong> pest abundance. Many non-specialised parasites also belong to this category <strong>of</strong> natural<br />

enemies. This barrier is especially important. 2) If <strong>the</strong> monitoring shows that <strong>the</strong> abundance<br />

<strong>of</strong> natural enemies is not sufficient, we can increase <strong>the</strong>ir number by mowing <strong>the</strong> nearest<br />

perennial herbs, meadows or grassy margins. 3) Next step is a release <strong>of</strong> some artificially<br />

reared enemies. However, too many mass releases may be dangerous for <strong>the</strong> stability <strong>of</strong> an<br />

agro-ecosystem due to competition with natural enemies and eradication <strong>of</strong> alternative preys<br />

(hosts). 4) In many cases <strong>the</strong> s<strong>of</strong>t (microbiological and any similar) pesticides may be used.<br />

The limitations are <strong>the</strong> same as in <strong>the</strong> previous case.<br />

Therefore, in contrast to IPM, an EPM strategy demands <strong>the</strong> interruption <strong>of</strong> <strong>the</strong> outbreak<br />

development in its very beginning when <strong>the</strong> pest abundance level is lower than <strong>the</strong> economic<br />

injury threshold.


111<br />

Advantages and shortcomings <strong>of</strong> EPM<br />

Advantages <strong>of</strong> EPM are <strong>the</strong> following: 1) no pollution <strong>of</strong> environment and food products; 2)<br />

no problems <strong>of</strong> resistance to chemical pesticides, 3) supporting <strong>of</strong> maximal biodiversity and<br />

conservation <strong>of</strong> rare species; 4) favourable conditions for natural pollinators; 4) long-term<br />

effect <strong>of</strong> agro-landscape management; 6) low expenses.<br />

The shortcomings <strong>of</strong> EPM are: 1) The main hindrance is a psychological barrier. All<br />

persons engaged in production are sure that it is impossible to grow <strong>the</strong> yield without<br />

chemical pesticides; 2) relatively lower reliability, especially in <strong>the</strong> beginning <strong>of</strong> landscape<br />

management procedure, <strong>the</strong>refore <strong>the</strong> replacement <strong>of</strong> <strong>the</strong> IPM by <strong>the</strong> EPM should be made<br />

gradually; 3) much more complex monitoring schemes for both <strong>the</strong> pests and <strong>the</strong> beneficial<br />

arthropods, including also <strong>the</strong>ir alternative preys (hosts), both in <strong>the</strong> field and around it (field<br />

margins and surrounding biotopes). Such monitoring should be made by well-educated<br />

specialists. Indeed, in many cases estimation <strong>of</strong> population density only <strong>of</strong> well distinguished<br />

species (indicator species - Sugonyaev & Monastyrsky, 1997) is sufficient; 4) elaboration <strong>of</strong><br />

<strong>the</strong> EPM-system for such pests which can also decrease <strong>the</strong> quality <strong>of</strong> production for instance<br />

for bugs Eurygaster integriceps Put., and for many orchard pests. In this case <strong>the</strong> economic<br />

injury level may be very close to level <strong>of</strong> slipping out; 5) elaboration <strong>of</strong> EPM-system for long<br />

distance migratory pests like some locust species; 6) elaboration such system for new alien<br />

species like <strong>the</strong> Colorado beetle, because <strong>the</strong> complex <strong>of</strong> <strong>the</strong>ir natural enemies has not been<br />

formed finally so far.<br />

The main directions <strong>of</strong> scientific research in EPM elaboration<br />

The main topics and stages <strong>of</strong> research are <strong>the</strong> following: 1) to determine <strong>the</strong> dominating<br />

species <strong>of</strong> predators and parasites in <strong>the</strong> field and <strong>the</strong> spectrum <strong>of</strong> <strong>the</strong>ir preys (pests and<br />

inhabitants <strong>of</strong> weeds and wild plants); 2) to elaborate <strong>the</strong> management <strong>of</strong> alternative preys<br />

(hosts) outside <strong>the</strong> field; 3) to provide additional food sources for entomophagous arthropods,<br />

e.g. nectar, pollen and honeydew, and to determine <strong>the</strong>ir preferred flowering plants; 3) to<br />

study interactions between natural enemies; 4) to investigate seasonal colonisation <strong>of</strong> <strong>the</strong> field<br />

by natural enemies, optimal places for <strong>the</strong>ir hibernation, aestivation and reproduction; 5)<br />

supporting <strong>of</strong> field margins on <strong>the</strong> stage <strong>of</strong> succession optimal for natural enemies; 6) to study<br />

interactions between natural enemies and artificially reared and released predators and<br />

parasites, microbiological and chemical pesticides; and 7) construction <strong>of</strong> optimal types <strong>of</strong><br />

agro-landscapes.<br />

Conclusion<br />

Ecological Pest Management is supposed to replace <strong>the</strong> concept <strong>of</strong> Integrated Pest<br />

Management (IPM) in <strong>the</strong> near future. We think that namely cereal crops will provide <strong>the</strong> best<br />

opportunity to elaborate <strong>the</strong> first EPM system. We suppose also that such a system <strong>of</strong> natural<br />

balance in agro-ecosystem may be also feasible for disease and weed control.<br />

References<br />

Food and Agriculture Organisation 1975: Report FAO Panel <strong>of</strong> Experts on Integrated Pest<br />

Control, 5 th October 15-25, 1974, Rome, Italy: FAO- UN. Meeting report 1975/m/2,<br />

41 pp. (cit. by M. Kogan, 1998)<br />

Isajev, A.S., Khlebopros, R.G., Nedorjezov, L.V., Kondakov, Ju, P. & Kiseljev, V.V. 1984:<br />

Population dynamics in forest insects. Novosibirsk, Nauka, 224 pp. (in Russian).


112<br />

Khamrajev, A.S., 1992: [Anthropogenous influence on <strong>the</strong> dominating complex <strong>of</strong> pests and<br />

entomophaguos insects in ecosystem <strong>of</strong> cotton crop in south-west <strong>of</strong> Uzbekistan.]<br />

Doklady <strong>of</strong> Acad. Sci. Uzbekistan 10-11: 85-87 (in Russian).<br />

Kogan, M. 1998: Integrated Pest Management: Historical Perspectives and Contemporary<br />

Developments. Ann. Rev. Entomol. 43: 243-270.<br />

Kovalenkov, V.G. & Tjurina, N.M. 1993: [Is it possible to rule <strong>the</strong> activity <strong>of</strong> entomophagous<br />

insects ?] Plant Protect. 8: 7-8 (in Russian).<br />

Kozar, F., 1992: Ecological plant protection in Hungary. In: Proc. <strong>of</strong> <strong>the</strong> Congress Agriculture<br />

and Environment in Eastern Europe and <strong>the</strong> Ne<strong>the</strong>rlands. Agric. University, Wageningen:<br />

283-291.<br />

National Research Council 1996: Ecologically Based Pest Management. New Solution for a<br />

New Century. Washington, DC. Natl. Acad. 160 pp.<br />

Nyazov, O.D. 1992: [Ecological principles <strong>of</strong> cotton crop protection.] Izvestia Acad. Sci.,<br />

Turkmenistan: 3-13 (in Russian).<br />

Sokolov, M.S., Monastyrsky, O.A. & Pikushova, E.A. 1994: [Application <strong>of</strong> ecological<br />

principles to plant protection.] Pushchino, RA Agric., 477 pp. (in Russian).<br />

Sugonyaev, E.S. & Monastyrsky, A.L. 1997: [Introduction to <strong>the</strong> management <strong>of</strong> insect<br />

populations - rice pests in Vietnam.] Khanoy, 291 pp. (in Russian).<br />

Tshernyshev, W.B. 1994: [Ecological plant protection.] Plant Protect. 8: 46-47 (in Russian).<br />

Tshernyshev, W.B. 1995: Ecological pest management: general approaches. J. Appl. Ent. 119:<br />

379-381.<br />

Tshernyshev, W.B. 1996: [Insect Ecology.] Moscow, Publ. House <strong>of</strong> Moscow University: 297<br />

pp. (in Russian).<br />

Victorov, G.A. 1976: [Ecology <strong>of</strong> entomophagous parasitic insects.] Moscow, Nauka, 127 pp.<br />

(in Russian).


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 113 - 118<br />

Hessian fly (Mayetiola destructor Say) damage in relay intercropping <strong>of</strong><br />

cereals in Finland<br />

Erja Huusela-Veistola, Arja Vasarainen and Jaana Grahn<br />

Agricultural Research Centre <strong>of</strong> Finland, Plant Production Research, FIN-31600 Jokioinen,<br />

Finland<br />

Summary<br />

Relay intercropping is a double-cropping system where a ‘cover crop’ (spring cereal or oil-seed rape)<br />

and a winter cereal are usually sown simultaneously in <strong>the</strong> spring. The ‘cover crop’ is harvested in <strong>the</strong><br />

first year and <strong>the</strong> winter cereal in <strong>the</strong> second year. Relay intercropping saves time and energy,<br />

improves <strong>the</strong> soil properties and keeps <strong>the</strong> fields green throughout <strong>the</strong> year. Severe damage <strong>of</strong> winter<br />

cereals has occurred in relay intercropping fields in SW Finland. Hessian fly, Mayetiola destructor<br />

Say (Diptera: Cecidomyiidae), has damaged several intercropped winter cereals so heavily that many<br />

farmers have given up <strong>the</strong> new promising cropping system. To date, damage caused by <strong>the</strong> Hessian fly<br />

in Finland has been overlooked although damage in rye and barley has been reported as early as in <strong>the</strong><br />

early 1900s. Since 1995, Hessian fly infestations have occurred especially in spring-sown winter<br />

wheat, but also in rye and barley in relay intercropped fields. In 1999, Hessian fly infestations were<br />

found also in ordinary spring cereal (barley, wheat) and winter wheat fields. In <strong>the</strong> USA, resistant<br />

wheat varieties are being used to decrease <strong>the</strong> damage caused by <strong>the</strong> Hessian fly. Delayed sowing<br />

time, crop rotation and tillage may also decrease crop losses. However, in relay intercropping, springsown<br />

winter cereals are more and longer susceptible to <strong>the</strong> Hessian fly attack because <strong>the</strong>ir growth is<br />

retarded. In addition, it is difficult or impossible to target pesticides to larvae inhabiting <strong>the</strong> base <strong>of</strong><br />

undersown winter cereals. The importance <strong>of</strong> Hessian fly as a pest in relay intercropping in Finland is<br />

discussed.<br />

Key words: double-cropping, Cecidomyiidae, Hessian fly, integrated control, intercropping<br />

Introduction<br />

The bird cherry-oat aphid (Rhopalosiphum padi L.) is <strong>the</strong> main pest <strong>of</strong> spring cereals in<br />

Finland. Generally, aphid outbreaks occur 2-3 times per decade. Thus, insecticides are not<br />

used every year in large scale in cereal production in Finland. The orange wheat blossom<br />

midge (Sitodiplosis mosellana [Géhin]) and <strong>the</strong> frit fly (Oscinella frit L.) occasionally cause<br />

local problems to winter cereals.<br />

Changes in farming systems, e.g. trends in conservation tillage and non-tillage, may make<br />

<strong>the</strong> conditions more suitable for new or previously minor pest species. Integrated or<br />

sustainable cereal production may have different pest problems than conventional production<br />

methods. Usually, large monoculture fields favour pests than more diversified intercropped<br />

fields (Andow, 1991).<br />

Relay intercropping<br />

Today, winter-green fields have become popular because <strong>of</strong> <strong>the</strong> environmental support<br />

provided by <strong>the</strong> EU. Relay intercropping is a double-cropping system where a ‘cover crop’<br />

(spring cereal or oil-seed rape) and a winter cereal are usually sown simultaneously in <strong>the</strong><br />

spring. The ‘cover crop’ is harvested in <strong>the</strong> first year and <strong>the</strong> winter cereal in <strong>the</strong> second year.<br />

113


114<br />

Relay intercropping saves time and energy, improves <strong>the</strong> soil properties and keeps <strong>the</strong> fields<br />

green throughout <strong>the</strong> year. Moreover, <strong>the</strong> sowing conditions are better and more certain in <strong>the</strong><br />

spring than in <strong>the</strong> autumn in Finland. Severe damage <strong>of</strong> winter cereals has occurred in relay<br />

intercropping fields in SW Finland. Hessian fly, Mayetiola destructor (Say), has damaged<br />

several intercropped winter cereals so heavily that many farmers have given up <strong>the</strong> promising<br />

new cropping system.<br />

Hessian fly<br />

Distribution<br />

The Hessian fly is a member <strong>of</strong> <strong>the</strong> dipteran family Cecidomyiidae which includes several<br />

pests <strong>of</strong> crop plants. It is an important pest <strong>of</strong> wheat in <strong>the</strong> USA, Canada, New Zealand and<br />

Morocco (Gagne et al., 1991; Harris et al., 1996). Although <strong>the</strong> Hessian fly is native to<br />

Europe it causes only minor crop losses <strong>the</strong>re (Skuhrava et al., 1984).<br />

To date, damage caused by <strong>the</strong> Hessian fly in Finland has been overlooked although<br />

damage in rye and barley has been reported as early as in <strong>the</strong> early 1900s (Vappula, 1965).<br />

Since 1995, Hessian fly infestations have occurred especially in spring-sown winter wheat,<br />

but also in rye and barley in relay intercropped fields (Fig. 1). In 1999, Hessian fly<br />

infestations were found also in ordinary spring cereal (barley, wheat) and winter wheat fields.<br />

Fig. 1. Damage caused by Hessian fly in intercropped fields on practical (circle) and<br />

experimental (square) farms in Finland.


115<br />

Life cycle<br />

Adult Hessian flies resemble small mosquitoes. They are weak fliers and only live for about<br />

three days. Soon after mating <strong>the</strong> female deposits reddish 0.4-0.5-mm-long eggs on <strong>the</strong> upper<br />

surface <strong>of</strong> leaves <strong>of</strong> host plants. The eggs hatch in 3-10 days, depending on <strong>the</strong> temperature.<br />

The first-stage larvae, which are first reddish but soon become white, move down to <strong>the</strong> base<br />

<strong>of</strong> <strong>the</strong> leaf. The two later larval stages are immobile and <strong>the</strong>y feed on plant juice, burrowing<br />

between <strong>the</strong> leaf sheath and <strong>the</strong> stem. A puparium or ‘flaxseed’ develops after two to four<br />

weeks. The Hessian fly overwinters as larvae or pupae inside <strong>the</strong> puparium.<br />

The Hessian fly may have from one to six (basically two) generations per year (Barnes,<br />

1956). The number <strong>of</strong> generations depends on <strong>the</strong> temperature, moisture and food plants<br />

available (Barnes, 1956). Especially <strong>the</strong> length <strong>of</strong> <strong>the</strong> pupal stage varies and, <strong>the</strong>refore, brood<br />

may be a better term than generation (Buntin & Chapin, 1990). In practice, <strong>the</strong>re are several<br />

overlapping broods per year; all life stages can be found simultaneously in <strong>the</strong> summertime in<br />

Finland.<br />

Damage<br />

Wheat and barley are favourite hosts <strong>of</strong> <strong>the</strong> Hessian fly, but it can feed on rye, triticale and<br />

many o<strong>the</strong>r grasses, e.g. Agropyron spp., Elymus spp. and Phleum spp., as well (Jones, 1939;<br />

Barnes 1956; Johnson et al., 1987). Oats seems to be resistant to Hessian fly (Morrill, 1982).<br />

The injury is caused by <strong>the</strong> larvae which feed near <strong>the</strong> base <strong>of</strong> <strong>the</strong> plant. Because Hessian fly<br />

is a sapfeeder, larvae cause little apparent physical damage to <strong>the</strong> plant tissue. Hessian fly<br />

infestation causes crop losses in several ways: 1) by killing or severely stunting seedlings and<br />

tillers, 2) by breaking <strong>the</strong> stems <strong>of</strong> mature plants and 3) by reducing <strong>the</strong> grain size and number<br />

(Barnes, 1956). Damage is related to <strong>the</strong> degree <strong>of</strong> infestation and <strong>the</strong> growth stage <strong>of</strong> <strong>the</strong> host<br />

plant. In some cases, plant tillering may partly compensate for <strong>the</strong> damage caused by <strong>the</strong><br />

Hessian fly even if <strong>the</strong> main stem was destroyed. Tillering response varies between wheat<br />

cultivars (Wellso & Hoxie, 1994).<br />

Crop protection<br />

Crop protection recommendations are based on those applied in <strong>the</strong> USA where Hessian fly is<br />

one <strong>of</strong> <strong>the</strong> most destructive pests <strong>of</strong> wheat.<br />

Resistant varieties<br />

In <strong>the</strong> USA, several resistant wheat varieties are being used to decrease <strong>the</strong> damage caused by<br />

<strong>the</strong> Hessian fly. A total 27 resistant genes have been found (Ratcliffe & Hatchett, 1997).<br />

Resistance is based on larval antibiosis, which means that larvae cannot live and develop in<br />

resistant plants and die in a few days. However, due to selective pressure, more virulent<br />

biotypes <strong>of</strong> Hessian fly, which are able to overcome plant resistance, are becoming common.<br />

Fly-free sowing dates<br />

With late sowing <strong>of</strong> winter wheat, damage caused by <strong>the</strong> Hessian fly may be avoided (Buntin<br />

et al., 1990). The recommended fly-free sowing dates applied in <strong>the</strong> USA are based on<br />

rainfall and temperature conditions that are connected to <strong>the</strong> occurrence <strong>of</strong> Hessian fly adults.<br />

Cultural practices<br />

The most important cultural controlling method <strong>of</strong> Hessian fly is crop rotation: continuous<br />

growing <strong>of</strong> wheat in <strong>the</strong> same field should be avoided. Moreover, deep ploughing and<br />

destruction <strong>of</strong> volunteer wheat and old crop residues are recommended.


116<br />

Chemical control<br />

Use <strong>of</strong> insecticides is not recommended because it is not usually economically feasible.<br />

Sprays should be applied before peak egg deposition (Buntin & Hudson, 1991). If Hessian fly<br />

has infested a field, chemical control is not effective anymore. Larvae are not susceptible to<br />

systemic insecticides once <strong>the</strong>y have moved to <strong>the</strong> leaf base. A single application <strong>of</strong><br />

insecticide is usually not enough, and repeated applications are economically not feasible<br />

(Buntin & Hudson, 1991). If fly problems are perennial, prophylactic systemic granular<br />

insecticides (phorate, disulfoton, carb<strong>of</strong>uran) may be applied at planting (Morrill & Nelson,<br />

1975; Buntin, 1992; Buntin et al., 1992).<br />

Natural enemies<br />

Although over 40 parasitoid species have been found in <strong>the</strong> Hessian fly (MacFarlane, 1989),<br />

<strong>the</strong> control effect <strong>of</strong> natural enemies is questionable. Homoporus destructor (Say) (Hym.:<br />

Pteromalidae), Eupelmus allynii (French) (Hym.: Eupelmidae) and Platygaster hiemalis<br />

(Forbes) (Hym.: Platigasteridae) are <strong>the</strong> most common parasitic wasps <strong>of</strong> Hessian fly in <strong>the</strong><br />

USA (Hill, 1953; Lidell et al., 1987). Usually, <strong>the</strong> natural enemies are not effective enough to<br />

prevent crop losses but <strong>the</strong>y may reduce fly populations.<br />

Crop protection problems in relay intercropping<br />

Relay intercropping <strong>of</strong> cereals favour <strong>the</strong> Hessian fly. Spring-sown winter cereals are more<br />

and longer susceptible to <strong>the</strong> Hessian fly attack because <strong>the</strong>ir growth is retarded. Hessian fly<br />

overwinter well in intercropped fields because no crop rotation or tillage is used every year.<br />

Eggs and pupa are not susceptible to insecticides. There are no distinct flying periods because<br />

<strong>of</strong> overlapping generations, which makes <strong>the</strong> control <strong>of</strong> flying or ovipositing adults difficult.<br />

In addition, it is nearly impossible to target pesticides to larvae inhabiting <strong>the</strong> base <strong>of</strong> undersown<br />

winter cereals. Chemical control <strong>of</strong> Hessian fly was tried out on a farm in Kuusjokí in<br />

1997. An intercropped (barley+winter wheat) field was sprayed with dimethoate and labdacyhalothrin<br />

three times (13 June, 28 June and 18 July), but <strong>the</strong> number <strong>of</strong> Hessian fly pupae<br />

did not differ between <strong>the</strong> sprayed and <strong>the</strong> control field. The granular insecticides used in <strong>the</strong><br />

USA are not available in Finland. The wheat varieties used in Finland are not known to be<br />

resistant to Hessian fly.<br />

Hessian fly as a pest in relay intercropping – preliminary results<br />

Preliminary studies to optimise <strong>the</strong> relay intercropping system have been conducted at<br />

Agricultural Research Centre <strong>of</strong> Finland (MTT) since 1998. Different cover crops and winter<br />

cereals have been tested. Moreover, different seed rates have been compared to find out an<br />

optimal seed mixture.<br />

Different ‘cover crops’ (barley, oats and oil-seed rape) and seed rates were tested in<br />

Ypäjä in 1998-1999 (Fig. 1). Although oil-seed rape and oats are resistant to Hessian fly, <strong>the</strong>ir<br />

use as ‘cover crops’ did not prevent infestation <strong>of</strong> spring sown winter cereals by <strong>the</strong> Hessian<br />

fly.<br />

Different winter cereals (wheat, rye, triticale) were compared in three locations<br />

(Mietoinen, Pälkäne and Ylistaro; see Fig. 1) in 1998-1999. The number <strong>of</strong> Hessian fly pupae<br />

was higher in spring-sown wheat and rye than in triticale in <strong>the</strong> autumn samples <strong>of</strong> Mietoinen.<br />

However, spring-sown triticale did not overwinter in any location. In <strong>the</strong> next spring, Hessian<br />

fly pupae were found in spring-sown winter wheat and rye but not in autumn-sown winter<br />

cereals. Unfortunately, yield results from <strong>the</strong> experiments are not yet available.<br />

Important practical knowledge was acquired during <strong>the</strong> experiments. Perennial weeds<br />

should be avoided in relay intercropped field. Crop residues should be managed carefully


117<br />

before <strong>the</strong>y smo<strong>the</strong>r winter cereals. Finding combinations <strong>of</strong> crops and varieties suitable for<br />

certain conditions increases <strong>the</strong> chances <strong>of</strong> successful relay intercropping.<br />

Sometimes, under favourable conditions, yields <strong>of</strong> winter cereals in relay intercropping<br />

have been fairly good in spite <strong>of</strong> <strong>the</strong> moderate numbers <strong>of</strong> Hessian flies. Never<strong>the</strong>less, relay<br />

intercropping cannot be recommended before damage caused by <strong>the</strong> Hessian fly to winter<br />

cereals can be prevented or decreased. An ecologically and economically feasible solution to<br />

<strong>the</strong> problem is not yet available. In future work, <strong>the</strong> susceptibility <strong>of</strong> different wheat and rye<br />

varieties practicable in Finland should be tested. Moreover, <strong>the</strong> importance <strong>of</strong> adjacent fields<br />

and uncultivated areas for <strong>the</strong> success and phenology <strong>of</strong> <strong>the</strong> Hessian fly in Finland should be<br />

studied.<br />

References<br />

Andow, D.A. 1991: Vegetational diversity and arthropod population response. Annu. Rev.<br />

Entomol. 36: 561-586.<br />

Barnes, H.F. 1956: Gall midges <strong>of</strong> economic importance: gall midges <strong>of</strong> cereal crops. Vol<br />

VII. London; Grosby Lockwood and Son. 261 pp.<br />

Buntin, G.D. 1992: Assessment <strong>of</strong> a microtube injection system for applying systemic<br />

insecticides at planting for Hessian fly control in winter wheat. Crop. Prot. 11: 366-370.<br />

Buntin, G.D & Chapin, J.W. 1990: Biology <strong>of</strong> Hessian fly (Diptera: Cecidomyiidae) in <strong>the</strong><br />

Sou<strong>the</strong>astern United States: geographic variation and temperature-dependent phenology.<br />

J. Econ. Entomol. 83: 1015-1024.<br />

Buntin, G.D. & Hudson, R.D. 1991: Spring control <strong>of</strong> <strong>the</strong> Hessian fly (Diptera: Cecidomyiidae)<br />

in winter wheat using insecticides. J. Econ. Entomol. 84: 1913-1919.<br />

Buntin, G.D., Bruckner, P.L. & Johnson, J.W. 1990: Management <strong>of</strong> Hessian fly (Diptera:<br />

Cecidomyiidae) in Georgia by delayed planting <strong>of</strong> winter wheat. J. Econ. Entomol. 83:<br />

1025-1033.<br />

Buntin, G.D., Ott, S.L. & Johnson, J.W. 1992: Integration <strong>of</strong> plant resistance, insecticides,<br />

and planting date for management <strong>of</strong> <strong>the</strong> Hessian fly(Diptera: Cecidomyiidae) in winter<br />

wheat. J. Econ. Entomol. 85: 530-538.<br />

Gagne, R.J., Hatchett, J.H., Lhaloui, S & EL Bouhssini, M. 1991: Hessian fly and Barley<br />

stem gall midge, two different species <strong>of</strong> Mayetiola (Diptera: Cecidomyiidae) in<br />

Morocco. Ann. Entomol. Soc. America 84: 436-443.<br />

Harris, M.O., Dando, J.L., Griffin, W. & Madie, C. 1996: Susceptibility <strong>of</strong> cereal and noncereal<br />

grasses to attack by Hessian fly (Mayetiola destructor). New Zealand J. Crop &<br />

Hortic. Sci. 24: 29-238.<br />

Hill, C.C. 1953: Parasites <strong>of</strong> <strong>the</strong> Hessian fly in <strong>the</strong> North Central States. United States<br />

Department <strong>of</strong> Agriculture, Circular 923, 15 pp.<br />

Johnson, J.W., Buntin, G.D., Foster, J.E., Roberts, J.J. & Raymer, P.L. 1987: Response <strong>of</strong><br />

triticale to <strong>the</strong> Hessian fly (Diptera: Cecidomyiidae). J. Entomol. Sci. 22: 51-54.<br />

Jones, E.T. 1939: Grasses <strong>of</strong> <strong>the</strong> tribe Hordeae as hosts <strong>of</strong> Hessian fly. J. Econ. Entomol. 32:<br />

505-510.<br />

Lidell, M.C., Schuster, M.F. & Turney, H.A. 1987: Biology and distribution <strong>of</strong> <strong>the</strong> Hessian<br />

fly and its parasitoids in Texas. Progress Report - Texas Agric. Exp. Station 4530, 7 pp.<br />

MacFarlane, R.P. 1989: Mayetiola destructor (Say), Hessian fly (Diptera: Cecidomyiidae).<br />

Techn. Commun. Commonw. Inst. Biol. Control 10: 101-104.<br />

Morrill, W.L. 1982: Hessian fly: host selection and behaviour during oviposition, winter<br />

biology, and parasitoids. J. Georgia Entomol. Soc. 17: 150-156.


118<br />

Morrill, W.L. & Nelson, L.R. 1975: Hessian fly control with carb<strong>of</strong>uran. J. Econ. Entomol.<br />

69: 123-124.<br />

Ratcliffe, R.H. & Hatchett, J.H. 1997: Biology and genetics <strong>of</strong> <strong>the</strong> Hessian fly and resistance<br />

in wheat. In: New Development in Entomology. Bondari, K. (ed.). Research Signpost,<br />

Scientific Information Guild, Trivandrum, India: 47-56.<br />

Skuhrava, M., Skuhravy, V. & Brewer, J.W. 1984: The distribution and long-term changes in<br />

population dynamics <strong>of</strong> gall midges on cereals in Europe. Cecidollogia Intern. 5: 1-7.<br />

Vappula, N.A. 1965: Pests <strong>of</strong> cultivated plants in Finland. Acta Entomologia Fennica 19.<br />

Wellso, S.G. & Hoxie, R.P. 1994: Tillering response <strong>of</strong> ‘Monon’ and ‘Newton’ winter wheats<br />

infested biotype L Hessian fly (Diptera: Cecidomyiidae) larvae. The Great Lakes<br />

Entomologist 24: 235-239.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 119 - 127<br />

Results <strong>of</strong> a nation-wide survey <strong>of</strong> spider assemblages in Hungarian cereal<br />

fields<br />

Ferenc Samu 1 , Ferenc Tóth 2 , Csaba Szinetár 3 , Géza Vörös 4 and Erika Botos 1<br />

1 Plant Protection Institute, Hungarian Academy <strong>of</strong> Sciences, PO Box 102, Budapest,<br />

H-1525 Hungary<br />

2 Department <strong>of</strong> Plant Protection, Gödöllõ University <strong>of</strong> Agricultural Sciences, 1 Páter K. u.,<br />

Gödöllõ, H-2103 Hungary<br />

3 Department <strong>of</strong> Zoology, Berzsenyi College, PO Box 170 Szomba<strong>the</strong>ly, H-9701 Hungary<br />

4 Plant Protection and Soil Conservation Station, Tolna County, Szekszárd, H-7100 Hungary<br />

Summery<br />

Within <strong>the</strong> framework <strong>of</strong> a national survey on spider assemblages <strong>of</strong> agricultural areas, and cooperating<br />

with <strong>the</strong> research project "IPM <strong>of</strong> cereals", since 1992 <strong>the</strong> arachn<strong>of</strong>auna <strong>of</strong> cereals has been<br />

systematically investigated in Hungary. Samples by pitfall trapping and/or suction sampling took place<br />

at 16 different locations, representing five smaller regions <strong>of</strong> Hungary. As a result over 10,000<br />

specimens, representing more than 150 species were caught. The present paper analyses <strong>the</strong><br />

similarities and differences between <strong>the</strong> spider assemblages found in <strong>the</strong> various fields, regarding<br />

dominant species, diversity and community structure. Data analyses indicate, that cereal spider<br />

communities in Hungary are dominated by a well defined and limited set <strong>of</strong> species, which are more or<br />

less ubiquitous in every field. These most dominant and widespread agrobiont species are: Pardosa<br />

agrestis, Meioneta rurestris, Oedothorax apicatus, Pachygnatha degeeri, Tibellus oblongus. Apart<br />

from <strong>the</strong> agrobionts, cereal fields had a diverse spider community, reaching an estimated species<br />

richness <strong>of</strong> 110 species at several locations. Indicator species analysis showed, that for <strong>the</strong> cereal<br />

fields, as such, <strong>the</strong>re were no strictly specific indicator spider species. Considering arable fields in<br />

general (i.e. cereals plus alfalfa) <strong>the</strong>se habitats had few indicator species when compared to natural<br />

grassy habitats. These species are mostly native to wetland habitats. We propose that <strong>the</strong>y could<br />

became widespread in agricultural habitats, because through dispersal and life history characteristics<br />

<strong>the</strong>y became pre-adapted to <strong>the</strong> ephemeral conditions <strong>of</strong> agricultural fields.<br />

Key words: spider, natural enemy, biological control, cereal, alfalfa, fauna origin, wetlands<br />

Introduction<br />

Spiders are one <strong>of</strong> <strong>the</strong> most important predatory group <strong>of</strong> arthropods in cereal ecosystems<br />

(Nyffeler & Benz, 1987). There are a number <strong>of</strong> studies <strong>of</strong> <strong>the</strong> cereal arachn<strong>of</strong>auna in Western<br />

Europe and North America which revealed that this special and strongly human influenced<br />

vegetation type can accommodate a great diversity and density <strong>of</strong> spiders. Despite <strong>the</strong><br />

potential <strong>of</strong> cereal fields to harbour a rich spider fauna, individual fields show great variation.<br />

In some fields <strong>the</strong> spiders were as diverse and abundant as in natural habitats (T<strong>of</strong>t, 1989),<br />

while from o<strong>the</strong>r fields impoverished and sparse spider communities were reported (Nyffeler<br />

et al., 1994).<br />

In Hungary prior to <strong>the</strong> studies reported here, <strong>the</strong> cereal spider community had not been<br />

studied. To obtain basic information about <strong>the</strong> spider communities in Hungarian cereal fields,<br />

and o<strong>the</strong>r agricultural areas, several studies were conducted in parallel from <strong>the</strong> early 1990's<br />

119


120<br />

onwards. The present paper provides a meta-analysis <strong>of</strong> <strong>the</strong>se studies, aiming at answering <strong>the</strong><br />

questions:<br />

1. Which are <strong>the</strong> dominant species <strong>of</strong> spiders in cereals?<br />

2. Are dominant species <strong>the</strong> same across cereal spider communities; are <strong>the</strong>re regional<br />

differences present?<br />

3. What are <strong>the</strong> structural characteristics <strong>of</strong> cereal spider communities; how similar are<br />

cereal spider communities to o<strong>the</strong>r arable and comparable natural communities?<br />

4. Which are <strong>the</strong> indicator species, that could signify <strong>the</strong> similarities or dissimilarities<br />

between <strong>the</strong> cereal and o<strong>the</strong>r habitats examined?<br />

Material and methods<br />

The present paper ga<strong>the</strong>rs data from several studies, which were part <strong>of</strong> two broad projects:<br />

"The survey <strong>of</strong> Hungarian agricultural spider fauna" (project leader: Plant Protection Institute,<br />

HAS), and "IPM <strong>of</strong> cereals" (project leader: Plant Protection Department <strong>of</strong> <strong>the</strong> Gödöllõ<br />

University). Within <strong>the</strong> framework <strong>of</strong> <strong>the</strong>se projects cereal fields were sampled at 16 different<br />

locations, representing five smaller regions <strong>of</strong> Hungary. Investigations reported here took<br />

place between 1992 and 1998, each lasting between one to three years. We have included<br />

only those sites into <strong>the</strong> analysis, where in total more than 300 spider individuals were caught<br />

(Tab. 2).<br />

Two standard collecting methods were applied in <strong>the</strong> studies. In four studies spiders were<br />

sampled by a hand-held suction sampler (Samu & Sárospataki, 1995), which proved to be<br />

very efficient in collecting spiders, although it is biased towards <strong>the</strong> juvenile stages and<br />

towards <strong>the</strong> foliage dwelling and/or web building species (Samu et al., 1997). At six study<br />

sites pitfall trapping was used to collect spiders; and <strong>the</strong>re were only two sites where both<br />

methods were applied. Pitfalls tend more to collect adult specimens, and is biased towards to<br />

cursorial spider fauna (Topping & Sunderland, 1992). Details <strong>of</strong> <strong>the</strong> sampling protocols are<br />

given in Samu et al. (1996) and Tóth et al. (1999). Considering <strong>the</strong> above mentioned<br />

characteristics <strong>of</strong> <strong>the</strong> methods applied, <strong>the</strong>y are more complemental than comparable.<br />

Therefore in <strong>the</strong> present meta-analysis <strong>of</strong> data from different studies all comparisons were<br />

made between data sets obtained by <strong>the</strong> same method.<br />

Results and discussion<br />

Dominant species<br />

Spider assemblages in Hungarian cereal fields are characterised by a well defined and limited<br />

set <strong>of</strong> species. The most dominant, so called agrobiont species were ubiquitous in every field,<br />

corroborating results obtained in o<strong>the</strong>r European cereal fields (Luczak, 1975). The first five<br />

species which had <strong>the</strong> largest joint dominance considering all studies comprised over 3/4 <strong>of</strong><br />

<strong>the</strong> total spider fauna in each <strong>of</strong> <strong>the</strong> fields (d-vac: 76.07%±5.500, pitfall: 77.97%±10.717,<br />

values are mean±S.D.). Species composition <strong>of</strong> agrobiont species are very similar across<br />

different sites as compared among data sets collected by similar methods (Tab. 1). In <strong>the</strong><br />

present studies <strong>the</strong> average similarity was 70% for both methods. Combining both methods, it<br />

can be said that <strong>the</strong> most widely and abundantly occurring species in <strong>the</strong> Hungarian cereal<br />

fields are: P. agrestis, M. rurestris, O. apicatus, P. degeeri, and T. oblongus.<br />

Considering <strong>the</strong> 10 most abundant species, regionality had no significant influence on<br />

spider species composition, i.e. more similar communities were not necessarily closer<br />

toge<strong>the</strong>r in physical distance. (The association between species composition distances and


121<br />

physical distances were tested by Mantel test separately for D-vac and pitfall trap data; pitfall:<br />

test stat.= 0.22, t=0.757, NS; D-vac: test stat.= 0.11, t=0.268, NS.)<br />

Community structure<br />

The above results showed that few species dominate Hungarian cereal fields, and similar<br />

dominance structures in cereals were reported in o<strong>the</strong>r countries (e.g. Sunderland, 1987;<br />

Nyffeler & Breene, 1992). In spite <strong>of</strong> <strong>the</strong> prevalence <strong>of</strong> few agrobionts, <strong>the</strong> remaining one<br />

quarter <strong>of</strong> <strong>the</strong> spider individuals, which do not belong to any <strong>of</strong> <strong>the</strong>se species, show<br />

surprisingly high diversity. Species numbers obtained in <strong>the</strong> various studies depended on<br />

sampling effort. Observed species numbers ranged up to 91, species richness estimations<br />

reached values <strong>of</strong> c. 110 species for a number <strong>of</strong> fields (Tab. 2).<br />

As a result <strong>of</strong> <strong>the</strong> above mentioned dominance patterns, spider communities <strong>of</strong> cereal<br />

fields have a typical community structure (Fig. 1). They are dominated by few super-abundant<br />

species, thus <strong>the</strong> initial part <strong>of</strong> <strong>the</strong> rank abundance curves are steep. The tail <strong>of</strong> <strong>the</strong> rank<br />

abundance curves can be variable. Rank abundance curves in o<strong>the</strong>r habitats were less steep. In<br />

meadow communities, medium abundant species were more frequent, and as well a greater<br />

number <strong>of</strong> rare species were present. Alfalfa seemed to represent an intermediate type<br />

community, where usually a ra<strong>the</strong>r high density <strong>of</strong> <strong>the</strong> dominant species were observable, and<br />

<strong>the</strong> declination <strong>of</strong> <strong>the</strong> curve was between <strong>the</strong> grass and <strong>the</strong> cereal communities.<br />

Tab. 1a. Ten most dominant species <strong>of</strong> cereal fields in Hungary (lower part) and Jaccard<br />

similarity between fields. Data from sites where pitfall trapping was applied.<br />

Felsõnána –<br />

Julianna-major 0.89 –<br />

Kartal 1 0.88 0.58 –<br />

Kartal 2 0.79 0.60 0.64 –<br />

József-major 0.94 0.50 0.63 0.76 –<br />

Szomba<strong>the</strong>ly 0.95 0.58 0.59 0.83 0.57 –<br />

Site<br />

Felsõnána Juliannamajor<br />

Kartal 1 Kartal 2 Józsefmajor<br />

Szombat<br />

hely<br />

Total<br />

Species<br />

Pardosa agrestis 96 1176 660 468 1321 1764 5485<br />

Oedothorax apicatus 16 81 516 20 354 988 1975<br />

Meioneta rurestris 45 27 59 616 19 766<br />

Pachygnatha degeeri 34 10 71 12 21 117 265<br />

Trichoncoides piscator 5 2 9 175 191<br />

Xysticus kochi 17 25 80 17 41 1 181<br />

Zelotes mundus 3 106 59 168<br />

Drassyllus pusillus 17 13 5 12 76 29 152<br />

Robertus arundineti 5 20 1 20 73 119<br />

Robertus lividus 110 110


122<br />

Tab. 1b. Ten most dominant species <strong>of</strong> cereal fields in Hungary (lower part) and Jaccard<br />

similarity between fields. Data from sites where suction sampling was applied.<br />

b)<br />

Bánk –<br />

Diósjenõ 0.58 –<br />

Felsõnána 0.72 0.67 –<br />

Julianna-major 0.69 0.85 0.81 –<br />

Site Bánk Diósjenõ Felsõnána Juliannamajor<br />

Total<br />

Species<br />

Meioneta rurestris 72 12 17 158 259<br />

Tibellus oblongus 18 26 39 156 239<br />

Pisaura mirabilis 33 36 1 27 97<br />

Pardosa agrestis 13 1 9 45 68<br />

Mangora acalypha 4 5 14 13 36<br />

Erigone dentipalpis 1 24 25<br />

Pachygnatha degeeri 7 4 12 23<br />

Oedothorax apicatus 15 4 19<br />

Aulonia albimana 13 13<br />

Neottiura bimaculata 4 2 6 12<br />

Indicator species<br />

Compared to o<strong>the</strong>r agricultural and non-agricultural spider communities, cereal assemblages<br />

have a very limited specificity. Indicator species analysis tries to find species which are<br />

indicators <strong>of</strong> a given habitat type in comparison with a defined set <strong>of</strong> alternative habitats. A<br />

species is a good indicator <strong>of</strong> a certain habitat if a large portion <strong>of</strong> its individuals are caught in<br />

<strong>the</strong> given habitat, and if at <strong>the</strong> same time <strong>the</strong> species occurs at most <strong>of</strong> <strong>the</strong> study sites<br />

representing that habitat (Dufrene & Legendre, 1997). Indicator species analysis (Tab. 3)<br />

showed that <strong>the</strong>re is virtually no indicator species <strong>of</strong> cereal spider communities. If <strong>the</strong><br />

compared habitats were limited only to certain types (e.g. cereal-meadow, cereal-margin), <strong>the</strong><br />

species given as indicators were in fact equally representative for o<strong>the</strong>r agricultural habitat<br />

types such as alfalfa. On <strong>the</strong> o<strong>the</strong>r hand, agricultural areas taken as one habitat type had good<br />

indicator species, especially for <strong>the</strong> ground fauna (P. agrestis, O. apicatus, Zelotes gracilis),<br />

<strong>the</strong> first two being a dominant species <strong>of</strong> <strong>the</strong>se areas, as well. No indicator species <strong>of</strong> cereal<br />

fields existed in <strong>the</strong> cereal-alfalfa comparison, while Pardosa hortensis, for instance, was a<br />

good indicator <strong>of</strong> alfalfa.<br />

The origin <strong>of</strong> cereal spider fauna<br />

The analysis <strong>of</strong> data from several cereal fields in Hungary revealed that <strong>the</strong>se spider<br />

communities are dominated by a limited number <strong>of</strong> species, which are <strong>the</strong> same species from<br />

field to field, with no significant regional differences within <strong>the</strong> country. It was also<br />

demonstrated that <strong>the</strong>se species are not unique to cereal fields, and <strong>the</strong>y cannot be regarded as<br />

indicator species for that specific habitat type. However, considering cereal fields with o<strong>the</strong>r


123<br />

123<br />

Tab. 2. Observed and extrapolated species richness and diversity statistics.<br />

Study Method Samples Spec.<br />

No.<br />

Individual<br />

s<br />

Singletons<br />

ACE* Chao1* Jack1* MM* Alpha Shannon Simpson<br />

Bánk D-vac 12 18 533 5 21.5 24.3 23.5 29.8 5.1 2.0 4.4<br />

Diósjenõ D-vac 7 22 333 16 106.6 150.0 36.6 83.3 8.6 2.1 5.0<br />

Felsõnána D-vac 11 17 402 7 28.8 29.3 24.3 47.9 5.5 2.1 6.0<br />

Felsõnána both 26 30 645 14 59.0 62.7 44.4 41.4 8.1 2.4 6.6<br />

Felsõnána pitfall 15 18 243 9 40.6 58.5 26.4 33.0 4.8 1.8 3.8<br />

Julianna-major D-vac 48 49 2254 27 100.2 109.8 80.3 63.7 13.3 2.3 5.1<br />

Julianna-major both 93 76 3988 29 110.7 102.3 112.6 88.6 15.6 1.8 2.6<br />

Julianna-major pitfall 45 45 1734 13 58.8 52.0 60.6 52.9 8.7 1.2 1.6<br />

Kartal 1 pitfall 21 53 1661 23 86.9 82.4 76.8 69.2 10.7 1.7 3.2<br />

Kartal 2 pitfall 58 56 1039 20 77.9 81.0 76.6 63.7 13.2 2.1 3.5<br />

Józsefmajor pitfall 83 91 3713 23 111.5 120.4 113.7 98.1 17.2 2.4 4.9<br />

Szomba<strong>the</strong>ly pitfall 9 36 4179 12 50.5 50.4 51.1 67.0 5.7 1.4 2.6<br />

*Estimated species richness. Estimation was performed by EstimateS, based on species accumulation over <strong>the</strong> sampling occasions. For details <strong>of</strong> <strong>the</strong><br />

method and description <strong>of</strong> estimators refer to Colwell and Coddington (1994) and Colwell (1999).


124<br />

124<br />

3<br />

Cereal field Alfalfa<br />

Meadow / grass edge<br />

2<br />

1<br />

log(abundance)<br />

Bánk<br />

0<br />

3<br />

2<br />

1<br />

0<br />

3<br />

2<br />

1<br />

0<br />

3<br />

2<br />

1<br />

0<br />

0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70<br />

Julianna-major Felsõnána Diósjenõ<br />

Species rank<br />

Fig. 1. Rank-abundance curves <strong>of</strong> spider communities at four different sampling sites, in three habitat types, sampled by hand-held suction<br />

sampler.


125<br />

Tab. 3. Indicator species in <strong>the</strong> comparison <strong>of</strong> various habitat types. Indicator values<br />

calculated with method <strong>of</strong> Dufrene and Legendre (1997). Indicator species listed had<br />

significant indicator values (P < 0.05), as tested by Monte Carlo simulation. Species names<br />

are abbreviated.<br />

Method<br />

D-vac<br />

D-vac<br />

D-vac<br />

D-vac<br />

D-vac<br />

pitfall<br />

pitfall<br />

pitfall<br />

pitfall<br />

Compared habitats<br />

(1) cereal<br />

(2) alfalfa<br />

(3) margin<br />

(4) meadow<br />

(1) cereal<br />

(2) meadow<br />

(1) cereal<br />

(2) alfalfa<br />

(1) cereal+alfalfa<br />

(2) margin+meadow<br />

(1) cereal+alfalfa+margin<br />

(2) meadow<br />

(1) cereal<br />

(2) alfalfa<br />

(3) margin<br />

(1) cereal<br />

(2) alfalfa<br />

(1) cereal<br />

(2) margin<br />

(1) cereal+alfalfa<br />

(2) margin+meadow<br />

araehumi<br />

pardagre<br />

pardagre<br />

oedoapic<br />

zelograc<br />

Indicator species <strong>of</strong> habitats<br />

1 2 3 4<br />

– pardagre – hahnnava<br />

auloalbi<br />

epistrun,<br />

etc.<br />

clubdive<br />

hahnnava<br />

meiosimp,<br />

etc.<br />

– pachdege<br />

pardagre<br />

– auloalbi<br />

clubdive<br />

minimarg,<br />

etc.<br />

– auloalbi<br />

epistrun<br />

hahnnava,<br />

etc.<br />

– pardhort<br />

zelograc<br />

erigdent<br />

– pardhort<br />

zeloapri<br />

tracpede<br />

zelolatr<br />

zeloapri<br />

agrocupr<br />

titaschi<br />

tracpede<br />

zeloapri<br />

types <strong>of</strong> arable fields jointly, <strong>the</strong>re are a couple <strong>of</strong> indicator species, which distinguish <strong>the</strong>se<br />

communities from o<strong>the</strong>r grassy habitats that were considered in <strong>the</strong> present study. Grassy<br />

habitats are much more characteristic than arable habitats, by having numerous species that<br />

regularly occur <strong>the</strong>re, but not in <strong>the</strong> latter habitat type.<br />

It is interesting to look at in which type <strong>of</strong> natural areas arable indicator species occur.<br />

According to <strong>the</strong> Central European habitat database <strong>of</strong> spiders (Hanggi et al., 1995) and<br />

Hungarian data (Szita et al., 1998), P. agrestis, O. apicatus and Araeoncus humilis, apart


126<br />

from widely occurring in arable and similar human disturbed habitats, all native to wetland<br />

areas, such as saline grasslands, coastal sand dunes, beaches. These areas are characterised by<br />

frequent, regular perturbances <strong>of</strong> seasonal flooding, sand movement, tides, etc.<br />

Marshall and Rypstra (1999) argues that one factor that determines which species can<br />

dominate structurally simple ecosystems, such as agricultural land or wetland areas, is strong<br />

interspecific competition. We think that, although intense competition may infrequently occur<br />

in those ecosystems, agricultural spider species are <strong>the</strong> likely losers in such situation. There is<br />

evidence that directly or indirectly supports that it is not superior competitive ability that<br />

makes <strong>the</strong>se species dominant in agro-ecosystems. For instance, <strong>the</strong> most dominant<br />

agricultural species, P. agrestis, does not occur in more permanent grasslands, where more<br />

resources, but also more competitor species can be found (Szita et al., 1998). A subspecies <strong>of</strong><br />

<strong>the</strong> same species, Pardosa agrestis purbeckensis, was experimentally shown to be inferior in<br />

interference competition with o<strong>the</strong>r wolf spiders in a salt marsh area (Schaefer, 1974). O<strong>the</strong>r<br />

littoral Pardosa species similarly were shown to be inferior in competitive situations (Döbel<br />

et al., 1990).<br />

Here we suggest that <strong>the</strong> similarity between wet land and agricultural ecosystems is more<br />

in structural similarity and perturbance pattern, which actually interrupts or resets competitive<br />

situations. Therefore <strong>the</strong> success <strong>of</strong> agricultural spider species should lie in <strong>the</strong>ir ability to<br />

cope with perturbance and find suitable niches under structurally simple conditions.<br />

Adaptations to <strong>the</strong>se factors might lie in good dispersal ability (Bishop & Riechert, 1990) and<br />

life history characteristics (T<strong>of</strong>t, 1989; Samu et al., 1998). Attempts to augment spider<br />

densities in cereals should consider <strong>the</strong>se characteristics, and devise agricultural practices<br />

with <strong>the</strong> structural needs, dispersal modes and life cycle patterns <strong>of</strong> spiders in mind.<br />

Acknowledgements<br />

The study was supported by OTKA grants F 17691 and 23627. FS was funded by a research<br />

grant from <strong>the</strong> Ecological Centre <strong>of</strong> HAS. FS and CsSz were Bolyai Fellows <strong>of</strong> HAS.<br />

References<br />

Bishop, L. & Riechert, S.E. 1990: Spider colonization <strong>of</strong> agroecosystems mode and source.<br />

Env. Entomol. 19: 1738-1745.<br />

Döbel, H.G., Denno, R.F. & Coddington, J.A. 1990: Spider (Araneae) community structure in<br />

an intertidal salt marsh: effects <strong>of</strong> vegetation structure and tidal flooding. Env. Entomol.<br />

90: 1356-1370.<br />

Dufrene, M. & Legendre, P. 1997: Species assemblages and indicator species: <strong>the</strong> need for a<br />

flexible asymmetrical approach. Ecol. Monogr. 67: 345-366.<br />

Hänggi, A., Stöckli, E. & Nentwig, W. 1995: Habitats <strong>of</strong> Central European spiders. Misc.<br />

Faun. Helvet. 4: 1-460.<br />

Luczak, J. 1975: Spider communities <strong>of</strong> crop fields. Pol. Ecol. Stud. 1: 93-110.<br />

Marshall, S.D. & Rypstra, A.L. 1999. Spider competition in structurally simple ecosystems. J.<br />

Arachnol. 27: 343-350.<br />

Nyffeler, M. & Benz, G. 1987: Spiders in natural pest control: a review. J. Appl. Entomol.<br />

103: 321-339.<br />

Nyffeler, M. & Breene, R.G. 1992: Dominant insectivorous polyphagous predators in winter<br />

wheat: high colonisation power, spatial dispersion patterns, and probable importance <strong>of</strong><br />

<strong>the</strong> soil surface spiders. Dtsch. Ent. Z., N. F. 39: 177-188.


Nyffeler, M., Sterling, W.L. & Dean, D.A. 1994: Insectivorous activities <strong>of</strong> spiders in United<br />

States field crops. J. Appl. Entomol. 118: 113-128.<br />

Samu, F., Németh, J. & Kiss, B. 1997: Assessment <strong>of</strong> <strong>the</strong> efficiency <strong>of</strong> a hand-held suction<br />

device for sampling spiders: improved density estimation or oversampling? Ann. Appl.<br />

Biol. 130: 371-378.<br />

Samu, F., Németh, J., Tóth, F., Szita, É., Kiss, B. & Szinetár, C. 1998: Are two cohorts<br />

responsible for bimodal life history pattern in <strong>the</strong> wolf spider Pardosa agrestis in<br />

Hungary? <strong>Proceedings</strong> 17th European Colloquium <strong>of</strong> Arachnology, Edinburgh: 215-221.<br />

Samu, F. & Sárospataki, M. 1995: Design and use <strong>of</strong> a hand-hold suction sampler and its<br />

comparison with sweep net and pitfall trap sampling. Fol. Entomol. Hung. 56: 195-203.<br />

Samu, F., Vörös, G. & Botos, E. 1996: Diversity and community structure <strong>of</strong> spiders <strong>of</strong> alfalfa<br />

fields and grassy field margins in South Hungary. Acta Phytopath. Entomol. Hung. 31:<br />

253-266.<br />

Schaefer, M. 1974: Experimentelle Untersuchungen zur Bedeutung der interspezifischen<br />

Konkurrenz bei 3 Wolfspinnen-Arten (Araneida: Lycosidae) einer Salzwiese. Zool. Jb.<br />

Syst. 101: 213-235.<br />

Sunderland, K.D. 1987: Spiders and cereal aphids in Europe. Bull. SROP/<strong>WPRS</strong> 10: 82-102.<br />

Szita, É., Samu, F., Bleicher, K. & Botos, E. 1998. Data to <strong>the</strong> spider fauna (Araneae) <strong>of</strong><br />

Körös-Maros National Park (Hungary). Acta Phytopath. Entomol. Hung. 33: 341-388.<br />

T<strong>of</strong>t, S. 1989: Aspects <strong>of</strong> <strong>the</strong> ground-living spider fauna <strong>of</strong> two barley fields in Denmark:<br />

species richness and phenological synchronization. Entomol. Meddl. 57: 157-168.<br />

Topping, C.J. & Sunderland, K.D. 1992: Limitations to <strong>the</strong> use <strong>of</strong> pitfall traps in ecological<br />

studies, exemplified by a study <strong>of</strong> spiders in a field <strong>of</strong> winter wheat. J. Appl. Ecol. 29:<br />

485-491.<br />

Tóth, F. 1999: Comparative analyses <strong>of</strong> epigeic spider assemblages in Nor<strong>the</strong>rn Hungarian<br />

Winter wheat fields and <strong>the</strong>ir adjacent margins. J. Arachnol. 27: 241-248.<br />

127


128


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 129 - 137<br />

Agrobiological and biocenological study <strong>of</strong> winter oats (Avena sativa L.)<br />

T. Gueorguieva and A. Mateeva<br />

Agricultural University, 12, Mendeleev str., 4000 Plovdiv, Bulgaria<br />

Summary<br />

During <strong>the</strong> last years, winter oat (Avena sativa L.) cultivation has substantially increased in Bulgaria.<br />

To assure sustainable yields <strong>of</strong> <strong>the</strong> crop, <strong>the</strong> development <strong>of</strong> a new plant protection system, focusing<br />

on <strong>the</strong> bio-cenological conditions <strong>of</strong> <strong>the</strong> agro-ecosystem, is <strong>of</strong> major importance. Between 1995- and<br />

1997, <strong>the</strong> Department <strong>of</strong> Plant Growing at <strong>the</strong> Agricultural University <strong>of</strong> Plovdiv conducted several<br />

agro-biological and bio-cenological studies on winter oats as bi-factorial field experiments in a block<br />

design. The three tested varieties showed different productive tillering and expressed strong<br />

compensatory reaction during <strong>the</strong> vegetation periods. In <strong>the</strong> course <strong>of</strong> <strong>the</strong> bio-cenological studies <strong>the</strong><br />

following pests were found to be <strong>of</strong> significant economic importance: Oulema melanopus L., Oulema<br />

gallaeciana (Heyden), Cephus pygmeus L., Mayetiola destructor Say, Oscinella frit L., Sitobion<br />

avenae Fab., Agriotes obscurus (Linnaeus), and Heterodera avenae Wollenweber. The population<br />

dynamics <strong>of</strong> S. avenae were influenced by <strong>the</strong> following natural enemies: Coccinella septempunctata<br />

L., Hippodamia variegata Goeze, Propylea quatuordecimpunctata L., Adalia decempunctata L.,<br />

Chrysoperla carnea Stephens, Episyrphus balteatus De Geer, Scaeva pyrastri (L.), and Nabis<br />

pseud<strong>of</strong>erus Remane.<br />

Key words: Avena sativa L., oat, seeding rate, pest, pest control, predators, prey.<br />

Introduction<br />

The obvious tendency towards increase <strong>of</strong> <strong>the</strong> fields sown with winter oats (Avena sativa L.)<br />

in Bulgaria is due to <strong>the</strong> better utilisation <strong>of</strong> <strong>the</strong> winter moisture, <strong>the</strong> much earlier ripening,<br />

and <strong>the</strong> obtaining <strong>of</strong> more sustainable yields in comparison with <strong>the</strong> spring oats. Various<br />

studies on <strong>the</strong> basic aspects <strong>of</strong> <strong>the</strong> technology <strong>of</strong> cultivation and <strong>the</strong> productive capacity <strong>of</strong> <strong>the</strong><br />

different genotypes <strong>of</strong> winter oats in <strong>the</strong> agro-climatic conditions in Bulgaria have being<br />

carried out (Gueorguieva et al., 1990, 1991, 1993, 1996, 1997).<br />

Like any crop, winter oats is attacked by many diseases and pests. According to Grigorov<br />

(1964, 1965, 1980), Ljubenov (1963), and Lazarov (1969) oats is mostly attacked by <strong>the</strong> frit<br />

fly Oscinella frit L. (Dipt.: Chloropidae). In Bulgaria <strong>the</strong> damage caused by this pest every<br />

year varies from 12 to 28%. According to Grigorov (1993) <strong>the</strong> serious damage caused by this<br />

pest is due to <strong>the</strong> simultaneity <strong>of</strong> <strong>the</strong> period <strong>of</strong> mass flight <strong>of</strong> <strong>the</strong> 2nd generation and <strong>the</strong> oats<br />

panicle emergence. Moudry et al. (1996) also reports <strong>of</strong> serious damage to <strong>the</strong> oats caused by<br />

<strong>the</strong> frit fly. Damage to <strong>the</strong> oats is also caused by Oulema melanopus L., O. gallaeciana<br />

(Heyden), (both Col.: Chrysomelidae) Cephus pygmeus L. (Hym.: Cephidae), Sitobion avenae<br />

Fab., and Rhopalosiphum padi L. (both Hom.: Aphididae) (Grigorov, 1993; Valoska et al.,<br />

1995; Hsia et al., 1997).<br />

The solution <strong>of</strong> <strong>the</strong>se plant protection problems <strong>of</strong> winter oats requires pr<strong>of</strong>ound studies.<br />

Therefore, <strong>the</strong> objective <strong>of</strong> <strong>the</strong> presented investigations was to study some agro-biological and<br />

bio-cenological peculiarities <strong>of</strong> winter oats in <strong>the</strong> regions around <strong>the</strong> town <strong>of</strong> Plovdiv in<br />

Bulgaria.<br />

129


130<br />

Material and methods<br />

Between 1995 and 1997, <strong>the</strong> Department <strong>of</strong> Plant Growing at <strong>the</strong> Agricultural University <strong>of</strong><br />

Plovdiv conducted several agro-biological and bio-cenological studies on winter oats as bifactorial<br />

field experiments in a block design.<br />

The agro-biological characteristics <strong>of</strong> three cultivars <strong>of</strong> winter oats were studied under<br />

<strong>the</strong> conditions <strong>of</strong> different seeding rates and sowing schemes: Cultivars - Dunav 1st for<br />

Bulgaria, Joker (Fr) and No 83106028 (USA); Seeding rates 400 seeds/m 2 (12/2 cm), 200<br />

seeds/m 2 (12/4 cm), 100 seeds/m 2 (12/8 cm), 100 seeds/m 2 (10/10 cm), 25 seeds/m 2 (20/20<br />

cm), 6 seeds/m 2 (40/40 cm).<br />

The following biological indices <strong>of</strong> <strong>the</strong> oats were traced: entry into different phases <strong>of</strong><br />

development - germination, tiller production, stem elongation, panicle emergence, wax and<br />

complete ripeness; productive tillering (number <strong>of</strong> productive tillers per plant), and yield<br />

(kg/da).<br />

The bio-cenological information was collected by means <strong>of</strong> cutting with an<br />

entomological bag and leaf samples. All insect species were determined in <strong>the</strong> Department <strong>of</strong><br />

Entomology at <strong>the</strong> Agrarian University <strong>of</strong> Plovdiv.<br />

Results and discussion<br />

I. Agrobiological studies on winter oats<br />

During <strong>the</strong> two years <strong>of</strong> study <strong>the</strong> phenological development <strong>of</strong> <strong>the</strong> different cultivars was<br />

similar. During <strong>the</strong> autumn vegetation no differences in this index, depending on <strong>the</strong> cultivars<br />

and <strong>the</strong> seeding rate, were observed. In spring, <strong>the</strong> panicle emergence in cultivar No<br />

83106028 occurred 6 to 10 days earlier than in <strong>the</strong> o<strong>the</strong>r two cultivars. The differences in <strong>the</strong><br />

panicle emergence remained relatively <strong>the</strong> same until <strong>the</strong> time <strong>of</strong> ga<strong>the</strong>ring. With <strong>the</strong> decrease<br />

<strong>of</strong> <strong>the</strong> seeding rate, <strong>the</strong> plants’ development slowed down.<br />

The ripening was most simultaneous in cultivar N o 83106028, and slowest in <strong>the</strong> standard<br />

cultivar Dunav 1 (Tab. 1).<br />

Data in Table 2 indicate that <strong>the</strong> highest productive tillering per plant at all seeding rates<br />

was observed in cultivar Dunav 1, and <strong>the</strong> lowest in cultivar N o 83106028. Depending on <strong>the</strong><br />

number <strong>of</strong> seeds and <strong>the</strong> sowing scheme, we recorded that with <strong>the</strong> change <strong>of</strong> <strong>the</strong> seeding rate<br />

from 400 to 25 seeds/m 2 , <strong>the</strong> productive tillering increased from 3.1 to 13.1, respectively. The<br />

decrease <strong>of</strong> <strong>the</strong> seeding rate to 6 seeds/m 2 did not lead to fur<strong>the</strong>r increase <strong>of</strong> <strong>the</strong> productive<br />

tillering. Cultivars Joker and N o 83106028 form, however, largest amount <strong>of</strong> tillers at <strong>the</strong><br />

lowest seeding rate.


131<br />

Tab. 1. Phenological development <strong>of</strong> winter oats, depending on cultivars and <strong>the</strong> seeding<br />

rates.<br />

Seeding<br />

rate<br />

seeds/m 2<br />

1995/1996<br />

Sowing<br />

Germination<br />

Tiller<br />

production<br />

Stem<br />

elongation<br />

Dunav 1 - st<br />

400, 200,<br />

100,<br />

50, 25<br />

9. X. 25. X. 14. XII. 29. IV.<br />

Joker<br />

400, 200,<br />

100, 50, 9. X. 25. X. 16. XII. 29. IV.<br />

25<br />

N o 83106028<br />

400, 200,<br />

100, 50, 11. X. 27. X. 16. XII. 27. IV.<br />

25<br />

1996/1997<br />

Panicle<br />

emergence<br />

17-23. V.<br />

18-26. V.<br />

20-28. V.<br />

17-21. V.<br />

18-25. V.<br />

20-28. V.<br />

10-15. V.<br />

12-17. V.<br />

14-19. V.<br />

Grain - filling<br />

Wax<br />

ripeness<br />

19. VI.<br />

2. VII.<br />

5. VII.<br />

19. VI.<br />

1. VII.<br />

4. VII.<br />

13. VI.<br />

17. VI.<br />

20. VI.<br />

Complete<br />

ripeness<br />

23. VI.<br />

7. VII.<br />

10. VII.<br />

23. VI.<br />

5. VII.<br />

9. VII.<br />

17. VI.<br />

21. VI.<br />

24. VI.<br />

Dunav 1 - st<br />

400, 200,<br />

100, 50,<br />

25<br />

15. X. 22. X. 10. XII 30. IV.<br />

Joker<br />

400, 200,<br />

100, 15. X. 22. X. 10. XII. 30. IV.<br />

50,25<br />

N o 83106028<br />

400, 200,<br />

100, 50, 11. X. 22. X. 10. XII. 25. IV.<br />

25<br />

14-16. V.<br />

15-18. V.<br />

18-20. V.<br />

14-15. V.<br />

15-17. V.<br />

18-20. V.<br />

8-10. V.<br />

10-12. V.<br />

13-16. V.<br />

15. VI.<br />

18. VI.<br />

22. VI.<br />

15. VI.<br />

18. VI.<br />

22. VI.<br />

9. VI.<br />

13. VI.<br />

16. VI.<br />

22. VI.<br />

29. VI.<br />

2. VII.<br />

22. VI.<br />

25. VI.<br />

29. VI.<br />

16. VI.<br />

20. VI.<br />

23. VI.<br />

The analysis <strong>of</strong> <strong>the</strong> results <strong>of</strong> grain yield (Tab. 2) <strong>of</strong> <strong>the</strong> three tested cultivars indicates<br />

that oats is a crop with very good compensatory capacity. At optimum seeding rates (400<br />

seeds/m 2 - 12/2 cm) <strong>the</strong> yield varies from 633 to 891 kg/da. Cultivar Joker had <strong>the</strong> highest<br />

productivity. In <strong>the</strong> case, even if <strong>the</strong> seeding rate is decreased fourfold, <strong>the</strong> yield slightly<br />

increases. The distance between <strong>the</strong> plants <strong>of</strong> 20/20 cm (25 seeds/m 2 ) or 40/40 (6 seeds/m 2 )<br />

lead to a sharp decrease <strong>of</strong> <strong>the</strong> yield, regardless <strong>of</strong> <strong>the</strong> fact that, in this case <strong>the</strong> number <strong>of</strong> <strong>the</strong><br />

tillers varied from 8 to 14.


132<br />

Tab. 2. Productive tillering and winter oats yield, depending on cultivars and seeding rates.<br />

Seeding rate end<br />

sowing scheme<br />

Productive tillers<br />

(number/plant)<br />

seeds/m 2 - cm/cm Dunav1 Joker N o 831<br />

06028<br />

Yield<br />

(kg/da)<br />

Means Dunav1 Joker N o 831<br />

06028<br />

Means<br />

1996<br />

400 - 12/2 cm 2.7 3.2 3.5 3.1 633 891 732 752<br />

200 - 12/4 cm 1.8 4.0 4.7 3.5 598 835 666 700<br />

100 - 12/8 cm 7.8 6.9 6.6 7.1 735 779 710 742<br />

100 - 10/10 cm 9.8 10.0 8.5 9.4 808 1059 601 822<br />

25 - 20/20 cm 14.7 12.7 12.0 13.1 147 259 100 169<br />

6 - 40/40 cm 10.8 7.9 7.8 8.8 32 20 14 22<br />

Means 7.9 7.5 7.2 492 639 470<br />

1997<br />

400 - 12/2 cm 3.3 2.3 3.4 3.0 856 967 996 940<br />

200 - 12/4 cm 5.0 4.5 5.5 5.0 784 929 889 867<br />

100 - 12/8 cm 8.2 7.1 7.8 7.7 862 926 830 872<br />

100 - 10/10 cm 7.8 6.1 9.0 7.6 625 873 823 774<br />

25 - 20/20 cm 13.7 12.1 12.1 12.6 525 673 672 623<br />

6 - 40/40 cm 8.1 13.2 17.6 13.0 86 166 216 156<br />

Means 7.7 7.5 9.2 623 756 738<br />

II. Biocenological Studies on Winter Oats<br />

Figure 1 represents <strong>the</strong> structure <strong>of</strong> <strong>the</strong> winter oats bio-cenosis as a percent correlation<br />

between classes <strong>of</strong> Insecta, Arachnoidea, and Nematoda for two successive years (i.e. 1996-<br />

1997).<br />

1996<br />

1997<br />

6%<br />

13%<br />

4%<br />

6%<br />

81%<br />

90%<br />

Insecta<br />

Arachnoidea<br />

Nematoda<br />

Fig. 1. Structure <strong>of</strong> <strong>the</strong> winter oats bio-cenosis in <strong>the</strong> experimental training field <strong>of</strong> <strong>the</strong><br />

Department <strong>of</strong> Plant Growing –at <strong>the</strong> Agricultural University <strong>of</strong> Plovdiv.<br />

It is evident that in <strong>the</strong> bio-cenosis structure during both years <strong>of</strong> <strong>the</strong> study, <strong>the</strong> class<br />

Insecta was predominant over <strong>the</strong> classes Nematoda and Arachnoidea.


133<br />

Figure 2 shows <strong>the</strong> results <strong>of</strong> <strong>the</strong> percent correlation between <strong>the</strong> orders in <strong>the</strong> entomoand<br />

acar<strong>of</strong>auna <strong>of</strong> winter oats.<br />

28%<br />

1996<br />

3%<br />

6%<br />

2%<br />

27%<br />

1997<br />

4%<br />

7%<br />

1%<br />

1.Homoptera<br />

2.Orthoptera<br />

3.Hemiptera<br />

4.Thysanoptera<br />

2%<br />

1%<br />

25%<br />

3%<br />

2%<br />

27%<br />

5. Neuroptera<br />

6. Coleoptera<br />

7. Hymenoptera<br />

31%<br />

2%<br />

28% 1%<br />

8. Diptera<br />

9.Arachnoidea<br />

Fig. 2. Proportions <strong>of</strong> insect orders and Arachnoidea in <strong>the</strong> experimental winter oat field <strong>of</strong><br />

<strong>the</strong> Department <strong>of</strong> Plant Growing at Plovdiv. – Clockwise arrangement <strong>of</strong> sectors starting at<br />

right with 25% Homoptera (1996) and 27% Homoptera (1997).<br />

The results indicate that species <strong>of</strong> <strong>the</strong> orders Homoptera (25-27%), Hemiptera (28-315),<br />

Coleoptera (27-28%), and Diptera (6-7%) predominantly occur <strong>the</strong> winter oats in comparison<br />

with species <strong>of</strong> <strong>the</strong> orders Orthoptera, Thysanoptera, Neuroptera, and Hymenoptera, which<br />

were present in considerably lower proportions. Lower proportions <strong>of</strong> mites and spiders than<br />

insects appeared in winter oats. In <strong>the</strong> Arachnoidea, 28 (2%) and 19 (1%) species were<br />

recorded in 1996 and 1997, respectively.<br />

Within <strong>the</strong> order <strong>of</strong> Coleoptera, chrysomeldids were <strong>the</strong> predominating beetle family with<br />

a proportion <strong>of</strong> about 50% in both study years (Fig. 3).<br />

100%<br />

80%<br />

60%<br />

40%<br />

Chrysomelidae<br />

Curcuniolidae<br />

Tenebrionidae<br />

Nitidulidae<br />

20%<br />

0%<br />

1996 1997<br />

Scarabaeidae<br />

Elateridae<br />

Fig. 3. Proportion <strong>of</strong> beetle families in <strong>the</strong> experimental winter oat field <strong>of</strong> <strong>the</strong> Department <strong>of</strong><br />

Plant Growing at Plovdiv. – Arrangement <strong>of</strong> <strong>the</strong> families from top to bottom.<br />

Oulema melanopus was <strong>the</strong> most abundant chrysomelid species, followed by O.<br />

gallaeciana, and Phyllotreta vitulla (Redt.).


134<br />

1996<br />

1997<br />

20% 9%<br />

25%<br />

8%<br />

Oulema melanopus L.<br />

Oulema gallaeciana (Heyden)<br />

Phyllotreta vitulla Redt.<br />

71%<br />

67%<br />

Fig. 4. Proportion <strong>of</strong> chrysomelid species in <strong>the</strong> experimental winter oat field <strong>of</strong> <strong>the</strong><br />

Department <strong>of</strong> Plant Growing at Plovdiv. – Clockwise arrangement <strong>of</strong> sectors starting at right<br />

with 9% Phyllotreta vitulla (1996) and 8% Phyllotreta vitulla (1997).<br />

Within <strong>the</strong> group <strong>of</strong> phytophagous Diptera in both study years <strong>the</strong> following species were<br />

recorded (in <strong>the</strong> order <strong>of</strong> <strong>the</strong>ir predominance): frit fly (O. frit), Hessian fly (Mayetiola<br />

destructor Say [Dipt.: Cecidomyiidae]), and red wheat gnat (Haplodiplosis marginata Roser<br />

[Dipt.: Cecidomyiidae]) (Fig. 5).<br />

100%<br />

80%<br />

60%<br />

40%<br />

20%<br />

Haplodiplosis marginata Roser<br />

Mayetiola destructor Say.<br />

Oscinella frit L.<br />

0%<br />

1996 1997<br />

Fig. 5. Proportion <strong>of</strong> phytophagous dipteran species in <strong>the</strong> experimental winter oat field <strong>of</strong> <strong>the</strong><br />

Department <strong>of</strong> Plant Growing at Plovdiv. – Arrangement <strong>of</strong> <strong>the</strong> species from top to bottom.<br />

100%<br />

80%<br />

60%<br />

40%<br />

20%<br />

Trachelus tabidus F.<br />

Cephus pygmeus L.<br />

Dolerus haematodes (Schrank)<br />

0%<br />

1996 1997<br />

Fig. 6. Proportion <strong>of</strong> phytophagous Hymenoptera in <strong>the</strong> experimental winter oat field <strong>of</strong> <strong>the</strong><br />

Department <strong>of</strong> Plant Growing at Plovdiv. – Arrangement <strong>of</strong> <strong>the</strong> species from top to bottom.<br />

The most important phytophagous wasp species was <strong>the</strong> wheat stem wasp C. pygmeus,<br />

followed by <strong>the</strong> black wheat stem wasp Trachelus tabidus F. (Hym.: Cephidae) and Dolerus


135<br />

haematodes (Schrank) (Hym.: Tenthredinidae) (Fig. 6). The climatic conditions at <strong>the</strong> end <strong>of</strong><br />

April and <strong>the</strong> beginning <strong>of</strong> May during both years were especially favourable for <strong>the</strong><br />

appearance, flight, and reproduction <strong>of</strong> C. pygmeus.<br />

Tab. 3. Insect predators recorded in <strong>the</strong> experimental winter oat field <strong>of</strong> <strong>the</strong> Department <strong>of</strong><br />

Plant Growing at Plovdiv.<br />

Order Family Registered species<br />

Coleoptera Coccinellidae Coccinella septempunctata L.<br />

Coccinula quatordecimpustulata L.<br />

Hippodamia variegata Goeze.<br />

Propylea quatuordecimpunctata L.<br />

Adalia decempunctata L.<br />

Hippodamia tredecimpunctata (L.)<br />

Neuroptera Chrysopidae Chrysoperla carnea Stephens<br />

Chrysopa harrisii (Fitch)<br />

Chrysopa pallens Rambur<br />

Diptera Syrphidae Episyrphus balteatus De Geer<br />

Syrphus vitripennis Meig.<br />

Eupeodes latifasciatus (Mac.)<br />

Scaeva pyrastri (L.)<br />

Sphaerophoria scripta (L.)<br />

Heteroptera Anthocoridae Anthocoris pilosus (Yakovlev)<br />

Nabidae<br />

Nabis pseud<strong>of</strong>erus Rm.<br />

Coleoptera Carabidae Calosoma sycophanta (L.)<br />

III. Presentation <strong>of</strong> beneficial insect species in <strong>the</strong> agro-cenosis <strong>of</strong> winter oats<br />

In modern systems for pest management one <strong>of</strong> <strong>the</strong> main approaches is to use and preserve<br />

beneficial species in certain agro-cenosis. It is thus necessary to know when <strong>the</strong>se beneficials<br />

appear, what <strong>the</strong>ir optimum conditions for reproduction and development are, what are <strong>the</strong>ir<br />

regulatory capacities, and how beneficials are influenced by <strong>the</strong> use <strong>of</strong> pesticides and o<strong>the</strong>r<br />

agro-technical interventions. Within <strong>the</strong> framework <strong>of</strong> our study, <strong>the</strong> species structure <strong>of</strong><br />

beneficials in <strong>the</strong> winter oat agro-cenosis was determined. Densities <strong>of</strong> most beneficials<br />

during <strong>the</strong> different phases <strong>of</strong> <strong>the</strong>ir development, were closely related to <strong>the</strong> occurrence <strong>of</strong><br />

aphids in oats, particularly with that <strong>of</strong> S. avenae. The following families <strong>of</strong> beneficials were<br />

predominant: Coccinellidae, Chrysopidae, Syrphidae, Anthocoridae and Nabidae. Table 3<br />

represents <strong>the</strong> predators from <strong>the</strong>se families that feed on aphids in oat. Coccinellids were<br />

found in both years most frequently after <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> second half <strong>of</strong> May. Ladybird<br />

beetles were very important in regulating <strong>the</strong> number <strong>of</strong> oat aphids. We recorded 126 and 103<br />

specimen <strong>of</strong> coccinellids in 1996 and 1997, respectively. High number <strong>of</strong> coccinellids were<br />

related to aphid density in <strong>the</strong> crop and by ecological conditions. In <strong>the</strong> course <strong>of</strong> <strong>the</strong> study,<br />

larvae and adults <strong>of</strong> C. septempunctata were most frequently found in <strong>the</strong> colonies <strong>of</strong> <strong>the</strong><br />

aphids in winter oats. The larvae <strong>of</strong> <strong>the</strong> seven-spot ladybird consume during <strong>the</strong>ir lifetime<br />

about 600-800 aphids, and <strong>the</strong> adult ladybirds at about 40-50 aphids a day. In addition to


136<br />

C. septempunctata, <strong>the</strong> fourteen-spot ladybird P. quatordecimpunctata and <strong>the</strong> mutable<br />

ladybird H. variegata were also frequently found. The mutable and <strong>the</strong> fourteen-spot<br />

ladybirds have a smaller prey capacity than <strong>the</strong> seven-spot ladybird, but are also important<br />

biological regulators <strong>of</strong> <strong>the</strong> wheat and oat aphids in <strong>the</strong> winter oat agro-cenosis. For example,<br />

single larvae <strong>of</strong> P. quatordecimpunctata and H. variegata consume during <strong>the</strong>ir development<br />

200-250 aphids on wheat, and <strong>the</strong> adult ladybirds during <strong>the</strong>ir lifetime 2100-2900 aphids, and<br />

from 15 to 73 aphids a day.<br />

The most frequently found chrysopid was Chrysoperla carnea Stephens. A single larva<br />

has a prey capacity <strong>of</strong> 720 aphids. O<strong>the</strong>r chrysopids recorded in our studies were Chrysopa<br />

harrisii (Fitch)and Chrysopa pallens Rambur.<br />

The predominant syrphid encountered was Scaeva pyrastri (L.). At temperatures <strong>of</strong> 20-<br />

22 o C a single larva can consume between 15 and 45 aphids daily, depending on <strong>the</strong> type <strong>of</strong><br />

aphid and <strong>the</strong> development stage <strong>of</strong> <strong>the</strong> syrphid larva.<br />

Predatory bugs <strong>of</strong> family Nabidae, in particular Nabis pseud<strong>of</strong>erus Rm., were also found<br />

in <strong>the</strong> agro-cenosis <strong>of</strong> winter oats. The larvae N. pseud<strong>of</strong>erus consume 14-20 aphids daily, and<br />

about 400 during <strong>the</strong>ir total lifetime.<br />

Curculionids and nitidulids were only rarely observed during our studies.<br />

The three tested cultivars have different vegetation periods in terms <strong>of</strong> duration, with N o<br />

83106028 being <strong>the</strong> earliest one. The seeding rate influences <strong>the</strong> length <strong>of</strong> <strong>the</strong> vegetation<br />

period, and crops with lower rate develop slower. With <strong>the</strong> decrease <strong>of</strong> <strong>the</strong> seeding rate from<br />

400 to 25 seeds/m 2 , <strong>the</strong> productive tillering increased from 3 to 13 tillers per plant. Cultivar<br />

Joker proved to be <strong>the</strong> most productive one. Decreasing <strong>the</strong> seeding rate from 400 to 100<br />

seeds/m 2 did not affect <strong>the</strong> yield <strong>of</strong> any <strong>of</strong> <strong>the</strong> three tested cultivars.<br />

In <strong>the</strong> course <strong>of</strong> our study in winter oat fields at <strong>the</strong> Department <strong>of</strong> Plant Growing at <strong>the</strong><br />

Agricultural University <strong>of</strong> Plovdiv, <strong>the</strong> following important pests and beneficials were found:<br />

Pests: Oulema melanopus, O. gallaeciana, C. pygmeus, M. destructor, O. frit, S. avenae,<br />

Agriotes obscurus L., and Heterodera avenae Wollenweber.<br />

Beneficials: C. septempunctata, H. variegata, P. quatuordecimpunctata, C. carnea, E.<br />

balteatus, S. pyrastry, and N. pseud<strong>of</strong>erus .<br />

The predominant pests were found in <strong>the</strong> orders <strong>of</strong> Coleoptera and Diptera.<br />

References<br />

Grigorov, St. 1964: Contribution to Bulgaria’s entom<strong>of</strong>auna. Plant Protection Magazine.<br />

Grigorov, St. 1965: Entomocenosis <strong>of</strong> cereals in <strong>the</strong> region <strong>of</strong> S<strong>of</strong>ia. Scientific Works, Vol.<br />

XVI.<br />

Grigorov, St. 1980: Greenflies and <strong>the</strong>ir control. Zemizdat Magazine.<br />

Grigorov, P. 1993: The wheat trips – a serious pest. Zemizdat Magazine.<br />

Gueorguieva, T., Mileva, D. & Kostov, K. 1990: Influence <strong>of</strong> nitrogen fertilizer and seeding<br />

rate on nitratreductase activity <strong>of</strong> winter oats. Physiology 8 (1): 112-115.<br />

Gueorguieva, T., Kostov, K. & Mileva, D. 1991: Catalase and peroxidase activity, depending<br />

on fertiliser and seeding rates. Plant metabolism regulation. Varna: 46-49.<br />

Gueorguieva, T. & Yankov, B.,1993: Formation <strong>of</strong> winter oat yield, depending on seeding<br />

date and rate. Kostinbrod: 107-113.<br />

Gueorguieva, T. & Kostov, K. 1996: Biological characteristics <strong>of</strong> new winter oat cultivars<br />

grown under conditions <strong>of</strong> Central South Bulgaria. Scientific works <strong>of</strong> <strong>the</strong> Academy <strong>of</strong><br />

Agriculture, Vol. VII: 313-317.


Gueorguieva, T. & Kostov, K. 1996: Economic characteristics <strong>of</strong> new winter oat cultivars<br />

grown under conditions <strong>of</strong> Central South Bulgaria. Scientific works <strong>of</strong> <strong>the</strong> Academy <strong>of</strong><br />

Agriculture, Vol. VII: 318-321.<br />

Gueorguieva, T.& Dimitrova, M. 1997: Economic characteristics <strong>of</strong> winter oat panicle<br />

according to <strong>the</strong> crop structure and <strong>the</strong> degree <strong>of</strong> tillering. Soil Science, Agrochemistry,<br />

and Ecology Magazine 32 (6): 61-63.<br />

Gueorguieva, T., Mokreva, T. & Yanchev, I. 1997: Tillering as a factor <strong>of</strong> productivity<br />

formation in winter oats under conditions <strong>of</strong> reduced plant density. Scientific works <strong>of</strong><br />

<strong>the</strong> Academy <strong>of</strong> Agriculture, Vol. 4 (1): 23-25.<br />

Hsiao, W.F. & Khachtourian, G.G. 1997: Impact <strong>of</strong> cereal hosts on <strong>the</strong> susceptibility <strong>of</strong><br />

Rhopalosiphum padi (Homoptera: Aphididae) to <strong>the</strong> entomopathogenic fungus.<br />

Verticillium lecanii. Plant Protection Bulletin Taipei 39 (4): 313-327.<br />

Lazarov, A. 1969: The frit fly – Oscinosoma frit (Diptera) – as a pest on cereals in Bulgaria<br />

and <strong>the</strong> means for its control.<br />

Ljubenov, Y. 1961: The wild cereal plants as a reserve and source <strong>of</strong> <strong>the</strong> frit fly reproduction.<br />

Agricultural Thought Magazine, Vol. 5.<br />

Moudry, J., Vozenilkova, B. & Votavova, O., 1996: Damage to oats by frit fly (Oscinella frit<br />

L.): 71-80.<br />

Valoska, B., Sadova, M.. & Liskova, M., 1995: Response <strong>of</strong> selected cereal cultivars to<br />

Heterodera avenae Woll. pathotype H 12. Slovac Republic: 215-218.<br />

137


138


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 139 - 142<br />

Side effects <strong>of</strong> some pesticides on aphid specific predators<br />

in winter wheat<br />

A. Mateeva, M. Vassileva and T. Gueorguieva<br />

Agricultural University - Plovdiv, 12 Mendeleev St., 4000 Plovdiv, Bulgaria<br />

Summary<br />

From 1996/1999 field studies in <strong>the</strong> region <strong>of</strong> Jambol, Bulgaria, were performed to investigate side<br />

effects <strong>of</strong> chemical control <strong>of</strong> cereal aphids on different beneficial arthropods in winter wheat. Our<br />

results indicate that prophylactic spraying immediately before flowering was <strong>of</strong>ten to early to prevent<br />

a later built-up <strong>of</strong> aphid populations. In some years an accelerated increase <strong>of</strong> aphid density in<br />

combination with reduced immigration <strong>of</strong> aphid specific predatory such as Syrphidae, Coccinellidae<br />

and Chrysopidae could de observed in sprayed plots compared to non-treated control areas. When<br />

aphicides were applied according to <strong>the</strong> already existing economic threshold, aphid predators were<br />

eliminated by broad spectrum insecticides like Decis 2,5 EK 0,03% and Talstar 10 EK 0,03%. Side<br />

effects <strong>of</strong> an insecticide treatment on certain natural enemies <strong>of</strong> aphids could be reduced for by using<br />

more selective pesticides like Pirimicarb 0,1%. However, applying Pirimicarb occasionally lead to<br />

detrimental effects on aphid specific predators, mainly because <strong>the</strong> prey population was completely<br />

eliminated. Studies on <strong>the</strong> development <strong>of</strong> syrphid and coccinellid larvae in relation to aphid density<br />

showed that relative high numbers <strong>of</strong> <strong>the</strong>se predators could survive and complete <strong>the</strong>ir development<br />

even with limited food supply.<br />

Key words: pests, predators, populations, cereal aphids, pest control<br />

Introduction<br />

Agro-climatic conditions in Bulgaria are favourable for growth <strong>of</strong> over 80 varieties <strong>of</strong><br />

agricultural crops. Cereals are cultivated on about 21 million <strong>of</strong> decares, and are attacked<br />

various pests and diseases (Grigorov 1959 1964 1965 1990; Lyubenov 1961 1969; Lazarov et<br />

al. 1969).<br />

Aphids are among <strong>the</strong> economically most important pests in wheat, with Sitobion avenae<br />

Fab. and Rhopalosiphum padi L. as <strong>the</strong> predominant species (Vickerman 1977 1980; Lowe<br />

1982). Increasing use <strong>of</strong> fertilisers in Bulgaria resulted in more frequent aphid outbreaks.<br />

However, increased aphid density in cereals also resulted in a higher diversity <strong>of</strong> predators.<br />

Most <strong>of</strong>ten aphid predators belong to <strong>the</strong> families <strong>of</strong> Coccinellidae, Syrphidae and<br />

Chrysopidae have been reported as beneficials in integrated control systems, <strong>of</strong>ten associated<br />

with <strong>the</strong> use <strong>of</strong> selective pesticides (Beackman 1967; Sunderland et al. 1982; Hagen 1987). A<br />

number <strong>of</strong> studies have already been carried out to investigate <strong>the</strong> effect <strong>of</strong> pesticides on like<br />

Coccinella septempunctata L. (Col.: Coccinellidae), Episyrphus balteatus De Geer (Dipt.:<br />

Syrphidae) and Chrysoperla carnea Stephens (Neurop.: Chrysopidae) (e.g. De Clerco &<br />

Pietraszko 1983).<br />

The objective <strong>of</strong> <strong>the</strong> present study was to investigate side effects <strong>of</strong> pesticides used for<br />

aphid control on beneficial coccinellids, syrphids and chrysopids.<br />

139


140<br />

Material and methods<br />

The experiments were carried out in 1996/1999 in three replicates <strong>of</strong> square plots <strong>of</strong> 100 m.<br />

Winter wheat was sown in October using <strong>the</strong> cultivar ‘Sadovo 1’ at a seed rate, and fertiliser<br />

and herbicides treatments typical for <strong>the</strong> Jambol region. All <strong>the</strong> plots were fenced in mid-May<br />

with 60 cm high plastic sheets.<br />

The following treatments were compared: (i) untreated control, (ii) Decis 2,5 EK 0,03 %,<br />

(iii) Talstar 10 EK 0,03 %, and (iv) Pirimicarb 0,1 %. Applications were carried out at <strong>the</strong><br />

beginning <strong>of</strong> flowering <strong>of</strong> <strong>the</strong> winter wheat, i.e. on <strong>the</strong> 14th May 1996, <strong>the</strong> 10th May 1997,<br />

and on <strong>the</strong> 16th May 1998, respectively. Immediately after treatments, in each replicate<br />

aphids and natural enemies were monitored on plants in a square meter. The data was<br />

analysed by means <strong>of</strong> descriptive statistics, i.e. mean and standard deviation.<br />

Results and Discussion<br />

Table 1 shows <strong>the</strong> species composition <strong>of</strong> <strong>the</strong> most widely distributed predators in wheat<br />

crops. The greatest species variety was reported for Coccinellidae. The most abundant<br />

predator species were C. septempunctata and E. balteatus. According to Grigorov (1980), <strong>the</strong><br />

latter species shows a preference to develop in wheat crops with a high aphid density.<br />

Tab. 1. Species composition <strong>of</strong> predators recorded in wheat crops in <strong>the</strong> region <strong>of</strong> Jambol.<br />

Order Family Species<br />

Coleoptera Coccinellidae Coccinella septempunctata L.<br />

Coccinula quatuordecimpustulata (L.)<br />

Hippodamia variegata Goeze<br />

Propylaea quatuordecimpunctata L.<br />

Adalia spp.<br />

Hippodamia tredecimpunctata (L.)<br />

Neuroptera Chrysopidae Chrysoperla carnea Stephens<br />

Chrysoperla harrisii (Fitch)<br />

Chrysopa pallens Rambur<br />

Diptera Syrphidae Episyrphus balteatus De Geer<br />

Syrphus vitripennis Meigen<br />

Eupeodes latifasciatus (Macquart)<br />

Scaeva pyrastri (L.)<br />

Figure 1 shows <strong>the</strong> results <strong>of</strong> <strong>the</strong> monitoring between 1996 and 1998. A Decis and Talstar<br />

treatment considerably decreased <strong>the</strong> total numbers <strong>of</strong> all predators, and in particular <strong>of</strong><br />

syrphids.<br />

A similar trend was observed during <strong>the</strong> three years <strong>of</strong> studies. However, <strong>the</strong> low<br />

predator densities in 1997 we most likely also associated with unfavourable agro-climatic<br />

conditions in June 1997, that resulted in reduced aphid densities.


141<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Decis 2,5<br />

EK 0,03%<br />

Coccinellidae<br />

Talstar 10<br />

EK 0,03%<br />

Pirimicarb<br />

0,1%<br />

1996<br />

1997<br />

1998<br />

Average<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Decis 2,5<br />

EK 0,03%<br />

Chrysopidae<br />

Talstar 10<br />

EK 0,03%<br />

Pirimicarb<br />

0,1%<br />

1996<br />

1997<br />

1998<br />

Average<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Decis 2,5<br />

EK 0,03%<br />

Syrphidae<br />

Talstar 10<br />

EK 0,03%<br />

Pirimicarb<br />

0,1%<br />

1996<br />

1997<br />

1998<br />

Average<br />

Fig. 1. Abundance (in % <strong>of</strong> untreated control) <strong>of</strong> coccinellids, chrysopids and syrphids in<br />

insecticide-treated field plots in <strong>the</strong> Jambol region <strong>of</strong> Bulgaria in 1996 1997 and 1998, and on<br />

average (<strong>the</strong> latter ± SD).<br />

The Pirimicarp application proved to have <strong>the</strong> fewest side effects on aphid predators.<br />

Compared to <strong>the</strong> untreated control predator abundance was only reduced by 4 - 12%.<br />

However, <strong>the</strong> insecticide treatments greatly reduced <strong>the</strong> aphid densities, and thus largely<br />

limited <strong>the</strong> amount <strong>of</strong> available prey. Yet predators were able to sustain in <strong>the</strong> treated plots<br />

even at predator:prey ratios as low as 1:7 or 1:9 (Tab. 2)


142<br />

Tab. 2. Predator : aphid ratios in wheat crops in <strong>the</strong> Jambol region <strong>of</strong> Bulgaria (data from<br />

1996 - 1998).<br />

Versions 1996 1997 1998<br />

Untreated control 1:31 1:24 1:38<br />

Decis 2,5 EK 0,03% 1:10 1:10 1:9<br />

Talstar 10 EK 0,03% 1:12 1:10 1:7<br />

Pirimicarb 0,1 % 1:20 1:19 1:21<br />

Based on our results <strong>the</strong> following conclusions can be drawn:<br />

Predators in <strong>the</strong> wheat biocenose were mainly representatives <strong>of</strong> <strong>the</strong> families Coccinellidae,<br />

Chrysopidae and Syrphidae. A Decis 2,5 EK 0,03% and a Talstar 10 EK 0,03%<br />

treatment considerably reduced <strong>the</strong> density <strong>of</strong> predators in wheat crops. The predators were<br />

able to survive and complete <strong>the</strong>ir development even at very close predator:prey ratios.<br />

References<br />

Grigorov, St. 1959: Prinos kam prouchvane na listnite vashki po zhitnite rastenija v Bulgaria,<br />

BAN. (Bulg.)<br />

Grigorov, S. 1964: Prinos kam entom<strong>of</strong>aunata na Bulgaria. sp. Rastitelna zashtita.<br />

Grigorov, S. 1965: Entomotsenozata na zhitnite sas sljata povarhnost v S<strong>of</strong>ijsko. Nauchni<br />

trudove v tom XVI.<br />

Grigorov, S. 1990: Listni vashki, sp. Zemedelie, br.2, br.5.<br />

Ljubenov, J.A. 1961 1969: Divata zhitna rastitelnost kato rezerv i iztochnik za namaljavane<br />

na shvedskata muha ; sp. Selska misal, br.5.<br />

Lazarov, A. 1969: Shvedskata muha (Oscinosoma frit Diptera) kato neprijatel po zhitnite.<br />

Rastitelna zashtita, br.6.<br />

Blackman, R.L. 1967: Selection <strong>of</strong> applied prey by Adalia bipunctata L. and Coccinella<br />

septempunctata L.. Annals <strong>of</strong> Applied Biology 59: 331-338.<br />

Hagen, K.S. 1987: Nutritional ecology <strong>of</strong> terrestrial insect predators. In: Slansky, Jr. F. &<br />

Rodrigues, J. (eds): Nutritional ecology <strong>of</strong> insects, mites, spiders and related<br />

invertebrates. New York, John Wiley & Sons: 553-577.<br />

De Clerko, R. & Piertraszko, R. 1983: The influence <strong>of</strong> pesticides on <strong>the</strong> epigeal arthropoda<br />

fauna in winter wheat. Integrated and complex plant protection <strong>of</strong> field crops. Budapest<br />

4th - 9th July 1983. Hungary.<br />

Lowe, H.J.B. 1982: Some observations on <strong>the</strong> susceptibility and resistance <strong>of</strong> winter wheat to<br />

<strong>the</strong> aphid Sitobion avenae (F.) in Britain. Crop Protection 1 (4): 431-440.<br />

Sunderland, K.D., Chambers, R.J. & Stacey, D.L. 1982: Polyphagous predators and cereal<br />

aphids. Annual Report Glasshouse Crops Research Institute.<br />

Vickerman, G.P. 1977: Monitoring and forecasting insect pests <strong>of</strong> cereals. <strong>Proceedings</strong> British<br />

Crop Protection Conference: 227-231.<br />

Vickerman, G.P. 1980: Important changes in <strong>the</strong> numbers <strong>of</strong> insects in cereal fields. Game<br />

Conservancy Annual Review: 67-72.


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 143- 145<br />

Influence <strong>of</strong> <strong>the</strong> stubble burning<br />

on some pests and earth-worms density<br />

A. Mateeva 1 , D. Svetleva 1 , D. Andonov 2 and St. Stratieva 2<br />

1<br />

Agricultural University, Plovdiv 4000, Bulgaria<br />

2<br />

COS, Yambol 8600, Bulgaria<br />

Summary<br />

In recent years more attention has been paid to <strong>the</strong> impact <strong>of</strong> <strong>the</strong> stubble burning on soil pest<br />

population densities, as well as on some indicators <strong>of</strong> soil fertility such as earth-worms. An impartial<br />

assessment <strong>of</strong> this technique and an analysis <strong>of</strong> <strong>the</strong> consequences for <strong>the</strong> phytosanitary conditions <strong>of</strong><br />

<strong>the</strong> soils is <strong>of</strong> paramount importance for cereal growers in Bulgaria. The main objective <strong>of</strong> this<br />

investigation was to study <strong>the</strong> influence <strong>of</strong> stubble burning on some pests and also on earth-worms<br />

densities. The study was conducted on a 12 ha cereal plot in <strong>the</strong> Yambol district <strong>of</strong> Bulgaria. Results<br />

show that <strong>the</strong> stubble burning leads to a reduction <strong>of</strong> total pest numbers and earth-worm biomass.<br />

Key words: cereals, stubble burning, plough in <strong>the</strong> stubble, earth-worms, harmful pests<br />

Introduction<br />

Grain cereals take up two thirds <strong>of</strong> <strong>the</strong> cultivated area in Bulgaria and constitute <strong>the</strong> base <strong>of</strong><br />

<strong>the</strong> Bulgarian agricultural production. It is a common practice to burn <strong>the</strong> stubble after<br />

harvest. However, this process <strong>of</strong>ten gets out <strong>of</strong> control with all <strong>the</strong> inherent dangers for <strong>the</strong><br />

environment. Moreover, this practice needs a re-thinking particularly because <strong>of</strong> <strong>the</strong> increased<br />

interest in different forms <strong>of</strong> agriculture such as conventional, organic or bio-ecological. With<br />

this respect an impartial analysis and assessment <strong>of</strong> <strong>the</strong> phytosanitary conditions <strong>of</strong> <strong>the</strong> soils<br />

are needed. In addition, a comprehensive understanding <strong>of</strong> <strong>the</strong> farmers’ reasoning is required.<br />

The stubble management and <strong>the</strong> consequences for <strong>the</strong> soil ecosystem has recently attracted a<br />

lot <strong>of</strong> scientific interest (Haines et al., 1990; Chan & Heenan, 1993; Double et al., 1994). The<br />

main objective <strong>of</strong> <strong>the</strong> present study was to investigate <strong>the</strong> influence <strong>of</strong> stubble burning on<br />

some pests and earth-worms densities in Bulgaria.<br />

Material and methods<br />

The study was conducted on a 12 ha cereal plot in <strong>the</strong> Yambol district <strong>of</strong> Bulgaria. The field<br />

was separated into three experimental treatments with 10 m <strong>of</strong> protection strip. The stubble<br />

burning (<strong>the</strong> first treatment) was conducted directly after harvest and after 15 ml/m 2<br />

precipitation. In <strong>the</strong> second treatment <strong>the</strong> stubble was ploughed in. The third treatment was as<br />

a control without any agricultural practices. In every field 10 samples were taken from 1 m 2<br />

marked sectors in <strong>the</strong> plough layer with a thickness <strong>of</strong> 10 - 25 cm. Pests were analysed and<br />

determined at <strong>the</strong> Department <strong>of</strong> Entomology in <strong>the</strong> Agricultural University in Plovdiv.<br />

Results and discussion<br />

In Table 1 data on <strong>the</strong> density <strong>of</strong> different life stages <strong>of</strong> harmful coleopteran and lepidopteran<br />

species in <strong>the</strong> superficial ploughing layer is presented.


144<br />

Tab. 1. Average density <strong>of</strong> harmful species <strong>of</strong> <strong>the</strong> orders Coleoptera and Lepidoptera.<br />

Average density per m 2<br />

Treatments Order Coleoptera Order Lepidoptera<br />

Larvae Pupae Adults Larvae Pupae<br />

Stubble burning 2 14 3 2 12<br />

Ploughing in <strong>the</strong> stubble 1 3 1 – 1<br />

Control 13 16 4 3 13<br />

Higher numbers <strong>of</strong> Coleoptera were recorded in <strong>the</strong> control treatment (i.e. without<br />

applying any agricultural practices). The greatest reduction in density <strong>of</strong> Coleoptera and<br />

Lepidoptera were observed in <strong>the</strong> second treatment (with ploughing <strong>of</strong> <strong>the</strong> stubble). Ploughing<br />

most likely killed a great proportion <strong>of</strong> <strong>the</strong> insects in <strong>the</strong> superficial ploughing layer. In<br />

contrast stubble burning did not greatly affect <strong>the</strong> density <strong>of</strong> coleopteran and lepidopteran pest<br />

in <strong>the</strong> soil surface, particularly that <strong>of</strong> <strong>the</strong> pupae. The latter is probably due to two important<br />

factors: first <strong>the</strong> pupal stages are situated in depth beyond <strong>the</strong> ploughing layer, and second<br />

<strong>the</strong>y are relatively <strong>the</strong> most stable life stages <strong>of</strong> insects, particularly with regard to various<br />

adverse ecological and anthropogenic effects. The control treatment was characterised with<br />

<strong>the</strong> relatively highest densities and diversities <strong>of</strong> forms and stages <strong>of</strong> harmful species. In this<br />

undisturbed environment all development stages <strong>of</strong> <strong>the</strong> two insect orders (certainly except for<br />

adult butterflies) were found. Our results on <strong>the</strong> influence <strong>of</strong> different agricultural practices in<br />

cereal fields after harvest coincide with those <strong>of</strong> Pankhurst et al. (1995).<br />

Data on biomass <strong>of</strong> earth-worms are presented in Table 2.<br />

Tab. 2. Average density <strong>of</strong> earth-worms.<br />

Average density per m 2<br />

Treatments on 25.07.1999 on 15.08.1999<br />

Numbers g/m 2 Numbers g/m 2<br />

Stubble burning 7 105 8 118<br />

Ploughing in <strong>the</strong> stubble 21 360 43 650<br />

Control 27 430 19 380<br />

On <strong>the</strong> first sampling occasion in June 1999, <strong>the</strong> highest density was recorded in <strong>the</strong><br />

control treatment, followed by <strong>the</strong> ploughing treatment. The differences in absolute values<br />

between <strong>the</strong> two treatments in June were not substantially high. However, on <strong>the</strong> second<br />

sampling occasion in August 1999 higher numbers <strong>of</strong> earth-worms were recorded in <strong>the</strong><br />

ploughing compared to <strong>the</strong> control treatment. This probably reflects <strong>the</strong> favourable effect <strong>of</strong><br />

ploughing on <strong>the</strong> density <strong>of</strong> earth-worms which are indicators for <strong>the</strong> intensive soil formation<br />

activity. Lowest numbers <strong>of</strong> earth were always recorded in <strong>the</strong> stubble burning treatment.<br />

Based on our data we thus conclude that stubble burning leads to reduction <strong>of</strong> density <strong>of</strong><br />

coleopteran and lepidopteran pests and also to sharp decrease in <strong>the</strong> earth-worms biomass. An


145<br />

especially favourable practice for <strong>the</strong> purposes <strong>of</strong> a more ecologically oriented agriculture we<br />

thus recommend <strong>the</strong> ploughing in <strong>of</strong> <strong>the</strong> stubble.<br />

References<br />

Chan, K. & Heenan, I.D.P. 1993: Surface hydraulic properties <strong>of</strong> a red earth under continuos<br />

cropping with different management practices (wheat-lupine rotation). Australian Journal<br />

<strong>of</strong> Soil Research 31(1): 13-24.<br />

Double, B.M., Kirkegaard, I.A. & Buckerfield, I.C. 1994: Short term effects <strong>of</strong> tillage and<br />

stubble management on earth-worm populations in cropping systems in sou<strong>the</strong>rn New<br />

South Waks. Australian Journal <strong>of</strong> Soil Research 45(7): 1587-1600.<br />

Haines, P.I. & Uren, N.C. 1990: Effects <strong>of</strong> conservation tillage farming on soil microbial<br />

biomass, organic matter and earth-worm populations in North-Eastern Victoria.<br />

Australian Journal <strong>of</strong> Experimental Agriculture 30(3): 365-371.<br />

Pankhurst, C.E., Hawke, B.G., McDonald, H.I., Kirkby, C.A., Buckerfield, I.C., Michelsen,<br />

P., Brien, K.A. & Gupta, V.V.Sr. 1995: Evaluation <strong>of</strong> soil biological properties as<br />

potential bioindicators <strong>of</strong> soil health. Australian Journal <strong>of</strong> Experimental Agriculture<br />

35(7):1015-1028.


146


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 147- 151<br />

A study <strong>of</strong> aphid predation by Coccinella septempunctata L.<br />

(Coleoptera: Coccinellidae) using gut dissection<br />

Holger Triltsch<br />

BBA, Institute for Integrated Plant Protection, Stahnsdorfer Damm 81,<br />

D-14532 Kleinmachnow, Germany<br />

Short communication<br />

Since 1994 a long-term investigation on <strong>the</strong> diet <strong>of</strong> Coccinella septempunctata L. under field<br />

conditions has been carried out. Because <strong>the</strong> essential aphid prey is present only during a<br />

restricted time period, aphidophagous ladybirds have to consume a lot <strong>of</strong> different alternative<br />

food types, e.g. pollen, nectar, or o<strong>the</strong>r arthropods (Putman, 1964; Hemptinne & Desprets,<br />

1986). Most <strong>of</strong> our knowledge on <strong>the</strong> diet <strong>of</strong> C. septempunctata is based on laboratory<br />

experiments or on a few observations <strong>of</strong> ladybird individuals. Therefore, little is known about<br />

<strong>the</strong> relative importance <strong>of</strong> <strong>the</strong> different food types under field conditions. Moreover, we are<br />

not able to quantify <strong>the</strong> amount <strong>of</strong> aphids eaten by a ladybird in <strong>the</strong> field.<br />

To get some more information about <strong>the</strong> diet composition <strong>of</strong> C. septempunctata during<br />

<strong>the</strong> course <strong>of</strong> <strong>the</strong> year and in different locations this long-term field survey was undertaken.<br />

Additionally, a method to quantify aphid consumption under field conditions was developed.<br />

To date, about 2,000 adults and 500 larvae <strong>of</strong> C. septempunctata have been dissected and<br />

<strong>the</strong>ir gut content analysed. Ladybirds were collected in three different agricultural areas. At<br />

<strong>the</strong> Berlin-Staaken (BS) site ladybird adults were collected throughout <strong>the</strong> whole year in<br />

different crop and non-crop habitats. At <strong>the</strong> two o<strong>the</strong>r sampling sites, Nor<strong>the</strong>rn Flaeming (NF)<br />

and Magdeburger Boerde (MB), adults <strong>of</strong> C. septempunctata were collected in June and July<br />

in cereals. Since 1995 ladybird larvae were also investigated.<br />

While capturing <strong>the</strong> ladybirds for fur<strong>the</strong>r investigation, o<strong>the</strong>r potential prey arthropods, as<br />

well as certain plant material were also sampled. In many cases this material was useful in<br />

identifying food items present in <strong>the</strong> alimentary canal. In addition, feeding experiments with<br />

prey types, previously never detected in <strong>the</strong> gut but mentioned as prey <strong>of</strong> predaceous<br />

coccinellids, namely ladybird eggs and larvae <strong>of</strong> Oulema spp. (Col.: Chrysomelidae) (Rogers<br />

et al., 1972; Agarwala & Dixon, 1992), were carried out under laboratory conditions. An<br />

illustrated small catalogue <strong>of</strong> common food remains found in <strong>the</strong> alimentary canal <strong>of</strong><br />

C. septempunctata has already been published (Triltsch, 1999).<br />

All ladybird individuals were captured by sweepnetting or after visual searches. They<br />

were rendered inactive at low temperatures (below -10°C) and killed in 80% ethanol to<br />

prevent any excretion. The ladybirds were dissected under a binocular microscope and <strong>the</strong><br />

alimentary canal <strong>of</strong> each individual was removed. The food remains were inspected in<br />

glycerol and classified as belonging to <strong>the</strong> following food types: (1) aphids, (2) non-aphid<br />

arthropods, (3) fungal spores, (4) pollen and o<strong>the</strong>r plant material, (5) inorganic material, and<br />

(6) unidentifiable items. Frequency <strong>of</strong> a certain food type was defined as <strong>the</strong> percentage (%)<br />

<strong>of</strong> individuals <strong>of</strong> each sample containing that type <strong>of</strong> food. The category “combined meal”<br />

was adapted from Sunderland et al. (1995) and includes cases where more than one food type<br />

was found in a single gut. Since 1995 a more detailed identification <strong>of</strong> prey species within <strong>the</strong><br />

category “non-aphid arthropods” was performed.<br />

147


148<br />

Gut dissection <strong>of</strong> C. septempunctata collected in <strong>the</strong> field or fed under laboratory<br />

conditions indicates a very different degree <strong>of</strong> food recovery in <strong>the</strong> alimentary canal. If<br />

ladybird adults were fed with conspecific eggs, no remains were recovered which allowed for<br />

<strong>the</strong>ir identification. After feeding on Oulema spp. larvae it was quite difficult to find any<br />

characteristic food remains. On <strong>the</strong> o<strong>the</strong>r hand, smaller prey like aphids or Thysanoptera were<br />

easily detectable during gut dissection. These smaller arthropods are being devoured<br />

completely by C. septempunctata, which makes it easy to find typical chitin fragments or in<br />

some cases nearly intact bodies (especially in Acari, Collembola and Thysanoptera). Larger<br />

arthropods are eaten only in parts and <strong>the</strong> success <strong>of</strong> prey recovery during gut dissection<br />

depends on <strong>the</strong> amount <strong>of</strong> sclerotised body parts consumed.<br />

Tab. 1. Frequency <strong>of</strong> food types in <strong>the</strong> gut <strong>of</strong> adult Coccinella septempunctata L. collected at<br />

Berlin-Staaken in 1994-1997.<br />

Month Habitat/ crop<br />

Number<br />

<strong>of</strong> adults<br />

dissected<br />

(samples)<br />

meal*<br />

remains <strong>of</strong><br />

aphids<br />

non-aphid<br />

arthropods<br />

fungal spores<br />

Proportion (%) with<br />

pollen<br />

inorganic<br />

material<br />

not identified<br />

emptz gut<br />

combined<br />

II hibernacula 42 (1) 0.0 0.0 0.0 0.0 0.0 16.7 83.3 0.0<br />

III hibernacula 98 (4) 0.0 0.0 0.0 0.0 1.0 2.0 97.0 0.0<br />

IV hibern ,fallow 65 (2) 10.8 10.8 15.4 9.2 9.2 16.9 43.1 9.2<br />

V fallow,oat,rye 110 (4) 50.0 27.3 45.5 22.7 49.1 10.9 9.1 70.0<br />

VI oat, wheat 261 (10) 88.5 12.6 56.3 7.3 39.5 3.1 1.5 68.9<br />

VII oat 138 (4) 82.6 34.8 86.2 14.5 24.6 5.8 2.9 84.4<br />

VIII fallow,oat,maize 67 (3) 44.8 23.9 68.7 13.4 41.8 6.0 7.5 75.8<br />

IX fallow,grassland 176 (6) 36.4 9.7 55.1 23.3 24.4 7.4 24.4 69.2<br />

X fallow,hibern. 125 (3) 4.8 0.8 12.0 1.6 12.0 8.0 75.2 51.6<br />

XI fallow,hibern. 32 (2) 0.0 0.0 0.0 3.1 0.0 0.0 96.9 0.0<br />

XII hibernacula 32 (2) 0.0 0.0 0.0 0.0 0.0 0.0 100.0 0.0<br />

* Number <strong>of</strong> individuals with more than one food type in gut as percentage <strong>of</strong> <strong>the</strong> number <strong>of</strong><br />

individuals with food in gut<br />

Table 1 shows <strong>the</strong> diet composition <strong>of</strong> C. septempunctata adults at location BS (Triltsch,<br />

1999). During <strong>the</strong> course <strong>of</strong> a year remarkable changes in <strong>the</strong> frequency <strong>of</strong> consumption <strong>of</strong><br />

certain food types were observable. Aphid consumption was detectable from April until<br />

October. In May and June aphids were <strong>the</strong> most frequent food type. The highest proportion <strong>of</strong><br />

C. septempunctata adults with aphid remains present in <strong>the</strong>ir guts was found in June (88.5%).<br />

At <strong>the</strong> same time <strong>the</strong> proportion <strong>of</strong> adults with empty guts was lowest. Fungal spores were<br />

found to be nearly as frequent as aphids. This was an unexpected result and difficult to<br />

explain. The phenomenon could indicate feeding on honeydew because Alternaria spp., <strong>the</strong>


149<br />

most frequently found fungi, usually grows on honeydew. Non-aphid arthropods were<br />

consumed generally not so frequently. Within that food type Thysanoptera, Collembola,<br />

Acari, Hymenoptera, Diptera (larvae), Coccinellidae (larvae), and Chrysomelidae (larvae)<br />

were identified. Two periods <strong>of</strong> frequent pollen feeding were found, i.e. a first peak in May<br />

and ano<strong>the</strong>r in September.<br />

In June and July, no difference in adult diet was found between <strong>the</strong> three sampling sites.<br />

Also <strong>the</strong> composition <strong>of</strong> <strong>the</strong> larval diet was relatively similar to that <strong>of</strong> <strong>the</strong> adults.<br />

As already mentioned, <strong>the</strong> gut dissection method has its limits because prey without<br />

sclerotised parts, like ladybird eggs, are not detectable. On <strong>the</strong> o<strong>the</strong>r hand, smaller prey, like<br />

aphids, are relatively easy to identify. Moreover, aphid remains present in <strong>the</strong> gut could be<br />

quantified. In combination with experiments to study <strong>the</strong> rate <strong>of</strong> recovery <strong>of</strong> a certain aphid<br />

meal, gut dissection could be a suitable method for estimating aphid consumption rates under<br />

field conditions.<br />

Therefor, <strong>the</strong> number <strong>of</strong> certain aphid body fragments present in <strong>the</strong> alimentary canal was<br />

additionally obtained from gut dissection <strong>of</strong> C. septempunctata individuals collected in <strong>the</strong><br />

field. Two aphid body parts, <strong>the</strong> tarsus and <strong>the</strong> ultimate rostral segment, were chosen for<br />

quantification. These two chitin fragments are easily detectable and were seldom crushed<br />

during food intake and digestion. Some preliminary results from <strong>the</strong>se investigations are<br />

given in Table 2. The mean number <strong>of</strong> recovered aphid fragments per individual increased<br />

from April to June. Although C. septempunctata males were able to find some aphid prey<br />

earlier in <strong>the</strong> season, mean numbers <strong>of</strong> recovered aphid fragments were three times higher in<br />

females compared with males. The largest amount <strong>of</strong> aphid remains was observed in third and<br />

fourth instar larvae.<br />

Tab. 2. Quantity <strong>of</strong> two aphid’s body fragments present in <strong>the</strong> gut <strong>of</strong> Coccinella<br />

septempunctata L. collected at Berlin-Staaken in 1998.<br />

Date<br />

Mean number <strong>of</strong><br />

ultimate rostral segment / tarsus (Individuals dissected) in<br />

Males Females Larvae<br />

April 14 0.3 / 1.3 (6) 0.0 / 0.0 (2) --<br />

April 22 0.4 / 2.1 (23) 1.2 / 5.8 (12) --<br />

April 30 0.1 / 0.7 (9) 1.7 / 6.9 (18) --<br />

May 12 1.2 / 5.3 (37) 2.7 / 11.6 (56) --<br />

May 20 1.5 / 5.5 (4) 4.4 / 16.4 (20) 2.2 / 5.0 (6) 2. instar<br />

June 02 1.3 / 6.5 (46) 3.5 / 14.6 (51) 4.7 / 20.4 (300) 3./4. instar<br />

June 13 2.5 / 8.7 (8) 4.5 / 17.3 (11) 5.5 / 23.8 (290) 3./4. instar<br />

In some preliminary experiments, C. septempunctata adults were fed a certain number <strong>of</strong><br />

aphids before dissection. These experiments took place under laboratory conditions with<br />

20±3 °C, 65% r.h., and L:D=16:8 h. The ladybirds were kept individually in small plastic<br />

boxes (100 ml), starved for 48 h and <strong>the</strong>n fed for 5 days with a certain amount <strong>of</strong> Sitobion<br />

avenae F. (Hom.: Aphididae) (3, 6, 9, 12, 15, 20, 25, or 30 aphids/day). Two significant linear<br />

relationships were calculated between <strong>the</strong> number <strong>of</strong> aphids consumed per day and <strong>the</strong><br />

number <strong>of</strong> aphid chitin fragments present in <strong>the</strong> gut from dissection <strong>of</strong> <strong>the</strong>se ladybird


150<br />

individuals (fig. 1). Although at this time relatively little is known about <strong>the</strong> digestion rate in<br />

ladybirds, <strong>the</strong> observed significant relations lead to <strong>the</strong> conclusion, that it may be possible to<br />

estimate aphid consumption rates for C. septempunctata derived from <strong>the</strong> number <strong>of</strong> aphid<br />

chitin fragments present in <strong>the</strong> gut. This would <strong>of</strong>fer a new chance to intensively study aphid<br />

feeding <strong>of</strong> predaceous ladybirds under natural conditions and may answer <strong>the</strong> question: How<br />

many aphids are consumed by one ladybird in <strong>the</strong> field? A comparison <strong>of</strong> gut dissection<br />

results with <strong>the</strong> estimated aphid abundance at <strong>the</strong> location <strong>the</strong> ladybirds were collected would<br />

be very useful. From such a study important information about <strong>the</strong> feeding habits <strong>of</strong> ladybirds<br />

and how <strong>the</strong>y utilise our agricultural landscape could be ga<strong>the</strong>red.<br />

Aphid body fragments present in gut [number/adult]<br />

100<br />

Tarsus<br />

y=1.82x (R²=0.46) n=56<br />

<br />

80<br />

<br />

60<br />

<br />

<br />

<br />

40<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

20<br />

<br />

<br />

0<br />

<br />

<br />

40<br />

Ultimate rostral segment<br />

20<br />

y=0.43x (R²=0.48) n=56<br />

<br />

<br />

<br />

0<br />

<br />

5 10 15 20 25 30<br />

Aphid consumption [aphids/d]<br />

Fig. 1. Recovery <strong>of</strong> certain aphid body fragments during gut dissection after feeding<br />

C. septempunctata adults with S. avenae under laboratory conditions.<br />

Acknowledgement<br />

The author acknowledges <strong>the</strong> „Deutsche Forschungsgemeinschaft“ (DFG) for funding <strong>the</strong><br />

researcher and <strong>the</strong> Federal Biological Research Centre for Agriculture and Forestry (BBA) for<br />

facilitating <strong>the</strong> research. I would like to thank Pr<strong>of</strong>. J.-L. Hemptinne and Dr. K.D. Sunderland<br />

for <strong>the</strong>ir helpful comments and suggestions to that field <strong>of</strong> work.<br />

References<br />

Agarwala, B.K. & Dixon, A.F.G. 1992: Laboratory studies <strong>of</strong> cannibalism and interspecific<br />

predation in ladybirds. Ecol. Entomol. 17: 303-309.<br />

Hemptinne, J.-L. & Desprets, A. 1986: Pollen as a spring food for Adalia bipunctata. In:<br />

Ecology <strong>of</strong> Aphidophaga. Hodek, I. (ed.). Academia, Prague: 29-35.<br />

Putman, W.L. 1964: Occurrence and food <strong>of</strong> some coccinellids (Coleoptera) in Ontario peach<br />

orchards. Can. Entomologist 96: 1149-1155.<br />

Roger, C.E., Jackson, H.B., Angalet, G.W. & Eikenbary, R.D., 1972: Biology and life history<br />

<strong>of</strong> Propylea 14-punctata (Col.: Coccinellidae), an exotic predator <strong>of</strong> aphids. Ann.<br />

Entomol. Soc. America 65: 648-650.


Sunderland, K.D., Lövei, G.L. & Fenlon, J. 1995: Diets and reproductive phenologies <strong>of</strong> <strong>the</strong><br />

introduced ground beetles Harpalus affinis and Clivinia australasiae (Coleoptera:<br />

Carabidae) in New Zealand. Aust. J. Zool 43: 39-50.<br />

Triltsch, H. 1999: Food remains in <strong>the</strong> guts <strong>of</strong> Coccinella septempunctata L. (Coleoptera:<br />

Coccinellidae) adults and larvae. Eur. J. Entomol. 96(4): 355-364.<br />

151


152


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 153- 163<br />

Arthropod complex <strong>of</strong> winter wheat crops and its seasonal dynamics<br />

Valentina M. Afonina, Wladimir B. Tshernyshev, Inna I. Soboleva-Dokuchaeva,<br />

Alexander V. Timokhov, Olga V. Timokhova and Rimma R. Seifulina<br />

Dept. <strong>of</strong> Entomology, Faculty <strong>of</strong> Biology, Moscow State University, 119899, Moscow, Russia<br />

Summary<br />

The investigations were carried out in Moscow region for two years and in Krasnodar region (<strong>the</strong><br />

South <strong>of</strong> Russia) for one year. The size <strong>of</strong> <strong>the</strong> fields were 12-14 ha and 64 ha, respectively. We used<br />

entomological sweep netting and pitfall traps simultaneously. Samples were taken in <strong>the</strong> field centre,<br />

on <strong>the</strong> field edges and in adjoining biotopes (forests, forest belts, perennial herbs crops). The site on<br />

<strong>the</strong> halfway to <strong>the</strong> centre was added in Krasnodar region. Spatial and temporal dynamics <strong>of</strong> mass<br />

species <strong>of</strong> arthropods was described. Some <strong>of</strong> species were field residents and able to overwinter <strong>the</strong>re,<br />

however, <strong>the</strong> most <strong>of</strong> species migrated to <strong>the</strong> field from its grassy margins or from o<strong>the</strong>r biotopes <strong>of</strong> a<br />

landscape.<br />

Key words: agro-ecosystem, pests, natural enemies, migrations, arthropods, insects, mites<br />

Introduction<br />

Successful management <strong>of</strong> natural enemies would be impossible without knowledge <strong>of</strong><br />

arthropod complex and its distribution both in <strong>the</strong> agro-landscape and within <strong>the</strong> field during<br />

<strong>the</strong> vegetation period. Such seasonal migrations to <strong>the</strong> field from field margins in spring and<br />

backwards in autumn were estimated for separate groups <strong>of</strong> arthropods, among o<strong>the</strong>rs for<br />

ground beetles, some staphylinid beetles, and many spiders (Dennis & Fry, 1992; Kiss et al.,<br />

1994; Kromp & Steinberger, 1992; Peter et al., 1999; So<strong>the</strong>rton, 1985; Thomas et al., 1992;<br />

Toth & Kiss, 1999; Wallin, 1985). Seasonal migrations <strong>of</strong> many insects inhabiting <strong>the</strong><br />

vegetation were described by Melnichenko (1949).<br />

Here we describe seasonal changes <strong>of</strong> distribution both on <strong>the</strong> ground surface and in <strong>the</strong><br />

vegetation level, <strong>of</strong> mass species populations <strong>of</strong> pests and <strong>the</strong>ir natural enemies inhabiting <strong>the</strong><br />

winter wheat crops in two regions <strong>of</strong> Russia. We consider not only migrations into or out <strong>of</strong><br />

<strong>the</strong> field but also seasonal changes <strong>of</strong> populations distribution within <strong>the</strong> field.<br />

Investigation regions and methods<br />

Our investigations were carried out in winter wheat crops in Moscow region in 1996 and 1998<br />

and in Krasnodar region (<strong>the</strong> South <strong>of</strong> Russia) in 1999. The size <strong>of</strong> <strong>the</strong> fields were 12-14 ha in<br />

Moscow and 64 ha in Krasnodar, respectively. The observations lasted from snow melting in<br />

<strong>the</strong> spring until <strong>the</strong> harvest or snow fall in autumn. We used pitfall traps and sweep netting at<br />

<strong>the</strong> same sites. Samples were taken every eight or ten days in <strong>the</strong> centre <strong>of</strong> <strong>the</strong> field, on <strong>the</strong><br />

field edges (7 - 10 m into <strong>the</strong> field), on <strong>the</strong> field grassy margins and in adjoined biotopes<br />

(forest, forest belts, crops <strong>of</strong> perennial herbs). In Krasnodar samples were taken on halfway<br />

from <strong>the</strong> edge to <strong>the</strong> centre as well. The distance between <strong>the</strong> nearest edge and <strong>the</strong> centre was<br />

about 180 m in Moscow and 400 m in Krasnodar. In Moscow region we used 71 pitfall traps<br />

in 1996 (15 traps in <strong>the</strong> centre <strong>of</strong> <strong>the</strong> field, 24 traps on <strong>the</strong> field edges and 32 traps on margins,<br />

153


154<br />

respectively) and 65 traps in 1998 (10 traps in centre, 20 on edges, 20 on margins and 15 in<br />

adjoining biotopes, respectively).<br />

In Krasnodar region we used 50 pitfall traps in 5 biotopes (centre, halfway from <strong>the</strong><br />

centre to <strong>the</strong> edge, field edges, margins, adjoining forest belts): 10 in every biotop. Plastic<br />

glasses (0,5 l) containing some amount <strong>of</strong> wet soil (in 1996) or 4 % water solution <strong>of</strong> formalin<br />

(in 1998 and 1999) were used as traps. In 1996, pitfall traps were exposed for four days and<br />

<strong>the</strong> following four days <strong>the</strong>y were closed. In 1998 and 1999, <strong>the</strong>se periods were<br />

correspondingly five and five days respectively.<br />

We used standard entomological sweep netting techniques. Every sample contained <strong>the</strong><br />

catches <strong>of</strong> 25 strokes by net. In 1996 we collected 15 samples in <strong>the</strong> centre, 24 on <strong>the</strong> field<br />

edges and 32 on <strong>the</strong> margins; in 1998 eight samples in <strong>the</strong> centre, 16 on <strong>the</strong> edges, 16 on <strong>the</strong><br />

margins and 12 in <strong>the</strong> adjoining biotopes. In 1999 we had eight samples in every site.<br />

Results<br />

Here we will consider behaviour <strong>of</strong> herbivores and predators. The density <strong>of</strong> parasites was<br />

very low in all samples. Our results concerning mass species <strong>of</strong> arthropods are presented in<br />

fig. 1-20. Average catches per sample are shown in ordinate axes.<br />

Herbivores<br />

Chrysomelid beetles, Phyllotreta spp. (fig. 1-2), and Oscinella spp. (Dipt.: Chloropidae)<br />

(fig. 3) colonise <strong>the</strong> field at once and prefer its central parts. We can not exclude that some <strong>of</strong><br />

<strong>the</strong>se insects were able to overwinter in field plant debris. Aphids, Sitobion avenae F. (fig. 4),<br />

began to colonise <strong>the</strong> field from its edges. The same preference <strong>of</strong> field edges may be<br />

observed in overwintering bugs, Eurygaster integriceps Put. (Hem.: Scutelleridae) (fig. 5-6)<br />

and Lygus rugulipennis Popp. (Hem.: Miridae) (fig. 7). However, <strong>the</strong> following generations <strong>of</strong><br />

<strong>the</strong>se aphids and bugs (adults as well as older larvae) were more abundant in <strong>the</strong> central part<br />

<strong>of</strong> <strong>the</strong> field. On <strong>the</strong> contrary, <strong>the</strong> chrysomelid Oulema melanopus L. (fig. 8), was concentrated<br />

on field margins in <strong>the</strong> beginning <strong>of</strong> spring and later <strong>the</strong>n colonised <strong>the</strong> field edges. During<br />

<strong>the</strong> whole season this beetle species was only very seldom found in <strong>the</strong> central part <strong>of</strong> <strong>the</strong><br />

field. The second peak <strong>of</strong> abundance on field margins was most likely caused by an<br />

insecticide in <strong>the</strong> middle <strong>of</strong> June.<br />

120 N<br />

100<br />

80<br />

60<br />

40<br />

Margin<br />

Edge<br />

Centre<br />

20<br />

0<br />

3.5<br />

11.5<br />

19.5<br />

28.5<br />

5.6<br />

13.6<br />

22.6<br />

30.6<br />

9.7<br />

18.7<br />

26.7<br />

Dates<br />

Fig. 1. Phyllotreta vittula (Redt.) Moscow, 1996. Sweeping.


155<br />

60<br />

N<br />

50<br />

40<br />

30<br />

20<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

10<br />

0<br />

27.3<br />

5.4<br />

15.4<br />

25.4<br />

6.5<br />

15.5<br />

26.5<br />

4.6<br />

12.6<br />

23.6<br />

3.7<br />

Dates<br />

Fig. 2. Phyllotreta spp. Krasnodar, 1996. Sweeping.<br />

N<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Margin<br />

Edge<br />

Centre<br />

3.5<br />

11.5<br />

19.5<br />

28.5<br />

5.6<br />

13.6<br />

22.6<br />

30.6<br />

9.7<br />

18.7<br />

26.7<br />

Dates<br />

Fig. 3. Oscinella frit L. & O. pusilla Mg. Moscow, 1996. Sweeping.<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

6.5<br />

15.5<br />

26.5<br />

4.6<br />

12.6<br />

23.6<br />

Dates<br />

Fig. 4. Sitobion avenae F. Krasnodar, 1999. Sweeping.


156<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

N<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

15.5<br />

26.5<br />

4.6<br />

12.6<br />

23.6<br />

3.7<br />

Dates<br />

Fig. 5. Eurygaster integriceps Put. (imago). Krasnodar. 1999. Sweeping.<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

N<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

4.6<br />

12.6<br />

23.6<br />

3.7<br />

Dates<br />

Fig. 6. Eurygaster integriceps Put. (larvae III-V). Krasnodar. 1999. Sweeping.<br />

350<br />

300<br />

N<br />

250<br />

200<br />

150<br />

100<br />

Margin<br />

Edge<br />

Centre<br />

50<br />

0<br />

3.5<br />

11.5<br />

19.5<br />

28.5<br />

5.6<br />

13.6<br />

22.6<br />

30.6<br />

9.7<br />

18.7<br />

26.7<br />

Dates<br />

Fig. 7. Lygus rugulipennis Popp. Moscow, 1996. Sweeping.


157<br />

12 N<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

27.3<br />

5.4<br />

15.4<br />

25.4<br />

6.5<br />

15.5<br />

26.5<br />

4.6<br />

Dates<br />

12.6<br />

23.6<br />

3.7<br />

Fig. 8. Oulema melanopus L. Krasnodar, 1999. Sweeping.<br />

Predators<br />

We observed many carnivorous red mites Trombidiiformes (fig. 9) in trap samples in spring<br />

in Krasnodar, especially in <strong>the</strong> centre <strong>of</strong> <strong>the</strong> field, indicating that <strong>the</strong>y can hibernate in <strong>the</strong><br />

field. Later <strong>the</strong>ir abundance decreased sharply. Some spiders living in vegetation such as<br />

Tetragnatha extensa (L.) (Araneae: Tetragnathidae) (fig. 10), Xysticus ulmi Hahn (Araneae:<br />

Thomisidae) (fig. 11) moved to <strong>the</strong> field from its margins. Later <strong>the</strong> former species distributed<br />

evenly all over <strong>the</strong> field and its margins, whereas <strong>the</strong> latter species preferred field edges and<br />

margins, although could be found even in <strong>the</strong> centre. Spiders living in <strong>the</strong> vegetation like<br />

Misumenops tricuspidatus Fab. (Araneae: Thomisidae) (fig. 12), remained on <strong>the</strong> margins and<br />

could hardly be caught in <strong>the</strong> field. On <strong>the</strong> contrary, Hypsosinga pygmaea (Sand.) (Araneae:<br />

Araneidae) (fig. 13) was found in <strong>the</strong> field mainly in <strong>the</strong> centre from <strong>the</strong> very beginning <strong>of</strong><br />

vegetation season and avoided its margins and forest belts. Ground dwelling spiders <strong>of</strong> <strong>the</strong><br />

family Lycosidae seem to overwinter in <strong>the</strong> field and prefer it’s <strong>the</strong> central part (for instance,<br />

Pardosa agrestis (Westr.) and Trochosa spp. in Krasnodar region). Ground spider<br />

Oedothorax apicatus (Blackw.) (Araneae: Linyphiidae) inhabited both <strong>the</strong> field and its<br />

margins only in spring, but were abundant within <strong>the</strong> field in summer. Good-flying carabid<br />

beetles, e.g. Agonum muelleri (Hbst.) (fig. 14), were able to colonise <strong>the</strong> whole field at once.<br />

These beetles always preferred <strong>the</strong> central part <strong>of</strong> <strong>the</strong> field. Ground beetles Brachinus spp.<br />

(fig. 15) emerged at first on margins and in forest belts. Later <strong>the</strong>y occupied <strong>the</strong> edge <strong>of</strong> <strong>the</strong><br />

field and <strong>the</strong>n <strong>the</strong> whole surface <strong>of</strong> <strong>the</strong> field. On <strong>the</strong> contrary, Agonum dorsale Pont. (Col.:<br />

Carabidae) (fig. 16) appeared in forest belts in early spring and <strong>the</strong>n on <strong>the</strong> field margins;<br />

however, this species was nearly absent in field catches during <strong>the</strong> whole season. The second<br />

ga<strong>the</strong>ring <strong>of</strong> beetles on margins in June (fig. 15-16) could be explained by <strong>the</strong>ir mass<br />

migrations out <strong>of</strong> <strong>the</strong> field after spraying <strong>of</strong> insecticides. The poor flying carabids<br />

Pterostichus cupreus L. (fig. 17) and Harpalus rufipes (De Geer) (fig. 19), gradually<br />

colonised <strong>the</strong> field in Moscow region from its margins; <strong>the</strong> latter species replaced <strong>the</strong> former<br />

in margins in spring and in <strong>the</strong> field in summer. Pterostichus cupreus was <strong>the</strong> most common<br />

beetle in traps catches in <strong>the</strong> field, especially in its central part. During <strong>the</strong> whole season in<br />

Krasnodar region P. cupreus (fig. 18) was more abundant on <strong>the</strong> field edges. It is possible that<br />

in <strong>the</strong> South <strong>of</strong> Russia <strong>the</strong>se beetles can overwinter within <strong>the</strong> field or occupy <strong>the</strong> field edges<br />

very early in spring (February - beginning <strong>of</strong> March). Density <strong>of</strong> H. rufipes populations in<br />

Krasnodar was relatively low (fig. 20) and in all sites was highest only in <strong>the</strong> beginning <strong>of</strong><br />

July.


158<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

31.3<br />

10.4<br />

20.4<br />

30.4<br />

11.5<br />

20.5<br />

30.5<br />

9.6<br />

19.6<br />

28.6<br />

8.7<br />

Dates<br />

Fig. 9. Trombidiformes. Krasnodar, 1999. Pitfall traps.<br />

N<br />

1,6<br />

1,4<br />

1,2<br />

1<br />

0,8<br />

0,6<br />

0,4<br />

0,2<br />

Margin<br />

Edge<br />

Centre<br />

0<br />

3.5<br />

11.5<br />

19.5<br />

28.5<br />

5.6<br />

13.6<br />

22.6<br />

30.6<br />

9.7<br />

18.7<br />

26.7<br />

Dates<br />

Fig. 10. Tetragnatha extensa (L.). Moscow, 1996. Sweeping.<br />

1,2<br />

N<br />

1<br />

0,8<br />

0,6<br />

0,4<br />

Margin<br />

Edge<br />

Centre<br />

0,2<br />

0<br />

3.5<br />

11.5<br />

19.5<br />

28.5<br />

5.6<br />

13.6<br />

22.6<br />

30.6<br />

9.7<br />

18.7<br />

26.7<br />

Dates<br />

Fig. 11. Xysticus ulmi (Hahn). Moscow, 1996. Sweeping.


159<br />

N<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

0<br />

27.3<br />

5.4<br />

15.4<br />

25.4<br />

6.5<br />

15.5<br />

26.5<br />

4.6<br />

12.6<br />

23.6<br />

Dates<br />

Fig. 12. Misumenops tricuspidatus Fab. Krasnodar, 1999. Sweeping.<br />

N<br />

1,8<br />

1,6<br />

1,4<br />

1,2<br />

1<br />

0,8<br />

0,6<br />

0,4<br />

0,2<br />

0<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

27.3<br />

5.4<br />

15.4<br />

25.4<br />

6.5<br />

15.5<br />

26.5<br />

4.6<br />

12.6<br />

23.6<br />

Dates<br />

Fig. 13. Hypsosinga pygmaea (Sud.). Krasnodar, 1999. Sweeping.<br />

N<br />

3,5<br />

3<br />

2,5<br />

2<br />

1,5<br />

1<br />

Margin<br />

Edge<br />

Centre<br />

0,5<br />

0<br />

29.4<br />

7.5<br />

15.5<br />

23.5<br />

1.6<br />

9.6<br />

17.6<br />

25.6<br />

4.7<br />

13.7<br />

22.7<br />

30.7<br />

Dates<br />

Fig. 14. Agonum muelleri (Hbst.) Moscow, 1996. Pitfall traps.


160<br />

N<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

31.3<br />

10.4<br />

20.4<br />

30.4<br />

11.5<br />

20.5<br />

30.5<br />

9.6<br />

19.6<br />

28.6<br />

8.7<br />

Dates<br />

Fig. 15. Brachinus elegans Chd. and B. psophia Serv. Krasnodar, 1999. Pitfall traps.<br />

N<br />

30<br />

25<br />

20<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

15<br />

10<br />

5<br />

0<br />

31.3<br />

10.4<br />

20.4<br />

30.4<br />

11.5<br />

20.5<br />

30.5<br />

9.6<br />

19.6<br />

28.6<br />

8.7<br />

Dates<br />

Fig. 16. Agonum dorsale Pont. Krasnodar, 1999. Pitfall traps.<br />

N<br />

50<br />

40<br />

30<br />

20<br />

Adjoined<br />

biotopes<br />

Margin<br />

Edge<br />

Centre<br />

10<br />

0<br />

28.4<br />

3.5<br />

13.5<br />

28.5<br />

7.6<br />

17.6<br />

27.6<br />

7.7<br />

17.7<br />

27.7<br />

6.8<br />

1.9<br />

11.9<br />

Dates<br />

Fig. 17. Pterostichus cupreus L. Moscow, 1998. Pitfall traps.


161<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

N<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

31.3<br />

10.4<br />

20.4<br />

30.4<br />

11.5<br />

20.5<br />

30.5<br />

9.6<br />

19.6<br />

28.6<br />

8.7<br />

Dates<br />

Fig. 18. Pterostichus cupreus L. Krasnodar, 1999. Pitfall traps.<br />

28.4<br />

3.5<br />

13.5<br />

28.5<br />

7.6<br />

17.6<br />

27.6<br />

7.7<br />

17.7<br />

27.7<br />

6.8<br />

1.9<br />

11.9<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

N<br />

Adjoined biotopes<br />

Margin<br />

Edge<br />

Centre<br />

Dates<br />

Fig. 19. Harpalus rufipes (De Geer) Moscow, 1998. Pitfall traps.<br />

6<br />

N<br />

5<br />

4<br />

3<br />

2<br />

Shelter belt<br />

Margin<br />

Edge<br />

Halfway<br />

Centre<br />

1<br />

0<br />

31.3<br />

10.4<br />

20.4<br />

30.4<br />

11.5<br />

20.5<br />

30.5<br />

9.6<br />

19.6<br />

28.6<br />

8.7<br />

Dates<br />

Fig. 20. Harpalus rufipes (De Geer) Krasnodar, 1999. Pitfall traps.


162<br />

Only some species <strong>of</strong> Staphylinidae were able to inhabit <strong>the</strong> field central parts in Moscow<br />

region (i.e. Tachyporus hypnorum (Fab.), Philonthus cognatus Stephens, P. rotundicollis<br />

(Ménétries), some species <strong>of</strong> Aleocharinae). Many specimens <strong>of</strong> <strong>the</strong> staphylinid Arpedium<br />

quadrum Grav. was found in <strong>the</strong> field in autumn and spring. All rove beetles being good fliers<br />

colonised <strong>the</strong> field at once.<br />

Densities <strong>of</strong> populations <strong>of</strong> aphidophagous insects were relatively low. Adults <strong>of</strong> all<br />

species <strong>of</strong> ladybirds inhabited forest belts and field margins in early spring. The most<br />

abundant coccinellid, Propylaea 14-punctata L., remained <strong>the</strong>re throughout <strong>the</strong> whole season.<br />

On <strong>the</strong> contrary, Coccinella septempunctata L., Coccinula quatuordecimpustulata (L.) and<br />

Hippodamia variegata Goeze occurred later in <strong>the</strong> central part <strong>of</strong> <strong>the</strong> field. Larvae were<br />

caught only in June; <strong>the</strong>y obviously preferred <strong>the</strong> places <strong>of</strong> aphids aggregation (fig. 4).<br />

Adults <strong>of</strong> most abundant lacewing, Chrysoperla carnea Stephens, and less numerous<br />

C. septempunctata occupied only forest belts and field margins. We did not observed <strong>the</strong>m in<br />

<strong>the</strong> field at day time but <strong>the</strong>y might migrate into <strong>the</strong> field at night. On <strong>the</strong> contrary, C. carnea<br />

occurred in <strong>the</strong> field at day time beginning from May. In early spring this species could too be<br />

found on margins and forest belts. Larvae <strong>of</strong> Chrysoperla spp. were found only in June, being<br />

evenly distributed throughout <strong>the</strong> field.<br />

Adults <strong>of</strong> Syrphidae were practically absent in <strong>the</strong> samples. Their larvae occurred mainly<br />

in <strong>the</strong> field, especially in its central part, but <strong>the</strong>ir density was low. It is not excluded that we<br />

missed <strong>the</strong> most part <strong>of</strong> <strong>the</strong>se larvae because <strong>the</strong>y are active usually at night (Berest, 1981).<br />

Discussion<br />

The arthropod complex in <strong>the</strong> field is influenced during <strong>the</strong> vegetative season by <strong>the</strong><br />

following factors: 1) arthropods which are able to overwinter in <strong>the</strong> field (red mites, some<br />

spiders and, possibly, ground beetles); 2) arthropods which overwinter in field margins and/or<br />

neighbouring biotopes and colonise <strong>the</strong> field "step by step", (<strong>the</strong> majority <strong>of</strong> ground beetles<br />

and spiders, <strong>the</strong> chrysomelid O. melanopus, ladybirds); 3) arthropods which overwinter also<br />

outside <strong>the</strong> field and migrate to <strong>the</strong> field ei<strong>the</strong>r by active flight (bugs Eurygaster spp. and<br />

Lygus spp., staphylinids like Philonthus spp.) or as a result <strong>of</strong> "air plankton" sedimentation<br />

(especially small chrysomelids like Phyllotreta spp., small staphylinids and frit flies Oscinella<br />

spp.). Such arthropods are able to colonise all parts <strong>of</strong> <strong>the</strong> field at once.<br />

Every species has its own spatial and temporal pattern <strong>of</strong> distribution in agro-ecosystem.<br />

For a successful management <strong>of</strong> natural enemies we have to devote our attention only to such<br />

species that are capable to inhabit <strong>the</strong> field.<br />

Our results show that field margins are <strong>the</strong> main source <strong>of</strong> many beneficial species. We<br />

should study <strong>the</strong> biology <strong>of</strong> <strong>the</strong>se species in details to elaborate optimal conditions for<br />

survival <strong>of</strong> <strong>the</strong>se species. It is extremely important also to know <strong>the</strong> conditions which are<br />

favourable for beneficial arthropods overwintering in <strong>the</strong> field.<br />

The field itself is not uniform. The arthropod complexes in <strong>the</strong> centre <strong>of</strong> <strong>the</strong> field and its<br />

edges differ. Our results show also that <strong>the</strong> arthropods can distinguish also parts <strong>of</strong> field at a<br />

distance <strong>of</strong> 10, 200 and 400 m from its edge.<br />

References<br />

Berest, Z.A., 1981: Diurnal migrations <strong>of</strong> aphidophagous predators on cereal crops. In:<br />

Behaviour <strong>of</strong> insects as a base for elaboration <strong>of</strong> pest control. Minsk: 24-26.


Dennis, G. & Fray, G.L.A. 1992: Field margins: can <strong>the</strong>y enhance natural enemy population<br />

densities and general arthropod diversity on farmland? Agric. Ecosystems and Environ.<br />

40: 95-115.<br />

Kiss, J., Kadar, F., Toth, J, Kozma, E. & Toth, F. 1994: Occurrence <strong>of</strong> predatory arthropods in<br />

winter wheat and in <strong>the</strong> field edge. Ecologie. 25: 127-132.<br />

Kromp, B. & Steinberger, K.H. 1992: Grassy field, margins and arthropod diversity: a case<br />

study on ground beetles and spiders in eastern Austria (Coleoptera: Carabidae;<br />

Arachnida: Aranei, Opiliones). Agric. Ecosystems and Environ. 40: 71-93.<br />

Melnichenko, A.N. 1949: Forest belts and beneficial and harmful animals reproduction.<br />

Moscow. MOIP: 1-358 (in Russian).<br />

Peter, G., Kiss, J., Toth, F., Sasvari, M. & Petz, A. 1999: Role <strong>of</strong> field margins in <strong>the</strong> winter<br />

phenophase <strong>of</strong> carabid beetles (Coleoptera: Carabidae) in winter wheat field. In:<br />

Integrated control in cereal crops. Abstracts. Gödöllő: 21.<br />

So<strong>the</strong>rton, N.W. 1985: The distribution and abundance <strong>of</strong> predatory Coleoptera overwintering<br />

in field boundaries. Ann. Appl. Biol. 106: 17-21.<br />

Thomas, M.B., So<strong>the</strong>rton, N.W., Coombes, D.S. & Wratten S.D. 1992: Habitat factors<br />

affecting <strong>the</strong> distribution <strong>of</strong> polyphagous predatory insects between field boundaries.<br />

Ann. Appl. Biol. 120: 197-202.<br />

Toth, F. & Kiss, J. 1999: Comparative analyses <strong>of</strong> epigeic spider assemblages in Nor<strong>the</strong>rn<br />

Hungarian winter wheat fields and <strong>the</strong>ir adjacent margins. J. <strong>of</strong> Arachnology. 27: 241-<br />

248.<br />

Wallin, H. 1985: Spatial and temporal distribution <strong>of</strong> some abundant carabid beetles (Coleoptera:<br />

Carabidae) in cereal fields and adjacent habitats. Pedobiologia. 28: 19-34.<br />

163


164


Integrated Control in Cereal Crops<br />

<strong>IOBC</strong> wprs Bulletin 24 (6) 2001<br />

pp. 165 - 173<br />

Studies <strong>of</strong> <strong>the</strong> pests <strong>of</strong> Canary-grass (Phalaris canariensis L.)<br />

Kozma, E., Gólya. G. & Z. Záhorszki<br />

Szent István University, Faculty <strong>of</strong> Agricultural. and Environmental Sciences Department <strong>of</strong><br />

Plant Protection, H-2100 Gödöllő, Páter K. u. 1. Hungary<br />

Summary<br />

The aim <strong>of</strong> <strong>the</strong> present study was to gain knowledge about insect pests <strong>of</strong> Canary-grass (Phalaris<br />

canariensis L.), because little information is available on its pests and diseases and <strong>the</strong>ir dynamics. We<br />

conclude, that Oulema melanopus L. could be one <strong>of</strong> <strong>the</strong> dangerous pests on Canary-grass. In Hungary<br />

under optimal ecological circumstances Rhopalosiphum padi L. and <strong>the</strong> newly occurring Diuraphis<br />

noxia (Mordwilko) increase rapidly on <strong>the</strong> leaves and ears <strong>of</strong> Canary-grass.<br />

Introduction<br />

Canary-grass (Phalaris canariensis L.) is grown in a relatively small area in Hungary and it is<br />

mostly exported as feed for birds. Canary-grass belongs to <strong>the</strong> Gramineae family. It has a<br />

poor fibrous root system and its stalk is thin straw. Its inflorescence is panicle shape (Fazekas,<br />

1997; Antal, 1987).<br />

In Hungary four species <strong>of</strong> barley beetles occur on cereals: Oulema melanopus L., O.<br />

rufocyanea (Suffrian), O. gallaeciana (Heyden) and O. septentrionis (Weise) (Kaszab 1962).<br />

These barley beetles originally lived on wild Gramineae spp. Later <strong>the</strong>y changed to cereals<br />

(Kadocsa, 1957). First <strong>the</strong>y caused significant damage on spring barley and oat, <strong>the</strong>n <strong>the</strong>y<br />

became <strong>the</strong> most important pest <strong>of</strong> winter wheat (Szilágyi et al., 1986). Labeyrie (cit.<br />

Balachowsky, 1963) mentioned Canary-grass as a host plant. Barley beetles are not only<br />

dangerous because <strong>of</strong> <strong>the</strong>ir direct damage to foliage, but also for <strong>the</strong>ir role as a virus vector<br />

(Princinger, 1991). The great grain aphid Sitobion avenae Fabr. is considered to be <strong>the</strong> most<br />

frequent cereal aphid by Szalay-Marzsó (1970). Kozma et al. (1994) found, that <strong>the</strong> most<br />

common aphid species near Gödöllő were Schizaphis graminum Rondani, Rhopalosiphum<br />

padi L., Metopolophium dirhodum Walker, Sitobium avenae and Diuraphis noxia<br />

(Mordvilko). D. noxia first appeared in a yellow pan trap in 1989 in <strong>the</strong> main cereal growing<br />

area <strong>of</strong> central Hungary (Basky & Eastop, 1991). In 1993 it was found in <strong>the</strong> territory <strong>of</strong> <strong>the</strong><br />

River Tisza, Transdanubia and on <strong>the</strong> Great Plain (Basky, 1993a, 1993b), and on maize<br />

(Kozma et al., 1995). D. noxia is indigenous in <strong>the</strong> Mediterranean area and in <strong>the</strong> Iranian<br />

Turkestanian mountain range (Mordvilko, 1929; Nevskii, 1929). It has been a pest on wheat<br />

and barley in sou<strong>the</strong>rn Russia since before 1900 (Grossheim, 1914). The first published<br />

outbreak <strong>of</strong> D. noxia was in <strong>the</strong> Crimea (Mokrzhetsky 1901) and was associated with <strong>the</strong><br />

increase in <strong>the</strong> wheat growing area in <strong>the</strong> Ukraine. It was recorded as a serious pest <strong>of</strong> cereals<br />

in Ukraine in 1972 (Dyadechko & Ruban, 1975), in <strong>the</strong> USA (Pike, 1991) and South-Africa<br />

(Aalbersberg et al. 1987). D. noxia overwinters in <strong>the</strong> egg stage on winter wheat in Hungary.<br />

After harvest it lives on volunteer wheat, on Sorghum halapense and o<strong>the</strong>r Gramineae weeds.<br />

The alatae developing on <strong>the</strong>se plants fly to winter wheat where <strong>the</strong> males and females<br />

develop, and after fertilisation <strong>the</strong> oviparae lay eggs which overwinter. It has six to eight<br />

generations per year (Burton, 1989; Pike, 1991). D. noxia causes white or purple streaks and<br />

rolled leaves (Hewitt et al. 1984), perhaps by injecting a toxin and by disturbing tissue water<br />

165


166<br />

balance during feeding (Fouche et al., 1984; Basky, 1993). Injured ears are below <strong>the</strong><br />

standard in growth and spikes may be empty. Their peaks are not able to emerge from <strong>the</strong><br />

curled leaf sheath. Typical damage symptoms include spikes curving orbital and wavy spike<br />

axle. The wingless form <strong>of</strong> D. noxia is spindle shaped, light green coloured with a length <strong>of</strong><br />

1,2-2,4 mm. The alate vivipara is 1,5-2 mm long; its ventral part is pale green. Antennae <strong>of</strong><br />

alatae are shorter than half <strong>the</strong> body length; siphunculi are mammariform, shorter, than wide<br />

(Basky, 1993c). A typical morphological character is <strong>the</strong> supracaudal process below <strong>the</strong><br />

tongue-shaped cauda (Blackman & Eastop, 1984; Basky, 1993b).<br />

Material and methods<br />

A survey was carried out in 1997 on an experimental farm in József major (Szent István<br />

University, Gödöllő) in a canary-grass field <strong>of</strong> 5 ha. Samples were taken every ten days from<br />

mid May till harvest (July)<br />

The population fluctuation, biology and damage <strong>of</strong> Cereal leaf beetles.<br />

We made five replications <strong>of</strong> ten sweeps both at <strong>the</strong> margin <strong>of</strong> <strong>the</strong> field and 50 m from <strong>the</strong><br />

field margin. The population fluctuation and biology <strong>of</strong> Cereal leaf beetles were recorded.<br />

Species during <strong>the</strong> observation were identified immediately. In order to assess <strong>the</strong> damage we<br />

measured <strong>the</strong> height <strong>of</strong> 100 seriously injured, and 100 healthy plants, and <strong>the</strong> length <strong>of</strong> ears.<br />

The method <strong>of</strong> Aphid population fluctuation and species identification<br />

At <strong>the</strong> field margin and 50 m from <strong>the</strong> margin <strong>of</strong> <strong>the</strong> field aphids were washed <strong>of</strong>f from 100<br />

(each location) plants (both <strong>the</strong> panicles and <strong>the</strong> vegetative parts <strong>of</strong> <strong>the</strong> plants) with a brush<br />

and stored in Oudemans solution until identification. Species were identified with a stereo<br />

microscope.<br />

Results<br />

The population fluctuation, biology and damage <strong>of</strong> Cereal leaf beetles<br />

At <strong>the</strong> margin <strong>of</strong> <strong>the</strong> field we found two O. melanopus and one O. gallaeciana adults in<br />

50 sweeps during <strong>the</strong> first survey. In <strong>the</strong> next survey we found a huge number <strong>of</strong> larvae and<br />

also observed significant damage symptoms <strong>of</strong> larvae (32 individuals / 50 sweeps). Larva<br />

population did not change until 27 May. At <strong>the</strong> beginning <strong>of</strong> June <strong>the</strong> field was treated with an<br />

insecticide, so we could not continue examining <strong>the</strong> larva population. New imagoes appeared<br />

on <strong>the</strong> plants in late June (fig. 1).<br />

Inside <strong>the</strong> field Cereal leaf beetles appeared after 10 May especially considering <strong>the</strong><br />

higher number <strong>of</strong> O. melanopus (7 individuals / 50 sweeps), and O. gallaeciana (1 individual<br />

/ 50 sweeps). By <strong>the</strong> second survey <strong>the</strong> larvae had increased intensively (four times <strong>the</strong><br />

population at <strong>the</strong> margin <strong>of</strong> <strong>the</strong> field, 112, 138 individuals / 50 sweeps). Intensive larval<br />

hatching occurred during <strong>the</strong> second half <strong>of</strong> May. We did not continue <strong>the</strong> observation <strong>of</strong> <strong>the</strong><br />

population due to <strong>the</strong> previously mentioned insecticide treatment. New imagoes appeared on<br />

plants inside <strong>the</strong> field at <strong>the</strong> end <strong>of</strong> June (fig. 2).<br />

The damage symptoms <strong>of</strong> Cereal leaf beetles: The average length <strong>of</strong> canarygrass<br />

which was seriously damaged by cereal leaf beetles was 70.8 cm. The average stalk<br />

length <strong>of</strong> 100 healthy canary grass stalk was 83.8 cm. The average panicle length <strong>of</strong> damaged<br />

plants was 2.9 cm, and <strong>of</strong> healthy ones 3.3 cm (measurements were done on 12 June).


167<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

Larvae<br />

Lema lichenis<br />

Lema melanopus<br />

10<br />

5<br />

0<br />

14.05 23.05 27.05 06.06 12.06 18.06 26.06 10.07<br />

Fig. 1. Population dynamics <strong>of</strong> Cereal leaf beetles at <strong>the</strong> edge <strong>of</strong> <strong>the</strong> field (1997, Gödöllö)<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

Larvae<br />

Lema lichenis V.<br />

Lema melanopus L.<br />

40<br />

20<br />

0<br />

14.05 23.05 27.05 06.06 12.06 18.06 26.06 10.07<br />

Fig. 2. Population dynamics <strong>of</strong> Cereal leaf beetles inside <strong>the</strong> field (1997, Gödöllö)<br />

Results <strong>of</strong> aphids population fluctuation on Canary-grass<br />

At <strong>the</strong> margin <strong>of</strong> <strong>the</strong> field aphids appeared on <strong>the</strong> vegetative parts <strong>of</strong> <strong>the</strong> plants on 23<br />

May. At that time M. dirhodum was <strong>the</strong> most abundant species (11 individuals / 100 plants),<br />

followed by R. padi (4 individuals / 100 plants) and D. noxia (3 individuals / 100 plants).<br />

Multiplication <strong>of</strong> R. padi and D. noxia became very intensive from <strong>the</strong> 2 nd week <strong>of</strong> June,


168<br />

while <strong>the</strong> number <strong>of</strong> M. dirhodum decreased during <strong>the</strong> same period. We also found some<br />

individuals <strong>of</strong> S. avenae during <strong>the</strong> 2 nd week <strong>of</strong> June. The aphid population reached its peak at<br />

<strong>the</strong> end <strong>of</strong> June (R. padi 813 individuals/100 pants; D. noxia 518 individuals/ 100 plants;<br />

M. dirhodum 29 individuals / 100 plants). By 10 July <strong>the</strong> aphid population collapsed (fig. 3).<br />

Species composition <strong>of</strong> aphids in <strong>the</strong> field margin during <strong>the</strong> vegetation period was: R. padi<br />

63 %, D. noxia 35 %, M. dirhodum and S. avenae 2 % (fig. 4).<br />

1600<br />

1400<br />

individual number/100 plants<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

Metopolophium dirhodum<br />

Sitobion avenae<br />

Diuraphis noxia<br />

Rhopalosiphum padi<br />

200<br />

0<br />

23.05. 27.05. 06.06. 12.06. 18.06. 26.06. 10.07.<br />

Fig. 3. Canges in proportions <strong>of</strong> Aphid species at <strong>the</strong> edge <strong>of</strong> <strong>the</strong> field (1997, Gödöllö)<br />

Sitobion avenae<br />

0,2%<br />

Metopolophium<br />

dirhodum<br />

2,2%<br />

Diuraphis noxia<br />

34,9%<br />

Rhopalosiphum<br />

padi<br />

62,6%<br />

Fig. 4. Composition <strong>of</strong> Aphid species at <strong>the</strong> edge <strong>of</strong> <strong>the</strong> field (100 plants) (1997, Gödöllö)


169<br />

Inside <strong>the</strong> field aphids appeared two weeks later, than in <strong>the</strong> field margin (6 June). At <strong>the</strong><br />

beginning R. padi was <strong>the</strong> most abundant aphid species (450 individuals / 100 plants). Its<br />

number increased slightly (555 individuals / 100 plants), while D. noxia appeared in low<br />

numbers (46 individuals / plants), but later it became more abundant (1003 individuals / 100<br />

plants). Metopolophium dirhodum and S. avenae were not present in significant numbers on<br />

plants. The aphid population collapsed at <strong>the</strong> beginning <strong>of</strong> July (fig. 5). Composition <strong>of</strong> aphid<br />

species on plants inside <strong>the</strong> field was as follows: D. noxia 53 %, R. padi 45 %, M. dirhodum<br />

and S. avenae 2 % (fig. 6).<br />

1800<br />

1600<br />

individual number/100 plant<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

Metopolophium dirhodum<br />

Sitobion avenae<br />

Diuraphis noxia<br />

Rhopalosiphum padi<br />

200<br />

0<br />

23.05. 27.05. 06.06. 12.06. 18.06. 26.06. 10.07.<br />

Fig. 5. Changes in proportions <strong>of</strong> Aphid species inside <strong>the</strong> field (1997, Gödöllö)<br />

Sitobion avenae<br />

0,06%<br />

Metopolophium<br />

dirhodum<br />

1,8%<br />

Rhopalosiphum<br />

padi<br />

45,4%<br />

Diuraphis noxia<br />

52,6%<br />

Fig. 6. Composition <strong>of</strong> Aphid species inside <strong>the</strong> field (100 plants) (1997, Gödöllö)


170<br />

In <strong>the</strong> field margin aphids appeared on <strong>the</strong> panicle (ears) at <strong>the</strong> time <strong>of</strong> <strong>the</strong> 6 th survey (18<br />

June). Rhopalosiphum padi was more abundant (214 individuals/panicles), followed by<br />

D. noxia (851 individuals / panicles). By <strong>the</strong> time <strong>of</strong> <strong>the</strong> next survey <strong>the</strong> number <strong>of</strong> D. noxia<br />

increased (851 individuals / 100 panicles), while R. padi showed a lower abundance (502<br />

individuals/100 panicles) (fig. 7). During <strong>the</strong> vegetation period at <strong>the</strong> field margin <strong>the</strong>re were<br />

56% D. noxia, 43 % R. padi, 1 % S. avenae on <strong>the</strong> panicles (fig. 8).<br />

1600<br />

1400<br />

individual number/100 ears<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

Sitobion avenae<br />

Diuraphis noxia<br />

Rhopalosiphum padi<br />

200<br />

0<br />

23.05. 27.05. 06.06. 12.06. 18.06. 26.06. 10.07.<br />

Fig. 7. Changes in proportions <strong>of</strong> Aphid species on <strong>the</strong> ears at <strong>the</strong> edge <strong>of</strong> <strong>the</strong> field (1997,<br />

Gödöllö)<br />

Sitobion avenae<br />

1%<br />

Diuraphis noxia<br />

56%<br />

Rhopalosiphum<br />

padi<br />

43%<br />

Fig. 8. Composition <strong>of</strong> Aphid species on <strong>the</strong> ears at <strong>the</strong> edge <strong>of</strong> <strong>the</strong> field (100 ears) (1997,<br />

Gödöllö)


171<br />

Inside <strong>the</strong> field aphids appeared at <strong>the</strong> same time as in <strong>the</strong> margin (18 June).<br />

Diuraphis noxia was more abundant than at <strong>the</strong> margin (99 individuals / 100 panicles; R. padi<br />

102 individuals / 100 panicles). By <strong>the</strong> time <strong>of</strong> <strong>the</strong> next assessment (26 June) <strong>the</strong>se numbers<br />

changed as follows: R. padi 379 / 100 panicles, D. noxia 1428 / 100 panicles (fig. 9). Inside<br />

<strong>the</strong> field 84 % were D. noxia and 16% R. padi (fig. 10).<br />

2000<br />

1800<br />

1600<br />

individual number/100 ears<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

Sitobion avenae<br />

Diuraphis noxia<br />

Rhopalosiphum padi<br />

200<br />

0<br />

23.05 27.05 06.06 12.06 18.06 26.06 10.07<br />

Fig. 9. Changes in proportions <strong>of</strong> Aphid species on <strong>the</strong> ears inside <strong>the</strong> field (1997, Gödöllö)<br />

Sitobion avenae<br />

0,2 %<br />

Rhopalosiphum<br />

padi<br />

16,4 %<br />

Diuraphis noxia<br />

83,4%<br />

Fig. 10. Composition <strong>of</strong> Aphid species on <strong>the</strong> ears inside <strong>the</strong> field (100 ears) (1997, Gödöllö)


172<br />

The localisation <strong>of</strong> aphids on <strong>the</strong> plant<br />

At <strong>the</strong> end <strong>of</strong> May most M. dirhodum and R. padi were found on <strong>the</strong> abaxial surface <strong>of</strong> <strong>the</strong><br />

leaves. Diuraphis noxia appeared later in higher numbers under <strong>the</strong> shelter <strong>of</strong> <strong>the</strong> rolled top<br />

leaves. After panicles developed it increased on <strong>the</strong> clusters. We also found some individuals<br />

<strong>of</strong> R. padi on <strong>the</strong> clusters.<br />

Discussion<br />

Oulema melanopus could be one <strong>of</strong> <strong>the</strong> most dangerous pests on Canary-grass, infesting<br />

initially in May and causing significant damage if wea<strong>the</strong>r conditions are suitable.<br />

Temperature about 25 °C and high humidity are favourable for oviposition and hatching <strong>of</strong><br />

Cereal leaf beetles. Due to <strong>the</strong> higher plant density in <strong>the</strong> middle <strong>of</strong> <strong>the</strong> field, <strong>the</strong> humidity<br />

was higher resulting in more abundant cereal leaf beetle populations. If <strong>the</strong> wea<strong>the</strong>r is cold<br />

and rainy, population development is slower. In this study <strong>the</strong> length <strong>of</strong> stems and that <strong>of</strong> <strong>the</strong><br />

ears <strong>of</strong> canary grass were 15,5 % and 11,2 % shorter, respectively, as a result <strong>of</strong> <strong>the</strong> damage<br />

caused by Oulema spp.<br />

In Hungary under optimal wea<strong>the</strong>r conditions (temperature around 20-25 °C) R. padi and<br />

<strong>the</strong> newly occurred D. noxia increase rapidly on <strong>the</strong> leaves and ears <strong>of</strong> Canary-grass. Inside<br />

<strong>the</strong> field temperature and humidity were more balanced, resulting in <strong>the</strong> greater number <strong>of</strong><br />

aphids. Diuraphis noxia has a perfect sense for finding host plants (Aalbersberg, 1987). This<br />

is supported by Kozma et al. (1995) observation finding D. noxia colonies on maize. Even if<br />

Canary grass is grown in a small area in Hungary, under suitable ecological circumstances it<br />

can be colonised by high numbers <strong>of</strong> D. noxia as well as <strong>the</strong> sparsely emerged spring barley<br />

(Basky, 1997) causing high damage.<br />

Acknowledgement<br />

We which to thank to Pr<strong>of</strong>essor Alan C. York, Purdue University, for his help in reviewing<br />

<strong>the</strong> manuscript.<br />

References<br />

Aalbersberg, Y.K. 1987: Ecology <strong>of</strong> <strong>the</strong> wheat aphid Diuraphis noxia (Mordvilko) in <strong>the</strong><br />

Eastern Orange Free State. M.Sc. <strong>the</strong>sis, Univ. <strong>of</strong> Orange Free State, Bloemfontain,<br />

South Africa.<br />

Basky, Zs. & Eastop, V.F. 1991: Diuraphis noxia in Hungary. Barley Yellow Dwarf Newsletter<br />

4: 34.<br />

Basky, Zs. 1993 a: The Abundance <strong>of</strong> indigenous cereal aphids and occurrence <strong>of</strong> Diuraphis<br />

noxia in Hungary. Hungarian Agricultural Research 2(2): 14-16.<br />

Basky, Zs. 1993 b: Incidence and population fluctuation <strong>of</strong> Diuraphis noxia in Hungary. Crop<br />

Protection 12 (8): 605-608.<br />

Basky, Zs. 1993 c: Identification key for alate aphids caught in yellow pan traps. Acta<br />

Phytophathol. Entomol. Hung. 28: 71-121.<br />

Basky, Zs. & Jordan J. 1997: Biotypic variation <strong>of</strong> Diuraphis noxia (Homoptera: Aphididae)<br />

between South Africa and Hungary. J. Econ. Entomol. 90: 623-627.<br />

Fazekas, M. (szerk.) 1997: Amit a cirok- és madáreleség-félékről tudni kell. Agroinform<br />

Kiadó és Nyomda Kft. Budapest: 97-98.<br />

Kozma, E., Kiss, J. & Tóth, I. 1994: Őszi búzán károsító levéltetű fajok vizsgálata Gödöllő<br />

térségében. 40. Növényvédelmi Napok kiadványa 57.


Kozma, E., Kiss, J. & Tóth, F. 1995: A kukoricán károsító levéltetvek denzitásának és faji<br />

összetételének változása az ökológiai viszonyok tükrében. Növényvédelem 31: 485-493.<br />

Kozma, E. 1996: Károsító levéltetvek és takácsatkák és a hozzájuk kapcsolódó hasznos élő<br />

szervezetek faunisztikai vizsgálata. Doktori (Ph.D) <strong>the</strong>zisek. Gödöllő.<br />

Milne, W.M. 1987: Notes on visit to South Africa for CSIRO Division <strong>of</strong> Entomology,<br />

November 1986. Commonwealth Scientific and Industrial Organisation, Division <strong>of</strong><br />

Entomology, Canberra, Australia, Mimeo Rep.<br />

Mordvilko, A.K. 1929: Food plant catalogue <strong>of</strong> <strong>the</strong> Aphididae <strong>of</strong> USSR. Works Appl.<br />

Entomol. Leningrád, 14(1): 101. [cited in: Kovalev, O.V., Poprawski, T.J., Stekolshchikov,<br />

A.V., Vereshchagina, A.B. & Grandrapur, S.A. 1991: Diuraphis aizenberg (Hom.:<br />

Aphididae): key to apterous females, and review <strong>of</strong> Russian language literature on <strong>the</strong><br />

natural history <strong>of</strong> Diuraphis noxia (Kurdjumov 1913). J. Appl. Entomol. 112: 425-436.<br />

Perez, B., Collar, J.L., Avilla, C., Duque, M. & Fereres, A. 1995: Estimation <strong>of</strong> vector<br />

propensity <strong>of</strong> potato virus Y in open field pepper crops <strong>of</strong> Central Spain. J. Econ.<br />

Entomol. 88 (4): 986-991.<br />

Szabolcs, J. 1974: Vizsgálatok a gabonaféléket károsító Lema (Col.: Chrysomelidae) –<br />

fajokkal kapcsolatban. Növényvédelem 10: 389-393.<br />

Szalay-Marzsó, L. 1970: Adatok a hazai gabonalevéltetvek ismeretéhez. Növényvédelem 6:<br />

244-250.<br />

173

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!